Synthesis of oxymethylene ethers

122
Synthesis of Oxymethylene Ethers Dissertation Zur Erlangung des akademischen Grades Doktor der Naturwissenschaften (Dr. rer. nat.) vorgelegt an der Fakultät für Chemie und Biochemie der Ruhr-Universität Bochum von Anna Grünert Bochum Oktober 2019

Transcript of Synthesis of oxymethylene ethers

Page 1: Synthesis of oxymethylene ethers

Synthesis of Oxymethylene Ethers

Dissertation

Zur Erlangung des akademischen GradesDoktor der Naturwissenschaften (Dr. rer. nat.)

vorgelegt an der Fakultät für Chemie und Biochemieder Ruhr-Universität Bochum

von

Anna Grünert

BochumOktober 2019

Page 2: Synthesis of oxymethylene ethers
Page 3: Synthesis of oxymethylene ethers

Die vorliegende Arbeit wurde in der Zeit von Februar 2016 bis Oktober 2019 in der Abteilung

für Heterogene Katalyse am Max-Planck-Institut für Kohlenforschung in Mülheim an der

Ruhr unter der Leitung von Prof. Dr. Ferdi Schüth angefertigt.

Referent: Prof. Dr. Ferdi Schüth

Korreferent: Prof. Dr. Martin Muhler

Page 4: Synthesis of oxymethylene ethers
Page 5: Synthesis of oxymethylene ethers

I

I Acknowledgements

Firstly, I would like to thank my team of supervisors, Prof. Dr. Ferdi Schüth andDr. Wolfgang Schmidt, and my co-examiner Prof. Dr. Martin Muhler.

Ferdi, I would like to thank you for your trust that is the basis of the exceptionalfreedom of work, which you grant not only to me, but to all of your PhD students. Yougave me the resources, time and liberty to develop my PhD project in my own way, tomake mistakes, to solve challenging problems and to grow as a person. I am gratefulfor your appreciation of a cooperative and welcoming atmosphere in the group, whichis most apparent in your yearly invitation to the group trip to Oberwesel.

Wolfgang, I owe my thanks to you for your advice on many topics including materialsynthesis, catalyst characterisation and manuscript writing. I very much appreciateyour welcoming and relaxed attitude.

I am also thankful that to you both that you enjoy sharing your knowledge andexperience with me and my colleagues in catalysis seminars and other technicalseminars and in the focused project meetings.

Prof. Muhler, thank you for taking interest in my work and for agreeing to be theco-examiner.

The practical realisation of this work would not have been possible without the help of thestaff of service departments, the technical staff of Schüth group and fellow PhD students.

This is why I feel fortunate to have worked in close cooperation with the team of theFeinmechanics workshop including Wolfgang Kersten, Dirk Ullner, Knut Gräfenstein,Jürgen Majer, Sebastian Plankert and Ralf Thomas. Big thanks to all of you for yourtechnical support and your exceptional willingness to help students in building andmaintaining catalytic test equipment.

Dirk, thank you so much for your fast and competent support when my set-up gave metroubles. It was also a pleasure to learn about set-up design, properties of steels andother materials and to get an introduction to the tools used in the workshop. Knut andDirk, I will truly miss your ability to cheer me up with your special humour. I alwaysfelt welcomed at the Feinmechanics workshop.

I would also like to acknowledge the contributions of locksmith and glassblowerworkshops to my project.

In the pressure lab, where my catalytic test set-up was located, Nils Theyssen, NiklasFuhrmann and Lars Winkel made sure that everyday operation of set-ups and lab

Page 6: Synthesis of oxymethylene ethers

II

infrastructure ran smoothly. I owe my thanks to you and to my box neighbours ReneAlbert, Dr. Robert Urbanczyk, Kateryna Peinecke and Özgül Sener for regularlyhelping me out with technical advice and spare parts.

In the main lab, the technical staff André Pommerin, Laila Sahraoui and Flo Baum andtheir team of apprentices made sure that the lab was always safe and functional. Thankyou for being approachable and helpful and for knowing where to find all kinds ofthings.

I also owe many thanks to the staff of NMR, electron microscopy, HPLC, GC and ITdepartments including Dr. Bodo Zibrowius, Silvia Palm, Georg Breitenbruch, FrankKohler, Philipp Schulze, Marjan Tomas and Marcus Hermes.

I learned a lot from my colleagues Dr. Nicolas Duyckaerts, Dr. Ioan-Teodor Trotus,Dr. Daniel Wendt, Dr. Cristina Ochoa-Hernández, Dr. Pit Losch and Hrishikesh Joshi.

Nico, Teo and Daniel, thank you for your patience in sharing your knowledge aboutSwagelok, flow set-ups and reactors. I am grateful for your input regarding the designand assembly of my test set-up at the beginning of my PhD project. Cristina, Pit andHrishi, I appreciated your interest in my project and our fruitful discussions aboutchemistry.

Of course, there is more to a PhD than working in the lab. I feel thankful for having gotten toknow so many inspiring, kind and open fellow researchers. You have made me feel deeplyconnected to this group and have made my PhD time very enjoyable.

I would like to express my gratitude towards my colleagues Özgül Agbaba Sener, ReneAlbert, Dr. Ghith Al-Shaal, Dr. Amol Amrute, Alex Bähr, Adrian Barranco, TugceBeyazay, Marius Bilke, Alex Bodach, Joel Britschgi, Eko Budiyanto, Dr. MatthewClough, Dr. Yitao Dai, Dr. Isabel de Freitas, Jacopo de Bellis, Dr. Michael Dierks,Dr. Georgios “Jorro” Dodekatos, Dr. Rene Eckert, Dr. Michael Felderhoff, JessicaGonzález, Alexander Hopf, Dr. Daniel Jalalpoor, Kai Jeske, Hrishikesh Joshi,Dr. Marco Kennema, Dr. Jonglack Kim, Klara Kley, Dr. Pit Losch, Dr. Gun-HeeMoon, Dr. Valentina Nese, Edward Nürenberg, Dr. Cristina Ochoa-Hernández, EzgiOnur Sahin, Dr. Seyma Ortatatli, Dr. Jochen Ortmeyer, Kateryna Peinecke, Dr. HilkePetersen, Dr. Christian Pichler, Dr. Gonzalo Prieto, Steffen Reichle, Dr. BidyutSarma, Dr. Hannah Schreyer, Dr. Stefan Schünemann, Niklas Stegmann, Dr. HarunTüysüz, Dr. Robert Urbanczyk, Dr. Olena Vozniuk, Dr. Claudia Weidenthaler,Frederik Winkelmann, Mingquan Yu and all other colleagues that I may have forgottento mention.

Page 7: Synthesis of oxymethylene ethers

III

Thank you for spending time with me in the lab, at Mensa, in the social room, at lunchgroup, at group and cake seminars and other social events. Also, I am glad that some ofyou never forgot how to celebrate and dance.

Edward, Niklas, Alex, Özgül and Cristina, it was great to share an office with you. Ihave appreciated the focused working atmosphere on the one hand and your opennessto share matters of every-day life on the other hand. Thank you for your moral supportand your valuable advice.

Annette Krappweis und Kirsten Kalischer, your assistance in organisational issues wasof great help.

For financial support of my PhD, I would like to acknowledge the Max Planck Society andFonds der chemischen Industrie.

My participation in the JungChemikerForum Mülheim during my time at the Max-Planck-Institut für Kohlenforschung was also a great experience. It made me feel more connected tocolleagues from the other research groups of this institute

Minh Dao, Lorenz Löffler, Jonas Börgel, Jens Rickmeier, Suzanne Willems, FabioCaló, Christine Schulz, Tobias Biberger, Julius Hillenbrand, Van Anh Tran, MarcHeinrich, Sebastian Beeg, Simon Musch, Christina Erken and those that I unfortunatelyforgot to mention, I would like to thank you for great team work and for your trustwhen I was your speaker.

Last, but not least, I would like to express my gratitude towards my parents Wolfgang andTanja Grünert and my partner Arne Schüttler. They have supported me during my bachelor’sand master’s degree and during my PhD project.

Dad, thank you for passing on your appreciation for sciences, literature and music andespecially for sharing your passion for chemistry.

Mum, thank you for your unconditional moral support in all situations and for teachingme openness and curiousness.

Arne, thank you for sharing your daily life with me, including its challenges and joyfulmoments. I am deeply grateful for your continuous trust and encouragement over theyears.

Page 8: Synthesis of oxymethylene ethers

IV

Page 9: Synthesis of oxymethylene ethers

V

II Table of contents

I Acknowledgements ............................................................................................................. III Table of contents ................................................................................................................ VIII Abbreviations ................................................................................................................. VIII1 Thesis abstract ..................................................................................................................... 12 Introduction ......................................................................................................................... 2

2.1 Synthetic fuels ........................................................................................................... 22.1.1 Chemical CO2 recycling .................................................................................... 22.1.2 Drop-in fuels ...................................................................................................... 32.1.3 Reduction of diesel emissions ............................................................................ 4

2.2 Oxymethylene ethers ................................................................................................. 52.2.1 Physicochemical and fuel related properties of OME ....................................... 52.2.2 Development and future challenges of OME research ...................................... 72.2.3 Synthesis ............................................................................................................ 8

2.3 Solid acid catalysis .................................................................................................. 122.3.1 Zeolites ............................................................................................................. 132.3.2 Supported liquid phase catalysts ...................................................................... 162.3.3 Characterisation methods ................................................................................. 18

2.4 Methanol dehydrogenation ...................................................................................... 202.4.1 Oxidative vs. non-oxidative route .................................................................... 202.4.2 Catalysts ........................................................................................................... 21

3 Motivation and research objectives ................................................................................... 234 Description of test set-up ................................................................................................... 25

4.1 Concept ................................................................................................................... 254.2 Technical implementation ....................................................................................... 26

4.2.1 Gas and pressure control .................................................................................. 274.2.2 Evaporator unit ................................................................................................ 274.2.3 Heating ............................................................................................................. 284.2.4 Reactor ............................................................................................................. 28

4.3 Product analysis ...................................................................................................... 294.4 Safety features ......................................................................................................... 304.5 Extensions for combined process ............................................................................ 31

5 Screening of reaction conditions ....................................................................................... 335.1 Temperature ............................................................................................................ 335.2 Pressure ................................................................................................................... 365.3 Reactant ratio .......................................................................................................... 37

Page 10: Synthesis of oxymethylene ethers

VI

5.4 Water content .......................................................................................................... 385.5 Pellet Size ................................................................................................................ 385.6 Catalyst activation protocol .................................................................................... 385.7 Reproducibility ....................................................................................................... 39

6 Preliminary catalyst screening .......................................................................................... 407 OME synthesis over zeolite catalysts ................................................................................ 42

7.1 Catalyst screening ................................................................................................... 427.2 Correlation between acid site properties and catalyst performance ........................ 437.3 Influence of particle size and external surface area ................................................ 497.4 Adaptation of reaction conditions ........................................................................... 507.5 Catalyst deactivation and regeneration ................................................................... 517.6 Comparison of siliceous materials .......................................................................... 527.7 Conclusions ............................................................................................................. 54

8 OME synthesis over supported phosphoric acid ............................................................... 558.1 Catalyst Characterisation ........................................................................................ 568.2 Preliminary studies of exemplary H3PO4/C catalyst ............................................... 598.3 Impact of H3PO4 loading ........................................................................................ 608.4 Sodium phosphates ................................................................................................. 618.5 Comparison with benchmark zeolite ...................................................................... 638.6 Conclusions ............................................................................................................. 65

9 Two-step synthesis of OME from methanol ..................................................................... 669.1 Thermal decomposition of formaldehyde ............................................................... 679.2 Catalyst screening ................................................................................................... 689.3 Combined process ................................................................................................... 71

10 Summary and final remarks .............................................................................................. 7411 Experimental ..................................................................................................................... 76

11.1 Commercial materials ............................................................................................. 7611.1.1 Gases ................................................................................................................ 7611.1.2 Chemicals ........................................................................................................ 7611.1.3 Catalysts and other solid materials .................................................................. 77

11.2 Synthesis of catalysts .............................................................................................. 7811.2.1 Supported silicotungstic acid ........................................................................... 7811.2.2 SBA-15-SO3H ................................................................................................. 7811.2.3 Silicalite-1 ........................................................................................................ 7811.2.4 Supported phosphoric acid .............................................................................. 7911.2.5 Methanol dehydrogenation catalysts ............................................................... 79

11.3 Modification procedures ......................................................................................... 8011.3.1 Catalyst activation ........................................................................................... 80

Page 11: Synthesis of oxymethylene ethers

VII

11.3.2 Sodium exchange of zeolites ........................................................................... 8011.3.3 Oxalic acid treatment of zeolites ..................................................................... 8111.3.4 Regeneration protocols .................................................................................... 81

11.4 Characterisation methods ........................................................................................ 8111.4.1 X-ray powder diffraction (PXRD) ................................................................... 8111.4.2 Temperature programmed desorption of ammonia (NH3-TPD) ...................... 8111.4.3 Pyridine adsorption followed by FTIR spectroscopy (Py-FTIR) .................... 8211.4.4 Magic-angle spinning nuclear magnetic resonance (MAS-NMR) .................. 8211.4.5 Thermogravimetric analysis coupled with mass spectrometry (TG-MS) ........ 8211.4.6 Diffuse reflectance infrared spectroscopy (DRIFTS) ...................................... 8311.4.7 Nitrogen physisorption .................................................................................... 8311.4.8 GC-MS ............................................................................................................. 8311.4.9 Scanning electron microscopy (SEM) ............................................................. 8311.4.10 Energy dispersive X-ray spectroscopy (EDX) ................................................. 8311.4.11 Transmission electron microscopy (TEM) ...................................................... 8411.4.12 Elemental analysis ........................................................................................... 84

11.5 Batch reactions ........................................................................................................ 8411.6 Wet-chemical analysis methods .............................................................................. 84

11.6.1 Preparation of methanolic formaldehyde solution ........................................... 8411.6.2 Iodometry ......................................................................................................... 8411.6.3 Karl-Fischer titration ........................................................................................ 85

12 Appendix ........................................................................................................................... 8613 References to laboratory journal entries .......................................................................... 10114 References ....................................................................................................................... 104

Page 12: Synthesis of oxymethylene ethers

VIII

III Abbreviations

Table 1.1: Abbreviations.

ALPO aluminophosphates pKa negative logarithm of aciddissociation constant

BET Brunauer-Emmett-Teller PM particulate matter

CI compression ignition POM polyoxymethylene (polymer)

Xmax maximum conversion POMDME polyoxymethylene dimethyl ethers

DME dimethyl ether PXRD X-ray powder diffraction

DRIFTS diffuse reflectance infraredspectroscopy

rpm revolutions/rotations per minute

EDX energy dispersive X-rayspectroscopy

S selectivity

EFAl extra-framework aluminium SAPO silicoaluminophosphates

FA formaldehyde SAR silica-to-alumina ratio

FID flame ionization detector SEM scanning electron microscopy

FTIR Fourier transform infrared(spectroscopy)

SI spark-ignition

GC gas chromatography orgas chromatograph

SILP supported ionic liquid phase

GHG greenhouse gas Si-OH silanol group

HC hydrocarbons SLP supported liquid phase

HPA heteropoly acid Smax maximum selectivity

HPLC high-performance liquidchromatography

SPA phosphoric acid supported on silica

IR infrared STP standard temperature and pressure

Ka dissociation constant synair synthetic air

MAS magic angle spinning T in TO4 metal in tetrahedral oxygenenvironment

MeOH methanol TCD Thermal conductivity detector

Page 13: Synthesis of oxymethylene ethers

IX

MFC mass flow controller TEM transmission electron microscopy

mol% mole fraction TG thermogravimetric analysis

MS mass spectrometry TPD temperature programmeddesorption

NMR nuclear magnetic resonance TRI trioxane

NOx nitrogen oxides WHSV weight hourly space velocity

OME oxymethylene ether wt% mass fraction

PF paraformaldehyde

Page 14: Synthesis of oxymethylene ethers

1 Thesis abstract

1 Thesis abstract

Oxymethylene ethers (OME) are a class of chain ethers that have been classified as pollutant

reducing synthetic diesel additives in the late 1990s. In view of growing efforts to reduce

hazardous emissions from the transport sector and to find alternatives to fossil-based fuels,

OME have seen a rise in academic and industrial interest. Various synthesis routes that have

methanol as a common intermediate have been reported. However, the production remains the

main challenge for introduction of OME as a synthetic fuel.

This work explores the gas-phase synthesis of OME from methanol and formaldehyde as an

alternative approach to common liquid-phase routes. In particular, the investigation of solid

catalysts for this reaction is the focus of the conducted studies.

For this purpose, a versatile test set-up was built and suitable reaction conditions were

identified. The comparison of a selection of solid acid catalysts highlighted the activity of

zeolites in gas-phase OME synthesis. In a systematic study of a broad range of zeolite

catalysts, a relation between catalyst reactivity and silica-to-alumina ratio was established. It

could be shown that low amounts of acid sites are favourable for OME selectivity and that

strong acid sites are linked to by-product formation. Carbon supported phosphoric acid was

furthermore found to be an active catalyst for the gas-phase synthesis of OME with a superior

lifetime as compared to benchmark zeolite catalysts. Steady-state conversion and selectivity

were comparable at the same loading of active centres. In view of the simple preparation and

low cost of H3PO4/C, it provides an attractive alternative to zeolites. In a final step, the

viability of the gas-phase synthesis of OME from methanol without separation of

intermediates was demonstrated.

Page 15: Synthesis of oxymethylene ethers

Introduction 2

2 Introduction

The aim of this chapter is to give background information that is relevant in the context of this

work on gas-phase synthesis of oxymethylene ethers, including socio-economic

considerations, properties of OME and catalysts, and state of the art of synthetic procedures,

production processes and characterisation techniques.

Firstly, the importance of research on synthetic fuels in general and the potential positive

contribution to the development of our future transportation sector will be discussed in chapter

2.1. Secondly, oxymethylene ethers as a promising class of synthetic fuels will be introduced

in chapter 2.2 with a focus on the development of OME research, physicochemical and fuel-

related properties as well as synthetic routes. As the investigations presented in this thesis rely

on solid acid catalysis, the two solid acid classes mainly employed in this work, namely

zeolites and supported phosphoric acid, will be presented in chapter 2.3 with an emphasis on

acid properties. Owing to the broad range of techniques available for acid characterisation, an

overview highlighting advantages and limits of the methods is also included. Finally, the

partial methanol dehydrogenation is a relevant intermediate reaction for a potential OME

synthesis from carbon dioxide (CO2) feedstock. Related reaction characteristics and catalyst

classes are summarised in chapter 2.4.

2.1 Synthetic fuels

In the research field of synthetic fuel, the academic and industrial interest is mainly driven by

socio-economic and health related aspects. These include the potential reduction of

anthropogenic CO2 emissions by large scale usage of CO2 as a feedstock for synthetic fuels as

discussed in chapter 2.1.1. Furthermore, the facilitated implementation of liquid synthetic

fuels as so-called drop-in fuels using existing infrastructure is reviewed in chapter 2.1.2. A

third important aspect highlighted in chapter 2.1.3 is the potential of emission reducing

synthetic fuels to alleviate health issues related to air pollution.

2.1.1 Chemical CO2 recycling

In current research, many efforts are directed towards mitigation of CO2 emissions and

development of CO2 recycling and storage strategies. This interest is sparked by the goal to

diminish the effect of global warming. The latter is the amplification of the naturally occurring

greenhouse effect by emission of greenhouse gases (GHG) such as carbon dioxide, methane

Page 16: Synthesis of oxymethylene ethers

3 Introduction

(CH4), nitrous oxide (N2O) and fluorinated compounds from anthropogenic sources. CO2 is

rated to have the largest impact on global warming with 76% of total anthropogenic GHG

emissions.1

The transformation of CO2 to synthetic fuels can be classified as a CO2 recycle approach.

While CO2 can also be recycled and used for the production of chemicals such as urea,

salicylic acid and polycarbonates, its transformation to fuels can have a greater leverage owing

to larger fuel demand.2 Synthetic fuels can contribute to CO2 reduction from the transport

sector, which accounted for 23% of global CO2 emissions in 2013. More specifically, fuels

have a major impact in the road sector that represented a three quarters share of the transport

emissions and was driving its growth.1, 3

As of today, most synthetic fuels are still produced from fossil feedstocks such as mineral oil,

natural gas and coal. However, processes for the supply of CO2 as a feedstock by isolation

from industrial exhaust gases, biomass or via direct air capture are currently in development.4

Hydrogen is also required for the valorisation of CO2. Similarly, environmentally benign

processes for its production are not employed in large scale yet, but water electrolysis is a

viable approach. Interestingly, the recent drop in electricity prices from renewables has

sparked economical interest in power-to-liquid technologies. For example, a Norwegian

company is targeting to produce synthetic “e-diesel” from electrolysis hydrogen and recycled

CO2 in industrially relevant quantities from 2020 based on hydro powder.5

2.1.2 Drop-in fuels

When discussing CO2 based synthetic fuels in general, methanol (MeOH), dimethyl ether

(DME), higher alcohols, hydrocarbons and oxymethylene ethers (OME) are of importance. It

is necessary to consider the different requirements towards distribution infrastructure and

combustion in engines. While DME is a suitable fuel with regards to many important engine

related parameters, its gaseous state of matter at standard conditions requires adaptation of

infrastructure and vehicles to liquefied gas handling.6 It is argued that drop-in fuels, which are

liquid synthetic fuels that can directly substitute conventional fuels with only minor

modifications, will have a lower market introduction barrier. Also in the long run, the

development of non-fossil based liquid fuels will be important for applications that cannot

easily be equipped with pressurized tanks, batteries or fuel cells, for example aviation and

marine transport.

Page 17: Synthesis of oxymethylene ethers

Introduction 4

In road transportation, spark-ignition (SI) and compression ignition (CI) engines are

commonly used. In terms of efficiency, operating cost and CO2 emission per distance

travelled, the diesel fuelled CI engine is clearly favourable.7, 8 For CI engines, hydrocarbons

and oxymethylene ethers (OME) are the most prominent examples for synthetic drop-in fuels

that can potentially be produced on the basis of CO2 feedstock. Hydrocarbons can be

synthesised via the Fischer-Tropsch process. Information of the synthesis of OME will be

supplied in chapter 2.2.2. In comparison to Fischer-Tropsch diesel, OME have two interesting

additional advantages, namely the low toxicity and the pollutant reducing properties as

discussed in the following section.

2.1.3 Reduction of diesel emissions

Even though technologies such as battery or fuel cell powered vehicles are emerging, internal

combustion engines still dominate road transport worldwide and will certainly continue to

represent a large share of vehicles also in the medium-term. It is therefore interesting to

analyse the impact of internal combustion engines on the environment and approaches to limit

harmful effects.

As mentioned above, compression ignition engines outperform spark-ignition engines in terms

of efficiency, operating cost and CO2 emission per distance travelled.7, 8 However, a major

disadvantage is increased exhaust gas pollutant emissions from CI engines. Emitted pollutants

include mainly particulate matter (PM, soot) and nitrogen oxides (NOx) as well as lower

amounts of carbon monoxide (CO) and hydrocarbons (HC). Particle emissions from diesel

engines are 6-10 times higher than from gasoline engines.7

The pollutants pose a severe risk to the human health as they can cause heart and lung diseases

as well as strokes. According to the World Health Organization, 4.2 million premature deaths

per year are a consequence of ambient air pollution.9 Further adverse effects include acid rain,

ground-level ozone and reduced visibility.

Legislation dealing with air pollutants has been developing towards lower emission levels in

many countries worldwide.10 In this context, approaches to reduce pollutants from road

transport have gained in importance.11 Common strategies include engine improvements and

exhaust aftertreatment. For exhaust gas treatment of CI engines, diesel oxidation catalyst for

removal of CO and HC, particulate filters and catalytic abatement of nitrogen oxides (selective

catalytic reduction (SCR) and nitrogen storage and reduction (NSR)) can be applied.7, 12 In

Page 18: Synthesis of oxymethylene ethers

5 Introduction

addition to the mentioned routes, potential also lies in reducing the initial formation of

pollutants by using emission reducing fuel additives and clean burning synthetic fuels.

Emission reducing diesel additives are mainly hydrocarbon compounds that contain oxygen

functionalities and are referred to as oxygenates.8 For SI engines methanol, ethanol and

methyl-tertbutyl ether can be used.13 These are however not suitable for CI engines due to

fundamentally differing working principles and therefore differing requirements for diesel

fuels. For CI engines, oxymethylene ethers (OME) are particularly suitable owing to their

pronounced pollutant reducing effect and favourable physicochemical properties.

2.2 Oxymethylene ethers

2.2.1 Physicochemical and fuel related properties of OME

OME are a series of homologous chain ethers with the chemical formula CH3O(CH2O)nCH3 ,

n denoting the length of the central ether chain in the abbreviation OMEn (see Figure 2.1).

Figure 2.1: Chemical structure of oxymethylene ethers, n denoting the number of repeating units.

With regards to chemical stability, OME are stable in alkaline and neutral medium and are

hydrolysed under acidic conditions as other acetal containing compounds. Their

physicochemical properties depend on chain length (see Table 2.1). With increasing length,

boiling and melting point, density, cetane number, flash point, viscosity, surface tension, and

oxygen content increase.6, 14, 15

Table 2.1: Selected physicochemical and fuel properties of OME1-5.14, 16

OMEn

oligomeroxygencontent(wt%)

density at20 °C

(g/cm3)

meltingpoint(°C)

boilingpoint(°C)

flashpoint(°C)

cetanenumber

OME1 42.1 0.868 -105 42 -32 28OME2 45.2 0.971 -70 105 16 68OME3 47.0 1.035 -43 156 54 72OME4 48.1 1.079 -10 202 88 84OME5 48.9 1.111 18 242 115 93

Page 19: Synthesis of oxymethylene ethers

Introduction 6

OME1, also referred to as dimethoxymethane or methylal, is the shortest homologue and is a

well-established solvent used for industrial applications in plastics and perfume industry17, 18

as well as in organic synthesis.19 Its industrial production is commonly based on reactive

distillation from methanol and aqueous formaldehyde solution.20-23 With respect to its use as

fuels, OME1 and OME2 are more volatile than conventional diesel, but could be used in

modified devices.15, 24 Owing to its low flash point, OME2 has a potential application as a pilot

fuel for methanol based spark-ignition engines.25

The physicochemical properties of the intermediate length homologues OME2-5 fulfil fuel

requirements fully, such as flash point and cetane number, or partially, for example lubricity,

kinematic viscosity and surface tension.14, 26 Due to the increased melting points of OME

chains with more than five repeating units, the risk of precipitation in the engine makes those

homologues unsuitable for use as fuel. Hence, OME2-5 or OME/diesel blends are regarded as

potential drop-in fuels that could be used in conventional motors with only minor adjustments

as highlighted in chapter 2.1.2.4

Owing to the oxygen content, the heating value of OME is lower than of diesel fuel (higher

heating value of OME3-5: approx. 21 MJ/kg, diesel: approx. 45 MJ/kg).14 This results in higher

gravimetric fuel consumption. The increase in volumetric fuel consumption will however be

less pronounced, due to the higher density of OME.

As mentioned in chapter 2.1.3, a major advantage of oxygenates are the pollutant reducing

properties. These were identified and patented by Moulton and Naegeli.24, 27 The most

pronounced reduction is achieved with regards to soot and particulate matter as has been

demonstrated for OME124, 28-31 as well as OME2-56, 32-40 by various research groups. It is

reasoned that neat OME combusts nearly soot-free due to lack of C-C bonds.41-43 While the

absolute values of soot reduction depend on blends, reference fuel, operating points and

engine type,28 it is evident that the soot reducing properties are not solely an effect of diesel

substitution. For example, for a diesel blend containing 20% OME3-4, a decrease in soot

emission of 60% and decrease in particulate mass by 40% in raw emissions of a six-cylinder

engine without after-treatment was reported.6 Due to reduced soot formation, engine

parameters such as exhaust gas recirculation can be adjusted in a wider range, allowing to

decease also harmful NOx emissions.31, 38

Page 20: Synthesis of oxymethylene ethers

7 Introduction

An additional general benefit of OME are the high cetane numbers, which can improve engine

efficiency.14 Cetane numbers of OME>1 are significantly higher than the defined lower limit of

51 (see Table 2.1).26 Finally, full miscibility of OME with diesel fuel and its low toxicity are

advantageous.18

2.2.2 Development and future challenges of OME research

The description of chain ethers with the chemical formula CH3O(CH2O)nCH3, referred to as

oxymethylene ethers (OME) or polyoxymethylene dimethyl ethers (POMDME), reaches back

as far as 1904, when Descudé reported the synthesis of OME2 from sodium methylate and

dichloro dimethyl ether.44 In the 1920, Staudinger published works on oligomeric and

polymeric OME with the aim of understanding the nature of polymeric materials.45 The first

patent on synthesis of oligomeric OME was assigned to DuPont in 1948.46 Of commercial

relevance was the development of production processes for polymeric homologues of OME.

These are named polyoxymethylene (POM) or polyacetals and were commercialized by

DuPont in the 1960s.47 POM is a thermoplastic polymer with high mechanical, thermal and

chemical resistance and is mainly applied in the automotive and electronics sectors. In the

following years, further studies on properties of short chain OMEs were published.48-50 For

example, Boyd reported physical properties, such as vapour pressure, density, melting points

of the homologues OME1-4 in 1961.51

More recently, the main research interest in oligomeric OME is based on the finding that

oxygenates have favourable combustion properties as mentioned in chapter 2.1.3.8 In a

potential chemical production network based on methanol as a platform chemical as proposed

by Olah,52 OME could furthermore be attractive components.

Since 1997, when OME was found to have favourable properties for combustion in CI

engines,27 several companies have filed patents for the synthesis of OME as diesel additives,

such as BP Amoco Corporation53, 54 and BASF.55, 56 OME are also of particular interest to

Chinese companies that have filed numerous patents in the past years. Relying on its large

reserves in coal, synthesis gas based chemical industry has strongly developed in China, and

an oversupply of methanol has been observed.57, 58 In this context, OME could help to balance

changes in supply of diesel fuel. In 2015, the first and currently the only large-scale OME

plant was installed by Shandong Yuhuang Chemical Company in China.

Page 21: Synthesis of oxymethylene ethers

Introduction 8

In terms of application development, it is interesting to note that the viability of OME for use

in conventional cars was demonstrated in 2017. The automotive manufacturing company

Continental successfully employed OME blends in test vehicles.59 Moreover, engineers from

TU Darmstadt retrofitted a standard car to run on neat OME fuel.60

The major challenge remains the economically feasible production of OME. The predicted

price of OME fuel depends strongly on variables, such as choice and price of raw materials

and the considered synthesis route.61, 62 With current technology and prices of raw materials,

OME are estimated to be 2-4 times more expensive than conventional diesel fuel.5, 60 In the

current synthesis process, 60% of the production cost are attributed to raw materials, while

energy demands are considered to account for 20% of the final OME price.61 To date, OME

are produced with an energy use as high as 10 MJ/kg OME in the Chinese production

facility.63, 64 This results from the energy-intensive production and separation of the

intermediate trioxane and OME1.

It is evident, that process development will be a key factor to realise wide-spread use of OME

as diesel additive or fuel. In this context, the in-depth study of the catalytic transformations

involved in OME formation is an interesting and relevant field of research.

2.2.3 Synthesis

The general formula CH3O(CH2O)nCH3 shows that OME are built from a chain of

formaldehyde-derived repeating units (CH2O)n and methyl end-groups. As pure monomeric

formaldehyde is not commercially available as a reagent due to its tendency to polymerise, it

needs to be introduced into the reaction via either paraformaldehyde (PF), trioxane (TRI) or in

solution. PF represents uncapped formaldehyde polymer chains, while TRI is a trimer of

formaldehyde. The monomeric formaldehyde is released via acid-catalysed or thermal

decomposition. Aqueous or methanolic formaldehyde solutions contain mainly short-chain

poly(oxymethylene) glycols or hemiformals, respectively, which are in equilibrium with

monomeric FA and readily react.65 There are also several options for supplying the end-group.

Methanol (MeOH), dimethoxymethane (OME1) and dimethyl ether (DME) are options for

capping agents.

An exemplary reaction scheme for the reactant combination methanol and formaldehyde is

depicted in Figure 2.2. It includes hemiformal formation, chain-growth and acetal formation

steps. The reaction steps are reversible and the presence of water can shift the equilibrium

Page 22: Synthesis of oxymethylene ethers

9 Introduction

towards the hemiactals.66 The presence of water in the reaction has been described to reduce

reaction rate and OME selectivity in various liquid-phase batch experiments.67-69

Figure 2.2: Schematic representation of OME formation from methanol and formaldehyde.

Analogous to other chain-growth reactions, a product distribution of different homologues is

obtained. It follows a Schulz-Flory distribution, which has been developed to describe

molecular weight distributions of linear condensation polymers.67, 70-72 In literature, no

conclusive indications are available on whether chain-growth in OME formation occurs via

oligomeric hemiformals (initiation, growth and termination mechanism) or via insertion of

formaldehyde into OMEn (sequential mechanism). It is suggested that the mechanism depends

on reactants and reaction conditions. Commonly, the probability of chain growth is observed

to be low in acid catalysed synthesis which leads to product distributions mainly comprised of

short-chain OME.64

The different reactant combinations can be classified according to the reaction

characteristics.64 Reactant combinations of methanol with any formaldehyde source are called

aqueous route as water is a stoichiometric side-product in the acetal formation step. In

aqueous synthesis mode, poly(oxymethylene) glycols and hemiacetals are potential by-

products. Also the combination of OME1 and PF is classified as aqueous, as low amounts of

water are released upon depolymerisation of PF. In contrast, no water is formed when OME1

is reacted with TRI. This route is hence called anhydrous route. Current industrial processes

are based on the latter route. Also the recently described reaction of DME with TRI is an

anhydrous transformation.73

As OMEs synthesis is in general a catalytic transformation, the various proposed synthetic

procedures can also be classified according to the catalysts used.

Page 23: Synthesis of oxymethylene ethers

Introduction 10

Homogeneous catalysis involving mineral acids55 and acidic ionic liquids74-76 can be applied

for the above described reactant combinations. Additionally, recent reports of OME1

formation from CO2, H2 and MeOH over organometallic catalysts in solution77, 78 and of OME

formation from OME1 and absorbed gaseous FA catalysed by trimethyloxonium salts79 can be

assigned to this group.

A broad range of synthetic procedures involving heterogeneous catalysis have been reported.

As this work is focussed on solid catalysts for OME formation, this class will be described in

more detail below. Also methods for direct synthesis of OME1 via either one-step oxidation of

MeOH80-90 or via DME oxidation91-94 are based on heterogeneous catalysis. Mainly transition

metal oxide containing solid catalysts and materials based on supported heteropoly acids have

been reported in this context.

As mentioned above, a wide range of synthetic procedures involving solid acid catalysts has

been published in the scientific literature. Table 2.2 gives an overview of the reported reaction

parameters, such as reactant combination and ratio, catalyst type and loading, reaction

temperature and time. Additionally, the maximum conversion and OME selectivity reported

are specified. Table 2.2 is limited to batch-mode reactions, which constitute the majority of

reports in the scientific literature. The various solid acid catalysts listed include ion-exchange

resins, zeolites, alumosilicates, sulphated metal oxides, graphene oxide as well as supported

ionic liquids and stabilized heteropoly acids.

When comparing the catalyst performance data summarized in Table 2.2, it is necessary to

consider the following aspects: In case of OME1 as a capping agent, OME1 is commonly

regarded solely as a reactant, while in case of the capping agent MeOH, OME1 is one of the

products and is hence also considered in the calculation of selectivity. The maximum

attainable selectivity of homologues longer than OME1 will therefore differ. It is also not

consistently specified whether selectivity is given in mol% or wt% within the reports.

Additionally, the selectivity towards different OME homologues is lumped into varying

groups. The feasibility of a direct comparison of the reported data is hence limited.

In addition to liquid-phase batch-reactions, flow reactions were mainly described in patent

literature. In the context of this work, processes based on methanol and formaldehyde as

reactants are of interest. For example, Zhang et al. described a continuous-flow synthesis of

Page 24: Synthesis of oxymethylene ethers

11 Introduction

OME over alumina supported zirconia catalysts and over ion-exchange resins95, 96 and Burger

et al. patented a continuous process with Amberlyst 46 ion-exchange resin.97

Table 2.2: Overview of batch-mode OME synthesis procedures based on heterogeneous catalysis. The ratio of capping agentand formaldehyde source is given as mass ratio (wt.) or molar ratio (mol.). Maximum conversion (Xmax) is indicated withrespect to FA source if not otherwise indicated. Maximum selectivity (Smax) is lumped for a group of OME homologues asspecified in the reference. The reactant combination MeOH/FA corresponds to a methanolic formaldehyde solution.

cap.agent

(1)

FAsource

(2)

ratio(1):(2)

catalysttype

cat.loading/ wt%

T/ °C

t/ min

Xmax

/ %Smax

/ %ref

OME1 TRI 2.5 : 1(wt.)

ion-exchangeresins

7.5 90 30 89 OME3-8: 64.2 98

OME1 TRI 1:1(mol.)

ion-exchangeresins

ca. 10 100 480 - OME2-4 :22.3 99

OME1 TRI 2:1(wt.)

ion-exchangeresins

ca. 5 50/90

> 60 93 OME2-8: 51OME3-8: 27

66

OME1 TRI 1:1(mol.)

sulfonic acidfunctionalisedsilica

2 100 60 95.6 OME2-8: 61.8 100

OME1 TRI 2:1(mol.)

supported ionicliquids

4 105 60 92 OME3-8 : 52 101

OME1 TRI 3:1(mol.)

zeolites 0.3 25 60 94.5 OME3-5: 21.3 102

OME1 TRI 2 (wt.) zeolites 5 120 45 85.3 OME2-8: 88.5 103

OME1 TRI 3.3:1(wt.)

zeolites 0.5 70 250 ca. 95 OME3-5: ca. 43 104

OME1 TRI 1:1(mol.)

zeolites 5 120 1200 92.4 OME2-8: 90.6OME3-8: 60.3

105

OME1 TRI 2.5 : 1(wt.)

alumosilicates 7.5 105 120 92.6 OME2: 45.1OME3-8: 53.5

106

OME1 TRI 1:1(mol.)

alumosilicates 2 100 60 92.7 OME2-8: 56 107

OME1 TRI 1:1(mol.)

sulphated TiO2 1 80 60 89.5 OME3-8: 54.8 108

OME1 PF 3:1(wt.)

ion-exchangeresins

5 80 120 84.7 OME3-5: 36.6 58

OME1 PF 2:1(wt.)

ion-exchangeresins

- 90 360 - OME2-5: 41.2OME3-5: 22.1

109

OME1 PF 1.25:1(mol.)

ion-exchangeresins

5 90 120 77.5 OME3-6: 41.5 110

OME1 PF 2:1(mol.)

ion-exchangeresins + LiBrpromotor

ca. 10 100 1440 - OME2-4: 33.0 99

OME1 PF 3:1(mol.)

sulphated TiO2 3 80 50 ca. 85 OME3-5: ca. 22 111

Page 25: Synthesis of oxymethylene ethers

Introduction 12

cap.agent

(1)

FAsource

(2)

ratio(1):(2)

catalysttype

cat.loading/ wt%

T/ °C

t/ min

Xmax

/ %Smax

/ %ref

MeOH TRI 2:1(mol.)

zeolites 5 120 600 100 OME3-8: 29.4 112

MeOH TRI 2:1(wt.)

Pd-modifiedH-ZSM-5

1 130 - 95.2MeOH

OME2-5: 62.9 113

MeOH TRI 2:1(mol.)

PVP-stabilisedheteropoly acids

2 140 240 95.4 OME2-5: 54.9 114

MeOH TRI 2:1(mol.)

graphene oxide 5 120 600 92.8 OME2-8:30.9 115

MeOH TRI 2:1(mol.)

H-MCM-22zeolite

5 120 600 39.8 OME2-8: 65.1OME3-8 : 39.4

105

MeOH TRI 2:1(wt.)

sulphatedFe2O3-SiO2

1.5 130 120 81.9 OME3-8: 23.3 116

MeOH FA various ion-exchangeresins

ca. 5 >100 39 OME1 :14.4 w%a

OME2: 10.3 wt%a

OME3: 6.7 wt%a

OME4 :4.0 wt%a

117

MeOH FA 0.67:1(wt.)

ion-exchangeresins

0.5 80 >180 44 OME1 :14.8 w%a

OME2: 10.4 wt%a

OME3: 5.9 wt%a

OME4 :4.0 wt%a

118

MeOH FA 0.67(wt.)

ion-exchangeresins andzeolites

1 40-120

> 20 - OME1 :15.8 w%a

OME2: 9.9 wt%a

OME3: 5.6 wt%a

OME4 :3.1 wt%a

68

MeOH FA 0.5:1 ion-exchangeresins

10 40-80 180 - - 119

DME TRI 4:1(mol.)

H-BEA zeolite 0.4 80 960 13.9DME

not indicated 73

a overall mass fraction at chemical equilibrium

2.3 Solid acid catalysis

Solid acids are widely applied in catalytic processes in the petrochemical industry and

chemical synthesis. Historically, solid acid catalysts have replaced liquid mineral acids in

various processes, owing to advantages, for example, in process engineering, handling,

separation and regeneration.120

Generally, solid acids catalysts are characterised by the presence of Brønsted (proton

donating) and/or Lewis acidic (electron pair accepting) groups. The acid groups can be located

at the surface and/or inside the solid. In case of supported acids, the acid sites are in the active

phase that is distributed on the internal and/or external surface of a support.

Page 26: Synthesis of oxymethylene ethers

13 Introduction

In contrast to acids in solution, acid sites are not mobile in the reaction medium and their

properties depend on the local environment within the solid structure. Also, interactions of

molecules with the acid sites are influenced by factors such as surface adsorption, diffusion to

or accessibility of acid sites. It is then not surprising that the experimental methods to measure

acid properties of acids in solution and solid acids differ profoundly as described in

chapter 2.3.

Common examples of solid acids are alumina, amorphous and crystalline alumosilicates

(zeolites), functionalised metal oxides, ion-exchange resins, activated carbons, supported

mineral acids, and heteropolyacids.120 In this work, mainly zeolites and supported phosphoric

acid catalysts have been studied. These classes of catalysts will hence be described in more

detail in the following.

2.3.1 Zeolites121, 122

The term “zeolites” traditionally includes naturally occurring or synthetic crystalline

alumosilicates with structure-inherent porosity, which consist of a three-dimensional

framework of corner-connected tetrahedral primary building units (SiO4 and AlO4). The

excess charge related to the Al-containing tetrahedra is compensated by cations.123 A general

formula such as

⁄ ∙ ∙ ∙ (1)

is used in order to indicate the composition of the material. Similar materials including other

primary units, for example Al-P based aluminophosphates (ALPO), Si-Al-P based

silicoaluminophosphates (SAPO) based frameworks, and zeolitic materials with incorporated

germanium, titanium and other metals have been reported.123

The compensating cations in the zeolite structures are exchangeable. This is, for example,

exploited in the application of zeolites as ion-exchangers in detergents, which is also the

largest field of application.124 For use in acid catalysis, the cation is typically a proton. The

pores in the inorganic framework have a regular periodic arrangement and are of molecular

size. They can form one- to three-dimensional networks depending on the structure type. The

characteristics of zeolite porosity are of importance in applications as adsorbents as well as in

catalysis.

Page 27: Synthesis of oxymethylene ethers

Introduction 14

The above described zeolite properties account for their use in a large range of applications,

for example as detergents, in agriculture, as adsorbents and pigments and in catalysis. In

catalysis, oil refining, the (petro)chemical industry and environmental catalysis are the most

important fields of application. Until today, a large variety of frameworks has been reported.

Out of the 248 listed framework types in 2018,125 only few are, however, of commercial

interest.

The acidic properties of zeolites are the basis for their use as catalysts in industrially relevant

acid-base reactions, such as isomerization, cracking, (de)alkylation, Friedel-Crafts reactions,

addition and elimination reactions as well as oligomerization reactions. Zeolites can contain

Brønsted and Lewis acid sites with a variable acid strength. Despite the very weak acid

strength, also silanol groups may play a role in zeolite reactivity. The concentration and

strength of the three types of acid sites depend on multiple factors, including structure type,

composition and treatment of the material.

The Brønsted acid site is the most common and well-studied type of acid sites in zeolites. It

occurs when a three-valent Al+ replaces a four-valent Si+ in the tetrahedral TO4-building unit

and when the charge is compensated by a proton. In this case, it is suggested that the proton is

present in form of a bridging hydroxyl group (see Figure 2.3 a). The bridging OH-groups are

commonly strong Brønsted acid sites. However, the strength of bridging hydroxyl groups is

influenced by the local environment. For example, acid sites at different positions within the

framework may differ in the degree of interaction with neighbouring atoms. Additionally, the

Al content has an impact on the acid strength. The more Al is located in close proximity of a

Brønsted acid site within the framework, the lower is the acid strength. This effect is most

prominent at high Al loading.

Figure 2.3: Schematic representations of a) a Brønsted acidic bridging hydroxyl group, b) a proposed form of a Lewis acidicextra-framework Al species and c) an isolated (terminal) silanol group in zeolites.

Page 28: Synthesis of oxymethylene ethers

15 Introduction

Lewis acidity may arise from various sources. Firstly, charge-compensating metal ions

commonly represent weak Lewis acid sites. Secondly, so-called extra-framework aluminium

(EFAl) can form when Al is removed from the framework, commonly showing Lewis acidity.

The removal can occur during steaming, acid leaching and thermal or hydrothermal treatment.

However, the term EFAl groups various Al-containing species and the elucidation of their

chemical nature and their distribution is challenging. An AlO+ complex is exemplarily

depicted in Figure 2.3 b. Alternatively, AlxOyn+ complexes, uncharged Al2O3 particles and

AlO(OH) have been proposed to form.126 It is evident that the Lewis acidity of EFAl will

depend on which species are present. It is interesting to note that Lewis acidic EFAl species

interacting with Brønsted acid have been reported to increase the acid strength of the Brønsted

acid site.127

Silanol groups (Si-OH) are very weakly acidic and are not typically considered in zeolite acid

catalysis. However, in few cases such as the Beckmann rearrangement, silanol groups act as

active sites. Siliceous zeolitic materials such as Silicalite-1 are active in the industrially

relevant rearrangement of cyclohexanone oxime to caprolactam.128 Generally, silanol groups

can occur on the external surface of zeolite crystals, or in the bulk on framework defects. Also,

the Si-OH groups can be either isolated (Figure 2.3 c), vicinal, or in clusters. The latter is

argued to occur at a silicon vacancy and has been proposed to be the active site in Beckmann

rearrangement.128, 129

From the above description of acid sites in zeolites, it is evident that the acidity of zeolitic

materials is, except for Si-OH groups, determined by the Al content. According to the

Löwenstein-rule, there is a limit of Al content at a Si/Al ratio of 1:1 as AlO4 tetrahedra are not

stable when directly connected to each other.130 A short overview of methods available for

studying the nature, density, and strength of acid sites in zeolites and other solids is given in

chapter 2.3.3.

In the following, selected aspects of zeolite structural properties that are important in the

context of acid catalysis will be discussed. Information on further structure-related aspects

such as secondary and tertiary building units, nomenclature, classification of structure types

and others can be found in reference 122.

In zeolite catalysis, the highly ordered pore system is the most important structural feature that

governs various catalyst properties. The pore size and volume influences, for example, the

Page 29: Synthesis of oxymethylene ethers

Introduction 16

surface area, sorption of guest molecules, the accessibility of reactants to active (acid) sites in

the inside of zeolite crystals, the stabilisation of transition states and intermediates, often

termed shape-selectivity, and deactivation tendency via pore blocking.

The porosity of idealised zeolite crystals is defined by the zeolite framework type. The latter

describes the connectivity of tetrahedral TO4 units and is classified by the International Zeolite

Association using three letter codes.125, 131 The framework type specifies the size of pore

openings and channels as well as the dimensionality of the pore network. Pore opening size is

typically indicated as number of TO4 units connected to a ring at the pore opening. It can be

classified according to ring size: 8-ring (small), 10-ring (medium) and 12-ring (large). For

elucidation and study of framework type, powder X-ray diffraction is a method of choice.

The textural properties of a real zeolite material are not only determined by framework type,

but chemical composition, defect density and presence of extra-framework species (cations,

water, organic compounds, adsorbed molecules, EFAl) have major impact on porosity and

surface area. It is therefore necessary to study the textural properties for individual zeolite

samples. For the investigation of zeolite textural properties, such as porosity and surface area,

gas physisorption is a well-established technique.

2.3.2 Supported liquid phase catalysts

A supported liquid phase (SLP) catalyst may be defined as a catalytically active material

dispersed in/on an inert (porous) solid that is dissolved or molten at reaction temperature.132,

133 Different systems have been described that involve molten salts as well as organic or

aqueous phases. When a supported salt has a melting point below 100 °C, the term supported

ionic liquid phase (SILP) is commonly used.

The first applied supported liquid phase catalyst was the silica-supported V2O5 alkali-

pyrosulphate catalyst for the oxidation of sulphur dioxide in sulphuric acid production. The

presence of a molten phase in the catalyst was, however, only elucidated three decades after its

introduction in 1914.133 Another prominent example is the Deacon catalyst, which is based on

supported CuCl2 with promotors, and which is used for the oxidation of hydrogen chloride and

for the oxidative chlorination of unsaturated hydrocarbons.132

While the two latter catalysts systems aim at oxidation and oxychlorination reactions, the so-

called solid phosphoric acid (SPA) catalyst is an industrially relevant SLP catalyst for acid-

catalysed reactions. As a supported phosphoric acid catalyst has been applied in this work, it

Page 30: Synthesis of oxymethylene ethers

17 Introduction

will be described in more detail below. In more recent reports, the immobilization of

catalytically active metal organic complexes via dissolution in a supported ionic liquid (SILP)

or aqueous phase (SAP) has been described and applied to various reactions such as

hydroformylation.133

It is important to note that while the synthesis of SLP catalysts is often simple, the state of the

catalyst under reaction conditions may be complex. For example, the catalytically active

component may either be the molten salt (chlorides in Deacon catalyst) or the dispersed liquid

itself (H3PO4 in SPA catalyst) or it may be dissolved in a molten salt (V2O5 in alkali

pyrosulphate in the sulphuric acid catalyst and SILP catalysts) or in an aqueous phase (SAP

catalysts).

Phosphoric acid supported on a silica matrix, often naturally occurring Kieselgur, is a purely

Brønsted acidic catalyst that has been used as a solid acid catalyst since the 1930s.134 It is

commonly referred to as solid phosphoric acid (SPA). The main applications of SPA in

industrial processes are the oligomerisation of low molecular weight alkanes to form high

octane gasoline and the synthesis of ethylbenzene and cumene by alkylation of benzene with

ethylene or propylene, respectively. Furthermore, SPA have found application in hydration

reactions such propene to propanol transformation.133

Prepared by simple impregnation of ortho-phosphoric acid (H3PO4) on silica and subsequent

calcination, the final catalyst comprises many components. Firstly, condensed phosphoric acid

species such as pyro- and polyphosphoric acid can be present.133 Furthermore, various silicon

phosphate phases form upon calcination when silica is used as support.134 It is challenging to

determine the distribution of ortho-, pyro- and polyphosphoric acids under reaction conditions.

As the performance of the catalyst is strongly dependent on the concentration of water in the

reactant feed, is has, however, been argued that free ortho-phosphoric acid is the main active

species in oligomerisation and alkylation reactions.135

The main advantages of SPA are the low cost and high selectivity for Brønsted acid catalysed

reactions. Disadvantages include limited lifetime and the fact that the catalyst cannot be

regenerated.133 Although phosphoric acid has a very low vapour pressure,136 a possible loss of

active phase via leaching, and equipment corrosion must also be considered. Therefore, SPA

has been substituted by zeolite catalysts in some processes such as cumene synthesis. It is,

however, still in wide-spread industrial use.133, 137

Page 31: Synthesis of oxymethylene ethers

Introduction 18

As described above, phosphoric acid is commonly supported on silica in industrial

applications. In scientific reports, also other supports, such as carbon (H3PO4/C)138 and

alumina (H3PO4/Al2O3)139 are reported. Another class of material, which is based on a

preparation route similar to H3PO4/C, are phosphorylated carbons. These materials are,

however, subjected to high-temperature treatment and subsequent washing, resulting in the

formation of C-P bonds and removal of free phosphoric acid.140 Phosphorylated carbons are

therefore not considered as SLP catalysts.

2.3.3 Characterisation methods126, 141

In general, four main aspects can be considered when studying solid acid catalysts and the

related reactivity. Firstly, the nature of acid sites including Brønsted and Lewis types is

important. Likewise, the strength of acid sites has an impact on catalyst activity. In case of

Brønsted acids, the strength relates to the readiness of proton transfer and can be expressed as

intrinsic or relative strength. Thirdly, the acid site concentration or density can be measured.

Finally, the impact of accessibility of active sites may not negligible for porous solid catalysts,

especially when bulky reactants or products are involved.

As supported liquid phase catalysts as well as conventional solid catalysts are used in this

work, it is of interest to briefly discuss the difference in acid characterisation for liquid and

solid acids.141, 142 This comparison also demonstrates, why the two main catalyst classes

studied in this work cannot be characterised with the same methods.

In proton donor-type (Brønsted) acids in aqueous solutions, acid sites are mobile. Their

strength can be related to the intrinsic acid strength of the H3O+ species whereas the acid

concentration is related to the acid dissociation constant Ka of the acidic molecule in solution.

The latter is often expressed in the logarithmic form, the pKa. Various methods for

determination of pKa in dilute aqueous solutions are known with some of the most established

being potentiometry, conductometry, and electrophoresis. In non-aqueous medium, for strong

acids or for concentrated solutions, a method relying on spectrophotometric measurement of

the dissociation degree of indicators according to works by Hammett is preferred. The latter

method is one of the few that has been applied to both liquid and solid acids. The concept of

acid strength is typically not applied to Lewis acids in solution. The Lewis acidic chemical

species are rather described according to their reactivity.

Page 32: Synthesis of oxymethylene ethers

19 Introduction

While liquid or dissolved acids typically feature either Brønsted or Lewis acidity, both acid

types can be present within a solid material. They may even interact, resulting in changed acid

strength. This renders a careful analysis of the nature of the acid site present in the catalyst

necessary. In Table 2.3, an overview of characterisation methods for solid acids is given. As

zeolites are of importance to this work and as the majority of literature on acid site

characterisation is focussed on zeolites, the table includes notes on the applicability of

methods to zeolite materials. An additional method not included in Table 2.3 is model

catalytic reactions.141

Table 2.3: Overview over methods for characterization of solid acids with a focus on zeolite analysis.126, 141

nature of acid sites(Brønsted / Lewis)

number/densityof acids sites

strength (distribution)of acid sites

computationalmethods Brønsted acids ---

computed deprotonationenergy, zeolites: topologicaldensity of Al tetrahedra

Hammettindicators &butylaminetitration

Brønsted acids

butylamine titration(constraint:accessibility of bulkacid sites)

acid strength via colourchange of indicators uponprotonation(constraint: only strength ofstrongest acid site is measured,properties of indicators may beinfluenced by surfaceadsorption143)

NH3-TPD non-selective via sum of desorbedNH3

via temperature of desorption(constraint: only approximatemeasure, preferably only usedfor ranking of similar samples)

1H MASNMR144

different Brønsted acidsites via chemical shiftof -OH groups, differentLewis acid sites viachemical shift of adsorbedprobe molecules

quantification viacomparison with astandard

Brønsted acids via chemicalshifts induced by probemolecules, not viable forLewis acid sites

MAS NMRwith othernuclei144

27Al MAS NMR:coordination environmentvia Al chemical shift,assignment to framework(mainly Brønsted) andextra-framework (mainlyLewis) sites in zeolites

---

Brønsted acids via chemicalshifts induced by probemolecules, (e.g. 31P MASNMR of trimethyl-phosphine),not viable for Lewis acid sites

Page 33: Synthesis of oxymethylene ethers

Introduction 20

nature of acid sites(Brønsted / Lewis)

number/densityof acids sites

strength (distribution)of acid sites

FTIRwithout probemolecules

different Brønsted acidsites via characteristicwavenumber of -OH groupstretching vibration

--- ---

FTIR withbasic probemolecule,e.g. pyridine

different Brønsted andLewis acid sites

Only in transmissionmode, via Lambert-Beer law (constraint:availability ofextinctioncoefficient, acid siteaccessibility)

band shift and temperature ofdesorption of probe molecule

Microcalori-metricmeasurements

non-selectivevia amount of basicprobe necessary toneutralise acid sites

via measurement ofdifferential heat of adsorptionof probe molecule (constraint:influenced by van der Waalsinteractions145)

2.4 Methanol dehydrogenation

In a production process of OME via methanol and formaldehyde as studied in this work, the

reactants can be supplied via partial methanol dehydrogenation. In order to combine methanol

dehydrogenation with OME synthesis, it is necessary to consider the characteristics and

available catalyst systems for this reaction.

2.4.1 Oxidative vs. non-oxidative route

For the catalytic transformation of methanol to formaldehyde, there are two approaches:

oxidative and non-oxidative dehydrogenation. The former is employed industrially and is the

main production pathway of formaldehyde. It is carried out in large scale owing to the

importance of formaldehyde as an intermediate for production of resins, pesticides,

disinfectants, dyes, preservatives, explosives, paper and others.65

It is an exothermic reaction (ΔHR = -159 kJ/mol)65 that follows the overall reaction presented

in equation (2-1).

+ 0.5 ⎯⎯⎯ + (2-1)

Two different process routes can be distinguished, mainly differing in catalysts used and in

reaction conditions. On the one hand, a methanol-rich feed stream can be converted over silver

Page 34: Synthesis of oxymethylene ethers

21 Introduction

catalysts at 600 - 720 °C. Water vapour is fed in order to maintain the catalyst activity. For

this catalyst system, a two-step reaction via methanol dehydrogenation and subsequent

oxidation of hydrogen to water is reported. Hence, this route is also referred to as

oxydehydrogenation. On the other hand, a catalyst based on molybdenum, vanadium and iron

oxides can be used for methanol conversion in excess oxygen at 300 - 450 °C. In this case, the

mechanism relies on a single oxidation step.146

The non-oxidative dehydrogenation of methanol (see equation (2-2)) is currently not used in

large-scale industrial processes. However, this reaction has been studied with the intent to

circumvent formaldehyde – water separation for applications in which anhydrous

formaldehyde is required, for example production of polyoxymethylene thermoplastics.147

Another advantage is that hydrogen is a more valuable by-product than water.

⎯⎯⎯ + (2-2)

As it is an endothermic reaction (ΔHR = 84 kJ/mol)65 it needs to be carried out at elevated

temperatures. Thermodynamically, the formation of formaldehyde is favoured at temperature

above 475 °C.147 In this temperature range, the decomposition of formaldehyde to carbon

monoxide and hydrogen is, however, a prominent side reaction, which occurs even without

catalyst.147 Therefore, short residence times in the reactor are crucial.

2.4.2 Catalysts

In this work, the non-oxidative approach to methanol transformation was applied. Therefore,

the following short overview over the most relevant catalyst systems is limited to non-

oxidative methanol dehydrogenation. It is based on the comprehensive review by Usachev et

al.148

A range of catalyst systems for non-oxidative methanol dehydrogenation is described in

academic publications.147-149 Typical highly active (de)hydrogenation catalysts containing, for

example, platinum or iron150 are not among the catalysts suitable for MeOH dehydrogenation

due to an excess activity in formaldehyde decomposition to CO and H2. Rather, catalytic

materials with moderate (de)hydrogenation activity such as silver, copper and zinc dominate.

Additionally, catalysts containing sodium ions are reported to be active.148

The majority of studies on non-oxidative methanol dehydrogenation are based on zinc in the

form of molten metal, zinc alloys, zinc oxide and Zn ion-exchanged catalysts. A large range of

Page 35: Synthesis of oxymethylene ethers

Introduction 22

supports and co-catalysts including oxides of silicon, lanthanum, iron, indium, cerium,

tellurium, chromium and sodium have been reported. Silica supported ZnO catalysts and Zn

exchanged zeolite show the highest formaldehyde yields. It should be noted that elemental

zinc or Zn2+ reduced under reaction conditions can be leached out of the reactor in the form of

Zn vapour.148 A range of studies describe copper catalysts for non-oxidative methanol

dehydrogenation. While pristine copper deactivates quickly, increased stability was claimed

for various catalytic systems based on copper alloys or supported copper (II) oxide. The best

performance was reported for CuZnS, CuZnSe and CuO-Cu3(PO4)2/SiO2 catalysts.148 In the

class of silver based catalysts, pristine Ag and alloys containing Cu, Zn and/or Te are

mentioned. In contrast to commercial methanol oxidation catalysts based on molybdenum and

iron oxides, which are not active in non-oxidative methanol dehydrogenation, silver is active

in both dehydrogenation pathways. As mentioned above, the oxydehydrogenation of methanol

over silver proceeds via initial dehydrogenation of methanol and subsequent oxidation of

hydrogen. It is interesting to note that under non-oxidative conditions, silver catalysts need to

be pre-treated and regularly reactivated with oxygen, suggesting that also in this case, an

oxydehydrogenation mechanism may be involved. High initial formaldehyde yields and fast

deactivation is commonly described.148

The presence of transition metals is not mandatory to yield an active catalyst for the formation

of anhydrous formaldehyde. Sodium exchanged zeolites were patented for methanol

dehydrogenation to formaldehyde. Also, reports on sodium metal catalysts are available.

Interestingly, simple salts containing sodium ions achieve comparable performance to the

above described transition metal based catalyst classes. Among the studied simple sodium

salts, including carbonate, tetraborate, phosphate, molybdate, sulphate and aluminate anions,

the highest formaldehyde yields are obtained over sodium carbonate.148

Page 36: Synthesis of oxymethylene ethers

23 Motivation and research objectives

3 Motivation and research objectives

Synthetic fuels based on oxymethylene ethers have the potential to play a key role in

decreasing road transport emissions and in building a fuel infrastructure independent of fossil

resources. As OME synthesis remains the main challenge for its commercialisation as an

additive or fuel for compression-ignition engines, it is of interest to study alternative

production pathways.

In current industrial synthesis routes,118 the major drawback is the large number of process

steps, including five main steps: (1) formation of MeOH, (2) production of aqueous

formaldehyde, (3) synthesis of the intermediate OME1, (4) synthesis of intermediate trioxane

or paraformaldehyde and (5) OMEn formation. The latter three process steps are based on

liquid-phase processes. Another disadvantage is the need for a highly energy demanding

separation of the intermediates.

In perspective of a future large-scale production of OME for supplying large enough quantities

to use OME as a fuel, gas-phase technology has the advantage of easier scalability, improved

process integration and easy implementation of continuous processes. Potentially, OME could

be produced in a complete continuous gas-phase process in only three process steps starting

from CO or CO2 and H2 (syngas) including (1) synthesis of methanol, (2) subsequent partial

dehydrogenation to formaldehyde to yield the FA/MeOH reactant mixture, and (3) formation

of OMEn (see Figure 3.1).

Figure 3.1: Schematic representation of the three steps involved in the targeted gas-phase process.

The transfer of the last process step, namely the formation of OME from methanol and

formaldehyde, from liquid to gas-phase is one of the two main objectives of this work. The

second focus is the in-depth study of solid acid catalysts for gas-phase OME synthesis.

For these aims, it was targeted to build a versatile catalytic test set-up and to perform

preliminary studies on reaction conditions and on solid acid catalyst classes. The gained

knowledge was envisaged to be applied in the in-depth study of structure-activity relations of

Page 37: Synthesis of oxymethylene ethers

Motivation and research objectives 24

active catalysts. Finally, the gas-phase synthesis of OME from methanol without separation of

intermediates was aimed to be implemented as a proof of concept for the viability of the above

described complete gas-phase OME synthesis.

Page 38: Synthesis of oxymethylene ethers

25 Description of test set-up

4 Description of test set-up

The construction of a flow set-up was a prerequisite to perform the targeted catalytic studies

described above. Design, assembly, implementation, and calibration of the set-up were an

integral part of the project and will therefore be described in this chapter. The set-up was

designed to allow for flexibility to explore reaction parameters in a large range, e.g.

temperature, pressure, reactant concentration, and residence time. Additionally, an emphasis

was laid on ensuring safe handling of the pressurized equipment and of hazardous and

flammable reagents involved.

The set-up built is a continuous flow set-up with a fixed-bed reactor. It allows performing

reactions at up to 400 °C and up to 25 bars.

4.1 Concept

The underlying concept is depicted in Figure 4.1. The reactants are supplied in liquid form and

transported to a heated evaporator unit by a pump. Inside the evaporator, the evaporated

components are mixed with a flow of inert gas. The latter is supplied from a gas bottle and is

adjusted using a pressure regulator and mass flow controller. The flow of diluted reactants can

be passed through the reactor containing the catalyst to be tested. Alternatively, the reactor can

be bypassed in order to analyse the reactant stream directly. The pressure inside the system is

controlled by a back pressure regulator placed at the downstream end of the pressurized zone.

At the outlet of the set-up, an online gas chromatograph for qualitative and quantitative

analysis of the gas stream is located.

Figure 4.1: Conceptual schematic representation of the test set-up.

Page 39: Synthesis of oxymethylene ethers

Description of test set-up 26

4.2 Technical implementation

A more detailed representation of the set-up is given in Figure 4.2 and Table 4.1. It includes

information on technical instrumentation, piping and gas supply.

Figure 4.2: Detailed schematic representation of the test set-up. The main pathway of carrier gas and reactants is marked inbold lines. The heated zone is marked in orange. Detailed information is listed in Table 4.1.

Table 4.1: Description of set-up components as displayed in Figure 4.2.

1 supply of carrier gas 15 line for reactor purge andpressurization 29 pressure transducer with elevated

temperature resistance2 supply of helium 16 line for reactant solution outgassing 30 additional reactor (R1) and circular oven

3 supply of synthetic air 17 venting system 31 reactor (R2) and circular oven

4 supply of nitrogen 18 proportional relief valve 32 reactor bypass line

5 supply of hydrogen 19 line towards vent 33 line for reactor pressure release and purge

6 pressure reducers 20 pressure transducer 34 adjustable back pressure regulator

7 particle filter 21 reservoir of reactants (FA/MeOH) 35 line towards gas chromatograph

8 mass flow controllers 22 HPLC pump 36 gas chromatograph bypass, towards vent

9 magnetic valves 23 ball valve 37 condensation trap

10 check valve 24 reservoir for reactant waste frompurge 38 online gas chromatograph

11 bypass line for fastsystem pressurization 25 capillary towards evaporator 39 gas-washing bottle filled with water

12 ball valve 26 start of elevated temperature zone 40 gas-washing bottle with Na2SO3 solution

13 needle valve 27 evaporator heated by circular furnace

14 bypass line for fastreactor pressurization 28 high pressure ball valves

Page 40: Synthesis of oxymethylene ethers

27 Description of test set-up

For the operation of any catalytic test set-up, the control of process parameters is essential.

This includes supply of gases and liquid reactants, regulation of pressure and gas flows as well

as heat control. The implementation is briefly described in the following sections.

4.2.1 Gas and pressure control

In the OME synthesis set-up, a premixed gas containing 5% methane in nitrogen is used as a

carrier gas (Figure 4.2: 1). Both components are inert in the studied reaction. While nitrogen is

employed for dilution, methane is added as an internal standard for GC analysis. It is supplied

in a high pressure gas cylinder (up to 200 bars) and is calibrated by the manufacturer. The

carrier gas is supplied to the reactor either via the main pathway highlighted in Figure 4.2 or

via several by-pass lines (Figure 4.2: 11,14,15). Additional gases such as helium, synthetic air,

nitrogen and helium (Figure 4.2: 2-5) are connected to the gas chromatograph. Helium is

additionally used for outgassing of the reactant solution (Figure 4.2: 16).

The gas flows into the system are regulated by mass flow controllers (MFC), which are

operated remotely via LabView software (Figure 4.2: 8). All MFCs have a flow range of up to

1 L/min (STP). The MFCs require a pressure gradient of approximately 10 bars between the

upstream and downstream outlet. The respective upstream pressure is adjusted by a pressure

reducer (Figure 4.2: 6). The downstream pressure in the system is set by the adjustable back

pressure regulator at the outlet of the set-up (Figure 4.2: 33). When the MFCs were calibrated

using a volumetric primary flow calibrator device, the same pressure gradient was applied. All

MFCs are protected by magnetic ball valves that separate the MFCs from the main system

when no reaction is running (Figure 4.2: 9). In the main feed line for the reactor, the MFC is

additionally equipped with a particle filter and a check valve (Figure 4.2: 7,10). The latter

allows flow only in the downstream direction and prevents pressure changes in the system to

affect the MFC.

The pressure in the system can be read from two pressure transducers, which are placed at the

inlet of the system (Figure 4.2: 20) and at the inlet of the reactor (Figure 4.2: 29), respectively.

4.2.2 Evaporator unit

The reactant solution contains approximately 60% FA, 38% MeOH and 2% H2O and is

obtained by dissolving paraformaldehyde in methanol via refluxing (see chapter 11.6.1). It is

supplied in a reservoir and outgassed with a flow of helium prior to use (Figure 4.2: 16,21).

Page 41: Synthesis of oxymethylene ethers

Description of test set-up 28

The liquid is transported to the evaporator unit using a HPLC piston pump (Figure 4.2: 22).

The pump achieves flows as low as 2 μL/min. The reactants methanol and formaldehyde are

compatible with the provided pump. Other components such as oxymethylene ethers,

however, can only be pumped for short periods due to a limited chemical resistance of the

sealing materials.

While all other tubing of the set-up has dimensions of 6 mm outer diameter, the liquid feed

enters the evaporator via a capillary with 1.6 mm (1/16 inch) outer diameter (Figure 4.2: 25).

The thin capillary facilitates a constant flow of liquid into the evaporator even at very low

flow rates. It ends inside the heated zone of the evaporator, where the liquid evaporates in a

stream of carrier gas (Figure 4.2: 27). The evaporator volume is filled with inert granular

silicon carbide in order to improve evaporation and mixing with the inert gas by providing

high surface area.

4.2.3 Heating

All tubes downstream of the evaporator are heated to 170 °C in order to avoid polymerization

of formaldehyde and condensation of methanol and reaction products. The main components

of the set-up, such as evaporator, reactor and back pressure regulator, are equipped with

circular ovens. The connecting tubes are tightly wrapped with heating tape. The heated

components are furthermore wrapped in glass wool mats and tape as well as aluminium foil

for insulation. The ovens and heating tapes are regulated by individual in-house built

temperature controller units. The directing temperature input is supplied by thermocouples

placed closely to the heated components.

4.2.4 Reactor

The reactor is a flow reactor composed of a steel tube with an outer diameter of 10 mm and an

inner diameter of 5.8 mm (Figure 4.2:31 and Figure 4.3). It is equipped with an internal grid

that holds the catalyst bed in place. A thermocouple placed inside the catalyst bed permits to

measure the local temperature. Plugs of quartz wool placed at the inlet and outlet of the reactor

prevent contamination of connecting tubes with catalyst particles.

The catalyst bed is comprised of catalyst pellets (300-400 μm size range) diluted with inert

silicon carbide (46 grit) in a mass ratio of catalyst to SiC of 1:6. For catalyst activation prior to

reaction, the reactor can be purged and pressurized with inert gas (Figure 4.2:15 and 32).

Page 42: Synthesis of oxymethylene ethers

29 Description of test set-up

Figure 4.3: Schematic representation of the reactor R2 for OME synthesis. View is turned from vertical to horizontal

orientation.

4.3 Product analysis

In addition to process control, the reliable analysis of feed as well as product composition is a

key prerequisite for catalytic testing.

In the constructed set-up, the analysis is based on gas-chromatography (GC). The GC device is

placed at the outlet of the set-up. It is equipped with a six-way valve for online sampling, a

polyethylene glycol based polar capillary column and two detectors connected in series, a

flame ionization detector (FID), and a thermal conductivity detector (TDC). For its operation,

various gases are supplied via the general laboratory gas system. While helium is used as the

carrier gas, synthetic air and hydrogen are required to sustain the hydrogen flame of the FID.

Nitrogen gas is connected to the pneumatic actuators of the 6-way sampling valve.

Prior to running a reaction, the feed concentration can be monitored via GC by setting the

respective valves in the set-up (Figure 4.2:28,32) to the bypass position. To start the reaction

and for product analysis, the valves are switched towards the reactor.

For quantification via GC, the retention times and response factors of the analytes need to be

determined prior to analysis. A range of compounds was available as pure substances, such as

methanol, methyl formate, trioxane, formic acid OME1, OME3 and OME4. This allowed

simple determination of retention times and response factors via injection of liquid aliquots of

pure components or prepared mixtures containing 1-butanol as an internal standard. In case of

DME, a calibration gas mixture was employed. For formaldehyde, the retention time was

determined via headspace sampling from paraformaldehyde heated at 100 °C.

Page 43: Synthesis of oxymethylene ethers

Description of test set-up 30

Other OMEn and the hemiacetal of methanol and formaldehyde (hemiformal, HF) were

however not available as pure substances. Hence, their retention times were identified from a

reference liquid-phase OME synthesis. Trioxane and OME1 were reacted in an autoclave over

an acidic ion-exchange resin (see chapter 11.5). The product mixture was then analysed via

GC coupled with mass spectrometry (GC-MS). Response factors of OME2 and OME>4 were

extrapolated from the data obtained with pure OME1, OME3 and OME4 according to a method

specified in literature.117

In the test set-up, methane is used as an internal standard. Hence, the reference of response

factors was converted from 1-butanol to methane. For this purpose, the response factors of

MeOH, OME1 and OME3 with respect to methane were determined by evaporation of a

calibrated liquid feed. As the ratios of the response factors in the gas-phase were in agreement

with ratios determined from liquid injections, the other response factors were recalculated

accordingly without further experimental determination.

Similarly, the response factor of formaldehyde was determined by evaporation and analysis of

a calibrated liquid feed of a methanolic formaldehyde solution (see chapter 11.6.1). The

composition of the latter was identified via iodometry and Karl-Fischer titration (see chapter

11.6.2 and 11.6.3). It is important to note that the quantitative evaluation of hemiformal was

not feasible. On the basis of the marginal amount detected, the hemiformal content was

neglected, resulting in a minor systematic undervaluation of reactant concentrations. This is,

however, not expected to have a significant impact on catalytic data.

4.4 Safety features

In the test set-up, a range of chemicals with hazardous properties is employed, such as

formaldehyde (toxic, corrosive, irritant and is presumed to be carcinogenic) and methanol

(flammable and toxic). This necessitates rigorous safety precautions. OME are rated to be non-

toxic.18

In the context of process control, several safety features are included in the set-up design.

Firstly, a proportional relief valve is installed in order to release pressure from the system if a

threshold value is reached (ca. 40 bars, Figure 4.2:18). The heating units for the main

components, such as evaporator, reactor and back pressure regulator, are connected to in-

house built security shut-off units with an adjustable temperature threshold. In addition to the

general venting of the laboratory compartment where the set-up is installed, a vent is installed

Page 44: Synthesis of oxymethylene ethers

31 Description of test set-up

above the set-up (Figure 4.2:17). It is connected to the central exhaust system of the

laboratory. In order to achieve a bottom-to-top venting, transparent polymeric curtains cover

the front and sides of the set-up. The curtains can be moved aside for operation and

maintenance of the set-up. Before entering the general venting system, the set-up exhaust

gases are passed through two gas wash bottles filled with water and aqueous sodium sulphite

solution in order to absorb formaldehyde and other organic components (Figure 4.2: 39,40).

The time spent inside the laboratory compartment is minimised through computer based

remote access to key process parameters. Gas flow rates are adjusted via a LabView software

panel. Also temperature and pressure can be monitored remotely.

For the prevention of hazards related to formaldehyde, additional personal protective

equipment is provided. A handheld formaldehyde meter based on electrochemical sensing is

allocated at the entry of the laboratory compartment for regular testing of formaldehyde

concentration. An acoustic alarm signal turns on at >0.3 ppm of formaldehyde in the air. The

use of a non-stationary detector allows localising a formaldehyde source, for example a leak, if

necessary. A full-face gas filtration mask is also provided at the entry of the compartment. It is

fitted with a filter targeting a range of organic substances including formaldehyde and

methanol.

4.5 Extensions for combined process

For the implementation of OME synthesis from methanol, a second reactor suited for

methanol dehydrogenation was introduced to the set-up.

Figure 4.4: Schematic representation of the reactor R1 for methanol dehydrogenation. Gas flow is directed from left to right

side.

The reactor is a flow reactor composed of a tube with an outer diameter of 10 mm. It is fitted

with a quartz tube with an inner diameter of 5 mm (Figure 4.2:30 and Figure 4.4). The quartz

Page 45: Synthesis of oxymethylene ethers

Description of test set-up 32

liner is included in order to avoid potential blind reactions induced from components of the

steel reactor walls (1.4571 type steel). The catalyst bed is held in place by quartz wool plugs.

A thermocouple is placed closely to the catalyst bed. The catalyst bed is comprised of catalyst

pellets diluted with inert silicon carbide. Similar to reactor R2, the reactor can be purged and

pressurized with inert gas prior to reaction (Figure 4.2:15 and 32). It holds pressure up to

20 bars. The reactor is heated in a circular oven unit that has a temperature rating of 650 °C.

The unit was designed and built in the fine mechanics workshop of the institute.

In contrast to OME synthesis from formaldehyde and methanol, methane is a potential by-

product and is hence not a suitable internal standard for GC quantification. The experiments

were carried out with pure N2 as carrier gas. Relative response factors of reactants and

products were calculated and used for data evaluation.

Page 46: Synthesis of oxymethylene ethers

33 Screening of reaction conditions

5 Screening of reaction conditions

Prior to systematic investigations of catalyst performance and structure-activity relations

described in chapters 6, 7 and 8, a range of suitable reaction conditions was determined using

an exemplary industrial catalyst. An H-form mordenite zeolite with a SiO2-to-Al2O3 ratio of

40, denoted as H-MOR-40, was used for all experiments described in this section if not

specified otherwise. In the following, the effect of various reaction parameters on the

conversion and selectivity will be presented and discussed. The studied parameters include

reaction temperature, total pressure and partial pressure of reactants, reactant ratio, water

content, pellet size and catalyst activation protocol. In this context, the reproducibility of the

test reaction was also assessed.

5.1 Temperature

The temperature dependence of catalyst activity was studied in the range of 130 – 270 °C

(Figure 5.1, left). Firstly, a test run with increasing temperature steps from 170 to 270 °C was

conducted. Upon return to the initial temperature level at 170 °C, the conversion and

selectivity had only changed to a minor extent. This indicates that (de)activation of the catalyst

did not occur on the time scale of the experiment and that the trends in catalyst performance

can be attributed to a temperature effect. Due to the superior performance at 170 °C, a data

point at 130 °C was collected in a supplementary test run.

Figure 5.1: Conversion and selectivity of H-MOR-40 (left) and Silicalite-1 (right) as a function of temperature. Reactionconditions: 0.5 g catalyst pellets in 3 g SiC 10 bar, 100 ml/min inert gas flow, 14 μl/min FA/MeOH mixture.

Page 47: Synthesis of oxymethylene ethers

Screening of reaction conditions 34

Under the studied conditions, varying amounts of the products OME1 and OME2 as well as by-

products methyl formate (MeFO) and dimethyl ether (DME) were observed. A pronounced

influence of temperature on both conversion and selectivity is evident. OME selectivity

decreased from 72% at 130 °C to 3% at 220 °C. No OME is detected at 270 °C using

H-MOR-40.

When discussing the observed decrease in selectivity towards OME at increasing temperature,

two aspects need to be taken into account: Firstly, the intrinsic thermodynamics of OME

synthesis and secondly, the competing side reactions. The latter will be described below.

Computational studies of the thermodynamics of OME formation demonstrate that the

equilibrium reaction between any set of reactants and OME is strongly shifted towards the

reactant side with increasing temperature.151 The latter effect can be examined independently

of by-product formation when Silicalite-1 catalyst is employed. It is a very weakly acidic

siliceous zeolite material that was identified to be active in OME synthesis during

investigations described in chapter 7. Silanol groups, which are the active species in

Silicalite-1, do not favour DME or MeFO formation. Over Silicalite-1, a steady decrease in

conversion upon increase in reaction temperature occurs. OME selectivity is not significantly

influenced. This emphasises that in case of H-MOR-40, the catalyst activity towards by-

products determines the product distribution. In this context, it is of interest to discuss by-

product formation and assess the reversibility of the competing reactions.

While MeFO is the major by-product in the lower measured temperature range, DME

formation is favoured with increasing reaction temperature. DME is formed via acid catalysed

dehydration of methanol (see equation (5-1)). DME formation is not an equilibrium reaction is

hence showing an increase in reaction rate upon increasing temperature. This is in line with

reports of methanol dehydration over zeolites or alumina having an increasing reaction rate in

the range of up to approx. 300 °C.152, 153

2 ⎯⎯ + (5-1)

For the formation of methyl formate, two reaction pathways need to be considered. Firstly, a

Cannizzaro-type reaction (equation (5-2)) and a subsequent esterification (equation (5-3)) can

occur.154

2 + ⎯⎯ + (5-2)

+ ⎯ (5-3)

Page 48: Synthesis of oxymethylene ethers

35 Screening of reaction conditions

Under the studied reaction conditions, formic acid was not detected irrespective of catalysts

and product distribution. Previously, the retention time of formic acid in the gas

chromatograph was determined. A response factor was not established for formic acid, for

which reason the formation of minor amount of formic acid cannot be ruled out. However, the

Cannizzaro reaction typically requires strong bases such as sodium hydroxide.154

It therefore appears more likely that the formation of methyl formate proceeds via the second

possible pathway, namely the Tishchenko reaction. It is a one-step disproportionation-

dimerization reaction (equation (5-4)).

2 ⎯ (5-4)

The Tishchenko transformation of aldehydes is a well-established reaction in organic

synthesis. It is commonly performed using aluminium alkoxide catalysts in solution and

proceeds via a catalyst-coordinated hemiacetal transition state and a hydride transfer reaction

step.155, 156 While methyl formate is mentioned to be a potential by-product in OME liquid

phase synthesis, the formation mechanism is seldom discussed. However, some authors

attributed it to the Tishchenko reaction.66 There are only few reports of gas-phase Tishchenko

reaction with formaldehyde, in which mainly binary oxides have been employed. These

studies emphasise the importance of the presence of both acidic and basic sites in the

catalyst.157 Interestingly, in the base catalysed gas-phase aldol condensation of the

formaldehyde homologue propanal over HY zeolites, the Tishchenko reaction was likewise

reported to occur as a side reaction.158 In summary, it may be assumed that the formation of

methyl formate also proceeds via Tishchenko reaction in the case of zeolite H-MOR-40.

Exemplarily, the reversibility of methyl formate formation over H-MOR-40 was studied (see

Figure 5.2). At 130 °C, no conversion occurred upon exposure of a flow of MeFO to the

zeolite catalyst indicating the irreversibility of the reaction. It could furthermore be confirmed

that the formation of OME1 in the gas-phase is reversible. A feed of OME1 and water was

transformed into a mixture of reactants FA and MeOH and minor amounts of side products

MeFO and DME as well as OME2. The results highlight that the activity of a catalyst towards

the irreversible formation of side products has a major impact on the product distribution as

indicated above.

Page 49: Synthesis of oxymethylene ethers

Screening of reaction conditions 36

Figure 5.2: Reversibility test for methyl formate (upper left) and OME1 (upper right) formation and respective reactionscheme (bottom). The dashed line indicates the switch from bypass to reactor. Reaction conditions: 130 °C, 10 bar, 0.5 gH-MOR-40, 100 mL/min inert gas flow, 28 μL/min feed of methyl formate or OME1 mixed with ca. 5% H2O. H2O is detectedbut not quantified.

In terms of the preliminary screening of suitable reaction conditions, it was necessary to make

a compromise between the favourable impact of a low reaction temperature and the required

temperature to keep all components in the gas-phase. The latter was estimated by taking into

account the saturation vapour pressure of reactants and potential products159, 160 as well as the

maximal expected partial pressures of the components in the test set-up. The temperature

130 °C was chosen and adopted for all further test runs.

5.2 Pressure

According to Le Chatelier’s principle, an equilibrium reaction that features a change in the

number of moles of its components will be affected by a change in reaction pressure. This

feature is valid for the initial hemiformal formation and for the chain-growth steps in OME

formation (see chapter 2.2.2 , Figure 2.2). When studying the pressure dependence of a

catalytic reaction, both the total pressure and the reactant partial pressure can be varied. To

test the sensitivity of the reaction to total pressure the reaction was performed at 10 and 20

bars while the reactant partial pressure was held constant. In both cases, no change in catalyst

Page 50: Synthesis of oxymethylene ethers

37 Screening of reaction conditions

performance was observed. Tests at atmospheric pressure could not be evaluated due to

unsatisfactory carbon balance.

In contrast to total pressure, the partial pressure of both reactants has an impact on the final

product distribution (see Figure 5.3). Holding the reactant ratio constant, the reactant partial

pressure was increased in two steps. It resulted in an increase of OME1-3 selectivity. This

observation can be applied in further reaction condition optimisation. It should, however, be

kept in mind that in a gas-phase process, the partial pressure is restricted by the saturation

pressure of the components at which they will condense.

Figure 5.3: Conversion and selectivity of H-MOR-40 as a function of total (FA+MeOH) reactant partial pressure (left) and ofreactant ratio (right). Reaction conditions: 0.5 g catalyst pellets in 3 g SiC, 10 bar, 100 ml/min inert gas flow. Left: FA/MeOHfeed of 14, 30 or 45 μL/min with reactant ratio of 3:2. Right: 1.52 bar reactant partial pressure, 30 μL/min flow of reactantsolution with varying composition.

5.3 Reactant ratio

The product distribution at three different weight-based reactant ratios (3:2, 1:1 and 2:3

FA/MeOH) was obtained by successive addition of methanol to the reactant mixture. With

decreasing formaldehyde partial pressure, less OME chain growth occurs. At the same time,

the reactant conversion increases. For synthesis of OME>1 homologues, a high FA/MeOH is

favourable. In this work, the reactant solution was prepared by dissolution of

paraformaldehyde in methanol. The solubility of formaldehyde in methanol limits the

FA/MeOH ratio to 3:2.

Page 51: Synthesis of oxymethylene ethers

Screening of reaction conditions 38

5.4 Water content

As described in chapter 2.2.2, the presence of water shifts the reaction equilibrium and hence

is expected to have an adverse effect on OME yield. When OME is synthesised according to

the so-called aqueous route from methanol and formaldehyde as starting materials, the

stoichiometric formation of water as a by-product cannot be circumvented.

For the catalytic tests described in the following chapters, it was of interest to assess the

impact of the water content of the reactant feed. In the test set-up, formaldehyde and methanol

are introduced via evaporation of a solution containing ca. 60 wt% FA, 37-38 wt% MeOH and

2-3 wt% H2O. The latter is prepared by dissolution of paraformaldehyde in methanol and the

water content is related to the end-groups of the paraformaldehyde polymers.

The impact of water addition was tested at 3.3, 17.7 and 22.8 wt% H2O. There was neither an

effect on conversion nor on product distribution (see Appendix, Figure 12.1). The results do

not allow extrapolating to lower water contents. A stronger correlation of initial catalyst

performance to water content may arise in this regime. Nevertheless, the reactant solution

obtained from paraformaldehyde dissolution with 2-3 wt% water content was found to be

suitable for the purpose of this work.

5.5 Pellet Size

It is important to verify that the catalyst performance is independent of pellet size for the

chosen reaction conditions in order to demonstrate the absence of mass transfer limitations.161

The latter include macroscopic effects such as channel formation inside the catalyst bed and

effects on the microscopic scale as for example diffusion limitations into the catalyst grain.

For this purpose, the catalyst activity of H-MOR-40 in two different pellet sieve fractions

(100 – 200 μm and 300 – 400 μm) was compared. There was only a marginal difference in

catalyst performance which may be attributed to the limits of reproducibility of the test set-up.

For further testing, the pellet size range of 300 – 400 μm was chosen for all powdered

catalysts.

5.6 Catalyst activation protocol

To study the impact of the in-situ catalyst activation procedure performed inside the test set-

up, two different protocols were compared. Interestingly, a thermal treatment at 300 °C for 2h

on the one hand and the heating of the catalyst at the reaction temperature 130 °C for 1h on the

Page 52: Synthesis of oxymethylene ethers

39 Screening of reaction conditions

other hand yielded the same catalyst performance. For the sake of time efficient catalyst

screening and testing, the milder treatment at 130 °C was adapted.

5.7 Reproducibility

The reproducibility of catalytic tests was evaluated by 5-fold repetition of a test run with

following reaction conditions: 10 bar, 130 °C, 0.5 g H-MOR-40, 100 mL/min inert gas flow,

14 μL/min FA/MeOH solution feed (see Figure 12.2). The deviation of the arithmetic mean of

obtained conversion and selectivity results was below 3%.

Page 53: Synthesis of oxymethylene ethers

Preliminary catalyst screening 40

6 Preliminary catalyst screening

In order to assess the potential of different classes of conventional solid acid catalysts for gas-

phase OME formation, representatives of proton form zeolites (H-MOR-40, H-FAU-5),

functionalised metal oxides (sulphated and tungstated zirconia, SO4-ZrO2 and WO3-ZrO2),

supported heteropoly acids (silicotungstic acid supported on alumina, HPA-Al2O3) and ion-

exchange resins (Amberlyst 36) were studied in the first stages of this work. In the group of

ion-exchange resins, Amberlyst 36 was chosen owing to its improved thermal stability up to

150 °C as specified by the supplier. Further information about the catalysts is supplied in

chapters 11.1.3 and 11.2.1.

Figure 6.1: Initial selectivity and conversion of solid acid catalysts determined in the interval of 40 - 70 min. reaction time.Residual activity measured without catalyst. Reaction conditions: 10 bar, 130 °C, 0.5 g of catalyst, 100 mL/min inert gas flow,14 μL/min FA/MeOH solution feed. WHSVFA: 1.1 g(FA)*g(cat) 1*h-1.

Comparing the product distributions (see Figure 6.1), it may be noted that OMEn yield is

decreasing with increasing n – a typical feature of chain-growth reactions – and that products

(OMEn) and by-products (MeFO, DME) are formed in varying ratios. While the latter may

appear trivial, this is not commonly the case in the analogous liquid phase batch reactions

where equilibrium compositions are reached irrespective of the catalyst used (see Table 2.2.

Catalysts however vary in the time for reaching this equilibrium.68 The reversibility of the side

Page 54: Synthesis of oxymethylene ethers

41 Preliminary catalyst screening

reactions was tested as described in chapter 5.1. It was confirmed that MeFO is formed

irreversibly while OME1 formation is reversible. It can therefore be supposed that the tested

catalysts differ in their activity towards the irreversible formation of MeFO and DME.

H-MOR-40, HPA/Al2O3 and Amberlyst 36 show the highest activity. However, the two latter

catalysts have a major drawback. Leaching of active species was observed, causing

contamination of the downstream components of the set-up. When the reactor was bypassed,

active species that had accumulated in the back pressure regulator caused transformation of

the feed stream (see Figure 6.1, residual activity).

Silicotungstic acid is water-soluble and it can be assumed that the heteropoly acid was leached

from the alumina support. In case of Amberlyst 36, which is a sulfonic acid functionalised ion-

exchange resin, the downstream contamination is presumably caused by decomposition of

sulfonic acid groups. It appears that the reaction temperature of 130 °C is too close to the

maximum operating temperature of 150 °C specified by the supplier. The decomposition of

sulfonic acid was confirmed by preparation of sulfonic acid functionalised silica (SBA-15-

SO3H, see chapter 11.2.2). Accordingly, residual activity occurred after testing the material.

Potentially, sulphur trioxide can be released. It must be noted that the residuals in the set-up

show good activity and an excellent selectivity towards OME. Even though a series of tests

were performed, the residuals could not be accumulated and extracted nor could their nature

be fully elucidated. The investigations involved extensive maintenance of the set-up. It was

therefore refrained from further testing of supported heteropoly acids and ion-exchange resins.

Instead, a systematic investigation of zeolites for gas-phase OME synthesis was targeted,

which is described in chapter 7.

Page 55: Synthesis of oxymethylene ethers

OME synthesis over zeolite catalysts 42

7 OME synthesis over zeolite catalysts1

7.1 Catalyst screening

On the basis of the promising performance of the preselected zeolite H-MOR-40 in the

preliminary screening described in the previous chapter and also owing to the well-known

variability of zeolitic materials with regards to structure and acidity, a systematic study of

OME synthesis over zeolites was envisaged. This study included a screening of commercial

and synthesised zeolitic materials and a detailed study of the deactivation and regeneration

behaviour using the two best performing benchmark catalysts.

In the zeolite screening, materials with four different framework types and varying

SiO2/Al2O3-ratios were chosen. The selected zeolites were used in protonated form. In analogy

to the before mentioned catalyst H-MOR-40, the catalysts are named to indicate the proton

form (prefix H-), the framework type (three letter code) and SiO2/Al2O3-ratio (suffix). Three

samples of zeolite Y (H-FAU-12/129/340), two of zeolite Beta (H-BEA-35/150), three of

ZSM-5 (H-MFI-27/90/∞, the latter will be referred to as Silicalite-1 throughout the study) and

two of Mordenite (H-MOR-14/40) were tested under the conditions derived in chapter 5. In

Table 7.1, an overview of structural parameters of the included framework types is presented.

Further information about the employed materials and activation procedures is presented in

chapter 11.1.3, 11.2.3 and 11.3.1.

Table 7.1: Characteristics of selected zeolite framework types.125

frameworktype threeletter code

exemplary trivial namesof related materials

maximum diameter of asphere that can diffuse

along / Å

largest poreopening ring

sizea

channeldimension-

nalityb

MFI ZSM-5 a: 4.70, b: 4.46, c: 4.46 10 3DMOR Mordenite a: 1.57, b: 2.95, c: 6.45 12 2D (1D)c

FAU Faujasite, Y-zeolite a: 7.35, b: 7.35, c: 7.35 12 3DBEA Beta zeolited a: 5.95, b: 5.95, c: 5.95 12 3D

a number of TO4 units connected to a ring at the pore openingb including channels with pore opening ring sizes larger than 6c transport along one axis is structurally hindered, therefore effectively 1D122

d partially distorted materials, parameters given for idealised framework

1 The major part of this chapter was published in Gas-phase synthesis of oxymethylene ethers over Si-rich zeolites, A. Grünert,P. Losch, C. Ochoa-Hernández, W. Schmidt and F. Schüth, Green Chem., 2018, 20, 4719-4728. Copyright 2018, RoyalSociety of Chemistry. In the following text, numerous quotations and reproductions of figures and tables from the publicationare included, but will not be marked individually.

Page 56: Synthesis of oxymethylene ethers

43 OME synthesis over zeolite catalysts

The results of the zeolite catalyst screening are presented in Figure 7.1. Similar to the general

catalyst screening discussed in chapter 6, varying product distributions and conversion levels

are observed. It is remarkable that there is a trend in the zeolite catalyst performance. For all

four structural classes of zeolites, an increase in selectivity to OME and a decrease in

conversion are observed with increasing SiO2/Al2O3 ratio (SAR). At increased SAR, the

amount of Al and hence the amount of Brønsted-acidic protons is decreased.

An influence of SAR on catalyst performance has been reported for H-ZSM-5,103, 113

H-MCM-22 112 and Al-SBA-15 107 when OME was synthesized in batch-mode from OME1 or

MeOH and trioxane. In these cases however, a maximum in OME yield was generally

observed with conversion drastically decreasing at higher SAR due to insufficient release of

formaldehyde by acid catalysed decomposition of trioxane. In our study, no constraints by

trioxane decomposition exist and we have confirmed the trend over a wide range of SAR and

for a broad range of samples.

Figure 7.1: Catalyst screening: Initial selectivity and conversion of zeolitic catalysts determined in the interval of 40 - 70 min.reaction time. Reaction conditions: 10 bar, 130 °C, 0.5 g of catalyst, 100 mL/min inert gas flow, 14 μL/min FA/MeOHsolution feed. WHSVFA: 1.1 g(FA)*g(cat) 1*h-1.

7.2 Correlation between acid site properties and catalyst performance

In order to study the suggested correlation of catalyst performance and the properties of its

acid sites, temperature programmed desorption of ammonia (NH3-TPD) was carried out for all

materials under investigation (Figure 12.3 to Figure 12.6). It has been discussed for liquid-

Page 57: Synthesis of oxymethylene ethers

OME synthesis over zeolite catalysts 44

phase OME synthesis over different zeolites that moderately strong acid sites are best suited

for OME syntheses. 103, 105 This cannot be confirmed in case of gas-phase synthesis from

MeOH and FA. In Figure 7.2, conversion and OME yield are presented as a function of the

total amount of ammonia desorbed in the NH3-TPD measurement, the latter being related to

the total amount of acid sites in the zeolite. In addition to the zeolitic catalysts, an amorphous

siliceous reference material (fumed silica Aerosil 200) is included. In accordance with the

above described correlation between SAR and conversion, an increased amount of acidic sites

correlates to higher conversion for zeolitic samples (Figure 7.2, left). It is also evident that

amorphous silica is not active. The NH3-TPD curves show relatively broad and/or flat signals.

Therefore, a deconvolution into low- and high-temperature contributions was not performed.

Figure 7.2: Left: Conversion and right: OME yield as a function of total amount of ammonia desorbed. Filled symbols denotezeolitic catalysts; the hollow symbol indicates the siliceous reference sample. Suffixes at H-MOR-40 samples indicatecalcination temperature as discussed below.

When the OME yield is related to the total amount of acid sites (Figure 7.2, right), zeolites

with a low acid site concentration seem to perform best. The highest OME yields of 42% and

43% are achieved by H-MOR-40_350 and Silicalite-1, respectively. The suffix refers to the

calcination temperature. Silicalite-1 is a siliceous zeolitic material that is characterized by the

presence of only very weakly acidic silanol groups (not detected in NH3-TPD). The described

high activity of Silicalite-1 is unexpected. Conventionally, classical Brønsted acid sites created

by Si-O-Al bridges, or Lewis acid sites are thought to be responsible for the formation of

OME. Since these are absent in Silicalite-1, another active site than hitherto thought must be

responsible for the high activity of this catalyst. The amorphous silica used as a reference has

no catalytic activity.

Page 58: Synthesis of oxymethylene ethers

45 OME synthesis over zeolite catalysts

Figure 7.3: FTIR spectra of the pyridine stretching vibration region for a) MFI-27 and b) MOR-14 based materials at differentdesorption temperatures: i) 150 °C, ii) 250 °C and iii) 350 °C. Above: H-form zeolites, below: Na-form zeolites.

In order to substantiate the finding that Brønsted acid sites are not necessary to catalyse the

formation of OME in the gas-phase, two Al-containing zeolite catalysts were prepared in their

sodium form. For this purpose, the respective zeolite in ammonia form was ion-exchanged

with sodium nitrate and subsequently calcined (see chapter 11.3.2). The completion of the

sodium exchange was verified via FTIR spectroscopy using pyridine as a probe molecule.

Indeed, the stretching vibrations of pyridine interacting with Brønsted acid groups at 1635 and

1545 cm-1 vanished upon ion-exchange (see Figure 7.3 a). The bands present in the spectra of

the exchanged samples are typical for pyridine adsorbed on Na-zeolites.162 In the catalytic

tests, both materials showed a significantly improved performance resulting in an increase of

OME yield of as high as 38% in case of the Na-MFI-27 zeolite (see Figure 7.4).

As mentioned above, the product ratio is influenced by the activity of the catalysts towards the

irreversible formation of by-products. The observations that the formation of by-products is

suppressed by Na-exchange in Al-containing zeolites and that Silicalite-1 shows high OME

selectivity suggest that by-product formation may be related to the presence of strong

Brønsted acid sites. Weakly acidic sites such as silanol-groups in framework defects or at pore

mouths seem to provide sufficient acidity for the formation of OME. Besides, the presence of

weakly Lewis acidic sodium ions in the framework does also not have an adverse impact on

the OME selectivity.

1650 1575 1500 1425 1350

1595

1623

(iii)(ii)(i)

Na-MFI-27

Abs

orba

nce

(a. u

.)

Wavenumber (cm-1)

0.5

H-MFI-27

1635

1545

1455

1442

(i)(ii)(iii)

1650 1575 1500 1425 1350

159115

9916

20

(iii)(ii)(i)

Na-MOR-14

Abs

orba

nce

(a. u

.)Wavenumber (cm-1)

0.5

H-MOR-14

1633

1544

1454

1443

(i)

(ii)(iii)

a) b)

Page 59: Synthesis of oxymethylene ethers

OME synthesis over zeolite catalysts 46

Figure 7.4: Initial selectivity and conversion of H- and Na-form zeolites determined in the interval of 40 - 70 min. reactiontime. Reaction conditions: 10 bar, 130 °C, 0.5 g of catalyst, 100 mL/min inert gas flow, 14 μL/min FA/MeOH solution feed,WHSVFA: 1.1 g(FA)*g(cat) 1*h-1.

When discussing the acidic properties of zeolites, it is also important to consider the influence

of extra-framework aluminium (EFAl), which is typically characterized by Lewis acidity. The

influence of the presence of EFAl on the formation of OME from MeOH and FA was

exemplarily studied using H-MOR-40. In a series of H-MOR-40 material calcined at varying

calcination temperatures, emergence of EFAl was induced at temperatures above 350 °C. This

was evidenced by 27Al-MAS-NMR (Figure 7.5).

The pristine H-MOR-40 shows mainly tetrahedrally-coordinated Al (signal centred at 57 ppm)

and only little Al in octahedral environment (signal centred at 0 ppm). Upon temperature

treatment, an increase in the asymmetric broadening of the signal related to tetrahedral

framework indicates the formation of Al in distorted tetrahedral environment and/or penta-

coordinated Al. Furthermore, a rise in the peak at 0 ppm and the additional emergence of a

broad peak centred at -5 ppm, assigned to various Al species in octahedral environment,

indicate the removal of Al from the mordenite framework and the formation of EFAl

species.163 When the sample calcined at 550 °C was washed with oxalic acid, which is known

to dissolve primarily EFAl species, the content of Lewis acid sites could be reduced.

Page 60: Synthesis of oxymethylene ethers

47 OME synthesis over zeolite catalysts

Figure 7.5: Al-MAS-NMR spectra of H-MOR-40: pristine, calcined at 350 °C, 450 °C, 550 °C and H-MOR-40 calcined at550 °C with subsequent acid wash using oxalic acid (OA). Stacked spectra (a) and overlapping spectra (b and c). Signals arenormalized to the signal at 57 ppm.

The change in the ratio of Brønsted- to Lewis-acidity as a result of EFAl formation was

confirmed by Pyridine-FTIR measurements (Table 7.2 and Figure 12.7). The concentration of

Brønsted and Lewis acid sites was quantified by analysis of the pyridine stretching vibrations

corresponding to interaction with Brønsted acid sites at 1545 cm-1 and Lewis acid sites at 1455

cm-1. Additionally, information about the strength of the acid sites could be gained by

adsorption of pyridine at different temperatures. As expected, a decrease in the ratio of

Brønsted to Lewis acidity with increasing calcination temperature is observed. Notably, a

constant Si/Al ratio was determined for all FTIR measurements, which is an indication for

good comparability of FTIR results over the series of studied catalysts. The Si/Al ratio

calculated from FTIR data is commonly lower than when calculated from aluminium content

as not all of the Al atoms are probed. For example, Al may be related to a very weak site that

does not retain pyridine at the adsorption temperature of 150 °C. Alternatively, it may be

buried inside an extra-framework cluster and is therefore not accessible to pyridine.

Page 61: Synthesis of oxymethylene ethers

OME synthesis over zeolite catalysts 48

Table 7.2: Acid sites concentration of selected samples after pyridine adsorption at 150 °C (CB: concentration of Brønstedacid sites; CL: concentration of Lewis acid sites). Suffixes denote the calcination temperature. Extinction coefficients obtainedfrom reference 164.

pyridine desorptiontemperature (°C)

CB (mmol/g) CL (mmol/g) B/L Si/Ala

H-MOR-40_350ºC150 0.34 0.09 3.8

30250 0.29 0.08 3.6

350 0.18 0.06 3.0

H-MOR-40_450ºC150 0.30 0.11 2.7

30250 0.25 0.09 2.8

350 0.14 0.07 2.0

H-MOR-40_550ºC150 0.28 0.15 1.9

28250 0.26 0.13 2.0

350 0.18 0.10 1.8a Calculated at 150 ºC

Figure 7.6: Initial conversion and selectivity of H-MOR-40 as a function of calcination temperature.

The H-MOR-40 samples were also characterized by NH3-TPD (Figure 12.8). An increased

amount of ammonia desorbed in the high-temperature range of 500 - 700 °C is evident in the

curves of samples calcined at 450 and 550 °C as compared to 350 °C suggesting that upon

temperature treatment stronger acid sites were created. These could be due to strongly acidic

EFAl sites and/or Brønsted acid sites with increased acidity due to interaction with EFAl.127

Page 62: Synthesis of oxymethylene ethers

49 OME synthesis over zeolite catalysts

The effect of EFAl formation and the resulting rise in the Brønsted- to Lewis-acid ratio is

reflected in the catalytic performance of H-MOR-40. A significant drop in OME selectivity

was observed when calcination temperatures above 350 °C were employed (Figure 7.6).

In summary, one may conclude that three different acidic species in zeolites – namely

Brønsted acid sites, Lewis acidic EFAl species as well as silanol groups – all affect the

catalytic performance of the zeolites. This complex interplay of acidic sites along with the

competition of OME formation with irreversible side-reactions render it difficult to exactly

determine specific contributions of each type of acid site. However, the general conclusion can

be drawn that catalysts characterized by a low number of Brønsted and/or EFAl acid sites

show better performance and that weakly acidic species such as silanol groups are sufficient to

catalyse the OME formation.

7.3 Influence of particle size and external surface area

Figure 7.7: Left: Mean particle size and standard deviation for a selection of commercial zeolites. Right: Conversion as afunction of external surface area.

In order to rule out an effect of crystallite size, scanning electron microscopy (SEM)

micrographs and in some cases additional transmission electron microscopy (TEM) images of

catalysts tested in the screening were collected. Exemplarily, an SEM image and histogram of

H-FAU-12 are presented in Figure 12.9 and Figure 12.10. However, the size distribution of

the commercial materials was too broad to allow reasonable correlation of particle size and

catalyst activity (see Figure 7.7, left).

Page 63: Synthesis of oxymethylene ethers

OME synthesis over zeolite catalysts 50

Additionally, external surface areas were determined from nitrogen sorption isotherms via

t-plot analysis and plotted against conversion (Figure 12.11) and OME yield (Figure 7.7,

right). However, no clear correlation with the external surface area was evident.

7.4 Adaptation of reaction conditions

In order to further investigate the catalyst performance for OME gas-phase formation, the two

best performing zeolites from the screening were tested under adapted conditions. For both

materials, an improved OME yield was achieved when the weight hourly space velocity

(WHSV) was increased from 1.1 to 6.4 g(FA)/g(cat)-1*h-1 by adapting reactant mass flow as well

as reactant partial pressure (Figure 7.8). Under the mentioned conditions, total OME

selectivity reaches 95% at a conversion of 49% (H-MOR-40) or 47% (Silicalite-1) and, in

contrast to screening conditions, OME3 was detected. Notably, trioxane is also observed as a

by-product. However, the amount of trioxane formed decreases strongly within the first 60

min. reaction time and subsequently remains at a stable level. Initial conversion and selectivity

under adapted conditions was therefore determined at 60-90 minutes reaction time.

Figure 7.8: : Initial conversion/selectivity of H-MOR-40 and Silicalite-1 at increased WHSV and reactant partial pressuredetermined in the interval of 60 - 90 min. reaction time. Reaction conditions: 10 bar, 130 °C, 1 g of catalyst, 400 mL/min inertgas flow, 168 μL/min FA/MeOH solution feed. WHSV for formaldehyde: 6.4 g(FA)/g(cat)-1*h-1.

Page 64: Synthesis of oxymethylene ethers

51 OME synthesis over zeolite catalysts

7.5 Catalyst deactivation and regeneration

Whereas catalytic properties of Silicalite-1 and H-MOR-40 with regards to conversion and

product distribution are very similar, a difference is observed in deactivation behaviour. In

tests that were performed under the same reaction conditions as the catalyst screening,

deactivation proceeded much slower for H-MOR-40 than for Sillicalite-1 (Figure 7.9, left).

The deactivation onset was defined as the time at which conversion has decreased to 85% of

the steady-state conversion level. Deactivation experiments were repeated three times and the

average deactivation onset time was determined to be 38.3 h for H-MOR-40 and 11.1 h for

Silicalite-1 with a broader spread of data in case of H-MOR-40 compared to Silicalite-1. It has

to be noted that after the defined deactivation onset, the conversion drops with a smaller slope

in case of H-MOR-40 as compared to Silicalite-1.

Figure 7.9: Left: Exemplary deactivation curve: Conversion and OME selectivity as a function of time for H MOR 40 andSilicalite-1. The defined deactivation onset is indicated as a dotted grey line. Reaction conditions: 10 bar, 130 °C, 0.5 g ofcatalyst, 100 mL/min inert gas flow, 14 μL/min FA/MeOH solution feed WHSV for formaldehyde: 1.1 g(FA)/g(cat)-1*h-1.Right: TG-MS curve of Silicalite-1 measured in argon.

Several factors can affect the starting point of deactivation. For example, the deactivation

mechanism will have a major impact on the deactivation behaviour of the catalyst. As the

formation of OME is a chain growth reaction, formation of higher, non-volatile OME

homologues in small quantities is expected and could lead to a surface, pore or active site

blocking of the catalyst. In the TG-MS curve of Silicalite-1 measured in an inert gas stream

the release of FA and MeOH along with CO2 and H2O in the range of 170 – 350 °C is evident

(Figure 7.9, right). Similar data is obtained when measured in a stream of air (Figure 12.12).

For H-MOR-40, the mass loss occurs in several stages, but also in this case, the release of the

Page 65: Synthesis of oxymethylene ethers

OME synthesis over zeolite catalysts 52

starting materials FA and MeOH along with CO2 and H2O is observed (Figure 12.13 and

Figure 12.14).

The release of FA and MeOH can either be related to a release of monomeric FA and MeOH

from the pores and/or active sites, or to the presence and decomposition of non-volatile OME

homologues or other non-volatile FA-containing species such as paraformaldehyde. When

pore or surface blocking is discussed as possible deactivation mechanism, several factors can

be considered effective to result in the differences in deactivation onset between Silicalite-1

and H-MOR-40. The two samples have a pronounced difference in crystallite sizes and size

distribution (Silicalite-1: approx. 42 x 8 μm, H-MOR-40 large size distribution with an

average of about 0.15 μm). The smaller external surface area of the Silicalite-1 could result in

a faster blocking of the surface or pore entrances. A further parameter possibly influencing the

deactivation behaviour is the difference in diameters of the micropores (ring size of largest

channel: 12 (MOR) vs. 10 (MFI); computed as 6.45 Å for MOR vs. 4.7 Å for MFI).125

Both catalysts could successfully be regenerated. Silicalite-1 was calcined in air at 550 °C to

restore activity. For H-MOR-40, such a treatment would be too harsh and result in decreased

OME selectivity (vide supra), and so the mordenite sample was regenerated in inert gas flow

at 350°C. Whether such a treatment would also be sufficient for the Silicalite-1 was not

explored. As a proof of principle, the restoration of full performance was demonstrated two

times for each catalyst (Figure 12.15).

7.6 Comparison of siliceous materials

As mentioned above, the amorphous silica reference material (Aerosil 200) was found to be

inactive for OME synthesis, while Silicalite-1 (crystalline zeolite with MFI structure) is one of

the best performing catalysts in this study. In order to investigate the difference between the

two siliceous materials, a FTIR-DRIFTS adsorbate study was performed.

For spectra of activated samples see Figure 12.16. The pristine Aerosil 200 shows only

isolated silanol groups [3746 cm-1]165. Signals in the IR spectrum of Silicalite-1 can be

attributed to unperturbed internal silanol groups [3723 and 3675 cm-1]166 and H-bonded

internal silanol groups and silanol groups interacting with water [broad signal at 3000 - 3600

cm-1]. No isolated external silanols are observed, which can be attributed to the large

dimensions of the Silicalite-1 crystals that feature a very low external surface area compared

Page 66: Synthesis of oxymethylene ethers

53 OME synthesis over zeolite catalysts

to the bulk volume. At the activation temperature, which is the maximal temperature

achievable in the DRIFTS set-up, water is not completely removed as evident from the

presence of a signal at 1634 cm-1 167 and the broadness of the peak at 3000 - 3600 cm-1. A

harsher treatment to completely remove water was not applied, as water being a by-product of

OME formation will also always be present under reaction conditions.

Figure 7.10: Difference spectra of Aerosil 200 and Silicalite-1 after adsorption of probe molecules. For non-substractedspectra, refer to Figure 12.17.

After exposure of samples to FA and MeOH vapour, no additional signals could be observed

in case of Aerosil 200 (Figure 7.10). The signal related to external silanol groups shows

decreased intensity, indicating that there is interaction with adsorbed species. As the reactant

molecules do not seem to adsorb on the Aerosil 200 surface, this decrease might be assigned

to adsorption of additional water molecules. In case of Silicalite-1, a distinct pattern of signals

in the range 2770 - 3000 cm-1 and a group of weak intensity signals at 1449, 1465 and

1475 cm-1 appear upon adsorption of the vapour containing FA and MeOH. Notably, the

spectrum after adsorption of OME1 shows the same features. The OME1 features agree well

with literature data (cf. liquid OME1 spectrum)168. As a reference, pure MeOH was adsorbed

on Silicalite-1. In the considered range, signals at 2950 and 2846 cm-1 are present in the

difference spectrum after adsorption. Considering IR data of formaldehyde from literature

[NIST database: 2785, 2850 and 2995 cm-1]169, the pattern arising after exposure to FA and

Page 67: Synthesis of oxymethylene ethers

OME synthesis over zeolite catalysts 54

MeOH vapour cannot be explained by a superposition of FA and MeOH signals. We assume

that the reactants FA and MeOH have already reacted to OME1 at 40 °C. This is in good

agreement with reports from literature describing liquid phase OME synthesis at temperatures

as low as 50 °C.66

At this point, a clear assignment of activity to certain silanol species in Silicalite-1 is difficult.

In case of the Beckmann rearrangement of cyclohexanone oxime to ε-caprolactam, for which

Silicalite-1 is also highly active and selective, internal silanol nests as well as external silanol

groups are discussed to be the active species.128, 170 The data obtained in this study does not

allow such a straightforward interpretation as IR signals for silanol nests are not well resolved

due to the presence of water. Furthermore, the decrease in signal intensity upon adsorption

could only be assigned to unperturbed internal silanol groups.

From the FTIR-DRIFTS adsorbate study, a clear difference in the adsorption behaviour of

Silicalite-1 compared to amorphous silica was shown. We assume that the high adsorption

potential as present in micropores of the crystalline zeolite may be a key factor for activity in

OME synthesis.

7.7 Conclusions

In summary, a broad range of zeolites was tested in the gas-phase synthesis of OME from

methanol and formaldehyde. It was demonstrated that catalysts characterised by a low number

of Brønsted acid sites and/or EFAl show a better performance and that very weakly acidic

species such as silanol groups can catalyse OME formation with a lower tendency for by-

product formation than strong acid sites.

With respect to catalytic activity, Silicalite-1 and H-MOR-40 showed the best performance.

Both catalysts allow producing OME with selectivity as high as 95%. A deactivation study

showed that H-MOR-40 features increased long-term stability compared to the all-silica

material Silicalite-1, while both catalysts could be fully regenerated by thermal treatment.

Page 68: Synthesis of oxymethylene ethers

55 OME synthesis over supported phosphoric acid

8 OME synthesis over supported phosphoric acid2

In this section, phosphoric acid supported on carbon was investigated as an alternative to

zeolite catalysts for gas-phase OME synthesis. In established supported phosphoric acid

catalysts, for which silica is used as a support, the presence of various phosphor containing

species including silicon phosphates 134 makes a correlation of structure or loading with

activity difficult. When phosphoric acid is supported on a porous carbon (H3PO4/C, see Figure

8.1) and used without thermal treatment at elevated temperatures, no mixed phases or

phosphorylation of the support is expected to occur. This renders analysis, e.g. via 31P MAS

NMR analysis, significantly more simple. Furthermore, the H3PO4/C catalysts can be

synthesised from cheap and readily available materials via simple synthesis protocols. To date,

only few reports of carbon supported phosphoric acid catalysts have been published.138

Alumina is not considered as a support as it suffers from formation of inactive aluminium

phosphate.

Figure 8.1: Schematic representation of supported phosphoric acid catalysts employed in this work.

In the following, the characterisation of prepared H3PO4/C catalysts will be presented.

Subsequently, the activity of H3PO4/C catalysts in the formation of oxymethylene ethers from

methanol and formaldehyde and the activity of related hydrogen phosphates H2PO4- and

HPO42- is evaluated. As zeolites constitute a common alternative to phosphoric acid based

systems in industrial processes,135 the performance of H3PO4/C is additionally compared to a

benchmark zeolite catalyst.

2 The major part of this chapter will be published as Carbon Supported Phosphoric Acid Catalysts for Gas-Phase Synthesis of

Diesel Additives, A. Grünert, W. Schmidt and F. Schüth, to be submitted.

Page 69: Synthesis of oxymethylene ethers

OME synthesis over supported phosphoric acid 56

8.1 Catalyst Characterisation

For the preparation of supported catalysts, it is of interest to firstly study the textural properties

of the chosen support. In this study, the commercial granular activated carbon TC303 supplied

by Silcarbon was used. The nitrogen physisorption isotherm of the pristine granular carbon

(denoted as C-granule) shows typical features of a micro- and mesoporous material (see

Figure 8.2). While the steep increase at low N2 pressure indicates the presence of micropores,

the occurrence of a hysteresis loop is typical for mesoporous materials. Accordingly, the shape

of the hysteresis loop is characteristic for materials containing micro- and mesopores and

corresponds to a H4-type hysteresis loop in IUPAC classification,.171 Furthermore, a total pore

volume of 1.1 cm3/g was determined from physisorption data. From thermogravimetric

analysis (see Figure 12.18), a water content of 6.2% and ash content of 2.1% of the granular

carbon, pretreated as described in chapter 11.3.1, was determined.

Figure 8.2: N2 physisorption isotherm of granular carbon support and the impregnated sample 0.9_H3PO4/ C.

The impact of phosphoric acid impregnation on the sorption properties of the materials was

exemplarily studied using catalyst 0.9_H3PO4/C. The impregnated samples are named

x_H3PO4/C, the prefix x indicating the H3PO4 loading in [g H3PO4 /g C]. The filling of a large

share of pores upon impregnation is reflected by a pronounced decrease of total pore volume

from 1.1 to 0.3 cm3/g. Accordingly, the BET surface area decreased from 1573 m2/g to

203 m2/g. As BET surface area has a limited physical validity for microporous materials, it is

only specified as a means for comparison of the materials.

Page 70: Synthesis of oxymethylene ethers

57 OME synthesis over supported phosphoric acid

For the interpretation of N2-physisorption of supported phosphoric acid, it should be kept in

mind that H3PO4 is solid at measurement temperature (-196 °C), but will be liquid under

reaction conditions (130 °C, melting point H3PO4: 42 °C).172 The decrease of pore volume

evidenced in physisorption analysis hence is not regarded as a pore blocking, but rather gives

information about the filling degree of the pores.

The distribution of phosphoric acid within the carbon support was studied via phosphorus

elemental mapping. The micrographs presented in Figure 8.3 were collected using a scanning

electron microscope (SEM) coupled with energy dispersive X-ray spectroscopy (EDX). It

could be confirmed that the active phase is well distributed within the support matrix.

Figure 8.3: EDX-SEM micrographs of 0.9_H3PO4/C. a)-c): Overview of a representative granule and d)-f) close-up view ofgranule edge. Micrographs show a)/d) secondary electron image b)/e) carbon elemental map and c)/f) phosphorus elementalmap.

In 31P MAS NMR spectra of the as synthesised catalysts, two lines are present (see Figure

8.4). The line at about 0 ppm can readily be assigned to ortho-phosphoric acid (H3PO4) as the

spectra are referenced to 85% H3PO4. The second line at about –13 ppm is related to a

condensed phosphoric acid species, supposedly pyrophosphoric acid. As the chemical shifts of

condensed phosphates are dependent on the local environment,173 it is not surprising that

different chemical shifts have been reported for supported pyrophosphoric acid (H4P2O7)

depending on the support material. While spectra of phosphated zeolites feature lines at both

–5 ppm and –13 ppm,174 in spectra of the SPA catalyst (H3PO4/SiO2) the line at –5 ppm is not

Page 71: Synthesis of oxymethylene ethers

OME synthesis over supported phosphoric acid 58

observed.134 It may be supposed that SPA is a valid reference to H3PO4/C. Hence, the

observed line at –13 ppm indicates the presence of pyrophosphoric acid. The absence of

further lines suggests that the sample contains no polyphosphates.

Figure 8.4: Stacked 31P MAS NMR spectra of H3PO4/C catalysts with varying loading. Positions of spinning side bands aremarked with asterisks. Prefix denotes loading in [g H3PO4/ g C].

As the integrals of the lines can be assumed to be proportional to the amount of phosphorus

(with 1 P per H3PO4 and 2P per H4P2O7), the major share of phosphorus is present in the form

of phosphoric acid.

Under reaction conditions, the catalyst is exposed to gaseous polar components including

water at elevated temperatures (130 °C). Both ortho- and pyrophosphoric acid have melting

points below the reaction temperature (H3PO4: 42.3 °C, H4P2O7: 71.5 °C) and are hygroscopic.

It is therefore expected that, firstly, the active components melt upon preheating the catalyst

and, secondly, a concentrated solution is formed on the catalyst in the presence of water. As

the condensation of phosphoric acid to pyrophosphoric acid is a reversible reaction,

pyrophosphoric acid can undergo hydrolysis to release ortho-phosphoric acid.172 This may be

expected to also occur under reaction conditions.

A linear correlation was found when the amount of phosphorus detected via 31P MAS NMR,

expressed by the integral over the range of [80 ppm, –80 ppm], was related to the expected

amount, as calculated from incipient wetness impregnation, for a range of different samples

Page 72: Synthesis of oxymethylene ethers

59 OME synthesis over supported phosphoric acid

(see Figure 12.19). This indicates that the complete amount of phosphorus is detected via 31P

MAS NMR.

It was not viable to use an external phosphorus standard for comparison due to very long

relaxation times of phosphates. For example, Na2HPO4 requires significantly more than 10

minutes relaxation time per scan (see Figure 12.20). Even ammonium dihydrogen phosphate

(NH4H2PO4), which is a typical standard for quantification in 31P MAS NMR, has a relaxation

time of more than 15 minutes per scan. For the purpose of this study, it was sufficient to

establish a linear correlation of NMR integral to phosphorus loading.

It was not feasible to use methods for acid site characterisation in analogy to the study of

zeolites presented in chapter 7. In both cases, the complete removal of water by thermal

treatment is a prerequisite for collection of meaningful data. In case of H3PO4/C, condensation

of orthophosphoric acid to pyro- and polyphosphoric acid is expected to occur upon thermal

activation. The species are characterised by differences in acidity (H3PO4: pKa1 of 2.16 and

H4P2O7: pKa1 of 0.91),172 which renders the characterisation via NH3-TPD and pyridine-FTIR

difficult.

8.2 Preliminary studies of exemplary H3PO4/C catalyst

Prior to catalytic tests of the impregnated catalysts, the inertness of the support was confirmed.

In a blank test run under reaction conditions, no conversion over the granular carbon was

evidenced. However, it was observed that the carbon material strongly interacts with the

reagents methanol and formaldehyde (see Figure 12.21). In the first data point after the start of

the reaction, it is normal to see a drop in reactant concentration due to inert gas flushing out of

the reactor. However, the detected reagent concentrations increase only slowly thereafter,

reflecting the ongoing adsorption of reagents inside the porous carbon material. It should be

noted that the carbon balance was calculated analogously for all test runs and the herein

described behaviour was only observed to a marginal extent when a reaction occurred.

As a model H3PO4/C catalyst, a material with a loading of 0.9 g H3PO4/ g C was prepared. It

was used to test the general suitability of H3PO4/C catalysts in the gas-phase synthesis of

OME from MeOH and FA. Conversion levels of 47% and OME selectivity of 95% were

measured at moderate WHSVFA = 1.1 (see Figure 8.5, left). OME1, OME2 and the side product

methyl formate was formed. While the initial conversion is lower as compared to benchmark

Page 73: Synthesis of oxymethylene ethers

OME synthesis over supported phosphoric acid 60

zeolites (see chapter 7), the initial selectivity is improved, leading to a comparable initial OME

yield.

The moderate acid strength of phosphoric acid (pKa1 = 2.1) appears to be sufficient to reach

the benchmark performance. In analogy to previously studied series of zeolite catalysts, the

absence of strongly acidic groups prevents the excessive formation of the by-product methyl

formate and the occurrence of dimethyl ether.

The experiences of the preliminary solid acid catalyst screening described in chapter 6

highlighted the need to monitor leaching of active sites. A blind run with an empty reactor was

therefore conducted right after testing the model catalyst. No residual activity due to

contamination of downstream set-up components was observed.

8.3 Impact of H3PO4 loading

In order to test the influence of phosphoric acid loading on the catalyst performance, two

further catalysts with varying phosphoric acid loading were prepared, representing a range of

0.04 to 0.9g H3PO4/ g C.

Figure 8.5: Left: Initial conversion and selectivity after 1 h reaction time of H3PO4/C catalysts with varying phosphoric acidloading. Reaction conditions: 500 mg catalyst, 10 bar, 130 °C, WHSVFA = 1.1. Right: Initial conversion and selectivity after 1h reaction time of 0.9_H3PO4/C catalysts at WHSV for formaldehyde of 1.1 g(FA)/g(cat)-1*h-1 and 42.7 g(FA)/g(cat)-1*h-1.

In contrast to zeolites, a significant change in acid strength upon variation of acid

concentration is not expected for supported phosphoric acid catalysts. This can be rationalised

Page 74: Synthesis of oxymethylene ethers

61 OME synthesis over supported phosphoric acid

by the assumption that unlike protons in zeolites, the phosphoric acid is mobile and therefore

the individual molecules are not significantly influenced by the local environment.

The catalysts showed similar performance over the whole range of phosphoric acid loading

(see Figure 8.5, left). For all test runs, the conversion and selectivity reached a constant level

after 30 minutes of reaction time and showed no change until the end of measurement after

1.5 h reaction time. It can be regarded as an (initial) steady-state conversion/selectivity. The

independence of performance with respect to the active phase loading may be an indication

that the reaction is running close to equilibrium.

In order to evaluate the sensitivity of the reaction towards decreased residence times, the

exemplary catalyst 0.9_H3PO4/C was also tested at the maximal WHSVFA that can be realised

in the test set-up without raising the reactant concentration (WHSVFA = 42.7 h-1). The catalyst

in its granular form showed a drop in conversion to 31%. Additionally, the side reaction

towards methyl formate was supressed (see Figure 8.5, right).

The observed drop in conversion upon increase in WHSV indicates that the transformation is

not running in equilibrium under the respective conditions. This can be a consequence of

reaching the limit of the intrinsic reaction rate or of running in a mass transfer limited regime.

The latter may be caused by microscopic effects within the catalyst granules or by

macroscopic effects such as channel formation in the catalyst bed. A hint towards mass

transfer limitations can be a change in catalyst performance upon change of catalyst particle

size. When the catalyst granules were ground to a fine powder prior to testing, the conversion

improved and a comparable performance to the moderate WHSVFA recurred. This is only a

first indication towards the cause of the observed drop in conversion. It may, however, be

concluded that the powdered catalysts apparently react in equilibrium even at high space

velocity. For diverse reasons, such as improved handling and reproducibility, better

comparability with previous results and to avoid potential interference with mass transfer or

other limitations, all further tests were carried out at moderate weight hourly space velocity

(WHSVFA = 1.1 h-1).

8.4 Sodium phosphates

As phosphoric acid is a polyprotic acid, it is of interest to also study the contributions of the

related proton donating species, namely dihydrogen phosphate (H2PO4-) and hydrogen

phosphate (HPO42-). Their varying deprotonation barriers are reflected in the different

Page 75: Synthesis of oxymethylene ethers

OME synthesis over supported phosphoric acid 62

dissociation constants of phosphoric acid of pKa1 = 2.16, pKa2 = 7.21 and pKa3 = 12.32 (at

25 °C).172 Monosodium phosphate and disodium phosphate were used as representative salts.

A series of catalysts containing equimolar amounts (700 μmol/g C) of H3PO4, NaH2PO4 or

Na2HPO4, respectively, was prepared by incipient wetness impregnation of the respective

components.

Out of the series of catalysts, H3PO4 shows the highest activity in OME formation. In

comparison to H3PO4/C catalysts, activity of the sodium phosphate based catalysts is

significantly reduced (Figure 8.6, left). Conversion drops successively from H3PO4 to

Na2HPO4. For NaH2PO4/C, the conversion is decreased to a level of 10%. The material

Na2HPO4/C is inactive in the studied reaction.

Figure 8.6: Initial conversion and selectivity of granular carbon loaded with 700 μmol /g C of phosphoric acid or sodium (left)and Na2HPO4/C catalysts with different loading (right) after 1h reaction time. Reaction conditions: 500 mg catalyst, 10 bar,130 °C, WHSVFA = 1.1.

The sodium exchange also influences the catalyst selectivity. In case of NaH2PO4/C, the

irreversible formation of methyl formate from formaldehyde is more pronounced which leads

to a decrease in overall OME selectivity. Interestingly, a larger share of OME2 is formed as

compared to the H3PO4/C system. The product distribution of Na2HPO4/C is not meaningful as

the overall conversion was below 1% not allowing reliable quantification of reaction products.

Page 76: Synthesis of oxymethylene ethers

63 OME synthesis over supported phosphoric acid

In analogy to the test runs using H3PO4/C, conversion and selectivity reached an (initial)

steady-state after 30 minutes of reaction time.

As the conversion over NaH2PO4/C is lower than over the H3PO4/C system, it may be argued

that the reaction does not reach equilibrium in this case. The conversion should hence be

sensitive to reaction conditions such as residence time and catalyst loading. Indeed, the

increase of NaH2PO4/C loading from 0.08 to 0.3 g/g C resulted in a doubling of conversion

and an increase in OME selectivity (see Figure 8.6, right).

8.5 Comparison with benchmark zeolite

In order to evaluate the general performance of the H3PO4/C system with respect to the well-

established group of zeolite catalysts, results were compared with those over H-MOR-40. The

catalyst H-MOR-40 was chosen as it shows a superior stability as compared to Silicalite-1.

Furthermore, H-MOR-40 is a classic Brønsted acidic zeolite and is hence more easily

compared to the Brønsted acidic phosphoric acid catalyst. The catalyst stability and

deactivation behaviour were chosen as parameters to be compared.

For a fair comparison, a similar amount of Brønsted acidic sites should be present in the

reactor. The total amount of acid sites of H-MOR-40 was determined via NH3-TPD to be

385 μmol acid sites per gram zeolite.175 An analogous characterisation of H3PO4/C catalysts

with NH3-TPD is difficult due to expected changes in the catalyst state upon thermal

activation. As reasoned above, phosphoric acid may be assumed to be the predominant active

species under reaction conditions. Therefore, the acid site concentration was directly related to

the phosphoric acid loading. It should be acknowledged that this comparison is based on the

before mentioned assumptions and will therefore not yield exact numbers, but acid site

concentrations in the same range.

A representative sample of H3PO4/C with 385 μmol acid sites /g sample was prepared by

dilution of the 0.3_H3PO4/C with the carbon support material. In order to achieve a

satisfactory mixing and to yield a particle size similar to H-MOR-40, the H3PO4/C sample was

ground in a mortar prior to catalytic testing.

In the deactivation study presented in Figure 8.7, the stability of the H3PO4/C catalyst was

identified to be superior to the benchmark zeolite H-MOR-40. The zeolitic catalyst reached

85 % of its steady-state conversion after 21 h reaction time and 50% after 36 h. Even though

Page 77: Synthesis of oxymethylene ethers

OME synthesis over supported phosphoric acid 64

the same reaction conditions were applied as in the deactivation study described in chapter 7.5,

the catalyst lifetime is lower. This may be a result of the low reproducibility of the

deactivation curve of H-MOR-40, which has also been mentioned in chapter 7.5. In contrast to

the zeolite catalyst, the H3PO4/C catalyst kept 87% of its steady-state conversion up to the end

of the measurement at 95 h reaction time. The selectivity is hardly affected by the

deactivation.

Figure 8.7: Conversion and overall OME selectivity as a function of time for H3PO4/C and H-MOR-40 catalysts with the sameconcentration of acid sites. Reaction conditions: 500 mg catalyst, 10 bar, 130 °C, WHSVFA = 1.1.

In the previous study, pore blocking was argued to likely be a main cause for deactivation of

zeolitic catalysts.175 The tested zeolitic structures were purely microporous and therefore

susceptible to deactivation via pore blocking. In contrast, the H3PO4/C catalysts used in this

study have a micro- and mesoporous carbon matrix, which facilitates the access to the active

species. Furthermore, phosphoric acid as the active phase is in a liquid state of matter under

reaction conditions and is mobile, presumably leading to a good dispersion of the active phase.

Additionally, reagents may diffuse into the active phase. In this respect, it is not surprising to

observe deviating deactivation behaviour of the two fundamentally different solid acid

catalysts. Potentially, a combination of the described properties of supported phosphoric acid

is the basis for the deactivation resistance of the catalyst.

Page 78: Synthesis of oxymethylene ethers

65 OME synthesis over supported phosphoric acid

8.6 Conclusions

Carbon supported phosphoric acid catalysts are easily synthesised from cheap materials. Due

to the presence of a limited number of phosphorus-containing species, the catalysts can readily

be analysed using 31P MAS NMR.

In this section, the activity of H3PO4/C catalyst in the gas-phase synthesis of OME from

methanol and formaldehyde was studied in detail. H3PO4/C catalysts have a comparable initial

OME yield and overall steady-state performance as benchmark zeolites.

As the acid loading of H3PO4/C catalysts can be changed independently of acid strength, it

could be demonstrated that the catalyst initial activity is not correlated with the acid loading

under the studied reaction conditions. This indicates that the reaction is running close to

equilibrium. This was supported by the investigation of sodium phosphate catalysts. The acid

strength of the monosodium phosphate (pKa = 7.21) was demonstrated to be too low to reach

reaction equilibrium. Accordingly, no reaction occurred over the disodium phosphate catalysts

(pKa = 12.32).

Two representative materials of H3PO4/C and zeolite catalyst families were compared in a

deactivation study. At a comparable acid site concentration, the H3PO4/C catalyst showed

superior stability over H-MOR-40 zeolite. It was argued that various factors including the

presence of mesopores in the matrix and the liquid state of the active phase may facilitate the

deactivation resistance of the H3PO4/C catalyst.

In this section, we highlighted the benefits of the long-known catalyst system of supported

phosphoric acid catalysts using activated carbon as a support. Although supported phosphoric

acid, typically supported on silica, has in many cases been replaced by zeolite catalysts in acid

catalysis, this work demonstrates a reaction in which the use of H3PO4 based catalyst may be

advantageous.

Page 79: Synthesis of oxymethylene ethers

Two-step synthesis of OME from methanol 66

9 Two-step synthesis of OME from methanol

In this chapter, the implementation of a process combining methanol dehydrogenation and

OME synthesis will be presented. The aim is to show the gas-phase formation of OME from

methanol without intermediate reactant separation.

It was targeted to implement the combined process of methanol dehydrogenation and OME

synthesis as a proof of concept, hence not involving comprehensive studies or optimisation of

reaction conditions and catalysts.

For the above studied second reaction step, namely the formation of OME from methanol and

formaldehyde, the H-MOR-40 zeolite catalyst was employed. For the methanol

dehydrogenation step, the partial non-oxidative methanol dehydrogenation was assessed to be

more suitable than the oxidative route for two reasons. Both routes were introduced in chapter

2.4.1. Firstly, it is advantageous in terms of safety as there is no oxygen involved. In contrast

to oxidative methanol dehydrogenation, explosion limits do not need to be accounted for.

Secondly, the implementation and operation of the non-oxidative route in a laboratory-scale

test set-up is simpler. For example, an additional feed system of oxygen and water vapour is

not required as opposed to the oxidative process.

Table 9.1: Selection of reported catalyst systems for non-oxidative methanol dehydrogenation.

catalyst conversion /selectivity148 / %

stability148 reaction temperature

Na2CO3176

Na2CO3 + C 177

60 / 5750 / 90

stable after 10h ≥ 650 °C

Na-ZSM-5(B) 178, 179 63 / 92 98h at 550 °C 500 - 750 °C

Zn-13X 180 66 / 95 10% deactivationwithin 400 h 550 °C

ZnO/SiO2181

ZnO/SiO2182

75 / 7861 / 94

> 50h450 - 650°C

550 °C

A preliminary selection of catalyst classes to be studied in this section was made on the basis

of literature reports (see chapter 2.4.2). Two alkali- and two zinc-based catalysts were chosen

(see Table 9.1). Sodium carbonate is a non-volatile and cheap catalyst for non-oxidative

methanol dehydrogenation and is easily prepared. Its performance can potentially be improved

Page 80: Synthesis of oxymethylene ethers

67 Two-step synthesis of OME from methanol

by addition of active carbon. Boron-substituted Na-ZSM-5 (Na-ZSM-5(B)) is a thermally

stable catalyst with a reported catalyst lifetime of 98h. Zn-13X and ZnO/SiO2 were also

selected owing to acceptable formaldehyde yield and long catalyst lifetime. Silver-based

catalysts were not included due to the reported recurring need for regeneration in the range of

a few hours.148 In case of copper-based catalyst, favourable catalyst performance and stability

was reported when co-feeding sulphur or selenium containing compound. This, however, adds

excessive complexity to the test set-up and may have adverse effects on the downstream

catalyst. Hence, these catalyst classes were not considered.

Typically, the non-oxidative methanol dehydrogenation is performed at atmospheric pressure,

which is favoured according to Le Chatelier’s principle. However, in the context of this

project the subsequent gas-phase OME formation needs to be taken into consideration. In

contrast to methanol dehydrogenation, increased reaction pressure is favourable.

In the following, the identification of appropriate reaction conditions with a focus on

preventing thermal formaldehyde decomposition is firstly discussed. In the following sections,

the screening of selected catalysts for methanol dehydrogenation and the implementation of

the combined process will be described. The experiments were carried out in a modified

reaction set-up including a reactor for methanol dehydrogenation (R1) and a reactor for OME

synthesis (R2) as indicated in chapter 4.2 and Figure 4.2. The reactor R1 is equipped with a

quartz inlet and can achieve reaction temperatures as high as 650 °C.

9.1 Thermal decomposition of formaldehyde

For non-oxidative MeOH dehydrogenation, a temperature range of 550-650 °C is required. In

this range, thermal formaldehyde decomposition is a relevant side reaction. Formaldehyde is

decomposed to carbon monoxide and hydrogen. In order to obtain meaningful data in the

catalyst screening, it is important to identify a set of reaction parameters at which thermal

decomposition is at a minimum.

For this purpose, the reactant mixture used for OME synthesis containing formaldehyde,

methanol and a minor amount of water, was used to quantify the extent of thermal

decomposition at varying residence times. Thermal decomposition tests were carried out at the

maximum temperature with a reactor filled with inert SiC material and quartz wool. As

permanent gases cannot be separated with the available gas chromatograph, the decomposition

products CO and H2 are not quantified.

Page 81: Synthesis of oxymethylene ethers

Two-step synthesis of OME from methanol 68

When comparing reactant concentrations in the feed stream vs. reactor effluent (see Figure

9.1), it was observed that methanol is not affected. However, pronounced formaldehyde loss

occurs at conditions derived from OME synthesis screening conditions (τ1 in Figure 9.1).

Only 40% of formaldehyde is recovered after passing the reactor. An improvement can be

achieved by reducing the residence time of formaldehyde in the heated zone. The latter was

modified by changing gas-flows and pressure while keeping the reactant concentration

constant. The maximal gas flow in the set-up is limited to 400 ml/min. In case of pressure, a

compromise with OME synthesis needs to be made. At the lowest residence time that can be

implemented in the set-up, an acceptable level of 87% formaldehyde is recovered.

Figure 9.1: Study of thermal decomposition of formaldehyde. Share of initial formaldehyde (green) and methanol (grey) feedrecovered in dependence of residence time. Ratios of residence times τ1:τ2:τ3 of 1:4:8. Reaction conditions: τ1) 14 μL/minFA/MeOH solution, 100 ml/min inert gas, 10 bar; τ2) 56 μL/min FA/MeOH solution, 400 ml/min inert gas, 10 bar; τ3) 56μL/min FA/MeOH solution, 400 ml/min inert gas, 5 bar.

9.2 Catalyst screening

As mentioned above, four materials were selected for the preliminary screening of methanol

dehydrogenation catalysts. Sodium carbonate (Na2CO3) and boron containing Na-MFI zeolite

(Na-ZSM-5(B)) are members of the alkali-mediated catalyst class. In case of ZnO precipitated

on SiO2 (ZnO/SiO2) and Zn2+-ion-exchanged commercial faujasite zeolite 13X (Zn-13X), zinc

is the active component. The catalysts were prepared as specified in chapter 11.2.5. The

successful formation of crystalline phases was verified by powder X-ray diffraction (see

Figure 12.22 to Figure 12.25).

Page 82: Synthesis of oxymethylene ethers

69 Two-step synthesis of OME from methanol

Figure 9.2: Initial conversion and selectivity of a) Na2CO3, b) Na-ZSM-5(B), c) Zn-13X and d) ZnO/SiO2 for non-oxidativemethanol dehydrogenation at 0.5g of 200-300 μm catalyst pellets in 1.5g SiC, 5 bars, 400 ml/min N2 flow, 28 μL/min MeOHfeed, p(MeOH): 0.18 bars. Reaction temperature is 650 °C for Na2CO3 and 550 °C for other catalysts.

For the screening of the selected catalysts, catalyst mass, reaction pressure and reactant

concentration were kept constant. However, different optimal temperature ranges are

recommended in literature. While 550 °C was specified for the majority of catalysts,178-180, 182

Na2CO3 requires a significantly higher reaction temperature of 650 °C and above.176 As the

identification of a suitable catalyst was the main objective of the preliminary screening, the

temperatures were adapted to the respective recommended ranges despite affecting the

Page 83: Synthesis of oxymethylene ethers

Two-step synthesis of OME from methanol 70

comparability of results. It should be noted that selectivity is calculated with reference to

carbon converted. Hence, H2 and H2O are not considered.

The results of the catalyst screening are summarised in Figure 9.2. At first glance, it stands out

that the catalyst performance varies greatly.

For the Na2CO3 catalysts, 35% methanol conversion and formaldehyde selectivity of 55% is

determined. A marginal amount of methane is additionally formed. The residual methanol is

expected to be converted to carbon monoxide and coke deposits. While coke formation is

mentioned in literature,176 it is also clearly evident when comparing catalyst appearance before

and after catalytic testing. The initially white catalyst pellets turn black within two hours of

exposure to reactants (see Figure 9.3). This applies to all catalysts tested in the preliminary

screening.

Figure 9.3: Exemplary images of the quartz reactor inlet before (top) and after (bottom) methanol dehydrogenation testing.The catalyst bed is made up of initially white catalyst pellets and dark SiC granules. The section of the catalyst bed is markedwith blue lines.

The second alkali-based catalyst Na-ZSM-5(B) initially induces complete methanol

conversion to methane and other undetectable products. After 40 minutes reaction time, the

selectivity towards methane is, however suppressed. Instead, dimethyl ether as the major

detectable product and formaldehyde as a minor product are formed at decreasing methanol

conversion. Both Zn-based catalysts exhibit low selectivity towards detectable products. For

ZnO based catalysts, activity is claimed to be strongly correlated to surface area. Accordingly,

high surface area leads to MeOH decomposition only.183 Potentially, the performance of the

ZnO/SiO2 catalyst could be improved by adaptation of the preparation procedure towards low

surface area. However, this is not within the scope of the catalyst screening. Also in case of

Zn-13X, a distinct impact of initial zeolite composition and catalyst preparation procedure is

reported.148 It is interesting to note that the formaldehyde selectivity of Zn-13X increased

Page 84: Synthesis of oxymethylene ethers

71 Two-step synthesis of OME from methanol

slowly but steadily until the end of the screening experiment. Potentially, improved

formaldehyde yield can be obtained at prolonged reaction times.

Although the optimal catalyst performance in terms of methanol conversion and formaldehyde

selectivity reported in literature, which amounts to >60% and 57-95% respectively, was not

reached in the preliminary screening, Na2CO3 could be identified as an adequate catalyst for

the proof of concept implementation of a methanol to OME process.

9.3 Combined process

With the selected catalyst sodium carbonate, a short study of reaction conditions limited to

pellet size and methanol partial pressure was conducted. Subsequently, the combined process

was tested with the derived reaction conditions.

Firstly, it was confirmed that catalyst performance is not significantly influenced by pellet

size. In a comparison of catalyst performance at two different sieve fractions, the same

formaldehyde yield of 20% was obtained. Nevertheless, conversion and selectivity varied to

some extent. At reaction conditions corresponding to the above described catalyst screening,

the 200-300 μm pellet size yielded 57% formaldehyde selectivity at 34% methanol conversion

and the 40-12 μm sieve fraction showed 76% formaldehyde selectivity at 25% methanol

conversion. In contrast, methanol partial pressure has a more pronounced impact on

formaldehyde yield. When the methanol partial pressure was increased, less formaldehyde is

formed while methanol conversion is affected only marginally (see Figure 9.4). A 3.5-fold

increase in methanol partial pressure resulted in a 40% decrease of FA selectivity and a 50%

decrease in FA yield. For the combination of both processes, the following reaction conditions

were chosen: 5 bar, 28 μL/min MeOH feed, 400 ml/min inert gas flow, p(MeOH): 0.18 bars,

reactor R1: 0.5g of 40-125 μm pellets of Na2CO3 in 1.5g SiC, 650 °C, reactor R2: 0.5g of 300-

400 μm pellets of H-MOR-40 in 3g SiC, 130 °C.

Page 85: Synthesis of oxymethylene ethers

Two-step synthesis of OME from methanol 72

Figure 9.4: Initial conversion and selectivity of Na2CO3 as a function of methanol partial pressure. Reaction conditions: at0.5g of 40-125 μm catalyst pellets in 1.5g SiC, 5 bars, 400 ml/min N2 flow, 650 °C.

Figure 9.5: Initial conversion and selectivity for gas-phase OME formation from methanol. Reaction condition: 400 ml/minN2 flow, 28 μL/min MeOH feed, p(total): 5 bars, p(MeOH): 0.18 bars, reactor R1: 0.5g of 40-125 μm pellets of Na2CO3 in1.5g SiC, 650 °C, reactor R2: 0.5g of 300-400 μm pellets of H-MOR-40 in 3g SiC, 130 °C.

Indeed, in the respective test run OME1 was successfully formed from methanol at 60%

conversion and 75% selectivity (see Figure 9.5). Other detectable products include minor

Page 86: Synthesis of oxymethylene ethers

73 Two-step synthesis of OME from methanol

amounts of formaldehyde and methyl formate. The formation of higher OME homologues is

hindered by the low partial pressure of reactants and low FA/MeOH ratio as discussed in

chapter 5.2 and 5.3. As mentioned above, this study was conducted as a proof of concept. A

range of approaches for improvement remain which include optimisation of catalyst

preparation, use of promotors (e.g. active carbon for Na2CO3 catalyst)184 and the more in-

depth study of reaction parameters, such as pressure, temperature, and residence time.

It is interesting to compare the above described results to the selective oxidation of methanol

to OME studied by various groups. The benchmark catalysts in methanol oxidation include

acid modified V2O5/TiO283 and FeMo based catalysts.185 OME1 yields of 46% and 50%,

respectively, have been reported for reactions in a fixed-bed reactor at atmospheric pressure.

In the above described combined process, a yield of 45% OME1 was achieved without further

optimisation of catalysts or reaction conditions.

Page 87: Synthesis of oxymethylene ethers

Summary and final remarks 74

10 Summary and final remarks

In this thesis, the gas-phase synthesis OME from methanol and formaldehyde was

implemented and studied in detail. The design and assembly of a versatile set-up with high

safety standards allowed investigating reaction conditions and the catalytic activity of various

solid acids. With regards to reaction conditions, it was found that low reaction temperature,

high partial pressure of reactants and a high formaldehyde to methanol ratio are key factors

that favour OME yield, especially of OME>1 homologues. For the gas-phase process, this

repeatedly raised the necessity to make compromises. Both reaction temperature and reactant

partial pressure are limited by the saturation pressures of reactants and products. Favourable

conditions can more easily be catered to in a liquid-phase process, which accounts for the

higher yield of oligomeric OME in the liquid phase. Nevertheless, a suitable set of reaction

conditions for gas-phase testing could be identified.

On the basis of a screening of various catalyst classes, zeolites were identified as suitable

catalysts for an in-depth study of structure-activity relations. A general trend that could be

derived from the systematic study of zeolites is that in gas-phase synthesis of OME, zeolitic

materials with a low amount of acid sites show the best performance. The presence of strongly

acidic sites was linked to by-product formation. It could be demonstrated that weakly acidic

functional groups such as silanol groups are sufficient to catalyse OME formation. A

drawback of zeolite catalysts was found in limited catalyst lifetime. In the latter aspect,

supported phosphoric acid outperforms zeolites while exhibiting comparable conversion and

selectivity. From the study of phosphoric acid impregnated on activated carbon, some general

conclusion could be drawn. For example, it was suggested that for highly active catalysts with

low selectivity towards by-products, the reaction runs close to equilibrium. Over active

catalysts, the reaction occurred at such a high rate that non-equilibrium conditions could not

be realised within the experimental limits of reaction parameters. This impeded the additional

comparison of highly active catalysts at low conversion levels in order to get insight into

activity differences. The comparison of benchmark catalysts was hence based on the study of

catalyst deactivation and lifetime. Even at an acid site loading comparable to zeolites,

supported phosphoric acid has an increased lifetime.

In addition to zeolites and supported phosphoric acid, species that leach from ion-exchange

resins and from supported heteropoly acid were identified to be interesting, but difficult to

Page 88: Synthesis of oxymethylene ethers

75 Summary and final remarks

study catalysts. The high yield in OME and the increased selectivity towards OME>1

oligomers may be a motivation for the future in-depth study of these catalyst systems.

Finally, the viability of the gas-phase synthesis of OME from methanol without separation of

intermediates was successfully demonstrated in this work. In future investigations, the

optimisation of the combined process with regards to both catalyst preparation and reaction

conditions may be of interest.

Page 89: Synthesis of oxymethylene ethers

Experimental 76

11 Experimental

11.1 Commercial materials

11.1.1 Gases

All gases were purchased from Air Liquide, including helium (99,999%), argon (99.999%),

hydrogen (99,999%), synthetic air (20.5% O2 in N2), nitrogen (99.999%), calibrated carrier

gas containing 5%CH4/N2 and calibration gas containing 5% DME/N2.

11.1.2 Chemicals

The chemicals used in this work are summarised in Table 11.1. All chemicals were used as

supplied.

Table 11.1: Overview of employed chemicals including specifications and suppliers.

compound specifications supplier

1-butanol >99,8% VWR-International4,4’-trimethylenebis(N-methyl,N-benzyl-pipe-ridinium) dihydroxide

- Provided by Dr. Losch

ammonia 20% aqueous solution Rothboric acid > 99.5% Flukadisodium phosphate >99.0% Sigma Aldrichformic acid >95,0% Sigma Aldrichhydrogen peroxide 35% solution Sigma AldrichLudox AS-40 40% aqueous solution Sigma Aldrichmercaptopropyltrimethoxysilane >95,0% Sigma Aldrichmethanol 99,8% Sigma (Schüth)methyl formate >99,0% Sigma Aldrichmonosodium phosphate >99.0% Sigma Aldrichnitric acid 60% aqueous solution J.T. BakerOME1 99% Sigma AldrichOME3,4 - provided by Dr. Djinovicparaformaldehyde prilled Sigma Aldrichphosphoric acid 85% aqueous solution Alfa Aesarpotassium iodate >98% Sigma Aldrichpotassium iodide >99,0% Sigma Aldrichsilicotungstic acid >99,9% Sigma Aldrichsodium bicarbonate >99,0% Acros

Page 90: Synthesis of oxymethylene ethers

77 Experimental

sodium nitrate >99,0% Sigma Aldrichsodium thiosulfate >98,0% Sigma Aldrichtetraethylorthosilicate >99,0% Sigma Aldrichtetra-n-propyl ammonium hydroxide 40% aqueous solution Sigma Aldrichtetrapropylammonium bromide 98% Sigma Aldrichtoluene >99,0% Sigma Aldrichtrioxane >99,0% Sigma Aldrichzinc nitrate hexahydrate >99,0% Sigma Aldrich

11.1.3 Catalysts and other solid materials

The zeolite catalysts were kindly supplied by Südchemie (now Clariant) and Degussa (now

Evonik Industries). The granular activated carbon TC303 was kindly supplied by Silcarbon

and sulphated zirconium hydroxide and tungstated zirconium hydroxide by MEL Chemicals.

Further materials were purchased by suppliers specified in Table 11.2.

Catalysts were activated as specified in chapter 11.3.1. All powdered catalysts were pressed

and sieved to 300-400 μm pellets after activation. The large crystals of Silicalite-1 were used

as synthesized. Carbon granules were sorted to 1-1.5 mm before use.

Table 11.2: Overview of commercial materials. For zeolite sample, the suffix denotes the SiO2/Al2O3-ratio.

entry material name /abbreviation

material type manufacturer

1 SO4-Zr(OH)4 sulphated zirconium hydroxide MEL chemicals

2WO3-Zr(OH)4 tungstated zirconium

hydroxideMEL chemicals

3PURALOXSCFa -140

γ-alumina Condea (now Sasol)

4Amberlyst 46 sulfonic acid functionalised

ion-exchange resinRohm and Haas

5Amberlyst 36 sulfonic acid functionalised

ion-exchange resinRohm and Haas

6 H-BEA-35 zeolite, Beta Südchemie (now Clariant)7 H-BEA-150 zeolite, Beta Südchemie (now Clariant)8 NH4-FAU-12 zeolite, Faujasite Alfa Aesar9 H-FAU-129 zeolite, Faujasite Degussa (now Evonik)10 H-FAU-340 zeolite, Faujasite Degussa (now Evonik)11 NH4-MOR-14 zeolite, Mordenite Südchemie (now Clariant)12 H-MOR-40 zeolite, Mordenite Südchemie (now Clariant)13 NH4-MFI-27 zeolite, Pentasil Südchemie (now Clariant)14 H-MFI-90 zeolite, Pentasil Südchemie (now Clariant)

Page 91: Synthesis of oxymethylene ethers

Experimental 78

15 13X zeolite, Faujasite Alfa Aesar

16Aerosil 200 hydrophilic fumed silica,

amorphousEvonik

17 TC303 activated carbon granules Silcarbon

18 SiC silicon carbide Alfa Aesar

19 Quartz wool quartz Roth

11.2 Synthesis of catalysts

11.2.1 Supported silicotungstic acid

The supported heteropoly acid catalyst was prepared by impregnation of a commercial

γ-alumina support with an aqueous solution of silicotungstic acid. Briefly, 1g of H4[W12SiO40]

was dissolved in 3ml deionized water. 3g alumina support were slowly wetted with 1.3 ml

impregnation solution and mixed with a spatula. The resulting material was calcined at 200 °C

for 3h under static conditions.

11.2.2 SBA-15-SO3H

The first step, the preparation of thiol functionalised SBA-15 was performed according to

reference 186. Briefly, 1.5 g of SBA-15 was activated for 24 hours under argon in a round

bottom flask. Then, 30 ml of toluene was added and purged with argon for 30 minutes. 1.5 ml

of mercaptopropyltrimethoxysilane was added and refluxed for 24 hours. The solid was then

recovered by centrifugation and dried at 80 °C.

The thiol functionalised SBA-15 was further oxidised with H2O2. For that purpose, 1.3 g of

material was dispersed in 30ml of a 35% H2O2 solution. It was then centrifuged and

resuspended in 40ml of a 1 M H2SO4 solution. Finally, the product was recovered by

centrifugation and dried at 50 °C.

11.2.3 Silicalite-1

31 mL deionised water and 10.672 g tetrapropylammonium bromide were mixed in a 150 mL

Erlenmeyer flask equipped with a magnetic stirring bar. 23.2 mL Ludox AS-40 were

successively added and stirred at 750 rpm for 10 minutes at room temperature. Thereafter, the

synthesis mixture was cooled to 0 °C on an ice bath. 30 mL of a 20% aqueous solution of

ammonia were added to the synthesis mixture at 0 °C. The formed gel was aged for 2 h at 0 °C

while stirring at 750 rpm. The gel was then transferred into three 30 mL Teflon lined stainless

Page 92: Synthesis of oxymethylene ethers

79 Experimental

steel autoclaves. The hydrothermal synthesis was performed in a preheated oven at 180 °C for

7 days. The final solid product was obtained in its pure form by centrifugation and washing

three times with deionised water, drying at 80 °C for 4 h and finally calcining at 550 °C under

static conditions for 7 h (2 °C/min). Large crystals with dimensions of approx. 42 x 8 μm as

determined with an optical microscope were obtained (Figure 12.26). The powder pattern is

presented in Figure 12.27 and nitrogen physisorption isotherms in Figure 12.28. In elemental

analysis, the aluminium content was below the detection limit of 50 ppm.

11.2.4 Supported phosphoric acid

1g of activated carbon granules was impregnated with 1.67 ml of aqueous solution,

corresponding to the filling of the support pore volume. The solution contained the active

component, H3PO4, NaH2PO4 or Na2HPO4, in varying concentrations, depending on the

targeted loading. The carbon granules were supplied in a centrifuge tube and the solution was

added dropwise with repeated shaking of the tube allowing homogeneous impregnation. The

impregnated samples are denoted x_H3PO4/C, the prefix x indicating the H3PO4 loading in [g

H3PO4 /g C]. Analogous nomenclature was used for NaH2PO4 and Na2HPO4 based sample. All

samples were dried at 130 °C overnight.

11.2.5 Methanol dehydrogenation catalysts

Sodium carbonate was prepared via calcination of sodium bicarbonate.177 Briefly, 6 g of

NaHCO3 were spread out in a crucible in a thin layer. The material was calcined in air under

static conditions at 250 °C for 90 minutes with a heating ramp of 2 °C/min. The resulting

powder was partially sieved to 40-125 μm and partially pressed and sieved to 200-300 μm

pellets.

The catalyst Na-ZSM-5(B) was synthesized following instructions from patent literature.179

62.5g of tetraethyl orthosilicate, 0.63g boric acid, 60g of a 20% aqueous solution of tetra-n-

propyl ammonium hydroxide and 5.1g sodium nitrate were mixed and transferred to an

autoclave. The autoclave was heated at 170°C for 72 h. The obtained powder was separated by

filtration and washed with deionised water. Subsequently, it was calcined in air at 450 °C for

8h with a heat ramp of 2 °C/min. Ion-exchange with sodium nitrate was performed at 100 °C

for 3h and repeated 5 times. Finally, the powder was washed with deionised water and dried at

150 °C.

Page 93: Synthesis of oxymethylene ethers

Experimental 80

Zn-13X was prepared via ion-exchange of a commercial 13X zeolite.180 2g of 13X zeolite was

stirred in excess 1N zinc nitrate solution for 1h at 80 °C, then separated by filtration and

washed with deionised water. The exchange was repeated 10 times. The obtained powder was

dried at 110 °C for 3h and subsequently calcined 500 °C for 5h in static air.

The catalyst ZnO-SiO2 was prepared according to patent literature.181, 182 Briefly, 4.35 g of

zinc nitrate hexahydrate were dissolved in 100 ml deionised water mixed with 4.8 ml of a

60 wt% nitric acid solution. Then, 2.5g of tetraethyl orthosilicate were added. The mixture was

heated to 80 °C and stirred for 1h. The resulting solid was dried in a rotary evaporator and

calcined in a stream of air at 600 °C for 5h.

11.3 Modification procedures

11.3.1 Catalyst activation

Commercial zeolites in NH4-form were calcined at 550 °C. 1-2 g of sample were spread out in

a crucible and calcined at 550 °C for 5h with a ramp of 2 °C/min in static air.

H-MOR-40 samples were calcined at varying temperatures (350, 450, 550 °C). For that

purpose, 1 g of H-MOR-40 was prepared in a thin layer in a crucible and was calcined for 4 h

in static air with a heat ramp of 2 °C/min.

Sulphated and tungstated zirconia was obtained by calcination of the respective doped

zirconium hydroxide precursors in static air. The thermal treatment was performed at 550 °C

in case of sulphated and 750 °C in case of tungstated zirconium hydroxide as proposed by the

manufacturer. A heating ramp of 3 °C/min was applied and the final temperature was held for

3h.

The granular carbon support was activated by repeatedly adding hot deionised water (90-

95 °C), vigorously shaking and decanting the supernatant until the supernatant was clear.

Subsequently, the granules were dried at 80 °C until no change in mass was observed.

11.3.2 Sodium exchange of zeolites

For sodium exchange, 2 g of a NH4-form zeolite (NH4-MFI-27 or NH4-MOR-14) were

suspended in 20 mL of 1M NaNO3 solution and stirred for 1 h. This step was repeated twice.

The zeolite powder was then suspended in 20 mL of 1M NaNO3 solution and stirred

overnight. The zeolite was further washed with another aliquot of NaNO3 solution for 1 h. It

Page 94: Synthesis of oxymethylene ethers

81 Experimental

was then separated by filtration, dried at 80 °C for 2 h and at 120 °C for 90 mins, then calcined

under static conditions at 550 °C for 5 h with a heating ramp of 1 °C/min.

11.3.3 Oxalic acid treatment of zeolites

For oxalic acid treatment of zeolites, 1.35 g of oxalic acid was dissolved in 30 ml deionised

water. 1.26 g of zeolite was slurried in oxalic acid solution overnight at room temperature. The

solid was then recovered by filtration and washed with deionised water. Subsequently, the

sample was dried at room temperature and calcined at 350 °C for 4h with a heating ramp of

2 °C/min.

11.3.4 Regeneration protocols

Silicalite-1 was regenerated by calcination under static air at 550 °C for 4 h with a heating rate

of 2 °C/min. H-MOR-40 was regenerated by thermal treatment at 350 °C (heating rate

1 °C/min) in a tube oven for 4 h under inert gas flow (50 mL/min Ar).

11.4 Characterisation methods

11.4.1 X-ray powder diffraction (PXRD)

PXRD data was either recorded in transmission or reflectance mode. Transmission PXRD data

was recorded with a Stoe STADI P transmission diffractometer in Debye–Scherrer geometry.

The device was equipped with a bent primary germanium monochromator for measurements

with monochromatic CuKα1 radiation and a position-sensitive detector made by Stoe.

Powdered samples were prepared in 0.5 mm borosilicate glass capillaries. Reflectance PXRD

data was measured on a Stoe STADI P diffractometer in Bragg-Brentano geometry with

CuKα1 radiation.

11.4.2 Temperature programmed desorption of ammonia (NH3-TPD)

NH3-TPD was performed on a Micromeritics Autochem II 2920 device. 80-100 mg of catalyst

were activated at 500 °C for 1h (heating ramp of 5 °C/min) and then cooled to 150 °C. The

sample was exposed to a flow of 5% NH3/He for 30 min and subsequently purged in He for

2 h. The desorption profile was collected in the range of 100 °C to 800 °C with a heating rate

of 10 °C min−1.

For the H-MOR-40 samples, a milder activation procedure was applied: 100 mg of catalyst

were activated at 350 °C for 5 h (heating ramp of 2 °C min−1) and then cooled to 150 °C.

Page 95: Synthesis of oxymethylene ethers

Experimental 82

11.4.3 Pyridine adsorption followed by FTIR spectroscopy (Py-FTIR)

The acidity of selected samples was determined by adsorbing pyridine inside an FTIR

spectroscopy device (Py-FTIR). Self-supporting wafers (ca. 10 mg/cm2) were activated under

vacuum at 350 °C for 5 h. Then, pyridine (3 mbar) was adsorbed at 150 °C for 20 min.

Thereafter, desorption was carried out under high vacuum at 150 °C, 250 °C and 350 °C for

20 min at each temperature. Spectra were recorded using a Nicolet iS50 equipped with a MCT

detector. The absorption bands centred at 1545 cm-1 (PyH+) and 1455 cm-1 (PyL) were

selected for Brønsted and Lewis acid sites (BAS and LAS) quantification applying their

corresponding integrated molar extinction coefficients, εB=1.67 cm/μmol and

εL=2.22 cm/μmol, respectively.164

11.4.4 Magic-angle spinning nuclear magnetic resonance (MAS-NMR)

The solid-state 27Al MAS-NMR spectra were recorded on a Bruker Avance III HD 500WB

spectrometer using a double-bearing MAS probe (DVT BL4) at a resonance frequency of

130.3 MHz. The spectra were measured by applying single π/12-pulses (0.6 μs) with a recycle

delay of 1 s (6,000 scans) at two different spinning rates (10 kHz and 13 kHz). Prior to the

measurement the samples were saturated with water vapour in a desiccator overnight. The

spectra were referenced to external 1M aqueous solution of AlCl3.

31P MAS NMR spectra were recorded at a resonance frequency of 202.5 MHz. The spectra

were measured by applying single π/2-pulses (3.0 μs) with a recycle delay of 10 s (32 scans) at

a spinning rate of 10 kHz. High-power proton decoupling (spinal64) was applied. Prior to the

measurements the samples were dried at 130 °C for 12 h. The spectra were referenced with

respect to 85% aqueous H3PO4 using solid NH4H2PO4 as secondary reference (δ = 0.81 ppm).

11.4.5 Thermogravimetric analysis coupled with mass spectrometry (TG-MS)

Thermogravimetric analysis (TG) was performed using a NETZSCH STA 449 F3 Jupiter

thermal analysis device. For the determination of ash and water content in carbons,

approximately 3 mg of sample were heated in a stream of 40 mL/min synthetic air with an

additional protective flow of 20 mL/min of argon at a heating rate of 10 °C/min. Data was

collected in the range of 45 - 1000 °C.

In case of zeolite catalysts, the TG method was coupled with mass spectrometry (TG-MS)

using a NETZSCH QMS 403 D Aëolos mass spectrometer. Approximately 5 mg of sample

Page 96: Synthesis of oxymethylene ethers

83 Experimental

were heated in 40 mL/min gas flow (argon or synthetic air) with an additional protective flow

of 20 mL/min of argon. The ramp rate was 10 °C/min in a temperature range of 40 - 900 °C.

Mass spectra were collected in scan mode or in multiple ion detection (MID) mode.

11.4.6 Diffuse reflectance infrared spectroscopy (DRIFTS)

The samples were activated under inert gas flow at 235 °C in a DRIFT cell. For adsorption of

probe molecules, an inert carrier gas flow was bubbled through a probe liquid at room

temperature (reactant mixture 60% FA, 38% MeOH, 2% H2O or neat MeOH or OME1) before

entering the DRIFTS-chamber tempered at 40 °C. The chamber was subsequently purged with

inert gas. All spectra were collected at 40 °C with a Nicolet Magna-IR 560 spectrometer.

11.4.7 Nitrogen physisorption

Nitrogen physisorption was studied using a Micromeritics 3 Flex device. Samples were

activated under vacuum at 250 °C for 8 h in a Smart VacPrep unit. Adsorption and desorption

isotherms were measured at 77.4 K. Data evaluation was performed using the MicroActive

software package by Micromeritics. The total pore volume was determined at p/p0 = 0.95.

11.4.8 GC-MS

For GC-MS measurements, a sample was separated using a Thermofisher Trace-GC Ultra

device equipped with a DB-WAXETR column and was subsequently analysed using a

Thermofisher ISQ mass spectrometer with EI-ionization method.

11.4.9 Scanning electron microscopy (SEM)

SEM micrographs of zeolite samples were measured on a Hitachi S-3500N scanning electron

microscope with 15kV acceleration voltage.

11.4.10 Energy dispersive X-ray spectroscopy (EDX)

SEM-EDX measurements of supported phosphoric acid catalysts were carried out on a Hitachi

S-5500 equipped with a Thermo Scientific NORAN System 7 X-ray Microanalysis System

and a Thermo Scientific UltraDry EDS Detector 30mm2 silicon drift detector. Experiments

were carried out at an acceleration voltage of 30 kV. Cross sections of catalyst granules were

prepared using a Hitachi Ion Milling System E-3500. For this purpose, catalyst granules were

fixed to an aluminium support with graphite based adhesive and milled with argon ions for

12h at 6 kV and 100μA ion current.

Page 97: Synthesis of oxymethylene ethers

Experimental 84

11.4.11 Transmission electron microscopy (TEM)

TEM micrographs of zeolite samples were collected on a Hitachi H-7100 microscope with

100kV acceleration voltage. The samples were physically applied to lacey carbon coated

copper grids.

11.4.12 Elemental analysis

Elemental analysis was performed via absorption spectroscopy at the external service provider

Mikroanalytisches Laboratorium Kolbe in Mülheim a.d. Ruhr.

11.5 Batch reactions

Batch reactions were carried out in a 100 ml stainless steel autoclave at 100 °C for 24 h.

Firstly, 9 g solid trioxane, 1.2 g Amberlyst 46 and a stirring bar were filled into the autoclave.

The latter was sealed and 17.7 ml of OME1 were introduced with a syringe via a ball valve.

The autoclave was then introduced into a heating block preheated to 100 °C. After 24 h, the

reactor was cooled down to room temperature. Samples were taken using a syringe equipped

with a filter.

11.6 Wet-chemical analysis methods

11.6.1 Preparation of methanolic formaldehyde solution

The reactant solution for evaporation in the test set-up was prepared by dissolution of

paraformaldehyde in methanol. Briefly, 120 g paraformaldehyde and 80 g methanol were

mixed in a round bottom flask and refluxed at 80 °C for 24h. Then, the mixture was cooled to

room temperature and filtered.

11.6.2 Iodometry

The formaldehyde content of methanolic formaldehyde solutions was determined via

iodometry. Firstly, a 0.1M thiosulfate (Na2S2O3) reference solution was prepared. For

calibration, a weighted amount of potassium iodate and excess potassium iodide were

dissolved and acidified. Iodine that formed according to equation (11-1) was titrated with

0.1M Na2S2O3 solution, which allowed calculating the exact Na2S2O3 content according to

equation (11-2). Secondly, a 0.05 M iodine solution was prepared by dissolution of iodine and

excess potassium iodide. Its concentration was determined with 0.1M Na2S2O3 solution

following equation (11-2).

Page 98: Synthesis of oxymethylene ethers

85 Experimental

+ 5 + 6 3 + 6 + 3 (11-1)

+ 2 → 2 + (11-2)

+ + 3 + 2 + 2 (11-3)

Formaldehyde analysis was carried out by mixing an aliquot of sample with a known amount

of iodine solution and an aqueous solution of sodium hydroxide. The mixture was allowed to

react according to equation (11-3) for 15 minutes. Then, the solution was acidified with

sulphuric acid and the residual iodine was titrated with 0.1M Na2S2O3 according to equation

(11-2).

11.6.3 Karl-Fischer titration

Water content of methanolic formaldehyde solutions was determined via Karl-Fischer

titration. A Metrohm 831 KF Coulometer equipped with a Metrohm 728 Stirrer was used for

this purpose. Hydranal Coulomar AK and CG-K titration solutions from Honeywell-Fluka,

which are suitable for analysis of ketones and aldehydes, were employed.

Page 99: Synthesis of oxymethylene ethers

Appendix 86

12 Appendix

Figure 12.1: Initial selectivity and conversion of catalysts determined in the interval of 40 - 70 min. reaction time. Reactionconditions: 10 bar, 130 °C, 0.5 g of H-MOR-40, 100 mL/min inert gas flow, 14 μL/min FA/MeOH solution feed, water feedvaried. Weight hourly space velocity (WHSV) for formaldehyde: 1.1 g(FA)*g(cat)-1*h-1.

Figure 12.2: : Study of reproducibility. Initial selectivity and conversion of H-MOR-40 determined in the interval of 40 - 70min. reaction time. Reaction conditions: 10 bar, 130 °C, 0.5 g of catalyst, 100 mL/min inert gas flow, 14 μL/min FA/MeOHsolution feed. WHSVFA: 1.1 g(FA)*g(cat) 1*h-1.

Page 100: Synthesis of oxymethylene ethers

87 Appendix

Figure 12.3: NH3-TPD profiles of H-BEA-35 and H-BEA-150.

Figure 12.4: NH3-TPD profiles of H-FAU-12, H-FAU-129 and H-FAU-340.

Page 101: Synthesis of oxymethylene ethers

Appendix 88

Figure 12.5: NH3-TPD profile of H-MFI-27, H-MFI-90, Silicalite-1 and Aerosil 200.

Figure 12.6: NH3-TPD profile of H-MOR-14 and H-MOR-40.

Page 102: Synthesis of oxymethylene ethers

89 Appendix

Figure 12.7: FTIR spectra of the H-MOR-40 at different calcination temperatures after activation by outgassing at 350 °C: a)ν(OH) vibrations and b) stretching vibration region of the pyridine interacting with the acid sites before adsorption (BA) andafter 20 min desorption at 150 °C, 250 °C and 350 °C.

In order to get more information about the influence of calcination temperature on the

distribution of Brønsted and Lewis acid sites (BAS and LAS, respectively), a pyridine

adsorption study was performed. Figure 12.7 presents the transmission spectra of H-MOR-40

calcined at 350 °C, 450 °C and 550 °C. In the OH stretching vibration region (Figure 12.7a),

five different absorption bands are observed in the spectra before adsorption (BA):187, 188 a)

3745 cm-1: characteristic band of terminal silanols, b) 3732 cm-1: corresponds to silanols

located at internal positions (internal defects), c) 3701 and 3659 cm-1: usually associated to

OH groups located on extra-framework species and d) 3609 cm-1: ascribed to bridging acidic

hydroxyl groups (Si-OH-Al). After adsorption of pyridine at 150 °C, the latter bands fade

away giving rise to the appearance of new bands in the pyridine vibration region (Figure

12.7b). The pyridine interaction with the protons of Brønsted sites leads to typical bands at

1636 and 1545 cm-1 characteristic of pyridinium ions (PyH+).162, 189 On the other hand,

pyridine adsorbed on Lewis acid sites (PyL) is responsible for the band at 1455 cm-1,

corresponding to the 19b vibration mode of the pyridine. Furthermore, analysing the 8a

4000 3800 3600 3400

370137

45

3659

H-MOR-40_550°C

3609

Wavenumber (cm-1)

0.537

32

BA

150 °C

250 °C

350 °C

1650 1575 1500 1425 135014

61

350 °C16

111636 16

21

Abs

orba

nce

(a. u

.)

0.2

1545 1455

150 °C

250 °C

4000 3800 3600 3400

3659

H-MOR-40_450°C

3609

Wavenumber (cm-1)

0.5

BA

150 °C

250 °C

350 °C

3745 37

3237

01

1650 1575 1500 1425 1350

1461

350 °C

161116

36 1621

Abs

orba

nce

(a. u

.)

0.2

1545

1455

150 °C

250 °C

4000 3800 3600 3400

3665

H-MOR-40_350°C

3609

Wavenumber (cm-1)

0.5

BA

150 °C

250 °C

350 °C

3745 37

3237

01

1650 1575 1500 1425 1350

1461

350 °C

161116

3616

21

Abs

orba

nce

(a. u

.)0.2

1545

1455

150 °C

250 °C

a) b)

a) b)

a) b)

Page 103: Synthesis of oxymethylene ethers

Appendix 90

vibration mode region is possible to distinguish two different Lewis species at 1621 and 1611

cm-1, which can be ascribed to the presence of unsaturated Al3+ ions with different

environments.190 In Figure 12.7b), a decrease of the band intensities with the temperature due

to the existence of acid sites with different strengths is also observed. Besides, the formation

of a new band at 1461 cm-1 is associated to iminium ions interacting with some PyL

complexes.187 Apparently, the existence of this band depends on calcination temperature and,

hence, the presence of acidic protons (CBAS, H-MOR-40_350°C > CBAS, H-MOR-40_450°C > CBAS, H-MOR-

40_550°C). This fact explains why this band is sharper for the sample calcined at 350 °C

(possessing higher initial concentration of BAS and more probabilities that some iminium ions

interact with the PyL) than at 450 °C or 550 °C.

Figure 12.8: NH3-TPD profiles of H-MOR-40 treated at varying calcination temperatures. In order to not subject the samplesto change in EFAl-content, a mild activation procedure was chosen: 100 mg of catalyst were activated at 623 K for 5 h(heating ramp of 2 K min−1) and then cooled to 423 K.

Page 104: Synthesis of oxymethylene ethers

91 Appendix

Figure 12.9: Exemplary SEM micrograph of commercial H-FAU-12 zeolite

Figure 12.10: Exemplary particle size histogram of commercial H-FAU-12 including 300 particles measured.

Page 105: Synthesis of oxymethylene ethers

Appendix 92

Figure 12.11: Conversion as a function of external surface area.

Figure 12.12: TG-MS curve of Silicalite-1 measured in synthetic air.

Page 106: Synthesis of oxymethylene ethers

93 Appendix

Figure 12.13: TG-MS curve of H-MOR-40 measured in synthetic air.

Figure 12.14: TG-MS curve of H-MOR-40 measured in argon.

Page 107: Synthesis of oxymethylene ethers

Appendix 94

Figure 12.15: Right: Initial conversion/selectivity determined in the interval of 1-3 h reaction time of fresh samples and ofregenerated samples. Reaction conditions: 10 bar, 130 °C, 0.5 g of catalyst, 100 mL/min inert gas flow, 14 μL/min FA/MeOHsolution feed WHSV for formaldehyde: 1.1 g(FA)/g(cat)-1*h-1.

Figure 12.16: DRIFT-FTIR-spectra of activated Aerosil 200 before and after adsorption of probe molecules.

Page 108: Synthesis of oxymethylene ethers

95 Appendix

Figure 12.17: DRIFT-FTIR-spectra of activated Silicalite-1 before and after adsorption of probe molecules.

Figure 12.18: TG-MS curve of granular carbon support measured in air after activation according to chapter 11.3.1.

Page 109: Synthesis of oxymethylene ethers

Appendix 96

Figure 12.19: Relation of 31P MAS NMR signal per catalyst weigth and the mass of phosphoric acid loaded onto the granularcarbon via incipenet wetness impregnation.

Figure 12.20: 31P MAS NMR of Na2HPO4 with varying scan repeating times of 30s, 60s, 120s, 600s and 8h. Full relaxation of31P nucleus is only achieved after >>600s.

Page 110: Synthesis of oxymethylene ethers

97 Appendix

Figure 12.21: Share of feed mass flow exiting the reactor as a function of time for granular carbon support. Reactionconditions: 500 mg sample, 10 bar, 130 °C, 100 ml/min inert gas flow, 14 uL/min FA/MeOH mixture. WHSV forformaldehyde of 1.1 g(FA)/g(cat)-1*h-1.

Figure 12.22: Powder pattern of Na-ZSM-5(B).

Page 111: Synthesis of oxymethylene ethers

Appendix 98

Figure 12.23: Powder patterns of commercial 13X zeolite and Zn-ion-exchanged 13X zeolite.

Figure 12.24: Powder pattern of ZnO-SiO2 catalyst and reference reflections of ZnO wurzite (PDF number 89-0511).

Page 112: Synthesis of oxymethylene ethers

99 Appendix

Figure 12.25 Powder pattern of Na2CO3 catalyst.

Figure 12.26: Light microscope image of Silicalite-1 crystals.

Page 113: Synthesis of oxymethylene ethers

Appendix 100

Figure 12.27: Powder patterns of Silicalite-1: a) experimental and b) calculated using cif-file from http://www.iza-structure.org/databases/ accessed on 22.02.2018.

Figure 12.28: N2-physisorption isotherm of Silicalite-1 crystals. Step and hysteresis of isotherm in the range of p/po= 0.1 - 0.2is related to a phase transition of adsorbate molecules inside the micropores as described by Müller and Unger.191

Page 114: Synthesis of oxymethylene ethers

101 References to laboratory journal entries

13 References to laboratory journal entries

Table 13.1: References to laboratory journal entries for commercial and synthesised materials.

material sample-ID material sample-ID

H-FAU-12 GRC-GB-011-15 H-MOR-40_550 GRC-GB-011-52H-FAU-129 GRC-GB-011-10 0.9_H3PO4/C GRC-GB-053-04H-FAU-340 GRC-GB-011-09 0.6_H3PO4/C GRC-GB-053-06H-BEA-35 GRC-GB-011-07 0.3_H3PO4/C GRC-GB-053-07H-BEA-150 GRC-GB-011-13 0.04_H3PO4/C GRC-GB-053-11H.MOR-14 GRC-GB-028-01 H3PO4/C (0.7mmol) GRC-GB-053-17H-MOR-40 GRC-GB-011-03 NaH2PO4/C (0.7mmol) GRC-GB-053-18H-MFI-27 GRC-GB-013-01 Na2HPO4/C (0.7mmol) GRC-GB-053-19H-MFI-90 GRC-GB-011-12 0.08_NaH2PO4/C GRC-GB-053-18Silicalite-1 GRC-GB-037-01 0.3_NaH2PO4/C GRC-GB-053-12SO4-ZrO2 GRC-GB-019-01 Na2CO3 GRC-GB-058-01WO3-ZrO2 GRC-GB-019-02 Na-ZSM-5(B) GRC-GB-049-01HPA/Al2O3 GRC-GB-012-01 Zn-13X GRC-GB-050-01Amberlyst 36 GRC-GB-011-66 ZnO/SiO2 GRC-GB-057-03Na-MFI-27 GRC-GB-044-04 SBA-15-SO3H JOI-JA-147Na-MOR-14 GRC-GB-044-03 Aerosil GRC-GB-011-17H-MOR-40_350 GRC-GB-011-53 Carbon support GRC-GB-011-61H-MOR-40_450 GRC-GB-011-51

Table 13.2: References to laboratory journal entries for catalytic tests.

figure description sample-ID

Figure 5.1 temperature dependence GRC-GB-031-04 (H-MOR-40)GRC-GB-031-14 (H-MOR-40)GRC-GB-031-91 (Silicalite-1)GRC-GB-062-15 (Silicalite-1)

Figure 5.2 reversibility GRC-GB-031-74 (MeFO Feed)GRC-GB-031-76 (OME1 + H2O feed)

Figure 5.3 partial pressure

reactant ratio

GRC-GB-031-09GRC-GB-031-31GRC-GB-031-38GRC-GB-031-09GRC-GB-031-10GRC-GB-031-11

Page 115: Synthesis of oxymethylene ethers

References to laboratory journal entries 102

Figure 6.1 preliminary catalyst screening GRC-GB-027-01 (H-MOR-40)GRC-GB-027-02 (H-FAU-5)GRC-GB-027-04 (SO4-ZrO2)GRC-GB-027-05 (WO3-ZrO2)GRC-GB-027-06 (HPA/Al2O3)GRC-GB-027-07 (HPA residual)GRC-GB-025-01 (Amberlyst)

Figure 7.1 zeolite catalyst screening GRC-GB-031-32 (H-FAU-12)GRC-GB-031-19 (H-FAU-129)GRC-GB-031-16 (H-FAU-340)GRC-GB-030-02 (H-BEA-35)GRC-GB-031-08 (H-BEA-150)GRC-GB-031-27 (H-MOR-14)GRC-GB-030-03 (H-MOR-40)GRC-GB-027-03 (H-MFI-27)GRC-GB-031-08 (H-MFI-90)GRC-GB-031-56 (Silicalite-1)

Figure 7.2 conversion and yield as afunction of total ammoniadesorbed

See Figure 7.1, additionally:GRC-GB-043-28 (H-MOR-40_350)GRC-GB-043-26 (H-MOR-40_450)GRC-GB-043-27 (H-MOR-40_550)GRC-GB-031-42 (Aerosil)

Figure 7.4 sodium exchanged zeolite GRC-GB-043-15 (Na-MOR-14)GRC-GB-043-16 (Na-MFI-27)

Figure 7.6 H-MOR-40 calcination See Figure 7.1 and Figure 7.2Figure 7.7 OME yield as a function of

surface areaSee Figure 7.1

Figure 7.8 adapted reaction conditions GRC-GB-031-82 (H-MOR-40)GRC-GB-031-84 (Silicalite-1)

Figure 7.9 deactivation curve GRC-GB-031-89 (H-MOR-40)GRC-GB-031-93 (Silicalite-1)

Figure 8.5 H3PO4/C loading

H3PO4/C WHSV

GRC-GB-043-59 (0.9_ H3PO4/C)GRC-GB-043-84 (0.3_ H3PO4/C)GRC-GB-043-88 (0.04_ H3PO4/C)GRC-GB-043-99 (WHSV = 1.1, granules)GRC-GB-043-79 (WHSV = 42.7, granules)GRC-GB-043-80 (WHSV = 42.7, powder)

Figure 8.6 Ssdium phosphates

NaH2PO4/C loading

GRC-GB-043-94 (H3PO4/C)GRC-GB-043-91 (NaH2PO4/C)GRC-GB-043-93 (Na2HPO4/C)GRC-GB-043-91 (0.08_NaH2PO4/C)GRC-GB-043-82 (0.3_NaH2PO4/C)

Page 116: Synthesis of oxymethylene ethers

103 References to laboratory journal entries

Figure 8.7 deactivation curve GRC-GB-043-98 (H-MOR-40)GRC-GB-043-99 (H3PO4/C)

Figure 9.1 thermal decomposition of FA GRC-GB-062-03Figure 9.2 methanol dehydrogenation

catalyst screeningGRC-GB-062-07 (Na2CO3)GRC-GB-062-11 (Na-ZSM-5(B))GRC-GB-062-12 (Zn-13X)GRC-GB-062-13 (ZnO/SiO2)

Figure 9.4 methanol partial pressure GRC-GB-062-04Figure 9.5 OME formation from MeOH GRC-GB-062-08Figure 12.1 water content GRC-GB-043-37

GRC-GB-043-38GRC-GB-043-39

Figure 12.2 reproducibility GRC-GB-027-01GRC-GB-030-03GRC-GB-031-14GRC-GB-031-30GRC-GB-031-25

Figure 12.11 conversion as a function ofexternal surface area

See Figure 7.1

Figure 12.15 regeneration GRC-GB-043-23 (H-MOR-40)GRC-GB-043-24 (H-MOR-40, 1. regeneration)GRC-GB-043-25 (H-MOR-40, 2. regeneration)GRC-GB-031-91 (Silicalite-1)GRC-GB-043-03 (Silicalite-1, 1. regeneration)GRC-GB-043-05 (Silicalite-1, 2. regeneration)

Figure 12.21 granular carbon support GRC-GB-043-52

Page 117: Synthesis of oxymethylene ethers

References 104

14 References

1. R. Sims, R. Schaeffer, F. Creutzig, X. Cruz-Núñez, M. D’Agosto, D. Dimitriu, M. J. FigueroaMeza, L. Fulton, S. Kobayashi, O. Lah, A. McKinnon, P. Newman, M. Ouyang, J. J. Schauer, D.Sperling and G. Tiwari, Climate Change 2014: Mitigation of Climate Change. Contribution ofWorking Group III to the Fifth Assessment Report of the Intergovernmental Panel on ClimateChange, Cambridge University Press, Cambridge, New York, 2014.

2. G. Centi and S. Perathoner, Catal. Today, 2009, 148, 191-205.3. International Energy Agency, CO2 Emissions From Fuel Combustion Highlights, 2015.4. S. Deutz, D. Bongartz, B. Heuser, A. Kätelhön, L. Schulze Langenhorst, A. Omari, M. Walters, J.

Klankermayer, W. Leitner, A. Mitsos, S. Pischinger and A. Bardow, Energy Environ. Sci., 2018,11, 331-343.

5. C. Vieweg, Zeit Online, 28.08.2017.6. B. Lumpp, D. Rothe, C. Pastötter, R. Lämmermannn and E. Jacob, MTZ worldwide, 2011, 72,

34-38.7. I. A. Reşitoğlu, K. Altinişik and A. Keskin, Clean Technol. Envir., 2014, 17, 15-27.8. M. Natarajan, E. A. Frame, D. W. Naegeli, T. Asmus, W. Clark, J. Garbak, M. A. Gonzalez D, E.

Liney, W. Piel and J. P. Wallace, III, SAE Technical Paper, 2001, 2001-01-3631.9. Ambient air pollution: Health impacts, https://www.who.int/airpollution/ambient/health-

impacts/en/, (accessed 02.08.2019).10. Worldwide Emission Standards and Related Regulations - Passenger Cars / Light and Medium

Duty Vehicles, https://www.continental-automotive.com/getattachment/8f2dedad-b510-4672-a005-3156f77d1f85/EMISSIONBOOKLET_2019, (accessed 25.07.2019).

11. W. Knecht, Energy, 2008, 33, 264-271.12. F. Can, X. Courtois, S. Royer, G. Blanchard, S. Rousseau and D. Duprez, Catal. Today, 2012,

197, 144-154.13. F. Inal and S. M. Senkan, Combust. Sci. Technol., 2002, 174, 1-19.14. L. Lautenschütz, D. Oestreich, P. Seidenspinner, U. Arnold, E. Dinjus and J. Sauer, Fuel, 2016,

173, 129-137.15. J. Burger, M. Siegert, E. Ströfer and H. Hasse, Fuel, 2010, 89, 3315-3319.16. D. Deutsch, D. Oestreich, L. Lautenschütz, P. Haltenort, U. Arnold and J. Sauer, Chem. Ing.

Tech., 2016.17. E. W. Flick, Industrial Solvent Handbook, Noyes Data Corporation, Westwood, 1998.18. F. L. Weaver, A. R. Hough, B. Highman and L. T. Fairhall, Brit. J. Industr. Med., 1951, 8, 279-

283.19. P. C. Kierkus, in e-EROS Encyclopedia of Reagents for Organic Synthesis, eds. P. L. Fuchs, J. W.

Bode, C. A. B., T. Rovis and L. A. Paquette, John Wiley & Sons, Hoboken, 2001.20. H. Hasse, J.-O. Drunsel, J. Burger, U. Schmidt, M. Renner and S. Blagov, US Patent Application,

US2014187823, 2014.21. J.-O. Weidert, J. Burger, M. Renner, S. Blagov and H. Hasse, Ind. End. Chem. Res., 2017, 56,

575-582.22. H. Liu, H. Gao, Y. Ma, Z. Gao and W. Eli, Chem. Eng. Technol., 2012, 35, 841-846.23. X. Zhang, S. Zhang and C. Jian, Chem Eng. Res. Des., 2011, 89, 573-580.24. K. D. Vertin, J. M. Ohi, D. W. Naegeli, K. H. Childress, G. P. Hagen, C. I. McCarthy, A. S. Cheng

and R. W. Dibble, SAE Technical Paper, 1999, 1999-01-1508.25. K. Gaukel, D. Pélerin, M. Härtl, G. Wachtmeister, J. Burger, W. Maus and E. Jacob, 37.

Internationales Wiener Motorensymposium, Wien, 2016.

Page 118: Synthesis of oxymethylene ethers

105 References

26. EN 590 Automotive fuels - diesel - requirements and test methods., European Committee ForStandardization, Brussels.

27. D. S. Moulton and D. W. Naegeli, US Patent, US5746785, 1998.28. M. Härtl, P. Seidenspinner, E. Jacob and G. Wachtmeister, Fuel, 2015, 153, 328-335.29. A. Omari, B. Heuser and S. Pischinger, Fuel, 2017, 209, 232-237.30. K. H. Song and T. A. Litzinger, Combust. Sci. Technol., 2006, 178, 2249-2280.31. M. Härtl, P. Seidenspinner, G. Wachtmeister and E. Jacob, MTZ, 2014, 75, 68-73.32. H. Liu, Z. Wang, J. Zhang, J. Wang and S. Shuai, Appl. Energ., 2017, 185, 1393-1402.33. J. Liu, P. Sun, H. Huang, J. Meng and X. Yao, Appl. Energ., 2017, 202, 527-536.34. S. E. Iannuzzi, C. Barro, K. Boulouchos and J. Burger, Fuel, 2017, 203, 57-67.35. Z. Wang, H. Liu, J. Zhang, J. Wang and S. Shuai, Energy Proced., 2015, 75, 2337-2344.36. L. Pellegrini, M. Marchionna, R. Patrini and S. Florio, SAE Technical Paper, 2013, 2013-01-

1035.37. Y. Ren, Z. Huang, H. Miao, Y. Di, D. Jiang, K. Zeng, B. Liu and X. Wang, Fuel, 2008, 87, 2691-

2697.38. A. Feiling, M. Münz and C. Beidl, MTZ 2016, 16-21.39. J. Liu, H. Wang, Y. Li, Z. Zheng, Z. Xue, H. Shang and M. Yao, Fuel, 2016, 177, 206-216.40. D. Pélerin, K. Gaukel, M. Härtl and G. Wachtmeister, 4. Internationaler Motorenkongress,

Baden-Baden, 2017.41. W. Maus, E. Jacob, M. Härtl, P. Seidenspinner and G. Wachtmeister, 35. Internationales

Wiener Motorensymposium, Wien, 2014.42. W. Sun, G. Wang, S. Li, R. Zhang, B. Yang, J. Yang, Y. Li, C. K. Westbrook and C. K. Law, P.

Combus. Inst., 2016, 1-10.43. S. E. Iannuzzi, C. Barro, K. Boulouchos and J. Burger, Fuel, 2016, 167, 49-59.44. M. Descudé, C. R. Hebd. Seances Acad. Sci., 1904, 138, 1703-1705.45. H. Staudinger, R. Singer, H. Johner, M. Lüthy, W. K. D. Russidis and O. Schweitzer, Liebigs.

Ann. Chem, 1929, 474, 145-275.46. W. F. Gresham and R. E. Brooks, US Patent, US2449469, 1944.47. S. Lüftl and P. M. Visakh, Polyoxymethylene Handbook: Structure, Properties, Applications and

Their Nanocomposites, Scrivener Publishing, Beverly, 2014.48. G. D. Smith, R. L. Jaffe and D. Y. Yoon, J. Phys. Chem., 1994, 98, 9078-9082.49. D. J. Stanonis, W. D. King and S. L. Vail, J. Appl. Polym. Sci., 1972, 16, 1447-1456.50. T. Uchida, Y. Kurita and M. Kubo, J. Polym. Sci., 1956, 19, 365-372.51. R. H. Boyd, J. Polym. Sci., 1961, 50, 133-141.52. G. A. Olah, A. Goeppert and G. K. S. Prakash, Beyond Oil and Gas: The Methanol economy,

Wiley-VCH, Weinheim, 2018.53. G. P. Hagen and M. J. Spangler, International Patent Application, WO0029365, 1998.54. G. P. Hagen and M. J. Spangler, US Patent, US5959156, 1999.55. H. Schelling, E. Stroefer, R. Pinkos, A. Haunert, G.-D. Tebben, H. Hasse and S. Blagov, US

Patent Application, US20070260094, 2007.56. E. Stroefer, H. Hasse and S. Blagov, International Patent Applicaton, WO2007051658, 2007.57. C.-J. Yang and R. B. Jackson, Energ. Policy, 2012, 41, 878-884.58. Y. Zheng, Q. Tang, T. Wang, Y. Liao and J. Wang, Chem. Eng. Technol., 2013, 36, 1951-1956.59. Klimaneutral fahren: Continental testet erfolgreich synthetischen Diesel-Ersatzkraftstoff OME,

https://www.continental.com/de/presse/pressemitteilungen/2017-08-01-klimaneutral-fahren-92062, (accessed 25.07.2019).

60. J. Winterhagen, Frankfurter Allgemeine, 07.06.2017.61. N. Schmitz, J. Burger, E. Ströfer and H. Hasse, Fuel, 2016, 185, 67-72.

Page 119: Synthesis of oxymethylene ethers

References 106

62. A. O. Oyedun, A. Kumar, D. Oestreich, U. Arnold and J. Sauer, Biofuel. Bioprod. Bior., 2018,DOI: 10.1002/bbb.1887.

63. R. H. Ahrend, VDI Nachrichten.64. C. J. Baranowski, A. M. Bahmanpour and O. Kröcher, Appl. Catal. B, 2017, 217, 407-420.65. J. F. Walker, Formaldehyde, Reinhold Publishing, New York, 1964.66. J. Burger, E. Ströfer and H. Hasse, Ind. Eng. Chem. Res., 2012, 51, 12751-12761.67. Y. Zheng, Q. Tang, T. Wang and J. Wang, Chem. Eng. Sci., 2015, 134, 758-766.68. D. Oestreich, L. Lautenschütz, U. Arnold and J. Sauer, Chem. Eng. Sci., 2017, 163, 92-104.69. N. Schmitz, J. Burger and H. Hasse, Ind. Eng. Chem. Res., 2015, 54, 12553-12560.70. Y. Zheng, Q. Tang, T. Wang and J. Wang, Chem. Eng. J., 2015, 278, 183-189.71. P. J. Flory, J. Am. Chem. Soc., 1936, 58, 1877-1885.72. Y. Zhao, Z. Xu, H. Chen, Y. Fu and J. Shen, J. Energy Chem., 2013, 22, 833-836.73. P. Haltenort, K. Hackbarth, D. Oestreich, L. Lautenschütz, U. Arnold and J. Sauer, Catal.

Comm., 2018, 109, 80-84.74. Q. Wu, M. Wang, Y. Hao, H. Li, Y. Zhao and Q. Jiao, Ind. Eng. Chem. Res., 2014, 53, 16254-

16260.75. D. Wang, F. Zhao, G. Zhu and C. Xia, Chem. Eng. J., 2018, 334, 2616-2624.76. Z. Yang, Y. Hu, W. Ma, J. Qi and X. Zhang, Chem. Eng. Technol., 2017, 40, 1784-1791.77. B. G. Schieweck and J. Klankermayer, Angew. Chem. Int. Ed., 2017, 56, 10854-10857.78. K. Thenert, K. Beydoun, J. Wiesenthal, W. Leitner and J. Klankermayer, Angew. Chem. Int. Ed.,

2016, 55, 12266-12269.79. A. Peter, S. M. Fehr, V. Dybbert, D. Himmel, I. Lindner, E. Jacob, M. Ouda, A. Schaadt, R. J.

White, H. Scherer and I. Krossing, Angew. Chem. Int. Ed., 2018, DOI: 10.1002/anie.201802247.80. S. Chen, S. Wang, X. Ma and J. Gong, Chem. Commun., 2011, 47, 9345-9347.81. Y. Fu and J. Shen, Chem. Commun., 2007, 2172-2174.82. H. Guo, D. Li, D. Jiang, W. Li and Y. Sun, Catal. Comm., 2010, 11, 396-400.83. X. Lu, Z. Qin, M. Dong, H. Zhu, G. Wang, Y. Zhao, W. Fan and J. Wang, Fuel, 2011, 90, 1335-

1339.84. O. A. Nikonova, M. Capron, G. Fang, J. Faye, A.-S. Mamede, L. Jalowiecki-Duhamel, F.

Dumeignil and G. A. Seisenbaeva, J. Catal., 2011, 279, 310-318.85. N. T. Prado, F. G. E. Nogueria, A. E. Nogueira, C. A. Nunes, R. Diniz and L. C. A. Oliveira, Energ.

Fuel., 2010, 24, 4793-4796.86. J.-M. Tatibouët and H. Lauron-Pernot, J. Mol. Catal. A, 2001, 171, 205-216.87. K. Thavornprasert, M. Capron, L. Jalowiecki-Duhamel, O. Gardoll, M. Trentesaux, A.-S.

Mamede, G. Fang, J. Faye, N. Touati, H. Vezin, J.-L. Dubois, J.-L. Couturier and F. Dumeignil,Appl. Catal. B, 2014, 145, 126-135.

88. E. Zhan, Y. Li, J. Liu, X. Huang and W. Shen, Catal. Comm., 2009, 10, 2051-2055.89. M. Li, Y. Long, Z. Deng, H. Zhang, X. Yang and G. Wang, Catal. Comm., 2015, 68, 46-48.90. H. Zhao, S. Bennici, J. Shen and A. Auroux, J. Catal., 2010, 272, 176-189.91. X.-J. Gao, W.-F. Wang, Y.-Y. Gu, Z.-Z. Zhang, J.-F. Zhang, Q.-D. Zhang, N. Tsubaki, Y.-Z. Han and

Y.-S. Tan, ChemCatChem, 2018, 10, 273-279.92. Q. Zhang, Y. Tan, G. Liu, J. Zhang and Y. Han, Green Chem., 2014, 16, 4708-4715.93. Q. Zhang, W. Wang, Z. Zhang, Y. Han and Y. Tan, Catalysts, 2016, 6, 43.94. Q. Zhang, Y. Tan, G. Liu, C. Yang and Y. Han, J. Ind. Eng. Chem., 2014, 20, 1869-1874.95. J. Zhang, M. Shi, D. Fang and D. Liu, React. Kinet. Mech. Cat., 2014, 113, 459-470.96. J. Zhang, D. Fang and D. Liu, Ind. Eng. Chem. Res., 2014, 53, 13589-13597.97. J. Burger, N. Schmitz, H. Hasse and E. Stroefer, German Patent, DE102016222657, 2016.98. L. Wang, W.-T. Wu, T. Chen, Q. Chen and M.-Y. He, Chem. Eng. Commun., 2014, 201, 709-717.99. M. Arvidson, M. E. Fakely and M. S. Spencer, J. Mol. Catal., 1987, 4, 391-393.

Page 120: Synthesis of oxymethylene ethers

107 References

100. Z. Xue, H. Shang, C. Xiong, C. Lu, G. An, Z. Zhang, C. Cui and M. Xu, RSC Adv., 2017, 7, 20300-20308.

101. Y. Wu, Z. Li and C. Xia, Ind. Eng. Chem. Res., 2016, 55, 1859-1865.102. L. Lautenschütz, D. Oestreich, P. Haltenort, U. Arnold, E. Dinjus and J. Sauer, Fuel Process.

Technol., 2017, 165, 27-33.103. J. Wu, H. Zhu, Z. Wu, Z. Qin, L. Yan, B. Du, W. Fan and J. Wang, Green Chem., 2015, 17, 2353-

2357.104. C. J. Baranowski, A. M. Bahmanpour, F. Héroguel, J. S. Luterbacher and O. Kröcher, Catal. Sci.

Technol., 2019, 9, 366-376.105. J. Cao, H. Zhu, H. Wang, L. Huang, Z. Qin, W. Fan and J. Wang, Ranliao Huaxue Xuebao, 2014,

42, 986-993.106. W. H. Fu, X. M. Liang, H. Zhang, Y. M. Wang and M. Y. He, Chem. Commun., 2015, 51, 1449-

1452.107. Z. Xue, H. Shang, Z. Zhang, C. Xiong, C. Lu and G. An, Energ. Fuel., 2017, 31, 279-286.108. H. Li, H. Song, F. Zhao, L. Chen and C. Xia, J. Energy Chem., 2015, 24, 239-244.109. X. Y. Li, H. B. Yu, Y. M. Sun, H. B. Wang, T. Guo, Y. L. Sui, J. Miao, X. J. Zeng and S. P. Li, Appl.

Mech. Mater., 2013, 448-453, 2969-2973.110. Y. Liu, Y. Wang and W. Cai, Trans. Tianjin Univ., 2019, 25, 1-8.111. F. Liu, T. Wang, Y. Zheng and J. Wang, J. Catal., 2017, 355, 17-25.112. Q. Zhao, H. Wang, Z. Qin, Z. Wu, J. Wu, W. Fan and J. Wang, J. Fuel Chem. Technol., 2011, 39,

918-923.113. X. Gao, W. Yang, Z. Liu and H. Gao, Cuihua Xuebao, 2012, 33, 1389-1394.114. X. Fang, J. Chen, L. Ye, H. Lin and Y. Yuan, Sci. China Chem., 2015, 58, 131-138.115. R. Wang, Z. Wu, Z. Qin, C. Chen, H. Zhu, J. Wu, G. Chen, W. Fan and J. Wang, Catal. Sci.

Technol., 2016, 6, 993-997.116. H. Li, H. Song, L. Chen and C. Xia, Appl. Catal. B, 2015, 165, 466-476.117. N. Schmitz, F. Homberg, J. Berje, J. Burger and H. Hasse, Ind. Eng. Chem. Res., 2015, 54, 6409-

6417.118. M. Ouda, G. Yarce, R. J. White, M. Hadrich, D. Himmel, A. Schaadt, H. Klein, E. Jacob and I.

Krossing, React. Chem. Eng., 2017, 2, 50-59.119. M. Shi, X. Yu, L. Wang, F. Dai, G. He and Q. Li, Kinet. Catal., 2018, 59, 255-261.120. A. Corma, Curr. Opin. Solid St. M., 1997, 2, 63-75.121. J. Weitkamp and M. Hunger, in Studies in Surface Science and Catalysis, eds. J. Čejka, H. van

Bekkum, A. Corma and F. Schüth, Elsevier, Amsterdam, 2007, vol. 168, pp. 787-835.122. L. B. McCusker and C. Baerlocher, in Studies in Surface Science and Catalysis, eds. J. Čejka, H.

van Bekkum, A. Corma and F. Schüth, Elsevier, Amsterdam, 2007, ch. 2, pp. 13-37.123. E. M. Flanigen, R. W. Broach and S. T. Wilson, Zeolites in Industrial Separation and Catalysis,

Wiley-VCH Weinheim, 2010.124. T. Maesen, in Studies in Surface Science and Catalysis, eds. J. Čejka, H. van Bekkum, A. Corma

and F. Schüth, Elsevier, Amsterdam, Oxford, 2007, vol. 168, pp. 1-12.125. Database of Zeolite Structures, http://www.iza-structure.org/databases/, (accessed

22.06.2018).126. H. G. Karge, in Handbook of Heterogeneous Catalysis, eds. G. Ertl, H. Knötzinger, F. Schüth and

J. Weitkamp, Wiley-VCH Weinheim, 2008, ch. 3.2.4. Acidity and Basicity, pp. 1096-1122.127. A. Corma and H. García, Chem. Rev., 2002, 102, 3837-3892.128. G. Dahlhoff, J. P. M. Niederer and W. F. Hoelderich, Cat. Rev. - Sci. Eng., 2001, 43, 381-441.129. S. Bordiga, P. Ugliengo, A. Damina, C. Lamberti, G. Spoto, A. Zecchina, G. Spanò, R. Buzzoni, L.

Dalloro and F. Rivetti, Top. Catal., 2001, 15, 43-52.130. W. Löwenstein, Am. Mineral., 1954, 39, 92-96.

Page 121: Synthesis of oxymethylene ethers

References 108

131. R. M. Barrer, Pure Appl. Chem., 1979, 51, 1091-1100.132. J. Villadsen and H. Livejerg, Cat. Rev. - Sci. Eng., 1978, 17, 203-272.133. A. Riisager, R. Fehrmann and P. Wasserscheid, in Handbook of Heterogeneous Catalysis, eds.

G. Ertl, H. Knötzinger, F. Schüth and J. Weitkamp, Wiley-VCH, Weinheim, 2008, ch. 2.4.11, pp.631-644.

134. T. R. Krawietz, P. Lin, K. E. Lotterhos, P. D. Torres, D. H. Barich, A. Clearfield and J. F. Haw, J.Am. Chem. Soc., 1998, 120, 8502-8511.

135. F. Cavani, G. Girotti and G. Terzoni, Appl. Catal. A, 1993, 97, 177-196.136. E. H. Brown and C. D. Whitt, Ind. End. Chem., 1952, 615-618.137. W. van der Merwe, Environ. Sci. Technol., 2010, 44, 1806–1812.138. A. Malaika, P. Rechnia-Gorący, M. Kot and M. Kozłowski, Catal. Today 2018, 301, 266-273.139. L. R. R. de Araujo, C. F. Scofield, N. M. R. Pastura and W. de Araujo Gonzalez, Mater. Res.,

2006, 9, 181-184.140. A. M. Puziy, Theor. Exp. Chem., 2011, 47, 277-291.141. E. G. Derouane, J. C. Védrine, R. R. Pinto, P. M. Borges, L. Costa, M. A. N. D. A. Lemos, F.

Lemos and F. R. Ribeiro, Catal. Rev., 2013, 55, 454-515.142. J. Reijenga, A. van Hoof, A. van Loon and B. Teunissen, Anal. Chem. Insights, 2013, 8, 53-71.143. J. F. Haw, Phys. Chem. Chem. Phys., 2002, 4, 5431-5441.144. M. Hunger, in Handbook of Heterogeneous Catalysis, eds. G. Ertl, H. Knötzinger, F. Schüth and

J. Weitkamp, Wiley-VCH, Weinheim, 2008, ch. 3.2.4. Acidity and Basicity, pp. 1163-1178.145. P. Losch, H. R. Joshi, O. Vozniuk, A. Grünert, C. Ochoa-Hernandez, H. Jabraoui, M. Badawi and

W. Schmidt, J. Am. Chem. Soc., 2018, 140, 17790-17799.146. H.-J. Arpe, Industrielle Organische Chemie, Wiley-VCH, Weinheim, 2007.147. S. Su, P. Zaza and A. Renken, Chem. Eng. Technol., 1994, 17, 34-40.148. N. Y. Usachev, I. M. Krukovskii and S. A. Kanaev, Petrol. Chem., 2004, 44, 379-394.149. W. Dai and L. Ren, in Handbook of Heterogeneous Catalysis, eds. G. Ertl, H. Knötzinger, F.

Schüth and J. Weitkamp, Wiley-VCH, Weinheim, 2008, ch. 14.9, pp. 3256-3265.150. B. V. Vora, Top. Catal., 2012, 55, 1297-1308.151. Y. Lei, Q. Sun, Z. Chen and J. Shen, Huaxue Xuebao, 2009, 67, 767-772.152. J. Bandiera and C. Naccache, Appl. Catal. B, 1991, 69, 139-148.153. J. J. Spivey, Chem. Eng. Commun., 2010, 110, 123-142.154. J. Clayden, N. Greeves and S. Warren, Organic Chemistry, Oxford University Press, New York,

2012.155. Y. Ogata and A. Kawasaki, Tetrahedron, 1969, 25, 929-935.156. K. A. Bernard and J. D. Atwood, Organometallics, 1988, 7, 235-236.157. M. Ai, Appl. Catal., 1984, 9, 371-377.158. T. C. Keller, S. Isabettini, D. Verboekend, E. G. Rodrigues and J. Pérez-Ramírez, Chem. Sci.,

2014, 5, 677-684.159. J. Burger, E. Ströfer and H. Hasse, Chem. Eng. Res. Des., 2013, 91, 2648-2662.160. Saturaed Vapor Pressure,

http://ddbonline.ddbst.de/AntoineCalculation/AntoineCalculationCGI.exe?, (accessed26.06.2019).

161. F. Kapteijn and J. A. Moulijn, in Handbook of Heterogeneous Catalysis, eds. G. Ertl, H.Knötzinger, F. Schüth and J. Weitkamp, Wiley-VCH, Weinheim, 2008, ch. 9, pp. 2019-2045.

162. G. Busca, Micropo. Mesopor. Mat., 2017, 254, 3-16.163. T.-H. Chen, K. Houthoofd and P. J. Grobet, Micropo. Mesopor. Mat., 2005, 86, 31-37.164. C. A. Emeis, J. Catal., 1993, 141, 347-354.165. S. Läufer, J. Mol. Struct., 1980, 60, 409-414.

Page 122: Synthesis of oxymethylene ethers

109 References

166. A. Zecchina, S. Bordiga, G. Spoto, L. Marchese, G. Petrini, G. Leofanti and M. Padovan, J. Phys.Chem., 1992, 96, 4985-4990.

167. K. Vikulov, G. Martra, S. Coluccia, D. Miceli, F. Arena, A. Parmaliana and E. Paukshtis, Catal.Lett., 1996, 37, 235-239.

168. J. K. Wilmhurst, Can. J. Chem., 1958, 36, 285-289.169. Formaldheyde, https://webbook.nist.gov/cgi/cbook.cgi?ID=50-00-0, (accessed 16.08.2019).170. C. Flego and L. Dalloro, Micropo. Mesopor. Mat., 2003, 60, 263-271.171. M. Thommes, K. Kaneko, A. V. Neimark, J. P. Olivier, F. Rodriguez-Reinoso, J. Rouquerol and K.

S. W. Sing, Pure Appl. Chem., 2015, 87.172. A. F. Hollemann, E. Wiberg and N. Wiberg, Lehrbuch der anorganischen Chemie, Walter de

Gruyter, Berlin, New York, 2007.173. M. M. Crutchfield, C. V. Callis, R. R. Irani and G. C. Roth, Inorg. Chem., 1962, 1, 813-817.174. G. Lischke, R. Eckelt, H.-G. Jerschekewitz, B. Parlitz, E. Schereier, W. Storek, B. Zibrowius and

G. Öhlmann, J. Catal., 1991, 132, 229-243.175. A. Grünert, P. Losch, C. Ochoa-Hernández, W. Schmidt and F. Schüth, Green Chem., 2018, 20,

4719-4728.176. A. Meyer and A. Renken, Chem. Eng. Technol., 1990, 13, 145-149.177. S. Su, M. R. Prairie and A. Renken, Appl. Catal. A, 1993, 95, 131-142.178. M. G. Howden, Zeolites, 1985, 5, 334-338.179. Y. Matsumara, Japanese Patent Application, JP9040237, 1990.180. M. Akyama and T. Yao, Japanese Patent Application, JP9114532, 1991.181. S. Sago, Japanese Patent Application, 8722737, 1987.182. S. Sago and H. Fuji, Japanese Patent Application, 87289540, 1987.183. M. Sagou, T. Deguchi and N. Nakamura, in Studies in Surface Science and Catalysis, ed. T. Inui,

Elsevier, Amsterdam, 1988.184. S. Su, M. R. Prairie and A. Renken, Appl. Catal. A, 1992, 91, 131-142.185. J. Gornay, X. Sécordel, G. Tesquet, B. de Ménorval, S. Cristol, P. Fongarland, M. Capron, L.

Duhamel, E. Payen, J.-L. Dubois and F. Dumeignil, Green Chem., 2010, 12.186. H. Joshi, D. Jalalpoor, C. Ochoa-Hernández, W. Schmidt and F. Schüth, Chem. Mater., 2018,

30, 8905-8914.187. B. Gil, S. I. Zones, S. J. Hwang, M. Bejblova and J. Cejka, J. Phys. Chem. C, 2008, 112, 2997-

3007.188. O. Marie, P. Massiani and F. Thibault-Starzyk, J. Phys. Chem. B, 2004, 108, 5073-5081.189. J. Fermoso, H. Hernando, P. Jana, I. Moreno, J. Přech, C. Ochoa-Hernández, P. Pizarro, J. M.

Coronado, J. Čejka and D. P. Serrano, Catal. Today, 2016, 277, 171-181.190. C. Poupin, R. Maache, L. Pirault-Roy, R. Brahmi and C. T. Williams, Appl. Catal. A, 2014, 475,

363-370.191. U. Müller and K. K. Unger, in Studies in Surface Science and Catalysis, eds. K. K. Unger, J.

Rouquerol, K. S. W. Sing and H. Kral, Elsevier, Amsterdam, 1988, vol. 39, pp. 101-108.