Sie sind auf Seite 1von 72

CATALYSIS DEVELOPMENT TO ACHIEVE SUSTAINABLE SYNTHESIS OF ADIPIC ACID PRODUCTION WITH HYDROGEN PEROXIDE AS OXIDANT A REVIEW

A Dissertation submitted to the University of Manchester for the degree of the Master of Science in the Faculty of Engineering and Physical Science

2010

ABDIL HALIMIS STANI

SCHOOL OF CHEMICAL ENGINEERING AND ANALYTICAL SCIENCE

TABLE OF CONTENTS
TABLE OF CONTENTS .................................................................................................................................. 2 LIST OF TABLES .............................................................................................................................................. 4 LIST OF FIGURES ............................................................................................................................................ 5 ABSTRACT ......................................................................................................................................................... 6 ACKNOWLEDGEMENT ................................................................................................................................. 7 DECLARATION................................................................................................................................................. 8 COPYRIGHT STATEMENT ........................................................................................................................... 9 CHAPTER 1 - INTRODUCTION ............................................................................................................... 10 1.1Background......................................................................................................................................... 10 1.2 Objectives and the Scope of the Review ............................................................................... 11 CHAPTER 2 THE CONVENTIONAL PRODUCTION OF ADIPIC ACID.................................... 14 2.1 Introduction ...................................................................................................................................... 14 2.2 Two-step Oxidation of Cyclohexane ........................................................................................ 14 2.3 Phenol Route ..................................................................................................................................... 16 2.4 Environmental Concerns ............................................................................................................. 17 CHAPTER 3- OXIDATION OF THE CYCLOHEXANE TO OL/ONE BY USING HOMOGENEOUS CATALYSTS ................................................................................................................. 19 3.1 Introduction ...................................................................................................................................... 19 3.2 Iron and Manganese Based Catalysts...................................................................................... 20 3.2.1 Non-heme Mononuclear Iron and Manganese Complexes .............................. 20 3.2.2 Non-heme Binuclear Iron and Manganese Complexes .................................... 23 3.2.3 Non-heme Hexanuclear Iron Complexes ........................................................... 24 3.2.4 Heme Iron Complexes ......................................................................................... 24 3.2.5 Iron Salts .............................................................................................................. 25 3.3 Copper Based catalysts ................................................................................................................. 26 3.4 Vanadium Based Catalysts .......................................................................................................... 29 3.5 Hetero- Metallic Complexes ........................................................................................................ 30 3.5 Polyoxometallates (POMs) .......................................................................................................... 30 3.6 Hybrid System Metal Complexes - Polyoxometallates .................................................... 33 3.7 An Overview of Homogeneous Catalysis Development .................................................. 33 3.7.1 Summary of the Recent Research Progress ...................................................... 33 3.7.2 Future Prospect in Homogeneous Catalysis Development .............................. 34

CHAPTER 4 OXIDATION OF CYCLOHEXANE TO OL/ONE BY USING HETEROGENEOUS CATALYSTS .................................................................................................................................................... 40 4.1 Molecular Sieve Catalysts ............................................................................................................ 40 4.1.1 Titanium Based catalysts .................................................................................... 41 4.1.2 Vanadium Based Catalysts.................................................................................. 41 4.1.3 Cobalt, Iron and Chromium Based Catalysts ..................................................... 43 4.1.4 Cerium Based Catalysts ...................................................................................... 44 4.1.5 Germanic Faujasite ............................................................................................. 45 4.2 Carbon Nitride Polymer ............................................................................................................... 46 4.3 Heterogenizing Metal Complexes into Solid Materials .................................................... 46 4.4 An Overview of Heterogeneous Catalysis developments ............................................... 49 4.4.1Summary of the Recent Research Progress ....................................................... 49 4.7.2 Future Prospect in Heterogeneous Catalysis Development ............................ 51 CHAPTER 5 DIRECT OXIDATION OF CYCLOHEXENE TO ADIPIC ACID............................. 53 5.1 Biphasic reacting System (Phase-Transfer catalysts) ...................................................... 53 5.2 Molecular Sieves catalysts ........................................................................................................... 56 5.3 An Overview: Summary and Future Prospect of Direct Oxidation of Cyclohexene to Adipic Acid ........................................................................................................................................... 57 CHAPTER 6 - CONCLUDING REMARKS AND OUTLOOK ............................................................. 61 GLOSSARY....................................................................................................................................................... 63 REFERENCES ................................................................................................................................................. 65

LIST OF TABLES Table 1-1 Principles of green chemistry ............................................................................. 11 Table 1-2 Common oxidising agent used in the chemical industries ...................... 12 Table 3-1 Comparison of activities of copper catalyst for the cyclohexane oxidation to Ol/One............................................................................................................ 28 Table 3-2 Cyclohexane oxidation catalysed by Polyoxometalates ........................... 31 Table 3-3 Comparison of catalysis performance with cobalt naphtalene versus various homogeneous catalysts .................................................................................... 39 Table 4-1 Summary of heterogeneous catalysis for the cyclohexane oxidation to Ol/One ..................................................................................................................................... 45 Table 4-2 Summary of the cyclohexane oxidation catalysed by immobilised metal complexes .................................................................................................................. 47 Table 5-1 Cyclohexene oxidation via water-organic bi-phase catalytic system in free organic solvent reactions........................................................................................ 55 Table 5-2 Oxidation of Cyclohexene to adipic acid catalysed by molecular sieves .................................................................................................................................................... 57

LIST OF FIGURES

Figure 2-1 Conventional route of commercial production of adipic acid .............. 16 Figure 3-1 Oxidation mechanisms for hydrocarbon in the presence of iron complex hydrogen peroxide........................................................................................ 21 Figure 3-2 In-situ preparation of Schiff-Base ligand ..................................................... 22 Figure 3-3 Molecular structures of complex [Fe (mqmp) (CH3OH) Cl2] ................ 23 Figure 4-1 Summary of researches on heterogenous catalysis developmen ....... 50 Figure 5-1 Direct oxidation of the cyclohexene to adipic acid ................................... 53 Figure 5-2 Reaction mechanisms of the direct oxidation of cyclohexene to adipic acid ........................................................................................................................................... 53 Figure 5-3 Summary of possible synthetic pathways to AA from direct oxidation of cyclohexene ...................................................................................................................... 59

ABSTRACT
Adipic acid is a valuable chemical intermediate used for production of a wide range of materials. Currently, the major adipic acid production uses twostep oxidation of cyclohexane, which proceeds using air and nitric acid as oxidants. However it is carried out under harsh operating conditions, requires cumbersome separations, and accounts for 5-8% of the total global anthropogenic emission of N2O. Hydrogen peroxide is an attractive alternative oxidant because it is capable of carrying the reaction under mild conditions and producing benign by-products. This paper therefore will discuss the catalysis development to achieve a more sustainable synthesis of adipic acid production using hydrogen peroxide as oxidant. This review consists of three main parts. The first two parts are concerned with the homogeneous and heterogeneous catalysis for the oxidation of cyclohexane to the cyclohexanol-cyclohexanone, whilst the last part concentrates on one-step oxidation of cyclohexene to adipic acid. Development of various ligands for homogeneous systems has facilitated significant improved of activities, but the catalysts are often unstable. As for heterogeneous catalysis, many transition metals substituted molecular sieve catalysts shown to be active, but they sometimes lead to leaching of metal. New families of catalysts, e.g. rare earth metals, germanium faujasite and carbon nitride polymer have been shown to be more stable. The biggest barrier to overcome is to combine the advantages of homogeneous and heterogeneous catalysts. Direct synthesis of adipic acid from cyclohexene can be carried out using molecular sieve or biphasic transfer catalysts. This pathway is advantageous because it eliminates one step process and avoids the N2O emission, but the drawback is the relatively cost of feedstock and hydrogen peroxide. It can be concluded that adipic acid synthesis using hydrogen peroxide is unlikely to substitute the current technologies because of low efficiencies of hydrogen peroxide utilisation arise in many cases. However, research in this field remains to attract a great interest and would not close the possibility of implementation in commercial level in the future.

ACKNOWLEDGEMENT

I would like to thank and praise to Allah Almighty for each of His generous blessing and may peace and blessings upon Prophet Muhammad SAW, my family for their everlasting support and my friends for our colourful days in Manchester. I express my sincere gratitude to my supervisor, Dr. Sven L.M Schroeder, which has reviewed my dissertation and broadened my knowledge. I would also thank to Centre for Environmental Technology BPPT Indonesia for giving me permissions in continuing my study.

DECLARATION

I declare that no portion of the work referred to in the dissertation has been submitted in support of an application for another degree or qualification of this or any other university or other institute of learning.

COPYRIGHT STATEMENT

Copyright in this text of this dissertation rests with the author. Copies (by any process) either in full, or extract, may be made only in accordance with instruction given by the author. Details may be obtained from the appropriate Graduate Office. This page must form part of any such copies made. Further copies (by any process) of copies made in accordance with such instructions may not be made without the permission (in writing) of the author. The ownership of any intellectual property rights which may be described in this dissertation is vested in the University of Manchester, subject on any prior agreement to the contrary, and may not be available for use by third parties without the written permission of the University, which will prescribe the terms and conditions of any such agreement. Further information on the conditions under which disclosures and exploitations may take place is available from the Head of the School of Chemical Engineering and Analytical Science.

CHAPTER 1 INTRODUCTION

1.1 Background

Adipic acid, a straight-chain dicarboxylic acid, is one valuable chemical as it is a raw material for the production of Nylon 6.6 and its derivatives can be used in a wide range of application such as resins, plasticisers, food acidulant, lubricant, and polyurethanes (1). In 2008 the global production capacity of adipic acid was achieved at about 2.6 million a metric tonne of which at about 60% of adipic acid is used for production of Nylon 6.6 and fibres, whereas polyutherethanes accounted for almost 24% of total consumption (2). Adipic acid market is growing rapidly with annual global growth is nearly 3%, whilst demand is growing in the range of 5% to 6% annually by 2010 (3). The important milestone of adipic acid production began in around 1935 when DuPont team invented Fibre 6.6 through laboratory research and then commercially introduced as Nylon 6.6 in 1938 (4). In around 1939 the first commercial adipic acid was operated in West Virginia by using phenol as a raw material, however the increase of phenol price in around 1942 led to start up a new route via two-step oxidation of cyclohexane (1). This method is produces a relatively high yield and the most economically viable route, however it potentially generates considerable amount of nitrous oxides, which account for 5-8% of the total global anthropogenic emission of N2O or at about 400, 00 metric tonnes N2O emission per year (5). Moreover the presence of homogeneous catalyst, and/or corrosive solvent initiator is likely to result in cumbersome separations and leaching of metals. Furthermore the oxidation of cyclohexane possess inherent hazard during the processes due to harsh operating conditions. Although the installation of N2O abatement technologies is likely to decrease the N2O emissions with the efficiency range is 90% to 99% reduction of N2O emissions (6), however the prevention of waste is strongly preferred over the treatment or clean up after it generated. Moreover the concepts of
10

green chemistry and cleaner production have been widely introduced to replace command and control approach for environmental protection since around 1990 (7). The twelve principles of green chemistry are shown in Table 1 (8). These offers a standpoint of the greener technologies can be viewed and now become important guidelines to reduce or eliminate the hazardous compounds in the design, manufacture, and application of chemical products (9). Table 1-1 Principles of green chemistry (8)
1. 2. It is better to prevent waste than to treat or clean up waste after it formed Synthetic method should be designed to maximise the incorporation of all materials used into the final products 3. Wherever practicable, synthetic methodologies should be designed to use and generate substances that posses little or no toxicity to the human health and the environment 4. Chemical products should be designed to preserve efficacy of function while reducing toxicity 5. The use of auxiliary substances (e.g. solvents, separation agents, etc.) should be made unnecessary wherever possible and, innocuous 6. Energy requirements should be recognised for their environmental and economic impacts and should be minimised. Synthetic methods should be conducted at ambient temperature and pressure 7. A raw material of feedstock should be renewable rather than depleting wherever technically and economically practicable 8. Unnecessary derivatisation (blocking group, protection/deprotection, temporary modification of physical/chemical processes) should be avoided whenever possible 9. Catalytic reagents (as selective as possible) are superior to stoichiometric reagents 10. Chemical products should be designed to preserve efficacy of function while reducing toxicity 11. Analytical methodologies need to be developed to allow for real-time, in-process monitoring and control prior to the formation of hazardous substances

1.2 Objectives and the Scope of the Review Oxidising agents may affect the formation of product and the rate transfer of oxygen to the substrate (10). The selection of oxidant, therefore, becomes a relevant factor in accomplishing greener processes. A number of different types of oxidants are displayed in Table 1. It can be seen that several oxidising agents such as KMnO4 and CrO3, are not interesting to be exploited because they may result in toxic salts (11), whereas N2O and HNO3 should be avoidable due to its potential aerial emission pollutant by-products. NaClO may offer relatively high oxygen content, but in several occasions, ClO- can form toxic

11

and carcinogenic chloro-carbons (11). C6H5IO, on the other hand, is frequently effective in metal catalysed oxidation, but it is quite expensive (11). Hydrogen peroxide, ozone, molecular oxygen and tert-butyl peroxide seem to be the most interesting choices to be exploited because their high oxygen content and leave no harmful residuals. Molecular oxygen would most often to be preferred because of the economical advantage arising from its abundance in air, but O2 commonly requires high temperatures and pressures to be activated (10). Tert-butyl peroxide has quite high solubility relative to hydrogen peroxide and molecular oxygen. However it contains a relatively low active oxygen species (12). Ozone, O3, is a powerful oxidant, but rarely investigated (13) since the cost of its production is prohibitive. Hydrogen peroxide, therefore, remains the most attractive oxidant from a combined environmental and economic point of view. The potential advantages intrinsic in the use of hydrogen peroxide are quite evident due to its ability to carry out reaction in the relatively mild conditions, producing a benign by-product (water) while having high oxygen content. Table 1-2 Common oxidising agent used in the chemical industries (11; 14)
Oxidant O2 O2/reductor H2O2 N2O O3 KMnO4 HNO3 CrO3 NaOCl CH3COOH tBuOOH C5H11NO2 (NMO) KHSO5 mClC6H4COOOH Me3SiOOSiMe3 NaIO4 PhIO4 C6H5IO Active oxygen content (%) 100 50 47 36.4 33.3 30.4 25 24 21.6 21.1 17.8 13.7 10.5 9.3 9 7.5 7.3 7.3 Product Nothing or H2O H2O H2O N2 O2 Mn(II) salts NOx Cr(III) salts NaCl CH3COOH tBuOOH C5H11NO (NMM) KHSO4 mClC6H4COOH Me3SiOOSiMe3 NaIO3 PhI C6H5I

In a number of studies, catalytic approaches have been attempted to develop greener and more sustainable process for adipic acid productions. This research paper is, therefore, intended to make a comprehensive review of the
12

development of catalysis to obtain adipic acid in high yields in a manner that is environmentally friendly and economically viable in the presence hydrogen peroxide as oxidant. We also limit the scope of this review to cyclohexane oxidation and single step oxidation of cyclohexene to produce adipic acid.

13

CHAPTER 2 CONVENTIONAL TECHNOLOGIES FOR ADIPIC ACID PRODUCTION

2.1 Introduction

Cyclohexane (C6H12) and phenol (C6H5OH) are the common raw materials from which adipic acid is produced. Two-step oxidation of cyclohexane accounts for about 95% of worldwide production and is therefore the common method for the commercial production of adipic acid (15). Adipic acid production based on phenol, on other hand, is rarely used due to several reasons, which include the availability of starting materials, plant size and capital investment (16). Figure 2-1 presents the commercial pathways of adipic acid production.

2.2 Two-step Oxidation of Cyclohexane The commercial operation of this process was first introduced in 1942 as a batch oxidation process, but as manufacturing process demand grew considerably; productivity became an important consideration. As a result, a continuous oxidation has become the preferred method of operation (1). The basic technology underpinning this route remains similar that to introduced in the early production plants. The variation of current commercial production is commonly associated with the manufacturing of KA oil- also referred to KA Mixture or Ol/One- (17). The term KA oil arises from the fact that it contains cyclohexanone, a ketone (K), and cyclohexanol, which is an alcohol (A) (18). In the first step, cyclohexane is oxidised by air in the liquid phase to form cyclohexanone-cyclohexanol at temperatures of 125 to 165C at pressures of 0.8 to 1.5 MPa (19). This range of temperature achieve favourable reaction rates, whilst operating pressures above 0.8 MPa are required in order to maintain cyclohexane in the liquid phase. Without a catalyst or an initiator, this step can be done within 30 min at 165C; however the presence of soluble metal
14

catalysts (e.g. cobalt naphthenate, cobalt 2-ethylhexanoate, manganese or chromium naphthenate) and, additionally, aldehyde or ketone initiators can shorten the reaction to about 5 min. The economical process is normally used at conversions between 6% to 9% in order to achieve selectivity in the range of 60% to 80% (1; 20). The low conversion of cyclohexane is necessary to prevent over-oxidation since KA oil reacts more readily than cyclohexane, inviting further reaction to form cyclohexyl radicals (1; 21). The unreacted cyclohexane, then, is separated and recycled by azeotropic distillation with steam or vacuum distillation (1; 16). Significant improvement of the process has been made since Baskirov and co-workers revealed in the late 1950s the importance of using boric acid (HBO2) to improve hydrocarbon oxidation (22). This method was subsequently adopted by many companies such as Halcon Inc, IFP, Stamicarbon Technology and Institut Francais du Petrole. With the aid of boric acid, more satisfactory selectivity (85% to 90%) at conversion levels of about 4% to 15% can be achieved (1; 16). The overriding reason of this improvement is the formation of cyclohexyl borate that is capable of protecting cyclohexyl group, thus, the formation of cyclohexyl radicals as intermediate can be avoided. However the advantages of this technology are to some extent offset by the necessity to synthesise boric acid, to treat cyclohexane-boric acid slurries and to recover boric acid through crystallisation, centrifugation, and dehydration (1). Nevertheless both methods (with and without boric acid) are commonly used in the current commercial productions. In the second stage, the KA mixture is converted to adipic acid by oxidation with nitric acid in the presence of copper and vanadium catalysts. As nitric acid is used as the oxidising agent, it is reduced to NO2, NO, N2O and N2. The manufacturing plant is only capable of capturing NO2 and NO, whilst N2O and N2 are considered as nitric acid loss as they can not be recovered within the system (5). This process is not only very rapid, but also highly corrosive and exothermic; thus heat removal and control of reaction are necessarily required. In order to accomplish the oxidation process, a high proportion of nitric acid must be maintained. The suitable ratio of nitric acid and KA mixture to complete

15

oxidation process is at about 40:1(1; 16). Adipic acid yields of 93% to 95% are typically achieved by this route (23).

Benzene +H2

OH
Phenol HBO2 air oxidation Cyclohexane Co, Cr, Mn air oxidation

Co, Ni, Cr Hydrogenation

OH +

Cyclohexanol Cyclohexanone (KA mixture)

Cyclohexanol

nitric acid oxidation

Co, V

Co, V

HOOC HOOC

Adipic Acid

Figure 2-1 Conventional route of commercial production of adipic acid (1; 16)

2.3 Phenol Route Phenol can be hydrogenated to either cyclohexanol or cyclohexanonecyclohexanol in the molten state, a reaction that usually takes place in the presence of nickel, copper or chromium catalyst at around 150C and 0.3 MPa (1; 16). The selection of catalyst and operating condition can affect the formation of cyclohexanol or cyclohexanone (16). Cyclohexanone is not commonly used to produce adipic acid as it leads to a low yield. It is widely used for the preparation of caprolactam, thus the hydrogenation process is designed

16

to produce a high selectivity to cyclohexanol. With proper control and catalyst, no less than 99% phenol can be converted to cyclohexanol for the selectivity range between 97% to 99%, whilst the small amount of unconverted product can be removed through distillation (16). 2.4 Environmental Concerns There are a number of environmental issues associated with the conventional two-step oxidation of cyclohexane. The issues are mainly associated with the use of boric acid or homogeneous Co, Cr or V metal salt catalysts and the HNO3 as oxidising agent, which can lead to both large amounts of atmospheric and liquid wastes. In the first step, oxidation process having inherent hazard since it is carried out in the relatively harsh condition (high temperature and pressure) and flammable and explosive potential of vapour mixture of cyclohexane with gas containing oxygen (1). This step commonly forms catalyst and organics waste from ketone-alcohol purification. Oily water, low pH streams containing adipic acid, boric, glutaric and succinic acid with copper, vanadium and sulphuric acid. Several special control techniques, therefore, may be required to handle these wastes (24): Ion exchange systems to remove organic salts such as copper or vanadium salts from organic catalyst Evaporation and crystallisation to recover boric acid and other byproducts High efficiency biological treatment to treat remaining wastes Optimised phase separation and extraction with incineration of organic phase to reduce organic loads These processes are typically consumes a large amount of energy because considerable amounts of solids need to be separated, decomposed and boric acid has to be recycled. Thus the operational costs of this process are commonly high (20). Moreover, the leaching of heavy metal (from homogeneous cobalt catalyst) is still frequently unavoidable (25) or, in case of using boric acid as

17

promoter, slurry contains mixture of cyclohexane boric-acid also becomes a great concerns (1). In the second stage, the KA mixture is catalytically (copper, vanadium salts) oxidised with nitric acid. This process is highly exothermic, and the presence of nitric acid solutions of organic acid, particularly with vanadium, may lead to extremely corrosive conditions (1). Several issues associated with this step are: The process releases considerable amount of nitrous oxide from the stripping columns and crystallisers which can be estimated to produce at about 0.3 kg N2O per kg of adipic acid (26). If N2O is re-used, it can be utilised either by burning at high temperatures in the presence of steam to manufacture adipic acid or using N2O to selectively oxidise benzene to phenol. If N2O is released to the atmosphere, either catalytic decomposition or thermal destruction is required as end-pipe techniques. Both heat from exothermic reaction and combustion process may be used to raise or produce steam [18]. Other airborne pollutants involved are adipic acid particulates generated from drying and handlings, organic material originating from feedstock, absorbers and purification columns on the KA section and fugitive-hydrocarbon emitted through normal venting in the hydrocarbon-containing tanks (1; 24). Strong odour is potentially produced from acid and handling storage since the process produces 10 to 20% by-product such caproic, adipic, valeric, butyric, acetic and propionic acid. Furthermore there are difficulties to separate and utilise the by-product obtained, and large amount of base typically required to neutralise the acid (20). Other liquid wastes are released from ketone-alcohol catalyst, plant cleaning, non-volatile organic residues and organic recovery tails from KA mixture production, boric acid sweepings, oxidiser residues, caustic wash residues, ketone alcohol sump dredging and wastes on shutdown such as tar-contaminates sand (24). Considering these impacts above, there are challenges for engineer and scientist currently is to find out more attractive and environmentally friendly routes that can address both economic and technical point of view.
18

CHAPTER 3- OXIDATION OF THE CYCLOHEXANE TO OL/ONE BY USING HOMOGENEOUS CATALYSTS

3.1 Introduction The selective oxidation of cyclohexane is a challenging problem due to its inert C-H bonds (27). The classical process uses relatively high temperatures and pressures to initiate the cleavage of C-H Bond. In the first step of the oxidation, the conversion of cyclohexane must be maintained in the range of 4% to 7% in order to prevent over oxidation and affords selectivity up to 90% (11; 16; 27). The concentration of a KA mixture of 0.4 mol/L and alcohol/ketone ratio of 60:40 is typically attained after 40 min (27). Besides its low conversion, the reaction commonly leads to undesirable features such as CrIII salts or boric acid and their problematic disposal (1). The greener routes of cyclohexane oxidation therefore is expected to find an ideal catalyst that is stable and capable of producing approach 100% conversion and selectivity toward Ol/One in relatively milder conditions. Scuchardts et al (27) and Cavani et al (3), in their reviews regarding the synthesis of adipic acid, pointed out that oxygen was the most preferred terminal oxidant to achieve an environmentally and economically sustainable process for the cyclohexane oxidation to Ol/One. This the case because of its lower price, higher yields and due to the fact that the reaction can be carried out in the free solvent reaction. These advantages may be offset with the harsh condition required to carry out the reaction. The oxidation under hydrogen peroxide is much less energy intensive than that of air or molecular oxygen. The reaction temperature can be as low as 60C, or even at room temperature, and pressures as low as 1 atm. This chapter will review greener oxidation of cyclohexane to cyclohexanol-cyclohexanone (Ol/One) using various homogeneous catalytic system such as various metal salts, metal complexes, and polyoxometallates in the presence of hydrogen peroxide as oxidant.

19

3.2 Iron and Manganese Based Catalysts 3.2.1 Non-heme Mononuclear Iron and Manganese Complexes

A review regarding of selective oxidation of hydrocarbons in biomimetic nonheme iron and manganese catalysts using hydrogen peroxide was published by Tanase and Bouwman in 2006 (28). Enzymatic monoxygenase is capable of oxidising several hydrocarbons and halocarbons (11; 29); biomimetic systems therefore have been used in attempts to perform the reaction under relatively mild conditions. Both non-heme and heme- (e.g. porphyrin group bearing) iron and manganese complexes are often to be used to stimulate the enzymatic action of organic substrates oxidation (28). Besides requiring low energy, these non-heme metalhydrogen peroxide systems have inherent environmental advantages since dioxygen atoms is split yielding one O atom that is added to a hydrocarbon molecule to produce a hydroxyl group, whilst the other atom O is used to produce water, a benign by-product (30). The substrate oxidation proceeds as follows (28) RH + H2O2 ROH + H2O (3.1) Manganese and iron (either mononuclear or binuclear) complexes are common metals used to conduct the reaction with hydrogen peroxide due to their accessibilities. Acetonitrile or acetone is required as an organic solvent in order to increase efficiency of hydrogen peroxide utilisation, but acetonitrile is frequently preferred because it facilitates higher efficiency and avoids hazards associated with the reaction between hydrogen peroxide and acetone, which can produce highly explosive acetone peroxide adduct as a by-product (32). The possible reaction patterns of cyclohexane oxidation with hydrogen peroxide by mononuclear iron complexes fall into two pathways. On one hand uncontrolled hydroxyl radicals as follows (eq. 3.2 and 3.3) (28) FeII + HOOH FeIII + HOOH FeIII + HOO + HOFeII + HOO + HO+ (3.2) (3.3)

20

On the other, the formation of high valent metal iron based oxidant can occur as presented in figure 3-1 below (28)
HO+ LFeV = O LFeII + H2O2 LFeIII - OOH RH LFeIV = O + HOo ROH

Figure 3-1 Oxidation mechanisms for hydrocarbon in the presence of iron complexhydrogen peroxide (28) The homolytic cleavages in the equations 2 and 3 are commonly recognised as the Fenton reaction (14). Various mechanisms have been observed, which depend on operating conditions and compete in different ways (33). The high valent LFeIV=O species, however, is the key feature of oxidation catalysed by non-heme metal complex-H2O2 since homolysis of the O=O bond may able to generate short lived HO that is rapidly reacted to the cyclohexane to form alcohol as predominant product (28). An A/K ratio > 1, therefore, indicates that short lived HO is formed by metal-based oxidant. Different kinds of ligands for the metal centre have to be prepared and designed to enhance the selective oxidation of alkenes. These ligands play an important role in controlling the reactivity of metal ions by binding one or two metal centres and donating electron to achieve higher oxidation states (28). In the case of iron and manganese complexes, the ancillary N-donor and O-donor ligands are able to stabilise high-valent manganese species in order to fasten CH bond activation through hydrogen abstraction (28). Current developments emphasise on modifying ligands properties or synthesising ligands that are stable under oxidation condition without suffering extensive degradation as well as capable of donating electron (28). Oxidation of cyclohexane using pyridine-based pentadentate ligands (N4py) has been reported by Feringa and co-workers (34). Using acetonitrile as the solvent, the selectivity of the complex [Fe(N4py)(CH3CN2](ClO4)2 gave a moderate conversion based on oxidant (31%) at 25C. Iron complexes and new
21

ligands based tetra dentate ligands with an amide donor inside the molecule, bis(2-pyridyl)methyl-2-pyridylcarboxamide (tpcaH) (35), gave a relatively low turnover and efficiency (12%). Better results were obtained by Chen and Que (36) using a [Fe(bpmen)(CH3CN)2](ClO4)2 [Bpmen = N,N-dimethyl-N,N-bis(2pyridylmethyl)-1,2-diaminoethane] catalyst. a conversion up to 63% with Ol/One ratio of 8. Fairly good results have also been reported by Fernandes et al (37) using iron complex [Fe (gma) (PBu3)] [where H2 (gma) = glyoxal-bis (2mercaptoanil)], bearing an {FeN2S2} centre. A yield of 11.7 % was obtained under mild conditions (25C), after 4 h in acetonitrile; with oxidant to cyclohexane ratio of 1.2. A co catalyst was used to carry out the reaction that performs under atmosphere of N2 and O2 (1 atm). The most impressive achievement of oxidation cyclohexene in the presence of mononuclear metal complexes was achieved recently by Nayak and co-workers (38). They used a coordination of transition metal iron (III) and a Schiff-base ligand, 2-methoxy-6-((quinolin-8-ylimino) methyl)phenol (mqmp). It can easily be synthesised in-situ from the reaction between o-vanillin and 8aminoquinoline as presented in Figure 3-2. The complex is capable of oxidising cyclohexane and leading to the corresponding alcohols and ketones in

NH2 OHC +
N

MeOH HO OCH3
N

N HO OCH3

(MQMPH)

Figure 3-2 In-situ preparation of Schiff-Base ligand (38) The structure of the catalyst is depicted in Figure 3-3. Three donor atoms of the ligand mqmp- (O, N, N) surrounding the iron (III) centre result in a rigid coordination. A total cyclohexane conversion and selectivity toward KA oil up to 91% can be attained in the presence of [Fe (mqmp) (CH3OH) Cl2] catalyst in acetonitrile at 50C after 24 h, with an oxidant to cyclohexane ratio of 1.5. A mechanistic pathway study using DMSO as hydroxyl-radical scavenger

22

indicated that highly active hydroxyl radicals were only involved in the least part of the reaction.

Figure 3-3 Molecular structures of complex [Fe (mqmp) (CH3OH) Cl2] (38)

3.2.2 Non-heme Binuclear Iron and Manganese Complexes Binuclear iron and manganese complexes are designed to mimic methane monoxygenase (MMO). Pioneering work on the oxidation reaction of various hydrocarbon catalysed by binuclear iron complexes was carried out by the researchers of the University of California in the early 1990s (39). The iron complexes [Fe2O(OAc)Cl2(bpy)2] and [Fe2O(OAc)(tmima)2](ClO4)3, (tmima = tris[(I-methylimidazol-2-yl)methyl]amine were tested to oxidise cyclohexane at ambient temperatures. However the reactions exhibited low activities and relativities with maximum turn over number (TON) of 15 and efficiencies as low as 8%. A better result was achieved by microwave assisted radiation in the presence of Fe complexes containing BMPA (bis-(2-pyridylmethyl) amine) and BMPA-derivative ligands (40). The experiment showed that the complexes of [Fe(BMPA)Cl3] was the most active systems, giving after 25 min at 160C under 100 W microwave irradiation, substrate conversions of 16.9% with the following product distribution: 28.9% cyclohexanol, 25.7% cyclohexanone and 34.5% adipic acid. Binuclear manganese complexes, on the other hand, have been extensively tested by Shulpin et al (41). The manganese(IV) complex catalyst with 1,4,7-trimethyl-1,4,7-triazacyclononane afforded a cyclohexane conversion of 46% at 25C after 2 h, but the efficiency was still relatively low (30%). More recently Shulpin (42) suggested the use of pyrazine-2, 3dicarboxylic acid (2, 3-PDCA) as co-catalyst to accelerate the reaction.
23

3.2.3 Non-heme Hexanuclear Iron Complexes

Trettenhahn et al (43) reported the catalysis performance of the hexanuclear iron p-nitro benzoate [Fe6O3 (OH) (p-NO2C6H4COO) 11(dmf)4] with an [Fe6 (3-O)
3((2-OH)] 11+

core. A fairly sophisticated yield of about 30% was obtained in

acetonitrile with temperature control (6.5, 15.4, 25.4 and 38.8C) after 1 h. Although this metal complex performs notable activities, an unfavourable feature of this system is a high ratio of HP-Cyclohexane (10 to 40 mmol) that is required to achieve such a yield. These operating conditions therefore seem to be rather cost prohibitive for commercial application. 3.2.4 Heme Iron Complexes

In 1987, Mansuy (44) reported that cytochrome P-450 can catalyse the oxidation of organic substrate through a monoxygenase reaction. The reaction is given in equation 3.4 as follows (44) R + 2 + 2+ + 2e
P -450

R + 2

In the presence of two electrons and two protons, one atom of dioxygen is inserted into an organic substrate, whilst another atom forms water as byproduct. As iron porphyrin is an active site of cytochrome P-450, the selective oxidation of cyclohexane catalysed by metalloporphyrin system has been a great interest for chemical scientist. The reaction catalysed by cytochrome P450 in the presence of hydrogen peroxide (metalloporphyrin/O2/H2O2 system), being analogue with biological system for hydrocarbon oxidation, can be called a shunt system because it utilises oxidising agent to donor an active oxygen species (45). Unlike non-heme metal complexes; molecular oxygen or t-butyl hydro peroxide (TBHP) are more often to be used as oxidising agent in this reaction. A number of publications reported unsatisfactory results of cyclohexane and hydrocarbon oxidation catalysed by metalloporphyrin-hydrogen peroxide systems. For instance, the catalysis performance of [Fe (TPP) Cl] (TPP=tetraphenylphorphyrin), R4PFeCl, and R4MnPFeCl (R= Complexes of
24

meso-tetrakis (3, 5-ditert-butyl-4-hydroxypenyl) porphyrin) that only gave a yield of Ol/One below 6% (46; 47). The reason of this poor performance has been delineated by Giorgio Strukul (14) and later reinvestigated by E.I Karasevich and Y.K Karasevich (45). They described the mechanistic pathway of the shunt system and demonstrated the dismutation of hydrogen peroxide in water and dioxygen by heme iron complexes that derive from the production of hydroxyl radical due to the homolytic cleavage of the peroxide O-O bound. The generation of HO then would result in mutagenic or lethal event to biological system, hence lead to iron porphyrin self destruction (14; 45). Two methods have been developed in attempt to enhance the catalysis stability (i) (ii) and prevent self oxidative destruction of the metalloporphyrin/O2/H2O2 system (14; 45; 48) Inserting substituent in the porphyrin ring in order to increasing the electrovity of iron porphyrin Absorbing ionic metalloporphyrin on an ion-exchange resin or using supported iron porphyrin in solid support (not encapsulated). The first approach has been developed by Moreira et al using tetrakis(2,3,5,6-tetrafluoro-N,N,N,-trimethyl-4-anilinium) ([FeTF4TMAPP]5+). Tetrakis porphyrin (2,3,4,5,6-pentafluorophenyl)porphyrin ([Fe(TFPP)]+) and Iron(III) mesoThey showed higher catalytic activities instead of iron

porphyrin that do not bear electron withdrawing substituent in the mesophenyl rings. The second approach will be discussed further in chapter 4.2.

3.2.5 Iron Salts The pioneering work of catalysis study upon oxidation of cyclohexane using hydrogen peroxide in the presence of iron salts (without added ligand) has been reported by Sawyer (49) in the mid 1990s and later reinvestigated by Shulpin et al (50) in 2004. The reaction took places at room temperature in acetonitrile using FeCl3 as catalyst, but afforded a very low conversion (3.2% based on H2O2) while efficiency is as low as 40% after 3 h reaction. More recently the
25

oxidation of cyclohexane using hydrogen peroxide in the presence iron salts, Fe (ClO4)2, at a relatively higher temperature (50C) has been studied by the researchers at Leiden University (33). The reaction performed at 50C in acetonitrile with a total conversion (100%) of cyclohexane and selectivity of 97% can be attained after 23 h. Interestingly the same conversion and selectivity can be achieved within 2 h when the oxidation was performed under argon atmosphere. The significant catalytic improvements seem to be attributed to the operating variations (temperature and Ar at atmospheric pressure). 3.3 Copper Based catalysts A number of different types of metal complexes have been widely studied as catalyst for the oxidation of cyclohexane under mild conditions. Copper is one interesting compound to be investigated due to its abundance present in nature (51; 52; 53). The summary of peroxidative oxidation of cyclohexane toward Ol/One is presented in Table 3-1. The biomimetic system is extensively investigated using copper as the metal centre to coordinate with a various type of ligands since copper is found in the active sites of a number of enzymes and plays substantial roles in the living system (51; 52; 53). For instance, an interesting oxidation properties of copper is reported by A.C Sylva et al using (51) BMPA (bis-(2-pyridilmethyl)amine) ligand incorporated Cu forms a mononuclear complex [Cu(BMPA)Cl2] and the four-centre complex {[Cu(BMPA)Cl2][Cu(BMPA)( H2O)Cl][Cu(BMPA)Cl][CuCl4]}. It has been found that complex {[Cu(BMPA)Cl2][Cu(BMPA)(H2O)Cl][Cu(BMPA)Cl][CuCl4]} exhibits a much higher activities and yield rather than complex [Cu(BMPA)Cl2]. It resulted in a yield up to 68% after 24 h at room temperature. The observation suggested that the more unit of copper in the complex, the higher yield can be obtained; it also raised the possibility of an important role of [CuCl4]-. The species [CuCl4]- is more easily reducible and oxidisable and thus capable of producing more active catalysis system. Detoni et al (54) reported oxidation of cyclohexane catalysed copper complex with a phenanthroline ligand. It was found that the higher Lewis
26

acidity results in higher catalysis performance. The catalytic activity of the 1,10phenantroline Cu(II) complexes followed the order: [Cu(phen)2Cl]Cl.5H2O > [Cu(phen)Cl2] [Cu(phen)3]Cl2. The optimum operating temperature is at about 70C, allowing a yield of 67% after 24 h under Ar atmosphere. Besides Ol/One, adipic acid and intermediate cyclohexyl peroxide were also formed in this reaction. It is clearly obvious that the temperature greatly affected the formation of the product since the cyclohexyl peroxide is decomposed rapidly at higher temperature and adipic acid can be obtained directly at the range of 50 to 80C. A marked growth of catalysis activity was also observed by addition of nitric acid in the medium. Kirillov et al (52) tested copper complex with various nuclearity (mono-, di-, tri-, tetra- and polynuclear copper) to oxidise cyclohexane. These reactions proceeded on acidic medium and formed biphasic liquid catalysis system at room temperature. The activity and stability of copper complex over polydentate triethanolaminate (H3tea) were investigated under various operating parameters such as amount of cyclohexane, reaction times, oxidants, solvents and nitric acid concentrations. The best result was obtained in the presence of [Cu2(H2tea)2{m-C6H4(COO)2-1,4}n] 2nH2O catalyst. A 37% yield of Ol/One can be attained after 6 h reaction in acetonitrile with HP/Cyclohexane ratio of 16. The catalyst did not suffer significant decrease of activity after it being reused 5 times. Its performances seem to be strongly dependent on the acidity of the medium, oxidant and solvent. It has been shown that the reactions practically do not proceed or only gave negligible yield in the absence or low content of nitric acid and solvent. The acidic medium to carry out the process may also play an important role in enhancing catalytic process through proton transfer step and preventing rapid decomposition of HP to water and oxygen. The choice of acetonitrile as a solvent offered a number of advantages including: (i) capability to solubilise cyclohexane and Ol/One (ii) stability under reaction condition (iii) similar boiling point with the reactant, thus facilitating an easy recirculation of cyclohexane and solvent mixture. A more environmentally friendly novel catalyst introduced by Silva et al (53) is a water soluble scorpionate complex. The hydro solubility of the catalyst may offer a possibility to replace any organic solvent as the oxidation of
27

cyclohexane can be carried out in the pure aqueous media. The catalyst can be synthesised from a commercially available metallic salt, CuCL2 and functionalized scorpionate 2,2,2-tris(1-pyrazolyl)ethanol and 2,2,2-tris(1pyrazolyl)ethyl methanesulfonate to form the corresponding water soluble Cu(II) complexes [CuCl2{HOCH2C(pz)3}] and [CuCl2{CH3SO2OCH2C(pz)3}]2. Although this catalyst is relatively less effective and gave low yield, this process possesses a strong green attribute as it also can proceed at room temperature. Table 3-1 Comparison of activities of copper catalyst for the cyclohexane oxidation to Ol/One
Cat. mol 0.8a 0.8a 0.8a 0.8a 0.8a 25 25 25 25 25 25 25 10 10 10 10 131 166 60 25 mmol HP: mmol ane 0.8/0.8a 0.8/0.8a 0.8/0.8a 0.8/0.8a 0.8/0.8a 10/0.63 10/0.63 10/0.63 10/0.63 10/0.63 10/0.63 10/0.63 10/0.63 10/0.63 10/0.63 10/0.63 24/8 24/8 24/8 10/0.63 Solvent (ml) MeCN (n.d) MeCN (n.d) MeCN (n.d) MeCN (n.d) MeCN (n.d) MeCN (5) MeCN (5) MeCN (5) MeCN (5) + MeCN (5) MeCN (5) MeCN (5) MeCN (3) H2O MeCN (3) H2O MeCN (10) MeCN (10) MeCN (10) MeCN (5) + T, t h Conv. (%) n.d n.d n.d n.d n.d n.d n.d n.d n.d n.d n.d 38.5 n.d n.d n.d n.d 10.5 58.6 73.1 6 Selectivity Ol/One (%) n.d n.d 79, AA 7 85, AA 6 93, AA 2 n.d n.d n.d n.d n.d n.d 100 n.d n.d n.d n.d 49.3 63.1 40.7 90 Yield Ol/One (%) 1.2,AA1.5 66.9, AA 2 61b 67b 60b 27.7 15.4 17 14.6 14.7 37.2 38.5 23.4 0.09 2 0.4 5.2 37 29.8 5.4

Catalyst [Cu(BMPA)Cl2] {[Cu(BMPA)Cl2][Cu(BMPA) (H2O)Cl][Cu(BMPA)Cl] [CuCl4]} [Cu(phen)3]Cl2 [Cu(phen)2Cl]Cl [Cu(phen)3]Cl2 [Cu2(H2tea)2 (N3)] [Cu2(H2tea)2(H4C6H4COO)2] 2H2O [Cu2(H2tea)2((CH3)4C6H4CO O)2 [Cu2H2tea)2 2(H 2O (Cl3C6H4COO)2]2H2O [Cu2(H2tea)2(XC6H4COO)2] 2H2O [OCu4(tea)4-(BOH)4][BF4]2 [Cu2(H2tea)2{mC6H4(COO)2-1,4}n] 2nH2O [CuCl2{HOCH2C(pz)3}] [CuCl2{HOCH2C(pz)3}] [CuCl2 {CH3SO2OCH2C(pz)3}]2 [CuCl2 {CH3SO2OCH2C(pz)3}] Cu3 (PO4)2 Cu2P2O7 Cu5 (P3O10)2 CuNO3
a in

oC,

Ref (51) (51) (54) (54) (54) (52) (52) (52) (52) (52) (52) (52) (53) (53) (53) (53) (55) (55) (55) (52)

RT, 24 RT, 24 70, 24 70, 24 70, 24 RT, 6 RT, 6 RT, 6 RT, 6 RT, 6 RT, 6 RT, 6 20, 6 20, 6 20, 6 20, 6 65, 12 65, 12 65, 12 RT, 6

Molar bYield include cyclohexylperoxide (CHHP)

28

Besides using copper complex systems, a number of researchers also developed interest to test the oxidation of cyclohexane catalysed by copper salts. Recently Y.Du et al (55) disclosed results of cyclohexane oxidations over three copper salts under various types of solvents, temperatures and organic acid initiators. The best result was obtained in the presence of copper pyrophosphate, Cu2P2O7, acetonitrile and temperature of 65C, with a conversion of 58.8% and selectivity toward KA up to 100% oil after 12 h. The modest hydrophobicity of the catalyst is believed to play an important role in this process since it is capable of enhancing desorption of polar product, cyclohexanone and cyclohexanol, from the active sites once they are formed. 3.4 Vanadium Based Catalysts Researchers of U.S Environmental Protection Agency (56) have tested catalysts containing vanadium ions to carry out the oxidation of cyclohexane under mild conditions. The calcined vanadium phosphorus oxide (VPO) catalyst, with a P/V ratio of 1.1, exhibited vanadyl pyrophosphate (VO)2P2O7 as the predominant phase and also low intensity peaks characteristic of VOPO4 phase. The reactions took place in homogeneous system in acetonitrile with the hydrogen peroxide to cyclohexane ratio of 4 and under nitrogen pressure. It afforded a conversion of 70% and 100% selectivity toward Ol/One after 4 h. When the reaction time was 8 h, the conversion of cyclohexane increased to 84%.

OH +

O [V5+] H2O2

[V4+]

H2 O

Figure 3-4 Oxidation pathway over VPO catalyst (56) The mechanistic pathway elucidates the advantage of using nitrogen pressure over oxygen air (oxygen). The appearance of V5+ sites derives from pyrophosphate frameworks that are stabilised by the excess phosphorus at the
29

surface. As depicted in Figure 3-4, the species V5+ plays a key role as a dynamic oxidising centre; however the presence of oxygen may hinder the reduction of V5+ species back to V4+. 3.5 Hetero- Metallic Complexes In 2006 Nesterov et al (57) introduced an innovative catalysis system using inorganic coordination compounds bearing Fe/Cu/Co metals and a Ni bearing Cu/Co core that can be prepared by self assembly. A 45% yield of Ol/One was achieved for a HP/Cyclohexane molar ratio of 3 or 5 after 6 h in the presence of [FeCuCo(L)3(NCS)2(MeOH)]2.3.H2O (H2L=diethanolamine) complex in acetonitrile, whereas complex [Ni(H2L)2][CuCo(H2L)(L)2(NCS)]2(A)2 {A =NCS or Br} failed to exhibit satisfactory results. It only produced a yield of 8% at the same operating conditions. The dissociation of the Ni Centre from Cu/Co core seems to be the overriding reason of this poor performance. This experiment not only pointed towards a fundamental role of iron and the synergistic effect of Fe/Cu/Co to enhance the catalytic activities. It was also found that the greater HP/Cyclohexane molar ratio, the higher yield and faster reaction time can be achieved; but that ratio below 5 would lead to over oxidation and reduce the yield of product. 3.5 Polyoxometallates (POMs)

Polyoxometallates (POMs) is incorporation of metal-oxygen compound which forms a unique cluster (58). The crystalline structures of POMs made up of a polyhedral cage structure or framework anions, which is balanced by positive charges that are external to the cage (58). Being analog of metalloporphyrins, POMs have a rigid coordination sites surrounding the metal centre and have the common formula (XM12O40)3- (11; 27). The molecular compounds as the structural unit include isopoly acids, heteropoly acid and their salts. The advantages of using POMs in oxidation reaction are the stability under reaction conditions and the possibilities to substitute metal in peripheral centre without losing the structural integrity. Keggin type polyoxometalates is the most studied
30

the alternative system to oxidise cyclohexane. For example, the use of dimanganese and diiron substituted polyoxometallates to facilitate the oxidation of cyclohexane using hydrogen peroxide at 83C has been reported by Mizuno et al (59). Under these conditions, a selectivity toward Ol/One up to 100% with an HP efficiency of 95% was obtained in the presence of diiron substituted [-SiW10{Fe(OH2)}2O38]6-. Table 3-2 Cyclohexane oxidation catalysed by Polyoxometalates
Catalyst [-SiW10{Fe(OH2)}2O38]6[-SiW11Fe(OH2)O39]5[-SiW9{Fe(OH2)}3O37]7 [-SiW10Mn2O38]6[-SiW12O40]4Co4(PW9)2 Mn4(PW9)2 Fe4(PW9)2 Fe4(PW9)2 (TBA)4 H3 PW11 O39 (TBA)4 HPW11 CuO39 (TBA)4 HPW11 Co(H2O)O39.2H2O (TBA)4 HPW11 Mn(H2O)O39.3H2O (TBA)4 HPW11 Ni(H2O)O39.H2O (TBA)4 HPW11 Fe(H2O)O39.2H2O TBA4H2BFe(H2O)W11O39 H2O TBA4H2BFe(H2O)W11O39 H2O TBA4H2BFe(H2O)W11O39 H2O
a

Cat. mol 8 8 8 1.5 1.5 1.5 1.5 1.5 1.5 39 40 38 38 39 42 1.5 1.5 1.5

mol ane: mol HP 1/1 1/1 1/1 1/1 1/1 4/1 4/1 4/1 4/2 19/39 19/29. 19/39 19/39 9/39 19/39 2/1 4/1 4/1

Solv. (ml) MeCN (6) MeCN (6) MeCN (6) MeCN (6) MeCN (6) MeCN(1.5) MeCN (1.5) MeCN (1.5) MeCN (1.5) MeCN (10) MeCN (10) MeCN (10) MeCN 10) MeCN (10) MeCN (10) MeCN (1.5) MeCN (1.5) MeCN (1.5)

T, t ,h

Conv. Ol/One (%) 66 7 5 <1 <1 91 98 98 100 35 11 5 8 20 76 87 99 98

Select. Ol/One (%) 100 100 100 100 100 100 100 48 30 93a 87a 100 100 100 25a 68a 47a 43a

Yield Ol/one (%) 66 7 5 <1 <1 91 98 47 30 32.6 9.6 5 8 20 19 59.2 46.5 42.1

Ref (59) (59) (59) (59) (59) (60; 61) (60; 61) (60; 61) (60; 61) (62) (62) (62) (62) (62) (62) (63) (63) (63)

83, 96 83, 96 83, 96 83, 96 83, 96 40, 12 40, 12 40, 6 40, 12 80, 12 80, 12 80, 12 80, 12 80, 12 80, 9 80, 12 80, 6 80, 12

the rest of product is cyclohexyl hydro peroxide (CHHP)

The high selectivity towards product is believed to be due to the noninvolvement of hydroxyl radical pathway during the reaction. Dimanganesesubstituted [-SiW10Mn2O38]6- and non-substituted Keggin anion -SiW12O404-, in contrast, showed no significant activity. Even though this system is carried out with a HP/Cyclohexane molar ratio equal to 1 and showed no significant

31

decrease of activity in the second running, however the long duration of reaction (96 h) is likely to be a major drawback of this system. Cyclohexane oxidation catalysed by Keggin type polyoxometallateshydrogen peroxide systems (60; 61) was reported with catalytic performance of tetrabutylammonium (TBA) salts of Keggin-type polyoxotungstates [PW11O39]7- and [PW11M(L)O39]7- (where Mm+ is a first row transition metal including Fe, Cu, Mn, or Ni and L is H2O or CH3CN). Conversions in order of 5% to 20% were observed when Cu, Mn or Ni was used, but interestingly only cyclohexane and cyclohexanol were produced as the final products. When the reaction was carried out in the absence of transition metals, the heteroplyanion PW11 showed a higher conversion but it also produce CHHP as by-product. CHHP may decompose to cyclohexanol and cyclohexanone but carboxylic acid was also generated as by-product at the end of the process. Using iron, a quite high conversion was obtained, but resulted in a significant amount of cyclohexyl hydro peroxide (CHHP); accounting for 75% of total product. The longer reaction time would result in significant higher conversion, but led to the formation of a complex mixture with a lower selectivity. The efficiency of H202 for iron substituted-POM, however, is as high as 100%. Later it was disclosed that iron metal substituted Keggin-type tungstoborates TBA4H2BFe (H2O) W11O39H2O (62) and HP/Cyclohexane ratio of 2mmol/1mmol, a conversion of 87% could be obtained, with an Ol/One product accounting for 68% and efficiency of H2O2 being more than 98%. When the ratio of HP/Cyclohexane was increased to 4 /1mmol, a much higher conversion was observed (99%), which led to a substantial decline in the H2O2 efficiency. More recently sandwich-type Mn and tungstophosphates B-[M4(H2O)2(PW9O34)2]10, MII = Co, [FeIII4(H2O)2(PW9O34)2]6

(abbreviated as M4(PW9)2) were shown to catalyse the same reaction (63). The manganese substituted (PW9)2 is capable of converting the cyclohexane up to 98% and exhibited 100% selectivity toward Ol/One after 12 h at relatively mild conditions. Fe substituted (PW9)2, on the other hand, also exhibited high selectivity with conversion up to 100% and turnover number more than 1300 (mmol oxidised product per mmol catalyst) after 12 h; however, the cyclohexyl peroxide was formed and accounts for 70% of total product. Catalysis stability
32

test using infrared spectra, unfortunately, indicated that the residue that the catalysts Mn, Co and Fe substituted POMs suffering extensive degradation. A summary of the oxidation of cyclohexane to Ol/One catalysed by POMs is presented in Table 3-2.

3.6 Hybrid System Metal Complexes Polyoxometallates Recently Mirkhani et al (64) disclosed an interesting novel strategy for the preparation of metalloenzymes by covalent linkage to a Keggin type polyoxometallates (POM).K8SiW11O39. They incorporated Copper (II) salen (where H2salen=N, NO-bis (salicylidene) ethylenediamine) with K8SiW11O39. Cu salen-POM proved to enhance the reaction, affording the cyclohexanone as the predominant product (87%) and a low proportion of cyclohexanol (13%). At 80C, with HP/Substrate ratio of 6, high conversion and turnover frequency were obtained, reaching 45% and 2.35, after 10 h reaction. This result is higher compared to the corresponding unhybridised Cu salen. The attachment of POM is believed able to avoid the formation of the -oxo bridged complex that may lead to decreased activity of Cu salen. 3.7 An Overview of Homogeneous Catalysis Development 3.7.1 Summary of the Recent Research Progress

A number of various homogeneous catalysts for peroxidative oxidation of the cyclohexane to Ol/One other than Co complexes have been reported in the recent literature. A summary is presented in Table 3-3. The homogeneous system based on heme iron complexes of H2O2 was found to be ineffective to carry out the oxidation of cyclohexane. The self-destruction of the catalysts within the operating conditions due to homolytic cleavage of the peroxide O-O bond and low ability to produce satisfactory yield of Ol/One make this system uninteresting for further investigation; whilst the rapid developments of various ligands in the non-heme iron and copper complexes lead to significant progress of peroxidative oxidation of cyclohexane. Schiff base ligand is one of

33

the impressive works that demonstrated the 100% conversion of cyclohexane and that selectivity more than 90% can be achieved, but the time of reaction (24 h) is too long. Acidic medium proved to be enhancing the catalysis reaction by copper complexes with triethanolamine (H3tea). These copper catalysts can be prepared self assembly and produce reproducible results, but the use of large amount of nitric acid in order to form acidic medium may not represent a step forward for a greener process. A marked growth of catalysis process was also observed in heterotrimetallic Fe/Cu/Co complexes, but the efficiency of H2O2 is as low as 9%. The use of metal salts of copper, vanadium or iron salts may provide an attractive alternative as those can be used without added ligand and showed fairly good activities. However some of them require inert atmosphere. Compared with biomimetic metal complex catalyst; POMs, on the other hand, exhibits more promising performance and offered a number of intrinsic advantages such as higher thermal stabilities and facilitation of rapid and reversible multi-electron redox transformation under operating conditions. These may lead to very efficient oxidant utilisation and substrate conversion (66). Unfortunately the catalysts are often unstable, suffering decomposition under operating conditions. 3.7.2 Future Prospect in Homogeneous Catalysis Development

Impressive results have been achieved, but those cutting-edge experiments can be reviewed critically from economical, technical and environmental point of view. Based on economical and technical consideration, it is generally found that the catalysis development recently only focuses on activity, selectivity and stability of the catalysts, but gives no adequate elucidation in terms of the efficiency of hydrogen peroxide utilisation. As amplified by Scuchardt et al (27), the relative prices of hydrogen peroxide (US$ 0.58/kg) compared to adipic acid (US$ 1.34-1.5/kg) (65) is critical, and a poor efficiency of the peroxidative oxidation (below 40%) would attract higher cost of the oxidant. Unfortunately, it is quite often the case a that high ratio of cyclohexane-hydrogen peroxide (2 to 6) is frequently necessarily required to achieve higher yield of Ol/One;
34

efficiency of HP utilisation is frequently not determined (or yield based on oxidant). Furthermore, the limitation of this catalyst in many cases derives from the fact that the catalyst suffers extensive degradation under reaction conditions. From environmental perspective, it is sometimes apparent that the catalysis reactions under experiments seems to be environmentally friendly, the material and energy used to synthesis and prepare the catalyst often involve unsustainable processes. Another drawback is the requirement of solvent to carry out the reaction. Since contrast polarity between cyclohexane and hydrogen peroxide, at which cyclohexane is highly hydrophobic and hydrogen peroxide is strong hydrophilic (55), a solvent therefore is essentially needed in order to facilitate homogeneity for those reactants. Unfortunately traditional volatile organic solvent with high vapour pressure and toxicity were quite often used in the reported literature. Acetonitrile, for example, possesses high toxicity to the environment due to its aquatic persistence and potentially leads to bioaccumulations. Moreover since acetonitrile is a common compound that widely used in chemical industry, the high demand and tight supply of this chemical may result in higher price in the future (67). A number of substantial features therefore may need to be considered and improved to make the development of homogeneous catalyst more economically and environmentally sustainable in the future. First factor is performing reaction in green or free solvent conditions. A various alternative reaction medium may consist of room temperature ionic liquids, supercritical carbon dioxide, or solvent free conditions (68). The use of supercritical CO2 may offer advantages in terms of the possibility of recycling catalyst and minimising the loss of catalyst. This is particularly important when high cost metal and ligands are used. The presence of supercritical carbon dioxide may also lead to higher mass transfer rate and the improvement of selectivity of peroxidative oxidation of cyclohexane has been reported by Olsen et al. However since this process is also energy intensive and hindered with high price of CO2 - the commercial application may come into reality if yields obtained under this reaction are much superior to those obtained under conventional liquid-phase condition. The same benefits may possibly be obtained in the presence of ionic liquids. Since ionic liquids have no vapour pressure and can be generated by
35

combining a wide range of anions and cations; they could replace the conventional organic solvents. But it is also noteworthy than the manufacturing of greener solvents such as ionic liquid should not involve a metathesis step that may decline their green attribute to some extent (69). The toxicity, thermal stability, biodegradability properties of a solvent also should be fully determined. The latter alternative has investigated by Sylva et al (53). With a hydro soluble complex catalyst, the oxidation can be carried out in pure aqueous media, but the yield is quite low. The further challenge therefore is to find a novel water-soluble catalyst that capable of producing more competitive yield. A second feature is minimising cost, complexity and toxicity of reactant. There are a number of drawbacks that are restraining the application of the oxidation of cyclohexane to Ol/One catalysed by metal complex-HP system. Those limitations include most of metal complexes are often very expensive, frequently involved complicated preparation and only few data available regarding the potential toxicity of the ligands to the environment. For instance, in case of a ligand synthesise for metalloporphyrin or non heme complex catalyst. It is generally found that synthesis of more robust ligands can be achieved by halogenations (11) that may use considerable amount of chlorinated compounds or acidic solutions, thus producing a catalyst that is not more environmentally sustainable compared with those of simple metal salt catalysts. The simple method of a ligand synthesis such as in-situ preparation of Schiff base complex as demonstrated by Nayak and co-workers (38) or other easy and simple preparation seems to be more favourable for future developments. In addition as homogeneous catalysts are soluble in the reaction mixture, from an environmental perspective, the toxicity level of the metals should be taken into account. For example, the high toxicity of vanadium salts may make it unacceptable to develop homogeneous system catalyst. Third, designing suitable reactors for catalyst recovery has yet been sufficiently addressed. If costly catalyst is inevitable, designing appropriate reactors and processes to recover both metals and ligands is likely to be a future challenge (70). One of the methods of separation technologies is heterogenised

36

homogenous catalyst that is discussed in the chapter 5. A general scheme to improve the reusability for homogeneous catalysis is presented in Figure 3-5. Fourth, investigating the mechanistic pathway of catalysis reaction is required. Although outstanding progress of cyclohexane oxidation has been achieved in the recent decade, however the experimental data regarding the mechanistic pathway of catalysis reaction are still limited. In many cases, particularly for biomimetic system catalyst, the materials are often unstable and suffering extensive degradation under reaction conditions. The further mechanistic studies to distinguish intra- from intermolecular mechanisms therefore seem to be substantial to achieve better understanding of catalysis process (28). This is particularly important to elucidate how exactly the underpinning performance.
Homogeneous Catalysts

catalysis

works, are,

what and

the how

reasons to

of

the the

catalysis catalysis

decompositions/deactivation

improve

Continuous recycle

Recycle after Catalyst lost in work-up product

On heterogeneous support

With liquid supports

Ionic liquid

Supercritical CO2

water

Figure 3-5 A general scheme to improve the reusability of homogeneous catalysts (3; 67)

Last, improving the efficiency of H2O2 utilisation is of paramount of importance. With a low ratio of HP/Cyclohexane, the efficiency of H2O2 utilisation is relatively higher, but the yield of Ol/One is commonly not satisfactory. Increasing the ratio of HP/Cyclohexane proved to be able to results
37

in higher conversion, but it can be problematic. It frequently decreases the efficiency of H2O2 utilisation (63) and may lead to unselective oxidation due to over oxidation (57), and more importantly, it is obviously constrained by cost of the HP. Optimisation of the reaction to find an appropriate ratio of HP/Cyclohexane or method maximise oxidant efficiency seems to be essential. Furthermore it is also substantial the catalysis performance in terms of both yield based on substrate and efficiency of HP (sometimes called yield based on oxidant) need to be considered more.

38

Table 3-3 Comparison of catalysis performance with cobalt naphtalene versus various homogeneous catalysts
Catalytic property
Temp.C

Pressure (atm) Time Conv. ane (%) Select. Ol/One (%) Yield Ol/One(%) Ol:One HP/ane HP eff. (%) TONb Solvent Ligand Co-catalyst/ additives
a

Co or Mn Saltsa (1; 3; 27) 150-180 10-20 40 mins 5-7 75-80 3.8 5.6 60:40 -

FeP (46) RT 1 atm (Ar) 2h n.d n.d 5.53 3.5:2 1 n.d 47 MeCN -

Fe Complex (38) 50 1 atm 24 h 100 90.5 90.5 39.5 : 51 1.5 n.d n.d MeCN Schiff-Base -

Cu complex (52) RT 1 atm 6 ho 38.5 100 38.5 23.4 : 14.2 16 n.d 380 MeCN Triethanolamine HNO3

V salt (56) 65 1 atm (N2) 4h 70 100 70 37 : 63 4 n.d 308 MeCN -

Fe Salt (33) 50 1 atm (Ar) 2h 87 100 87 45 : 42 1.5 n.d n.d MeCN -

Cu Salt (55) 65 1 atm 12 h 58.6 100 37 37 :63 3 n.d n.d MeCN -

Fe/Cu/Co (57) RT 1 atm 6h n.d n.d 45 40:3 8.9% 47 MeCN Diethanolamine HNO3

POMs (63) 40 1 atm 6h 98 48 47 14 : 34 2 72% 657 MeCN -

Hybrid Cu SalenPOMs (64) 80 1 atm 10 h 47 100 47 13 : 87 6 n.d 2.35l MeCN Schiff-Base -

Currently used in commercial application with oxygen or air as terminal oxidant over number (mmol oxidised product per mmol catalyst) cTurn over frequency (mmol oxidised product per mmol catalyst per hour)
bTurn

39

CHAPTER 4 OXIDATION OF CYCLOHEXANE TO OL/ONE BY USING HETEROGENEOUS CATALYSTS

4.1 Molecular Sieve Catalysts

The difficulties to separate the catalyst from the mixture KA oil may be eliminated if the catalyst can be in a heterogeneous system. The development of heterogeneous catalysis system therefore may eliminate the cons of homogeneous system, which, when used on conventionally, results in a process with unavoidable leaching of metal salts or acid slurry during the oxidation of cyclohexane to Ol/One. Moreover, continuous processing is easily implemented with a heterogeneous system since the catalyst may be reusable and give reproducible performance (11; 70). The use of molecular sieves catalysts may offer distinct advantages due to their exchangeability, reactant or product selectivity, thermal stability and reusability (70). The term molecular sieve refers to a material containing a crystalline structure and a range of compositions which are capable of performing shapeselective adsorption and reaction properties (71) The development of heterogeneous catalysis of this cyclohexane oxidation in the last two decades predominantly using active transition metals such as Ti, V, Co, Cr, and Fe that incorporated in a variety of molecular sieves structures (e.g. APO, MCM-41, HMA, SBA etc) by hydrothermal synthesis. Few researchers also reported the catalytic performance of transition metal oxide such as V2O5TiO2 prepared via the sol-gel method. A summary of literature reporting on the heterogeneous oxidation of cyclohexane with hydrogen peroxide is given in Table 4-1.

40

4.1.1 Titanium Based catalysts Since the discovery of a microporous solid silicate structure materials, especially those containing SiO2 and TiO2 modified by isomorphous substitution of Si (IV) with Ti (IV), TS-1; extensive studies have been attempted to investigate TS-1 in various catalytic reactions (72). TS-1 can be prepared via hydrothermal crystallisation of a silicontitanium gel containing tetrapropylammonium salt (73). The reaction of cyclohexane in the presence of TS-1 in acetone showed stable activity, resulting in cyclohexane conversion of 6.3% and 64.9% selectivity toward KA oil after 5 h (74). In the other experiments using acetic acid as a solvent, TS-1 showed higher activities, leading to a conversion of cyclohexane and selectivity toward Ol/One of 16 and 79%, respectively, after 4 h reaction. In this case, the author pointed out the ability of acetic acid to facilitate the complexation of active sites within the titanium structure to enhance catalytic performance (73). However, some titanium leaching was observed during the experiment. A number of Ti-based catalysts were also tested by several researchers. A mixed phase material made up of both micro porous and meso porous Si/Ti materials (Ti-MMM-1) exhibited fairly good cyclohexane oxidation with acetone as a solvent (75). The most excellent result of heterogeneous Ti-Based system, however, was disclosed by Selvam and Mohapatra (76) using meso porous TiMCM-41 and Ti-HMA. Ti substituted hexagonal mesoporous aluminophosphates (HMA) with a narrow pore distribution and meso pores diameter of 2.9 nm. It exhibited a conversion and selectivity more than 90% at 100C after 12 h in the presence of acetic acid as a solvent and MEK initiator. At the same operating conditions, Ti-MCM-41 showed activities only slightly lower than that Ti-HMA.

4.1.2 Vanadium Based Catalysts Selvam and Dapurkar (77) investigated the synthesis and catalytic activity of various vanadium sources. The hydrothermal synthesis using different Si/V involving incipient wetness and calcination led to pore diameter of VMCM-1 in a range of 2.4 2.7 nm. The activity of these systems is greatly affected by the
41

sources of vanadium, which determines the extent of vanadium incorporation in the mesoporous matrix and thus catalytic performance. It has been found that tetravalent vanadium is preferred instead of pentavalent vanadium since it forms monomeric VO43- and exhibits maximum vanadium incorporation in the mesoporous matrix MCM-41. It afforded a conversion and selectivity to Ol/One more than 95% after 12 h in acetic acid. The catalytic reusability test showed that V-MCM-1 were more stable than VS-1 and V205/MCM-41, which suffer leaching of active vanadium ions from the matrix. With the same range of pore size diameter (2.5-2.7 nm), a vanadium framework substituted HMA (V-HMA) was performing catalysis activities as high and stable as V-MCM-41 (78). The presence of both isolated V4+/V5+ ions in tetrahedral framework sites may be attributed to its stable activity upon cycling. One of reason behind the decomposition of the catalyst is caused by the presence of water as by-product and accelerates by carboxylic acid. Another developments of vanadium substituted molecular sieve was reported by Jermy et al (79) using three dimensional (3-D) cubic V-KIT-6. The catalyst was successfully synthesised with a surface area at about 1000 m2/g and narrow size distribution of pore diameter (5-6 nm). It has been demonstrated that the higher vanadium content enhances the concentration of weak Bronsted acid sites due to the presence of tetrahedral vanadium. The activity of cyclohexane oxidation increased with the vanadium content due to higher acidity, however the excess vanadium content implies on the presence V2O5 species at the external surface that may decrease the efficiency of the oxidant. The optimum ratio of Si/V was found to be 49 (1.89% wt vanadium), which affords a cyclohexane conversion of 83.6% with Ol/One selectivity of 93.8% at 100C in acetonitrile and MEK initiator after 9 h. Moderate yields were obtained in the presence of incorporation of two transition metal ions in the APO framework, (Co-V) APO-5 catalyst (80). The cyclohexane oxidation was very selective, but the conversion was rather low. Furthermore, although the catalyst (Cr-V) APO-5 lost approximately 25% wt after five catalytic cycle, it did not suffer from significant decrease of catalytic performance. This fact indicates that leached metal ions may play a fundamental role in retaining the catalyst activity.
42

Transition metal oxides also attract great interest and are the subject of intensive research. For instance, a mixed oxide, V2O5TiO2, has been reported by Choukchou-Braham and his co-workers (81). The catalyst was successfully prepared and synthesised by an acid-catalysed sol-gel process, which affords materials containing three morphology types including: small agglomerated grains of about 20 nm, smooth middle-seized grains of about 85 nm and smooth large grains of about 300 nm diameters. Acetonitrile, acetic acid and methanol were tested as solvents to carry out the oxidation of cyclohexane with hydrogen peroxide. Best results were obtained with acetic acid as a solvent and acetone as initiator presented with a conversion of 8% and selectivity of 76% towards cyclohexanol after 8 h at 70C.

4.1.3 Cobalt, Iron and Chromium Based Catalysts A number of authors reported catalytic activities of cyclohexane oxidation using heterogeneous cobalt based catalyst. Selvam and Mohapatra (82) compared the catalytic properties of cobalt framework-substituted HMA, cobalt containing mesoporous silicate (Co-MCM-41), microporous cobalt aluminophosphate (CoAPO-5), and cobalt silicate (Co/S-1). Using a Co-HMA as catalyst, they obtained a conversion of 92.4% with the selectivity toward Ol/One of 96.5% after 12 h. Under the same conditions, Co MCM-41, Co APO-5, and Co S-1 showed activities lower than that Co-HMA. It was found that hydrothermally synthesised Co-HMA proved to be a stable heterogeneous catalyst since only a negligible amount of leaching of cobalt was observed. The catalytic performance of cobalt doped mesoporous SBA type silica materials, Co-SBA 15 and Co-SBA 3 also have been studied recently (83). The larger pore diameter of Co-SBA-15 (4.5nm) compared to that those of Co-SBA15 (3.6 nm) and Co-MCM-41 (2.8 nm) is believed to play a key role in enhancing selectivity toward desired product (Ol/One). Co- SBA 3 gave a nearly 100% selectivity to Ol/One at about 91% conversion, in 8 h reaction time at 100C. Moreover, when the filtered Co-SBA-3 was reused, its activity was not significantly diminished.

43

The incorporation of cobalt into mesoporous crystalline structured titanium, followed by cogellation method using 1-dodecylamine as template, led to formation of a Co-MTiO2 catalyst with average pore diameter of 10.8 nm. It afforded a 93.5% yield of Ol/One and turnover number of 136, using acetic acid as a solvent (84). The recycling of the catalyst in successive experiments gave reproducible results. A lower activity, on the other hand, was observed in the Co/TiO2 catalyst due to dispersion of Co on the surface of TiO2 instead of the inside pore channel of TiO2 by impregnation. The investigation of catalytic properties of iron metal substituted molecular sieve is somewhat difficult due to the instability of iron species in the matrix during heat treatment. Selvam and Mohapatra (85) therefore used a number of different characterisation methods to investigate the stability of trivalent iron in the hexagonal frameworks. It has demonstrated that the dislodgement of trivalent iron from the framework structure during calcination lead to high thermal stability of Fe-HMA catalyst. A yield of 84.1% with turn over number of 756 can be obtained after 12 h in acetonitrile. Since the catalyst retained its structured and porosity after recycling, the repeated running exhibited reproducible results and no significant leaching of active iron species was observed during catalytic stability test (85). Satisfactory results were also achieved using chromium doped HMA and MCM-41 (86). The ability of mesoporous silicates and aluminophosphates to stabilise chromium in the framework led to stable catalyst that do not suffer leaching under reaction conditions. The calcined Cr HMA possesses an average pore size of 2.8 nm and showed an excellent selectivity to Ol/One (100%) after 12 h reactions at 100C in acetic acid. 4.1.4 Cerium Based Catalysts

Unlike transition metals, the investigation of catalytic properties of rare earth element has been rather limited. One example was disclosed by researchers at Yunnan University (87). Cerium-doped MCM-41 catalyst showed a conversion of cyclohexane up to 94.6% and predominantly produced cyclohexanol at 100C. At reaction temperature of 50C, it was found that cyclohexanone was the major
44

product. More interestingly, the catalyst was recyclable since it showed stable performance during repeated running. Table 4-1 Summary of heterogeneous catalysis for the cyclohexane oxidation to Ol/One
Cat. Catalyst Na-GeX Sulphated GeX Ce-MCM-41 Cr-HMA Cr-MCM-41 TS-1 TS-1 Ti-MMM-41 Ti-MCM-41 Ti- HMA V2O5TiO2 Co-HMA Co-MCM-41 Co-APO-5 Co-S-1 Co-SBA-3 Co-SBA-15 Co-MTiO2 Co-TiO2 Co-MCM-41 Fe-HMA VMCM-41 V205-MCM-41 VS-1 V-HMA V-KIT-6 (Cr-V)APO-5 (g) 0.1 0.1 0.4 0.05 0.05 0.1 0.27 0.1 0.05 0.05 0.03 0.05 0.05 0.05 0.05 0.1 0.1 0.15 0.15 0.1 0.05 0.05 0.05 0.05 0.1 0.1 0.1 mol HP/ mol ane 2 2 2 1 1 1.3 3 3 1 1 0.27 1 1 1 1 2.2 2.2 2 2 2 1 1 1 1 1 4 1 without without CH3COOH CH3COOH CH3COOH (CH3)2CO CH3COOH (CH3)2CO CH3COOH (+MEK) CH3COOH (+MEK) CH3COOH CH3COOH (+MEK) CH3COOH (+MEK) CH3COOH (+MEK) CH3COOH (+MEK) CH3COOH CH3COOH CH3COOH CH3COOH CH3COOH CH3COOH CH3COOH CH3COOH CH3COOH CH3COOH (+MEK) CH3COOH (+MEK) (CH3)2CO Solv. T, t
oC,

Conv. ane(%) 66.5 42 94.6 93.5 95.6 6.3 16 9.2 90 88 7 92.4 58.2 33.5 28 91.6 42.8 100 43.7 87.4 84 99 97.3 36.8 95 83.6 10

Select. to Ol/One (%) 100 100 89 100 97.1 64.9 79 89.8 95 99 76 96.5 87 83.7 85.3 99.9 91.8 93.5 85.2 88.6 93 96.4 93.9 97.6 94.4 93.8 90

Yield Ol/One (%) 66.5 42 84.1 93.5 92.8 4.1 12.6 8.2 85.5 87.1 5.3 89.1 50.6 28 23.9 91.5 39.3 93.5 37.2 77.4 78.1 95.4 91.4 35.9 89.7 78.4 9 (88) (88) (87) (86) (86) (74) (73) (75) (76) (76) (81) (82) (82) (82) (82) (83) (83) (84) (84) (83) (85) (77) (77) (77) (78) (79) (80) Ref

70, 3 70, 1.5 100, 12 100, 12 100, 12 77, 5 60, 4 95, 8 100, 12 100, 12 70, 8 100, 12 100, 12 100, 12 100, 12 100, 8 100, 8 100, 12 100, 12 100, 12 100, 12 100, 12 100, 12 100, 12 100, 12 100, 9 60, 40

(CH3)2CO CH3COOH (+MEK)

4.1.5 Germanic Faujasite

A research of catalytic activities of a germanic near faujasite and its sulphated form molecular sieve has been reported by Prvulescu et al (88). The Na-GeX catalysts was synthesised from Al-Ge gels at low temperature, whilst sulphated
45

Ge-X can be prepared by treating Na-GeX with NH4SO4 and then the sample heated gradually up to 400C. Impressively, The Na-GeX systems exhibited an excellent catalytic activity. A cyclohexane conversion of 66% and 100% selectivity toward KA oil can be attained after 3 h at temperature of 70C under autogenic pressure and free solvent reaction. The H2O2 efficiency was also fairly good (58.7%); whilst the sulphatation of Na-GeX able to enhance the catalytic activity, but the H2O2 efficiency decreased to 25%. The successive experiments showed that catalyst did not suffer significant decrease of activity. 4.2 Carbon Nitride Polymer

Some work has recently focussed on carbon nanostructures due to its wide range of application. An example is bulk CxNy materials. Wang et al (89) reported a novel metal free-catalyst using boron- and fluorine enriched mesoporous polymeric carbon nitride (CNBF) that can be environmentally synthesised in the presence of an ionic liquid. This catalyst preparation potentially replaces the conventional method of removing silica structures that typically involving metathesis step using aqueous hazardous ammonium bifluoride (NH4HF2) or hydrogen fluoride (HF). Impressively, this polymer proved to be a heterogeneous catalyst since it can be separated easily by simple filtration and gave reproducible result without loss of significant activity when repeated catalyst was used. The oxidation is highly selective; with a 100% of Ol/One can be attained after 4 h in acetonitrile, but conversion is as low as 5.7%. 4.3 Heterogenising Metal Complexes into Solid Materials

Latest catalysis developments that have been recently investigated are immobilisation of homogenous catalysts into supported solid material or inorganic matrixes in order to improve the recoverability and reusability of the catalysts and to produce more stable active site species. A summary of research development in this field is presented in Table 4-2.

46

The encapsulation of metal complexes in solid support is, however, often leads to unsatisfactory results. For example, the cyclohexane and hydrocarbon oxidation catalysed by iron porphyrin encapsulated in zeolite [FeIII(TPP)Y] or in silica matrix (FePES) showed poor performance and produced a yield of Ol/One below 1% (46; 48)[24c, 49]. Moreira et al [24c] and Olsen et al [49] then suggested the problem associated with the encapsulation of iron heme complex as follows: The oxygen rebound mechanism required for oxidation of the inert cyclohexane may be impeded with the polar environment of the FeP active site The encapsulation of the metalloporphyrin catalyst led to an electronic effect that capable of hindering the reaction Table 4-2 Summary of the cyclohexane oxidation catalysed by immobilised metal complexes
Catalyst FeIII(TPP)- Zeolite Y [Co2([]8-44)]- [i2([]8-44)]- [Cu2([H]8-N4O4)]a-NaY [Mn2([H]8-N4O4)]a-NaY Fe-Montmorillonite Mn-Montmorillonite Fe-Silica mol HP: mol ane 1 2 2 2 2 0.05 0.05 0.05 Solvent MeCn MeCn MeCn MeCN MeCN DCM-MeCN DCM-MeCN DCM-MeCN DCM-MeCN MeCN MeCN
oC,h

T, t

Conv. ane (%) n.d 18.6 3.7 39.8 8.7 n.d n.d n.d n.d 45.5 45.5

Select. Ol/One(%) n.d 100 100 100 100 n.d n.d n.d n.d 100 100

Yield Ol/One(%) 0.72 18.6 3.7 39.8 8.7 n.d n.d n.d n.d 45.5 45.5

HP ff(%) n.d n.d n.d n.d n.d 13 8 16 <5 n.d 37.8

ref (46) (90) (90) (90) (90) (91) (91) (91) (91) (92) (93)

, 2 60, 2 60, 2 60, 2 60, 2 RT, 24 RT, 24 RT, 24 RT, 24 60, 12 70, 8

Mn-Silica 0.05 Salen VO complex4 MCM-41 Fe Complex - Zn/Al LDH 1 aAverage value after 12 times reuse

Better results were obtained through encapsulated catalysts in non heme metal complexes. Salavati-Niasari et al (90) reported the preparation and catalytic performance of Mn (II), Co (II), Ni (II) and Cu (II) and complexes of octahydro-schiff base immobilised in nanopores of zeolite in the presence of acetonitrile solvent. Copper complex has found to be the most active system, whereas cobalt, manganese and nickel exhibited weaker activities. It is noticeable that the temperature greatly affect the catalysis performance,
47

showing the most selective and stable performance in the presence of [Cu2 ([H]8-N4O4)]-NaY catalyst at 60C. Compared with [Cu2([H]8-N4O4)], without encapsulation with zeolite framework, this catalytic performance is relatively lower. These results agree with the other experiments that demonstrated the decrease activity of encapsulated metal complex [24c, 49]. This disadvantage may be offset by the benefit of the recyclability of the catalyst. It has been shown that the catalyst shows excellent activity after 4 times running. In this case the authors hypothesised that: (i) (ii) (iii) The nanocavities are capable of immobilising the complexes The formation of inactive species in the nanocavities can be diminished due to steric effect of zeolite framework The interaction of zeolite framework with zeolite lattice. Immobilisation of homogeneous catalysts by other methods also has been reported by a number of authors. The structure of inorganic solids is capable of preventing the intermolecular aggregation of the active surface of homogenous catalyst due to their rigid structure (72). Covalent bonding is one of the simple methods to achieve this purpose. For instance, an interesting supported catalyst based on incorporation of iron and manganese porphyrin and aminofunctionalized montmorillonite K10 or silica by covalent binding has been recently investigated by the researchers from Universidade de So Paulo (91). Mineral clay is selected because of its wide range of chemical properties, its ability to expand and undergo pillarisation, its intercalation chemistry and ion change properties (91). Unfortunately the yield of cyclohexanone and cyclohexanol of these supported metallophorphyrin-HP system is by far lower compared to the catalysis system in the presence PhIO as terminal oxidant. The authors assumed that the difficulty with supported MePs-H2O2 system is the fact that the inorganic matrix can result in peroxide dismutation. In addition, the low proportion of HP/Cyclohexane and low temperature during the experiment may become other significant reasons of this poor performance. Recently a better result was obtained by using a salen oxovanadium complex immobilised in MCM-41 via multi grafting method (92). The uniformity of the mesoporous catalyst MCM-41 allows the bulky cyclohexane molecule to diffuse with minimum steric hydrance, resulting in a conversion of 45.5% with an
48

excellent selectivity (100%) after 12 h at 60C. More recently Parida et al (93) reported the oxidation of cyclohexane with iron-Schiff base complex intercalated Zn-Al layer double hydroxide (LDH). This system, unlike the encapsulated method, interestingly showed higher performance compared to the homogenous iron-Schiff base complex (without immobilisation). The authors suggested the well separation and isolation of metal centres from each other is capable of enhancing the oxidation reaction. But in case of homogenous system, the formation of inactive oxo dimer species may responsible for the lower catalytic performance. 4.4 An Overview of Heterogeneous Catalysis development 4.4.1 Summary of the Recent Research Progress Heterogeneous oxidation of the cyclohexane to Ol/One, transition metals (e.g. cobalt, iron, or chromium) and rare earth metal (cerium) substituted mesoporous HMA, SBA-3, SBA 15, KIT and MCM-41 showed active and stable performances. Titanium and vanadium ions are also active and stable, but much resistant to leaching in some cases. The physical properties of the catalyst (e.g. pores size) and the operating conditions (temperature, solvents) greatly affect the catalytic performance of molecular sieves catalyst. In many cases, acetic acid has found to be the most suitable solvent to carry out the reactions, but crystalline faujasite and its sulphated form, Na-GeX and Sulphated Ge-X, remarkably are capable of performing reactions under free solvent conditions. The emerging field of catalytic properties using boron- and fluorine-containing mesoporous carbon nitride polymers also shows fairly good results. This metal free catalyst is active and stable under the reaction conditions. Heterogenising homogeneous catalyst through immobilisation is also extensively investigated. Covalent bonding is the simplest method to achieve this, but the performances are not satisfactory. Better results were attained when the catalyst was placed inside of a zeolite or a mesoporous material. This is particularly successful for nonheme metal complexes, but failed to exhibit satisfactory results for metalloporphyrin catalysts. In addition, in most cases, the activities of the attached catalyst are somewhat reduced, but uncommon
49

results achieved in the presence of Fe(III)-Schiff base intercalated Zn-Al layer double hydroxide. The catalyst performance was slightly better than that homogeneous metal complex. However the H2O2 efficiency is rather low. A summary of literature reporting on the heterogeneous oxidation of cyclohexane with hydrogen peroxide is displayed in Figure 4.1

Molecular sieves

Rare earth metals

Cerium based catalyst Ce-MCM-41

Other metals

Germanium based catalysts Na-Gex, Sulphated Ge-X

Transition metals

Ti based catalysts TS-1, Ti-MCM-41, Ti-HMA V2O5-TiO2 Co based catalysts Co-HMA, Co-S-1, Co-APO-5 Co-SBA-3, Co-SBA-15 Co-MTiO2, CoTiO2, Co-MCM-41

Heterogeneous catalysts

Non-metals

Boron, fluorinecarbon nitride polymer (CNBF)

Cr based catalysts Cr-HMA, Cr-MCM-41

V based catalysts V-MCM-41, V-HMA, VS-1, V-KIT-6 Fe based catalysts Fe-HMA

Heterogenisation metal complexes on solid support

Intercalation

Fe Complex - Zn/Al LDH

Multi grafting

Salen VO complexMCM-41

Covalent bonds

Fe-montmorillonite Mn-montmorillonite Fe-Silica, Mn-Silica

Encapsulation

FeIII(TPP)- Zeolite Y [Ni2([]8-44)]- [Co2([]8-44)]- [Cu2([]8-44)]- [Mn2([]8-44)]-

Figure 4-1 Summary of researches on heterogeneous catalysis development

50

4.7.2 Future Prospect in Heterogeneous Catalysis Developments

Generally there are a number of pros and cons found in the oxidation of cyclohexane to Ol/One catalysed by heterogeneous system, particularly transition metal substituted molecular sieves catalysts. First, low stability of the catalysts under operating conditions was observed in some cases. The examples are TS-1, VS-1, APO-5 and V MCM-41 that are much less resistant to leaching. As pointed out by Centi et al (11), leaching of metal may lead to irreversible catalyst deactivation and it also may greatly affect all of the observed catalysis activity. The development of alternative metals such as germanic faujasite and rare earth materials is likely to be a good alternative in the near term. Second, the preparation and synthesis of many heterogeneous catalysts based on silicates-aluminates such as MCM-41, HMA, and TS-1 may involve energy-intensive hydrothermal procedures. The physical properties of the catalysts such as pore size, surface area and pore volume seem greatly affect the catalytic performance of molecular sieves catalysts. The formation of pores, however, involves drying or calcining precipitates of hydrous oxides that is carried out under extreme conditions and rather time consuming. Development of new materials of the heterogeneous catalysts that can be environmentally synthesised, e.g. carbon nitride polymers, seems to have attracts interest for the future research. Third, like the homogeneous system, the selection of the solvent plays fundamental roles in determining the polarity of medium. Various initiators also have to be found capable of enhancing the reactions. The catalysis reactions therefore usually require traditional solvents and additives such as acetonitrile, acetone and methyl ethyl ketone; with an exception of crystalline faujasite NageX and its modified sulphated form that perform the oxidation of cyclohexane under free solvent reaction. Lastly, the experimental data of H2O2 efficiency is also little bit scarce. Some of those disadvantages could be cancelled out by the notable achievements of research progress in the recent years. Reproducible result of the repeated catalysts can be an important feature for the industrial application. Compared with homogeneous catalysts, the distillation used to separate
51

reaction products and unused reactants from the homogeneous conventional cobalt catalyst is typically energy-intensive process and requires high range of temperatures. This is definitely crucial factor that may result in high downstream cost. This problem can be eliminated by using heterogeneous system since catalyst-product separation can be easily done by simple filtration. However, since the activity that can be obtained with molecular sieve catalyst is typically inferior to that homogeneous catalyst; this type of catalyst seems to be not favoured for recent industrial application. The improvement to allow the homogeneous catalyst separation, recovery and recycle through immobilisation is likely to be a trend and more preferable in the future development since it can combine the advantages of homogeneous and heterogeneous catalysts. The method immobilisation and support material seem to be greatly affected the activity and the stability of the catalysts. The important milestone of developing chemically homogeneous catalyst but physically heterogeneous catalyst has achieved in a recent study. Fe(III)-Schiff base intercalated Zn-Al layer double hydroxide catalyst succeed in demonstrating that the intercalation does not reduce the activity of the bound catalyst.

52

CHAPTER 5 DIRECT OXIDATION OF CYCLOHEXENE TO ADIPIC ACID

5.1 Biphasic reacting System (Phase-Transfer catalysts) Noyori and co-workers (5) suggested an innovative route to adipic acid by oxidising cyclohexene with 30% hydrogen peroxides via phase-transfer catalytic reaction. Hydrogen peroxide is a moderate inorganic oxidant. It can not form homogenous solutions with most organic substrates; therefore phase transfer catalyst is required to conduct the reaction. Noyori reported the use of [CH3(n-C8H17)3N]HSO4 as a phase transfer catalyst (PTC), and sodium tungstate, Na2WO4 to carry out the reaction as shown in Figure 5.1
Na2WO4 [CH3(n-C8H17)N]HSO4 HOOC

+ Cyclohexene

4 H202

HOOC

Adipic Acid

Figure 5-1 Direct oxidation of the cyclohexene to adipic acid (5)

The intermediates to define the reaction mechanism involves 6 steps (Figure 5-2) including epoxidation, hydrolysis of epoxide, ring opening, alcohol, Baeyer-Villiger oxidations and followed by oxidation to obtain anhydride. In the last step, hydrolysis of anhydride produces adipic acid at about 93% yield. This reaction can be done without the presence of any organic and halide solvent.
[H20] OH [O]

O
OH

O
[O]

O
[O]

O
[O]

O
OH

O
O
[H20] HOOC

OH

HOOC

Adipic Acid

Figure 5-2 Reaction mechanisms of the direct oxidation of cyclohexene to adipic acid (5)

53

Even though organic solvents are not used in this process, the cocatalysts, quaternary ammonium compounds, are not environmentally benign. Moreover, phase transfer catalysis is a relatively expensive process for industrial application. Further research therefore suggested the replacement of phase-transfer catalysts with a complex catalyst peroxytungstate, [W(O)(O2)2L(2)]2-, using oxalic acid as a ligand (94). Oxalic acid is able to perform reaction phase-transfer. It forms an oil phase of peroxytungstateoxalic acid complex to prevent immiscibility between cyclohexane and the catalysts. The stronger the acidity, the more oleophilic a peroxytungstateorganic complex system can be formed, and the higher yield can be produced. It has been shown that an adipic acid yield of 86 % can be obtained after 24 h in the presence of 0.5% mol (relative to the amount of cyclohexene) peroxytungstate-oxalic acid complexes catalyst. The yield will increase to 96.6% by increasing the amount of catalyst to 1% mol. For amount of catalyst up to 1.5% mol, an adipic acid yield of 93.5% can be produced after 8 h. When the same experiment was carried out using amphiphilic oxodiperoxo tungsten complex (WO(O2)22QOH) with 8-quinolinol (QOH) as a ligand, an adipic acid yield of 89.8% was obtained after 24 h (95). Success with peroxometallates has stimulated investigations of surfactant type peroxytungstates and peroxomolybdates (96) in order to stabilise the emulsion droplet and achieve more satisfactory activities. The dimeric anion of polyoxometallates, [M2O3(O2)4]2(M = W and Mo), served as a catalyst, and was combined with a long chain lipophilic cation. It possessed phase-transfer function in the W/O emulsion system to form surfactant type catalyst (STC) and tuning the hydrophile-liphophile balance of surfactant. The experimental results showed that the activities of peroxytungstates, [C16H33N(CH3)3]2[W2O3(O2)4] and [C5H5NC16H33]2[W2O3(O2)4, gave 77.8% and 78.3% yields, respectively. Molybdenum analogous catalysts, [C16H33N(CH3)3]2[Mo2O3(O2)4 and [C16H33N(CH3)3]2[Mo2O3(O2)4] resulted in low activities, affording adipic acid yields of 0% and 15.2% after 20 h. Despite reports of higher yields of adipic acid under laboratory conditions, the use of these synthesised surfactant type polyoxometallates are still expensive for commercial application. The emulsion may also result in
54

problematic handling. Another alternative is offered by using a small molecule, glicine, as a ligand. Heteropoly complexes, glycine phosphotungstate [HGly]3[PW12O40] 5H2O exhibited an excellent yield of adipic acid (95.1%) after 12 h and it has been shown that the reaction is strongly dependent of pH value of the solution (97). Table 5-1 Cyclohexene oxidation via water-organic bi-phase catalytic system in free organic solvent reactions
Catalyst [CH3(n-C8H17)3N]HSO4 and Na2WO4 [W(O)(O2)2L(2)]2[W(O)(O2)2L(2)]2[W(O)(O2)2L(2)]2[W(O)(O2)2L(2)]2WO(O2)22QOH WO(O2)22QOH WO(O2)22QOH [C16H33N(CH3)3]2[W2O3(O2)4] [C5H5NC16H33]2[W2O3(O2)4 [C16H33N(CH3)3]2[Mo2O3(O2)4 [C16H33N(CH3)3]2[Mo2O3(O2)4] [HGly]3[PW12O40] 5H2OA (NH4)6Mo7O24 4H2O (NH4)3PMo12O40 xH2O Na3[P(Mo3O10)4] xH2O [CH3(n-C8H17)3N]HSO4 and Na2WO4 (NH4)6H2W12O40 . xH2O Na3[P(W3O10)4] aq 3Na2WO4. 9WO3 (NH4)6Mo7O24 catalyst (mmol) 1 0.5A 1A 1.5A 2A 0.9 0.9 0.9 0.6 0.6 0.6 0.6 0.2 1.2 1.2 1.2 1.2 1.2 1.2 1.2 0.17 T (oC) 75-90 94 94 94 94 100 110 90 90 90 90 90 90 70-90B 70-90B 70-90B 70-90B 70-90B 70-90B 70-90B 90 Reaction time (h) 8 24 24 8 8 10 10 24 20 20 20 20 12 0.75 0.75 0.75 0.75 0.75 0.75 0.75 6 Yield of AA (%) 93 86 96.6 93.5 94.2 88.1 86.0 89.8 77.8 78.3 0 15.2 95.1 0 0 0 21 30 40 45 90 Ref (5) (94) (94) (94) (94) (95) (95) (95) (96) (96) (96) (96) (97) (98) (98) (98) (98) (98) (98) (98) (100)

A= in % mol (relative to the amount of cyclohexene) B=under microwave heating programs

Another strategy for improving the catalytic performance was proposed by Freitag et al (98). The abovementioned Noyori oxidation system was reinvestigated under microwave (MW) assisted radiations. The microwave energy is capable of enhancing reaction as a 21% yield of adipic acid was rapidly formed within 45 min. Furthermore, a various commercially available tungsten salts and molybdenum salts were also tested to substitute sodium tungstate. It has been demonstrated that sodium polytungstate, 3Na2WO4.9WO3, afford the highest yield; resulting a 45% yield of adipic acid at the same operating conditions. Some of molybdenum salts and ammonium paratungstate did not produce adipic acid. The major difficulties of the coupling of microwave
55

energy for commercial applications are the constraint to the energy input and the irradiation time to control and setting of the reaction parameters (99). Even though the application of heteropoly compound and polyoxometallates in two-phase catalytic systems have been shown to result in high yields, a highly dispersed emulsion is formed during reaction that behaves like a homogenous catalyst and is therefore difficult to be recycled. This may results in significant catalyst loss because of the complexity of the recovery process. In order to allow easier separation of the catalyst from the product mixture, designing a suitable reactor to immobilise the phase-transfer catalyst may remove this disadvantage. Recently, researchers at the University of Calabria (100) described a symmetric hydrophobic membrane (M3), characterised by a high water contact angle (R > 110) for both layers. It facilitated the oxidation of cyclohexene based on (NH4)6Mo7O24 catalyst and succinic acid as a ligand. The experiments showed that selectivity toward adipic acid of 90% can be obtained at 90C after 6 h. The pros of this reactor are: (i) providing the compartmentalisation of two reaction phases by means of polymeric membranes that result in contacting cyclohexene with the catalytic active species in the aqueous phase (ii) optimisation of catalytic performance in terms of H2O2 efficiency (iii) separation of the product from organic phase can also be achieved by this configuration. However membrane stability could be an important issue for commercial applications. A summary of cyclohexene oxidation by biphasic catalytic systems is presented in Table 5.1 5.2 Molecular Sieves catalysts As heterogeneous systems offer the possibility of continuous processing, a number of researchers have sought another method for liquid phase oxidation of cyclohexene by using molecular sieve catalysts. A summary of these experiments is presented in Table 5-2. The researchers at the University of Cambridge (101) disclosed a result of cyclohexene oxidation using a molecular sieve catalyst based on TAPO-5 under free-organic solvent conditions. Cyclohexane conversion of 100% with the selectivity toward adipic acid of 30.3% was obtained after 72 h at 80C. Better results were obtained by
56

Taiwanese researchers (102) using a periodic mesoporous catalyst (PMS) based on a mixed-valence oxotungstensilica mesoporous structure (WSBA-15) incorporating tetrahedral tungsten units. They obtained a conversion of cyclohexane up to 100% after 13 h with selectivity toward adipic acid of 30%. The selectivity to adipic acid was increased at about 45% after 30 h. However because these results do not address the possibility of recycling of catalysts, Timofeeva et al (103) investigated catalytic activity and stability by using titanium- and cerium containing mesoporous silicate TI-MMM-2 and CeSBA-15. The catalyst from the first run was filtered and washed and used in the second catalytic cycle. The results show that the yield of adipic acid decreased considerably in the second catalytic cycle, despite the fact that 100% conversion of cyclohexane can be reached. For TI-MMM-2, the yield of adipic acid decreased from 15% to 5%; also the yield of adipic acid decreased considerably from 10% to 1% in the presence of Ce-SBA-15 (103). In addition, both catalysts suffer leaching of metal due to interaction of the surface Ti sites with the reaction products. Partial loss of mesoporous structure and destruction of the active sites due to aggressive reaction medium and polar reaction products result in irreversible catalyst deactivation (103). Table 5-2 Oxidation of Cyclohexene to adipic acid catalysed by molecular sieves
(%) AA Diol Othera TAPO-5 0.5 80 24 50 13.1 64.4 22.5 (101) TAPO-5 0.5 80 72 100 30.3 30 39.7 (101) WSBA-15 0.2 85 13 100 30 59.5 10.5 (102) WSBA-15 0.2 85 30 100 45.9a 42.8 11.3 (102) TI-MMM-2b 0.2 80 24 100(100) 15(5) 8(27) 77(68) 76(59) (103) Ce-SBA-15b 0.2 80 72 100(100) 10(1) 12(15) 78(84) 67(55) (103) aother products including 2-cyclohexene-ol, 2-cyclohexenone, 2-hydroxycyclohexanone, cyclohexandione. cyclooxoheptane-1-ol-6-one and adipic acid anhydride. b Filtered catalyst was used in the second run
C

Catalyst

Weight (g)

t (h)

Ene Conv. (%)

Selectivity (%)

HP. Eff

Ref.

5.3 An Overview: Summary and the Future Prospect of Direct Oxidation of Cyclohexene to Adipic Acid For the direct oxidation of cyclohexene to adipic acid, biphasic reacting system (phase transfer catalyst) and molecular sieves catalyst have been extensively investigated to carry out the reaction. In biphasic reacting system tungsten or
57

molybdenum is the typical metals salts used as catalysts. Since the pH plays a fundamental role in determining solubility in two phases, a ligand therefore substantially required to prevent immiscibility of the reactants. The presence of surfactants may replace the use of a ligand because its properties may lower the interfacial areas. However, the generated emulsions may result in problematic handling problem. The development of a membrane reactor may resolve the problem of separating reactant and product. Microwave assisted radiation, on the other hand, successfully reduces the reaction time and affords a yield of 45% after 45 min. The use of molecular sieves mesoporous Si, Ti catalysts, whereas, is less satisfactory since catalyst deactivation is frequently found; thus it can not afford reproducible results. However, the major disadvantage of this route is, as shown in the stoichiometric reaction, that the formation a mole of adipic acid route will require 4 moles of hydrogen peroxides (figure 2), the relatively high cost of hydrogen peroxide, thus, is likely become the major drawback of commercial application (27) The new development of synthesis of hydrogen peroxide lowering the cost and to afford commercially viable solutions in small- to medium size plants therefore seems to be crucial to make this process become more attractive (104). The indirect anthraquinone process, which is only economically applicable for relatively large-scale production, accounts for 95% of global total H2O2 production. However, recently there is a significant progress of inexpensive and more environmentally benign production based on direct synthesis of H2O2 in the presence of Pd and Au-Pd catalyst that are capable of affording selectivity more than 80% (105). Another drawback of this process relates to feedstock availability. As cyclohexene is not a readily available raw material, the production of cyclohexene must be synthesised from another materials. Cyclohexene can be obtained from (27; 106): (i) (ii) (iii) (iv) The dehydration of cyclohexanol The oxidative dehydrogenation of cyclohexane Partial hydrogenation of benzene and Dehydrohalogenation of cyclohexyl aldehydes.
58

The first method may not be preferred as cyclohexanol can be oxidised directly to yield adipic acid, whereas the oxidative dehydrogenation of cyclohexane involves energy-intensive process and the yield is typically not high. Tavolaris and Keane (107) proposed an innovative strategy through dehydrohalogenation of cyclohexyl aldehydes. Despite having the potential advantages of recycling halogenated compound (27), this process suffers serious catalytic problem due to coke formation during the reaction. The most viable option is partial hydrogenation of benzene developed by Asahi Chemical (108; 109). This method is currently implemented for commercial use and affords yield of cyclohexane up to 60%. However this process is also somewhat problematic since the hydrogenation to cyclohexene (G = 98 kJ/mol, at 298 K) is thermodynamically favoured instead of the partial hydrogenation to cyclohexene (G = 23 kJ/mol) (106).

Benzene

Cyclohexane

Cyclohexyl halides Cyclohexanol Cl OH

Partial Oxidative Hydrogenation Dehydrogenation

Dehydro halogenation

Dehydration

Cyclohexene Direct (one-step) oxidation

Phase transfer catalysts Molecular sieve catalysts

HOOC Adipic Acid

HOOC

in commercial use in research phase

Figure 5-3 Summary of possible synthetic pathways to AA from direct oxidation of cyclohexene (27; 106)

59

Nevertheless this innovative route abides by the rule of green chemistry and would eliminate the conventional pathway via cyclohexane or cyclohexanol and thus it also avoid both atmospheric and non-atmospheric waste formation generated from two steps oxidation of cyclohexane. Since study in this area is still active, the step toward a new and sustainable adipic acid synthesis from this route remains possible, but it is highly dependent on the research progress of the preparation of hydrogen peroxide and cyclohexene. A summary of possible synthetic pathways to adipic acid from this route is presented in Figure 5-3.

60

CHAPTER 6 - CONCLUDING REMARKS AND OUTLOOK

The oxidation of the cyclohexane to Ol/One and direct oxidation of cyclohexene to adipic acid in the presence of hydrogen peroxide have been reviewed. The use of homogenous catalyst offers advantages in terms of possibilities to tune the performance, through modifying the properties of ligands. It can gain better level of understanding on a molecular level and generally results in better activities compared to the solid or heterogeneous catalyst (3). However, these catalystligand systems seem to be quite expensive and they can be unstable under reaction conditions. The preparation of transition metal substituted molecular sieve catalysts, on the other hand, is typically time consuming and energy intensive, but it would not result in a considerable problem as long as the catalyst can be easily recovered and give reproducible result. However, leaching of metal, arise in some cases, can lead catalytic deactivation. Nevertheless research in this area is still active, apparently driven by the potential enormous benefit of realising success and encouraged by development of various new classes of catalysts. The examples are germanic faujasite, rare earth materials, and carbon nitride nanomaterials. The biggest barrier to overcome is to combine the advantages of homogeneous catalyst and heterogeneous catalyst. There are two alternatives to answer this challenge; either immobilise homogenous catalyst on heterogeneous support or with liquid support such as water, supercritical fluid or ionic liquids. The latter method may also offer great advantage in terms of substitute the use of traditional and unacceptable solvents such as acetonitrile. The further studies related to support materials and methods of immobilisation are likely to attract great interest in near future. Direct oxidation of cyclohexene, on the other hand, is also attractive because it eliminates one step process and avoids the nitrous oxide emissions to the atmosphere and aqueous nitric acid waste and handling. Biphasic transfer catalysis is the most promising subject of research in this field. Maximising catalyst hydrophobicity/oleophilicity is the successful key in this biphasic
61

reaction. It can be achieved by modifying ligands to achieve desirable properties that are able to prevent immiscibility between cyclohexene and catalyst. Another approached proposed to improve the contact between reactants are the use of a membrane reactor and surfactants type catalysts. Microwave assisted radiation, on the other hand, is hugely beneficial in reducing reaction times, but a detail cost-benefit analysis may also be required. This route is, however, prohibitive due to the relatively cost of feedstock and hydrogen peroxide. In today adipic acid industry, the current production technologies are considered mature, where extensive research efforts is not expected to substitute the conventional methods. Furthermore the nature of hydrogen peroxide that can easily decompose to water and oxygen under the reaction conditions may decline the efficiency of hydrogen peroxide utilisation; thus hinder its industrial application. However the oxidation under hydrogen peroxide is much less energy intensive than that of air, molecular oxygen or nitric acid; and the profit margin can be obtained by significant catalysis performances. Therefore, the possibility of implementing synthesis of adipic acid using hydrogen peroxide at the commercial level is still widely open; although to date, it seems to be highly unlikely.

62

GLOSSARY
AA Ane
APO

adipic acid cyclohexane


aluminium phosphate

A/K BMPA Bpmen CNBF CHHP Conv. DMSO DCM Eff. Ene gma HMA HP H3tea h IL KA
KIT

alcohol/ketone (bis-(2-pyridilmethyl)amine) N,N-dimethyl-N,N-bis(2-pyridylmethyl)-1,2 diaminoethane] Carbon nitride boron fluorine cyclohexyl peroxide conversion dymethyl sulfoxide dichloromethane efficiency cyclohexene glyoxal-bis (2-mercaptoanil) hexagonal mesoporous aluminophosphates hydrogen peroxide triethanolaminate hour(s) ionic liquid ketone/alcohol
Korea Advanced Institute of Science and Technology

LDH MCM MeCN MEK MMM MMO MW min mqmp

layer double hydroxide mobil composition of matter acetonitrile


methyl ethyl ketone

microporous mesoporous material methane monoxygenase microwave minute(s) 2-methoxy-6-((quinolin-8-ylimino) methyl)phenol


63

NCS n.d

N-Chlorosuccinide not determined

Ol/One PDCA Phen POMs Pz TAPO Temp. TBHP TF4TMAPP Tmima TON tpcaH TPP SBA Select. Solv. STC QOH

cyclohexanol/cyclohexanone Pyrazine-2, 3-dicarboxylic acid phenanthroline Polyoxometallates (1-pyrazolyl) titanium aluminophosphate temperature t-butyl hydro peroxide (2,3,5,6-tetrafluoro-N,N,N,-trimethyl-4-anilinium) porphyrin tris[( I-methylimidazol-2-yl)methyl]amine] turnover number bis(2-pyridyl)methyl-2-pyridylcarboxamide tetraphenylphorphyrin Santa Barbara amorphous Selectivity Solvent surfactant type catalyst 8-quinolinol

64

REFERENCES
1. McKetta, J.J. Encyclopedia of Chemical Processing and Design. New York : Marcel Dekker, 1977. 2. Sri Consulting. World Petrochemicals Report on Adipic Acid. [Online] 2010. [Cited: 2 August 2010.] http://www.sriconsulting.com/WP/Public/Reports/adipic_acid. 3. Cavani, Fabrizio., Centi, Gabriele., Perathoner, Siglinda and Trifiro, Ferruccio. Sustainable Industrial Processes. Weinheim : Wiley-VCH Verlag GmbH & Co, 2009. 4. Myers, L. Richard. The 100 Most Important Chemical Compounds: A Reference Guide. 2007 : Westport. Greenwood Publishing. 5. Noyori, Ryoji, Aoki, M and Sato, K. 1998, Science, Vol. 281, pp. 1646 1647. 6. Mainhardt, Heike and Kruger, Dina. N2O Emissions from Adipic Acid and Nitric Acid Production - Good Practice and Uncertainty Management in National Greenhouse Gas Inventories. [Online] 2000. [Cited: 21 June 2010.] 7. Anastas, P.T. Bartlett, L.B. Kirchoff, M.M and Williamson, T.C. 2000, Catalysis Today, Vol. 55 , pp. 1122. 8. Anastas, P.T and Warner, T.J. Green Chemistry: Theory and Practice. New York : Oxford University Press, 1998. 9. Collins, T.J. Encyclopedia of Chemistry. New York : Macmilan, 1997. 10. Goti, Andrea and Cardona, Fransesca. Hydrogen Peroxide in Green Oxidation Reactions: Recent Catalytic Processes. Netherlands : Springer, 2008. 11. Centi, Gabriele, Cavani, Fabrizio and Trifiro, Ferruccio. Selective Oxidation by Heterogenous Catalysis. New York: Kluwer Academic : Plenum Publisher, 2001. 12. Einaga, Hisahiro and Futamura, Shigeru. 2005, Applied Catalysis B: Environmental, Vol. 60, pp. 49-55. 13. Kumar, Ranjit, Sithambaramand, Shanthakumar and Suib, Steven L. 2009, Journal of Catalysis, Vol. 262, pp. 304-313. 14. Strukul, Giorgio. Catalytic oxidations with Hydrogen Peroxide as Oxidant. Dordrecht : Kluwer Academic Publisher, 1992.

65

15. Chauvel, A and Lefebvre, G. Petrochemical Processes: Major Oxy- Genated, Chlorinated and Nitrated. Paris : Editions Technip, 1989. 16. Castellan, A, Bart, J. C. J and Cavallaro, S. 1991, Catalysis Today, Vol. 9, pp. 237-254. 17. Musser, Michael Tuttle. Ullmanns Encyclopedia of Industrial Chemistry. s.l. : Wiley-VCH Verlag GmbH & Co. KGaA, 2000. 18. Hassan, Abbas., Bagherzadeh Ebrahim., Anthony, Rayford G., Borsinger, Gregory and Hassan, Aziz. High shear process for cyclohexanol production. 7,952,493 B2 US Patent, 22 September 2009. 19. Weissermel, K and Arpe, H. J. Industrial Organic Chemistry. Heppenheim : Wiley VCH, 2003. 20. Schuchardt, U, Carvalho, W.A and Spinace, E.V. 1993, Synlett, Vol. 10, pp. 713-718. 21. Wiseman, Peter. An Introduction to Industrial Organic Chemistry. London : Applied Science Publisher, 1979 . 22. Stanley, H.M. 1968. Review Lecture. Novel Organic Chemical Processes. Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences. Vol. 303 (1474), pp. 259-273. 23. Wittcoff, H.A. Reuben and B.G. Plotkin, J.S. Industrial Organic Chemicals. New Jersey : Wiley & Sons, 2004. 24. European Comission. Integrated Pollution and Control (IPPC) reference document on best available techniques in the large organic chemical. [Online] 2003. [Cited: 11 June 2010.] http://www.bvt.umweltbundesamt.de/archive/lvocbref-e.pdf. 25. Clark, James and Macquarrie, Duncan. Handbook of Green Chemistry and Technology. London : Blackwell Science Ltd, 2002. 26. Hocking, Martin B. The Handbook of Chemical Technology and Pollution Control (3rd Edition). British Columbia : Academic Press, 2005. 27. Schuchardt, Ulf., Cardoso, Dilson and Sercheli, Rocardo., Pereira, Ricardo., Da Cruz, Rosenira S., Guerreiro, Mario C., Mandelli, Dalmo., Spinace, Estevam V and Pires, Emerson L. 2001, Applied Catalysis A: General, Vol. 211, pp. 1-17. 28. Tanase, Stefania and Bouwman, Elisabeth. 58, 2006, Advances in Inorganic Chemistry, pp. 29-75.

66

29. Mansuy, D. 1990, Pure and Applied Chemistry, Vol. 62, pp. 741-746. 30. Dzierzak, Joanna, Lefenfeld, Michael and Raja, Robert. 2009 , Topics in catalysis, Vol. 52, pp. 1669-1676. 31. Berkessel, A, et al. 1997, Journal of Molecular Catalysis A: Chemical, Vol. 117, pp. 339-346. 32. Milas, N.A and Golubovic, A. 1959, Journal of the American Chemical Society, Vol. 81, pp. 6461-6462. 33. Retcher, Brandon., Costa, Jos Snchez., Tang, Jinkui., Hage, Ronald., Gamez, Patrick and Reedijk, Jan. 2008, Journal of Molecular Catalysis A: Chemical, Vol. 286, pp. 15. 34. Ben L, Feringa., Roelfes, Gerard., Lubben, Marcel., Hage, Ronald and Que, Lawrence Jr. 2000, Chemistry - A European Journal, Vol. 6, pp. 2152 2159. 35. Gutkina, E. A, Rubtsova, T. B and Shteinman, A. A. 2003, Kinetics and Catalysis, Vol. 44, pp. 106111. 36. Chen, Kui and Que, Lawrence Jr. 2001, Journal of the American Chemical Society, Vol. 123 , pp. 63276337. 37. Fernandes, R.Ricardo, Kirillova , M.V., da Silva, J.A.L., da Silva J.J.R.F and Pombeiro, A.J.L. 2009, Applied Catalysis A: General, Vol. 353, pp. 107-112. 38. Nayak, Sanjit., Gamez, Patrick., Kozlevcr, Bojan., Pevec, Andrej., Roubeau, Olivier., Dehnen, Stefanie and Reedijk, Jan. 2010, Polyhedron, Vol. 29, pp. 22912296. 39. Fish, Richard H., Konings, Mark S., Oberhausen, Kenneth J., Fong, Raymond H., Winnie M. Yu., Christou, George., John B. Vincent, DeAnna K. Coggin and Buchanan, Robert M. 1991, Inorganic Chemistry, Vol. 30, pp. 3002-3006. 40. Carvalho, Nakdia M.F., Alvarez, Heiddy M., Horn Jr, Adolfo and Antunes, Octavio A.C. 2008, Catalysis Today, Vols. 133-135, pp. 689-694. 41. Shulpin, G.B, Sss-Fink, G and Smith, John R.L. 1999, Tetrahedron, Vol. 55, pp. 5345-5358. 42. Shul'pin, Georgiy B and Nizova, Galina V. 2007, Tetrahedron, Vol. 63, pp. 7997-8001.

67

43. Trettenhahn, Gnter., Nagl, Michael., Neuwirth, Norbert., Arion, Vladimir B., Jarr, Walther., Pchlauer, Peter., and Schmid, Walter A. 2006, Angewandte Chemie International Edition, Vol. 45, pp. 2794 - 2798. 44. Mansuy, Daniel. 1987, Pure and Applied Chemistry, Vol. 59, pp. 759770. 45. Karasevich, E. I and Karasevich, Yu. K. 2002, Kinetics and Catalysis, Vol. 43, pp. 1928. 46. Olsen, M.H.N., Salomo, G.C., Drago, V., Fernandes, C., A. Horn, Jr., Filho, L.Cardozo and Antunes, O.A.C. 2005 , The journal of Supercritical Fluids, Vol. 34, pp. 119-124. 47. Gerasimova, O. A., Shpakovskii, D. B., Milaeva, E. R., Louloudi, M. and Hadjiliadis, N. 2007, Moscow University Chemistry Bulletin, Vol. 62, pp. 264268. 48. Moreira, Maria Silvia Monsalves., Martins, Patrcia R. , Curi, Rebeca B., Nascimento, Otaciro R and Iamamoto, Yassuko. 2005, Journal of Molecular Catalysis A: Chemical, Vol. 233, pp. 7381. 49. Sawyer, Donald T. 1997, Coordination Chemistry Reviews, Vol. 165, pp. 297-313. 50. Shulpin, Georgiy B., Nizova, Galina V., Kozlov, Yuriy N., Cuervo, Laura Gonzalez and Sss-Fink, G. 2004, Advanced Synthesis & Catalysis, Vol. 346 , pp. 317-332. 51. Silva, Aires C., Fernndez, Tatiana Lopez., Carvalho, Nakdia M.F., Herbst, Marcelo H., Bordinho, Jairo., Horn Jr, Adolfo., Wardell, James L., Oestreicher, Enrique G and Antunes, O.A.C. 2007, Applied Catalysis A: General, Vol. 317, pp. 154160. 52. Kirillov, Alexander M., Kopylovich, Maximilian N., Kirillova, Marina V., Karabach, Evgeny Yu., Haukka, Matti., Silva, M. Ftima C. Guedes da and Pombeiro, Armando J. L. 2006, Advances Synthetic Catalyis, Vol. 348, pp. 159 174. 53. Silva, Telma F. S., Mishra, Gopal S. , Silva, M. Ftima Guedes da., Wanke, Riccardo., Martins, Lsa M. D. R. S and Pombeiro, Armando J. L. 2009, Dalton Transaction, Vol. 42, pp. 9207-9215. 54. Detoni, Chaline., Carvalho, Nakdia M.F., Aranda., Donato A.G., Lousi, B and Antunes, O.A.C. 2009, Applied Catalysis A: General, Vol. 365, pp. 281-286. 55. Du, Ying., Xiong, Yonglian., Li, Jing and Yang, Xiangguang. 2009, Journal of Molecular catalysis A: Chemical, Vol. 298, pp. 12-16.
68

56. Pillai, Unnikrishnan R and Sahle-Demessie, Endalkachew. 2002, Chemical Communication, pp. 2142-2143. 57. Nesterov, Dmytro S., Kokozay, Volodymyr N., Dyakonenko, Viktoriya V., Shishkin, Oleg V., Jezierska, Julia., Ozarowski, Andrew., Kirillov, Alexander M., Kopylovich, Maximilian N and Pombeiro, Armando J. L. 2006, Chemical Communication, p. 4605. 58. Kortz, Ulrich and Mal, Sib Sankar. Transition Metal Substituted Polyoxometallates and Process for their Preparation. 7,645,907 B2 US Patent, 12 January 2010. 59. Mizuno, Noritaka., Nozaki, Chika., Kiyoto, Ikuro and Misono,. Makoto. 1998, Journal of the American Chemical Society , Vol. 120, pp. 92679272. 60. Simes, M. M. Q., Conceio, C. M. M., Gamelas, J. A. F., Domingues, P. M. D. N., Cavaleiro, A. M. V., Cavaleiro, J. A. S., Ferrer-Correia, A. J. V and Johnstone, R. A. W. 1999, Journal of Molecular catalysis A: Chemical, Vol. 144, pp. 461-468. 61. Simes, Mrio M.Q., C.M.S, Santos., Balula, M. Salete S., Gamelas, Jos A.F., Cavaleiro, Ana M.V., Neves, M. Graa P.M.S and Cavaleiro, Jos A.S. 91 92, 2004, Catalysis Today, pp. 211214. 62. Santos, Isabel C. M. S., Balula, M. Salete S., Simes, Mrio M. Q., Neves, M. Graa P. M. S., Cavaleiro, Jos A. S., and Cavaleiro, Ana M. V. 2003, Synlett, Vol. 11, pp. 16431646. 63. Santos, Isabel C.M.S., Gamelas, Jos A.F., Balula, M. Salete S., Simes, Mrio M.Q., Neves, M. Graa P.M.S. , Cavaleiro, Jos A.S and Cavaleiro, Ana M.V. 2007, Journal of Molecular Catalysis , Vol. 262, pp. 41-47. 64. Mirkhani, Valiollah., Moghadam, Majid., Tangestaninejad, Shahram., Mohammadpoor-Baltork, Iraj and Rasouli, Nahid. 2008 , Catalysis Communications, Vol. 9, pp. 24112416. 65. Qi, Wei., Wang, Yizhan., Li, Wen and Wu, Lixin. 16, 2010, Chemistry A European Journal, pp. 1068 1078. 66. ICIS Chemical Business. Chemical Profile - Adipic Acid. [Online] 2007. [Cited: 9 August 2010.] http://www.chemplan.com/chemplan_demo/sample_reports/Adipic_acid.pdf. 67. Sri Consulting. Acetonitrile. [Online] 2008. [Cited: 7 July August 2010.] http://www.sriconsulting.com/CEH/Public/Reports/605.5000/.

69

68. Ballini, Roberto. Eco-Friendly Synthesis of Fine Chemicals. Cambridge : Royal Society of Chemistry, 2009. 69. Prvulescu, Vasile I and Hardacre, Christopher. 2007, Chemical Reviews, Vol. 107, pp. 2615-2665. 70. Bartholemomew, C.H and Farrauto, R.J. Fundamental of Industrial Catalytic Processes (2nd Edition). New Jersey : John Wiley & Sons, 2006. 71. Szostak, R. Handbook of molecular sieves. New York : Van Nostrand Reinhold, 1992. 72. Smith, Gerard V and Notheisz, Ferenc. Heterogeneous catalysis in Organic Chemistry. London : Academic Press, 1999. 73. Sooknoi, Tawan and Limtrakul, Jumras. 2002, Applied Catalysis A: General, Vol. 233, pp. 227237. 74. Zahedi-Niaki, M. Hassan., Kapoor, Mahendra Parkash and Kaliaguin, Serge. 1998, Journal of Catalysis, Vol. 177, pp. 231-239. 75. Poladi, Raja H.P.R and Landry, Christopher C. 2002, Microporous and Mesoporous Materials, Vol. 52, pp. 1118. 76. Selvam, Parasuraman and Mohapatra, Susanta K. 2004, Microporous and Mesoporous Materials, Vol. 73, pp. 137-149. 77. Selvam, P and Dapurkar, S.E. 2004 , Journal of Catalysis , Vol. 229, pp. 64 71. 78. Susanta K, Mohapatra and Parasuraman, Selvam. 2004, Catalysis Letters, Vol. 93, pp. 47-53. 79. Jermy, Balasamy Rabindran., Cho, Dal-Rae., Bineesh, Kanattukara Vijayan., Kim, Sang-Yun and Park, Dae-Won. 2008, Microporous and Mesoporous Materials, Vol. 115, pp. 281292. 80. Fan, Weibin., Fan, Binbin., Song, Minggang., Chen, Tiehong., Li, Ruifeng., Dou, Tao., Tatsumi, Takashi and Weckhuysen, Bert M. 2006, Microporous and Mesoporous Material, Vol. 94, pp. 348357. 81. Bellifa, A., Lahcene, D., Tchenar, Y.N., Choukchou-Braham, A., Bachir, R., Bedrane, S and Kappenstein, C. 2006, Applied Catalysis A: General, Vol. 305 , pp. 16. 82. Selvam, P and Mohapatra, S.K. 2005, Journal of Catalysis , Vol. 233, pp. 276287.

70

83. Liu, Xiaochen., He, Jiao., Yang, Lijun. Wang, Yunan., Zhang, Shihong., Wang, Wei and Wang, Jiaqiang. 2010, Catalysis Communications, Vol. 11, pp. 710714. 84. Yao, Wenhua., Fang, Hua., Ou, Encai., Wang, Jiaqian and Yan, Zhiying. 7, 2006, Catalysis Communications, pp. 387390. 85. Selvam, P and Mohapatra, S.K. 2006, Journal of Catalysis , Vol. 238, pp. 88 99. 86. Mohapatra, Susanta K., Hussain, Firasat and Selvam, Parasuraman. 2003, Catalysis Letters, Vol. 85, pp. 217-222. 87. Yao, Wenhua., Chen, Yongjuan., Min, Liang., Fang, Hua., Yan, Zhiyin., Wang, Honglin., & Wang, Jiaqiang. 246, 2006, Journal of Molecular Catalysis A: Chemical, pp. 162166. 88. Parvulescu, V.I., Dumitriu, D and Poncelet, G. 1999, Journal of Molecular Catalysis A: Chemical, Vol. 140, pp. 91-105. 89. Wang, Yong., Zhang, Jinshui., Wang, Xinchen., Antonietti, Markus and Li, Haoran. 2010, Angewandte Chemie, Vol. 122, pp. 3428 3431. 90. Salavati-Niasari, Masoud., Salimi, Zohreh., Bazarganipour and Mahdiand Davar, Fatemeh. 2009, Inorganica Chimica Acta , Vol. 362, pp. 37153724. 91. L. Faria, Andr., Leod, Tatiana O. C. Mac, Barros, Valria P and Assis, Marilda D. 2009, Journal of the Brazilian Chemical Society, Vol. 20, pp. 895-906. 92. Zhao, Jiquan., Wang, Weiyu and Zhang, Yuecheng. 2008 , Journal of Inorganic and Organometallic Polymers, Vol. 18, pp. 441-447.
93. Parida, K.M., Sahoo, Mitarani and Singha, Sudarshan. s.l. :

doi:10.1016/j.molcata.2010.06.010, Journal of Molecular Catalysis A:Chemical. 94. Deng, Youquan., Ma, Zufu., Wang, Kun and Chen, Jing. 1999, Green Chemistry, Vol. 1, pp. 275-276. 95. Li, Huaming., Zhu, Wenshuai., He, Xiaoying., Zhang, Qi., Pan, Jianming and Yan, Yongsheng. 2007, Reaction Kinetics and Catalysis Letters, Vol. 92, pp. 319-327. 96. Li, Huaming., Zhu, Wenshuai., He, Xiaoying., Zhang, Qi., Shu, Huomingn and Yan, Yongsheng. 2008, Catalysis Communication, Vol. 9, pp. 551-555. 97. Ren, Shuiying., Xie, Zhengfeng., Cao, Liqin., Xie, Xiaopeng., Qin , Gaofei and Wang, Jide. 2009, Catalysis communications, Vol. 10, pp. 464-467.
71

98. Freitag, Janine., Nchter, Matthias and Ondruschka, Bernd. 2003, Green Chemistry, Vol. 5, pp. 291295. 99. Nchter, M., Ondruschka, B., Bonrath, W and Gum, A. 2004, Green Chemistry, Vol. 6, pp. 128 141. 100. Drioli, Enrico., Buonomenna, Maria G., Golemme, Giovanni and De Santo, Maria P. 2010, Organic process research & Development, Vol. 14, pp. 252-258. 101. Lee, Sang-Ok., Raja, Robert., D. M. Harris, Kenneth., Thomas, John Meurig., Johnson, Brian F. G and Sankar, Gopinathan. 2003, Angewandte Chemie International Edition, Vol. 42, pp. 1520-1523. 102. Cheng, Ching-Yuan., Lin, Kuan-Jiuh., Prasad, Muppa R., Fu, Shu-Juan., Chang, Sheng-Yueh., Shyu, Shin-Guang., Sheu, Hwo-Shuen., Chen, Chia-Hao., Chuang, Cheng-Hao and Lin, Minn-Tsong. 2007, Catalysis Communication, Vol. 8, pp. 1060-1064. 103. Timofeeva, M.N., Kholdeeva, O.A., Jhung, S.H and Chang, J.-S. 2008, Applied Catalysis A: General, Vol. 345, pp. 195-200. 104. Perathoner, Siglinda and Centi, G. 2003, Catalysis Today, Vol. 77, pp. 287 297. 105. Edwards, Jennifer K. and Hutchings, Graham J. 2008, Angewandte Chemie International Edition, Vol. 47, pp. 9192-9198. 106. Hu, Sung-Cheng and Chen, Yu-Wen. 1997, Industrial Engineering & Chemistry Research, Vol. 36, pp. 51535159. 107. Tavoularis, George and Keane, Mark A. 1999, Applied Catalysis A: General, Vol. 182, pp. 309-316. 108. Nagahara, Hajime and Konishi, Mitshuo. Process for producing cycloolefins. 4,734,536 US Patent, 29 March 1988. 109. Nagahara, H., Ono, M., Konishi, M and Fukuoka, Y. 1997, Applied Surface Science, Vols. 121-122, pp. 448-451.

72

Das könnte Ihnen auch gefallen