Вы находитесь на странице: 1из 173

Synthesis of Crystalline Mo-V(-W-Cu)-O Complex Oxides and Their Application to Selective

Title Oxidation of Acrolein to Acrylic Acid

Author(s) ,

Issue Date 2015-09-25

DOI 10.14943/doctoral.k12035

Doc URL http://hdl.handle.net/2115/59901

Type theses (doctoral)

File Information Chuntian_Qiu.pdf

Instructions for use

Hokkaido University Collection of Scholarly and Academic Papers : HUSCAP


Synthesis of Crystalline Mo-V(-W-Cu)-O Complex Oxides and Their

Application to Selective Oxidation of Acrolein to Acrylic Acid

Chuntian Qiu

2015

Molecular Chemistry and Engineering

Graduate School of Chemical Sciences and Engineering

Hokkaido University

Table of Contents

Chapter 1 General Introduction ............................................................. 1

1.1 Selective oxidation process ........................................................................................ 1

1.1.1 Catalysts for selective oxidation ....................................................................... 1

1.1.2 Catalytic selective oxidation with molecular oxygen ....................................... 8

1.2 Selective oxidation for Acrylic acid production ...................................................... 11

1.2.1 Production of acrylic acid from propylene ..................................................... 11

1.2.2 Production of acrylic acid from propane ........................................................ 13

1.2.3 Production of acrylic acid from glycerol ........................................................ 16

1.3 Selective oxidation of acrolein to acrylic acid ......................................................... 19

1.3.1 Mo-V-based catalysts for the selective oxidation of acrolein to acrylic acid . 20

1.3.2 Mars-van-Krevelen Mechanism ..................................................................... 22

1.3.3 Effects of additives ......................................................................................... 24

1.3.4 Influence of water ........................................................................................... 27

1.3.5 Activation and deactivation of acrolein oxidation .......................................... 28

1.4 Outline of this thesis ................................................................................................ 29

Chapter 2 Synthesis of crystalline Mo-V-W-O complex oxides with orthorhombic and

trigonal structures for acrolein oxidation to acrylic acid ................................ 53


2.1 Introduction .............................................................................................................. 55

2.2 Experimental ............................................................................................................ 58

2.2.1 Synthesis of four distinct crystalline Mo-V-O ................................................ 58

2.2.2 Synthesis of W-containing Mo-V-O ............................................................... 58

2.2.3 Characterization .............................................................................................. 59

2.2.4 Gas-phase catalytic oxidation ......................................................................... 60


2.3 Results and discussion ............................................................................................. 60

2.3.1 Acrolein oxidation over four different pure crystalline Mo-V-O catalysts and

W-containing analogues ..................................................................................... 60

2.3.2 Acrolein oxidation over W-containing Mo-V-O catalysts ............................. 62

2.3.3 Effect of water on the acrolein oxidation over W-containing Mo-V-O

catalysts .............................................................................................................. 64

2.4. Conclusions ............................................................................................................. 65

Chapter 3 Effects of W on the selective oxidation of acrolein to acrylic acid over

crystalline Mo-V-W-O catalysts ........................................................... 75

3.1 Introduction .............................................................................................................. 77

3.2 Experimental ............................................................................................................ 78

3.2.1 Synthesis of three distinct Mo-V-O complex oxides ...................................... 78

3.2.2 Synthesis of amorphous and orthorhombic Mo-V-W-O complex oxides ...... 78

3.2.3 Synthesis of trigonal Mo-V-W-O complex oxides ......................................... 79

3.2.4 Characterization .............................................................................................. 79

3.2.5 Gas-phase catalytic oxidation ......................................................................... 80

3.3 Results and discussion ............................................................................................. 81

3.3.1 Structural characterization of Mo-V-W-O catalysts ....................................... 81

3.3.2 Catalytic activity for oxidation of acrolein to acrylic acid over Mo-V-W-O . 83

3.4 Conclusions .............................................................................................................. 84

Chapter 4 Synthesis of orthorhombic Mo-V-Cu-O complex oxide as highly selective

catalyst for selective oxidation of acrolein to acrylic acid ............................... 103

4.1 Introduction ............................................................................................................ 105

4.2 Experimental .......................................................................................................... 106

4.2.1 Synthesis of catalysts. ................................................................................... 106

2
4.2.2 Characterization ............................................................................................ 107

4.2.3 Gas-phase catalytic oxidation ....................................................................... 108

4.3 Results and discussion ........................................................................................... 109

4.3.1 Characterization of orthorhombic Mo-V-Cu-O catalyst .............................. 109

4.3.2 Effect of Cu on the orthorhombic Mo-V-O structure ................................... 110

4.3.3 Catalytic performance of orthorhombic Mo-V-Cu-O catalyst for the

selective oxidation of acrolein to acrylic acid. ................................................. 111

4.3.4 Effect of Cu content on the catalytic selectivity to acrylic acid ................... 112

4.3.5 Investigation of structure-activity relationships ........................................... 113

4.4 Conclusions ............................................................................................................ 114

Chapter 5 Synthesis of highly efficient orthorhombic Mo-V-W-Cu-O catalysts for the

selective oxidation of acrolein to acrylic acid ............................................ 127

5.1 Introduction ............................................................................................................ 129

5.2 Experimental .......................................................................................................... 130

5.2.1 Synthesis of orthorhombic Mo-V-W-Cu-O complex oxides. ...................... 130

5.2.2 Characterization. ........................................................................................... 130

5.2.3 Gas-phase catalytic oxidation. ...................................................................... 131

5.3 Results and discussion ........................................................................................... 132

5.3.1 Characterization of orthorhombic Mo-V-W-Cu-O catalysts. ....................... 132

5.3.2 Catalytic performance of orthorhombic Mo-V-W-Cu-O catalysts for the

selective oxidation of acrolein to acrylic acid. ................................................. 133

5.3.3 Highly effective orthorhombic Mo-V-W-Cu-O catalysts for the selective

oxidation of acrolein to acrylic acid. ................................................................ 134

5.4. Conclusions ........................................................................................................... 134

Chapter 6 Synthesis of orthorhombic Mo-V-M-O (M=Mn, Co, Ni, Zn, Fe) catalysts for

3
acrolein oxidation to acrylic acid ......................................................... 147

6.1 Introduction ............................................................................................................ 149

6.2 Experimental .......................................................................................................... 150

6.2.1 Synthesis of orthorhombic Mo-V-M-O complex oxides.............................. 150

6.2.2 Characterization. ........................................................................................... 150

6.2.3 Gas-phase catalytic oxidation. ...................................................................... 151

6.3 Results and discussion ........................................................................................... 152

6.3.1 Structural analysis of Orth-Mo-V-M-O........................................................ 152

6.3.2 Catalytic performance of acrolein oxidation to acrylic acid over

Orth-Mo-V-M-O catalysts ................................................................................ 152

6.3.3 Conclusions ..................................................................................................... 153

Chapter 7 General conclusion ............................................................ 161

Acknowledgment .......................................................................... 165

4
Chapter 1 General Introduction
2
1.1 Selective oxidation process

Selective oxidation is one of the key technologies for transformation of hydrocarbon

feed stocks (olefins, alkanes and aromatics) to value-added oxyfunctionalized chemicals in

modern chemical industry. About a quarter of organic production are synthesized via the

selective oxidation process.[1-3] Obtained oxygenated derivatives such as alcohols, aldehydes,

ketones, and carboxylic acids are widely used as raw materials for manufacturing plastics,

synthetic fibers, polyesters, etc. Some important oxidation processes are listed in Table 1-1.

Owing to the economic importance of this transformation process, selective oxidation is

highly developed by academic and industrial research in the past years. Figure 1-1 shows a

massive increase in the number of scientific publications on selective oxidation in the last

decades.[4] However, due to the complexity of interaction of many parameters that are

involved in the oxidation process, oxidations are still poorly understood chemical

transformations. Therefore, selective oxidation remains highly desirable to be developed.

1.1.1 Catalysts for selective oxidation

In parallel with the development of the modern industrial technology, traditional

energy-costly and environmentally unacceptable processes are expected to be replaced by

environment-friendly alternatives. Catalytic selective oxidation is one of the most efficient

candidates. Catalyst, as it should be the key issue of catalytic selective oxidation process, can

be roughly classified to metal catalysts, oxides catalysts, coordination complexes catalysts,

molecular sieve catalysts, etc.

Metal catalysts

Metal catalysts such as Ag, Au, Pt, and Pd are widely used in the selective oxidation

process, for example Ag catalyst for ethene oxidation to ethylene oxide, gold catalyst for

alcohol and aldehyde oxidation[5-7]. The catalytic performances of those metal catalysts

markedly depend on size/dispersion, supports, and preparation methods.

Size effect: gold, for example, was regarded to be the least reactive metal. As a
heterogeneous catalyst it showed poor activity. However, when gold is prepared to

nanoscale particles, its chemistry is dramatically changed.[8-11] High dispersed gold catalysts

are really active in many important reactions for industry production and environmental

catalysis.[12-15] For example, Corma et al. found out a gold supported CeO2 catalysts which

is active for the selective aerobic oxidation of aliphatic and aromatic aldehydes. The activity

is mainly due to the particle size of Au and CeO2.[5] Ketchie et al. evaluated

carbon-supported Au particles with a size range from 5 to 42 nm and an unsupported Au

powder as catalysts in the aqueous phase oxidation of glycerol.[10] Results showed that the

fine particles are much more active than the catalysts containing 20 nm or larger particles in

both CO and glycerol oxidation. Miller et al. further studied Au catalysts with different

particle sizes and supporters of SiO2, Al2O3, TiO2, CeO2, ZrO2.[9] As the dispersion increases

or the particle size decreases, a contraction of the Au-Au bond length occurs, by which the

activity for CO oxidation increased. Therefore, to obtain a good metal catalyst, control of

particle size and dispersion on the supports is important.

Supports: in most of cases, metal catalysts are loaded on a mono oxide or multiple

oxides supporter for instance SiO2, Al2O3, or TiO2. At the early time, supporters are only

considered to be contributed to improvements of physic properties like surface area and

pore volume. However, deepening the cognition on the role of supporters, increasing

evidences manifest that a support-metal strong interaction extensively exists, and it is

related to the reaction activity. Abad et al. observed a collaborative effect between gold and

a support, which induces the selective oxidation of alcohols.[6, 7] The turnover frequencies

for the oxidation of 3-octanol mainly depends on the kind of supporters and in the order of

AuCeO2 (nanaocrystalline CeO2) > Au/TiO2 > Au/CeO2 > Au/Fe2O3 > Au/C. Schubert et

al. grouped metal oxide-supported Au catalysts into two categories based on the study of

gold catalysts with different support materials.[16] The first group gold catalysts are

supported on inert materials, such as SiO2, Al2O3, or MgO. Those supported gold catalysts

are less active for oxidation, and show a strong dependence on the gold particle size. They

lose their activity rapidly with the increase of the gold particle size. The supporters in the
2
second group are reducible transition metal oxides for instance Fe2O3. Owing to their ability

to provide reactive oxygen, the second group supported gold catalysts exhibit significantly

enhanced activities than that of first group.

Preparation methods: metal catalysts are mostly prepared by an impregnation method,

because of the very low loading of metals, especially noble metals such as Au, Pt, Rh.

Taking the Three-Way Catalysts for example, those catalysts are used to eliminate CO,

hydrocarbons (HC), and NOx to control exhaust emissions. The loading of Pt, Pd or/and Rh

on the catalysts is as low as 0.5 wt.%.[17] However, in other high loading cases, the

impregnation method can hardly result in high dispersion of metals on the supports.

Following methods may help to obtain small metal particles with high dispersion on a

variety of supports:

(1) Co-precipitation.[18] For example Pt nanoparticles can be prepared by following

procedures. Appropriate of H2PtCl6 and cerium (III) nitrate are dissolved into deionized

water to obtain a Pt precursor solution. Then a solution of KOH was added into the Pt

solution until the pH value increases to 9. The mixture was aged for several days with a

constant stirring. After that, the mixture was centrifuged to collect the product. The product

was washed with ethanol and deionized water and dried in air overnight. Finally, the

powder was calcined in air condition.

(2) Deposition-precipitation.[19] Typically, a calculated amount of HAuCl44H2O

aqueous solution was added into a three-neck bottle containing the support. And then

ammonia water was slowly added to adjust the pH between 9 and 10 with stirring at 60 C

for 24 h. Then the precipitate was filtered, washed and dried overnight. After that, obtained

solids are calcined under the ambient condition.

(3) Co-sputtering.[20] Metals such as Au or Pt was sputter-deposited on the surface of a

substrate to form thin film. Then this film is exposed in an atmosphere condition containing

or out of molecular oxygen, and finally calcined with or without of air. The thickness of the

coating can be controlled from nanometers to hundreds of nanometers. To achieve

homogeneous metal distributions, the sample holder is rotated throughout the deposition
3
process.

(4) Chemical vapor deposition.[21] A vapor of organometallic precursor for example

(C3H5)Pd(C5H5) (C3H5 = allyl; C5H5 = cyclopentadienyl) was deposited an oxide support.

Then obtained organometallic compound is calcined in inert gas condition for a required

time to completely decompose the organometallic complex.

Oxides catalysts

Solid oxides catalysts also can be conducted by supported and unsupported oxides. As

been showed in Table 1-2, both supported and unsupported oxides catalysts are useful for a

series of important oxidation reactions. Compared to metal catalysts it is much more

challenge for scientific research on the oxides catalysts, because of the complexity of metal

ions, oxygen species, as well as oxygen and metal ion vacancies existing at the surface of

oxides. Three major factors are considerable in analyzing oxides catalysts for the selective

oxidation reactions namely redox capability, acid-base properties and phase structure.

Oxides catalysts usually contain valence alterable metal components such as Bi, Fe,

Mn, V, and Mo. They have removable lattice oxygens, and thus exhibit redox properties.

Mars-Van-Krevelen mechanism which is used for explaining a selective oxidation process

suggests that catalyst is involved in a redox cycle. The solid oxides can act as oxygen donors

in the Mars-van-Krevelen mechanism. After transfer one lattice oxygen atom from the oxide

to the substrate, the oxygen vacancy in the solid oxide is regenerated by O 2 from the gas

phase. In order to become a useful cyclic oxidant, regeneration of the reduced metal oxide

by molecular O2 must be faster than its reduction.

For the catalytic selective oxidation, activity and selectivity also depend on the

acid-base properties.[22, 23] Busca et al. measured the adsorption of ammonia, pyridine, and

acetonitrile of two (VO)2P2O7 catalysts, and suggested that strong Brnsted sites as well as

medium strong Lewis sites are present on the vanadyl pyrophosphate surface. The presence

of very strong Lewis sites coupled to V=O bonds are the active sites for n-butane selective

oxidation.[22] Shen et al. revealed the importance of surface acidity for the synthesis of
4
dimethoxymethane from the direct oxidation of methanol by poisoning the surface acidity

using K2CO3.[24] Results clearly showed that the addition of K2CO3 greatly decreased the

conversion of methanol and selectivity to dimethoxymethane. Generally, optimization of

acid-base property will only influences the adsorption properties of catalyst instead of

altering functional groups in the reactant. A strong acidity or basicity usually leads to a

strong adsorption which will thus prolong the detention time of reactants and products on

the surface of catalyst. For an organic compound, the longer time it remains on the catalysts

surface, it takes higher possibility to be completely oxidized to CO2 and H2O. On the

contrary, weak adsorption is beneficial to desorption of products, and results in higher

selectivity; however, it decelerates the oxidation process, and thus decreases activity.

Therefore, the acidity and basicity of catalysts needs to be monitored to a modest level.[23]

Phase structure is another key issue in selective oxidation processes. Industry catalysts,

in most of cases are amorphous phase or mixed phases because of the existence of various

active components and additives. Study on those sophisticated multiple phases oxides is

completely challenging. In contrast, catalysts with uniform morphologies and well-defined

structures are preferable to establish good research models and pursue a fundamental study

on catalytic behaviors. Hua et al.[25] succeed in synthesis of uniform Cu2O nanocrystals with

different morphologies. Those distinct Cu2O nanocrystals are demonstrated to exhibit

morphology-dependent catalytic activity in selective oxidation of propylene to acrolein.

Results revealed crystal plane-controlled selectivity of Cu2O catalysts. That is Cu2O

octahedra exposing crystal planes are most selective for the formation of acrolein; Cu2O

cubes exposing {100} crystal planes are most selective for the formation of CO 2; Cu2O

rhombic dodecahedra exposing {110} crystal planes are most selective for the formation of

propylene oxide. Therefore, the expression of crystal plane-depended nature is a very useful

strategy for developing selective catalysts. Another important example showing the

importance of phase structure is the discovery of crystalline Mo-V-O complex oxides for the

acrolein oxidation to acrylic acid. Mo-V-based oxides catalysts are widely used in the

partial oxidation and ammonium oxidation of light alkanes, olefins, and aldehydes since
5
1960s. Most of the industrial catalysts are composed by multiple phase oxides. With the

successful synthesis of a series of crystalline Mo-V-O complex oxides by Ueda et al.,[26] the

real active phase (orthorhombic and trigonal) for acrolein oxidation to acrylic acid was

revealed. The results clearly show that different phase structures with a similar element

constitution may exhibit much different activities under the same conducted reaction

condition. All those improvements of crystal-plane engineering of oxide catalysts provide a

very useful strategy for catalysts preparation and selection, as well as a fundamental

understanding of catalytic reactions at the molecular level.

Despite the main factors above, consideration of a metal-oxygen bond with

intermediate strength under reaction condition is also important for the section of oxides

catalysts for oxidation reactions. If the metal-oxygen bond is too strong, reaction will not

occur, while a weak metal-oxygen bond will result in over-oxidation and produce undesired

products.

Coordination complexes

Coordination complexes catalysts are also important for the oxidation reactions for

example PdCl2/CuCl2 for Wacker reaction,[27-29] Fe and Mn coordination complexes for

oxidation of Alkanes[30]. Taking the PdCl2/CuCl2 catalysts used in Wacker reaction for

example (Scheme 1-1), it is an important industrial application of coordination complexes

catalyst. Oxidation of ethylene with water to form acetaldehyde is typically promoted by

this PdCl2/CuCl2 catalyst in the presence of aqueous hydrochloric acid and an oxidizing

agent. The ultimate oxidizing agent present in the reaction mixture is often molecular

oxygen, and it serves to reoxidize Cu(I) to Cu(II) thus keeping the catalytic system active.

The reaction pathways of Wacker reaction are as follows:

PdCl2 + CH2=CH2 + H2O 2HCl+CH3CHO + Pd

Pd + CuCl2 2CuCl + PdCl2

2CuCl + HCl + O2 CuCl2 + H2O

This redox process can be divided into two stages: oxidation over PdCl2 catalyst, and
6
regeneration of PdCl2. In the first stage PdCl2 is the catalyst oxidizing ethane. In the second

stage, PdCl2 regenerate by CuCl2 and O2. From an aspect of whole oxidation mechanism,

the direct oxidizing agent is PdCl2, and O donor is H2O instead of O2. The role of O2

molecule is to help the regeneration of PdCl2 catalyst. Most of reactions employing

coordination complexes catalysts are conducted in a liquid phase condition; therefore

conferring good solubility of catalysts in the solvent is important for high activity.

Molecular sieve catalysts

Molecular sieves are crystals with an intracrystalline system of micropores of

molecular dimensions. They have been widespread application in diverse areas such as ion

exchange, separation, and catalysis since 1960s.[31-35] As a catalyst, it has many advantages

for instance high surface area, selective adsorption property,[36] and ion exchange

capacity.[37] Most zeolites discovered by now barely have strong redox property, so that they

are of no interest for redox reactions. However, chemical composition of zeolites can be

modified by transition metals such as Fe, Cu, Co, Pt, and Rh with the incorporation either

into the crystal lattice or into the micropore. Ion-exchange which has been extensively used

in the modification of zeolites is the most common method. Reducible cation pairs at ion

exchange sites in Si-Al zeolites are able to activate benzene,[38] alkanes,[35] olefins,[39] and

also inorganic molecular such as NO, and SO2. Beside ion-exchange, hydrothermal method

is also used to the incorporation of transition metals into the zeolites. For example Zhan et

al. successfully incorporated nanometer-sized hydrous RuO2 into super cages of faujasite

zeolite using a one-step hydrothermal synthesis.[32] This catalyst displays extraordinarily

high activity and selectivity in the oxidation of alcohols to the corresponding aldehydes and

ketones under aerobic conditions. And it is much more active than bulk RuO2 in the aerobic

oxidation of benzyl alcohol under the same conditions. Panov et al. found that compared to

metal oxides catalysts, Fe stabilized ZSM-5 possesses remarkable ability to generate a new

form of surface oxygen (-oxygen).[40] At room temperature, this -oxygen can react with

various organic molecules and exhibits a high reactivity for selective oxidation reaction
7
such as benzene oxidation to phenol.

Biological catalysts

Despite those traditional catalysts for selective oxidation, biological catalysts are

highly expanded in parallel with development of biotechnology. For the selective oxidation

reactions, the concerted reaction between O2 and carbon in organic compounds is difficult.

Although some traditional catalysts have been developed to catalyze specific oxygenations,

efficient and specific insertion of oxygen atom into an organic substrate is still challenge. In

the past decades, many biocatalysts have been found to be efficient for oxygenating organic

substrates using molecular oxygen. Classification of those oxidizing enzymes is showed in

Figure 1-2.[41] As one of the representative biocatalysts, oxygenase is active for

hydroxylation, epoxidation and Baeyer-Villiger reaction.[42] To activate molecular oxygen,

transition metal often used in an enzyme for example Fe-containing CYP450 enzymes.[43]

This type of enzymes can catalyze diverse specific oxidation reactions under mild reaction

conditions.

Generally, biological catalysts can catalyze reaction with extremely high reaction rate

and selectivity at a mild condition (room temperature and atmospheric pressure). However,

considering the industrial processes, only a small number of biological catalysts have been

applied in the industry production. This is partly due to the cost and difficulties of enzyme

expression or isolation. Another practical problem which hinders biocatalysts applications is

the fact that most of the enzymes show their activities depending on coenzymes. Those

coenzymes are also expensive and thus increase the cost of biocatalysts. Besides, sensitive

to temperature and pH value, and easy deactivation are also drawbacks for the wide

applications of biological catalysts.

1.1.2 Catalytic selective oxidation with molecular oxygen

In the past years, a great number of oxidizing agents are developed for selective

oxidation such as oxygen, hydrogen peroxide, ozone, NaOCl, nitrogen oxides. Nevertheless,
8
molecular oxygen is convinced to be the most common route to oxyfunctionalize

hydrocarbon because of its non-toxic and abundant. For reactions employing molecular

oxygen as oxidant to selective oxidize diverse organic substrates, the main challenges are

activation of the C-H bond in company with the activation of O2 molecule.

The O2 molecule has a double bond with a dissociation energy of 497 kJ/mol.[44] This

bond has to be broken during the oxidation process, while new bonds will be formed. On the

surface of catalysts there are many types of oxygen spices can be formed. Those oxygen

spices are important and highly related to the oxidation process. Up to date, considerable

adsorbed oxygen spices are O2, O2-, O-, and O2-. Owing to the different electronegativities of

O2-, O-, and O2-, those oxygen species play divers roles in the oxidation process. Generally,

those oxygen species can be grouped into two different type namely electrophilic and

nucleophilic oxygen which are responsible for total and partial oxidation, respectively.

Electrophilic oxygen comprises electron deficient adsorbed species such as superoxide O2-,

and oxide O-. They are inclined to attack the centers with high density of electron cloud, and

thus take effects on the total oxidation. Differently, nucleophilic oxygen includes saturated

species such as terminal oxygen groups M=O, or bridging groups M-O-M, both with the

oxygen atom in a nominal O2- state and without oxidative property. Therefore, O2- is only

related to the selective oxidation. Taking the propylene oxidation for example, the two

different oxidation processes are illustrated in Figure 1-3. Nucleophilic O2- is responsible for

the selective oxidation of propylene to acrolein then to acrylic acid, while electrophilic

oxygen species result in the total oxidation to H2O and COx. Although it is still difficult to

obtain the accurate thermodynamic data of those oxygen spices, we can roughly compare the

formation energy of them, which are showed below:

1/2 O2 (g) O (g) H1 = 248 kJ/mol

O (g) + e O- (g) H2 = -148 kJ/mol

O- (g) + e O2- (g) H3 = 844 kJ/mol

1/2 O2 (g) + 2e O2- (g) H4 = 944 kJ/mol

Those equations suggest that large amount of energy are needed for the transformation of a
9
molecular O2 to 2O2-. That is to say, O2- is the most unstable form in the gas phase condition,

and it only can be stable in the lattice. Therefore, O2- is usually considered to be lattice

oxygen instead of surface oxygen species. In contrast, O- and O2- can exist stably on the

surface of catalysts.

Although to activate both O2 molecule and the C-H bond can be overcome by using

transition metal complexes, it is still difficult to control the whole oxidation process to

chemo-, regio- and stereoselective under sustainable and environmentally safe conditions.

Therefore, it is necessary to further develop both the catalysts and theories for selective

oxidation.

10
1.2 Selective oxidation for Acrylic acid production

Acrylic acid (AA), as the simplest unsaturated mono-carboxylic acid, is an important

intermediate for the production of methyl-acrylic acid, ethyl-acrylic acid, n-butyl-acrylic acid,

and 2-ethylhexyl-acrylic acid which are widely used in polymer industry, for instance

absorbent materials, coatings, and additives in textile production.[45-48] The world

production of AA is more than 4 million tons per year in 2008, and it is increasing with a

highest annual growth rate of 5% compared to other unsaturated carboxylic acids.[48-50]

AA has been produced commercially since 1920s, employing an Ethylenecyanohydrin

process which was developed by Rhm and Haas in Darmstadt.[48] After that, other three

technical processes which are so called Acrylonitrile process, Reppe process, and

Ketene-process (or propion lactone process) have been carried out. Those four processes are

showed in Scheme 1-2. Until 1970s, AA was mainly produced by those four processes.

Apparently, all those historical processes have the same drawbacks that are usage of

expensive starting materials and involve toxic chemicals. These shortages promoted the

development of a low cost and environment-friendly process. Catalytic selective oxidation

with molecular oxygen is one of the best alternatives.

1.2.1 Production of acrylic acid from propylene

In the 1970s, the change of raw material from coal to fossil oil and the increase of less

expensive propylene made the force for the development of propylene partial oxidation to

acrylic acid. The first attempts to convert propylene to acrylic acid in one step only

achieved 35% yields in 1960s,[48] process of which is illustrated in Scheme 1-3. A

temperature requirement for this one-step oxidation process limits the total yield achievable.

Apparently, for one-step process, high temperature up to 380 oC is requisite to activate

propylene and to afford a sufficient conversion. However, under such high temperature, not

only the desired reaction of acrolein to acrylic acid takes place, but also further deep

oxidation of acrolein and acrylic acid occurs. Thus, the overall acrylic acid yield of one-step

process is low. As a result, the one-step process has not been a choice for commercial
11
production. In the following years, a great number of research work promoted the catalysts

and production process of propylene selective oxidation. With todays technology about

90% yields can be achieved based on a two-step propylene oxidation process which is

showed in Scheme 1-4. This two-step process goes through acrolein as the intermediate to

make acrylic acid [51] :

Step 1: H2C=CH-CH3 + O2 H2C=CH-CHO + H2O,

H = -81.4 kcal/mol

Step 2: H2C=CH-CHO + 1/2O2 H2C=CH-COOH,

H = -60.7 kcal/mol

In the industry production of acrylic acid from propylene, propylene was firstly selectively

converted to acrolein over Mo-Bi-based oxides catalyst; subsequently acrolein without

further purification was further selectively oxidized to acrylic acid employing Mo-V-based

oxides catalyst.

For first step of propylene oxidation to acrolein, the industrial catalytic formulations

mainly based on multi-component Mo-Bi-O metal oxides modified by various metal

elements such as Fe, Co, Ni. The conversion and selectivity are in the range of 87-98% and

90-97% respectively, and obtained in the temperature range of 250~450C due to different

catalysts. After years of study, mechanism of the catalytic oxidation of propylene to

acrolein on metal mixed oxides has been widely revealed (Scheme 1-5).[52] Isotope studies

showed that the partial oxidation of propylene to acrolein is via a formation of a symmetric

-allylic species as intermediate. And then the allyl radical is converted to acrolein via a

succession of redox and acid-base process.[53-55] The -allyl anion is firstly formed by

abstracting proton from propylene on a basic active site, and then it is oxidized to the -allyl

cation on a redox active site. Subsequently, -allylic species are formed by a nucleophilic

attack at -allyl cation with lattice O2- anion. Finally, a redox step via a hydride abstraction

result in the formation of acrolein.[53] The obtained labeled acrolein molecules reflect that

the isomeric -allylic species are formed via a nucleophilic attack of the O2- anion on either

the C1 or the C3 of the -allylic species, and the interconversion between and species is
12
reversible. [54]

Kinetic and mechanistic studies further implied that the reaction obeys the

Mars-Van-Krevelen mechanism, and is controlled by the catalyst reduction and reoxidation.

Mechanism of selective oxidation of propylene to acrolein over bismuth molybdate is

illustrated in Scheme 1-6.[56-59]. It was suggested that the lattice oxygen were involved in the

oxidation process, and the active sites for propylene chemisorption are molybdenum dioxo

groups.[54] Haber et al. studied the kinetic on the same bismuth molybdate system and

suggested that acrolein is formed on Bi III sites.[60] The slow reaction step is the activation

of propylene and oxygen. The formation of carbon oxides and acetaldehyde occurs as a

parallel reaction.

About the second step of acrolein oxidation to acrylic acid, a detailed discussion will

be carried out in the later section.

1.2.2 Production of acrylic acid from propane

Propane oxidation ascribed in Scheme 1-7 is another important process for the acrylic

acid production. It is widely investigated in the past decades, duo to the increasing

economical efficiency of propane. With the increase of natural gas exploitation, propane

which is a component with about 10 % fraction in natural gas can be obtained abundantly.

The total oxidation or combustion as a fuel to generate heat is the primary use of propane.

However, considering the average price per ton of propane is only about 1/2 compared to

propylene, increasing numbers of companies and research groups paid much more attentions

on the transformation of propane to value-added fine chemicals, for instance acrylic acid. A

great deal of investigations on catalysts and reaction systems for the selective oxidation of

propane to acrylic acid are reported in the past decades.[51, 52, 59, 61-67]

The one-step oxidation of propane with molecular oxygen to acrylic acid is listed as

follow equation [51]:

C3H8 + 2O2 CH2=CH-COOH + 2H2O

H = 171 kcal
13
As a saturated hydrocarbon, propane has low reactivity under a mild reaction conditions,

because of its high strength of C-H bonds, especially those in the terminal methyl groups. In

addition, propane is much less reactive than its partial oxidation products such as propylene,

acrolein, acetic acid and acrylic acid.[68] When temperatures are controlled to be sufficient to

activate propane, all its partial oxidation products can easily be further oxidized to carbon

oxides releasing large amount of heat. Therefore, without proper catalysts, propane is either

inactive or totally oxidized to COx while generating large quantity of heat. The challenge

for propane oxidation to acrylic acid is to activate propane and to prevent valuable C3

intermediates from the total oxidation at the same time. To overcome this challenge, it is quite

apparent that a catalytic process is indispensable. Such a catalytic process should selectively

activate the strong C-H bonds of propane while avoiding the breaking of weaker C-C bonds.

The catalysts for one-step selective oxidation of propane to acrylic acid can be

conducted to three typical systems: (1) vanadium pyrophosphate (VPO) type catalysts,

which are effective in n-butane oxidation to maleic anhydride; (2) heteropoly acids and the

salts (HPC), which are efficient catalysts for alkane oxidative dehydrogenation; (3)

multi-component mixed metal oxides, which are utilized in propylene oxidation to acrylic

acid, and are effective in propane ammoxidation and alkane oxidative dehydrogenation. To

date, most of the effective catalysts developed for selective oxidation of propane to acrylic

acid belong to these three systems and significant progress has been achieved in the past

years, especially the mixed metal oxide catalysts.

VPO catalysts have been successfully used in the industrial process from n-butane to

maleic anhydride since the 1970s. Those catalysts are very effective and perform high

conversion and high selectivity to maleic anhydride. However, VPO catalysts are not very

efficient catalysts for propane selective oxidation to acrylic acid. Additionally, some results

are sometimes difficult to reproduce due to the complexity of preparation method and

additives.

HPC refer to inorganic, heteropoly acids and the corresponding salts which are

effective in alkane oxidative dehydrogenation. A HPC often have well defined cage-like
14
structures with a central cation and surrounding polyanions. The central cation can be

cations of heteroatoms, such as P, As, Si, Ge, and B. The surrounding polyanions are often

oxoanions of Mo or W. Unlike VPO or other metal oxide catalysts, heteropolyacids

catalysts can contain acidic protons and have strong acidity. Recently, increasing efforts

have been devoted to the study of HPCs as catalysts for one-step conversion of propane to

acrylic acid with molecular oxygen as the oxidant. The representatives are H3PMo12O40,

(NH4)3PMo12O40, (PyH)3PMo12O40.[63, 69]


To date, the highest acrylic acid yield (13%)

achieved on this type of catalyst is H1.26Cs2.5Fe0.08PVMo11O40 reported by Mizuno and

co-workers.[70] However, those HPC catalysts will be rapidly deactivated (in about 12 h)

due to decomposition of the Keggin anion structure. Different from other catalyst systems,

the preparation of HPCs usually does not involve a calcination step. Therefore, the thermal

stability of the HPCs mainly depends on the formation of cage structures and arrangement

of the constituent elements of polyanion and counter-cations. Normally, for Keggin type

HPC, at the temperature below 400 oC in air condition structural decomposition will occur.

Mixed metal oxides are effective catalysts for propane oxidation to acrylic acid,

especially Mo-V-based. Since the further oxidation of acrylic acid increases greatly in

parallel with the increase of reaction temperature, mixed metal oxides catalysts have the

ability to minimize the further oxidation of acrylic acid after its formation by incorporation

of specific elements. Therefore, they exhibit excellent propane conversion and acrylic acid

selectivity. In addition, mixed metal oxide catalysts are found to be thermally stable under

the reaction conditions, because they are usually prepared via calcination process at high

temperatures. One of the major drawbacks of the mixed metal oxide catalysts is the

difficulty of controlling the preparation conditions which will greatly affect the catalytic

properties. Apparently, different preparation conditions lead to many differences in the

catalyst chemical composition, acid-base properties, morphology, structures, etc. Therefore,

the development of synthetic methods on obtaining reproducible preparation parameters is

important for the future of mixed metal oxides catalysts.

Mixed metal oxides catalysts for propane oxidation are Mo-based and most of them
15
contain V as a major component. Generally, they are prepared by a calcination method at

high temperatures. So that different from the HPC catalysts, the mixed metal oxides have

excellent thermal stability. Up to now, the most effective oxides catalysts for one-step

propane oxidation to acrylic acid are Mo-V-Te-Nb-O catalysts which were reported by

Gerhard Mestl.[66] Those catalysts achieved 50% yield of acrylic acid. In the propane

oxidation, Mo and V are responsible for the activation of propane, while Nb and Te are part

of the active sites responsible for the formation or release of C3 oxygenates. Moreover, the

existence of Nb in these mixed oxide catalysts was believed to stabilize the structure of

Mo-V oxides, and was contributed to return to its original state from a strongly oxidized or

reduced catalyst.[71]

The oxidation of propane can take place via many different pathways, as illustrated in

scheme 1-8. [51, 72] The major processes are based on the progressive oxidation from propane

to propylene, and then to acrolein, and finally to acrylic acid. As been noticed in section

1.2.1 selective oxidation of propylene, different temperature requisite in each oxidation step

limits the one-step oxidation method. Therefore, multi-step oxidation process is more

considerable to achieve a high acrylic acid yield from propane.

1.2.3 Production of acrylic acid from glycerol

In parallel with petroleum reserves diminish, much more attention has been focused on

the conversion of bio-renewable feedstocks into commodity chemicals and clean fuels.

Glycerol, as a potentially important bio-refinery feedstock, is one of the most promising

building blocks for the production of fine chemicals. Nearly two-thirds of glycerol is

produced as a byproduct of biodiesel via transesterification of vegetable oils or animal fats

with alcohol (Scheme 1-9). For every 9 kg of biodiesel produced, about 1 kg of glycerol

byproduct is formed. More than 350,000 tons of glycerol is produced per annum in the USA,

and about 600,000 tons is produced per annum in Europe.[73, 74]. High functionality and

relatively low price make it a potential precursor to produce value-added derivatives. Several

processes, so called deoxygenation, were applied to eliminate its oxygen content.


16
Moreover, the dehydrated product, acrolein, can be readily converted to other C3

petrochemicals for example acrylic acid.

A variety of multifunctional materials have been demonstrated to be effective catalysts

for the direct transformation of glycerol into acrylic acid, that is POM[75], zeolites[76, 77], and

metal oxides[78] catalysts. Thanasilp et al.[75] prepared a series of alumina-supported

polyoxometalate (Al2O3-supported POM) catalysts by impregnation method for the liquid

phase catalytic oxydehydration of glycerol to acrylic acid at a low temperature of 90C.

Supported Al2O3 with 4 wt.% H4SiW12O40 loading exhibited a highest glycerol conversion of

about 84% with a yield of acrylic acid of 25%. Pestana et al.[76] studied the oxidative

dehydration of glycerol to acrylic acid over vanadium-impregnated zeolite Beta. Catalysts

were prepared by wet impregnation of ammonium metavanadate over ammonium-exchanged

zeolite Beta, followed by air calcination at 823 K. Impregnation reduced the specific surface

area, but did not significantly affected the acidity (Brnsted and Lewis) of the zeolites.

Glycerol conversion achieved at about 75% while acrylic acid was formed with

approximately 25% selectivity at 548 K.

Similar with the production of acrylic acid from propylene and propane, the conversion

of glycerol to acrylic acid can be preceded in two steps (Scheme 1-10). In the first step,

glycerol is dehydrated to acrolein. In the second step, acrolein is selective oxidized to acrylic

acid of in the presence of molecular oxygen.

For the first step of glycerol dehydration in Scheme 1-10, acid catalysts such as

H-ZSM5 zeolite,[79] H4SiW12O40/Al2O3 solid acid,[80] W-Nb-O complex metal oxides,[81] and

vanadium pyrophosphate oxide (VPO)[82] are found to be highly effective. Although those

catalysts have been identified to be active for the acrolein production from glycerol, fast

deactivation of catalyst is a major problem that needs to be overcome. For example, the acid

zeolite catalysts are easy deactivated by the obstruction of micropores. Up to 30 wt.%

products can be deposited on the catalysts during glycerol dehydration. This is presumably

because acrolein is highly reactive to form other products after its formation in the

dehydration process. Omata et al. reported a W-Nb complex metal oxides catalyst for
17
glycerol dehydration to acrolein.[81] This catalyst showed more than 70% yield of acrolein.

More importantly, the deactivation rate in the W-Nb-O catalyst was less than that in

WO3/ZrO2 or H-ZSM-5.

For the second step of acrolein oxidation to acrylic acid, a redox catalyst is needed,

basically, Mo-V-based complex oxides which will be discussed in detail in the next section.

18
1.3 Selective oxidation of acrolein to acrylic acid

Selective oxidation of acrolein to acrylic acid can be considered as the final step in the

multi-step process transformation of propylene, propane and glycerol to acrylic acid.

Although one-step transformation of those feedstocks is available, yields of one-step

method are not satisfactory. A rigid temperature requirement for the one-step oxidation

limits the total yield achievable. In the one-step reaction, the operating temperature has to

be controlled to be high to afford a sufficient conversion due to the high activation energy of

C-H bond. However, under such high temperatures, not only the desired reaction takes place,

side reaction and deep oxidation are also likely to occur. Thus, the overall acrylic acid yield

of the one-step process is lower than the two-step process. In recent years, the dehydration

of glycerol to acrolein is one of the attractive approaches for glycerol utilization (Scheme

1-10).[79-83] Owing to the development of glycerol dehydration technology, abundant of

acrolein can be produced. Therefore, the transformation of acrolein to acrylic acid is more

attractive.

Since the commercialization of the process of acrolein oxidation to acrylic acid in

1960s, a great number of catalysts have been prepared and studied, for example Mo-V-based

oxides, Mo-Co-based oxides, and V-Sb-based oxides. Up to date, the most effective catalysts

for the selective oxidation of acrolein to acrylic acid are Mo-V-based multi-component

oxides modified by transition metals. Early catalysts are phase-mixed oxides and a large

number of elements such as Na, Mg, Ni, Ce, Ti, and Cu were introduced into the catalysts as

promoters. Those additional metals enhanced activity, selectivity to target products and also

the thermal stability. However, multiple elements in those catalysts increased the difficulty of

scientific research on the nature of catalysts and reaction.

Recently, a series of crystalline Mo-V-O complex oxides were successfully synthesized,

and the application of those crystalline Mo-V-O on the acrolein oxidation gained remarkable

results (Figure 1-4). The temperature of 100% acrolein conversion was decreased from about

300 oC to less than 250 oC. More importantly, the crystalline structure supplied more simple

and reliable models for the further discussion on the effects of promoters, structure-activity
19
relationships.

1.3.1 Mo-V-based catalysts for the selective oxidation of acrolein to acrylic acid

Mo-V oxides with a diverse of promoters such as W, Fe, Ni and Cu are efficient catalysts

for acrolein oxidation to acrylic acid. The catalytic performance of those catalysts highly

depends on the composition, oxidation state, acid-base properties, and preparation method.

Andrushkevich et al. intensively investigated the catalytic property of binary Mo-V

oxides system.[84] His study shows that MoO3 and binary compositions with low vanadium

content have very low activity. With the increase of vanadium, high activity and selectivity

to acrylic acid were achieved. However, catalysts enrich with vanadium (> 40 mol%)

enhanced the total oxidation of acrolein. Apparently, vanadium is concerned with the active

component, and the ratio of Mo/V plays an important role in the acrolein transformation.

Normally, Mo/V ratios in the region of 3 to 4 deliver the best catalytic performance.[85]

Oxidation state of each element is also one of key issues in the catalysts system. In the

case of vanadium, contrast test of acrolein oxidation over V2O4-MoO3 and V2O5-Mo3 shows

that V2O4 has a high selectivity to acrylic acid, while V2O5 only catalyzes complete

oxidation. One possible reason is that V4+ can serve as the centers for stabilization of

acylate which is the intermediate compound for acrylic acid formation. Oxidation state also

affects the redox property with the involvement of lattice oxide ions in the reaction. For the

partial oxidation reactions, a redox mechanism usually proceeds by the repetition of

reduction and oxidation of catalysts. Oxidation state of both bulk and surface element will

influence the immigration of lattice oxygen, which plays a very important role in the redox

process.
[86, 87]
Ai et al. investigated the acid-base properties of numerous oxidation catalysts.

He summarized that the activity and selectivity in mild oxidation can be well interpreted in

connection with the acid-base properties of catalysts. The acidic sites contribute to the

activation of electron-donor-type reactants such as olefins. The basic sites are connected

with the oxidizing sites. The electron-donating (basic) reactants such as olefinic and
20
aromatic hydrocarbons are correlated with the acidic nature of the catalyst surface. Catalytic

activity for the oxidation of acidic reactants such as carboxylic acids is correlated with the

basic nature of the catalyst.

Beside the chemical composition, oxidation state, the structure and morphologies also

have considerable influences on the catalytic performance. All those factors will differ by

various preparation routes. With the passage of time, preparation method of the catalysts for

acrolein oxidation is greatly improved. There are two major methods named as high

temperature calcination and spray drying process. In the early time, the most common

method for the preparation of Mo-V catalysts is high temperature calcination. Mo-V oxides

were prepared from vanadium oxalate or ammonium metavanadate and paramolybdate under

a moderate temperature around 400 oC, or anneal a mixed solid with MoO3, V2O5 and

promoter precursors in the desired stoichiometry at 600 oC for several days.[48] The second

preparation method spray drying is based on aqueous solution, in which ammonium salts of

Mo and V were dissolved. The precursor solution was dried via a spray drying process.

Subsequently, dried catalysts were calcined at about 400 oC.[48, 85] Recent years, with the

discovery of crystalline Mo-V-O complex oxides, hydrothermal synthesis method is widely

employed. For hydrothermal method, soluble Mo and V precursors were dissolved in an

aqueous solution, which were then transferred to an autoclave with a Teflon inner tube and a

Teflon sheet, and hydrothermally treated at 175 oC for 48 h.[26, 88, 89]
For the industrial

catalysts, the shape of achieved Mo-V-O catalysts are changed from spheres to tablets, and

then to rings. Owing to the constant improvements of catalysts and preparation method, about

90% yields can be achieved over Mo-V-based catalysts.

More recently, a series of single crystalline Mo-V-O complex oxides were discovered by

Ueda et al.[88, 90, 91]


These Mo-V-O complex oxides namely orthorhombic, trigonal,

tetragonal, and amorphous were synthesized by a hydrothermal method. Synthesized

materials are rod-shaped and have a layered structure with slabs. The slabs are consisted of

pentagonal {Mo6O21} building units and {MO6} (M = Mo, V) octahedra but with different

arrangement. Slab grows along c-axis. The distance between each slab is 0.4 nm.[88, 89, 92, 93]
21
The pentagonal {Mo6O21} building units contain an {MoO7} pentagonal bipyramidal center

surrounded by five edge-shared {MoO6} octahedra. High-angle annular dark-field

(HAADF) scanning transmission electron microscopy (STEM) imaging clearly showed the

different structure of orthorhombic symmetry (M1-type), Mo5O14-type tetragonal symmetry,

or trigonal symmetry (Figure 1-5).[94] For the M1-type phase, HAADF analysis suggests

that both Mo and V occupy the octahedral sites linking the pentagonal units.[90] Owing to

the existence of channels in the orthorhombic and trigonal structure, small molecules such

as methane, ethane, and CO2 can enter into the pore and be selectively separated.

The application of those crystalline Mo-V-O materials especially orthorhombic and

trigonal phases on the selective oxidation of acrolein to acrylic acid gained remarkable

results.[26] Catalytic activities of those crystalline materials are illustrated in Figure 1-4. Due

to the simplicity of component and uniformity of the crystalline structure, those Mo-V-O

complex oxides provide suitable models for understanding relationships between

composition, structure and catalytic functions.

For synthesis of orthorhombic and trigonal Mo-V-O mixed-metal oxides, ammonium

heptamolybdate ((NH4)6Mo7O244H2O), and vanadyl sulfate (VOSO4nH2O n=5.4, 3.28 g,

Mitsuwa Chemicals) are firstly mixed in a water solution. After adjusting the pH value of

this precursor solution, the orthorhombic or trigonal Mo-V-O was selectively synthesize

under hydrothermal conditions.[89] Latest results showed that when the pH value of

precursor solution is controlled to higher than 1.7, a ball-shaped polyoxometalate

[Mo72V30O282(H2O)56(SO4)12]36 ({Mo72V30}) is produced in the reaction precursor solution.

This polyoxometalate contains 12 pentagonal [Mo6O21]6- units and 30 [V=O]2+ units as

linkers. Hydrothermal synthesis of the precursor solution will produce crystalline

orthorhombic or trigonal Mo-V oxides which also contains pentagonal polyoxomolybdate

unit with other Mo and V ions.[95].

1.3.2 Mars-van-Krevelen Mechanism

Partial oxidation of acrolein to acrylic acid is not only a commercially important


22
reaction but also a convenient reaction for scientific research. Because the numbers of

products (CO, CO2, acrylic acid) is moderate, and the catalysts for this reaction can be

simplified to binary compositions while keeping a high activity and selectivity to target

product. The study of this model reaction supplies useful and reliable information that helps

us to summary the fundamental theory of catalysts actions and reaction dependencies.

The selective oxidation of acrolein to acrylic acid over a heterogeneous oxide catalyst

usually follows a Mars-van-Krevelen Mechanism. This Mechanism is a redox process. For

acrolein oxidation, it can be divided into four procedures: (1) acrolein is adsorbed on the M1n+

(M= metal) center on the surface of catalyst forming an adsorption intermediate; (2)

intermediate takes reaction with the neighboring lattice oxygen to form products, at the same

time, electrons transfer from M1n+ to M2m+; (3) desorption of products; (4) electrons transfer

back from M2m+ to M1n+, simultaneously, molecular oxygen is adsorbed on M2m+ center, and

thus transforms to lattice oxygen. Catalysts are recovered through the four procedures which

are ascribed in Figure 1-6.

More detailed mechanism of acrolein oxidation to acrylic acid is illustrated in Figure 1-7.

In the first acrolein adsorption step, acrolein produces an interaction with Mo6+ cation. This

interaction proceeds via a donor-acceptor mechanism by the transfer of the lone electron

pair of carbonyl oxygen atom to the vacant orbital of cation on the oxide surface.

Subsequently, proton transfer results in producing the CH2=CH-C=O- fragment which is a

good electron donor. Then the nucleophilic attack of this fragment on the oxygen of catalyst

leads to the formation of the carbonyl bound compound SI-III and it is very fast transfer into

the acrylate SI-IV. In the case of vanadium-molybdenum oxide catalyst the surface acrylate

is stabilized on V4+. The decomposition of surface acrylate which produces acrylic acid is

the slowest step. The decomposition rate mainly depends on the acid-base properties of the

centers and the bond energy of the surface oxygen. Thus, a catalyst which can perform high

acrolein conversion and high selectivity to acrylic acid should contain centers with optimum

acid-base properties and strong bond energy of nucleophilic surface oxygen.

23
1.3.3 Effects of additives

The incorporation of additives into the catalyst can modify properties such as acidity,

thermal stability, and especially catalytic activity. Numerous components such as K, Mg, Cu,

W, and Fe have been applied to optimize the structure and catalytic performance for

Mo-V-based oxides catalysts.[48, 61, 96-101] According to the mechanistic studies of oxidation of

acrolein to acrylic acid discussed in section 1.3.2, catalytic activity mainly depends on two

parameters [84, 102]: (1) the bond strength of the intermediates with the catalyst surface; (2) the

bond energy of lattice oxygen. The activity and selectivity over the Mo-V oxide catalysts can

be optimized by adding additives, by which above two parameters can be controlled.

Apparently, it is necessary to select a catalyst with weak and reversible adsorption of

both acrolein and acrylic. If strong bonds exist on the surface structures, desorption of

acrolein and acrylic acid will take place at higher temperatures and thus the carbon chain may

be broke. Due to the different acidic properties of acrolein and acrylic acid, it is possible to

modify the surface acid-base properties in order to affect the binding strength of

intermediates by incorporation of additives.[103] There are two types of additives, alkaline

additives and acid additives: (1) Alkaline additives, such as Cs, K, Na, do not affect the initial

selectivity. The adsorption of acrylic acid is stronger than acrolein when alkaline promoted

the catalyst. (2) Acid additives, such as phosphorus, will sharply decrease the initial

selectivity. The adsorption of acrolein is stronger than acrylic acid when the catalyst is

promoted by an acid component. Therefore, the selectivity of Mo-V oxide catalyst can be

optimized by using alkaline additives or weak acid additives.

The second way to control the catalytic properties of the Mo-V oxide catalyst is to

optimize the bond energy of lattice oxygen. The bond energy of oxygen in the catalyst lattice

must be controlled to optimum, because the lattice oxygen takes part in the transformation of

intermediates and is responsible for regeneration of surface structures. Weakly bonded

oxygen forms complexes with a higher oxygen content, which results in the destruction of the

adsorbed molecule and then lead to a nonselective oxidation. With the increase of bond

energy of oxygen, reaction rate will decrease while selectivity may increase.
24
Industrial Mo-V oxide catalysts always contain W and Cu as additives for selective

oxidation of acrolein to acrylic acid. However, the functions of those additives so far are

poorly understood. So here, I only focus on the role of W and Cu.

Tungsten

W is widely used in catalysts for diverse of reactions, for example alcohol oxidation [104,
105]
, light alkanes oxidation [106-110], and also acrolein oxidation to acrylic acid [48, 49, 85, 111, 112].

The introduction of W can influence the activity of catalysts by changing the phase structure,

adjusting the surface composition, acid-base properties, etc:


[113]
(1) W acts as a structural promoter. Safikhani et al. mentioned that adding a low

amount of W to traditional metal alloys is effective to control the elemental diffusivity and to

improve the mechanical property of the high temperature strength, oxidation resistance. H.

Vogel et al. [48, 112] synthesized a series of Mo8V2WcOx catalysts with a varying amount of W

and conducted those catalysts in acrolein oxidation to acrylic acid. Results showed that WO3

is inert for acrolein oxidation and can act as a structural promoter to improve long-term

stability. With an increase of W content, temperature stable amorphous or nanocrystalline

structures formed. However, the amount of inert is also increased.


[109]
(2) W as an electrophilic additive optimizes acidity. Pierini et al. investigated the

promotion effect of W in VPO catalyst in the partial oxidation of n-butane to maleic

anhydride and reveal the promoting effect W is to increase the concentration of very strong

Lewis acid sites by creating surface defects. The presence of trace amounts of W at the
[114]
surface may contribute to the CH bond breaking of the starting paraffin. Yang et al.

confirmed the presence of strong Brnsted acid sites and Lewis acid sites when tungsten

oxide species is incorporating into a MCM-48 material. The optimized acid properties are

beneficial to the selective oxidation of cyclopentene to glutaraldehyde.


[110]
(3) W contributes to electron transfer. Zhu et al. reported that the addition of W to

Pt/ZSM-5 catalysts can increase the oxidation resistance of platinum by an electron transfer

from tungsten to platinum at the WPt interface. Based on the interaction between platinum
25
and tungsten oxide, the catalytic activity of the total oxidation of propane is improved.

Pierini et al. [109] also suggested that the promotion effect of W in VPO catalyst in the partial

oxidation of n-butane to maleic anhydride is connected with the generation of V5+ sites which

is close to the abundant V4+ center of the pyrophosphate structure.

Although W is an excellent additive, the effect of W on the acrolein oxidation to acrylic

acid is poorly understood due to the complexity of multiple components and structures in the

industrial catalysts. A possible approach to overcome this issue is to prepare a highly efficient

catalyst with simple structure and component. The crystalline Mo-V-O materials can supply

convenient models to investigate the roles of additional metals. However, it is a challenge

work to introduce heteroatom into a catalyst keeping the crystalline structure, because

additional metals usually affect the normal crystalline growth process. In this thesis I succeed

in incorporating W into the framework of crystalline Mo-V-O structure, and revealing the

functions of W on the selective oxidation of acrolein to acrylic acid.

Copper

Cu is also important for the selective oxidation process, and a great deal of studies on the

effect of Cu in oxidation reaction process have been reported. For example, Coloma et al. [115]

studied the competitive adsorption of crotonaldehyde and CO on Cu/TiO2 using infrared

spectra. Results showed that the co-existence on catalyst surfaces of Cu(0) and Cu(I) sites are

particularly active for hydrogenation of carbonyl (C=O) bonds. Byesen et al.[116, 117]

investigated the synergistic effect between copper and vanadium in AlPO-5 during

the selective oxidation of propylene. Their studies suggest that the activity is directly
related to the reducibility of vanadium, and Copper(II) showed a clear promotion effect on the

reduction of V(V) V(IV)/V(III). Hua et al. [25] revealed that selectivity of Cu2O catalysts

highly depends on Cu2O morphology in the catalysis of propylene oxidation with O2.
[66]
Although Cu may show detrimental effect on acrylic acid yields , it is still widely

used in industrial Mo-V oxides catalysts for acrolein oxidation to acrylic acid. It is

worthwhile to notice that in the Andrushkevichs report[84], the catalysts with CuO showed
26
the highest selectivity to acrylic acid. In the acrolein oxidation to acrylic acid, the adsorption

of acrylic acid is stronger than acrolein when alkaline promoted the catalyst. CuO as an

alkaline additive may optimize the surface acid-base properties of catalyst, and thus

improve the catalytic selectivity. Introduction of Cu may optimize the bond energy of lattice

oxygen. In addition, Cu has strong redox ability which may contribute to the redox cycle

and is responsible for regeneration of lattice oxygen.

To incorporate Cu into a crystalline structure, the common approach is ion-exchange

method. However, using ion-exchanged method so far cannot obtain Mo-V-Cu-O catalyst

with a uniform crystalline structure. So, in this thesis, I employ a hydrothermal method to

synthesize crystalline Mo-V-Cu-O catalysts with an orthorhombic structure.

1.3.4 Influence of water

Water is a co-product of propylene or propane oxidation as well as glycerol dehydration

to acrolein. In the industrial process, after acrolein production process, water, acrolein and

additional air are further oxidized to acrylic acid without further purification. For acrolein

oxidation, although water is neither reactant nor product of the oxidation of acrolein to acrylic

acid, its presence in the gas phase speeds up the reaction. The influence of water on the

reaction can be summarized as follows:

(1) A physical reason is that as an inert mass, water has a high heat capacity. It can

reduce hot-spots, in which way water reduces the formation of coke layer and decelerates the

catalyst deactivation.

(2) The addition of water in the reaction results in the formation of a great number of

hydroxyl groups on the surface of metal oxide catalyst, by which Lewis-acid centers are

transformed to Brensted-acid centers. These electrophilic centers are available for the partial

oxidation of acrolein to acrylic acid. Landi et al. have found a linear dependency between the

acrylic acid formation rate and the number of acid groups on the catalysts surface.[64]

(3) In the catalytic oxidation of acrolein process, acrylate is an intermediate on the

surface, it is also possible to consider the following process: [102]


27
CH2=CH-COO- + H2O (g) = CH2=CH-COOH (g) + OH-

Water accelerates the step and is contributed to desorption of acrylic acid from the surface of

catalyst, and thus active sties can be exposed again.

There is also possibility that water combines with the reactants to form intermediates

that adsorb more easily and faster on the catalyst surface. Although many mechanisms

described the partial oxidation of acrolein to acrylic acid, a mechanism with the participation

of water still not clear and it must be further investigated.

1.3.5 Activation and deactivation of acrolein oxidation

Normally, before the reaction catalysts are activated in a reaction gas which contains

acrolein, O2, H2O and balance gas (usually N2 or He) at appropriate temperature for several

hours. The catalyst is accompanied to its steady state by the activation process.

Deactivation of catalysts is one of the top issues that matters to catalyst section.

Numerous of factors are highly related to the catalysts deactivation, for example lose of

active components, structure decomposition of catalyst, shape change of forming-catalyst,

and competing reaction. For acrolein oxidation, the deactivation of catalyst is mainly caused

by the structure decomposition because of the high reaction temperature. In the case of Mo-V

phase-mixed oxides catalysts, in parallel with the increase of reaction time, agglomeration of

catalyst gradually happened. And it results in diminish of active sites and change of surface

properties, by which activity significantly decreases. Similarly, for crystalline Mo-V-based

complex oxides, with the increase of reaction time, crystalline structures are gradually

decomposed to less active phases. Therefore, thermal stability is one of the key issues of a

good catalyst for acrolein oxidation.

28
1.4 Outline of this thesis

Crystalline Mo-V-Os were found to be very efficient for selective oxidation to acrylic

acid. The addition of transition metals W and Cu into the crystalline structure is expected to

further improve the activity and selectivity. Those crystalline Mo-V-Os are suitable models

to identify the effects of W and Cu as well as the fundamental investigation of

structure-activity relationships.

This thesis mainly focuses on the synthesis of highly efficient crystalline

Mo-V(-W-Cu)-O catalysts for the selective oxidation of acrolein to acrylic acid. A series of

crystalline Mo-V-W-O, Mo-V-Cu-O, and Mo-V-W-Cu-O catalysts were successfully

synthesized by a hydrothermal method. Selective oxidation of acrolein to acrylic acid was

performed in gas phase over these catalysts. Effects of W and Cu on acrolein oxidation were

also revealed.

In chapter 2, W was successfully introduced into Mo-V-O with keeping the

orthorhombic, trigonal, and amorphous structures. Synthesized crystalline Mo-V-W-O with

orthorhombic and trigonal structures, both of which possess heptagonal channels, showed

catalytic activity for the gas phase acrolein selective oxidation to acrylic acid is superior to

amorphous Mo-V-W-O and to tetragonal Mo-V-O. The results suggest that W-containing

catalysts showed less water dependency on acrolein oxidation. Incorporation of W into the

orthorhombic structure can increase catalytic activity of acrolein oxidation to acrylic acid.

In chapter 3, a series of crystalline Mo-V-W-O complex oxides with the orthorhombic or

trigonal structure were synthesized by a hydrothermal method, and were characterized by

inductively coupled plasma atomic emission spectroscopy, TEM, STEM-EDX, X-ray

diffraction, Rietveld analysis, and a N2 adsorption method. It was found for the first time that

W can be successfully incorporated into the trigonal Mo-V-O structure by using

(CH3CH2NH3)2Mo3O10. The alkylammonium cation acted as a structural stabilizer that was

requisite for the formation of a trigonal structure when additional metal ions were present.

For the orthorhombic Mo-V-W-O structure, introduction of W into the orthorhombic

structure caused a rod segregation effect by which nanoscale crystals formed and the external
29
surface area greatly increased. These Mo-V-W-O materials were applied as catalysts to the

gas phase selective oxidation of acrolein to acrylic acid. The best catalyst was assigned to

orthorhombic Mo-V-O-W7.5%, which possessed an ordered arrangement of heptagonal and

hexagonal channels and a large external surface area. Additionally, effect of water

concentration on acrolein oxidation was investigated.

In chapter 4, a series of crystalline Mo-V-Cu-O complex oxides with an orthorhombic

structure were synthesized for the first time by a hydrothermal method using

(CH3NH3)6Mo7O24. Cu ions were determined to be in the heptagonal channels by scanning

transmission electron microscopy (STEM), X-ray diffraction (XRD), Rietveld refinement,

and N2-adsorption methods. Addition of Cu into the orthorhombic structure can enhance

acrylic acid selectivity to more than 99% in the gas phase acrolein selective oxidation to

acrylic acid.

In chapter 5, a series of crystalline Mo-V-W-Cu-O complex oxides with an

orthorhombic structure were synthesized. These crystalline Mo-V-W-Cu-O complex oxides

are efficient catalysts for selective oxidation of acrolein to acrylic acid. Compared to Mo-V-O

catalysts, both catalytic activity and selectivity to acrylic acid were enhanced. As the best

catalyst, Orth-MoVW7.5%Cu7.5%O achieved more than 90% acrolein conversion and 99%

selectivity to acrylic acid at 230 oC.

In chapter 6, a variety of crystalline Mo-V-M-O (M= Mn, Fe, Co, Ni, Zn) complex

oxides with an orthorhombic structure were synthesized by a hydrothermal method using

(CH3NH3)6Mo7O24 as Mo precursor. Transition metals were supposed to be in the heptagonal

channels. These materials are good catalysts for the acrolein oxidation to acrylic acid, and

they are expected to expand the application domain of crystalline Mo-V-based complex

oxides.

In chapter 7, important results and general conclusions were drawn.

30
References

[1] Grasselli R.K. Catalysis Today, 1999, 49, 141-153.

[2] Mestl G. Topics in Catalysis, 2006, 38, 69-82.

[3] Guo Z., Liu B., Zhang Q.H., Deng W.P., Wang Y., Yang Y.H. Chemical Society

Reviews, 2014, 43, 3480-3524.

[4] Hermans I., Spier E.S., Neuenschwander U., Turra N., Baiker A. Topics in Catalysis,

2009, 52, 1162-1174.

[5] Corma A., Domine M.E. Chemical Communications, 2005, 4042-4044.

[6] Abad A., Concepcion P., Corma A., Garcia H. Angewandte Chemie-International

Edition, 2005, 44, 4066-4069.

[7] Abad A., Corma A., Garcia H. Chemistry-a European Journal, 2008, 14, 212-222.

[8] Ketchie W.C., Fang Y.L., Wong M.S., Murayama M., Davis R.J. Journal of Catalysis,

2007, 250, 94-101.

[9] Miller J.T., Kropf A.J., Zha Y., Regalbuto J.R., Delannoy L., Louis C., Bus E., van

Bokhoven J.A. Journal of Catalysis, 2006, 240, 222-234.

[10] Grunwaldt J.D., Kiener C., Wogerbauer C., Baiker A. Journal of Catalysis, 1999, 181,

223-232.

[11] Haruta M. Catalysis Today, 1997, 36, 153-166.

[12] Song W.Y., Ferrandez D.M.P., van Haandel L., Liu P., Nijhuis T.A., Hensen E.J.M.

Acs Catalysis, 2015, 5, 1100-1111.

[13] Hughes M.D., Xu Y.J., Jenkins P., McMorn P., Landon P., Enache D.I., Carley A.F.,

Attard G.A., Hutchings G.J., King F., Stitt E.H., Johnston P., Griffin K., Kiely C.J.

Nature, 2005, 437, 1132-1135.

[14] Biella S., Prati L., Rossi M. Journal of Catalysis, 2002, 206, 242-247.

[15] Whiting G.T., Kondrat S.A., Hammond C., Dimitratos N., He Q., Morgan D.J.,

Dummer N.F., Bartley J.K., Kiely C.J., Taylor S.H., Hutchings G.J. Acs Catalysis,

2015, 5, 637-644.
31
[16] Schubert M.M., Hackenberg S., van Veen A.C., Muhler M., Plzak V., Behm R.J.

Journal of Catalysis, 2001, 197, 113-122.

[17] He H., Dai H.X., Ng L.H., Wong K.W., Au C.T. Journal of Catalysis, 2002, 206, 1-13.

[18] Guo D.J., Jing Z.H. Journal of Power Sources, 2010, 195, 3802-3805.

[19] Qian K., Fang J., Huang W.X., He B., Jiang Z.Q., Ma Y.S., Wei S.Q. Journal of

Molecular Catalysis a-Chemical, 2010, 320, 97-105.

[20] Zaporojtchenko V., Podschun R., Schurmann U., Kulkarni A., Faupel F.

Nanotechnology, 2006, 17, 4904-4908.

[21] Dossi C., Psaro R., Bartsch A., Brivio E., Galasco A., Losi P. Catalysis Today, 1993,

17, 527-535.

[22] Busca G., Centi G., Trifiro F., Lorenzelli V. Journal of Physical Chemistry, 1986, 90,

1337-1344.

[23] Bordes-Richard E. Topics in Catalysis, 2008, 50, 82-89.

[24] Fu Y.C., Shen J.Y. Chemical Communications, 2007, 2172-2174.

[25] Hua Q., Cao T., Gu X.-K., Lu J., Jiang Z., Pan X., Luo L., Li W.-X., Huang W.

Angewandte Chemie International Edition, 2014, 53, 4856-4861.

[26] Chen C., Kosuke N., Murayama T., Ueda W. Chemcatchem, 2013, 5, 2869-2873.

[27] Takacs J.M., Jiang X.T. Current Organic Chemistry, 2003, 7, 369-396.

[28] Keith J.A., Henry P.M. Angewandte Chemie-International Edition, 2009, 48,

9038-9049.

[29] Kotov V., Scarborough C.C., Stahl S.S. Inorganic Chemistry, 2007, 46, 1910-1923.

[30] Vincent J.M., Menage S., Lambeaux C., Fontecave M. Tetrahedron Letters, 1994, 35,

6287-6290.

[31] Choi M., Na K., Kim J., Sakamoto Y., Terasaki O., Ryoo R. Nature, 2009, 461,

246-U120.

[32] Zhan B.Z., White M.A., Sham T.K., Pincock J.A., Doucet R.J., Rao K.V.R.,

Robertson K.N., Cameron T.S. Journal of the American Chemical Society, 2003, 125,

2195-2199.
32
[33] Cundy C.S., Cox P.A. Chemical Reviews, 2003, 103, 663-701.

[34] Davis M.E., Lobo R.F. Chemistry of Materials, 1992, 4, 756-768.

[35] Chang C.D., Silvestri A.J. Journal of Catalysis, 1977, 47, 249-259.

[36] Csicsery S.M. Zeolites, 1984, 4, 202-213.

[37] Boyd G.E., Adamson A.W., Myers L.S. Journal of the American Chemical Society,

1947, 69, 2836-2848.

[38] Kharitonov A.S., Sheveleva G.A., Panov G.I., Sobolev V.I., Paukshtis Y.A.,

Romannikov V.N. Applied Catalysis a-General, 1993, 98, 33-43.

[39] Li X.Y., Ramamurthy V. Journal of the American Chemical Society, 1996, 118,

10666-10667.

[40] Panov G.I., Uriarte A.K., Rodkin M.A., Sobolev V.I. Catalysis Today, 1998, 41,

365-385.

[41] Burton S.G. Trends in Biotechnology, 2003, 21, 543-549.

[42] van Berkel W.J.H., Kamerbeek N.M., Fraaije M.W. Journal of Biotechnology, 2006,

124, 670-689.

[43] Julsing M.K., Cornelissen S., Buhler B., Schmid A. Current Opinion in Chemical

Biology, 2008, 12, 177-186.

[44] Roduner E., Kaim W., Sarkar B., Urlacher V.B., Pleiss J., Glaser R., Einicke W.D.,

Sprenger G.A., Beifuss U., Klemm E., Liebner C., Hieronymus H., Hsu S.F., Plietker

B., Laschat S. Chemcatchem, 2013, 5, 82-112.

[45] Saarikoski E., Rautkoski H., Rissanen M., Hartman J., Seppala J. Journal of Applied

Polymer Science, 2014, 131,

[46] Zhang M.Y., Cheng Z.Q., Liu M.Z., Zhang Y.Q., Hu M.J., Li J.F. Journal of Applied

Polymer Science, 2014, 131,

[47] Ben Fradj A., Lafi R., Ben Hamouda S., Gzara L., Hamzaoui A.H., Hafiane A.

Journal of Photochemistry and Photobiology a-Chemistry, 2014, 284, 49-54.

[48] Kampe P., Giebeler L., Samuelis D., Kunert J., Drochner A., Haass F., Adams A.H.,

Ott J., Endres S., Schimanke G., Buhrmester T., Martin M., Fuess H., Vogel H.
33
Physical Chemistry Chemical Physics, 2007, 9, 3577-3589.

[49] Jekewitz T., Blickhan N., Endres S., Drochner A., Vogel H. Catalysis

Communications, 2012, 20, 25-28.

[50] Sarkar B., Pendem C., Konathala L.N.S., Tiwari R., Sasaki T., Bal R. Chemical

Communications, 2014, 50, 9707-9710.

[51] Lin M.M. Applied Catalysis a-General, 2001, 207, 1-16.

[52] Bettahar M.M., Costentin G., Savary L., Lavalley J.C. Applied Catalysis a-General,

1996, 145, 1-48.

[53] Burrington J.D., Kartisek C.T., Grasselli R.K. Journal of Catalysis, 1980, 63,

235-254.

[54] Burrington J.D., Kartisek C.T., Grasselli R.K. Journal of Catalysis, 1984, 87,

363-380.

[55] Brazdil J.F., Suresh D.D., Grasselli R.K. Journal of Catalysis, 1980, 66, 347-367.

[56] MoroOka Y., Ueda W. Advances in Catalysis, Vol 40, 1994, 40, 233-273.

[57] Burrington J.D., Kartisek C.T., Grasselli R.K. Journal of Catalysis, 1983, 81,

489-498.

[58] Adams C.R., Jennings T.J. Journal of Catalysis, 1964, 3, 549-558.

[59] Concepcion P., Botella P., Nieto J.M.L. Applied Catalysis a-General, 2004, 278,

45-56.

[60] Bruckman K., Haber J., Wiltowski T. Journal of Catalysis, 1987, 106, 188-201.

[61] Chaudhari C.S., Sable S.S., Gurav H., Kelkar A.A., Rane V.H. Journal of Natural Gas

Chemistry, 2010, 19, 593-599.

[62] Ueda W., Vitry D., Kato T., Watanabe N., Endo Y. Research on Chemical

Intermediates, 2006, 32, 217-233.

[63] Ueda W., Suzuki Y. Chemistry Letters, 1995, 541-542.

[64] Landi G., Lisi L., Volta J.C. Catalysis Today, 2004, 91-2, 275-279.

[65] Kum S.S., Jo B.Y., Moon S.H. Applied Catalysis a-General, 2009, 365, 79-87.

[66] Mestl G., Margitfalvi J.L., Vegvari L., Szijjarto G.P., Tompos A. Applied Catalysis
34
a-General, 2014, 474, 3-9.

[67] Ueda W., Endo Y., Watanabe N. Topics in Catalysis, 2006, 38, 261-268.

[68] Millet J.M.M., Roussel H., Pigamo A., Dubois J.L., Jumas J.C. Applied Catalysis

a-General, 2002, 232, 77-92.

[69] Li W., Oshihara K., Ueda W. Applied Catalysis a-General, 1999, 182, 357-363.

[70] Mizuno N., Tateishi M., Iwamoto M. Applied Catalysis a-General, 1995, 128,

L165-L170.

[71] Thorsteinson E.M., Wilson T.P., Young F.G., Kasai P.H. Journal of Catalysis, 1978,

52, 116-132.

[72] d'Alnoncourt R.N., Csepei L.I., Havecker M., Girgsdies F., Schuster M.E., Schlogl R.,

Trunschke A. Journal of Catalysis, 2014, 311, 369-385.

[73] Zheng Y.G., Chen X.L., Shen Y.C. Chemical Reviews, 2008, 108, 5253-5277.

[74] Wang F., Dubois J.L., Ueda W. Journal of Catalysis, 2009, 268, 260-267.

[75] Thanasilp S., Schwank J.W., Meeyoo V., Pengpanich S., Hunsom M. Journal of

Molecular Catalysis a-Chemical, 2013, 380, 49-56.

[76] Pestana C.F.M., Guerra A.C.O., Ferreira G.B., Turci C.C., Mota C.J.A. Journal of the

Brazilian Chemical Society, 2013, 24, 100-105.

[77] Possato L.G., Cassinelli W.H., Garetto T., Pulcinelli S.H., Santilli C.V., Martins L.

Applied Catalysis a-General, 2015, 492, 243-251.

[78] Chieregato A., Soriano M.D., Garcia-Gonzalez E., Puglia G., Basile F., Concepcion P.,

Bandinelli C., Nieto J.M.L., Cavani F. Chemsuschem, 2015, 8, 398-406.

[79] Decolatti H.P., Dalla Costa B.O., Querini C.A. Microporous and Mesoporous

Materials, 2015, 204, 180-189.

[80] Liu L.C., Wang B., Du Y.H., Borgna A. Applied Catalysis a-General, 2015, 489,

32-41.

[81] Omata K., Izumi S., Murayama T., Ueda W. Catalysis Today, 2013, 201, 7-11.

[82] Feng X.Z., Yao Y., Su Q., Zhao L., Jiang W., Ji W.J., Au C.T. Applied Catalysis

B-Environmental, 2015, 164, 31-39.


35
[83] Haider M.H., D'Agostino C., Dummer N.F., Mantle M.D., Gladden L.F., Knight D.W.,

Willock D.J., Morgan D.J., Taylor S.H., Hutchings G.J. Chemistry-a European

Journal, 2014, 20, 1743-1752.

[84] Andrushkevich T.V. Catal.Rev.-Sci.Eng., 1993, 35, 213-259.

[85] Drochner A., Kampe P., Menning N., Blickhan N., Jekewitz T., Vogel H. Chemical

Engineering & Technology, 2014, 37, 398-408.

[86] Ai M. Journal of Catalysis, 1977, 49, 313-319.

[87] Seiyama T., Sakamoto T., Aso I., Egashira M. Journal of Catalysis, 1972, 24, 76-&.

[88] Sadakane M., Kodato K., Kuranishi T., Nodasaka Y., Sugawara K., Sakaguchi N.,

Nagai T., Matsui Y., Ueda W. Angewandte Chemie-International Edition, 2008, 47,

2493-2496.

[89] Sadakane M., Watanabe N., Katou T., Nodasaka Y., Ueda W. Angewandte

Chemie-International Edition, 2007, 46, 1493-1496.

[90] Pyrz W.D., Blom D.A., Sadakane M., Kodato K., Ueda W., Vogt T., Buttrey D.J.

Chemistry of Materials, 2010, 22, 2033-2040.

[91] Pyrz W.D., Blom D.A., Sadakane M., Kodato K., Ueda W., Vogt T., Buttrey D.J.

Proceedings of the National Academy of Sciences of the United States of America,

2010, 107, 6152-6157.

[92] Ishikawa S., Murayama T., Ohmura S., Sadakane M., Ueda W. Chemistry of

Materials, 2013, 25, 2211-2219.

[93] Konya T., Katou T., Murayama T., Ishikawa S., Sadakane M., Buttrey D., Ueda W.

Catalysis Science & Technology, 2013, 3, 380-387.

[94] Qiu C.T., Chen C., Ishikawa S., Murayama T., Ueda W. Topics in Catalysis, 2014, 57,

1163-1170.

[95] Sadakane M., Endo K., Kodato K., Ishikawa S., Murayama T., Ueda W. European

Journal of Inorganic Chemistry, 2013, 1731-1736.

[96] Grasselli R.K., Lugmair C.G., Volpe A.F., Andersson A., Burrington J.D. Catalysis

Letters, 2008, 126, 231-240.


36
[97] Haggblad R., Wagner J.B., Deniau B., Millet J.M.M., Holmberg J., Grasselli R.K.,

Hansen S., Andersson A. Topics in Catalysis, 2008, 50, 52-65.

[98] Korovchenko P., Shiju N.R., Dozier A.K., Graham U.M., Guerrero-Perez M.O.,

Guliants V.V. Topics in Catalysis, 2008, 50, 43-51.

[99] Sankaranarayanan T.M., Ingle R.H., Gaikwad T.B., Lokhande S.K., Raja T., Devi

R.N., Ramaswamy V., Manikandan P. Catalysis Letters, 2008, 121, 39-51.

[100] Botella P., Dejoz A., Abello M.C., Vazquez M.I., Arrua L., Nieto J.M.L. Catalysis

Today, 2009, 142, 272-277.

[101] Pyrz W.D., Blom D.A., Shiju N.R., Guliants V.V., Vogt T., Buttrey D.J. Catalysis

Today, 2009, 142, 320-328.

[102] Tichy J. Applied Catalysis a-General, 1997, 157, 363-385.

[103] Ai M., Ikawa T. Journal of Catalysis, 1975, 40, 203-211.

[104] Bhat B.R., Choi J.S., Kim T.H. Catalysis Letters, 2007, 117, 136-139.

[105] Fang Z.T., Li Z.J., Kelley M.S., Kay B.D., Li S.G., Hennigan J.M., Rousseau R.,

Dohnalek Z., Dixon D.A. Journal of Physical Chemistry C, 2014, 118, 22620-22634.

[106] de Lucas A., Valverde J.L., Canizares P., Rodriguez L. Applied Catalysis a-General,

1998, 172, 165-176.

[107] de Lucas A., Valverde J.L., Canizares P., Rodriguez L. Applied Catalysis a-General,

1999, 184, 143-152.

[108] Kobayashi T., Guilhaume N., Miki J., Kitamura N., Haruta M. Catalysis Today, 1996,

32, 171-175.

[109] Pierini B.T., Lombardo E.A. Catalysis Today, 2005, 107-08, 323-329.

[110] Zhu Z.Z., Lu G.Z., Guo Y., Guo Y.L., Zhang Z.G., Wang Y.Q., Gong X.Q.

Chemcatchem, 2013, 5, 2495-2503.

[111] Drochner A., Kampe P., Kunert J., Ott J., Vogel H. Applied Catalysis a-General, 2005,

289, 74-83.

[112] Endres S., Kampe P., Kunert J., Drochner A., Vogel H. Applied Catalysis a-General,

2007, 325, 237-243.


37
[113] Safikhani A., Aminfard M. International Journal of Hydrogen Energy, 2014, 39,

2286-2296.

[114] Yang X.L., Dai W.L., Gao R.H., Chen H., Li H.X., Cao Y., Fan K.N. Journal of

Molecular Catalysis a-Chemical, 2005, 241, 205-214.

[115] Coloma F., Bachiller-Baeza B., Rochester C.H., Anderson J.A. Physical Chemistry

Chemical Physics, 2001, 3, 4817-4825.

[116] Byesen K.L., Mathisen K. Catalysis Today, 2014, 229, 14-22.

[117] Boyesen K.L., Kristiansen T., Mathisen K. Physical Chemistry Chemical Physics,

2014, 16, 20451-20463.

[118] Witsuthammakul A., Sooknoi T. Applied Catalysis a-General, 2012, 413, 109-116.

38
Figures and tables

Table 1-1. Important oxidation processes of the chemical industry.

Substrates products Usage

Methanol Formaldehyde Polymers, fine chemicals

Ethene Ethylene oxide PET, ethylene glycol

Propene Acrylic acid Polymers

Propene Propylene oxide Propylene glycol

n-Butane Maleic anhydride THF, unsaturated Polyester

Table 1-2. Oxidation processes employing oxides catalysts.

Substrates Products Catalysts Supporter

Fe2O3-MoO3
CH3OH + 1/2O2 HCHO + H2O -
Fe2O3-MoO3-TiO2

Bi2O3-MoO3
CH3CH=CH2 + O2 CH2=CHCHO + H2O SiO2
CoO-MoO3

CoO-Bi2O3-MoO3
CH3CH=CH2 + 3/2O2 CH2=CHCOOH + H2O SiO2
Sb2O5-V2O5-MoO3

Bi2O3-MoO3
CH3CH=CH2 + 3/2O2 +NH3 CH2=CHCN + 3H2O SiO2
Fe2O3-Sb2O5

MgO-Fe2O3 Al2O3
CH3CH2CH=CH2 + 1/2O2 CH2=CHCH=CH2 + H2O
ZnO-Fe2O3 TiO2

Fe2O3
Al2O3
+ 1/2O2 + H2O MgO-V2O5
SiO2
MgO-Fe2O3

39
Figure 1-1. Number of Publications on selective oxidations in the last decades.[4]

40
Figure 1-2. Classification of oxidizing enzymes. [41]

Figure 1-3. Electrophilic and nucleophilic oxidation processes with different oxygen species

41
Figure 1-4. Catalytic activities of () Orthorhombic, () Trigonal, () Tetragonal, and ()

Amorphous Mo-V-O catalysts.[26]

42
Figure 1-5. HAADF-STEM images of orthorhombic (a), trigonal (b), tetragonal (c), and

amorphous (f) Mo-V-O and structural models for orthorhombic (d), trigonal (e)phases.

43
Figure 1-6. Selective oxidation of acrolein to acrylic acid follows a Mars-van-Krevelen

Mechanism.

44
Figure 1-7. Mechanism of acrolein selective oxidation to acrylic acid. [102]

45
Scheme 1-1. Catalyst used in Wacker reaction.

Scheme 1-2. Methods of acrylic acid production: (I) Ethylenecyanohydrin process; (II)

Acrylonitrile process; (III) Reppe process; and (IV) Ketene-process. [48]

46
Scheme 1-3. Selective oxidation of propylene to acrylic acid with a one-step process.

Scheme 1-4. Selective oxidation of propylene to acrylic acid with a two-step process.

47
Scheme 1-5. Mechanism of the catalytic oxidation of propylene to acrolein on metal mixed

oxides.[52]

48
Scheme 1-6. Mechanism of selective oxidation of propylene to acrolein over bismuth

molybdate.[56, 57]

49
Scheme 1-7. Selective oxidation of propane to acrylic acid.

Scheme 1-8. Proposed oxidation pathway over a Mo-V-Te-Nb-O catalyst. [72]

50
Scheme 1-9. Glycerol production via transesterification of vegetable oils or animal fats with

alcohol.[118]

Scheme 1-10. Transformation of glycerol to acrylic acid via one- and two-step process.[76]

51
52
Chapter 2 Synthesis of crystalline Mo-V-W-O complex oxides with

orthorhombic and trigonal structures for acrolein oxidation to acrylic acid

53
54
2.1 Introduction

Catalytic oxidation is a key technology for converting petroleum-based feed stocks to useful

chemicals. The obtained oxygen-containing products such as alcohols, aldehydes, ketones, or acid

are widely used as raw materials for manufacturing plastics, synthetic fiber, polyesters, etc.[1-3] As for

solid-state oxidation catalysts, transition metal oxides are most used as a key component of various

heterogeneous oxidation catalysts because of their diversity in composition and valence state.

Among various metal oxide and complex metal oxides, Mo-V-based complex metal oxides have long

attracted much attention, back to the sixties in the last century.

Mo-V-based complex metal oxide is very well known as the selective oxidation catalyst of

acrolein to acrylic acid. Acrylic acid is one of the most important chemicals largely employed by the

chemical industry for the production of super absorber, polymer, adhesive, paint, plastic, rubber

synthesis, detergent, etc. Currently all acrylic acid is manufactured via two steps propylene based

oxidation processes, where initial stage is the formation of acrolein followed by oxidation of acrolein

to acrylic acid in which Mo-V-based complex metal oxide is used. During the long period of

optimization of industrial catalysts, a large number of metal promoters (Me) such as W, Cu, Mn, Fe,

Sb, Cr, and Sr were introduced to modify the catalytic performance of these complex oxides. [4] A

brief characteristic of the industrial catalysts is in Table 2-1. In the selective catalysts, Mo is always

richer than V in content and these components are usually in lower oxidation states. Since the

catalysts contain various elements for achieving necessary catalyst performance, materials are

basically X-ray amorphous but show a diffraction of 0.4 nm d-spacing. Though these conventional

investigations apparently facilitated an increase in the efficiency of the catalysts, little has been

provided in terms of the understanding of the real active material and structural phase and underlying

mechanisms so far due to the complexity of the industrial catalysts.

Nevertheless, there are some useful and important researches. The most important is a report by

Andrushkevich[5]. She nicely listed up important characteristics of Mo-V-based oxide catalyst on the

basis of the reported results and information on mixed Mo-V oxides, which are shown in Table 2-1.

One can easily see common features between industrial catalysts and mixed Mo-V oxides in the table.

Her most important conclusion amongst the features is the elemental composition of Mo3VO11 as the
55
catalytically active material. However, detailed information like crystal structure about this material

has been lacking.

After the report by Andrushkevich, there are an appreciable number of reports on the topics of

mixed Mo-V oxide catalysts. However, there seemno results confirming active material phase for the

selective acrolein oxidation. For example, Mestl et al. suggested (MoVW)5O14 with a tetragonal

structure as the active and selective phase[6], while it was concluded by Vogel et al. that only X-ray

amorphous Mo/V mixed oxides contained selective oxidation centers in contrast to the crystalline

samples[7]. Obviously, mixed Mo-V oxide catalysts with tetragonal structure or with amorphous

structure could be a candidate real active material. The situation, however, is still complicated owing

to the diversity of components and the lack of full structural information of the catalysts. One of the

possible ways to reach the real state is to make a catalytically highly active material with simpler

components and well characterized.

Meanwhile, Mitsubishi Chemicals developed a complex Mo-V-O based catalyst, which has an

orthorhombic structural phase, so-called M1, for propane oxidation to acrylic acid and the

ammoxidation of propane to acrylonitrile.[8-11] The catalysts commonly contain various levels of Nb

and Te or Sb which are considered able to affect catalyst activity either through promotion of product

selectivity as with Te or structurally as with Nb. This development not only contributes to realize

catalytic propane ammoxidation processes but also helps to understand the active phase of acrolein

oxidation. More recently, four different pure crystalline Mo-V-O materials assuming trigonal,

orthorhombic, tetragonal, and amorphous systems (denoted as Tri-MoVO, Orth-MoVO,

Tetra-MoVO, and Amor-MoVO, respectively) without containing a third metal elements[12] were

reported and eventually it is found that Tri-MoVO and Orth-MoVO are extremely active and

selective for ethane oxidation and for acrolein oxidation to acrylic acid, of which acrolein oxidation

performance is superior to the industrial catalyst as far as compared under laboratory scale

experiments[13].

HAADF-STEM images of these four materials and crystal structure models of Tri-MoVO and

Orth-MoVO have been reported as shown in Figure 2-1[14, 15]. The materials have a rod-like

morphology along its c- axis direction. All of the complex Mo-V-O based catalysts are layered with
56
a lattice constant of 0.4 nm, in which each layer is comprised of MO6 octahedra and pentagonal

bipyramidal M(M5O27) units[16-19]. The units are arranged forming heptagonal and hexagonal

channels in the a-b plane in Tri-MoVO, Orth-MoVO and Amor-MoVO but without these channels

in Tetra-MoVO. All the above structural characterizations were fully confirmed by HAADF-STEM

images shown in Figure 2-1. Owing to the simplicity of the components and the uniformity of the

structures of these materials, study of the structureactivity relationship between the properties of the

catalysts and their performance is possible.

In this chapter, four different pure crystalline Mo-V-O phase materials and W-containing

analogues were synthesized and tested as catalysts in the selective oxidation of acrolein to acrylic

acid. By these experiments, real active structural phase for acrolein oxidation will be identified.

57
2.2 Experimental

2.2.1 Synthesis of four distinct crystalline Mo-V-O


[12, 13]
The catalysts were synthesized as previously reported . Briefly for Orth-MoVO, a

solution of VOSO4nH2O (n=5.4, 3.28 g, Mitsuwa Chemicals) in deionized water (120 mL) was

added to a solution of (NH4)6Mo7O244H2O (8.82 g, Mo: 50 mmol, Wako) in deionized water (120

mL) with stirring. The mixture was stirred for 10 min and then transferred into an autoclave with a

300 mL-Teflon inner tube and Teflon thin sheet enough length for filling about half of Teflon inner

tube space. The reaction mixture was purged with N2 for 10 min and then hydrothermally treated at

175 C for 48 h. The procedure for Tri-MoVO was the same as that above except that the pH value of

the mixture was adjusted to 2.2 with H2SO4 (2 molL-1). Duration of hydrothermal reaction was 20 h.

As-synthesized Orth-MoVO and Tri-MoVO was purified by the treatment in a solution of oxalic acid

(Wako) (0.4 mol L1, 25ml/ 1g solids) at 60 C for 30 min to remove amorphous impurities.

Synthesis of Tetra-MoVO was done through heat treatment of Orth-MoVO (400 C in air for 2 h and

575 C under an atmosphere of N2 for 2 h). Amor-MoVO was obtained by the same procedure of

Orth-MoVO synthesis except the twofold precursor concentration, without use of the Teflon sheet,

and no N2 bubbling.

2.2.2 Synthesis of W-containing Mo-V-O

Orthorhombic Mo3VW0.25Ox (denoted as Orth-MoVWO) was synthesized with

(NH4)6Mo7O244H2O, (NH4)6[H2W12O40]6H2O, and VOSO4nH2O (n=5.4) under hydrothermal

conditions. Firstly, solution A was obtained with 8.38 g of (NH4)6Mo7O244H2O (Mo: 50 mmol,

Wako) and 0.64 g of (NH4)6[H2W12O40]6H2O (W: 2.5 mmol, Nippon Inorganic Colour & Chemical)

being dissolved in 120 ml of deionized water, and solution B was obtained with 3.28 g of

VOSO4nH2O (n=5.4, 3.28 g, Mitsuwa Chemicals) being dissolved in another 120 ml of deionized

water. Secondly, solution B was poured into solution A under stirring conditions. The other synthetic

procedures are the same as described above. Hydrothermal reaction was conducted at 175 oC for 48 h.

As-synthesized material was treated with oxalic acid solution (0.4 molL-1) at 60 C for 2 times. The

58
treatment was kept for 30 min every time.

Amorphous Mo3VW0.25Ox (denoted as Amor-MoVWO) was also synthesized with the same

precursor with twice high concentration and by the same procedure but without Teflon sheet.

For synthesizing trigonal Mo3VW0.25Ox (denoted as Tri-MoVWO), ethylammonium

trimolybdate (EATM, (CH3CH2NH3)2Mo3O10) was used as a Mo source instead of

(NH4)6Mo7O244H2O used to synthesizing Tri-MoVO. EATM was first prepared as follows. 22.594 g

of MoO3 (Kanto; 0.150 mol) was dissolved in 28.0 mL of 70% ethylamine solution (ethylamine:

0.300 mol, Wako) diluted with 28.0 mL of distilled water. The reason for the addition of distilled

water is to reduce the viscosity of the mixed solutions. After being completely dissolved, the solution

was evaporated under vacuumed condition at 70 C and then solid powder was obtained. The powder

was dried in air at 80 C overnight.

8.995 g of EATM (Mo: 50 mmol) was dissolved in 120 mL of distilled water. Separately, an

aqueous solution of VOSO4 was prepared by dissolving 3.28 g of VOSO4nH2O (n=5.4, 3.28 g,

Mitsuwa Chemicals) in 120 mL of distilled water. The two solutions were mixed at ambient

temperature and stirred for 10 min before the addition of 2.40 g of (NH4)6[H2W12O40]6H2O (W: 9.1

mmol, Nippon Inorganic Colour & Chemical). Then, the obtained mixed solution was introduced

into an autoclave with a 300 mL-Teflon inner vessel and 4000 cm2 of Teflon thin sheet to occupy

about half of Teflon inner vessel space. After being introduced, N2 was fed into the solution in the

tube in order to remove residual oxygen. At this stage, the pH of the solution was 2.4. Then the

hydrothermal reaction was started at 175 C for 48 h under static conditions in an electric oven. Gray

solids formed on the Teflon sheet were separated by filtration, washed with distilled water, and dried

in air at 80 C overnight. Purification with oxalic acid was conducted to the obtained solids in order

to remove amorphous type materials contained as an impurity. To 25 mL aqueous solution (0.4

molL1, 60 C) of oxalic acid (Wako), 1 g of the dried material was added and stirred for 30 min,

then washed with 500 mL of distilled water after filtration.

2.2.3 Characterization

X-ray diffraction patterns were measured by using an X-ray diffractometer (RINT-Ultima III,
59
Rigaku) with Cu K to study their crystalline structure. Scanning electron microscopy (SEM) images

were taken using an electron microscope (JSM-6360LA, JEOL). Scanning transmission electron

microscope (STEM) image and metal element mapping of Mo, W, and V were obtained on an

HD-2000 (HITACHI). The chemical compositions of the catalysts determined by ICP-AES method

with a VISTAPRO apparatus (Varian). N2 adsorption-desorption isotherms measurements were

carried out on an auto-adsorption system (BELSORP MAX, Nippon BELL) using t-plot method to

obtain the external surface area and micropore volume. Before measurements, the samples were

treated at 400 C for 2 h under air and out gassed at 300Cunder vacuum for 2h.

2.2.4 Gas-phase catalytic oxidation

Catalytic oxidation of acrolein to acrylic acid was carried out using a fixed bed stainless tubular

reactor at an atmospheric pressure. The amount of catalyst was 0.25 g (with 2.5 g of carborundum) or

0.125 g (with 2.5 g of carborundum). The catalysts were firstly heated from room temperature to 400

C under N2 of 50 mlmin-1 and were kept for 2 h at this temperature. Then the temperature was

decreased to a desired reaction temperature. The reaction was conducted under two different water

pressures. With higher water pressure, the feed composition was

acrolein/O2/H2O/N2+He=2.3/7.4/25.2/65.1 (mol%), and total flow rate was 107.6 mlmin-1. With

lower water pressure, the fed amount of H2Owas reduced to 12.6mol%, and the total flow rate was

kept constant with the increase of N2 as balance gas. The quantitative analysis was carried out using

three on-line gas chromatographs with columns of Molecular Sieve 13X, Gaskuropack 54, and

Porapak Q. Blank runs showed that no reaction took place without catalysts under the experimental

conditions. Carbon balance was always ca. 96100%.

2.3 Results and discussion

2.3.1 Acrolein oxidation over four different pure crystalline Mo-V-O catalysts and

W-containing analogues

Gas-phase selective oxidation of acrolein to acrylic acid was conducted over crystalline

Mo-V-O catalysts with four different crystal structure phases and W-containing Mo-V-O catalysts.

The results are summarized in Table 2-2 with structural parameters and catalytic performance
60
changes as the function of reaction temperature are illustrated in Figure 2-2. As been reported [13, 20],

there is a large difference in the performance of these four Mo-V-O catalysts with the different crystal

phases. Tri-MoVO was the most active catalyst, giving appreciable conversion of acrolein even at

175 C and achieving almost 100% conversion at 215 C with high acrylic acid selectivity more than

97%. This catalytic performance is surprisingly high and totally not comparable to other reported

data [4, 6, 7]. This catalyst is followed by Orth-MoVO in terms of the conversion as shown in Figure 2-2.

By taking it into account that the external surface area of Tri-MoVO is about three times higher than

that of Orth-MoVO (Table 2-2), the activity of Orth-MoVO giving 68% at 215 C is reasonable and

thus it should be concluded that both Tri-MoVO and Orth-MoVO assume a similar intrinsic catalytic

property for the acrolein selective oxidation. Amor-MoVO, on the other hand, was less active and

gave only 25% conversion at 223 C which is a little higher than those applied to Tri-MoVO and

Orth-MoVO. To reach 95% conversion with the Amor-MoVO catalyst, the reaction temperature

needed to be elevated to nearly 300 C, as can be seen in Figure 2-2.The temperature necessary for 95%

conversion is approximately 90 C higher than that for Tri-MoVO catalyst. Tetra-MoVO which has
[21]
often been examined as an active phase for the reaction was found almost inactive under the

temperature condition below 250 C.

In order to understand the observed clear relationship between crystal structure and catalytic

activity, various determining factors of acrolein oxidation activity should be considered. The four

catalysts having different crystal structures were carefully prepared phase-purely, so that any effects

by impurity materials can be ruled out. One may notice that the Mo/V values of Orth-MoVO and

Tri-MoVO are slightly higher than those of Amor-MoVO and Tetra-MoVO as listed in Table 2-2.

This is a structure requisite and such difference may not cause big difference in the activity. External

surface area should be a direct effect on the conversion as already stated above in the activity

comparison between Orth-MoVO and Tri-MoVO catalysts. In the case of a comparison between

Orth-MoVO and Amor-MoVO catalysts, however, twice amount of Amor-MoVO to compensate the

external surface area of Orth-MoVO could not cover the reaction temperature difference of about 30

C between Orth-MoVO and Amor-MoVO (Figure 2-2). There are no ways to compensate in the case

of Tetra-MoVO. These may allow us to consider the activity-determining factor on the basis of
61
crystal structure straightforwardly. Tri-MoVO, Orth-MoVO, and Amor-MoVO all possess

heptagonal channels and hexagonal channels in their structures, whereas Tetra-MoVO does not. This

simple fact can suggest that the catalytically active sites could exist around the heptagonal channels

and the hexagonal channels. Then next it has to be explained why Amor-MoVO was inferior to

Tri-MoVO and Orth-MoVO. From structural point of view, Amor-MoVO has the disordered

arrangement of the pentagonal units in a-b plane, so that the number ofthe heptagonal channels and

the hexagonal channels should be decreased. This causes lower number of the active sites associated

with the heptagonal channels and the hexagonal channels than those of Tri-MoVO and Orth-MoVO.

This could be the reason why Amor-MoVO had the lower catalytic activity than the other two

catalysts.

The above discussion may suggest that the acrolein oxidation mainly takes place on the surface

of the a-b plane of the catalysts where the heptagonal channels and the hexagonal channels locate and

not on the side surface of the rod-shaped catalysts. In order to clarify which contributes more to the

oxidation activity, the side surface or the section surface of the rod-shaped crystals, unground

Orth-MoVO catalyst was tested for the selective oxidation of acrolein. The acrylic acid selectivity

(96%)of the unground catalyst was almost the same with the ground catalyst but the acrolein

conversion (9.8 %) was far less compared with the ground catalyst (37.8%) in spite of similar external

surface areas (ground; 6.2 m2g-1, unground; 4.0 m2g-1). This fact clearly indicates that the section

surface exposed by the grind treatment is far more active for the selective oxidation of acrolein than

the side surface of the rod-shaped crystals. As a consequence, acrolein oxidation to acrylic acid can

proceed over the external surface of the rod-shaped crystals but dominantly on the section surface of

the rod-shaped crystals.

2.3.2 Acrolein oxidation over W-containing Mo-V-O catalysts

The author succeeded in introducing W in Mo-V-O with keeping the orthorhombic, trigonal, and

amorphous structures by the preparative procedure shown in the experimental section. The W

contents in Orth-MoVWO, Tri-MoVWO,and Amor-MoVWO are determined by ICP-AES method

to be 4.7%, 5.2%, and 5.8% respectively, as shown in Table 2-2, manifesting that W was successfully
62
introduced into the synthesized materials. The crystalline structure of the catalysts was studied by

XRD characterization. Diffraction peaks at 22.2 and 45.4 was observed for all the catalysts, which

indicates that the present catalysts are a kind of layered-type material with a layer lattice distance of

about 0.4 nm (Table 2-2). These two peaks were ascribed to (001) and (002) plane reflections.

Besides these two peaks, diffraction peaks corresponding to the orthorhombic crystal system

emerged in the pattern of Orth-MoVWO at 6.6, 7.9, 9.0, and 27.3 (Cu K), which were ascribed

to the plane of (020), (120), (210), and (630), respectively. These diffraction peaks at the low angel

region less than 10 indicates that Orth-MoVWO was well crystallized along a- and b-axis as well as

c-axis. Similarly, diffraction peaks corresponding to the trigonal crystal system emerged in the

pattern of Tri-MoVWO at 4.6, 8.2, and 9.5(Cu K), which were ascribed to the plane of (100),

(110), and (200), respectively. These are the most distinct difference compared with the amorphous

samples. Amor-MoVWO showed only a broad peak below 10with sharp peaks ascribed to (001)

and (002) plane reflections. These data satisfactory support the introduction of W in the lattice of

orthorhombic and amorphous structures. Further details on the state of W in Orth-MoVWO and

Amor-MoVWO will be reported elsewhere.

The morphology analyses of the present catalysts were conducted through SEM

characterization. Orth-MoVWO presented as rod-shaped crystals, which is similar to that of

Orth-MoVO. The average diameter of Orth-MoVO was about 300 nm, while that of Orth-MoVWO

was apparently smaller to be less than 200 nm. Since the surface area of Orth-MoVWO was quite

high, the actual diameter of Orth-MoVWO rods should be much smaller, which is likely caused by

the rod-segregation of the rod-shaped crystals due to lattice contraction generated by the

incorporation of W. Accompanying the decrease of the rod diameter and the length, the external

surface area of Orth-MoVWO reached 28.9 m2.g-1, which is 4.8 times as high as that of Orth-MoVO

(6.2 m2.g-1, Table 2-2).The rod shape was also observed by STEM characterization in Tri-MoVWO

and Amor-MoVWO. However, the introduction of W in Tri-MoVO and Amor-MoVO had no clear

effect on the external surface area and the rod size did not change obviously by the introduction of W.

Table 2-2 lists the acrolein conversion and product selectivity over three W-containing catalysts,

Orth-MoVWO, Tri-MoVWO, and Amor-MoVWO. It can be clearly seen that the catalytic activity of
63
Orth-MoVWO is higher than that of Orth-MoVO and higher than those of Tri-MoVWO and

Amor-MoVWO..Total conversion of acrolein could be realized at much lower temperature when

Orth-MoVWO is used, which is comparable to Tri-MoVO (Table 2-2). For example, 91% conversion

of acrolein could be obtained at the temperature of 215 C over Orth-MoVWO catalyst, while 68%

conversion was only obtained at the same temperature of 215 C over Orth-MoVO catalyst. On the

contrarily, the addition of W introduction into the trigonal and amorphous type catalysts caused a

negative effect on the conversion as can be seen in Figure 2-2. Nevertheless, nearly 100 %

conversion of acrolein could be obtained at the temperature of 280C over Amor-MoVWO with a bit

higher selectivity to acrylic acid that that over Orth-MoVWO and Tri-MoVWO catalysts.

The extremely high catalytic activity was achieved over Orth-MoVWO catalyst compared to those

over Tri-MoVWO and Amor-MoVWO catalysts. This is obviously due to the largely increased

surface area by the introduction of W, but at the same time this fact suggests that the introduction of

W the introduction may affect the orthorhombic structure material specifically. In fact, the

introduction of W showed the clear negative promotion effect on the catalytic activity for the trigonal

and amorphous system. It is considered that at the present stage that only when W is incorporated into

the orthorhombic crystalline structure with bringing about the increase of active surface area and

structural strain, excellent catalytic activity can be realized at lower temperature. On the contrary, the

negative effect of W addition observed in the Tri-MoVO is presumably due to structural stabilization

by W in the structure with decreasing structural strain and thus active surface area.

2.3.3 Effect of water on the acrolein oxidation over W-containing Mo-V-O catalysts

In order to understand the addition effect of W from other aspect, water addition effect on the

acrolein oxidation was tested. Although the promotion mechanism of added water in reactant feed is

still not clear, it is well accepted that water is able to act as an important promoter for the conversion

of acrolein[22,23]. For example, with Mo/V/W-mixed oxides as catalyst, the conversion of acrolein

decreased with the reduction of fed amount of water at lower temperature[22].

Figure 2-3 illustrates the acrolein conversion and product selectivity as the function of the reaction

temperature over four catalysts, Orth-MoVWO, Orth-MoVO, Tri-MoVWO, Tri-MoVO, under two
64
different water feed concentrations. Table 2-3 also summarizes water feed effect on the conversion

and the selectivityover Orth-MoVWO, Orth-MoVO, Tri-MoVWO, Tri-MoVO, Amor-MoVWO, and

Amor-MoVO catalysts in different catalyst weights. As expected, the acrolein conversion over

Orth-MoVO and Tri-MoVO catalysts appreciably decreased when the water concentration decreased

to half. Higher reaction temperature is required when the water concentration is lower (see arrows in

Figure 2-3). On the other hand, the conversion of acrolein over the W-containing catalysts,

Orth-MoVWO and Tri-MoVWO, showed only small decreases even when the fed amount of water

decreased by half. These results clearly demonstrate that the role of W in the oxidation of acrolein

over Mo-V-O catalysts is to moderate the effect of water.

It should be noted in Table 2-3 that different from the other two catalysts, no clear effect of W on

the water addition effect was observed when the catalysts are amorphous. The author speculate that

the W effect can appear only when W is incorporated into the framework of the orthorhombic or the

trigonal crystalline structure. Obviously further experiments are necessary, which are ongoing.

2.4. Conclusions

The present work on acrolein oxidation catalysts based on Mo-W-VO was conducted in three

different approaches; first, structure-activity relationship by using different crystal phase materials as

catalysts, second, effect of W addition effects in Mo-V-Oon the conversion and third, effect of water

addition depending on the presence of W in the catalyst structures. By combining all collected data in

these three approaches, it was concluded that the network of the pentagonal units forming the

heptagonal and hexagonal channels is indispensable for creating active sites on the surface the

Mo-V-O catalyst and that the a-b plane with the heptagonal and hexagonal channels on the surface is

responsible for acrolein oxidation activity. This active surface formation is assisted by the

W-addition in the case of orthorhombic Mo-V-O catalyst. It was clearly demonstrated that the active

surface containing W can work even under low partial pressure of water in the reactant feed.

Many years have passed since the discovery of active Mo and V-based oxide catalyst for acrolein

selective oxidation to acrylic acid. Now, the present work proposes that crystalline Mo-V-W-O with

orthorhombic or trigonal structure is the real active phase of industrial Mo-V-W-O catalysts for

acrolein oxidation.
65
Reference

[1] Punniyamurthy T, Velusamy S, Iqbal J. Chemical Reviews, 2005, 105, 2329-2364

[2] Centi G, Cavani F, Trifiro F. Selective Oxidation by Heterogeneous Catalysis (Eds.: M. T. Twigg,

M. S. Spencer), Kluwer Academic/Plenum Publishers, New York, 2001

[3] Hodnett BK, Heterogeneous Catalytic Oxidation, Wiley, New York, 2000

[4] Nojiri N, Sakai Y, Watanabe Y, Catalysis Reviews: Science and Engineering, 1995, 31, 145-178

[5] Andrushkevich TV, Catalysis Reviews: Science and Engineering, 1993, 35, 213-259

[6] Dieterle A, Mestl G, Jager J, Uchida Y, Hibst H, Schlogl. Journal of Molecular Catalysis A:

Chemical, 2001, 174, 169-185

[7] Vogel H, Bohling R, Hibst H. Catalysis Letters, 1999, 62, 71-78

[8] Ushikubo T, Oshima K, Kayou A, Umezawa T, Kiyono K, Sawaki I. Mitsubishi Chem. Corp,

1993, Patent EP 529853,

[9] Ushikubo T, Oshima K, Kayou A, Vaarkamp M, Hatano M. Journal of catalysis, 1997, 169,

394-396

[10] Tsuji H, Koyasu Y. Journal of the American Chemical Society, 2002, 124, 5608-5609

[11] Ueda W. Journal of the Japan Petroleum Institute, 2013, 56, 122-132

[12] Konya T, Katou T, Murayama T, Ishikawa S, Sadakane M, Buttrey D, Ueda W. Catalysis Science

& Technology, 2013, 3, 380-387

[13] Chen C, Nakatani K, Murayama T, Ueda W. ChemCatChem, 2013, 5, 2869-2873

[14] Sadakane M, Watanabe N, Katou T, Nodasaka Y, Ueda W. Angewandte Chemie International

Edition, 2007, 46, 1493-1496

[15] Sadakane M, Kodato K, Kuranishi T, Nodasaka Y, Sugawara K, Sakaguchi N, Nagai T, Matsui Y,

Ueda W. Angewandte Chemie International Edition, 2008, 47, 2493-2496

[16] Sadakane M, Yamagata K, Kodato K, Endo K, Toriumi K, Ozawa Y, Ozeki T, Nagai T, Matsui Y,

Sakaguchi N, Pyrz WD, Buttrey D J, Blom D A, Vogt T, Ueda W. Angewandte Chemie

International Edition, 2009, 48, 3782-3786

[17] Sadakane M, Endo K, Kodato K, Ishikawa S, Murayama T, Ueda W. European Journal of


66
Inorganic Chemistry, 2013, 10-11, 1731-1736

[18] Pyrz WD, Blom DA, Sadakane M, Kodato K, Ueda W, Vogt T, Buttrey DJ. Proceedings of the

National Academy of Sciences, 2010, 107, 6152-6157

[19] Pyrz WD, Blom DA, Sadakane M, Kodato K, Ueda W, Vogt T, Buttrey DJ. Chemistry of

Materials, 2010, 22, 2033-2040

[20] Ishikawa S, Yi X, Murayama T, Ueda W. Applied Catalysis a: General, 2014, 474:1017

[21] Mestl G. Topics in Catalysis, 2006, 38, 69-82

67
Figures and tables

Table 2-1 Typical features of Mo-V-O catalysts for gas-phase acrolein oxidation

Features Industrial catalysts Mixed Mo-V oxide

Composition Mo/V = 2~8 Mo3VO11-x

Additives Cu, W, Fe, Sb etc. none

XRD X-ray amorphous Phase mixture

(with peak of d=0.4 nm) (with peak of d=0.4 nm)

Valence state Highly reduced Mo6+, V4+

IR band (cm-1) not available 870(st), 917(sh), 824(sh)

68
Table 2-2 Structural information of Mo-V-O catalysts and W-containing analogues, and their

catalytic performance in acrolein oxidation

Elemental External Reac.


Lattice parameter (nm) Acrolein Selectivity (%)
Catalyst composition (%)a surface temp.
conv.(%)c
Mo V W a b c area(m2/g)b (oC) AAd AcOHe COx

Orth-MoVO 75.6 24.4 0 2.108 2.657 0.3997 6.2 215 68.2 98.1 0.1 1.8

Orth-MoVWO 69.8 25.5 4.7 2.102 2.648 0.3994 28.9 215 91.1 97.1 0.3 2.6

Tri-MoVO 75.8 24.2 0 2.127 0.4011 16.3 215 99.3 97.3 0.3 2.4

Tri-MoVWO 74.1 20.7 5.2 2.129 0.3996 18.0 215 53.9 97.5 0.2 2.3

Amor-MoVO 72.2 27.8 0 0.3996 3.0 223 25.2 98.4 0.0 1.6

Amor-MoVWO 71.9 22.3 5.8 0.3994 3.4 223 16.0 98.8 0.0 1.2

a
Determined by ICP-AES
b
Measured by N2adsorption and determined by t-plot
c
Reaction conditions: 107.6 mLmin-1 of reactant gas with the composition of acrolein/O2/H2O/N2+

He = 2.3/7.4/25.2/65.1(mol%) was fed in 0.25 g of catalysts diluted with carbonrundum (2.5 g)


d
Acrylic acid
e
Acetic acid

69
Table 2-3 Water-addition effect on acrolein oxidation over Orth-MoVO, Orth-MoVWO, Tri-MoVO,

Tri-MoVWO, Amor-MoVO, Amor-MoVWO catalysts.

Catalyst Water Reac. Acrolein Selectivity

Catalyst weight content temp. Conv. /%

/g /mol% /oC /% AAc AcOHd COx

Orth-MoVO 0.25 25.2a 215 68.2 98.1 0.1 1.8

Orth-MoVO 0.25 12.6b 215 27.0 98.2 0.1 1.7

Orth-MoVWO 0.15 25.2 215 73.6 97.5 0.2 2.3

Orth-MoVWO 0.15 12.6 215 58.9 97.6 0.2 2.2

Tri-MoVO 0.10 25.2 215 50.7 98.1 0.1 1.8

Tri-MoVO 0.10 12.6 215 29.2 97.9 0.1 2.0

Tri-MoVWO 0.25 25.2 215 53.9 97.5 0.2 2.3

Tri-MoVWO 0.25 12.6 215 44.9 97.5 0.2 2.3

Amor-MoVO 0.25 25.2 245 67.4 98.7 0.1 1.2

Amor-MoVO 0.25 12.6 245 58.8 98.8 0.1 1.1

Amor-MoVWO 0.25 25.2 245 47.2 98.9 0.1 1.0

Amor-MoVWO 0.25 12.6 245 45.2 99.0 0.1 0.9

a
Reaction conditions: 107.6 mLmin-1 of reactant gas with the composition of

acrolein/O2/H2O/N2+He=2.3/7.4/25.2/65.1 (mol%) was fed in 0.25 g of catalysts diluted with carbonrundum

(2.5 g).
b
Reaction conditions: 107.6 mLmin-1 of reactant gas with the composition of

acrolein/O2/H2O/N2+He=2.3/7.4/12.6/77.7 (mol%).
c
Acrylic acid.
d
Acetic acid.

70
Figure 2-1. HAADF-STEM images of (a) orthorhombic, (b) trigonal, (c) tetragonal, and (f)

amorphous Mo-V-O and structural models for (d) orthorhombic, (e) trigonal phases.

71
3.0
Selectivity to COx/%

2.5
2.0
1.5
1.0
0.5
0.0
160 180 200 220 240 260 280 300 320
100
Selectivity to AA/%

98

96

94

92

90
160 180 200 220 240 260 280 300 320
100
Conversion/%

80

60

40

20

0
160 180 200 220 240 260 280 300 320
o
Reaction temperature/ C

Figure 2-2 Acrolein oxidation over Orth-MoVO (filled square), Orth-MoVWO (square), Tri-MoVO

(filled diamond), Tri-MoVWO (diamond), Amor-MoVO (filled circle), and Amor-MoVWO (circle)

catalysts under H2O: 25.2 mol%.

72
Selectivity to COx/%
3.0
2.5
2.0
1.5
1.0
0.5
0.0
170 180 190 200 210 220 230 240 250 260
Selectivity to AA/%

100

98

96

94

92

90
170 180 190 200 210 220 230 240 250 260
100
Conversion/%

80

60

40

20

0
170 180 190 200 210 220 230 240 250 260
o
Reaction temperature/ C

Figure 2-3 Water-addition effect on acrolein oxidation over Orth-MoVO (H2O filled: square: 25.2

mol%, square: 12.6 mol%), Orth-MoVWO (H2O filled: circle: 25.2 mol%, circle: 12.6 mol%),

Tri-MoVO (H2O filled: triangle: 25.2 mol%, triangle: 12.6 mol%), and Tri-MoVWO (H2O filled: star:

25.2 mol%, star: 12.6 mol%) catalysts (0.25 g).

73
74
Chapter 3 Effects of W on the selective oxidation of acrolein to acrylic acid

over crystalline Mo-V-W-O catalysts

75
76
3.1 Introduction

Mo-V-based complex oxides catalysts have a long history of use in selective oxidation and

ammoxidation of light alkanes since the 1960s.[1-5] Numerous metal components such as Te, Nb, Cu,

W, and Fe have been applied to optimize the structure and catalytic performance. [6-10] Additional

metals modified the structure and improved the catalytic activity; however, multiple components

increased the complexity of structure and composition. W as one of the most effective promoters is

widely employed in industrial catalysts. However, the effect of W is so far poorly understood. As

described in chapter 1, W-containing Mo-V-O catalysts with crystalline structure were synthesized,

and a positive effect of W on acrolein oxidation to acrylic acid was observed. Further study on these

crystalline Mo-V-W-Os will give advantages to reveal the structure-activity relationship.

For the molecular sieve-type porous materials, the incorporation of a transition metal into the

framework could modify properties such as acidity, thermal stability, and especially catalytic

activity.[11, 12] However, introduction of heteroatom usually affects the normal crystalline growth

process, and it is a challenge to introduce a heteroatom metal while maintaining the crystal structure.

From the view point of structure, the reported crystalline Mo-V-O complex oxides are layered

material with micro porosity.[13-15] Each layer contains pentagonal {Mo6O21} building units and

{MO6} (M = Mo, V) octahedra that are arranged to form heptagonal and hexagonal channels and

then grow in the direction of the c-axis. The introduction of W into a crystalline structure will affect

the crystal growth and modify the structural properties which will thus impact the catalytic

performance.

In this chapter, orthorhombic and trigonal phase MoVWO complex oxides with various

amounts of W were synthesized by a hydrothermal method. W is successfully incorporated into the

framework of the orthorhombic and trigonal structure. The introduction of W into orthorhombic

structure leads to a rod segregation effect which contributes to increase the external surface area.

Caused by increases of the external surface area, more active phase exposed, and thus catalytic

activity was obviously enhanced.

77
3.2 Experimental

3.2.1 Synthesis of three distinct Mo-V-O complex oxides

The catalysts were synthesized according to a previous report.[15, 16] For orthorhombic Mo-V-O

(denoted as Orth-MoVO-W0), a solution of VOSO4nH2O (n=5.4, 3.28 g, Mitsuwa Chemicals) in

deionized water (120 mL) was added to a solution of (NH4)6Mo7O244H2O (8.82 g, Mo: 50 mmol,

Wako) in deionized water (120 mL) with stirring. The mixture was stirred for 10 min and then

transferred into an autoclave with a Teflon inner tube and a Teflon sheet enough length for filling

about half of Teflon inner tube space. This sheet is necessary for formation of well-crystallized

sample, because solids are formed on the sheet. The mixture was purged with N2 for 10 min (pH=3.2)

and then hydrothermally treated at 175 oC for 48 h. The procedure for synthesis of trigonal Mo-V-O

(denoted as Tri-MoVO-W0) was the same as that for Orth-MoVO-W0 except that the pH value of the

mixture was adjusted to 2.2 with H2SO4 (2 molL1). As-synthesized Orth-MoVO-W0 and

Tri-MoVO-W0 were purified by treatment in a solution of oxalic acid (Wako) (0.4 molL-1, 25ml/ 1g

solids) at 60 oC for 30 min to remove amorphous impurities. Amorphous Mo-V-O (denoted as

Amor-MoVO-W0) was obtained by the same procedure as that for Orth-MoVO-W0 synthesis but

with two-fold higher precursor concentrations, without the use of a Teflon sheet, and no N2 bubbling.

3.2.2 Synthesis of amorphous and orthorhombic Mo-V-W-O complex oxides

Orthorhombic Mo-V-W-O catalysts (denoted as Orth-MoVWO) with various contents of W

were synthesized with (NH4)6Mo7O24H2O, VOSO4nH2O (n=5.4) and (NH4)6[H2W12O40]6H2O

under a hydrothermal condition. Firstly, solution A was obtained with 8.82 g of (NH4)6Mo7O244H2O

(Mo: 50 mmol, Wako) and a certain amount of (NH4)6[H2W12O40]6H2O (Nippon Inorganic

Colour & Chemical) (with adjustment of W content to 2.5 at.-%, 5.0 at.-%, 7.5 at.-%, and 10 at.-% in

the synthesized Mo-V-W-O, determined by ICP-AES method) being dissolved in 120 mL of

deionized water, and solution B was obtained with 3.28 g of VOSO4nH2O (n=5.4, Mitsuwa

Chemicals) being dissolved in another 120 mL of deionized water. Secondly, solution B was poured

into solution A under a stirring condition. The obtained mixed solution was introduced into an

autoclave with a Teflon sheet in the Teflon inner vessel. The reaction mixture was purged with N2 for
78
10 min (pH=3.2) and then hydrothermally treated at 175 oC for 48 h. As-synthesized Orth-MoVWO

catalysts (denoted as Orth-MoVO-W2.5, Orth-MoVO-W5.0, Orth-MoVO-W7.5, and

Orth-MoVO-W10, respectively) were treated with oxalic acid solution (0.4 mol L-1) 2 times at 60 oC.

The treatment was maintained for 30 min each time.

Amorphous Mo-V-W-O catalysts (denoted as Amor-MoVWO) with various contents of W

(denoted as Amor-MoVO-W2.5, Amor-MoVO-W5.0, Amor-MoVO-W7.5, and Amor-MoVO-W10,

respectively) were also synthesized with the same precursor at a two-fold higher concentration and

by the same procedure but without a Teflon sheet.

3.2.3 Synthesis of trigonal Mo-V-W-O complex oxides

A trigonal Mo-V-W-O (denoted as Tri-MoVWO) catalyst was obtained by the same procedure

as that for Orth-MoVWO synthesis except that ethylammonium trimolybdate (EATM:

(CH3CH2NH3)2Mo3O10) was used as a Mo source instead of (NH4)6Mo7O244H2O. The synthesis

procedure for EATM was as follows. 22.594 g of MoO3 (0.15 mol, Kanto) was dissolved in 28.0 mL

of 70% ethylamine solution (ethylamine: 0.30 mol, Wako) diluted with 28.0 mL of deionized water

followed by the stirring for 30 min. After being completely dissolved, the solution was evaporated

under vacuumed condition of P/P0= 0.03 at 70 oC for 30 min and white solid was obtained. Time

was measured from the time that white powder was started to precipitate. The powder was dried in

air at 80 oC overnight. As-synthesized materials were denoted as Tri-MoVO-W2.5, Tri-MoVO-W5.0,

Tri-MoVO-W7.5, and Tri-MoVO-W10, respectively.

3.2.4 Characterization

X-ray diffraction patterns were measured by using an X-ray diffractometer (RINT-Ultima III,

Rigaku) with Cu K to study their crystalline structure. Crystallite Size Broading Analysis of

Rietveld program with Powder Reflex (Material Studio 5.5.3, Accelrys) was utilized to fit the

experimental XRD pattern of Orth-MoVWO. Raman spectra (inVia Reflex Raman spectrometer,

RENISHAW) were taken in air on a static sample with Ar laser power. Scanning transmission

electron microscopy (STEM) images and metal element mapping of Mo, W, and V were obtained on
79
an HD-2000 (HITACHI). Transmission electron microscopy (TEM) images were taken with a 200

kV transmission electron microscope (JEOL JEM-2010F). The chemical composition of the catalysts

was determined by the ICP-AES method with a VISTA-PRO apparatus (Varian). N2

adsorption-desorption isotherm measurements were carried out on an auto-adsorption system

(BELSORP MAX, Nippon BELL) to obtain the external surface area and micropore volume using

the t-plot method. Prior to adsorption measurements, the samples were treated at 400 oC for 2 h under

air and out-gassed at 300 oC under vacuum for 2 h.

3.2.5 Gas-phase catalytic oxidation

Gas phase catalytic oxidation of acrolein to acrylic acid was performed using a fixed bed

stainless tubular reactor at atmospheric pressure. The catalysts (0.25 g) were firstly ground for 5

minutes, then diluted with 2.5 g Carborundum and pretreated at 400 oC under N2 of 50 mL min-1 for 2

h. Reactant gas was conducted with change in the water content from 0 to 25.2 vol.-% while keeping

other feeding gases constant: acrolein= 2.5 mL min-1, O2= 8 mL min-1 and N2 balance (total: 107.6

mL min-1). Quantitative analysis was performed using three on-line gas chromatographs with

columns of Molecular Sieve 13X, Gaskuropack 54, and Porapak Q. Blank runs showed that no

reaction took place without catalysts under the experimental conditions. Carbon balance was always

over 95%, and selectivity was calculated on the basis of products of sum.

80
3.3 Results and discussion

3.3.1 Structural characterization of Mo-V-W-O catalysts

3.3.1.1 Location of W in Mo-V-W-O

Table 3-1 shows the chemical compositions of the Mo-V-W-O complex oxides determined by

the ICP-AES method. Atomic ratios of tungsten in the Amor-, Orth- and Tri-MoVWO groups were

adjusted approximately to 2.5, 5.0, 7.5, and 10 at.-%, respectively. Different Mo and V atomic ratios

derived from the distinction of structure requisite. Element mapping of Mo, V and W of Amor-, Orth-

and Tri-MoVO-W2.5 showed that metal elements were distributed evenly along the rod-shaped

materials (Figure 3-1). In the Raman spectra (Figure 3-2), the main band at 872 cm-1 that was

ascribed to pentagonal units gradually shifted to a low wave number.[16] All of the results indicated

that W was successfully introduced into the three different structures.

The crystalline structure of the materials was investigated by XRD characterization (Figure 3-3).

Diffraction peaks at 22.2o and 45.4 o were ascribed to (001) and (002) plane reflections, respectively.

Concentrated at the (001) peak, an obvious peak shift in a high angle direction was observed in all of

the three different types of Mo-V-W-O, after being calibrated with silicon as an internal standard

(Figure 3-3). With increase of W content, the lattice parameter decreased gradually (Table 3-1),

implying that lattice contraction happened with the incorporation of W. Based on the above results,

W was considered to be incorporated into the framework. Orthorhombic, trigonal and amorphous

Mo-V-O contained the same pentagonal {Mo6O21} units and {MO6} (M=Mo, V) octahedra. The

pentagonal unit was constructed with Mo only, and octahedra contained V as well as Mo. [16, 17] It is

notable that, as shown in Table 3-1, with an increase of tungsten content, the ratio of V in Mo-V-W-O

decreased gradually, strongly suggesting that tungsten replaced not only Mo but also the V ions.

Therefore, there was a high probability that W formed {WO6} octahedra and acted as linkers

connecting {Mo6O21} pentagonal units.

3.3.1.2 Influence of W on Mo-V-O structure

In the XRD pattern of Amor-MoVWO (Figure 3-2-I), there were two broad peaks centered at 8o

and 27o, implying that Amor-MoVWO was only crystalline along the c-axis, while a disordered
81
arrangement of pentagonal {Mo6O21} and {MO6} octahedra was formed in the a-b plane. This is the

most distinct difference from the orthorhombic and trigonal structures. With an increase of W content,

the XRD patterns of Amor-MoVWO barely changed except for a shift of the (001) peak in a higher

angle direction. This revealed that although lattice contraction occurred, the disordered arrangement

in the a-b plane was not affected by the introduction of W. Therefore, crystal size (Figure 3-7-I and

Table 3-2) and external surface area (Table 3-1) of Amor-MoVWO slightly changed.

In the XRD pattern of Orth-MoVWO (Figure 3-3-II), main diffraction peaks corresponding to

the orthorhombic structure emerged at 6.6o, 7.9o, 9.0o and 27.3o, etc., which were ascribed to the

planes of (020), (120), (210), and (630), respectively.[18] The emergence of diffraction peaks at a low

angle below 10o indicated that Orth-MoVWO was well-crystallized along the a- and b-axes.

However, with an increase of W content, the rod-segregation effect proceeded. That could be clearly

observed from the TEM image (Figure 6). Orth-MoVWO crystals partially split into smaller rods,

and those nanoscale rods had sizes of only several tens of nanometers or even smaller (Average

crystal size was calculated and is shown in Figure 3-7-II and Table 3-2). For crystalline materials,

lattice expansion or contraction usually results in an unstable structure. Lattice contraction occurred

with the incorporation of W, and that might have caused cleavage of Mo-O and V-O bonds and

decrease in the long-range order of the a-b plane, thus facilitating dehiscence. The diffraction peaks

(especially below 10o) of Orth-MoVWO decreased and broadened gradually with increase of W

content. Rietveld analysis using the Crystallite Size Broadening procedure was carried out, and

results are shown in Figure 3-4. It was confirmed that broadening of diffraction peaks was not caused

by the formation of an amorphous phase but by the dehiscence of orthorhombic crystals. Although

the addition of W affected the crystallinity of the orthorhombic structure, decrease of crystal size

contributed to a larger external surface area (Table 3-1), and more active sites were exposed, which is

particularly important for the activity of acrolein oxidation to acrylic acid.[15] When W content rose to

10 at.-%, external surface area increased to 35.5 m2 g-1, almost 6-times larger than that of the

well-crystallized Orth-MoVO-W0.

Different from the synthesis process of Amor- and Orth-MoVWO, an organic Mo source

(CH3CH2NH3)2Mo3O10 (EATM) was used to synthesize Tri-MoVWO (Section 2.3). As noted above,
82
additional metal ions usually affect the self-assembly process. In the trigonal structure, there is a

linker unit of triple-octahedron and the occupancy of this triple-octahedron is only 0.6, while in the

orthorhombic structure, the occupancy of quintuple-octahedron (instead of triple-octahedron) is

1.0.[13, 14] This revealed that the trigonal structure is not as stable as the orthorhombic structure.

Therefore, simply adding W and other metal precursors into the hydrothermal process only resulted

in amorphous phase instead of trigonal structure. The diameter of the heptagonal channel is about 0.4

nm, thus being limited to counter cation with a small molecular size such as an ammonium cation

with a size of 0.28 nm. In the self-assembly process, an ethylamine cation with a larger size than that

of an ammonium cation balanced the trigonal framework and played an important role as a

structure-directing agent and stabilizer for the trigonal structure, and it could be easily removed by

calcination in air. In the XRD pattern of Tri-MoVWO (Figure 3-3-III), peak intensity of

W-containing trigonal Mo-V-W-O was higher than that of Tri-MoVO-W0, suggesting that the use of

EATM as Mo source can provide trigonal Mo-V-W-O crystals with better crystallinity (average

crystal size shown in Figure 3-7-III and Table 3-2). Due to the good crystallinity of Tri-MoVWO,

only a slight difference in the external surface area of Tri-MoVWO was observed (Table 3-1).

3.3.2 Catalytic activity for oxidation of acrolein to acrylic acid over Mo-V-W-O

Acrylic acid has become a widely used chemical in recent years because of its extensive

applications in super absorbent materials, coatings and additives in textile production.[19-21] It can be

acknowledged that the crystalline structure was maintained and a much larger external surface was

achieved for Orth-MoVWO, which was expected to show excellent performance in the selective

oxidation of acrolein to acrylic acid.

Catalytic performance without H2O feeding was investigated in order to eliminate the effect of

H2O (Figure 3-8, Figure 3-9 Figure 3-10 and Figure 3-11). Crystalline Mo-V-W-O catalysts had more

active sites because the a-b plane of orthorhombic and trigonal structures was constructed into a high

degree of ordered arrangement. Therefore, for acrolein oxidation, crystalline Mo-V-W-O catalysts

showed much higher activity than amorphous catalysts did, while selectivity to acrylic acid was

almost the same. Figure 3-5 shows reaction temperatures of 50% acrolein conversion over Amor-,
83
Orth-, and Tri-MoVWO catalysts under different H2O partial pressure and W content conditions.

There is no doubt that water acts as an important promoter for the conversion of acrolein. And under

the same water feeding condition, Orth- and Tri-MoVWO catalysts always showed higher catalytic

activity than that of Amor-MoVWO. Moreover, a positive effect on acrolein conversion was

observed when W was incorporated into the orthorhombic structure. However, in the case of trigonal

and amorphous structures, W did not show any positive effect because of the slight changes in

structure and morphology. The highest catalytic activity was achieved over the Orth-MoVO-W7.5

catalyst, which had a crystalline structure and large external surface area. Therefore, crystalline

structure is the guarantee for good catalytic performance, and external surface area is another

important factor for oxidation of acrolein to acrylic acid.

3.4 Conclusions

The author succeeded in introducing W into the trigonal structure using EATM as Mo source.

For Orth-MoVWO, tungsten acted as a structural promoter that resulted in a rod-segregation effect

and increase of external surface area. Crystalline Orth- and Tri-MoVWO achieved showed very high

catalytic activity compared with that of Amor-MoVWO. It was found that crystalline structure and

external surface area were responsible for good catalytic activity.

84
References

[1] Tu X.L., Furuta N., Sumida Y., Takahashi M., Niiduma H. Catalysis Today, 2006, 117,

259-264.

[2] d'Alnoncourt R.N., Csepei L.I., Havecker M., Girgsdies F., Schuster M.E., Schlogl R.,

Trunschke A. Journal of Catalysis, 2014, 311, 369-385.

[3] Amakawa K., Kolen'ko Y.V., Villa A., Schuster M.E., Csepei L.I., Weinberg G., Wrabetz S.,

d'Alnoncourt R.N., Girgsdies F., Prati L., Schlogl R., Trunschke A. Acs Catalysis, 2013, 3,

1103-1113.

[4] Millet J.M.M., Roussel H., Pigamo A., Dubois J.L., Jumas J.C. Applied Catalysis a-General,

2002, 232, 77-92.

[5] Drochner A., Kampe P., Menning N., Blickhan N., Jekewitz T., Vogel H. Chemical

Engineering & Technology, 2014, 37, 398-408.

[6] Mestl G., Margitfalvi J.L., Vegvari L., Szijjarto G.P., Tompos A. Applied Catalysis a-General,

2014, 474, 3-9.

[7] Nakazawa Y., Matsumoto S., Kobayashi T., Kurakami T. Catalyst and method for producing

acrylic acid: United States. August 22, 2013-.

[8] Welker-Nieuwoudt. C.A., Karpov. A., Rosowski. F., Mueller-Engel. K.J., Vogel. H., Drochner.

A., Blickhan. N., Duerr. N., Jekewitz. T., Menning. N., Petzold. T., Schmidt. S. Process for

heterogeneously catalyzed gas phase partial oxidation of (meth)acrolein to (meth)acrylic acid:

United States. January 16, 2014-.

[9] Ushikubo T., Oshima K., Kayou A., Vaarkamp M., Hatano M. Journal of Catalysis, 1997, 169,

394-396.

[10] Tsuji H., Koyasu Y. Journal of the American Chemical Society, 2002, 124, 5608-5609.

[11] Sadakane M., Watanabe N., Katou T., Nodasaka Y., Ueda W. Angewandte

Chemie-International Edition, 2007, 46, 1493-1496.

[12] Sadakane M., Kodato K., Kuranishi T., Nodasaka Y., Sugawara K., Sakaguchi N., Nagai T.,

Matsui Y., Ueda W. Angewandte Chemie-International Edition, 2008, 47, 2493-2496.


85
[13] Chen C., Kosuke N., Murayama T., Ueda W. Chemcatchem, 2013, 5, 2869-2873.

[14] Hartmann M., Kevan L. Chemical Reviews, 1999, 99, 635-663.

[15] Thomas J.M., Raja R., Sankar G., Bell R.G. Nature, 1999, 398, 227-230.

[16] Konya T., Katou T., Murayama T., Ishikawa S., Sadakane M., Buttrey D., Ueda W. Catalysis

Science & Technology, 2013, 3, 380-387.

[17] Sadakane M., Yamagata K., Kodato K., Endo K., Toriumi K., Ozawa Y., Ozeki T., Nagai T.,

Matsui Y., Sakaguchi N., Pyrz W.D., Buttrey D.J., Blom D.A., Vogt T., Ueda W. Angewandte

Chemie-International Edition, 2009, 48, 3782-3786.

[18] Ueda W., Vitry D., Kato T., Watanabe N., Endo Y. Research on Chemical Intermediates, 2006,

32, 217-233.

[19] Saarikoski E., Rautkoski H., Rissanen M., Hartman J., Seppala J. Journal of Applied Polymer

Science, 2014, 131,

[20] Zhang M.Y., Cheng Z.Q., Liu M.Z., Zhang Y.Q., Hu M.J., Li J.F. Journal of Applied Polymer

Science, 2014, 131,

[21] Ben Fradj A., Lafi R., Ben Hamouda S., Gzara L., Hamzaoui A.H., Hafiane A. Journal of

Photochemistry and Photobiology a-Chemistry, 2014, 284, 49-54.

86
Figures and tables

Table 3-1 Chemical compositions, external surface areas and lattice parameters of Mo-V-W-O

a
Composition (at.-% ) External surface Lattice parameter (nm)
Sample b 2 -1
Mo V W area (m g ) a b c

Amor-MoVO-W0 76.1 23.9 0.0 4.7 - - 0.3998

Amor-MoVO-W2.5 74.0 23.3 2.7 4.7 - - 0.3993

Amor-MoVO-W5.0 70.5 23.6 5.9 4.7 - - 0.3991

Amor-MoVO-W7.5 70.1 21.9 8.0 3.9 - - 0.3990

Amor-MoVO-W10 67.9 21.8 10.3 4.0 - - 0.3989

Orth-MoVO-W0 71.5 28.5 0.0 5.0 2.1279 2.6634 0.4002

Orth-MoVO-W2.5 71.8 25.6 2.6 15.6 2.1205 2.6631 0.3999

Orth-MoVO-W5.0 70.0 25.2 4.8 23.3 2.1203 2.6629 0.3995

Orth-MoVO-W7.5 69.7 23.1 7.2 31.4 2.1183 2.6612 0.3994

Orth-MoVO-W10 68.2 21.6 10.2 35.5 2.1053 2.6592 0.3991

Tri-MoVO-W0 74.8 25.2 0.0 17.4 2.1382 - 0.4030

Tri-MoVO-W2.5 74.9 22.1 3.0 17.5 2.1382 - 0.4022

Tri-MoVO-W5.0 73.2 21.4 5.4 14.4 2.1381 - 0.4018

Tri-MoVO-W7.5 71.6 20.7 7.7 18.3 2.1381 - 0.4014

Tri-MoVO-W10 69.8 20.1 10.1 18.9 2.1380 - 0.4011


a
Determined by ICP-AES
b
Measured by N2 adsorption and determined by t-plot method

87
Table 3-2 Average crystal sizes of Mo-V-W-O particles

Catalyst Average diameter (m) Average length (m)

Amor-MoVO-W0 0.31 1.48


Amor-MoVO-W2.5 0.29 1.30
Amor-MoVO-W5.0 0.22 1.20
Amor-MoVO-W7.5 0.23 1.12
Amor-MoVO-W10 0.20 1.11
Orth-MoVO-W0 0.35 1.97
Orth-MoVO-W2.5 0.18 0.74
Orth-MoVO-W5.0 0.10 0.33
Orth-MoVO-W7.5 0.08 0.30
Orth-MoVO-W10 0.08 0.37
Tri-MoVO-W0 0.14 0.82
Tri-MoVO-W2.5 0.14 0.72
Tri-MoVO-W5.0 0.18 0.99
Tri-MoVO-W7.5 0.16 1.06
Tri-MoVO-W10 0.17 1.14

88
Figure 3-1 STEM images (a) and element mapping of Mo (b), V (c) and W (d) of

Amor-MoVO-W2.5 (I), Orth-MoVO-W2.5 (II) and Tri-MoVO-W2.5 (III).

89
Figure 3-2 Raman spectra of Amor-MoVWO (I), Orth-MoVWO (II) and Tri-MoVWO (III) with

different W contents: (a) W0, (b) W2.5, (c) W5.0, (d) W7.5, and (e) W10.

90
Figure 3-3 XRD patterns of Amor-MoVWO (I), Orth-MoVWO (II) and Tri-MoVWO (III) with

different W contents: W0 (a), W2.5 (b), W5.0 (c), W7.5 (d), and W10 (e).

91
Figure 3-4 Rietveld analysis of Orth-MoVWO catalysts: Orth-MoVO-W2.5 (I), Orth-MoVO-W5.0

(II), Orth-MoVO-W7.5 (III), and Orth-MoVO-W10 (IV).

92
Figure 3-5 Temperatures of 50% acrolein conversion over Amor-MoVWO (I), Orth-MoVWO (II),

and Tri-MoVWO (III) catalysts with different tungsten contents and water partial pressures.

93
Figure 3-6 TEM images of Orth-MoVWO with different W contents: W0 (a), W2.5 (b), W5.0 (c),

W7.5 (d), and W10 (e).

94
95
96
Figure 3-7. Distribution of diameters and lengths of Amor-MoVWO(I), Orth-MoVWO (II) and

Tri-MoVWO (III) particles.

97
Figure 3-8. Acrolein conversion of Amor-MoVWO (I), Orth-MoVWO (II) and Tri-MoVWO (III)

with different W contents: W0 (), W2.5 (), W5.0 (), W7.5 (), W8.5 (), and W10 ()

under the condition of water-0.

98
Figure 3-9 Selectivity to acrylic acid over Amor-MoVWO (I), Orth-MoVWO (II) and Tri-MoVWO

(III) with different W content: W0 (), W2.5 (), W5.0 (), W7.5 (), W8.5 (), and W10 ()

under the condition of water-0.

99
Figure 3-10 Selectivity to acetic acid over Amor-MoVWO (I), Orth-MoVWO (II) and

Tri-MoVWO (III) with different W content: W0 (), W2.5 (), W5.0 (), W7.5 (), W8.5 (),

and W10 () under the condition of water-0.

100
Figure 3-11. Selectivity to COx over Amor-MoVWO (I), Orth-MoVWO (II) and Tri-MoVWO (III)

with different W content: W0 (), W2.5 (), W5.0 (), W7.5 (), W8.5 (), and W10 ()

under the condition of water-0.

101
102
Chapter 4 Synthesis of orthorhombic Mo-V-Cu-O complex oxide as highly

selective catalyst for selective oxidation of acrolein to acrylic acid

103
104
4.1 Introduction

In chapter 2 and 3, I described the synthesis of crystalline Mo-V-W-O complex oxides, and

revealed the effect of W on the acrolein oxidation. Although orthorhombic and trigonal phases both

show high efficiency for acrolein oxidation, the catalytic activity of orthorhombic Mo-V-W-O is

superior to trigonal Mo-V-W-O. So in this chapter, I only focus on the introduction of Cu into

orthorhombic structure.

Cu as important promoter is widely used in industrial catalysts. Unlike W, Cu is expected to

be incorporated in the channels instead of framework of orthorhombic Mo-V-O structure because

of the completely different chemistry. Much work has been done trying to introduce Cu into the

heptagonal or hexagonal channels of orthorhombic Mo-V-O. The most common approach is

ion-exchange method which has been extensively used in the modification of zeolites. However, it

is confronted with great difficulties in synthesizing uniform Cu-containing crystalline Mo-V-O

catalysts by using ion exchange method. Therefore, I tried hydrothermal method to incorporate Cu

into the pores in the orthorhombic structure.

In the orthorhombic structure, diameter of the heptagonal channel is about 0.4 nm, while the

size of an ammonium cation occupied the channels is only 0.28 nm which is a little small for a

charge balance counterpart. Therefore, organic structure-directing agents which can act as a

stabilizer will contribute more advantages to form orthorhombic crystals.

In this chapter, I introduce a method of using (CH3NH3)6Mo7O24 to synthesize Mo-V-Cu-O

complex oxide with high dimensional crystalline orthorhombic structure. Cu ions were determined

to be in the heptagonal channels. Incorporation of Cu into the orthorhombic structure can enhance

acrylic acid selectivity to more than 99% in the gas phase acrolein selective oxidation to acrylic

acid. Additionally this carefully prepared Mo-V-Cu-O supplied an excellent model to clear and

indentify the effects of Cu on the acrolein oxidation.

105
4.2 Experimental

4.2.1 Synthesis of catalysts.

4.2.1.1 Synthesis of orthorhombic Mo-V-O complex oxides.

Orthorhombic Mo-V-O (denoted as Orth-MoVO-AHM) was synthesized according to a

previous report.[1, 2] Briefly, a solution of VOSO4nH2O (n=5.4, 3.28 g, Mitsuwa Chemicals) in

deionized water (120 ml) was added to a solution of (NH4)6Mo7O244H2O (8.82 g, Mo: 50 mmol,

Wako) in deionized water (120 ml) with stirring. The mixture was stirred for 10 min and then

transferred into an autoclave with a 300 mL-Teflon inner tube and a Teflon sheet enough length for

filling about half of Teflon inner tube space. This sheet is necessary for the formation of

well-crystallized sample, because solids are formed on the sheet. The mixture (pH=3.2) was purged

with N2 for 10 min and then hydrothermally treated at 175 oC for 48 h. As-synthesized sample was

purified by treatment with a solution of oxalic acid (Wako) (0.4 molL-1 25ml/ 1g solids) at 60 oC

for 30 min to remove amorphous impurities. Obtained solid was dried in air at 80 oC overnight.

4.2.1.2 Synthesis of orthorhombic Mo-V-Cu-O complex oxides.

Methylammonium heptamolybdate (MAHM: (CH3NH3)6Mo7O24) was used as the Mo source

for synthesis of orthorhombic Mo-V-Cu-O. The synthesis procedures for MAHM were as follows
[3]
. 22.594 g of MoO3 (Kanto, 0.15 mol) were slowly dissolved in 33.2 ml of 40% methylamine

solution (methylamine: 0.300 mol, Wako) followed by the stirring for 30 min. After being

completely dissolved, the solution was evaporated under vacuumed condition of P/P0= 0.03 at 70
o
C for 30 min and white solid was obtained. Time was measured from the time that white powder

was started to precipitate. Then the solid was dried in air at 80 oC overnight. XRD pattern of

MAHM is showed in Figure 4-0.

Orthorhombic Mo-V-Cu-O was synthesized by using MAHM, VOSO4nH2O and

CuSO45H2O under a hydrothermal condition. Firstly, solution A (colorless) was obtained with 8.9

g of MAHM being dissolved in 120 ml of deionized water, and solution B (blue color) was

obtained with 3.28 g of VOSO4nH2O (n=5.4, Mitsuwa Chemicals) being dissolved in another 120

ml of deionized water. Secondly, solution B was poured into solution A under a stirring condition
106
to obtain a mixed solution, color of which is purple. After 10 min stirring, 0.60 CuSO45H2O (2.4

mmol, Wako) was added into the mixture solution under the stirring condition for another 10 min.

Then, the obtained mixed solution (pH=3.2) was introduced into an autoclave with a 300

mL-Teflon inner tube and Teflon sheet enough length for filling about half of Teflon inner tube

space. The reaction mixture was purged with N2 for 10 min and then hydrothermally treated at 175
o
C for 48 h. As-synthesized orthorhombic Mo-V-Cu-O sample (denoted as

Orth-MoVO-MAHM-Cu5%, Cu content was determined by ICP-AES method) was treated with

oxalic acid solution (0.4 molL1, 25ml/ 1g solids) (Wako) at 60 oC for 30min. Then obtained

samples were dried in air at 80 oC overnight.

For comparison, orthorhombic Mo-V-O sample without Cu using MAHM as precursor was

also synthesized following the same procedure in section 2.1.1. As-synthesized sample was denoted

as Orth-MoVO-MAHM.

All of the samples were calcined at 400 oC under air condition for 2h to remove ammonium

and organic part before characterization and activity testing.

4.2.2 Characterization

X-ray diffraction patterns (XRD) were measured by using an X-ray diffractometer

(RINT-Ultima III, Rigaku) with Cu K to study their crystalline structure. Rietveld analysis

program with Powder Reflex (Material Studio 5.5.3, Accelrys) was utilized to simulate the

structural model of Orth-MoVCuO. Scanning transmission electron microscopy (STEM) images

and metal element mapping of Mo, V, W and Cu were obtained on an HD-2000 (HITACHI). The

chemical composition of the catalysts was determined by the inductively coupled plasma atomic

emission spectroscopy (ICP-AES) method with a VISTA-PRO apparatus (Varian). N2

adsorption-desorption isotherm measurements were carried out on an auto-adsorption system

(BELSORP MAX, Nippon BELL) to obtain the external surface area and micropore volume using

the t-plot method. Prior to adsorption measurements, the samples were treated at 400 oC for 2 h

under air and out-gassed at 300 oC under vacuum for 2 h.

107
4.2.3 Gas-phase catalytic oxidation

Gas phase catalytic oxidation of acrolein to acrylic acid was performed using a fixed bed

stainless tubular reactor at atmospheric pressure. Catalysts (0.25 g) were firstly ground for 5

minutes, then diluted with 2.5 g Carborundum and pretreated at 400 oC under N2 of 50 mlmin-1 for

2 h. Reactant gases were H2O= 27.1 mlmin-1, acrolein= 2.5 mlmin-1, O2= 8 mlmin-1 and N2

balance (total: 107.6 mlmin-1). Quantitative analysis was performed using three on-line gas

chromatographs with columns of Molecular Sieve 13X, Gaskuropack 54, and Porapak Q. Blank

runs showed that no reaction took place without catalysts under the experimental conditions.

Carbon balance was always over 95%, and selectivity was calculated on the basis of products of

sum.

108
4.3 Results and discussion

4.3.1 Characterization of orthorhombic Mo-V-Cu-O catalyst

The crystalline structures of all catalysts were investigated by XRD characterization, and XRD

patterns are showed in Figure 4-1. Diffraction peaks emerged at 6.6o, 7.9o, 9.0o, 22.2o, 27.3o, 45.4o,

etc. are ascribed to the planes of (020), (120), (210), (001), (630), (002), etc. in the orthorhombic

structure, respectively (Figure 4-1-I).[4, 5] An obvious peak shift of Orth-MoVO-MAHM-Cu5% in

lower angle direction was observed after being calibrated with silicon as an internal standard

(Figure 4-1-II), suggesting that Cu was incorporated into the structure. In the XRD pattern of

Orth-MoVO-MAHM, broaden peaks were observed while in the cases of Orth-MoVO-AHM and

Orth-MoVO-MAHM-Cu5% sharp diffraction peaks emerged. This can be ascribed to the decrease

of Orth-MoVO-MAHM crystal size. Accompanying with the decrease of crystal size, external

surface area of Orth-MoVO-MAHM increased obviously to 29.0 m2g-1 which was nearly 5 times

larger than that of Orth-MoVO-AHM and Orth-MoVO-MAHM-Cu5% (Table 4-1). It is reported

that the external surface area of crystalline Mo-V-based catalyst is an important factor for acrolein

conversion, so Orth-MoVO-MAHM with high external surface area was expected to show good

activity for acrolein oxidation. [1, 6]

Table 4-1 shows the chemical compositions of all catalysts determined by the ICP-AES

method. Cu content of Orth-MoVO-MAHM-Cu5% was 5.3 at.-% indicating that Cu was

successfully introduced into the catalyst. Different Mo and V atomic ratios are derived from the

structure requisite. Orth-MoVO-MAHM-Cu5% was rod-shaped (STEM image, Figure 4-2-I) and

the diameters of crystals were hundreds of nanometers. Element mapping of Mo, V and Cu showed

that Cu evenly dispersed in the material (Figure 4-2-II). This can be another evidence of that Cu

was incorporated into the orthorhombic structure. it is worthwhile to notice that, although external

surface area 4.9 m2g-1 of Orth-MoVO-MAHM-Cu5% is the same with that of Orth-MoVO-AHM,

the pore volume of Orth-MoVO-MAHM-Cu5% diminished to only 0.810-3 mlg-1. Orthorhombic

Mo-V-O material shows microporosity because of the existence of heptagonal channels. The

decrease of pore volume of Orth-MoVO-MAHM-Cu5% strongly suggests that Cu ions are located

in the heptagonal channels.


109
To confirm the location of Cu in Orth-MoVO-MAHM-Cu5%, Rietveld refinement was carried

out. Structural model and Rietveld analysis of Orth-MoVO-MAHM-Cu5% are showed in Figure

4-3. Simulated data fit experimental XRD pattern very well and Rwp value is 6.27%, manifesting

that the structure model is reasonable. Cu ions were simulated to be incorporated into the

orthorhombic structure and located in the heptagonal channels neighboring to the

quintuple-octahedron. The short distance between Cu and quintuple-octahedron indicates that a

strong interaction between Cu and neighboring oxygen may exist. When the heptagonal channels

are fully occupied by Cu, atomic ratio of Cu is about 6.3 at.-%. As exhibited in Figure 4-6-I, with

the increase of Cu content from 0 to 10 at.-% in the catalysts, the orthorhombic structure was

maintained. It is worthwhile to notice that, when Cu content was lower than 5 at.-%, obvious peak

shift in lower angle direction was observed. On the contrary, there was almost no peak shift when

Cu content was beyond 5 at.-% (Figure 4-6-II). This indicated that a limited amount of Cu ions can

be incorporated into the orthorhombic structure. Although catalysts with high Cu content could also

be obtained for example Orth-MoVO-Cu10%, in this case, Cu maybe not only located in the

channels but also formed other species.

4.3.2 Effect of Cu on the orthorhombic Mo-V-O structure

Interestingly, when methylammonium cation ((CH3NH4)+) or Cu ions were presented

individually in the precursor solution, well crystallized orthorhombic structure cannot be obtained

through the hydrothermal synthesis. For example, broaden peaks emerged in the XRD pattern of

Orth-MoVO-MAHM in the Figure 4-1-I; and only amorphous phase was formed when Cu and

NH4+ both existed in the precursor solution. However, when Cu ions and (CH3NH4)+ both exist in

the precursor solution, high dimensional orthorhombic Mo-V-Cu-O was synthesized. These results

reveal that the self-assembly process is seriously affected by the ammonium species and Cu ions. A

co-effect of (CH3NH4)+ and Cu ions may exist, and this co-effect will contribute to the formation of

a high dimensional orthorhombic structure. Similar with the functions of ammonium cation in the

synthesis of aluminosilicate zeolites, ammonium cation also played very important roles in the

formation of the orthorhombic structure.[3] In the self-assembly process, {Mo6O21} units and
110
{MO6} (M = Mo or V) octahedra are firstly formed, and then assemble the orthorhombic structure

with ammonium cation occupying the heptagonal channels. Therefore, the ammonium cation acts

not only as a charge balance counterpart but also structure-directing agent and stabilizer. In

addition, pH value affects the formation of Mo species and self-assembly process in the

hydrothermal condition.[3, 7] For example, orthorhombic structure is formed under the condition of

pH=3.2, while for the formation of trigonal phase the pH value must be decreased to 2.2. The pH

value of mixed solution after adding CuSO4 was around 3.2 which is an available for the formation

of the orthorhombic structure avoiding any pH adjustment by using ammonium solution.

4.3.3 Catalytic performance of orthorhombic Mo-V-Cu-O catalyst for the selective oxidation

of acrolein to acrylic acid.

Previous work showed that crystalline orthorhombic structure and external surface area played

extraordinary important roles in the selective oxidation of acrolein to acrylic acid.[1, 6] Acrolein

molecules are not small enough to enter the channels which are only 0.4 nm, so acrolein selective

oxidation is only a surface reaction. Compare with amorphous and mixed phase oxides catalysts,

the a-b plane of the orthorhombic structure was constructed into an ordered arrangement. Hence,

orthorhombic Mo-V-O catalysts can supply more active sites and show much better activity. In this

work, orthorhombic Mo-V-Cu-O catalyst was successfully synthesized. Different from the

orthorhombic Mo-V-W-O catalysts in which W ions were incorporated into the framework,[6] Cu

ions are in the heptagonal channels of orthorhombic Mo-V-O. When this orthorhombic

Mo-V-Cu-O was applied as a catalyst to the acrolein oxidation, the catalytic performance may be

affected.

Gas-phase selective oxidation of acrolein to acrylic acid was conducted over crystalline

Orth-MoVO-AHM, Orth-MoVO-MAHM and Orth-MoVO-MAHM-Cu5% catalysts. The acrolein

conversions as the function of reaction temperature are illustrated in Figure 4-4.

Orth-MoVO-MAHM was the most active catalyst achieving 93.9% acrolein conversion at the

temperature of 220 oC. Orth-MoVO-AHM and Orth-MoVO-MAHM-Cu5% catalysts at this

temperature, on the other hand, were less active and gave only 57.5% and 29.5%, respectively. As
111
noticed in the structure analysis, all those catalysts showed the same orthorhombic structure, but

with different external surface areas. Orth-MoVO-MAHM catalyst had the highest external surface

area of 29.0 m2g-1 (Table 4-1). Therefore, it showed the best activity. For Orth-MoVO-AHM and

Orth-MoVO-MAHM-Cu5% catalysts, they had the same orthorhombic structure and external

surface area, but Cu containing catalyst showed less catalytic activity indicating that Cu had a

negative effect on the acrolein conversion. Although Cu containing catalyst showed less activity

than that of Mo-V-O catalyst, it showed extremely high selectivity to acrylic acid. As been showed

in Figure 4-5, selectivity to acrylic acid over Orth-MoVO-MAHM-Cu5% is enhanced to more than

99% which is surprisingly high and totally not comparable to other reported data.[1, 8, 9] It was

acknowledged that 100% selectivity to targeted product is a challenge but highly desired for

heterogeneous catalytic reactions. High selectivity will simplify the after-treatment and refinement

in the industrial production. For acrolein oxidation, almost 100% conversion can be obtained by

adjusting the reaction temperature, but it is a challenge to achieve 100% selectivity to acrylic acid

while keeping high acrolein conversion. So the promotion effect of Cu on the selectivity to acrylic

acid is very helpful for the industrial production of acrolein oxidation to acrylic acid.

4.3.4 Effect of Cu content on the catalytic selectivity to acrylic acid

Acrolein conversion and selectivity to acrylic acid as function of temperature are illustrated in

Figure 4-7. It is noticed that with the increase of Cu content in the orthorhombic structure, activity

decreased gradually while selectivity to acrylic acid increased. The selectivity to acrylic acid over

Orth-MoVO-MAHM catalyst was 97.5%. With the increase of Cu content from 0 to 6 at.-%, the

selectivity was gradually enhanced to 99.3%. However, when Cu content was increased to more

than 6 at.-%, selectivity to acrylic acid showed a plateau (Figure 4-8). This percentage (6 at.-%)

was coincident with the maximum capacity of Cu (6.3 at.-%) in the orthorhombic structure

calculated by the structural analysis. This strongly indicates that only the Cu that is incorporated

into the structure may enhance the catalytic selectivity.

For selective oxidation reaction, according to the Mars-van-krevelen mechanism, lattice

oxygen is involved in the redox cycle and plays an important role in the selective oxidation process.
112
In the Orth-MoVO-MAHM-Cu5% catalyst, Cu ions occupy the heptagonal channels and strong

interaction between Cu and neighboring oxygens exist. So the lattice oxygens nearby Cu are, to

some extent, deactivated by the strong interaction with Cu. They are not easy to be translated to

substrate, and the molecular oxygen in the reactant gas can more easily transform to lattice oxygen

due to the requirement of completing a lattice defect. That is to say, the strong interaction slows

down the migration of the lattice oxygen, and results in the decrease of acrolein conversion rate

(Figure 4-7) on one hand. On the other hand, the strong interaction between the incorporated Cu

and heptagonal 7-memberd ring is beneficial to the integrality of the orthorhombic structure which

is highly selective for the partial oxidation of acrolein to acrylic acid. Therefore, the addition of Cu

into the orthorhombic structure can enhance the selectivity but decrease acrolein conversion.

4.3.5 Investigation of structure-activity relationships

The defined orthorhombic Mo-V-Cu-O structure offers a suitable model for revealing the

structure-activity relationships. To investigate what happened with the structure during the acrolein

oxidation process, an acrolein-treatment over the Orth-MoVO-MAHM-Cu5% catalyst was carried

out. Processes of the acrolein-treatment are as follows. Orth-MoVO-MAHM-Cu5% was firstly

treated with acrolein (2.5 mlmin-1) and N2 (20mlmin-1) mixed gas at 350 oC for 30min. Then XRD

and Rietveld analysis method were carried out to define the structure of the treated sample. After

that, acrolein-treated Orth-MoVO-MAHM-Cu5% was calcined again under air condition at 400 oC

for 2h. XRD patterns of Orth-MoVO-MAHM-Cu5% before and after acrolein treatment were

showed in Figure 4-9. After acrolein-treatment, the orthorhombic structure was maintained, but

relative intensities of diffraction peaks changed. For example, intensity ratio of (210) to (120)

diffraction peak was decreased obviously; and peak shift of (001) diffraction peak in high angle

direction was observed. However, XRD pattern of Orth-MoVO-MAHM-Cu5% can recover after

being calcined again in the air condition. These results suggest that the orthorhombic structure was

partially changed by the acrolein-treatment, and acrolein can take reactions with the catalyst

without the presence of molecular air. To determine the structure changes during the

acrolein-treatment, Rietveld refinement method was used for the construction of the reduced
113
orthorhombic structure.

Rietveld refinement results of Orth-MoVO-MAHM-Cu5% treated by acrolein were showed in

Figure 4-10. From Figure 4-10, it is clearly noticed that Cu ions migrated to the center of

heptagonal channels after acrolein-treatment. This indicated that acrolein reacted with lattice

oxygen, and strong interaction between Cu and lattice oxygen exists. The bond energy of oxygen in

the catalyst lattice is related to the catalytic activity and selectivity. Weakly bonded oxygen will

contribute to form products with higher oxygen content, and then lead to a nonselective oxidation.

Strong bonded oxygen will decelerate the oxygenation process, and thus give advantages to the

selective oxidation.[8, 10, 11] Therefore, incorporation of Cu strengthens the metal-oxygen bond and

results in a higher selectivity to acrylic acid; on the other hand, strong bonded oxygen will decrease

the acrolein conversion. Comparing the original position of Cu and the position after the

acrolein-treatment, it can be proposed that the oxygen atom belonging to the linker {VO6}

octahedra (OI) in the quintuple-octahedron plays a very important role during the reaction process.

It was found that the real active phase was on the a-b plane of the orthorhombic structure.[6]

Here, the results suggest that, for orthorhombic Mo-V-Cu-O catalyst, the 7-membered ring is the

active site for acrolein oxidation to acrylic acid.

4.4 Conclusions

Orthorhombic Mo-V-Cu-O catalysts with different Cu content were successfully synthesize by

a hydrothermal method using (CH3NH3)6Mo7O24 as Mo precursor. The addition of Cu into the

heptagonal channels of orthorhombic structure significantly enhanced the selectivity in the

selective oxidation of acrolein to acrylic acid. Succeeding in the synthesis of crystalline

orthorhombic Mo-V-Cu-O complex oxide is important example for incorporation of metals into the

channels. Using the similar synthetic method, it is available to prepare crystalline Mo-V-O with

other transition metals such as Mn, Fe, Co, Ni and Zn. These metals can optimize the crystalline

structure as well as the surface properties, and are expected to expand the application domain of

Mo-V-based molecular sieve.

114
References

[1] Chen C., Kosuke N., Murayama T., Ueda W. Chemcatchem, 2013, 5, 2869-2873.

[2] Konya T., Katou T., Murayama T., Ishikawa S., Sadakane M., Buttrey D., Ueda W. Catalysis

Science & Technology, 2013, 3, 380-387.

[3] Ishikawa S., Murayama T., Ohmura S., Sadakane M., Ueda W. Chemistry of Materials, 2013, 25,

2211-2219.

[4] Sadakane M., Kodato K., Kuranishi T., Nodasaka Y., Sugawara K., Sakaguchi N., Nagai T.,

Matsui Y., Ueda W. Angewandte Chemie-International Edition, 2008, 47, 2493-2496.

[5] Ueda W., Vitry D., Kato T., Watanabe N., Endo Y. Research on Chemical Intermediates, 2006,

32, 217-233.

[6] Qiu C.T., Chen C., Ishikawa S., Murayama T., Ueda W. Topics in Catalysis, 2014, 57,

1163-1170.

[7] Sadakane M., Endo K., Kodato K., Ishikawa S., Murayama T., Ueda W. European Journal of

Inorganic Chemistry, 2013, 1731-1736.

[8] Andrushkevich T.V. Catal.Rev.-Sci.Eng., 1993, 35, 213-259.

[9] Kunert J., Drochner A., Ott J., Vogel H., Fuess H. Applied Catalysis a-General, 2004, 269,

53-61.

[10] Jekewitz T., Blickhan N., Endres S., Drochner A., Vogel H. Catalysis Communications, 2012,

20, 25-28.

[11] Tichy J. Applied Catalysis a-General, 1997, 157, 363-385.

115
Figures and tables

Table 4-1 Chemical composition, external surface area and pore volume of Orth-MoVO-AHM,

Orth-MoVO-MAHM and Orth-MoVO-MAHM-Cu5%.

Compositiona (at.-% ) External surface Pore volumeb


Catalyst
Mo V Cu areab (m2g-1) (10-3mlg-1)

Orth-MoVO-AHM 71.5 28.5 - 5.0 13.7

Orth-MoVO-MAHM 76.6 23.4 - 29.0 6.3

Orth-MoVO-MAHM-Cu5% 69.0 25.8 5.3 4.9 0.8

a
Determined by ICP-AES
b
Measured by N2 adsorption and determined by t-plot method

Figure 4-0 XRD pattern of Methylammonium heptamolybdate (MAHM: (CH3NH3)6Mo7O24)

116
Figure 4-1 (I) XRD patterns of Orth-MoVO-AHM, Orth-MoVO-MAHM and

Orth-MoVO-MAHM-Cu5%; (II) after calibrated with silicon as an internal standard.

117
Figure 4-2 (I) STEM images and (II) element mapping of Mo, V and Cu of

Orth-MoVO-MAHM-Cu5%.

118
Figure 4-3. Structural model and Rietveld analysis of Orth-MoVO-MAHM-Cu5%.

119
Figure 4-4 Acrolein (ACR) conversion over Orth-MoVO-AHM (), Orth-MoVO-MAHM () and

Orth-MoVO-MAHM-Cu5% () catalysts.

120
Figure 4-5 Selectivity to acrylic acid (AA) (a), acetic acid (HAC) (b) and COx (c) over

Orth-MoVO-AHM (), Orth-MoVO-MAHM() and Orth-MoVO-MAHM-Cu5% () catalysts.

121
Figure 4-6 (I) XRD patterns of Orth-MoVO-MAHM-Cu with different Cu content:

Orth-MoVO-MAHM (a), Orth-MoVO-MAHM-Cu2.5% (b), Orth-MoVO-MAHM-Cu5% (c),

Orth-MoVO-MAHM-Cu7.5% (d) and Orth-MoVO-MAHM-Cu10% (e); (II) after calibrated with

silicon as an internal standard.

122
Figure 4-7 Acrolein (ACR) conversion and selectivity to acrylic acid (AA) over

Orth-MoVO-MAHM-Cu with different Cu content: Orth-MoVO-MAHM (),

Orth-MoVO-MAHM-Cu2.5% (), Orth-MoVO-MAHM-Cu5% (),

Orth-MoVO-MAHM-Cu7.5% () (d) and Orth-MoVO-MAHM-Cu10% () .

123
Figure 4-8 Selectivity to acrylic acid over Orth-MoVO-MAHM-Cu catalysts with different Cu

contents at 90% acrolein conversion.

124
Figure 4-9 XRD patterns of Orth-MoVO-MAHM-Cu5% before and after acrolein-treatment.

125
Figure 4-10 Structural model and Rietveld analysis of acrolein-treated

Orth-MoVO-MAHM-Cu5%.

126
Chapter 5 Synthesis of highly efficient orthorhombic Mo-V-W-Cu-O

catalysts for the selective oxidation of acrolein to acrylic acid

127
128
5.1 Introduction

Owing to the development of glycerol dehydration technology, abundant of acrolein can be

produced by new approach glycerol dehydration to acrolein.[1-5] So the catalysts for acrolein

oxidation to acrylic acid achieve much more attentions. Compared to industrial Mo-V-based

catalysts which are mixed phases with numerous promoters, so far the crystalline Mo-V-W-O and

Mo-V-Cu-O catalysts were only conducted in laboratory scale. However, these crystalline

Mo-V-based catalysts show excellent application prospect in production of acrylic acid from

acrolein.

In the chapter 2 and 3, orthorhombic Mo-V-W-O catalysts with various amount of W were

synthesized. The introduction of W leads to a rod segregation effect by which the external surface

area greatly increased, and thus catalytic activity of acrolein oxidation to acrylic acid increased. In

the chapter 4, a series of orthorhombic Mo-V-Cu-O catalysts were synthesized and the addition of

Cu largely enhanced selectivity to acrylic acid. Based on these work, it is expected to combine both

positive effects of W and Cu to obtain a highly effective orthorhombic Mo-V-W-Cu-O catalyst for

the selective oxidation of acrolein to acrylic acid. So in this chapter, I synthesized a series of

orthorhombic Mo-V-W-Cu-O catalysts with different W/Cu ratio by a hydrothermal method.

Contents of W and Cu in the catalysts were adjusted from 0 to 10mol %, respectively. These

catalysts were performed in gas phase acrolein oxidation to acrylic acid. The results showed that

both catalytic activity and selectivity to acrylic acid over orthorhombic Mo-V-W-Cu-O catalysts

were enhanced compared to orthorhombic Mo-V-O and those industrial mixed phases Mo-V-O

catalysts. The best catalyst for acrolein oxidation is assigned to orthorhombic

Mo-V-W7.5%-Cu7.5%-O catalyst which exhibit more that 90% conversion and 99% selectivity to

acrylic acid at 230oC.

129
5.2 Experimental

5.2.1 Synthesis of orthorhombic Mo-V-W-Cu-O complex oxides.

A series of orthorhombic Mo-V-W-Cu-O (denoted as Orth-MoVO-WxCuy; x and y are atomic

ratios of W and Cu, respectively) catalysts with different W/Cu ratios were obtained according to a

previous report.[6, 7] Briefly, a solution of VOSO4nH2O (n=5.4, 3.28 g, Mitsuwa Chemicals) in

deionized water (120 ml) was added to a solution of Methylammonium heptamolybdate (MAHM:

(CH3NH3)6Mo7O24, 8.9g) in deionized water (120 ml) under stirring condition for 10 min. Then

certain amount of (NH4)6[H2W12O40]6H2O (Nippon Inorganic Colour & Chemical) and

CuSO45H2O (Wako) were added into the mixed solution (W and Cu content were adjusted from 0 to

10 at.-% respectively, determined by ICP-AES method, Table 5-2). The mixture was stirred for

another 10 min and then transferred into an autoclave with a 300 mL-Teflon inner tube and a

Teflon sheet enough length for filling about half of Teflon inner tube space. This sheet is necessary

for the formation of well-crystallized sample, because solids are formed on the sheet. The mixture

(pH=3.2) was purged with N2 for 10 min and then hydrothermally treated at 175 oC for 48 h.

As-synthesized sample was purified by treatment with a solution of oxalic acid (Wako) (0.4 mol L1,

25ml/ 1g solids) at 60 oC for 30 min to remove amorphous impurities. Obtained solid was dried in

air at 80 oC overnight.

The synthesis procedures for MAHM were as follows.[8] 22.594 g of MoO3 (Kanto; 0.15 mol)

were slowly dissolved in 33.2 ml of 40% methylamine solution (methylamine: 0.300 mol, Wako)

followed by the stirring for 30 min. After being completely dissolved, the solution was evaporated

under vacuumed condition of P/P0= 0.03 at 70 oC for 30 min and white solid was obtained. Time

was measured from the time that white powder was started to precipitate. Then the solid was dried

in air at 80 oC overnight.

All of the samples were calcined at 400 oC under air condition for 2h to remove ammonium

and organic part before characterization and activity testing.

5.2.2 Characterization.

X-ray diffraction patterns (XRD) were measured by using an X-ray diffractometer


130
(RINT-Ultima III, Rigaku) with Cu K to study their crystalline structure. Scanning transmission

electron microscopy (STEM) images and metal element mapping of Mo, V, W and Cu were

obtained on an HD-2000 (HITACHI). The chemical composition of the catalysts was determined

by the inductively coupled plasma atomic emission spectroscopy (ICP-AES) method with a

VISTA-PRO apparatus (Varian). N2 adsorption-desorption isotherm measurements were carried out

on an auto-adsorption system (BELSORP MAX, Nippon BELL) to obtain the external surface area

and micropore volume using the t-plot method. Prior to adsorption measurements, the samples were

treated at 400 oC for 2 h under air and out-gassed at 300 oC under vacuum for 2 h.

5.2.3 Gas-phase catalytic oxidation.

Gas phase catalytic oxidation of acrolein to acrylic acid was performed using a fixed bed

stainless tubular reactor at atmospheric pressure. Catalysts (0.25 g) were firstly ground for 5

minutes, then diluted with 2.5 g Carborundum and pretreated at 400 oC under N2 of 50 mlmin-1 for

2 h. Reactant gases were H2O= 27.1 mlmin-1, acrolein= 2.5 mlmin-1, O2= 8 mlmin-1 and N2

balance (total: 107.6 mlmin-1). Quantitative analysis was performed using three on-line gas

chromatographs with columns of Molecular Sieve 13X, Gaskuropack 54, and Porapak Q. Blank

runs showed that no reaction took place without catalysts under the experimental conditions.

Carbon balance was always over 95%, and selectivity was calculated on the basis of products of

sum.

131
5.3 Results and discussion

5.3.1 Characterization of orthorhombic Mo-V-W-Cu-O catalysts.

A hydrothermal method using MAHM as Mo precursor was used for the synthesis of

orthorhombic Mo-V-W-Cu-O catalysts. A group of typical Mo-V-W-Cu-O catalysts which contain

a constant Cu content (5 at.%) and different W content were characterized. XRD patterns of these

Orth-MoVO-WCu (Cu5%) catalysts are showed in Figure 5-1. In the patterns, diffraction peaks

emerged at 6.6o, 7.9o, 9.0o and 27.3o, etc. corresponding to the orthorhombic structure.[9] This

indicated that Mo-V-W-Cu-O catalysts with an orthorhombic structure were successfully

synthesized. With the incorporation of W and Cu, a lattice contraction happened and a peak shifted

of (001) in high angle direction can be observed in Figure 5-1-II. Apparently, the lattice contraction

affected the long-range order of building blocks, and it made against the crystal growth. Herein,

smaller crystals were easily formed and broaden diffraction peaks emerged with the increase of W

content (Figure 5-1-I). In the Figure 5-2, STEM image clearly shows that crystal size of the

Orth-MoVO-W5%Cu5% decreased to only tens of nanometers. In chapter 3, it was revealed that

the broad peaks were ascribed to the rod-segregation effect which was caused by the addition of

W.[10] Here, the introduction of W showed the same effect in the Mo-V-W-Cu-O catalysts. The

external surface area of Orth-MoVO-W10%Cu5% largely increased to 34.1 m2g-1 (Table5-1).

However, pore volumes of these catalysts are much smaller than that of Orth-MoVO inferring that

Cu ions are located in the heptagonal channels.

Chemical compositions of catalysts were determined by ICP-AES method, and results are

listed in Table 5-1. The results showed that W and Cu were introduced into the Mo-V-W-Cu-O

catalysts. In company with the increase of W content, both Mo and V content gradually decreased.

That is because W is incorporated into the framework, not only Mo but also V was replaced by W.

Additionally, elemental mapping of Mo, V, W, and Cu in Figure 5-2 showed that all elements were

evenly dispersed in the rod-shaped crystals. All these results indicate that W and Cu were

successfully incorporated into the structure of orthorhombic Mo-V-W-Cu-O catalysts.

Normally, when more than 5 at.-% W was introduced into the orthorhombic structure, broaden

peaks were observed in the XRD pattern (Figure 3-3-II). But in the case of
132
Orth-MoVO-W5%Cu5%, clear diffraction peaks emerged in the XRD pattern while, in this case

the content of hetero atoms (W and Cu) in the Orth-MoVO-W5%Cu5% catalyst is more than 10

at.-%. This indicates that although W goes against to the crystal growth, the co-effect of Cu and

methylammonium cation tends to stabilize the orthorhombic structure. The

Orth-MoVO-W10%Cu5% catalyst with large external surface area and crystalline orthorhombic

structure were expected to show good activity and high selectivity for the selective oxidation of

acrolein to acrylic acid.

5.3.2 Catalytic performance of orthorhombic Mo-V-W-Cu-O catalysts for the selective

oxidation of acrolein to acrylic acid.

Synthesized orthorhombic Mo-V-W-Cu-O (Cu5%) catalysts were conducted in gas phase

acrolein oxidation to acrylic acid. Acrolein conversion and selectivity to acrylic acid, acetic acid,

and COx were illustrated in Figure 5-3 and Figure 5-4, respectively. Compared to

Orth-MoVO-MAHM-Cu5% catalyst, the introduction of W into the catalysts enhances acrolein

conversion. For example, acrolein conversion of Orth-MoVO-MAHM-Cu5% catalyst is only

29.5% at the temperature of 220 oC, while conversion of Orth-MoVO-W10%Cu5% rises up to

63.1% which is even higher than that of Orth-MoVO-AHM catalyst. With the increase of W

content, activities are gradually enhanced with keeping high selectivity to acrylic acid. At the

conversion of 90%, more than 99% of selectivity to acrylic acid is achieved. All these results show

that the addition of W and Cu enhances both catalytic activity and selectivity for the selective

oxidation of acrolein to acrylic acid.

It is worthwhile to notice that, with the increase of W content, pore volumes of Cu-containing

samples slightly increase from 1.2 to 3.110-3 mlg-1 (Table 2). It seems that the capacity of Cu ions

in the structure was expanded by the introduction of W. If so, more Cu can be incorporated into the

structure.

133
5.3.3 Highly effective orthorhombic Mo-V-W-Cu-O catalysts for the selective oxidation of

acrolein to acrylic acid.

Trying to obtain the best formulation, a series of orthorhombic Mo-V-W-Cu-O catalysts with

different amount of W and Cu were synthesized. The W and Cu contents were adjusted from 0 to

10 at.-%, respectively. All these catalysts were conducted in the gas phase selective oxidation of

acrolein to acrylic acid. Details of chemical composition, external surface area and pore volume are

summarized in Table 5-2. Temperature of 90% acrolein conversion and selectivity to acrylic acid at

90% acrolein conversion are illustrated in Figure 5-5and Figure 5-6, respectively.

In the Figure 5-5, catalytic activity decrease with the increase of Cu content from 0 to 10 at-%

on the one hand. On the other hand, with the increase of Cu content, selectivity to acrylic acid

gradually increases to more than 99%. In the case of W, less than 7.5 at.-% of W in the catalyst

enhanced the acrolein conversion, but more than that percentage, the activity gradually decreased.

Because too much W will affect the formation of crystalline orthorhombic crystals, and amorphous

phase easily forms. This amorphous phase is much less active for acrolein oxidation.[6, 10] Catalyst

with the highest activity is assigned to Orth-MoVO-W7.5%Cu0, which possesses orthorhombic

structure and high external surface area. Although the activity of Orth-MoVO-W7.5%Cu0 catalyst

is high, selectivity is not good enough from the view point of industrial production. Interestingly,

there is a bottom in the Figure 5-5 when Cu content increased to a higher level. That is

Orth-MoVO-W7.5%Cu7.5%. This catalyst achieved 90% acrolein conversion at about 230 oC, and

more importantly the selectivity to acrylic acid is more than 99% (Figure 5-6). Therefore, the best

catalyst for selective oxidation of acrolein to acrylic acid is assigned to

Orth-MoVO-W7.5%Cu7.5% which posses both high catalytic activity and acrylic acid selectivity

at a low reaction temperature (230 oC).

5.4. Conclusions

In this chapter, a series of crystalline Mo-V-W-Cu-O complex oxides with an orthorhombic

structure were successfully synthesized. The addition of W and Cu enhanced both catalytic activity

and selectivity. The best catalyst is assigned to Orth-MoVO-W7.5%Cu7.5% which achieves more
134
than 90% acrolein conversion and 99% selectivity to acrylic acid at 230 oC.

135
References

[1] Haider M.H., D'Agostino C., Dummer N.F., Mantle M.D., Gladden L.F., Knight D.W., Willock

D.J., Morgan D.J., Taylor S.H., Hutchings G.J. Chemistry-a European Journal, 2014, 20,

1743-1752.

[2] Liu L.C., Wang B., Du Y.H., Borgna A. Applied Catalysis a-General, 2015, 489, 32-41.

[3] Decolatti H.P., Dalla Costa B.O., Querini C.A. Microporous and Mesoporous Materials, 2015,

204, 180-189.

[4] Omata K., Izumi S., Murayama T., Ueda W. Catalysis Today, 2013, 201, 7-11.

[5] Feng X.Z., Yao Y., Su Q., Zhao L., Jiang W., Ji W.J., Au C.T. Applied Catalysis

B-Environmental, 2015, 164, 31-39.

[6] Chen C., Kosuke N., Murayama T., Ueda W. Chemcatchem, 2013, 5, 2869-2873.

[7] Konya T., Katou T., Murayama T., Ishikawa S., Sadakane M., Buttrey D., Ueda W. Catalysis

Science & Technology, 2013, 3, 380-387.

[8] Ishikawa S., Murayama T., Ohmura S., Sadakane M., Ueda W. Chemistry of Materials, 2013, 25,

2211-2219.

[9] Ueda W., Vitry D., Kato T., Watanabe N., Endo Y. Research on Chemical Intermediates, 2006,

32, 217-233.

[10] Qiu C.T., Chen C., Ishikawa S., Murayama T., Ueda W. Topics in Catalysis, 2014, 57,

1163-1170.

136
137
Figures and tables

Table 5-1 Chemical composition and external surface area of Orth-MoVO-WCu (Cu 5%) with

different W/Cu ratios.

Compositiona (at.-% ) External surface Pore volumeb


Catalyst
Mo V W Cu areab (m2g-1) (10-3mlg-1)

Orth-MoVO-AHM 71.5 28.5 - - 5.0 13.7

Orth-MoVO-MAHM-Cu5% 69.0 25.8 - 5.3 4.9 1.2

Orth-MoVO-W2.5%Cu5% 67.9 24.5 2.0 5.7 12.7 1.4

Orth-MoVO-W5.0%Cu5% 66.6 21.5 5.4 6.5 21.2 2.8

Orth-MoVO-W7.5%Cu5% 65.4 19.9 8.7 6.0 29.8 2.4

Orth-MoVO-W10%Cu5% 63.5 18.6 11.2 6.7 34.1 3.1

a
Determined by ICP-AES
b
Measured by N2 adsorption and determined by t-plot method

138
Table 5-2 Chemical compositions and external surface areas of Orth-MoVO-WCu catalysts with

different W/Cu ratios.

Compositiona (at.-% )
External surface areab Pore volumeb
Sample
Mo V W Cu (m2g-1) (10-3mlg-1)

Orth-MoVO-AHM 71.5 28.5 - - 5.0 13.7

Orth-MoVO-MAHM 76.6 23.4 - - 29.0 6.3

Orth-MoVO-W2.5%Cu0 71.8 25.6 2.6 - 15.6 12.0

Orth-MoVO-W5.0%Cu0 70.0 25.2 4.8 - 23.3 8.1

Orth-MoVO-W7.5%Cu0 69.7 23.1 7.2 - 31.4 8.0

Orth-MoVO-W10%Cu0 68.2 21.6 10.2 - 35.5 4.7

Orth-MoVO- W0Cu2.5% 71.6 26.9 - 1.5 10.1 2.0

Orth-MoVO- W0Cu5% 69.3 25.4 - 5.3 4.7 0.8

Orth-MoVO- W0Cu7.5% 70.4 23.5 - 6.1 4.3 0.9

Orth-MoVO- W0Cu10% 66.4 22.5 - 11.1 5.4 1.2

Orth-MoVO-W2.5%Cu2.5% 70.5 24.0 2.5 3.0 11.8 1.8

Orth-MoVO-W5.0%Cu2.5% 70.9 21.7 5.2 2.2 13.9 2.5

Orth-MoVO-W7.5%Cu2.5% 68.0 21.6 8.0 2.4 17.6 1.6

Orth-MoVO-W10%Cu2.5% 66.0 20.0 10.9 3.1 23.3 1.9

Orth-MoVO-W2.5%Cu5% 67.9 24.5 2.0 5.7 12.7 1.4

Orth-MoVO-W5.0%Cu5% 66.6 21.5 5.4 6.5 21.2 2.8

Orth-MoVO-W7.5%Cu5% 65.4 19.9 8.7 6.0 29.8 2.4

Orth-MoVO-W10%Cu5% 63.5 18.6 11.2 6.7 34.1 3.1

Orth-MoVO-W2.5%Cu7.5% 65.6 25.2 2.0 7.2 7.9 1.9

Orth-MoVO-W5.0%Cu7.5% 64.5 21.1 5.5 8.9 20.6 2.5

Orth-MoVO-W7.5%Cu7.5% 64.4 20.1 8.7 6.8 36.9 2.8

Orth-MoVO-W10%Cu7.5% 62.2 19.2 11.1 7.6 32.1 2.7

Orth-MoVO-W2.5%Cu10% 63.9 22.1 2.4 11.6 7.0 1.4


139
Orth-MoVO-W5.0%Cu10% 62.2 20.6 5.1 12.0 12.7 1.8

Orth-MoVO-W7.5%Cu10% 60.5 18.7 7.9 13.0 23.3 1.9

Orth-MoVO-W10%Cu10% 60.0 18.3 10.1 11.6 24.2 1.9

a
Determined by ICP-AES
b
Measured by N2 adsorption and determined by t-plot method

140
Figure 5-1 (I) XRD patterns of Orth-MoVO-WCu (Cu5%) with different W/Cu ratios:

Orth-MoVO-Cu5% (a), Orth-MoVO-W2.5%Cu5% (b), Orth-MoVO-W5%Cu5% (c),

Orth-MoVO-W7.5%Cu5% (d), Orth-MoVO-W10%Cu5% (e); (II) after calibrated with silicon as

an internal standard.

141
Figure 5-2 STEM image and element mapping of Mo, V, W and Cu in Orth-MoVO-W5%Cu5%.

142
Figure 5-3 Acrolein (ACR) conversion over Orth-MoVO-WCu (Cu5%) catalysts with different

W/Cu ratios: Orth-MoVO-AHM (), Orth-MoVO-MAHM-Cu5% (), Orth-MoVO-W2.5%Cu5%

(), Orth-MoVO-W5%Cu5% ( ), Orth-MoVO-W7.5%Cu5% ( ) and

Orth-MoVO-W10%Cu5% ().

143
Figure 5-4 Selectivity to acrylic acid (AA) (a), acetic acid (HAC) (b) and COx (c) over

Orth-MoVO-WCu (5%) catalysts with different W/Cu ratios: Orth-MoVO-AHM (),

Orth-MoVO-MAHM-Cu5% (), Orth-MoVO-W2.5%Cu5% (), Orth-MoVO-W5%Cu5% (),

Orth-MoVO-W7.5%Cu5% (), and Orth-MoVO-W10%Cu5% ().

144
Figure 5-5 Acrolein (ACR) conversion over Orth-MoVO-WCu catalysts with different W/Cu

ratios

145
Figure 5-6 Selectivity to acrylic acid (AA) over Orth-MoVO-WCu catalysts with different W/Cu

ratios.

146
Chapter 6
Synthesis of orthorhombic Mo-V-M-O (M=Mn, Co, Ni, Zn, Fe) catalysts
for acrolein oxidation to acrylic acid

147
148
6.1 Introduction
Besides W and Cu, many other transition metals such as Fe, Mn, Ni and Co are also important
additives for the catalysts of acrolein oxidation to acrylic acid. However so far, few reports have
indentified the effects of these metals on acrolein oxidation. In chapter 4 and 5, I succeeded in
introducing Cu into the orthorhombic structure by a hydrothermal method using (CH3NH3)6Mo7O24.
This synthesis method supplies an available approach to incorporate other metals into the
orthorhombic structure. So in this chapter, a series of orthorhombic Mo-V-M-O (M= Mn, Co, Ni, Zn,
Fe) were synthesized by a hydrothermal method using (CH3NH3)6Mo7O24 as Mo precursor.
Synthesized catalysts were conducted in gas phase selective oxidation of acrolein to acrylic acid. The
effects of each element were briefly investigated.

149
6.2 Experimental
6.2.1 Synthesis of orthorhombic Mo-V-M-O complex oxides.
A series of orthorhombic Mo-V-M-O (M= Mn, Co, Ni, Zn, Fe) catalysts were obtained
according to a previous report.[1, 2]
Methylammonium heptamolybdate (MAHM:
(CH3NH3)6Mo7O24) was used as the Mo source. The synthesis procedures for MAHM were as
follows.[3] 22.594 g of MoO3 (Kanto, 0.15 mol) were slowly dissolved in 33.2 ml of 40%
methylamine solution (methylamine: 0.300 mol, Wako) followed by the stirring for 30 min. After
being completely dissolved, the solution was evaporated under vacuumed condition of P/P0= 0.03 at
70 oC for 30 min and white solid was obtained. Time was measured from the time that white
powder was started to precipitate. Then the solid was dried in air at 80 oC overnight.
Orthorhombic Mo-V-M-O (M= Mn, Co, Ni, Zn, Fe) were synthesized by using MAHM,
VOSO4nH2O (n=5.4) and MSO4nH2O under a hydrothermal condition. Firstly, solution A was
obtained with 8.9 g of MAHM being dissolved in 120 ml of deionized water, and solution B was
obtained with 3.28 g of VOSO4nH2O (n=5.4, 3.28 g, Mitsuwa Chemicals) being dissolved in
another 120 ml of deionized water. Secondly, solution B was poured into solution A under a stirring
condition for 10min. Then 2.5 mmol of MSO4 (M= Mn2+, Co2+, Ni2+, Zn2+, Fe2+, Wako) were added
into the mixture under stirring condition for another 10min. The obtained mixed solution (purple,
pH3.2) was introduced into an autoclave with a Teflon sheet in the Teflon inner vessel. The
reaction mixture was purged with N2 for 10 min and then hydrothermally treated at 175 oC for 48 h.
As-synthesized Orthorhombic Mo-V-M-O samples (denoted as Orth-MoVO-Mn, Orth-MoVO-Co,
Orth-MoVO-Ni, Orth-MoVO-Zn, and Orth-MoVO-Fe) were treated with oxalic acid solution oxalic
acid (Wako) (0.4 mol L1, 25ml/ 1g solids) at 60 oC for 30min. Then obtained samples were dried in
air at 80 oC overnight.
All of the samples were calcined at 400 oC under air condition for 2h to remove ammonium
and organic part before characterization and activity testing.

6.2.2 Characterization.
X-ray diffraction patterns (XRD) were measured by using an X-ray diffractometer
(RINT-Ultima III, Rigaku) with Cu K to study their crystalline structure. Scanning transmission
electron microscopy (STEM) images and metal element mapping of Mo, V, and M (M= Mn, Co, Ni,
Zn, Fe) were obtained on an HD-2000 (HITACHI). The chemical composition of the catalysts was
determined by the inductively coupled plasma atomic emission spectroscopy (ICP-AES) method
with a VISTA-PRO apparatus (Varian). N2 adsorption-desorption isotherm measurements were

150
carried out on an auto-adsorption system (BELSORP MAX, Nippon BELL) to obtain the external
surface area and micropore volume using the t-plot method. Prior to adsorption measurements, the
samples were treated at 400 oC for 2 h under air and out-gassed at 300 oC under vacuum for 2 h.

6.2.3 Gas-phase catalytic oxidation.


Gas phase catalytic oxidation of acrolein to acrylic acid was performed using a fixed bed
stainless tubular reactor at atmospheric pressure. Catalysts (0.25 g) were firstly ground for 5
minutes, then diluted with 2.5 g Carborundum and pretreated at 400 oC under N2 of 50 mlmin-1 for
2 h. Reactant gases were H2O= 27.1 mlmin-1, acrolein= 2.5 mlmin-1, O2= 8 mlmin-1 and N2
balance (total: 107.6 mlmin-1). Quantitative analysis was performed using three on-line gas
chromatographs with columns of Molecular Sieve 13X, Gaskuropack 54, and Porapak Q. Blank
runs showed that no reaction took place without catalysts under the experimental conditions.
Carbon balance was always over 95%, and selectivity was calculated on the basis of products of
sum.

151
6.3 Results and discussion
6.3.1 Structural analysis of Orth-Mo-V-M-O
Crystalline structures of Orth-Mo-V-M-O (M= Mn, Co, Ni, Zn, Fe) were conducted through
XRD characterization and the results were illustrated in Figure 6-1. In these patterns, diffraction
peaks emerged at 6.6o, 7.9o, 9.0o and 27.3o, etc. which are assigned to the orthorhombic structure.[4]
Compare to the position of Orth-MoVO-MA (001) diffraction peak, (001) peak in other
Orth-Mo-V-M-O catalysts shifted in lower angle direction (Figure 6-1-II). However, broad peaks
are observed in the XRD patterns of Orth-MoVO-MA and Orth-MoVO-Mn. Similar with
orthorhombic Mo-V-W-O catalysts, broad peaks emerges in company with the decrease of crystal
size, and thus external surface area obviously increased. From Table 6-1 it is clearly noticed that the
external surface area of Orth-Mo-V-O-Mn and Orth-MoVO-MA catalyst increases to 18.5 m2g-1
and 29.0 m2g-1, respectively. Chemical composition of each catalyst can also be obtained from
Table 6-1. Contents of additional metals were adjusted to around 5 at.-%, and different Mo and V
compositions are attributed to the structural requirement of each catalyst. ICP results showed that
additional metals are surely incorporated in the catalysts. Moreover elemental mapping, as been
showed in the Figure 6-2, showed that all these metals are evenly dispersed in the catalyst. These
results strongly indicate that Mo-V-M-O (M= Mn, Co, Ni, Zn, Fe) catalysts with an orthorhombic
structure were successfully synthesized. Pore volumes of synthesized Mo-V-M-O were listed in the
Table 6-1. The pore volume of Orth-MoVO-AHM is 13.710-3mlg-1. It shows the microporosity
because of the existence of heptagonal channels in the orthorhombic structure. However, all pore
volumes of Mo-V-M-O decreased to about 110-3mlg-1. This indicated that the additional metals
are located in the heptagonal channels.

6.3.2 Catalytic performance of acrolein oxidation to acrylic acid over Orth-Mo-V-M-O


catalysts
Catalytic performance of acrolein oxidation to acrylic acid over Orth-Mo-V-M-O catalysts was
conducted in the gas phase acrolein oxidation to acrylic acid. The results were illustrated in Figure
6-3 and Figure 6-4. Orth-MoVO-MA shows the highest catalytic activity. More than 90% of
acrolein conversion is achieved at 215 oC. In contrast, the temperature of 90% acrolein conversion
over Orth-MoVO-AHM is 235 oC. These two catalysts both own orthorhombic structure. However,
external surface area of Orth-MoVO-MA (29.0 m2g-1), is much higher than that of
Orth-MoVO-AHM (5.0 m2g-1). As noticed in previous chapter, external surface area is of crucial
importance for crystalline Mo-V-based catalysts in acrolein oxidation to acrylic acid. Here, I also

152
consider the external surface area of Orth-Mo-V-M-O as an important factor to impact the catalytic
activity. With the increase of external surface area, activity gradually increased. For example,
Orth-MoVO-Mn and Orth-MoVO-Ni catalysts, that possess external surface area of 18.5 m2g-1 and
13.1 m2g-1 respectively, show moderate activity. Orth-MoVO-Zn and Orth-MoVO-Co catalysts,
that possess the external surface area of 5.7 m2g-1 and 7.3 m2g-1 respectively, exhibit less activity.
Certainly, despite the external surface area, there are other factors to impact the catalytic
activities of Orth-Mo-V-M-O catalysts, for example redox property. Oxides of valence alterable
metals such as Cu, Fe and Mn have removable lattice oxygens, so that they exhibit redox properties.
In a selective oxidation process ,catalyst is involved in a redox cycle. Strong redox properties of
metals will contribute to accelerate the redox cycle.[5, 6] This may explain Orth-MoVO-Co showed a
higher activity than that of Orth-MoVO-Zn, although the external surface area of Orth-MoVO-Co
(7.3 m2g-1) is close to that of Orth-MoVO-Zn (5.7 m2g-1).
Unlike the catalytic activity, the selectivity to acrylic acid, acetic acid and COx are only
slightly changed by the introduction of additional metals (Figure 6-4). All catalysts showed high
selectivity to acrylic acid, which means these catalysts are effective for acrolein oxidation to acrylic
acid. For the selectivity to acetic acid and COx, although Orth-MoVO-Co and Orth-MoVO-Zn
showed less catalytic activity, selectivity of these two catalysts are little higher than that of other
catalysts. This may caused by the lower formation rate of acrylic acid over these two catalysts.
Acrolein oxidation to acrylic acid is a strong exothermic reaction. Higher external surface area
supplies more active sites. The formation rate of acrylic acid will rise in company with the increase
of external surface area. So a large amount of heat from the oxidation process was produced, and it
leads to deep oxidation of acrylic acid to acetic acid and COx.[7] This explains the reason why the
most active catalyst Orth-MoVO-MA exhibits the poorest selectivity.

6.3.3 Conclusions
Using (CH3NH3)6Mo7O24) as Mo precursor, crystalline Mo-V-M-O (M= Mn, Co, Ni, Zn, Fe)
catalysts with an orthorhombic structure were successfully synthesized by a hydrothermal method.
Additional transition metals are located in the heptagonal channels. These catalysts are preformed in
the gas phase selective oxidation of acrolein to acrylic acid. Addition of metals into crystalline
structure leads to the change of external surface area, by which Mo-V-M-O catalysts showed
different catalytic activity while selectivity to acrylic acid was remained. External surface area is
not only related to the catalytic activity but also the selectivity. Transition metals are reported to be
effective for many other oxidation reactions. Succeeded in adding these metals into the crystalline

153
structure will contribute to expanding the application domain of crystalline Mo-V-based complex
oxides.

154
References

[1] Chen C., Kosuke N., Murayama T., Ueda W. Chemcatchem, 2013, 5, 2869-2873.
[2] Konya T., Katou T., Murayama T., Ishikawa S., Sadakane M., Buttrey D., Ueda W. Catalysis
Science & Technology, 2013, 3, 380-387.
[3] Ishikawa S., Murayama T., Ohmura S., Sadakane M., Ueda W. Chemistry of Materials, 2013, 25,
2211-2219.
[4] Sadakane M., Kodato K., Kuranishi T., Nodasaka Y., Sugawara K., Sakaguchi N., Nagai T.,
Matsui Y., Ueda W. Angewandte Chemie-International Edition, 2008, 47, 2493-2496.
[5] Andrushkevich T.V. Catal.Rev.-Sci.Eng., 1993, 35, 213-259.
[6] Tichy J. Applied Catalysis a-General, 1997, 157, 363-385.
[7] Roduner E., Kaim W., Sarkar B., Urlacher V.B., Pleiss J., Glaser R., Einicke W.D., Sprenger
G.A., Beifuss U., Klemm E., Liebner C., Hieronymus H., Hsu S.F., Plietker B., Laschat S.
Chemcatchem, 2013, 5, 82-112.

155
Figures and tables
Table 6-1 Chemical compositions and external surface areas of Orth-MoVO-M catalysts.

a
Composition (at.-% ) External surface area Pore volume
Sample b 2 -1
Mo V M (m g ) (10-3mlg-1)

Orth-MoVO-AHM 71.5 28.5 - 5.0 13.7


Orth-MoVO-Fe 72.8 23.7 3.5 9.5 1.8
Orth-MoVO-Zn 72.9 22.8 4.3 5.7 0.8
Orth-MoVO-Ni 77.6 18.0 4.5 13.1 1.3
Orth-MoVO-Co 76.0 18.6 5.4 7.3 1.2
Orth-MoVO-Mn 76.1 19.2 4.6 18.5 1.6
Orth-MoVO-MAHM 76.6 23.4 - 29.0 6.3

a
Determined by ICP-AES
b
Measured by N2 adsorption and determined by t-plot method

156
Figure 6-1 XRD patterns of Orth-MoVO-M (M=Mn, Co, Ni, Zn, Cu, Fe)

157
Figure 6-2 STEM image and elemental mapping of Orth-MoVO-M (M=Mn, Co, Ni, Zn, Fe)

158
Figure 6-3 Acrolein conversion over Orth-MoVO-M catalysts: Orth-MoVO-AHM(),
Orth-MoVO-Fe(), Orth-MoVO-Zn (), Orth-MoVO-Ni( ), Orth-MoVO-Co ( ),
Orth-MoVO-Mn (), Orth-MoVO-MAHM().

159
Figure 6-4 Selectivity to acrylic acid (a), acetic acid (b), and COx (c) over Orth-MoVO-M catalysts:
Orth-MoVO-AHM(), Orth-MoVO-Fe(), Orth-MoVO-Zn (), Orth-MoVO-Ni( ),
Orth-MoVO-Co ( ), Orth-MoVO-Mn (), Orth-MoVO-MAHM().

160
Chapter 7 General conclusion

161
162
In this thesis, author mainly focuses on the synthesis of highly efficient crystalline
Mo-V(-W-Cu)-O catalysts for the selective oxidation of acrolein to acrylic acid. A series of
crystalline Mo-V-W-O, Mo-V-Cu-O, and Mo-V-W-Cu-O catalysts were successfully synthesized by
a hydrothermal method. Selective oxidation of acrolein to acrylic acid was performed in gas phase
over these catalysts. Effects of W and Cu on acrolein oxidation were also revealed. To obtain the most
efficient Mo-V-W-Cu-O catalysts for acrylic acid production, experiments in this thesis were
conducted through three stages:
(1) Synthesis of orthorhombic, trigonal, and amorphous Mo-V-W-O catalysts with various W
contents. In this stage, crystalline structure was found to be very important for the transformation of
acrolein to acrylic acid in the gas phase condition. Catalytic activity for acrolein selective oxidation
over orthorhombic and trigonal Mo-V-W-O is much higher than amorphous Mo-V-W-O and
tetragonal Mo-V-O. Addition of W into an orthorhombic structure exhibits positive effect in the
reaction that is superior to trigonal case. W is incorporated into the framework of crystalline Mo-V-O
structure. W-containing catalysts showed less water dependency on acrolein oxidation. The
introduction of W leads to a rod-segregation effect on orthorhombic structure, by which external
surface area greatly increases and catalytic activity is enhanced.
(2) Synthesis of orthorhombic Mo-V-Cu-O catalysts with various Cu contents. In this stage, Cu
is incorporated into the heptagonal channels. Addition of Cu was found to be effective to enhance the
acrylic acid selectivity in the gas phase selective oxidation of acrolein to acrylic acid.
(3) Introducing both W and Cu into orthorhombic Mo-V-O structure to synthesize highly
efficient catalyst for acrolein oxidation. In this stage, a large number of orthorhombic Mo-V-W-Cu-O
catalysts with different W and Cu contents were synthesized and conducted in gas phase selective
oxidation of acrolein to acrylic acid to obtain the best formulation. Balance the catalytic activity and
selectivity, Orth-Mo-V-W7.5%-Cu7.5%-O is proposed to be the most efficient catalyst for the
selective oxidation of acrolein to acrylic acid.
In summarize, industrial Mo-V oxide catalysts always contain W and Cu as additives for
selective oxidation of acrolein to acrylic acid. The present work clearly showed the functions of W
and Cu for acrolein oxidation to acrylic acid. Using a hydrothermal method, highly efficient
orthorhombic Mo-V-W-Cu-O catalyst for acrolein oxidation to acrylic acid was successfully
synthesized. This orthorhombic Mo-V-W-Cu-O catalyst with high catalytic activity and selectivity
shows good application prospect for acrolein oxidation to acrylic acid.
Crystalline structure of Mo-V-based catalysts is a guarantee for good catalytic performance in
acrolein oxidation to acrylic acid, and external surface area is important factor to impact the catalytic

163
activity. The synthesis approach of using (CH3NH3)6Mo7O24 in the hydrothermal condition is
available to introduce other transition metals such as Mn, Fe, Co, and Ni into the orthorhombic
structure to optimize the structure. These orthorhombic Mo-V-M-O complex oxides are also good
catalysts for the acrolein oxidation to acrylic acid. More importantly, they are expected to expand the
application domain of crystalline Mo-V-based complex oxides.

164
Acknowledgment
This thesis is finished under Professor Wataru Uedas supervision. Most of the researches

presented in this thesis were conducted at Catalysis Research Center, Hokkaido University in three

years (from October, 2012 to September, 2015)

First of all, I would like to express my sincere thanks to Professor Ueda. His wide knowledge

and patient guidance make me learn more when doing research. His discussion often provides

important information to solve the problems and gives interesting ideas for further investigation.

His encouragement also makes me recovered when I suffered from setback.

I also would like to thank Professor Kenichi Shimizu. His kind help on manuscript preparation

is very important to finish this thesis.

I also would like to thank Assistant Professor Toru Murayama for his kind assistance on daily

experiments and discussion.

I am grateful to Associate Professor Chen, Mr. Zhang, Mr. Ishikawa, Mr. Goto, Mr. Nakamura,

Ms. Omata, and other members in Ueda Lab for their kind help, suggestion, and discussion.

I would like to thank CSC for financial support.

Finally, I would like to thank my family. They are my strong shield all the time.

Chuntian Qiu

165

Вам также может понравиться