Sie sind auf Seite 1von 249

Partial Oxidation of Ethene

to Ethylene Oxide in
Microchannel Reactors

von der Fakultat fur Naturwissenschaften der Technischen


Universitat Chemnitz genehmigte Dissertation zur
Erlangung des akademischen Grades

doctor rerum naturalium

(Dr. rer. nat.)

vorgelegt von Dipl.-Chem. Ansgar Kursawe

geboren am 25. April 1971 in Essen


eingereicht am 13. Januar 2009

Gutachter Prof. Dr.-Ing. Elias Klemm


Prof. Dr. Werner A. Goedel

Tag der Verteidigung 11. Dezember 2009


http://archiv.tu-chemnitz.de/pub/2010/0121
Bibliographische Beschreibung und
Referat

Ansgar Kursawe

Partial Oxidation of Ethene to Ethylene Oxide in Micro-


channel Reactors

Technische Universitat Chemnitz, Fakultat fur Naturwissenschaften

Dissertation, 2009, 249 Seiten

In der vorliegenden Arbeit wurde die heterogen katalysierte Oxida-


tion von Ethen zu Ethylenoxid an Silberkatalysatoren untersucht.
Ziel dieser Arbeit war es, Mikrostrukturreaktoren fur schnelle und
stark exotherme Oxidationsreaktionen zu erproben und diese Epoxi-
dation diente als Modellreaktion. Gleichzeitig wurden explosions-
und flammhemmende Eigenschaften des Mikrostrukturreaktors aus-
genutzt, um die partielle Oxidation von Ethen zu Ethyenoxid im
Explosionsbereich (> 9% Sauerstoff) ohne Sicherheitsprobleme zu
ermoglichen.
Um die gesteckten Ziele zu erreichen wurden zwei parallele
Losungswege beschritten. Zunachst wurden modulare Mikrostruk-
turreaktoren und geeignete mikrostrukturierte Katalysatortrager
entwickelt, um Untersuchungen verschiedener katalytischer Be-
schichtungen in dieser neuen und nicht allgemein verfugbaren
Reaktorbauweise zu ermoglichen. Zur katalytische Erprobungen
dieser Konstruktion war es notwendig, geeignete Beschichtungstech-
niken zur Immobilisierung katalytisch aktiver Spezies zu entwickeln.
Durch die Bauweise dieser Reaktoren als Wandreaktor erschien
es anfanglich nicht moglich, kommerziell verfugbare pellet-artige
Katalysatoren zu verwenden.
Daher wurden, parallel zur Konstruktion der modularen Mikrore-
aktoren, verschiedene auf Silber basierende Beschichtungstechniken
hinsichtlich ihrer Eignung fur diese Reaktion erprobt. Zur Erprobung
kamen u.a. Silberimmobilisierung in einem durch anodischen Oxida-
tion erzeugen Porensystem bzw. einer durch Sol-Gel Beschichtung
erzeugten -Aluminiumoxid Schicht und die Abscheidung von metal-
lischem Silber per Vakuumbeschichtung auf einem Tragermaterial.
Zuletzt wurde die Immobilisierung eines gemahlenen, kommerziel-
len Katalysators per elektrostatischer Pulverabscheidung auf einem
entsprechend praparierten Tragermaterial vorgenommen und erfolg-
reich erprobt.
Die wichtigste Erkenntnis dieser Arbeit aus chemisch-katalytischer
Sicht ist der enorm positive Einfluss hoher Sauerstoffkonzentratio-
nen auf die Selektivitat und gleichzeitig den Umsatz des umzuset-
zenden Ethylens. Wird die Sauerstoffkonzentration von unter 10%
auf bis zu 80% erhoht, so steigt die Selektivitat zu Ethylenoxid um
ca. 10% an und simultan kann der Umsatzgrad abhangig vom Kata-
lysator und der Reaktortemperatur um den Faktor 2 bis 10 gestei-
gert werden. Diese Beobachtung wurde fur jeden funktionierenden
Katalysator gemacht, unabhangig von dessen Herstellung. Mittels
Promotoren wie Cs-Salzen und Stickoxiden konnte die Selektivitat
in Abwesenheit gangiger Moderatoren wie Chlorkomponenten auf
70% (Cs-Salze) bzw. 75% (NOx ) gesteigert werden.
Verfahrenstechnisch ist festzuhalten, dass Mikrostrukturreaktoren
gleich welcher Bauweise unter allen Reaktionsbedingungen ther-
misch stabil und beherrschbar blieben. Es wurden Umsatzgrade bis
99% bezuglich Ethen erzielt bzw. Reaktionstemperaturen von uber
630 K bei einem binaren Ethylen-Sauerstoff Gemisch (20%/80%)
angewendet. Angesichts adiabater Temperaturerhohungen von mehr
als 3000 K konnte dennoch ein stabiler Betrieb des Reaktors festge-
stellt werden. Diese thermische Stabilitat war bei Katalysatoren in
Festbettreaktoren nicht gegeben.

Stichworte: Ethen, Ethylenoxid, Epoxidation, Mikrostrukturre-


aktor, Silber, Stickoxid, Caesium, Explosionsbereich, Promotor.
Dank

Mein besonderer Dank gilt Herrn Prof. Honicke fur die hochinteres-
sante Aufgabenstellung, seine stete konstruktive Diskussionsbereit-
schaft wahrend der Erstellung der Arbeit, wertvolle Ratschlage und
Hinweise sowie fur die auergewohnliche Freiheit bei der Planung
und Durchfuhrung der Versuche. Besonders mochte ich mich fur sei-
ne Unterstutzung bezuglich des Entwurfs und des Baus der modu-
laren Reaktoren sowie der dazugehorigen mikrostrukturierten Bau-
teile mit den an einer Universitat verfugbaren Mitteln bedanken.
Ohne diese Ruckendeckung ware die schnelle und unburokratische
Umsetzung eines spontanen interdisziplinaren Entwicklungsprojek-
tes im Niemandsland zwischen Chemie und Fertigungstechnik nicht
moglich gewesen.
Diese modulare Reaktorenentwicklung und die dadurch ermoglichten
umfangreichen katalytischen Erprobungen und Erkenntnisse waren
ohne die konstruktive Mithilfe und den Erfahrungsschatz der me-
chanischen Werkstatten der TU-Chemnitz nicht moglich gewesen.
Deswegen gilt mein Dank hier stellvertretend fur alle Beteiligten
dem Werkstattleiter, Herrn Arnold.
Dem Lehrstuhl fur Fertigungstechnik und Schweitechnik und da-
mit den Herren Professor Durr, Professor Matthes sowie Dr. Pilz
danke ich fur die Anfertigung der durch Drahterosion hergestellten
mikrostrukturierten Wafer und die gewahrte unburokratische Un-
terstutzung. Den Herren Dipl.-Ing. Letsch und Dipl.-Ing. Meyer dan-
ke ich fur ihr Engagement bei Laserschweiss- und Schneidarbeiten
an Aluminiumbauteilen und diversen Dichtungsmaterialien.
Der Fa. CRI-Cataysts und damit Dr. McAteer, Dr. Rubinstein und
Herrn TeRaa danke ich fur die Bereitstellung eines kommerziellen
Ethylenoxid-Katalysators.
Diese Arbeiten wurden im Rahmen des vom BMWI geforderten AiF
Projektes Heterogen katalysierte Gasphasenoxidationen in Mikro-
reaktoren durchgefuhrt. Fur die finanzielle Unterstutzung sei an
dieser Stelle gedankt.
Fur die Anfertigung der elektronenmikroskopischen SEM Aufnah-
men sowie Hilfestellungen bei der Bewaltigung der Tucken einer
Inline-Gaschromatographie mochte ich mich besonders bei Herrn Dr.
Enrico Dietzsch bedanken. Herrn Thomas Kittel vom Otto-Schott
Institut fur Glaschemie der Universitat Jena sei fur aufschlussreiche
elektronenmikroskopische Cs-Backscatter Aufnahmen gedankt. Oh-
ne Ihre Hilfe ware eine schnelle und aussagekraftige Kontrolle der
beschichteten Wafer nur schwer moglich gewesen.
Mein besonderer Dank gilt an dieser Stelle Frau Benndorf und
Frau Wienzek fur die Betreuung der automatisiert arbeitenden Ver-
suchsanlage sowie die GC/MS Kontrolle der Kuhlfallenkondensate
und Frau Reichardt fur die Praparation dutzender Meter anodisch
oxidierter Aluminiumdrahte bzw. Wafer. Herrn Schauer danke ich
fur die Hilfestellung beim elektrischen und elektronischen Aufbau
der Versuchsanlage und sowie die erfolgreiche Eliminierung diverser
Kupferwurmer.
Nicht zuletzt bedanke ich mich bei meinen Eltern, Verwandten und
Freunden fur die moralische Unterstutzung dieser Arbeit.
Contents

1 Introduction 13
1.1 Industrial production of ethylene oxide . . . . . . . . 14
1.1.1 Wurtz-process . . . . . . . . . . . . . . . . . . 14
1.1.2 Direct oxidation process . . . . . . . . . . . . 16
1.2 Handling of ethene oxide . . . . . . . . . . . . . . . . 19
1.3 Usage of ethylene oxide . . . . . . . . . . . . . . . . 21

2 Objectives 23

3 Theory 25
3.1 Reaction mechanism of the direct oxidation process . 25
3.2 Kinetics of the ethene epoxidation . . . . . . . . . . 27
3.3 Catalyst design . . . . . . . . . . . . . . . . . . . . . 30
3.4 Reactor design and heat management . . . . . . . . 34
3.4.1 Reactor design in industrial plants . . . . . . 34
3.4.2 Laboratory and microreactor design . . . . . 41
3.4.3 Advantages of using microchannel reactors . 46

7
8 Contents

4 Results 51
4.1 Epoxidation of ethene in microchannel reactors . . . 53
4.1.1 MCR1: Bulk silver microchannel reactor . . 53
4.1.2 Silver supported on aluminum wafers:
MMCR1-5, MCR2, MCR3 . . . . . . . . . . . 62
4.1.2.1 Test of Ag/Al as a suitable, catalytic
active coating (MMCR1) . . . . . . 62
4.1.2.2 Short term aging of an Ag/Al acti-
vated microchannel reactor (MMCR2) 65
4.1.2.3 Thermal stability of an Ag/Al acti-
vated microchannel reactor (MMCR2) 69
4.1.2.4 Influence of the Ag layer thickness
on selectivity and conversion degree
(MMCR3) . . . . . . . . . . . . . . 72
4.1.2.5 Influence of the aluminum pretreat-
ment on selectivity and conversion
degree (MMCR4,MMCR5) . . . . . 76
4.1.2.6 Silver supported on anodic
preoxidized aluminum surface
(Ag/Al2 O3 /Al, MCR2) . . . . . . . 78
4.1.2.7 Silver supported on metallic Al
(Ag/Al, MCR3) . . . . . . . . . . . 83
4.1.3 Silver supported on -Al2 O3 surface . . . . . 93
4.1.3.1 Silver sputtered on ANOF prepared
-Al2 O3 surface (MMCR6) . . . . . 93
4.1.3.2 Silver sputtered on -Al2 O3 (sol-gel
coating) . . . . . . . . . . . . . . . 95
4.1.3.3 Silver impregnated on -Al2 O3 sur-
face by sol-gel coating (MMCR8) . 96
4.1.3.4 Usage of a commercial SHELL-800
Series, -Al2 O3 based EO silver
catalyst in microchannel reactors
(MMCR9, MMCR10) . . . . . . . . 104
Contents 9

4.1.4 Silver supported on stainless steel surfaces


(MMCR11) . . . . . . . . . . . . . . . . . . . 110
4.1.5 Influence of promotors on Ag/Al coatings
(MMCR12, MMCR13, MCR2Cs) . . . . . . . 112
4.1.5.1 Influence of NO2 and Cs on an
Ag/Al coated microchannel reactor
(MMCR12, MMCR13) . . . . . . . 112
4.1.5.2 Regeneration of MCR2 by immobi-
lization of Cs (MCR2Cs) . . . . . . 120
4.2 Epoxidation in traditional tube-type fixed bed reactors124
4.2.1 Bulk silver catalysts (FBR1, similar MCR1) . 124
4.2.2 Silver supported on Al2 O3 /Al (FBR2, similar
MCR2) . . . . . . . . . . . . . . . . . . . . . 127
4.2.3 Silver on Al (FBR3, similar to MCR3) . . . . 129
4.2.4 Silver supported on -Al2 O3 . . . . . . . . . 132
4.2.4.1 Silver immobilized on -Al2 O3 by
impregnation . . . . . . . . . . . . . 132
4.2.4.2 Use of a commercial SHELL-800 Se-
ries, -Al2 O3 based EO silver cata-
lyst in a fixed bed reactor (FBR4) . 132
4.3 Heat management in microchannel reactors . . . . . 136
4.3.1 Changes of the temperature profile applying
stationary reaction conditions . . . . . . . . . 137
4.3.1.1 Temperature gradients caused by
changes in the reactor temperature 138
4.3.1.2 Temperature gradients caused by
changes of the flow rate . . . . . . . 140
4.3.1.3 Temperature gradients caused
by changes in the ethene feed
concentration . . . . . . . . . . . . . 143
4.3.2 Changes of the temperature profile applying
dynamic reaction conditions . . . . . . . . . 145
10 Contents

4.4 Design aspects of modular microchannel reactors . . 149

5 Discussion 153
5.1 Computation of selectivity vs. conversion behavior . 154
5.1.1 Computation based on published kinetic data 154
5.1.2 Computation based on a triangular reaction
scheme . . . . . . . . . . . . . . . . . . . . . . 157
5.2 Catalytic performance of different silver coatings . . 159
5.3 Influence of the reaction conditions on catalytic per-
formance . . . . . . . . . . . . . . . . . . . . . . . . . 160
5.3.1 Impact of different oxygen concentrations . . 160
5.3.2 Impact of different ethene concentrations . . 162
5.3.3 Impact of the total pressure . . . . . . . . . . 163
5.3.4 Selectivity and conversion behavior of differ-
ent silver catalysts . . . . . . . . . . . . . . . 164
5.3.4.1 Experimental findings . . . . . . . . 164
5.3.4.2 Computations . . . . . . . . . . . . 166
5.3.5 Calculation of activation energies . . . . . . . 173
5.4 Heat effects and reaction engineering aspects . . . . 176
5.4.1 Temperature profiles and heat effects in mi-
crochannel reactors . . . . . . . . . . . . . . . 176
5.4.2 Reactor construction . . . . . . . . . . . . . . 179
5.4.3 Performance parameters of fixed-bed and mi-
crochannel reactors . . . . . . . . . . . . . . . 180
5.4.3.1 Comparison of tube type and cor-
responding microchannel reactors
(FBR1-3, MCR1-3) . . . . . . . . . 181
5.4.3.2 Comparison of reactor performance 182

6 Experimental 187
6.1 Flow apparatus . . . . . . . . . . . . . . . . . . . . . 187
Contents 11

6.1.1 Flow control . . . . . . . . . . . . . . . . . . 187


6.1.2 Analytics . . . . . . . . . . . . . . . . . . . . 189
6.2 Reactor design . . . . . . . . . . . . . . . . . . . . . 191
6.2.1 Pressure resistant laboratory tube type reac-
tor (FBR1-4) . . . . . . . . . . . . . . . . . . 191
6.2.2 Commercial microchannel reactors
(MCR1-3) . . . . . . . . . . . . . . . . . . . . 192
6.2.3 Modular microchannel reactors . . . . . . . . 194
6.2.3.1 Modular Microchannel Reactor
Type I . . . . . . . . . . . . . . . . 194
6.2.3.2 Modular Microchannel Reactor
Type II . . . . . . . . . . . . . . . . 198
6.3 Design and manufacturing of microstructured wafers 202
6.3.1 Wire electro discharge machining . . . . . . . 202
6.3.2 Parallel multiple milling method . . . . . . . 204
6.4 Catalyst preparation and coating procedures . . . . 206
6.4.1 Physical immobilization methods . . . . . . . 207
6.4.2 Anodic oxidation & impregnation . . . . . . . 207
6.4.3 Sol-gel coatings . . . . . . . . . . . . . . . . . 209
6.4.4 Electrostatic powder deposition . . . . . . . . 210

7 Appendix 213
7.1 Chemical properties of ethylene oxide . . . . . . . . 213
7.2 Environmental effects of ethylene oxide . . . . . . . 215
7.3 Physical properties of ethylene oxide . . . . . . . . . 217
7.4 Additional figures and measurements for MCR2 . . . 218
7.5 Additional figures and measurements for MCR3 . . . 222
7.6 Additional figures and measurements for MMCR8 . 226
7.7 Additional figures and measurements for MMCR9/10 229
7.8 Additional figures and measurements for MCR2Cs . 233
12 Contents

7.9 Additional figures and measurements for FBR4 . . . 237

Bibliography 239
Chapter 1
Introduction

Ethylene oxide, otherwise known as ethene oxide or oxirane, is the


most simple cyclic ether and is very reactive. Its highly strained ring
(Fig. 1.1) with a COC angle of only 61.62o can be opened easily.
Thus, it is one of the most versatile chemical intermediates and has a
widespread use in the pharmaceutical and chemical industry. Ethy-
lene oxide (EO) was first discovered by Wurtz [1] in 1859 by liquid
phase oxidation using potassium hydroxide to eliminate hydrochlo-
ric acid from ethylene chlorohydrin and nowadays, a heterogeneous
catalyzed gas phase oxidation is exclusively used. The worldwide
annular production was about 11 million tons in 1986 and raised
to 16 million tons in 2000. With approximately 4 million tons it
ranked in the top 25 among all chemicals produced in the United

Figure 1.1: Structure, binding angels and atomic distances of ethylene


oxide.

13
14 Chapter 1. Introduction

States [2, 3]. In Germany, approximately 715.000t ethylene oxide


were produced in Ludwigshafen (BASF), Dormagen (Erdolchemie),
Gendorf (Clariant) and Marl (DEA) [3].
The epoxidation of ethene with oxygen is a highly exothermic
reaction, which requires cooling of the reactor in order to prevent
a runaway. Furthermore, ethylene and oxygen may form explosive
mixtures, which is in combination with potential hot spots a severe
thread to safe operation.
The microchannel reactor (MCR) is a very new reactor type,
which emerged in the 80s and found soon a widespread use in re-
search departments. This reactor concept emerged from the devel-
opment of small counter-current or cross-flow micro heat exchangers.
The very small distance between heat source and coolant with typ-
ically only some 10 to some hundred micrometers and channel sizes
in the same range result in large surface areas per volume, even low
temperature gradients allow high heat transfer rates.
The principle of a wall reactor eliminates heat transfer problems
between catalyst and heat exchanger surface. Furthermore, the high
ratio of the reactors inner surface to the total volume of the reac-
tor combined with the small channel size proved to function as a
flame and explosion stopping construction, very similar to the well
known 600-800 mesh sieves i.e. for acetylene or petroleum lamps
used in mining since 1816 [4] and invented by Sir Humphrey Davy.
Those lamps (see Fig. 1.2) allow the use of an open flame even in
atmospheres with potentially explosive hydrocarbon concentrations,
utilizing the flame stopping and quenching effect of narrow sieves
on the radical combustion reaction, preventing flame propagation
through the sieve.

1.1 Industrial production of ethylene ox-


ide

1.1.1 Wurtz-process

The industrial production of ethylene oxide started in 1914 [5] with


the chlorohydrin process, a route very similar to Wurtz original
1.1. Industrial production of ethylene oxide 15

Figure 1.2: Miners lamp with flame stopping 600-800 mesh sieve for use
in explosive atmospheres.

preparation method. In the first step, ethene reacts in the liquid


phase with an alkalized, aqueous chlorine solution in order to form
the chlorohydrin. In the following step, H-Cl is eliminated in an
alkaline solution and ethylene oxide formed:

H O
Cl2 + 2OH
2
Cl + ClO + H2 O
ClO + C2 H4 + H2 O Cl CH2 CH2 O + OH
2 Cl CH2 CH2 OH + CaO 2 C2 H4 O + CaCl2 + H2 O

The main disadvantages of this process are the formation of


alkali-chlorides in stoichiometric amounts (such as CaCl2 , if CaO/
Ca(OH)2 (lime) is used), the need for stoichiometric amounts
of chlorine and the formation of chlorinated byproducts. Those
unwanted hydrocarbons are formed by the by radical reaction
of chlorine with ethene (addition and substitution) or the alkali
16 Chapter 1. Introduction

catalyzed hydrolysis of ethylene oxide to ethylene glycol.

Cl2 + C2 H4 C2 H4 Cl2
HCl + C2 H4 C2 H3 Cl
C2 H4 O + H2 O HOCH2 CH2 OH

Therefore, the selectivity to ethylene oxide was about 80%, based on


ethene in this process. The typical requirements and product yields
for the production of ethylene oxide by the chlorohydrin process
are listed in table 1.1 Furthermore, the chlorohydrin process has
also some practical disadvantages due to reactor corrosion by wet
chlorine vapors, which strongly affected capital cost for equipment
[6]. The process had general ecological problems caused by the use of
molecular chlorine and the unwanted production of the chlorinated
by-products. Therefore, this process was abandoned as soon as the
direct oxidation process became available.

1.1.2 Direct oxidation process


In 1931, the direct gas phase oxidation of ethene to ethene oxide was
discovered by Lefort [7]. In this process, the oxidation of ethene is
performed directly with oxygen or air and silver as catalyst. This
heterogeneously catalyzed gas phase oxidation is much easier to han-
dle than the liquid phase reaction of the chlorohydrin process, be-
cause air is free of charge in contrast to chlorine and thus, the direct

Table 1.1: Requirements, yields and side-products of the chlorohydrin


process [6].

Requirements per kg EO Product and Yield


ethene 0.8 kg 80% selectivity towards EO based on
chlorine 2.0 kg ethylene, 95% on chlorohydrin
lime 1.6 kg Side products per kg EO:
electricity 0.02 kWh 0.1-0.15 kg ethylene dichloride
steam 12 kg 0.08 kg 2,2-dichlorodiethylether
water 30 kg 0.0065 kg acetaldehyde
0.01 kg other chlorinated products
1.1. Industrial production of ethylene oxide 17

oxidation is more cost efficiant than the chlorohydrin route. There-


fore, the chlorohydrin route was abandoned in the 1950s [6, 8]. Up
to now, silver is the only catalytic active component for this oxida-
tion process. Initially, about 50% selectivity to ethene oxide were
achieved. The discovery of alkali promoters such as cesium salts
and gaseous moderators as vinylchloride and 1,2-dichloroethane im-
proved the selectivity to nearly 68% in the 1960s and with ongo-
ing research, the selectivity increased to 75-85% within the 70 and
80s [8]. Presently, the application of highly promoted silver cata-
lysts allows selectivities of initially 90% [3, 9].
Historically, there have been two types of direct oxidation pro-
cesses used for the production of ethylene oxide. The first and former
widespread process type uses air, the other and newer one is based
on oxygen for the epoxidation.
The air-based process requires low ethene concentrations in the
feed, which is converted with air to ethylene oxide. In this setup,
two (or in larger plants three) reactors are subsequently arranged
with increasing degree of ethene conversion until nearly complete
conversion of ethene (about 95%) is achieved. Therefore, the first
reactor operates at low conversion degree, but higher selectivities
and the last reactor operates at high conversion degrees, but low
selectivities [10]. The advantage of this process is, that air is free of
charge and as already mentioned, the investment cost is low. The
disadvantage is, that several reactors are required with lower ethene
selectivities. Only low ethene concentrations (< 5%) can be applied
[6] in order to stay out of the explosion range.
In contrast, the oxygen-based process requires only a single re-
actor. In this setup, high concentrations of ethene (up to 40% [5]) are
mixed with oxygen and an inert gas resulting in a mixture having less
than 8% oxygen in order to stay below the lower explosion bound-
ary. This mixture is passed through a single reactor at low ethene
(and oxygen) conversion degrees [10]. After absorption of ethylene
oxide and venting a small purge stream to prevent inert gas enrich-
ment, some carbon dioxide is washed out. The remaining ethylene,
oxygen and inert gas containing stream is enriched with fresh ethy-
lene and oxygen and recycled into the reactor again. Therefore, the
epoxidation takes place in a single reactor with a limited conversion
degree of ethene per cycle and therefore, high selectivities to ethene
18 Chapter 1. Introduction

Figure 1.3: Structural damage to an ethylene oxide purification column


caused by an autoignition of ethylene oxide due to an external hot spot.
(a) Photo of the surrounding installations. (b) Photo of the former column
[12].

oxide. Thus, ethene is better utilized than in the air based process,
although an additional air separation unit for oxygen enrichment is
required. Despite the higher investment costs, nowadays the oxygen
based process is the only one left.
1.2. Handling of ethene oxide 19

1.2 Handling of ethene oxide


Due to the reactivity of ethylene oxide even in absence of other
chemicals, the handling of pure ethylene oxide and the use of it
are potentially dangerous, despite its toxic effects. There had been
numerous incidents in the past with great damage caused by the ex-
plosion or decomposition of ethylene oxide or its vapor. Sometimes
small leaks led to severe explosions. In 1987, a catastrophic explo-
sion of an ethylene oxide purification column occurred [12]. Due to a
leak of a manhole flange, ethylene oxide got in contact with mineral
wool used as insulating material. The following exothermic isomer-
ization, disproportionation and decomposition of ethylene oxide and
reactions with moisture caused an external hot spot, which heated
the column up to temperatures above the autoignition temperature
of ethylene oxide, resulting in an ignition / explosion of the com-
plete column. The great structural damage is shown in figure 1.3.
The column itself was completely destroyed and severe structural
damage to the surrounding installations within a radius of several
hundred meters had to be noted.
In another incident, small amounts of ethylene oxide exhibited
an extraordinarily high destructive potential. Due to a leakage of
two blocked discharge valves, about 300 g of ethylene oxide got into
the head of a high speed pump, which was normally idle and used
as reserve pump. A fault within the electrical installation caused
the blocked pump to start and due to internal friction, the pump
heated up to ethylene oxide decomposition temperature within a
few minutes. The resulting explosion (Fig. 1.4) caused twelve 3/4
stainless steel bolts to fail and the motor of the pump, having a
weight of approx 1/2 ton was catapulted over a distance of 20 meters
[12].
Therefore, ethylene oxide is a dangerous and poisonous chemical
and its production, purification and transport should be handled
with extreme care in order to avoid ignition or emission.
20 Chapter 1. Introduction

Figure 1.4: Damage of a pump ( 500 kg), caused by 300g decomposed


ethylene oxide. (a) Head of the pump with ruptured 3/4 stainless steel
bolts. (b) Photo of the pumps motor, which was catapulted over a dis-
tance of 20 meters [12].
1.3. Usage of ethylene oxide 21

1.3 Usage of ethylene oxide


Despite its toxic and explosive properties, ethylene oxide is used in
many industrial products because of its high reactivity and versa-
tility (see Appendix, page 213 for an overview). A brief overview
about the usage of ethylene oxide and its main products is given in
table 1.2 and 1.3
Ethylene glycols are produced by thermal hydration of ethylene
oxide into a mixture of mono- (MEG), di- (DEG), tri- (TEG) and
polyethylene glycols (PEG) [10, 11, 13]:

n C2 H4 O+H2 O HO(CH2 CH2 O)n H Hn=1 = 79.4 kJ/mol


The main usage of MEG is the production of PET (PolyEthyleneTereph-
thalate) and use as anti-freeze in automotive cooling systems. DEG

Table 1.2: Usage of ethylene oxide in Western Europe in 1980 [10].

product usage / %
ethylene glycols 55
polyglycols 4
ethanolamines 7
glycol ethers 12
surface active agents 12
polyols 4
other 6

Table 1.3: Conversion of ethylene oxide to glycols by region [11], World-


wide production 18 million metric tons in 2006

North America 73%


Western Europe 44%
Japan 63%
Other Asia 90%
Africa 99%
22 Chapter 1. Introduction

and TEG are used for gas treatment as absorbent and TEG is
furthermore needed for production of cellophane. PEG is used
in the cosmetic and pharmaceutical industry as base material for
carrying the active ingredients [10].
Ethanolamines, which are also widely used in pharmaceutical
and cosmetic products are obtained by the reaction of ethylene oxide
and ammonia:

n C2 H4 O + NH3 Hn3 N(C2 H4 OH)n n = 1 3

Glycolethers are produced by the reaction of ethylene oxide and


ethers such as Dimethylether [5]:

CH3 OCH3 + nC2 H4 O CH3 O(CH2 CH2 O)n CH3


n=14

Those ethers are mainly used as solvents, detergents, brake fluid and
extracting agent for SO2 , H2 S, CO2 and mercaptanes from natural
gas. Minor amounts of ethylene oxide are directly used as sterilizing
agent in hospitals and the food industry [12, 14] (see Appendix, page
215 for toxicity and environmental data) .
Chapter 2
Objectives

At the end of 1996, there were few successful performed reactions in


microchannel reactors, and most of them concentrated on testing the
concept of the microchannel reactor with relatively simple reactions
such as total combustion or hydrogenation reactions. The former
lacked any selectivity problem and the latter lacked sufficient heat
production to justify the use of a microchannel reactor. Therefore,
microchannel reactors had to be characterized for their suitability
of performing highly exothermic reactions with potential selectivity
problems.
In the present study, the partial oxidation of ethene as an
exothermic and fast reaction was chosen as a model reaction in order
to evaluate the performance of microchannel reactors. Furthermore,
there is few data available in the open scientific literature about the
partial oxidation of ethene applying oxygen feed concentrations in
the explosion range, although this reaction has been under investi-
gation for now more than 50 years. It is normally too dangerous to
apply suchlike reaction conditions in conventional reactors, because
a single hot spot will result in an uncontrolled runaway, likely ending
in an explosion.
Therefore, the exothermic partial oxidation of ethene was uti-
lized both to evaluate the performance of microchannel reactors and
to investigate the ethene epoxidation at oxygen concentrations in
the explosion range.

23
24 Chapter 2. Objectives

In order to achieve these goals, this work concentrated on a few,


but important steps. First of all, it was necessary to develop a new
catalyst preparation method, because normal silver based and indus-
trially available catalysts are meant to be used in a fixed bed and
seemed unsuitable for the use at the walls of a microchannel. Second,
a suchlike immobilized catalyst had to be used and characterized in
a fixed bed before similar experiments applying comparable reaction
conditions, had to be performed in corresponding microchannel re-
actors having the same catalytic surface. In parallel, it was intended
to increase the oxygen concentration within the explosion range in
order to gain experience with the catalytic properties of silver cata-
lysts using high oxygen concentrations in the feed.
Due to the limited number of available, commercially manufac-
tured microchannel reactors for solid-gas catalysis, the initial cat-
alyst screening should be performed in a conventional fixed bed
reactor. In parallel, it was tried to manufacture reactors having
removable microstructured parts in order to allow more and non de-
structive tests. As a side effect, the flexibility allowed access and
exchange of the activated microstructured parts for analytical inves-
tigations applying surface science techniques without destroying the
expensive welded commercial reactors.
Chapter 3
Theory

In this chapter, assumptions about the potential reaction mechanism


of the direct epoxidation of ethene are made. The influence of reac-
tion parameters like reactant partial pressure, reactor temperature
and selectivity enhancing additives is discussed and the preparation
of commonly used catalysts described. Due to the exothermic re-
action, the reactor design plays an important role in the industrial
performance of this reaction. Therefore, results from a Dutch re-
search group dealing with the heat management of a conventional,
industry size tubular reactor are presented. Based on their kinetic
models, the influence of reaction parameters like ethene and oxygen
concentration on the selectivity and conversion degree was calcu-
lated. Eventually, the concept of a microchannel reactor, its use in
industry with its advantages and disadvantages is discussed.

3.1 Reaction mechanism of the direct ox-


idation process
The reaction mechanism of the ethene epoxidation has been exten-
sively studied, but up to now there is no common agreement about
the mechanism. In the 1970, a reaction mechanism strongly sup-
ported by Kilty and Sachtler [15] was proposed, who predicted the
following mechanism involving atomically (Ag O) and molecularly

25
26 Chapter 3. Theory

(Ag O2 ) absorbed oxygen species on on catalytic active silver sur-


faces (Ag ):

6 O2 + 6 Ag 6 (Ag O2 )
3 O2 + 6 Ag 6 (Ag O)

6 (Ag O2 ) + 6 C2 H4 6 C2 H4 O + 6 (Ag O)
6 (Ag O) + C2 H4 2 CO2 + 2 H2 O + 6 Ag

It is suggested, that chlorine blocks the dissociative adsorption of


oxygen and therefore, enhances selectivity. Assuming a completely
blocked dissociative adsorption of ethene on silver, the former re-
action steps can be summarized and the following equation for the
formation of ethylene oxide is obtained:

7 C2 H4 + 6 O2,ads 6 C2 H4 O + 2 CO2 + 2 H2 O

Thus, the maximum selectivity to ethene oxide is 6/7 or 85.7% if the


assumption of this reaction mechanism is correct. Basically, there
are numerous TPD-studies [16, 17] dealing with the state of oxy-
gen adsorbed on silver surfaces and it was shown, that both types
of adsorption take place when silver is exposed to oxygen at higher
temperatures. Furthermore, this reaction mechanism is supported
by a study of Herzog [18], who performed the epoxidation with ni-
trous oxide instead of oxygen. The decomposition of N2 O yields
atomically adsorbed oxygen and it showed, that the only products
observed were CO2 and H2 O.

6 N2 O + 6 Ag 6 (Ag O) + 6 N2
6 (Ag O) + C2 H4 2 CO2 + 2 H2 O + 6 Ag

Later, a scheme of a different reaction mechanism emerged af-


ter several authors found by surface science and especially isotope
exchange techniques, that atomically adsorbed oxygen rather than
molecular is involved in the selective oxidation of ethene [19, 20].
Furthermore, the role of subsurface oxygen, which is definitely in-
volved in the epoxidation [21] remained unconsidered in this model.
Eventually, the 6/7 selectivity barrier predicted by the old mech-
anism had been substantially broken and thus, the mechanism was
obsolete.
3.2. Kinetics of the ethene epoxidation 27

Results from TAP-experiments1 [22] indicated, that the forma-


tion of ethylene oxide requires a pretreatment of the clean silver sur-
face with oxygen. An oxygen free silver powder proved to be initially
inactive for ethene epoxidation. Only after several oxygen / ethylene
gas-pulses, ethylene oxide formation was observed and the initially
dominating total oxidation suppressed. This is ascribed to the for-
mation of subsurface oxygen, which allows the stabilization or for-
mation of a certain oxygen surface species, which is in turn required
for the stabilisation of an epoxidation active and selective surface-
oxygen species. Results obtained by Grant & Lambert [23] support
this theory, because it showed necessary to pretreat an Ag(111) sin-
gle crystal with oxygen in order to activate it for ethene epoxida-
tion. Isotope exchange experiments with 16 O2 and 18 O2 by Bertole
& Mims showed, that the selectivity to ethylene oxide is directly
correlated with the availability of subsurface oxygen [24].
Another hint of atomically absorbed oxygen being the crucial
epoxidation active species are calculations performed by Salazar
et.al [25]. The authors found, that caesium lowers the dissocia-
tion barrier between molecular and atomically absorbed oxygen and
proved a correlation between their calculated barrier height, which
is a function of the caesium coverage, and the observed selectiv-
ity toward ethylene oxide, again as a function of caesium coverage.
Therefore, current investigations aim toward a better understand-
ing of surface/subsurface oxygen interactions and the influence of
known promoters such as Cs and Cl on the specific oxygen ad- and
absorption on silver [26].

3.2 Kinetics of the ethene epoxidation


In commercial ethylene oxide plants, the partial pressures of ethene,
oxygen and reaction-modifier may be varied. In many scientific pub-
lications, the ethene epoxidation was investigated in absence of chlo-
rine or other reaction modifiers. The selectivity is expected to in-
crease with increasing oxygen partial pressure as shown by Akella &
Lee [27], Cant & Hall [28] and Klugherz & Harriott [8]. It was ob-
served the epoxidation reaction rate increasing faster with increasing
1 Temporal Analysis of Products, a fast and time resolved product detection

by mass spectroscopy of gas-pulses passing a catalyst in a vacuum chamber.


28 Chapter 3. Theory

oxygen partial pressure than the rate of the total oxidation. There-
fore, higher oxygen partial pressures improve the degree of conver-
sion as well as the selectivity to ethylene oxide Contrary, Borman &
Westerterp [29] found in more recent investigations, that there is no
difference in the reaction rate whether 7% or 12% oxygen are used,
applying 1% C2 H4 in the feed.
Furthermore, the reaction rate of ethene to ethylene oxide on
the one hand and to carbon dioxide on the other increase nearly
simultaneously with increasing ethene partial pressure, showing a
maximum in both reaction rates [8]. After passing the maximum,
the reaction rate of the total oxidation drops faster the rate of the
epoxidation, suggesting increased selectivities with increasing ethene
partial pressures. Similar observations were made by Khasin [30]
as well as Cant & Hall [28]. Again contrary, Borman & Westerterp
found, that the selectivity to ethylene oxide decreases with increasing
ethene partial pressure [29].
There is also little common agreement about the influence of
the reaction products on selectivity and conversion degree, although
many authors (but not all) agree, that in absence of chlorine com-
pounds, carbon dioxide enhances selectivity and decreases the overall
reaction rate [31, 32]. In order to get a brief insight of the influence of
C2 H4 and O2 partial pressures on selectivity and conversion degree,
a kinetic model had to be adapted. Recently, a Dutch research group
performed extensive investigations [63, 64] and published a kinetic
model [29, 31], which was derived from experiments with a tube-type
(single pass) and a Berty / Bobo reactor (internal gas recirculation)
using a commercial ethylene oxide catalyst based on Ag/-Al2 O3 .
The authors tested four different kinetic models, generally assuming

a parallel reaction of ethene to ethylene oxide and carbon diox-


ide having the same mechanism and therefore, the same kinetic
model,
no consecutive combustion of ethylene oxide supporting the
plain parallel reaction scheme and
an adsorption of the reaction products ethylene oxide, water
and carbon dioxide on the surface of the catalyst.

Two different kinetic models in each two variations were tested,


3.2. Kinetics of the ethene epoxidation 29

based on an Eley-Rideal and a Langmuir-Hinshelwood mechanism.


In each kinetic model, one was formulated for atomic oxygen and
the other for molecular oxygen as the active component involved in
the rate determining step. All tests were performed in several reac-
tor types to exclude arbitrary reactor specific effects supporting one
model in favor of the other. It was found, that the kinetic equations
resulting from the Langmuir-Hinshelwood mechanism showed better
fitting quality than those derived from the Eley-Rideal one. Eventu-
ally, the authors decided to use an equation assuming the Langmuir-
Hinshelwood mechanism having molecular oxygen involved in the
rate determining step:
k KC2 H4 KO2 pC2 H4 pO2
r= P (3.1)
(1 + Ki pi )2

The authors provided computed constants for the selective oxi-


dation and total combustion of ethene in two reactor types as listed
in table 3.1. These constants will be used in the discussion to com-
pute reaction rates and from these a selectivity / conversion curve
for the described catalyst within both reactor types.

Table 3.1: Kinetic constants for the selective oxidation and to-
tal combustion of ethene for a Berty- and tubular-type reac-
tor [31]

Reactor type Berty tubular


Reaction type sel. ox. total comb. sel. ox. total comb.
kr 0.50 x 106 7.34 x 106 0.92 x 106 16.1 x 109
Tact 9.2 x 103 10.7 x 103 8.8 x 103 12.6 x 103
KC2 H4 0.0130 0.222 9.6 x 103 1.6 x 103
Tads,C2 H4 3.5 x 103 2.5 x 103 3.0 x 103 3.4 x 103
KO2 5.8 4.9 1.5 3.2
KC2 H4 O 10 93 - -
KCO2 101 55 27 96
KH2 O 55 14 50 43

The industrially most important parameter on selectivity and


conversion degree is the concentration of the reaction modifier chlo-
rine, which is used in form of a few ppm 1,2-dichloroethane (DCE) or
vinylchloride as moderator in order to increase selectivity. Therefore,
numerous studies and patents (e.g. [33, 39]) dealt with the influence
30 Chapter 3. Theory

of chlorine on the selectivity of the epoxidation or application of


chlorine in order to attain as high selectivities as possible. The se-
lectivity of pure silver powder or crystals in absence of promoting
agents is reported to be between 30% and 40% at low conversion
degrees [40]. Unmodified and unpromoted Ag/-Al2 O3 catalysts ex-
hibit higher selectivities between 40% and 60% [35, 41], depending
on reaction conditions. With increasing degree of chlorine coverage,
the selectivity increases to 75-85%, depending on the investigated
silver crystallite surface and reactor temperature [35, 36, 40, 42].
Actually, chlorine promoted industrial catalysts exhibit selectivities
of initially slightly better than 90% [9].
Ag/-Al2 O3 catalyst gain approximately 20% selectivity
in presence of DCE. Unfortunately, the TON2 and thus, the
degree of conversion decreases with increasing chlorine cover-
age [35, 42, 43]. It is believed, chlorine enhances the concentration
of an epoxidation-selective oxygen surface species, which is required
for the epoxidation and exists predominantly at high surface and
subsurface oxygen coverage [8]. Campbell and Paffett [43] showed,
an increasing chlorine coverage decreased the rate of dissociative
oxygen absorption. Furthermore, Tan, Grant and Lambert [44]
found an enhancement of oxygen diffusion into the bulk silver by
chlorine absorption.

3.3 Catalyst design


All modern industrial ethylene oxide catalysts are based on -Al2 O3 ,
having surface areas below 2 m2 /g. Such low surface areas are ob-
tained when alumina is fired at high temperatures for a long time.
Experiments from Vannice et. al. [41, 45] proved, that an -Al2 O3
support material specially prepared with a high surface area up to
50 m2 and impregnated with silver yields no ethylene oxide but only
carbon dioxide. Even when such a catalyst is placed behind a selec-
tive ethylene oxide catalyst, the ethylene oxide produced by the first
catalyst is completely oxidized to carbon dioxide and water by the
following high surface area catalyst. Therefore, high specific sur-
face areas are to be avoided, even if the support material itself is
2 Turn-Over-Number
3.3. Catalyst design 31

suitable. It is generally assumed, Lewis acid sites support the iso-


merization and consecutive combustion of ethylene oxide. Therefore,
many support materials are treated with HCl or alkali hydroxides /
halogenides to minimize the number of those sites before immobi-
lization of silver is performed [36, 46]. Sometimes up to 10% TiO2
is used as co-support material [47]. Early investigations revealed,
that silicon carbide and silica can be also used as support. -Al2 O3 ,
pumice, silica gel, carbon, magnesia and high surface supports in
general are not suitable [6].
Silver is a unique catalyst for the epoxidation of ethene [48].
In order to yield high selectivities, there are many promoters and
co-promoters in use, which have a strong effect on the selectivity as
soon as chlorine is involved. Caesium [49] and/or rubidium combined
with other alkali like barium are used in nearly every promoted cat-
alyst. An example for such a promoter / co-promoter combination is
Rhenium [50], which may be co-promoted with phosphorus and/or
boron [9, 37]. Niob and/or tantal, which is co-promoted by sulfur
in presence of alkali is another example [51]. Several patents and
publications such as [52, 53] describe a process to reactivate a silver
catalyst for ethene epoxidation by passing a Rb and/or Cs salt con-
taining solution at ambient temperatures through the reactor. It can
be assumed, that -Al2 O3 surface is re-passivated by alkali. Mina-
han [54] showed, the surface of a fresh (and selective) ethylene oxide
catalysts is nearly completely covered with silver, whereas uncovered
-Al2 O3 surface is observed on an aged (and less selective) catalyst
due to agglomeration of the silver particles, exposing the support
materials surface, effectively lowering the selectivity towards ethy-
lene oxide.
Another important point affecting strongly the selectivity and
activity of any catalyst is the effective size of the silver particles. A
patent filed already in 1975 used an Ag/polyacrylonitrile-complex
as precursor to deposit fine silver particles of about 150 nm on the
catalysts surface. After polymerization, the plastic support was de-
stroyed by calcination [55]. Another patent [56] claimed preferable
particle sizes of 50 to 400 nm. Investigations of Balzhiminaev et.
al. [57, 58, 59] showed, the rate of ethene epoxidation increased with
increasing silver particle size up to 60 nm (Fig. 3.1). Therefore, a
high dispersion of silver on the catalysts surface has to be avoided
in order to yield a selective and active catalyst. A REM picture
32 Chapter 3. Theory

18,0

17,5
lg rC H O / a.u.

17,0
4
2

Ag/Al 2O3
16,5 Ag/SiO 2
Ag powder

16,0
0 20 40 60 80 100 1000

particle diameter / nm

Figure 3.1: Reaction rate of ethylene oxide formation as a function of


the average silver particle size [57].

of a commercial ethylene oxide catalyst with silver particles on its


surface is depicted in figure 3.2.
Another important point affecting the selectivity and conver-
sion degree of a catalyst is the distribution of silver within the pel-
let. Several investigations [35, 36, 60, 61] showed, a non uniform
silver distribution within the catalyst pellet is preferable as soon as
the mass- and heat transfer effects in industrial catalyst pellets are
taken into account. The authors proved experimentally, that the po-
sition of a catalyst layer within a pellet has a strong influence on the
observed selectivity / conversion behavior [35]. This can be under-
stood only, when large temperature gradients within the pellet are
assumed and mass transfer through a catalytically inactive barrier
restricts the specific reaction rate and therefore, the heat production
lowering the hot spot temperature and enhancing selectivity.
It is known, that silver catalysts used for the partial oxidation of
ethene, require an activation procedure before reproducible results
are observed. In this activation period, selectivity and conversion
3.3. Catalyst design 33

Figure 3.2: REM picture of a commercial ethylene oxide catalyst (-


Al2 O3 ) having evenly distributed silver particles on its surface [5].

degree increase with increasing time on stream. Schouten et.al. [62]


reported activation times of about 3-4 days for a commercial ethylene
oxide catalyst before reproducible results were obtained.
34 Chapter 3. Theory

3.4 Reactor design and heat manage-


ment
This section deals with the reactor design and heat effects within
larger tubular reactors. All findings are results previously published
( [63, 64]), but some data points are plotted in a different way to
point out interesting temperature gradients and heat effects (Fig.
3.3, 3.4, 3.6, 3.8).

3.4.1 Reactor design in industrial plants


The partial oxidation of ethene is a highly exothermic reaction. The
exothermy of the total combustion is more than ten times higher
than that of the selective oxidation. Therefore, there is always a risk,
that local hot spots emerge, combusting precious ethylene. Thus,
this reaction requires a good heat management as already stated in
a patent by Law et al. in 1942 [33]. Generally, there are two types of
reactors suitable for highly exothermic reactions. The first reactor
type is the multi-tube, fixed-bed reactor. In this reactor type, the
catalyst is located in tubes having a length of several meters, but
only a few cm in diameter. The tubes are kept in a liquid medium
such as molten salts, metals, water or high boiling liquids to remove
the heat produced by the reaction. The second reactor type is the
fluidized bed or bubbling bed reactor. The advantage of this reactor
type is a very high heat transfer ability, but at the cost of a non uni-
form residence time distribution due to formation of bubbles in the
fluidized bed and high attrition of the catalyst. Today, all ethylene
oxide plants use multi-tube, fixed bed reactors, because the attrition
problem remained unsolved for this reaction [10].
In order to study heat effects in industrial multitube reactors,
the Dutch research group around Westerterp constructed a reactor
having a design similar to a single industrial reactor tube [62]. This
reactor consisted of a single tube, having a length of 1.19 m and
a diameter of 53 mm. The reactor was packed with an industrial
ethylene oxide catalyst and cooled with pressurized, boiling water
to remove the heat. In order to measure temperature profiles, the
reactor was equipped with 24 thermocouples to monitor local tem-
peratures in axial and radial direction.
3.4. Reactor design and heat management 35

In the following figures, the temperature data provided by the


Dutch group was used and transformed from 2D plots as published
to pseudo-3d plots in order to achieve a better overview of the tem-
perature distribution in the tube-reactor. Furthermore, some minor
calculations such as the adiabatic temperature rise and the overall
heat production of this reaction were performed in order to allow a
better understanding of the temperature effects in this reactor type.
The axial and radial temperature profile for a certain flow rate
and reactor temperature is depicted in figure 3.3. Despite the very
low wall temperature, a hot-spot of 40 K above the wall temperature
is located at about 1/3 of the catalysts bed height. It is obvious, that
the highest temperatures at any given axial position are measured
in the center of the reactor (r=0 mm).
The influence of the wall temperature on the axial tempera-
ture profile is depicted in figure 3.4. A hot spot can be observed
at 220 mm nearly in the middle of the catalysts bed. The temper-
ature of the hot spot depends on the adjusted wall temperature.
An increased wall temperature at a constant flow rate causes higher
degrees of conversion and therefore, the heat generated by the reac-

510

400
504
476
499
300
axial position / mm

488 482
493
temperature / K

493 488
200 504
499 482

100 476

471
476 471
0 465
0 5 10 15 20 25
radial position / mm

Figure 3.3: Axial and radial temperature profile of an industry-like re-


actor tube [62]. Reaction conditions: Twall = 466 K, c(C2 H4 )=0.715% in
air, p=0.5 MPa.
36 Chapter 3. Theory

475 50
42
470
33 42
465 17 25
wall temperature / K

33
460

temperature / K
455 8 25

450
17
445 8
8
440

435 0
0 100 200 300 400

axial position / mm

Figure 3.4: Axial reactor temperature rise (at r=0 mm) within an
industry-like reactor tube as a function of the adjusted wall tempera-
ture [62]. Reaction conditions: GHSV=0.82 kg/m2 s, c(C2 H4 )=0.5% in
air, p=0.5 MPa.

tion increases. This results in hot-spot temperatures of up to 50 K


above the adjusted wall temperature. The position of the hot spot
is obviously independent from the wall temperature.
The temperature rise, which would be observed in an ideal adi-
abatic reactor is calculated by the following equations:
P
4Tad = P (3.2)
nc cp,c
X
P= ni Xi Si 4HR,i (3.3)
P
P is the total thermal power of the reactor in Watt, nc cp,c is
the average heat capacity of the product stream consisting of the
components c and the index i denotes the involved reactions,
which produce the mentioned heat.
Therefore, the heat production by the reaction as well as the
heat removal by the hot effluent are both linear correlated with the
flow rate. Thus, the adiabatic temperature rise is independent of
the total amount of heat produced in the reactor, but depends on
3.4. Reactor design and heat management 37

the concentrations / partial pressures, the enthalpy, selectivity and


conversion degree of the involved reactions.
Assuming such an ideal adiabatic reactor, the corresponding
temperature rise and heat production was calculated for the pub-
lished selectivities and conversion degrees. The plot of the mea-
sured hot-spot temperature rise and the calculated adiabatic tem-
perature rise versus the wall temperature (Fig. 3.5) yields two par-
allel straight lines applying a logarithmic scale. This indicates a
constant proportion of those two parameters. The overall absolute
heat production of the reactor varies between 30 and 141 Watts,
corresponding to volume specific heat production of 26 to 125 kW
per m3 catalyst.
100
140

120

100

temperature rise / K
heat production / W

80
10

60

40

heat production
20 calculated adiabatic temperature rise
measured temperature rise
1
430 440 450 460 470 480

wall temperature / K

Figure 3.5: Heat production, calculated adiabatic temperature rise and


measured hot spot temperature rise of an industry-like reactor tube as
a function of the adjusted wall temperature [62]. Reaction conditions:
GHSV=0.82 kg/m2 s, c(C2 H4 )=0.5% in air, p=0.5 MPa.

The influence of the flow rate on the reactors axial tempera-


ture profile is depicted in figure 3.6. It can be clearly seen, that the
highest temperatures are obtained at low flow rates and therefore,
at high conversion degrees. With increasing flow rate, the hot spot
38 Chapter 3. Theory

505
250
475 500
470 475
200 495

temperature / K
490
flow rate / l/min

150 480
485

480
100

485 475
495
50 490 500 470

20 465
0 100 200 300 400

axial position / mm

Figure 3.6: Axial reactor temperature profile (at r=0 mm) of an


industry-like reactor tube as a function of the mass flow rate [62]. Reac-
tion conditions: Twall = 465K, c(C2 H4 )=0.5% in air, p=0.5 MPa.

temperature is lowered and shifted towards the middle of the reac-


tor. At low flow rates of 30 l/min, the hot spot is located at about
100-150 mm (1/3 of the beds length) with temperatures more than
35 K above the wall temperature, whereas the hot-spot is shifted
at higher flow rates of 200 l/min to 200-250 mm (1/2 of the beds
length), showing only 10-15 K temperature rise. Based on the selec-
tivities and conversion degrees, the adiabatic temperature rise and
the heat production was calculated (Fig. 3.7) and plotted as a func-
tion of the flow rate. With increasing flow rate, the heat production
of the reactor increases, whereas the conversion degree and therefore,
the adiabatic temperature rise decreases. It can be seen, that at high
flow rates and therefore, at high heat production rates, the measured
hot-spot temperature and the calculated adiabatic temperature rise
are the same. At lower flow rates, the measured hot-spot tempera-
ture is always lower than the calculated adiabatic temperature rise.

The last important parameter having strong influence on the


temperature profile is the ethene concentration. Therefore, the au-
thors changed this concentration, keeping the wall temperature and
3.4. Reactor design and heat management 39

120 70

110 60

100 50
heat production / W

(Tmax-TWall) / K
90 40

80 30

70 20

heat production
60 10
calculated adiabatic temperature rise
measured temperature rise
50 0
0 50 100 150 200 250

flow rate / l/min

Figure 3.7: Heat production, calculated adiabatic temperature rise and


measured hot spot temperature of an industry-like reactor tube as a func-
tion of the mass flow rate [62]. Reaction conditions: Twall = 465K,
c(C2 H4 )=0.5% in air, p=0.5 MPa.

mass flow rate constant. The measured axial temperature profile


in the middle of the reactor (r=0 mm) is depicted in figure 3.8.
The authors reported nearly constant degrees of conversion for all
ethene concentrations, indicating a reaction order for ethene close
to one. Therefore, the heat production increases with increasing
ethene concentration and thus, the hot-spot-temperature increases
with increasing ethene concentration too. Even at low ethene con-
centrations of only 0.9% in air, the hot spot is nearly 70 K above
the wall temperature.
In order to summarize the results, there are three important
rules for temperature profiles in this huge industry-like reactor as
long as the reactor is in a steady state mode of operation:

1. When the wall temperature is increased, the hot spot tem-


perature increases proportionally to the calculated adiabatic
temperature rise.
2. Increased flow rates lower the hot-spot temperature, although
the absolute heat production of the reactor is increasing.
40 Chapter 3. Theory

0,9 540
520

490 530
0,8
ethene concentration / vol%

510 520
0,7

temperature / K
500 510
0,6
500
0,5
490
0,4
480

0,3 480
470 470
470
0,2 460
0 100 200 300 400

axial position / mm

Figure 3.8: Axial reactor temperature profile (at r=0 mm) of an


industry-like reactor tube as a function of the ethene concentration [62].
Reaction conditions: Twall = 465K, GHSV= 0.82 kg/m2 s, balance air,
p=0.5 MPa.

Therefore, the hot-spot temperature is not correlated with


the absolute heat production rate.
3. Increased hydrocarbon concentrations lead to higher hot-spot
temperatures.

The highest absolute heat production of 163 Watt within this reactor
was obtained at a wall temperature of 465K, 22% conversion degree,
64% selectivity and a mass flow rate of 2.51 kg/m2 s. Therefore,
the highest volume specific heat production rate was 144 kW/m3
catalyst. At this point, the authors pointed out the reactor was
operated very close to a runaway.
3.4. Reactor design and heat management 41

3.4.2 Laboratory and microreactor design

In laboratory reactors, there are typically fewer problems with hot-


spots. This is mainly ascribed to the smaller dimensions of the
reactor, because the heat conductivity of the material is the same,
but the distance from the reactor axis to the wall is much shorter.
Furthermore, there is the possibility to dilute the catalyst with lots
of inert material in order to lower the number of active sites per vol-
ume and therefore, the heat production and with it the temperature
gradient. Thus, it is much easier to diminish heat gradients in small
laboratory type reactors than in larger pilot-plant sized tubes. The
standard reactor for laboratory use is the tube reactor, consisting of
a simple tube (steel or glass), which is heated by a suitable device.
In most applications, a resistance heating wire fitted directly to the
tube is applied, but other heat sources like indirectly heated high
boiling liquids or bubbling beds may be also applied. A disadvan-
tage of the reactor is still the low heat conductivity of the catalytic
active material, which guarantees temperature gradients if high spe-
cific heat production rates are applied and cannot be avoided.
A very interesting and promising reactor type for reactions re-
quiring high heat transfer is the concept of the microchannel reactor
(Fig. 3.9 and 3.10 ). In this reactor type, the catalytic active
coating is located on the walls of a most likely metallic micro heat

Figure 3.9: Scheme of a simple microchannel reactor. The walls of the


microchannels are covered with a catalytic active coating.
42 Chapter 3. Theory

Figure 3.10: Scheme of a cross flow microchannel heat exchanger with


alternating layers for high heat transfer.

exchanger. Therefore, microchannel reactors are wall reactors. The


surface area is still high enough to provide adequate catalytic active
surface areas (Fig. 3.11). Small channel dimensions in the range of
500 to 50 m result in surface areas of 47 to 470 cm2 /cm3 . With
a surface enlarging coating, the total catalytic active surface area
per volume is comparable to conventional reactor and designs, con-
taining common catalysts. Heat transfer coefficients of 18 to 54
kW/m2 K have been reported [65]. The heat transport is generally
described by the following equation:

dT
Q = A (3.4)
dx
Therefore, the higher the heat conductivity , the higher the heat
exchange area A and the lower the distance x, the lower is the result-
ing temperature gradient T for a given heat flux Q. A thin coating
with only a few microns will not cause a major heat transfer resis-
tance between the presumably ceramic like coating to the metal core
of the reactor. Furthermore, the metallic framework provides a su-
perb heat conductivity, which is typically one dimension higher than
that of a ceramic frame. For example, the heat transfer coefficient
of glass is in the range of to 1 W/mK, ceramic frames such as -
Al2 O3 exhibit 25 W/mK, whereas metals like steel, aluminum and
silver have heat transfer coefficients of 70 to 400 W/mK [66]. And
3.4. Reactor design and heat management 43

Figure 3.11: Calculated surface per volume ratio as a function of the


channel dimension of a microchannel reactor. Assumptions: square chan-
nels, wall thickness is 0.3 channel width.

finally, the low wall thickness helps to establish a low temperature


gradient. Therefore, the heat generated by the reaction is removed
very efficiently. Even when a single stack (Fig. 3.9) without al-
ternating reaction/ coolant layers (see Fig. 3.10, 3.12) is used, the
metal framework still provides a superb heat dissipation. Assuming
a volume specific heat production of 500 kW/m3 , a 1 cm x 1 cm
x 1 cm stack of microstructured aluminum wafers will produce 0.5
Watts of heat. This heat is removed through four outer walls of the
microstructured stack in equal shares and therefore, about 0.125 W
heat have to be removed through each wall. Assuming, the whole
heat of reaction has to be transported from the core of the reactor
to the wall (distance 0.5 cm), the frame is made of aluminum having
44 Chapter 3. Theory

Figure 3.12: Photographs of a microstructured reactor / heat exchanger


with and without tube fitting, developed and constructed by the Karl-
sruhe Research Center (a). SEM micrograph showing a corner view of
the reactor (b). The larger channels (140 x 200 m2 ) appear to be going
into the darker face of the reactor with the smaller channels (70 x 100
m2 ) running perpendicular [67].

a heat conductivity of 230 W/mK and 1/3 of the stacks area is


in contact with the next layer through its nose-piece, the estimated
worst case temperature gradient from the center of the stack to the
outer wall is calculated to 0.081K. Practically, this is negligible and
even much higher volume specific heat production rates do not cause
large temperature gradients.
A disadvantage of the high heat conductivity is a decreased
effectiveness of a microchannel reactor. Normally, the axial heat
conductivity of a i.e. a multitube heat exchanger can be neglected
and a temperature gradient between the inlet and outlet can be es-
3.4. Reactor design and heat management 45

tablished. This gradient is responsible for a high overall efficiency


of countercurrent operated heat exchangers - the temperature dif-
ference between the outlet of the process side and the inlet of the
coolant side is high. With their typically strong axial heat dissipa-
tion, metallic microchannel reactors act as a thermal shortcut and
eliminate the differences between concurrent and countercurrent op-
erated heat exchangers. Although this affects the economy of a
commercial heat exchanger, this effect helps to dissipate heat gener-
ated in potential local hot spots across the device. Computations of
Stief [68] proof this general consideration. He showed, the maximum
efficiency of 84.5% for a microstructured heat exchanger is achieved
at heat conductivities of 0.5 to 1 W/mK, which is typical for glass.
Copper as an excellent heat conductor allows the expected 50% ef-
ficiency. Stainless steel having a heat conductivity of 15 W/mK
allows efficiencies of approximately 65%.
Actually, microreactors3 and microchannel reactors are widely
spread in the analytical chemistry. A typical example is the meth-
anizer, which is used since the 1970ies in the analytical chemistry

Figure 3.13: (a) Reactor module for screening of catalysts, consisting of


35 stacked frames. The catalyst wafers are mounted and removed in the
directions of the arrow. (b) Microstructured catalyst wafers made of alu-
minum either by mechanical micromilling or (c) by wet etching (channel
radius: 130 m) [69].

for the total hydrogenation of carbon monoxide and carbon diox-


3 An extremely small reactor containing the catalytic active material in small

pellets, similar to a laboratory reactor


46 Chapter 3. Theory

ide (utilizing hydrogen from the carrier gas) in order to make them
visible in the sensitive flame ionization detector [70]. Another ex-
ample is the application of microreactors and microchannel reactors
in screening devices for rapid discovery and evaluation of catalysts.
In order to keep the effort for synthesizing huge numbers of catalyst
samples low, microchannel reactor have successfully been utilized for
suchlike systems (Fig. 3.13) [69, 71].
Another interesting feature of a microchannel reactor is its suit-
ability for handling explosive mixtures. Recently, the oxidation of
hydrogen was performed in a 1 cm3 microchannel reactor (Fig. 3.12)
providing a crossflow cooling and having a microchannel cross sec-
tion of 140 x 200 m2 [67]. In this reactor, a stoichiometric mix-
ture of undiluted H2 and O2 was converted neither with runaway
nor explosion. The authors reported, that the microchannels are
smaller than the quenching distance of hydrogen, which is reported
to be 1 mm in capillary tubes. Therefore, possible explosions are
quenched immediately in the narrow channels of the reactor.

3.4.3 Advantages of using microchannel reactors


Besides safety considerations, microchannel reactors may have re-
markable advantages in the following applications:

Screening devices: For most screening devices, material con-


sumption is a critical design issue. The lower the material con-
sumption and the higher the number of experiments, the better the
devices efficiency. Microreactors as well as microchannel reactors
with their low volume and sharp residence time distribution are very
suitable for screening of i.e. catalysts or active pharmaceutical inter-
mediates. Especially for the latter, starting materials are typically
limited and precious [72].

Catalytic investigations: The potential and likely hot spot for-


mation of tube type reactors make measurements of reaction kinetics
more complex than necessary. Having an exothermic and fast reac-
tion resulting in a temperature gradient across the catalysts bed,
any reaction rate is affected by the local temperature profile. Hav-
3.4. Reactor design and heat management 47

ing a practically isothermal reactor type available, the rates may


be measured without having temperature gradients to be taken into
account.

Continuous production of chemicals: Besides these former


two research applications, microchannel reactors may be used even
for continuous production of fine chemicals or pharmaceutical
compounds. Traditionally, these compounds are synthesized in
multi-purpose batch reactors.
Despite the small volume of the microchannel reactor, continu-
ous operation allows remarkable production capacities. For example,
having a 10wt% product mixture, synthesis of 1000 kg product re-
quires a throughput of approximately 10.000 kg. Assuming 8000 h
as time on stream per year and a density of 0.8 g/ml, the result-
ing flow rate is as low as 26 ml/min. Therefore, even small devices
may be used to synthesize continuously pharmaceutical components
or intermediates by making use of the advantages of microchannel
reactors.
Besides their small volume, the high heat transfer capacity of
microchannel reactors and their capability of continuous operation
may result in a pay off within a production process. This will be
illustrated in the following example:
Given a typical organic reaction, which requires 5 minutes for
95% completion. Furthermore, a starting material concentration of 2
moles per liter for each reagent will be assumed as well as 100 kJ/mol
heat release by the reaction and a heat capacity of 2.2 J/gK. Having
this data on hand, the adiabatic temperature rise is calculated to be
86 K.
Assuming, this model reaction follows a second order power law
kinetic, the rate constant k is calculated to be 0.033 l/(mol*s) (Fig.
3.14). Having the rate available, the initial heat of reaction can
be calculated. At the beginning of the reaction, 0.033 l/(mol*s) *
2 mol/l * 2 mol/l * 100 kJ/mol = 13.2 kW per liter are released,
which is 13.2 MW/m3 .
Standard batch vessels providing 100 l Volume (i.e. cylindrical
vessel having a diameter 40 cm and 80 cm height) have a surface to
volume ratio of approximately 6 to 12 m2 /m3 . With an area specific
48 Chapter 3. Theory

1,0

0,8
degree of conversion

0,6

0,4

0,2

0,0

0 1 2 3 4 5

time / min

Figure 3.14: Conversion degree as a function of time for a second order


reaction.r = k c2 , k=0.033 mol/(l*s), c0 =2 mol/l each.

heat exchange coefficient of up to 1000 W/m2 K, such a vessel has


a maximum volume specific heat transfer capacity of 12 kW/m3 K.
By using glass or vessels having less favorable heat exchanger con-
structions, the heat transfer capacity may be even much lower4 and
close to 1 kW/m3 K.
Therefore, the assumed reaction releases initially 1000 times
more more heat than a standard pilot plant sized vessel can remove.
As a consequence for the latter, the dosing speed of at least one
reagent has to be slowed down in order to keep thermal control
of this vessel. As a drawback, this prolonged dosing time spreads
the residence time distribution of a perfect batch vessel, allowing
potential side- and follow up reaction to take place and to affect
product quality.
Even a small 100 ml flask providing a surface to volume ra-
tio of approximately 120 m2 /m3 and a heat transfer coefficient of
4 Buchi glass reactor type CR101, volume 100l, 150..300 W/m2 K, heat ex-

change area 0.6 m2 [73]


3.4. Reactor design and heat management 49

200 W/m2 K would allow only 36 kW/m3 K. Thus, smaller lab-sized


equipment is not applicable for safe continuous production of fine
chemicals as soon as high reaction rates are expected and such a
reaction requires better heat transfer to ensure safe and trouble-free
operation.
Due to these advantages, highly exothermic chemistry with
short living intermediates may be performed safely in a microchan-
nel reactor and typically performed at much higher reaction tem-
peratures than in batch production. Some examples for handling
of short-living intermediates were already published [74, 75], taking
advantage of shortening the residence time for sensitive and reac-
tive intermediates. In a first example, C2 F5 I was converted to an
active intermediate by either MeMgCl or n-BuLi as metal M ,
avoiding an intramolecular elimination yielding C2 F4 and MF of the
intermediate.

In a second example, a Lithium-halide exchange was performed


using n-BuLi, forming a metal organic intermediate eventually
quenched with a nucleophilic reagent like DMF:

Currently, a patent for a similar two step synthesis for glycosides


by utilizing a high temperature lithiation (-10 to +20o C) in the first
step with n-BuLi or t-BuLi is requested by Bristol-Myers Squibb [77].
Normally, a suchlike synthesis is performed at -78o C and difficult to
scale up because of the highly exothermic reaction in combination
with a short living and sensitive intermediate.
50 Chapter 3. Theory

Microreactors may be even designed for handling of slurries /


suspensions as proven by Golbig [76]. Using several microreactors in
parallel, a throughput of 30t per year was obtained. A photo of this
pilot plant sized micro reaction system is depicted in figure 3.15.

Figure 3.15: Photo of a 30 t per year pilot plant system for pigment
synthesis [89].
Chapter 4
Results

In this chapter, results of the partial oxidation of ethene to ethene


oxide in microchannel reactors (MCR), modular microchannel reac-
tor (MMCR) and traditional tube type fixed bed reactors (FBR) are
presented.
The presentation is divided into four main parts. First, the cat-
alytic examination of potentially suitable silver coating methods for
microchannel reactors and closer catalytic examinations of selected
coatings were performed (chapter 4.1).
Second, similar investigations were performed with the same
type of catalyst/coating but utilizing a traditional fixed bed reactor
to point out differences between the two competing reactor concepts
(chapter 4.2)
Chapter 4.3 and 4.4 deal with reaction engineering issues of
reactor construction. Heat effects and temperature gradients in mi-
crochannel reactors were studied as well as the influence of reactor
construction on temperature distribution and catalytic properties
such as unwanted product combustion.

51
52 Chapter 4. Results

Table 4.1: Overview of the investigated microchannel & fixed bed re-
actors and their main purpose. MCR = MicroChannelReactor, MMCR
= Modular MicroChannelReactor

Denomination Purpose
MCR1 Benchmark reactor using bulk silver

MMCR1 Proof of principle reactor for sputtered Ag on Al


MMCR2 Tests up to 633 K for thermal runaway / aging
MMCR3 Impact of Ag layer thickness and its optimization
MMCR4 Test reactor for MCR3, Ag/Al coating
MMCR5 Test reactor for MCR2, Ag/-Al2 O3 /Al coating

MCR2 Ag/Al2 O3 /Al coating in a commercial reactor


MCR3 Ag/Al coating in a commercial reactor

MMCR6 Ag sputtered on -Al2 O3 prepared by ANOF


MMCR7 Ag sputtered on -Al2 O3 prepared by sol-gel
MMCR8 Ag impregnated on a sol-gel prepared -Al2 O3
MMCR9 Commercial ethylene oxide catalyst
immobilized in a microstructure
MMCR10 see MMCR9, but different reactor type

MMCR11 Ag sputtered on stainless steel

MMCR12 Test system for NO as gaseous promoter


MMCR13 MMCR12 promoted/ regenerated with Cs
MCR2Cs MCR2 promoted/ regenerated with Cs

MMCR14 Pair of test reactors for comparing two different


MMCR15 types of modular microchannel reactors

FBR1 a/b Fixed bed reactor, similar to MCR1 using two


different suppliers for silver foil
FBR2 Fixed bed reactor similar to MCR2
FBR3 Fixed bed reactor similar to MCR3
FBR4 Commercial ethylene oxide catalyst
in a fixed bed reactor
4.1. Epoxidation of ethene in microchannel reactors 53

4.1 Epoxidation of ethene in microchan-


nel reactors

In this section, experimental results of potential suitable catalytic


active silver coatings in the partial oxidation of ethene are presented.
In general, the selectivity / conversion behavior was monitored by
varying the residence time. Furthermore, the influence of the oxygen
and the ethene partial pressure as well as the impact of reactor
temperature and operating pressure on selectivity and conversion
was investigated at different reactor temperatures.
Standard investigations were performed in absence of purposely
added promoters such as NOx and Cs to avoid different impacts on
different types of catalytic active coatings. In some cases, investiga-
tions were performed also in presence of these promoters to identify
the impact on the catalytic properties.
A disadvantage of microchannel reactors is the time consuming
immobilization of suitable catalytic active species on the walls of the
channels. Therefore, suitable catalytic active coatings and coating
methods had to be developed.
The easiest way having silver as surface material of microchan-
nels is to manufacture these structures directly from silver sheets.
Therefore, the whole microstructure is made of silver and there is
no need to develop a coating method. Such a reactor will be used
as a benchmark for all following reactor types and coatings.

4.1.1 MCR1: Bulk silver microchannel reactor

The first microchannel reactor, denominated as MCR1, is made of


microstructured silver wafers having a purity of 99.97%. Its chan-
nel size was chosen as 200 x 200 m2 with a channel length of
50 mm. Each wafer (width 10 mm) has 33 channels and 26 wafers
are mounted in one stack forming the microchannel reactor. A more
detailed reactor description including a photo of the assembled re-
actor is given in the experimental section of this paper. The most
important geometric parameters are listed in table 4.2.
54 Chapter 4. Results

Table 4.2: Geometric parameters of the bulk-silver made mi-


crochannel reactor MCR1.

Basic reactor type FZK made MCR [90]


Channel width 200 m
Channel height 200 m
Channel length 50 mm
Number of channels per wafer 33
Number of wafers 26
Wafer height 300 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 343 cm2
Total channel volume 1.72 cm3
Total stack volume 3.9 cm3

Catalytic activation: The activation of the bulk-silver catalyst


was performed under reaction conditions. A mixture of 20% ethy-
lene, 20% oxygen with methane as balance was used, applying a
residence time of 1.1 s referring to flow rates measured at standard
temperature and pressure. The residence time was calculated ac-
cording to the following equation:

Vreactor,geometric
= (4.1)
VSTP

The geometric reactor volume Vgeometric was calculated from the


total channel volume without taking the volume of the diffusers into
account.
A reactor temperature of 523 K was chosen, because initially
small amounts of carbon dioxide as an indicator for catalytic activity
were observed. Higher reactor temperatures and high degrees of
conversion were avoided during the activation period in order to
prevent rapid aging of the catalyst. The dependence of selectivity
and conversion degree on the time on stream (TOS) is depicted in
figure 4.1. Initially, selectivities to ethene oxide close to 70% could
be observed. No other organic byproducts such as acetaldehyde were
found. After a time on stream of one day at 523 K a selectivity of
little more than 62% at 1.3% conversion degree was obtained.
4.1. Epoxidation of ethene in microchannel reactors 55

0,016 0,8

0,014 0,7

selectivity to ethene oxide


degree of conversion

0,012 0,6

0,010 0,5

0,008 0,4

0,006 0,3

0,004 0,2

0,002 0,1

0,000 0,0
0 5 10 15 20 25 30 35

time on stream / h

Figure 4.1: Selectivity to ethylene oxide and conversion degree of ethene


for the microchannel reactor MCR1 as a function of the initial time on
stream. Reaction conditions: 20% C2 H4 , 20% O2 , 60% CH4 , p= 0.3 MPa,
T= 523 K

Influence of the oxygen concentration: The influence of the


oxygen concentration on selectivity and conversion degree was inves-
tigated applying a constant concentration of 20% ethene and oxygen
concentrations varying between 6.6% and 80% (Fig. 4.2), keeping
the residence time constant. Methane was used as balance. At
low reactor temperatures of 503 K and therefore, low degrees of
conversion, the selectivity was nearly independent from the oxygen
concentration, varying only slightly between 63.7 and 62.1%. In con-
trast, the conversion degree was strong correlated with the oxygen
concentration. The higher the oxygen concentration, the higher the
conversion degree. At low oxygen concentrations of 6.6%, the con-
version degree was 0.4%, whereas 80% oxygen in the feed improved
the conversion degree nearly by factor 10 up to 4.0% at 503 K. With
increasing reactor temperature and therefore, higher degrees of con-
version, the selectivity became correlated with the oxygen concen-
tration. At a reactor temperature of 543 K and low oxygen concen-
trations of 6.1%, the selectivity was 50.6% at a conversion degree of
1.5%, whereas an increased oxygen concentration of 80% improved
the selectivity to 55.9% and the degree of conversion to 11.6%.
56 Chapter 4. Results

0,65

selectivity to ethylene oxide


503K

0,60 523K

543K
0,55

0,50
degree of conversion

543K
0,10

523K
0,05
503K

0,00
0 10 20 30 40 50 60 70 80 90
oxygen concentration / %

Figure 4.2: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the oxygen concentration for the microchannel reactor
MCR1. Reaction conditions: 20% C2 H4 , balance CH4 , =1.1 s (STP),
p= 0.3 MPa, T= 503 K / 523 K / 543 K.

Influence of the ethene concentration: In order to investi-


gate the influence of the ethene concentration, a mixture of ethylene
with oxygen and no inert component was passed through the re-
actor at two different, but in each experiment constant residence
times (Fig. 4.3). In the first experiment, the ethene concentra-
tion was varied between 0.63% and 28% using a residence time of
1.1s. At low ethene concentrations such as 0.63%, the conversion
degree was 47%, whereas at higher ethene concentrations, the con-
version degree declined down to 3.3% at 28% ethene. The selectiv-
ity depended slightly on the ethene concentrations under the chosen
reaction conditions. At low ethene concentration of 0.63% and a
corresponding high conversion degree of nearly 50%, the selectivity
was 51.7%. At higher ethene concentrations and lower conversion
degrees, the selectivity increased to 58.7%.
Having a steady rise in selectivity, it had to be assumed, the se-
lectivity was still more increasing. Due to experimental limitations
4.1. Epoxidation of ethene in microchannel reactors 57

0,62

0,5

0,60

selectivity to ethylene oxide


0,4
degree of conversion

0,58

0,3

0,56

0,2 =1.1s =3.4s


0,54

0,1

0,52

0,0

0 10 20 30 40 50 60 70

C H concentration / %
2 4

Figure 4.3: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the ethene concentration for the microchannel reactor
MCR1. Reaction conditions: balance O2 , p= 0.3 MPa, T= 503 K.

in the available range of massflow controllers, the residence time


had to be increased. In the second experiment, the residence time
was prolonged to 3.4 s. With access to ethene concentrations be-
tween 31% and 63%, the selectivity was still increasing from 57.6%
to 61.4%, respectively. The corresponding degree of conversion de-
clined from 5.9% to 1.56%.

Selectivity / conversion behavior: The selectivity and conver-


sion behavior of this reactor was investigated for ethene concentra-
tions of 4% and 20% in oxygen and at different reactor temperatures
by variations of the residence time in order to attain different de-
grees of conversion. The results of those experiments are depicted
in figure 4.4 and 4.5.
At 4% ethene in oxygen and temperatures of 503 K, a high
selectivity to ethylene oxide (EO) of 59.6% at a conversion degree
of 15.2% was observed (Fig. 4.4). With decreasing flow rate and
therefore, increasing degree of conversion, the selectivity decreased
slightly to 57.0% at conversion degrees of 37.4%. At higher reactor
58 Chapter 4. Results

0,60
503K
0,58 523K
543K
selectivity to ethylene oxide

0,56

0,54

0,52

0,50

0,48

0,46

0,15 0,20 0,25 0,30 0,35 0,40 0,45 0,50 0,55 0,60

degree of conversion

Figure 4.4: Selectivity to ethylene oxide for the microchannel reactor


MCR1 as a function of the conversion degree, adjusted by residence time
variation. Reaction conditions: 4% C2 H4 in O2 , p= 0.3 MPa, T= 503 K,
523 K, 543 K

temperatures, higher degrees of conversion and lower selectivities


were observed. At 543 K, the initial selectivity was 55.2% at a
conversion degree of 31.8%. With an increased conversion degree up
to 58.4%, the selectivity decreased to 49.5%. Therefore, three nearly
parallel curves were obtained with the highest selectivities observed
for the lowest reactor temperature.
At 20% ethene in oxygen (Fig. 4.5), a different selectivity/ con-
version behavior was observed. At a reactor temperature of 503 K,
initially 62.0% selectivity at a conversion degree of 4.1% were ob-
served. With increasing degree of conversion, the selectivity dropped
down to 53.7% at 14.5% conversion degree. Surprisingly, the in-
creased reactor temperature did not result in a parallel curve as
observed for 4% C2 H4 in O2 . At low conversion degrees of 6.0%,
lower selectivities of 59.6% were observed. With increasing degree
of conversion, the selectivity decreased to 49.2% at a conversion
degree of 23.8%. At conversion degrees higher than 10%, the selec-
tivities at 523 K were surprisingly higher than at 503 K. The same
4.1. Epoxidation of ethene in microchannel reactors 59

0,64
0,62 503K
0,60 523K
543K
selectivity to ethylene oxide

0,58
0,56
0,54
0,52
0,50
0,48
0,46
0,44
0,42
0,40
0,05 0,10 0,15 0,20 0,25 0,30 0,35

degree of conversion

Figure 4.5: Selectivity to ethylene oxide for the microchannel reactor


MCR1 as a function of the conversion degree, adjusted by residence time
variation. Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa, T= 503 K,
523 K, 543 K.

behavior was observed for a reactor temperature of 543 K. At con-


version degrees above 20%, the selectivity at 543 K was higher than
at 523 K.
In all those experiments, no temperature gradient caused by the
reaction was observed, even when the flow rate was increased to 1100
ml/min by factor 10 to produce more heat. In order to evaluate the
heat production in this microchannel reactor, similar calculations to
those for the industrial type reactor (see page 36) were made and
additional experiments at higher flow rates performed. The results
for the selectivity, adiabatic temperature rise and heat production as
a function of the conversion degree are depicted in Fig. 4.6. It was
observed, that there was a small initial increase in the selectivity at
very high flow rates and therefore, at low degrees of conversion. At
a conversion degree of 2.3%, the initial selectivity was 53.2%, which
increased slightly to a maximum of 57.2% at a conversion degree
of 8.3%, after which the selectivity decreased steadily to 42% at
30.7% conversion degree. The latter behavior was already observed
60 Chapter 4. Results

/ K
1400

ad
heat production / W

T
1200

adiabatic temp. rise


2 1000

800

600
1
400

200

0 0
selectivity to ethylene oxide

0,65

0,60

0,55

0,50

0,45

0,40
0,00 0,05 0,10 0,15 0,20 0,25 0,30 0,35

degree of conversion

Figure 4.6: Selectivity to ethylene oxide, heat production and adiabatic


temperature rise as a function of the degree of conversion (attained by
residence time variation) for the microchannel reactor MCR1. Reaction
conditions: 20% C2 H4 in O2 , p= 0.3 MPa, T= 543 K.

in former experiments (see fig. 4.5).


The heat production rate is correlated with the flow rate. With
the flow rate increasing faster than the conversion degree decreas-
ing, the heat production within the reactor by the epoxidation and
total oxidation was the highest at the highest flow rate and thus,
the lowest degree of conversion. At a conversion degree of 2.3%, the
heat production was 2.43 W, decreasing to 0.28W at a conversion
degree of 30.7%. Contrary, the adiabatic temperature rise Tad in-
creased with increasing degree of conversion. At the highest heat
production of 2.43 W, Tad was 77 K, whereas at 0.28 W and 30.7%
conversion degree, an adiabatic temperature rise 4Tad of 1350 K
was calculated.

Influence of the total pressure: The influence of the reactors


total pressure on selectivity and conversion degree, applying 4%
ethene in oxygen, is depicted in figure 4.7. The degree of conversion
is increasing with increasing total pressure and therefore, with in-
4.1. Epoxidation of ethene in microchannel reactors 61

residence time / s (ST)

0 5 10 15 20

selectivity to ethylene oxide


0,60

0,58

0,56

0,54

0,30
degree of conversion

0,25

0,20

0,15
0,0 0,5 1,0 1,5 2,0

pressure / MPa

Figure 4.7: Selectivity to ethylene oxide and degree of conversion as a


function of the total pressure / residence time (calc. based on standard
temperature) for the microchannel reactor MCR1. Reaction conditions:
4% C2 H4 in O2 , STP = 1.1 s, T= 523 K.

creasing hydrodynamic residence time. At 0.1 MPa, the conversion


degree was about 18% and increased to 30.6% at a pressure of 2 MPa.
The influence of the total pressure on the selectivity showed initially
increasing selectivities. At 0.1 MPa, the selectivity was 53.7% and
increased to 59.5% at a pressure of 0.9 MPa. Higher pressures of
1.5 to 2 MPa resulted in slightly lower selectivities of 59.1% and
58.1%, respectively. It can be clearly seen, that conversion as well
as selectivity show no direct correlation with the with the residence
time. Raising the pressure by factor 20 increases the conversion by
factor 1.7.

Influence of the reactor temperature: The influence of the


reactor temperature on the selectivity and conversion degree was
investigated at low degrees of conversion (<10%) and thus, at high
flow rates in order to allow a calculation of the reaction rates accord-
ing to the kinetic model of an ideal differential reactor. As shown
62 Chapter 4. Results

selectivity to ethylene oxide


0,64

0,62

0,60

0,58

0,56
degree of conversion

0,10

0,05

0,00
480 490 500 510 520 530 540

reactor temperature / K

Figure 4.8: Selectivity to ethylene oxide and degree of conversion as a


function of the reactor temperature for the microchannel reactor MCR1.
Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa, =1.0s.

in figure 4.8, the selectivity seemed to decrease linear with increas-


ing reactor temperature, indicating a selectivity loss of roughly 1.4
percent per 10 K increase in reactor temperature, whereas the de-
gree of conversion seemed to increase exponentially as expected by
an Arrhenius equation. At low reactor temperatures, the conversion
degree doubles each 20 K increase in reactor temperature.

4.1.2 Silver supported on aluminum wafers:


MMCR1-5, MCR2, MCR3

4.1.2.1 Test of Ag/Al as a suitable, catalytic active coating


(MMCR1)

In order to evaluate the suitability of silver coated aluminum wafers,


the modular microchannel reactor MMCR11 was used with 20 sim-
1 see construction details for the modular microchannel reactors (MMCR)

type I and type II on page 194


4.1. Epoxidation of ethene in microchannel reactors 63

Table 4.3: Geometric parameters of the modular microchannel reactor


MMCR1, having Ag/Al microstructured wafers as catalyst.

Basic reactor type MMCR type I


Channel width 200 m
Channel height 200 m
Channel length 50 mm
Number of channels per wafer 33
Number of wafers 20
Wafer height 300 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 264 cm2
Total channel volume 1.32 cm3

Ag coating thickness 200 nm (each side)


Coating method PVD, once, perpendicular

ple microstructured AlMg3 wafers2 (3% Mg alloyed in Al), which


were coated once with 200 nm silver deposited by physical vapor de-
position. The most important geometric parameters of this reactor
are listed in table 4.3.
After an initial activation, which was performed in the same way
as for the bulk-silver microchannel reactor MCR1 (see page 54), the
selectivity and conversion degree as a function of the residence time
were monitored for ethene concentrations of 20% in oxygen. The
results are depicted in figure 4.9.
The conversion degree increased with increasing residence time
from 16.4% at 0.8s to 41.7% at a residence time of 5.6s. The selec-
tivity initially increased from 53.0% (0.8s) to a maximum of 59.9%
(3.4s). Higher residence times lead to slightly decreased selectiv-
ities. Again, the adiabatic temperature rise was calculated. At
16.4% conversion degree and a selectivity to EO of 53.0%, the adia-
batic temperature rise was 598 K. With increasing degree of conver-
sion, the adiabatic temperature rise increased to 1400 K (X=41.7%,
S=57.5%).
Surprisingly, small amounts of acetaldehyde besides the main
2 supplied by Forschungszentrum Karlsruhe
64 Chapter 4. Results

0,45 0,65

0,40

selectivity to ethene oxide


degree of conversion

0,35 0,60

0,30

0,25 0,55

0,20

0,15 0,50

0 2 4 6

residence time / s (STP)

Figure 4.9: Selectivity to ethylene oxide and degree of conversion as a


function of the residence time (STP) for the modular microchannel reactor
MMCR1. Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa, T= 523 K.

210
ratio EO / acetaldehyde
concentration

200

190

180

0,004
selectivity to acetaldehyde
conc. / %

0,025

0,020
acetaldehyde

0,015 0,002

0,010

0,005

0,000 0,000
0 1 2 3 4 5 6

/ s (STP)

Figure 4.10: Concentration of acetaldehyde and ratio of ethylene oxide


to acetaldehyde for the modular microchannel reactor MMCR1 as a func-
tion of the residence time (STP). Reaction conditions: 20% C2 H4 in O2 ,
p= 0.3 MPa, T= 523 K.
4.1. Epoxidation of ethene in microchannel reactors 65

products EO and carbon dioxide/ water were observed in this re-


actor. The concentration of acetaldehyde and the ratio of the EO
to acetaldehyde concentration as a function of the residence time is
depicted in figure 4.10.
The ratio of both concentrations was used in order to have a di-
rect comparison of the reaction rates of both products, the resulting
selectivity to acetaldehyde was approximately 0.3%. The concentra-
tion of acetaldehyde increased with increasing residence time from
0.009% at 0.8 s to 0.027% at 5.6 s. The ratio of EO to acetaldehyde
concentration varied only slightly between 177 and 208.
After the catalytic experiments in this modular microchannel
reactor had been finished, the silver coated aluminum wafers were
removed from the reactor. The color of the wafers attracted the
attention, because the initially shiny metallic surface of the silver
coated aluminum wafers changed to grayish-white. Furthermore,
the silver wafers were sticking firmly together. In order to examine
the change of the surface, two REM pictures were made. The first
picture (Fig. 4.11a) shows the surface of the unused silver coated
aluminum wafer, looking onto the bottom of a channel. Apparently,
the surface is unstructured with only some dust or splinters, which
had been coated with silver during the coating process. The surface
of a channel bottom after the catalytic experiments were performed
is depicted in the second picture (Fig. 4.11b). It is obvious, that the
surface topography changed dramatically, obviously roughened and
agglomerates were formed. The average diameter of those particles
is approximately 800 nm.

4.1.2.2 Short term aging of an Ag/Al activated mi-


crochannel reactor (MMCR2)

In order to examine the thermal stability of this Ag/Al microchan-


nel reactor at high degrees of conversion and to get an insight into
aging effects of this catalytic system, another microstructured wafer
stack was prepared and mounted in the modular microchannel re-
actor type I. This time, the wafers were coated with a thicker silver
layer. The silver coating was performed by PVD for two times at an
angle of plus and minus 45o perpendicular to the channel structure in
order to achieve a better coating quality within the rectangular mi-
66 Chapter 4. Results

Figure 4.11: REM surface picture of the 200 nm silver coated (PVD)
modular microchannel reactor MMCR1 before (a) and after (b) use.

crochannel structure. The geometric parameters of this microchan-


nel reactor MMCR2 are listed in table 4.4. The reactor MMCR2 was
activated using the default activation method at a reactor tempera-
ture of 503 K as already described for MCR1 and MMCR1. Directly
after the activation, the following reference point of the fresh silver
coating was determined, applying a reactor temperature of 503 K,
20% ethene in oxygen and a residence time of 1.05 s:

degree of conversion: 13.2% +/-0.2% (503 K)


4.1. Epoxidation of ethene in microchannel reactors 67

Table 4.4: Geometric parameters of the modular microchannel reactor


MMCR2 having an Ag/Al microstructure as catalyst.

Basic reactor type MMCR type I


Channel width 200 m
Channel height 200 m
Channel length 30 mm
Number of channels per wafer 33
Number of wafers 25
Wafer height 300 m
Wafer width 10 mm
Wafer length 30 mm
Total geometric surface area 198 cm2
Total channel volume 0.99 cm3
Total stack volume 2.25 cm3

Ag coating thickness 400 nm


Coating method PVD, struct. side 2 times
perpendicular +/-45o ,
unstruct. side coated once

selectivity to EO: 69.7% +/- 0.5%

constant results for 12h

After taking this reference point, the selectivity / conversion behav-


ior of the fresh catalyst was determined in subsequently performed
experiments by a variation of the flow rate and increasing the max-
imum degree of conversion by lowering the total flow rate and thus,
prolonging the residence time. The results are shown in figure 4.12.
In the first experiment, the selectivity was initially 72.7% at a con-
version degree of 7.5% when a residence time of 0.53 s was applied.
With subsequently increasing residence time, the conversion degree
increased to 26.9%, yielding a lower selectivity of 69.9% when apply-
ing 5.0 s residence time. In the next experiment, applying the same
reaction conditions (24h later), the initial selectivity was only 69.9%
compared to 72.2% at a residence time of 0.53 s, yielding a slightly
higher degree of conversion of 8.8%. Again, with increasing residence
time, the conversion degree raised to 29.1%, showing a selectivity of
67.2% at a residence time of 5.0s. The S/X curve was shifted nearly
parallel to lower selectivities and slightly higher degrees of conver-
68 Chapter 4. Results

first run

second run
0,72
selectivity to ethylene oxide

0,70

0,68

0,05 0,10 0,15 0,20 0,25 0,30

degree of conversion

Figure 4.12: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the microchannel reac-
tor MMCR2 at subsequently performed variations of the residence time.
Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa, T= 503 K.

sion after the first run. Each experiment took about 12 hours. Small
amounts of acetaldehyde were detected only at high conversion de-
grees and the concentration ratio of EO to acetaldehyde was close
to 600, resulting in selectivities close to 0.1%.
Internal repeatability test: Nearly any experiment in the
present work was performed such, that the data point with the high-
est flow rate was repeated at the end of each run. Thus, the run was
started with the highest available flow rate and the flow rate step-
wise lowered until the lowest flow rate was reached. Eventually, the
first data point at the highest flow rate was adjusted again to check
for a potential deviation by a not fully stabilized / aged catalyst.
If this check for repeatability was unsuccessful, the whole run was
performed again. Each run took approximately 12 hours, therefore
this procedure ensured a short term repeatability.
4.1. Epoxidation of ethene in microchannel reactors 69

4.1.2.3 Thermal stability of an Ag/Al activated mi-


crochannel reactor (MMCR2)

In following experiments, this reactor (MMCR2) was heated up from


between 463 K in steps to 633 K and cooled down back, again step-
wise to the starting temperature of 463 K. The flow rate was kept
constant, a 20% ethene in oxygen mixture was applied and the to-
tal pressure 0.1 MPa. The selectivity and corresponding degree of
conversion were constantly monitored.
This experiment should answer two important questions. The
first and most important question deals with the thermal stability
of this reactor. With a constant reagent flow and increasing reactor
temperature, the conversion degree is expected to increase whereas
the selectivity decreases, causing large heat effects within the explo-
sion range of the ethylene/ oxygen mixture. The second question is
about the thermal stability of that coating. With increasing reactor
temperature and steadily increasing amounts of water (formed by the
total oxidation), the coating is chemically and thermally stressed. If
the selectivity at the beginning and at the end is still the same, this
would prove the usefulness of this catalytic coating.
Unfortunately, it proved practically impossible to perform the
experiments at elevated pressures of 0.3 MPa, because after passing
the dew-point of the by-product water, the pressure controller was
plugged by condensate and therefore, the experiment failed. Due
to the technical specifications of the pressure controller, it was im-
possible to heat the pressure controller to temperatures higher than
333 K to avoid this condensation problem. Therefore, only pressures
of 0.1 MPa were applied in this experiment.
The results of those temperature stress tests are depicted in
figure 4.13. The experiment with increasing reactor temperature is
marked with a solid line and open, up-pointing triangles, whereas
the experiment with decreasing reactor temperatures is marked with
a dashed and a solid triangle, pointing down. At the beginning and
therefore, at low reactor temperatures of 463 K, the selectivity was
69.5% at a conversion degree of 3.8%. Initially, the selectivity de-
creased with increasing reactor temperature. In the middle of the
investigated temperature range between 523 K and 583 K, the se-
lectivity was quite constant and even slightly increasing from 55.2%
70 Chapter 4. Results

selectivity to ethylene oxide


increasing reactor temperature
0,7
decreasing reactor temperature

0,6

0,5

0,8
degree of conversion

0,6

0,4

0,2

0,0

440 460 480 500 520 540 560 580 600 620 640

reactor temperature / K

Figure 4.13: Selectivity to ethylene oxide and degree of conversion as a


function of the reactor temperature for the microchannel reactor MMCR2.
Reaction conditions: 20% C2 H4 in O2 , p= 0.1 MPa, =1.6 s.

to 56.3%, despite the increasing degree of conversion, which raised


from 20.4% to 47.4%. At this point, the selectivity decreased again
with increasing temperature. At the highest applied reactor tem-
perature of 633 K, the selectivity was down to 47.1% at a conversion
degree of 75.0%. At this point, the adiabatic temperature rise is
calculated to be 3065 K and the total heat production rate is esti-
mated to be 9.2 Watt. A further increase of the conversion degree by
increasing the reactor temperature was impossible due to technical
limitations by the reactors heating device. With a stepwise cooling
down, the selectivity and conversion degree was monitored again.
This time, slightly lower selectivities were observed. In the temper-
ature range of 523-583 K, the selectivity dropped by approximately
4% after exposition to high temperatures. At even lower tempera-
tures, the selectivity difference between both experiments increased
to 5.1-6.7%. Smaller differences between both experiments were ob-
served regarding the degree of conversion. As shown in figure 4.13,
the conversion degree is despite a minor deviation at 583 K very
much the same.
4.1. Epoxidation of ethene in microchannel reactors 71

ratio EO/acetaldehyde
600

concentration
400

200

0
conc. of acetaldehyde / %

0,08

0,06

0,04

0,02

0,00

440 460 480 500 520 540 560 580 600 620 640

T/K

Figure 4.14: Ratio of ethylene oxide to acetaldehyde concentration and


absolute acetaldehyde concentration as a function of the reactor temper-
ature for the microchannel reactor MMCR2. Reaction conditions: 20%
C2 H4 in O2 , p= 0.1 MPa, =1.6 s.

In these experiments, the dependence of the acetaldehyde for-


mation as a function of the reactor temperature was monitored. Re-
sults are shown in figure 4.14. At reactor temperatures below 523 K,
no acetaldehyde formation was observed. At higher temperatures,
the acetaldehyde formation increased seemingly exponentially, yield-
ing 0.082% acetaldehyde at a reactor temperature of 633 K. The
ratio of EO to acetaldehyde decreased with increasing temperature.
At temperatures of 523 to 543 K, the ratio was close to 500 and
decreased steadily to approximately 100 at 633 K.
After this high temperature treatment at high degrees of con-
version, the selectivity / conversion behavior was monitored again.
The results of this experiment are depicted in figure 4.15, using
closed pentagons. Datapoints from the first and second run of this
reactor are plotted for comparison purposes (see Fig. 4.12). Com-
pared to the results obtained for the second run (open squares), the
selectivity dropped down from 69.7% to 62.1% at low degrees of
conversion, whereas the degree of conversion increased from 8.8% to
72 Chapter 4. Results

0,72

selectivity to ethylene oxide

0,70

0,68

0,66

0,64
first run

second run

after high
0,62
temperature exp.

0,05 0,10 0,15 0,20 0,25 0,30 0,35

degree of conversion

Figure 4.15: Selectivity to ethylene oxide as a function of the conversion


degree attained by residence time variation for the Ag/Al microchannel
reactor MMCR2 before and after a high temperature treatment. Reaction
conditions: 20% C2 H4 in O2 , p= 0.3 MPa, T= 503 K.

10.2% at residence times of 0.53s. Contrary to the trend of the first


two experiments, the selectivity increased with increasing degree of
conversion to a maximum of 64.8-65.0% at conversion degrees above
25%. At a residence time of 5.0s, the conversion degree increased
from 29.1% before to 33.1% after the high temperature experiments.
The amount of acetaldehyde did not change with the treatment and
only small, compared to the fresh catalyst unchanged amounts of
acetaldehyde were detected at very high degrees of conversion.

4.1.2.4 Influence of the Ag layer thickness on selectivity


and conversion degree (MMCR3)

It can be easily derived from the different results of the bulk-silver


microchannel MCR1 on the one hand and the silver coated aluminum
microchannel reactors MMCR1 and MMCR2 on the other, that both
catalytic systems have different properties. In order to investigate
the effect of the silver layer thickness on the selectivity / conver-
4.1. Epoxidation of ethene in microchannel reactors 73

sion behavior of the microchannel reactor, a modular microchan-


nel reactor was equipped with three microstructured AlMg3 wafers,
which have been coated on the structured site with 50 nm Ag by
sputtering. The geometric parameters of this reactor are listed in
table 4.5. The activation of this catalyst was performed under reac-
tion conditions as long as the selectivity / conversion changed with
time on stream. After typically 24 to 48h and a constant level of
activity, the flow rate was decreased stepwise in order to monitor
the selectivity / conversion behavior of this reactor. Afterward, the
catalyst was removed from the modular microchannel reactor and
coated with an additional amount of silver. Thus, a dense silver
layer was added to an already activated catalyst and the proce-
dure, including the initial activation, was performed again to track
changes. The selectivity / conversion behavior of the Ag/Al reactor
MMCR3 obtained by variation of the residence time is depicted in
figure 4.16. It is noteworthy that at a very low layer thickness of
about 50 nm and conversion degrees as low as 5%, the selectivity
dropped down sharply with increasing degree of conversion. With
an increasing layer thickness, much higher degrees of conversion up
to 40% were attained. The maximum selectivity for each catalyst

Table 4.5: Geometric parameters of the modular microchannel reactor


MMCR3 having an Ag/Al microstructure as catalyst

Basic reactor type MMCR type I


Channel width 300 m
Channel height 700 m
Channel length 50 mm
Number of channels per wafer 14
Number of wafers 3
Wafer height 1000 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 42 cm2
Total channel volume 0.29 cm3
Total stack volume 1.5 cm3

Ag coating thickness 50, 400, 800, 1400 nm


Coating method sputtering on
structured side
74 Chapter 4. Results

0,60

0,55

0,50
selectivity to ethylene oxide

0,45

0,40

0,35

0,30

0,25 Ag layer thickness


50 nm
0,20
400 nm
0,15 800 nm
1400 nm
0,10
0,0 0,1 0,2 0,3 0,4

degree of conversion

Figure 4.16: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/Al microchannel
reactor MMCR3 for different Ag layers. Reaction conditions: 20% C2 H4
in O2 , p= 0.3 MPa, T= 523 K.

increased slightly with increasing layer thickness, the overall highest


selectivity was provided by the thickest silver layer. Selectivities of
initially 56% observed at 50 nm Ag and low degrees of conversion
may be ascribed to a not sufficient initial activation period, because
every catalyst showed the specific highest selectivities at the begin-
ning of each aging / activation procedure, but before reaching its
steady state. Furthermore, every catalyst showed nearly constant
selectivities at low degrees of conversion. At a certain point, the se-
lectivity dropped with increasing degree of conversion and for a layer
thickness of 400 nm, this point was located at about 23%, for 800
nm at 27% and for 1400nm Ag layer thickness conversion degrees
up to 32% were attained without losing much selectivity. Again,
the formation of the by-product acetaldehyde was monitored (Fig.
4.17). The lower the Ag layer thickness, the higher the acetaldehyde
concentration at the same degree of conversion. Using high Ag layer
thicknesses, the concentration kept finally constant.
With 1400 nm Silver sputtered on the the structured side,
4.1. Epoxidation of ethene in microchannel reactors 75

0,030
50 nm Ag
Figure 4.17: Concentration
concentration of acetaldehyde / %

0,025 400 nm Ag of acetaldehyde as a function


800 nm Ag of the conversion degree for
0,020 1400 nm Ag the Ag/Al microchannel reac-
tor MMCR3 at different Ag
0,015 layer thicknesses. Reaction
conditions: 20% C2 H4 in O2 ,
0,010
p= 0.3 MPa, T= 523 K.
0,005

0,000

0,0 0,1 0,2 0,3 0,4


degree of conversion

0,65

Figure 4.18: Selectivity to


ethylene oxide as a func-
selectivity to ethylene oxide

0,60
b tion of the conversion degree
(attained by residence time
0,55 variation) for the Ag/Al mi-
crochannel reactor MMCR3
a
(1400 nm Ag). Unstructured
0,50
backside of wafers (a) with-
out coating (b) with 400 nm
0,45 Ag sputter coating. Reaction
uncoated unstructured side conditions: 20% C2 H4 in O2 ,
400 nm Ag on unstructured side
p= 0.3 MPa, T= 523 K.
0,40
0,1 0,2 0,3 0,4

degree of conversion

400 nm were added to the unstructured backside of the wafers in


order to investigate the influence of the uncoated aluminum surface
area within the reaction zone. The wafer was mounted again in
the modular microchannel reactor and the selectivity / conversion
behavior was monitored (Fig. 4.18). At a conversion degree of
16%, the selectivity is 55% for the wafers having an uncoated
backside, whereas the wafers with the coated backside exhibit
61.3% selectivity. At higher conversion degrees, the wafers with
the coated backside perform even better. At a conversion degree of
30%, the selectivity is 55.5% compared to 62.7%, respectively.
76 Chapter 4. Results

4.1.2.5 Influence of the aluminum pretreatment on selec-


tivity and conversion degree (MMCR4,MMCR5)

In order to evaluate the influence of a pretreatment of the aluminum


surface on the catalytic performance of the silver coated structure,
two different catalysts were prepared and tested in the modular mi-
crochannel reactor. The first catalyst (MMCR4) was made by sput-
tering silver onto a microstructured aluminum wafer on both sides.
Therefore, silver was sputtered on bare aluminum, having an ex-
tremely thin natural Al2 O3 layer. The second catalyst (MMCR5)
is made in the same way with the same silver layer thickness, ex-
cept that the aluminum wafer was anodically oxidized for 20 min-
utes in oxalic acid before the sputtering took place. Therefore, an
Al2 O3 layer having a controlled thickness of about 1 m was pre-
served. Both experiments were performed using a single microstruc-
tured wafer, which was tested in the modular microchannel reactor
type I. The geometric parameters of both reactors with their en-
closed wafers are listed in table 4.6.
The selectivity and the conversion degree as a function of the
residence time is depicted in figure 4.19. At residence times between
78 and 590 ms, the conversion degree of the Ag/Al microchannel
reactor MMCR4 increased from 3.3 to 11.3%, with the selectivity
increasing in the same manner from 45.2% to 51.1%, respectively.
The anodically oxidized MMCR5 exhibited initially much higher se-
lectivities. Adjusting a residence time of 78 ms, the selectivity was
50.7%. With increasing residence time, the degree of conversion in-
creased to 18.7%, yielding lower, but still high selectivities of 47.1%.
Thus, the Ag/Al2 O3 /Al microchannel reactor exhibited higher
degrees of conversion at the same flow rate and initially nearly con-
stant selectivities, whereas the Ag/Al reactor showed increasing se-
lectivities with increasing degree of conversion and its best perfor-
mance at high residence times. Therefore, the more active anodic
preoxidized type was tested in a commercial microchannel reactor.
The results of this test and its catalytic properties are described in
the following section.
4.1. Epoxidation of ethene in microchannel reactors 77

Table 4.6: Geometric parameters of the modular microchannel reactors


MMCR4 and MMCR5.

Basic reactor type MMCR type I


Channel width 300 m
Channel height 700 m
Channel length 50 mm
Number of channels per wafer 14
Number of wafers 1
Wafer height 1000 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 14 cm2
Total channel volume 0.098 cm3
Total stack volume 0.5 cm3
Ag coating thickness 450 nm
Coating method MMCR4: Ag on Al
MMCR5: Ag on
1 m Al2 O3 (GX 285.50.20)

0,52
selectivity to ethylene oxide

0,50

0,48

0,46

0,44
degree of conversion

0,15

0,10

0,05
MMCR4: Ag on unmodified Al surface

MMCR5: Ag on anodic oxidized Al surface


0,00
0 100 200 300 400 500 600

residence time / ms

Figure 4.19: Selectivity to ethylene oxide and degree of conversion


as a function of the residence time for the microchannel reactors
MMCR4 (4) and MMCR5 (2). Reaction conditions: 20% C2 H4 in O2 ,
p= 0.3 MPa, T= 523 K.
78 Chapter 4. Results

4.1.2.6 Silver supported on anodic preoxidized aluminum


surface (Ag/Al2 O3 /Al, MCR2)

According to the high activity observed with the modular mi-


crochannel reactor MMCR5, a commercial made microchannel
reactor was ordered. The activation of the microstructured AlMg3
wafers was performed by an anodic oxidation of the bare aluminum
followed by immobilization of silver (800 nm layer thickness) by
sputtering. To compare the results obtained in the first bulk silver
made microchannel reactor MCR1, the same reactor and channel
geometry was chosen. The most important parameters of this
microchannel reactor MCR2 are listed in table 4.7
The activation of this microchannel reactor MCR2 was per-
formed using a similar method as for MCR1, except the reactor
temperature was set to 503 K to take the expected higher activity
into account Further experiments showed a typical behavior for a
variation of oxygen concentration, ethene concentration and the to-
tal pressure are depicted and described in the appendix, see Fig.7.1
to 7.4.

Table 4.7: Geometric parameters of the Ag/Al2 O3 /Al microchannel re-


actor MCR2.

Basic reactor type FZ K made MCR


Channel width 200 m
Channel height 200 m
Channel length 50 mm
Number of channels per wafer 33
Number of wafers 26
Wafer height 300 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 343 cm2
Total channel volume 1.72 cm3
Total stack volume 3.9 cm3
Ag coating thickness 800 nm
Coating method sputtering of Ag
1 m Al2 O3 (GX 285.50.20)
wafers. Coating on both sides.
4.1. Epoxidation of ethene in microchannel reactors 79

Selectivity / conversion behavior: The selectivity and con-


version behavior of this reactor was investigated for ethene concen-
trations of 4% and 20% in oxygen at different reactor temperatures
by variations of the residence time in order to attain different de-
grees of conversion. The results of those experiments are depicted
in figure 4.20 and 4.21.
Adjusting 4% ethene in oxygen, the highest selectivities are ob-
served at the lowest adjusted reactor temperature of 463 K. The
selectivity to EO seems to be nearly independent from the degree of
conversion. Between 16% and 33% conversion degree, the selectivity
was nearly constant at 65.7% to 64.9%, respectively. At higher reac-
tor temperatures, and therefore, at higher degrees of conversion, the
selectivity was still nearly independent from the conversion degree.
At 483 K, having conversion degrees from 27.3% to 55.8%, the selec-
tivity varied from 64.2% to 62.1% and at 503 K, the selectivity was
within a range of 59.6% to 57.7% at conversion degrees of 45.0% to
70.4%. Although there were only small changes in the selectivity, the
selectivity seemed to decrease slightly with increasing degree of con-
version. At any given degree of conversion, the selectivity decreased
with increasing reactor temperature.
Different results are obtained when the ethene concentration
was increased to 20%, still using oxygen as balance. At a reac-
tor temperature of 463 K, the selectivity to EO was between 68.4 -
65.1% at conversion degrees of 4.2% and 15.1%, respectively. The
selectivity decreased with increasing conversion degree, especially at
long residence times and therefore, at high degrees of conversion.
Adjusting higher reactor temperatures, the initial selectivity to EO
decreased as expected. At the highest investigated reactor temper-
ature of 503 K, the selectivity was in the range of 61.2 - 55.1% at
conversion degrees between 15.4% and 39.7%, respectively. Surpris-
ingly, the selectivity at higher reactor temperatures was not neces-
sary lower than at lower reactor temperatures as observed for 4%
ethene in fig. 4.20. Having conversion degrees above 25%, the selec-
tivity was at 503 K unexpectedly higher than at 483 K. A similar
observation has to be expected, when the steep decline in selectiv-
ity at a reactor temperature of 463 K is cautiously extrapolated for
conversion degrees greater than 15%. Such an observation was also
made before with MCR1 (Fig. 4.5, p. 59).
80 Chapter 4. Results

0,70

0,68
selectivity to ethylene oxide
0,66
463K

0,64
483K
0,62

0,60
503K
0,58

0,56

0,2 0,3 0,4 0,5 0,6 0,7

degree of conversion

Figure 4.20: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the microchannel reactor
MCR2. Reaction conditions: 4% C2 H4 in O2 , p= 0.3 MPa, T= 463 / 483 /
503 K.
0,70

0,68
463K
0,66
selectivity to ethylene oxide

0,64
483K
0,62

0,60

0,58 503K
0,56

0,54

0,52

0,50
0,05 0,10 0,15 0,20 0,25 0,30 0,35 0,40

degree of conversion

Figure 4.21: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the microchannel reactor
MCR2. Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa, T= 463 /
483 / 503 K.
4.1. Epoxidation of ethene in microchannel reactors 81

Influence of the reactor temperature: The influence of the


reactor temperature on selectivity and conversion degree is depicted
in figure 4.22. At a reactor temperature of 433 K, the selectivity
was 72.7% at a conversion degree of 1.4%. With increasing reactor
temperature, a nearly linear decrease of the selectivity was observed.
The slope of the regression is -0.200, i.e. the selectivity decreased by
2% with an increase in reactor temperature of 10 K. The conversion
degree increased as expected with increasing reactor temperature by
approximately factor 1.5 per 10 K.
selectivity to ethylene oxide

0,70

0,65

0,60

0,55
degree of conversion

0,15

0,10

0,05

0,00
430 440 450 460 470 480 490 500 510

reactor temperature / K

Figure 4.22: Selectivity to ethylene oxide and degree of conversion for


the microchannel reactor MCR2 as a function of the reactor temperature.
Reaction conditions: 20% C2 H4 in O2 , = 0.94 s.

Aging of the Ag/Al2 O3 /Al microchannel reactor MCR2:


In order to investigate the effect of aging on this catalyst, the very
first stable X/S run (fresh catalyst/coating, see Fig. 4.21) and a
final control experiment has to be compared. The result of this
comparison is shown in figure 4.23.
The fresh catalyst exhibited initial selectivities of 65.4% at a
conversion degree of 8.25% ( = 0.94 s). With increasing degree
82 Chapter 4. Results

0,66
fresh catalyst
aged catalyst
selectivity to ethylene oxide

0,64

0,62

0,60

0,58

0,56

0,05 0,10 0,15 0,20 0,25 0,30

degree of conversion

Figure 4.23: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the microchannel reactor
MCR2 for a fresh and aged catalyst. Reaction conditions: 20% C2 H4 in
O2 , p= 0.3 MPa, T= 483 K.

of conversion, the selectivity decreased, resulting in a selectivity of


57.4% at a conversion degree of 27.1%. The aged catalyst exhibited
slightly higher degrees of conversion, starting with 60.3% selectivity
at a conversion degree of 10.7% ( = 0.94 s). With increasing de-
gree of conversion of up to 20%, the selectivity was initially nearly
unchanged and close to 60%. Applying higher degrees of conversion,
the selectivity decreased, resulting in 56.3% selectivity at a conver-
sion degree of 28.1%. Therefore, the aged catalyst exhibited nearly
the same selectivity at high degrees of conversion compared to the
fresh one. In contrast, the fresh catalyst offered 4 5% higher
selectivities within a conversion range of 10-17%.

Formation of acetaldehyde: The formation of acetaldehyde was


observed only at 503 K, 20% ethene in oxygen at a residence time
of 5 s in very small amounts of 0.008 vol% (see Fig. 4.21). At lower
temperatures or lower residence times, the acetaldehyde concentra-
4.1. Epoxidation of ethene in microchannel reactors 83

tion was always below 0.004 vol% and thus, below the detection limit
of the online gas chromatograph for this component.

4.1.2.7 Silver supported on metallic Al (Ag/Al, MCR3)

In the following section, results of the Ag/Al microchannel reactor


MCR3 are presented. Again, the same reactor geometry as already
used for MCR1 and MCR2 was used. This time, the microstructured
aluminum wafers were activated by sputtering of silver without ap-
plication of the anodic oxidation. The Ag-layer thickness was in
principle comparable with MCR2, but it was presumed the aging
might be suppressed by waiving the anodic oxidation and decreas-
ing the amount of exposed alumina support material. Preliminary
experiments (see MMCR4/MMCR5, Fig. 4.19, p.77) indicated a
similar X/S behavior of silver on aluminum.
The most important geometric parameters are listed in table
4.8. Aiming to increase the silver layer thickness in order to im-

Table 4.8: Geometric parameters of the Ag/Al2 O3 /Al microchannel re-


actor MCR3.

Basic reactor type FZK made MCR


Channel width 200 m
Channel height 200 m
Channel length 50 mm
Number of channels per wafer 33
Number of wafers 26
Wafer height 300 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 343 cm2
Total channel volume 1.72 cm3
Total stack volume 3.9 cm3
Ag coating thickness 1200 nm
Coating method 400 nm sputtering of Ag
followed by PVD, struct.
side 2 times perpendicular
+/-45o , each 800nm.
Back side 1 m Ag by PVD
84 Chapter 4. Results

prove the selectivity / conversion behavior, the coating process of


the cleaned aluminum (AlMg3) wafers was performed in three steps:

1. Deposition of 400 nm Silver by sputtering on the structured


side. The sputter process is comparably slow, but provides a
good coating quality.

2. Deposition of additional 800 nm Silver by PVD. The rapid


coating by PVD was performed two times using different an-
gles. Having shadow effects near the rectangular walls of the
micromachined channels within the PVD process, the coat-
ing was performed twice using plus and minus 45 degrees to
perpendicular, respectively. Therefore, the whole channel sur-
face including the perpendicular channel walls were coated by
PVD.

3. Deposition of 1 m Silver by PVD in a single step on the


flat backside of the wafers.

As a result, the thickness of the silver layer deposited by sputtering


and PVD is 1200 nm.
The activation of this microchannel reactor MCR3 was per-
formed using a similar method as for MCR1 and MCR2. The reactor
temperature was initially at 503 K and the same as for MCR2, but
raised to 523 K in order to accelerate the process. This experiment
as well as further experiments showing a typical behavior for a vari-
ation of oxygen concentration, ethene concentration and the total
pressure are depicted and described in the appendix, see Fig.7.5 to
7.8.

Selectivity / conversion behavior at low reactor tempera-


tures: The selectivity / conversion behavior of this Ag/Al coated
microchannel reactor MCR3 at low reactor temperatures and low
degrees of conversion using the fresh catalyst is depicted in figure
4.24 for an ethene concentration of 4% in O2 and in figure 4.25 for
20% in O2 .
Using 4% C2 H4 in O2 , selectivities of up to 68.2% at low degrees
of conversion and low reactor temperatures of 483 K were observed.
With increasing degree of conversion of up to 27.3%, the selectivity
4.1. Epoxidation of ethene in microchannel reactors 85

0,68
483K
selectivity to ethylene oxide
0,66

503K
0,64

0,62

0,60 523K

0,58

0,56
0,10 0,15 0,20 0,25 0,30 0,35 0,40 0,45 0,50 0,55

degree of conversion

Figure 4.24: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/Al microchan-
nel reactor MCR3. Reaction conditions: 4% C2 H4 in O2 , p= 0.3 MPa,
T= 483 / 503 / 523 K.

decreased to 66.4%. At higher reactor temperatures and therefore,


higher degrees of conversion, the loss in selectivity increased with in-
creasing conversion degree. With an increase of the reactor temper-
ature to 523 K, selectivities of 64.8% to 57.3% at conversion degrees
between 32.0% and 55.5% were observed.
At higher ethene concentrations of 20% in O2 , lower conversion
degrees but higher selectivities were observed (Fig. 4.25). Applying
a reactor temperature of 483 K, selectivities between 70.5 - 67.1%
were observed at conversion degrees between 3.5 - 13.3%. With in-
creasing reactor temperature, increasing degrees of conversion were
observed. At 523 K, selectivities of 66.1 - 59.7% at conversion de-
grees of 11.8% to 26.0%, respectively, were observed. Again, the
selectivity at higher reactor temperatures was not necessarily lower
than at lower reactor temperatures as observed within an experiment
using 4% ethene (see Fig. 4.24). Having conversion degrees above
16.5%, the selectivity was at 523 K again higher than at the same
conversion degree at 503 K. The same phenomenon is expected for
86 Chapter 4. Results

0,70

selectivity to ethylene oxide 483K


0,68

503K
0,66

0,64

523K
0,62

0,60

0,05 0,10 0,15 0,20 0,25

degree of conversion

Figure 4.25: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/Al microchannel
reactor MCR3. Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa,
T= 483 / 503 / 523 K.

reactor temperatures of 483 and 503 K, if conversion degrees above


13.3% are adjusted and the graph for the X/S curve at 483 K is cau-
tiously extrapolated. Exactly at this point, the selectivity to ethene
oxide is at 483 K and 503 K very much the same.

Selectivity / conversion behavior at high reactor tempera-


tures: In order to apply high heat production rates by adjusting
high flow rates and high ethene concentrations and to determine the
optimal ethene concentration for this reactor MCR3, the selectiv-
ity / conversion behavior was subsequently examined at different
ethene concentrations of up to 60% using reactor temperatures up
to 563 K. Due to the previous experiments, the reactor was already
aged.
The results of those experiments at 563 K are depicted in fig-
ure 4.26. The highest selectivities of 52.7% were obtained at ethene
concentrations of 20% and conversion degrees of approximately 30%.
The lowest selectivity was observed at 60% ethene and the highest
4.1. Epoxidation of ethene in microchannel reactors 87

0,53
0,52 4% C 2H 4
0,51
selectivity to ethylene oxide
0,50 20% C 2H 4
0,49
0,48
0,47 40% C 2H 4
0,46
0,45

0,44
0,43
0,42 60% C 2H 4

0,1 0,2 0,3 0,4 0,5 0,6 0,7

degree of conversion

Figure 4.26: Selectivity to ethylene oxide for the Ag/Al microchannel


reactor MCR3 (aged catalyst) as a function of the conversion degree at
different ethene concentrations in O2 . Reaction conditions: 4/20/40/60%
C2 H4 in O2 , p= 0.3 MPa, 563 K.

conversion degree in this experiment of 23.2%. Generally, the selec-


tivity increased initially with increasing degree of conversion in each
experiment. At high conversion degrees, decreasing selectivities were
observed, except for 4% ethene in O2 . It may be assumed, that even
a conversion degree of 69.9% was still not sufficient to cause a loss
in selectivity.
To estimate the space-time-yield of the reactor, the EO concen-
tration for each experiment and ethene concentration was plotted
as a function of the residence time (STP). This plot is depicted in
figure 4.27. At any given residence time, the highest EO concentra-
tions were achieved when 40% ethene in oxygen was applied. Higher
ethene concentrations of 60% led to decreased selectivities (see fig.
4.26) accompanied by lower downstream EO concentrations. Al-
though the highest selectivities to EO were observed at 20% ethene,
the EO concentration at any given residence time is lower than at
40% C2 H4 .
Due to the quite constant EO selectivity within the experiment
88 Chapter 4. Results

40% C H
2 4
7
EO concentration / %

6 60% C H
2 4

20% C H
2 4
5

4% C H
2 4

0,0 0,2 0,4 0,6 0,8 1,0 1,2 1,4 1,6 1,8 2,0 2,2 2,4

residence time / s

Figure 4.27: Downstream concentration of ethylene oxide for the Ag/Al


microchannel reactor MCR3 (aged catalyst) as a function of the residence
time (STP) at different ethene concentrations in O2 . Reaction conditions:
4/20/40/60% C2 H4 in O2 , p= 0.3 MPa, 563 K.

using 4% ethene, the selectivity was monitored as a function of the


conversion for conversion degrees up to > 99%. The reaction temper-
ature was decreased by 20 K to 543 K as higher residence times were
applied because of the high activity of this catalyst. Results of this
experiment are depicted in figure 4.28. At this reactor temperature,
the selectivity showed a slight increase from 52.2% at 26% conversion
degree to 53.1% at a conversion degree of 44 - 53%. Higher conver-
sion degrees of 80% lead to decreased selectivities of 51%, finally
resulting in 20.6% selectivity at a conversion degree of 99%.

Influence of the reactor temperature: The influence of the


reactor temperature on selectivity and conversion degree was investi-
gated using (a) the fresh catalyst before the high temperature / high
throughput experiments were performed and (b) the aged catalyst
after these experiments at very high reactor temperatures. Results
for (a) are depicted in figure 4.29. The selectivity seems to decrease
nearly linear with increasing reactor temperature, indicating a selec-
4.1. Epoxidation of ethene in microchannel reactors 89

0,55

0,50
selectivity to ethylene oxide
0,45

0,40

0,35

0,30

0,25

0,20

0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1,0

degree of conversion

Figure 4.28: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/Al microchannel
reactor MCR3 (aged catalyst). Reaction conditions: 4% C2 H4 in O2 ,
p= 0.3 MPa, T= 543 K.

tivity loss of roughly 1.3% per 10 K increase in reactor temperature.


The degree of conversion seems to increase exponentially as expected
for the Arrhenius equation. At 483 K, the selectivity is 66.1% at a
conversion degree of 4.86%. A similar behavior was expected as this
experiment was repeated after the catalyst / reactor was exposed
to high temperatures (Fig. 4.30) and high conversion degrees (aged
catalyst). Due to aging effects - which will be described in detail
within the following section - the selectivity decreased and the con-
version degree increased. At 523 K, the conversion degree increased
to to 8.41%, yielding a selectivity of 61.5%. With increasing reactor
temperature, the selectivity loss per 10 K increased reactor temper-
ature increased from 1.3% to 1.75%. The degree of conversion seems
to increase exponentially as expected.

Aging effect of the Ag/Al microchannel reactor MCR3: In


order to investigate the effect of aging on this catalyst, one of the
very first X/S experiments (fresh catalyst/coating) and one of the
last is compared. In all experiments, the same residence times were
90 Chapter 4. Results

selectivity to ethylene oxide


0,66

0,64

0,62

0,60

0,58
degree of conversion

0,20

0,15

0,10

0,05

480 490 500 510 520 530 540 550

reactor temperature / K

Figure 4.29: Selectivity to ethylene oxide and degree of conversion for


the fresh Ag/Al microchannel reactor MCR3 as a function of the reactor
temperature. Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa, =1.0 s.

0,70
selectivity to ethylene oxide

0,65

0,60

0,55

0,25
degree of conversion

0,20

0,15

0,10

0,05

0,00
450 460 470 480 490 500 510 520 530
reactor temperature / K

Figure 4.30: Selectivity to ethylene oxide and degree of conversion for


the aged Ag/Al microchannel reactor MCR3 as a function of the reac-
tor temperature. Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa,
=1.0 s.
4.1. Epoxidation of ethene in microchannel reactors 91

0,66

increasing TOS
first experiment
0,64
selectivity to ethylene oxide
0,62

0,60 last experiment

0,58

0,56

0,54

0,52

0,50

0,48

0,46

0,10 0,15 0,20 0,25 0,30 0,35 0,40 0,45 0,50 0,55 0,60 0,65

degree of conversion

Figure 4.31: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the microchannel reactor
MCR3 at different times on stream. Reaction conditions: 20% C2 H4 in
O2 , p= 0.3 MPa, T= 523 K.

adjusted. Therefore, the very first and the very last data point in
each experiment refers to the same residence time. The results of
those experiments are shown in figure 4.31.
With the fresh catalyst, the selectivity decreased with increas-
ing reactor temperature from a range of initially 66.1% to 59.7%
at conversion degrees of 11.7% and 26.0%, respectively (see Fig.
4.25). With increasing time on stream, including some experiments
at very high reactor temperatures, the selectivity dropped to a range
of 55.8% to 46.2% at higher conversion degrees of 22.3% to 60.4%,
respectively. Therefore, the conversion degree nearly doubled as a re-
sult of aging of this catalyst, whereas the selectivity to EO decreased
by 10.3% (abs.). Furthermore, the shape of the X/S curves changed.
Although the selectivity of the fresh catalyst was higher than that
of the aged one, the selectivity of the fresh dropped sharply at con-
version degrees of about 20 - 25%. In contrast, the aged catalyst
showed no remarkable dependence of the selectivity on conversion
degree within that range.
92 Chapter 4. Results

Formation of acetaldehyde: In this Ag/Al microchannel reac-


tor MCR3, no formation of acetaldehyde was observed. Even at high
reactor temperatures of 563 K and EO concentrations close to 9%,
the acetaldehyde concentration was below 0.002%. Therefore, the
rate of acetaldehyde formation must have been at least 4500 times
slower than the rate of EO formation.
4.1. Epoxidation of ethene in microchannel reactors 93

4.1.3 Silver supported on -Al2 O3 surface

-Al2 O3 is the only support material used in industrial EO catalysts.


Therefore, efforts were made to prepare a silver catalyst supported
on -Al2 O3 surface in microchannels. In the following section, three
different methods are described, which result in the immobilization
of silver supported on -Al2 O3 surface in a microchannel reactor.
Similar to the experiments with the Ag/Al system, no promoters
were applied in order to study the basic performance of this silver
catalyst and to avoid errors by different impacts of promoters on
different catalysts.

4.1.3.1 Silver sputtered on ANOF prepared -Al2 O3 sur-


face (MMCR6)

One method to prepare an alumina layer on aluminum is to make


use of a suitable plasma-chemical oxidation of the aluminum surface
in order to yield -Al2 O3 . Therefore, the ANOF-method3 [91] was
used to coat a WEDM4 made aluminum wafer with a thin alumina
film. On this -Al2 O3 layer, silver was immobilized by sputtering.
In order to test this catalytic system, three aluminum wafers,
each having 14 channels of 50 mm x 0.3 mm x 0.7 mm, were plasma-
chemically oxidized and coated with 450 nm silver by sputtering.
Those coated wafers were used in the modular microchannel reactor
type I in order to evaluate the performance of this coating. The acti-
vation of this system (MMCR6) was performed at 548 K (first traces
of CO2 as indicator for catalytic activity), using 20% C2 H4 and 20%
O2 in CH4 at a residence time of 260 ms. The resulting selectivity to
EO and the degree of conversion as a function of the time on stream
during the activation is depicted in figure 4.32. With increasing time
on stream, the selectivity declined steadily from 55 -57% directly at
the beginning down to 41% after nearly 12 hours time on stream.
In contrast, the degree of conversion initially increased as expected
from 0.1 - 0.5% at the beginning to a maximum of 1.4% at 3.3h,
after which decreasing conversion degrees were observed. After 12
hours, 0.7% conversion degree were left, still with a decreasing trend.
3 Anodic oxidation by spark discharge
4 Wire electro discharge machined
94 Chapter 4. Results

Table 4.9: Geometric parameters of the modular microchannel reactor


MMCR6 having an Ag/-Al2 O3 surface as catalyst

Basic reactor type MMCR type I


Channel width 300 m
Channel height 700 m
Channel length 50 mm
Number of channels per wafer 14
Number of wafers 3
Wafer height 1000 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 42 cm2
Total channel volume 0.29 cm3
Total stack volume 1.5 cm3

Ag coating thickness 450 nm


Coating method sputtering on structured side

0,016

0,6
0,014
degree of conversion

0,012

0,010
0,5

0,008

0,006

0,4
0,004

0,002
0 5 10 15

time on stream / h

Figure 4.32: Selectivity to ethylene oxide and degree of conversion for


the Ag/-Al2 O3 microchannel reactor MMCR6 as a function of the initial
time on stream during the activation. Reaction conditions: 20% C2 H4 ,
20% O2 , balance CH4 , = 260 ms, p= 0.3 MPa, T= 548 K.
4.1. Epoxidation of ethene in microchannel reactors 95

At this point, the experiment was stopped. Despite the low conver-
sion degree and low EO selectivity, small amounts of acetaldehyde
(0.002 to 0.004%) were observed within the first five hours of oper-
ation. Due to these disappointing results, this coating method was
not further investigated.

4.1.3.2 Silver sputtered on -Al2 O3 (sol-gel coating)

Another method to provide an -Al2 O3 layer on aluminum wafers


is to make use of a sol-gel derived process. The experiments were
performed using four WEDM made wafers, which were coated with
an -Al2 O3 layer by the HITK5 . In order to activate those wafers for
the ethene epoxidation, 400 nm silver were sputtered on both sides of
the wafers, which were mounted in the modular microchannel reactor
type I and tested for their catalytic performance. The geometric
parameters of this microchannel reactor are listed in table 4.10. The
activation of the silver coating was performed using 20% C2 H4 and
20% O2 in CH4 at 523 K, applying a residence time of 350 ms.

Table 4.10: Geometric parameters of the Ag/-Al2 O3 microchannel re-


actor MMCR7.

Basic reactor type MMCR type I


Channel width 300 m
Channel height 700 m
Channel length 50 mm
Number of channels per wafer 14
Number of wafers 4
Wafer height 1000 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 56 cm2
Total channel volume 0.59 cm3
Total stack volume 2.0 cm3

Ag coating thickness 1200 nm


Coating method 400 nm sputtering of Ag
on sol-gel prepared -Al2 O3

5 Hermsdorfer Institut fur Technische Keramik e.V., http://www.hitk.de


96 Chapter 4. Results

0,07 0,7

0,06 0,6

selectivity to ethylene oxide


0,05 0,5
degree of conversion

0,04 0,4

0,03 0,3

0,02 0,2

0,01 0,1

0,00 0,0

0,0 0,5 1,0 1,5 2,0 2,5 3,0

/ s

Figure 4.33: Selectivity to ethylene oxide and degree of conversion as


a function of the residence time for the Ag/-Al2 O3 (sol-gel derived)
microchannel reactor MMCR7. Reaction conditions: 20% C2 H4 in O2 ,
p= 0.3 MPa, T= 503 K.

After the activation, the selectivity and degree of conversion was


monitored as a function of the residence time, using 20% C2 H4 in
O2 . The results are depicted in figure 4.33. At low residence times
of 340 ms, 3.85% conversion at a selectivity of 66.2% to EO was
observed. With increasing residence time, the selectivity dropped
nearly linear down to 27% at a residence time of 3.2 s. Surpris-
ingly, the conversion degree did not increase with increasing time on
stream, but showed a maximum at = 1.1 s. Higher residence times
resulted in lower selectivities and lower degrees of conversion. At
= 3.2 s, 4.8% conversion degree were left. This obvious deactivation
proved to be irreversible. Due to the poor performance this catalytic
system was abandoned and no further investigations performed.

4.1.3.3 Silver impregnated on -Al2 O3 surface by sol-gel


coating (MMCR8)

Another method to immobilize silver on -Al2 O3 is to make use of


conventional impregnation methods. Therefore, three WEDM made,
4.1. Epoxidation of ethene in microchannel reactors 97

-Al2 O3 (sol-gel) coated microstructured wafers were impregnated


with silver by use of an Ag-lactate solution, applying an immobiliza-
tion method already described by Minahan et. al. [78]. After disap-
pointing results with a single impregnation step (exemplary X/S be-
havior depicted in fig. 4.34), resulting in low to nearly zero selectiv-
ities even at low degrees of conversion, the impregnation procedure
was repeated two more times to increase the immobilized amount of
silver on the catalysts surface. Therefore, an activation diagram of

0,30
Figure 4.34: Selectivity to
ethylene oxide as a func-
selectivity to ethylene oxide

0,25
tion of the conversion degree
0,20 (attained by residence time
variation) for the Ag/-
0,15 Al2 O3 microchannel reac-
0,10
tor MMCR8 (impregnated
once). Reaction condi-
0,05 tions: 20% C2 H4 in O2 ,
p= 0.3 MPa, T= 523 K.
0,00

0,00 0,05 0,10 0,15 0,20


degree of conversion

the fresh catalyst is unavailable. After the additional impregnation


steps, the catalyst showed satisfying selectivities and much better
degrees of conversion. Therefore, the influence of the reaction con-
ditions on this improved catalytic system had been investigated and
the results will be described in the following paragraphs. The most
important geometric parameters of this reactor are listed in table
4.11.

Selectivity / conversion behavior: The selectivity / conver-


sion behavior of this Ag/-Al2 O3 catalyst was investigated at a reac-
tor temperature of 483 K and 503 K using 20% C2 H4 in O2 . Results
are depicted in figure 4.35. The highest selectivities of 54% were
observed at 483 K and low degree of conversion. Initially, the selec-
tivity was nearly constant at conversion degrees between 3.2% and
10.6%. Higher conversion degrees led to decreased selectivities. At
98 Chapter 4. Results

Table 4.11: Geometric parameters of the Ag/-Al2 O3 /Al microchannel


reactor MMCR8.

Basic reactor type MMCR type I


Channel width 300 m
Channel height 700 m
Channel length 50 mm
Number of channels per wafer 14
Number of wafers 3
Wafer height 1000 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 42 cm2
Total channel volume 0.44 cm3
Total stack volume 1.5 cm3
-Al2 O3 coating thickness 1 m
Coating method Sol-gel coating with -Al2 O3
and triple impregnation
with silverlactate solution

0,55

0,50
selectivity to ethylene oxide

0,45

0,40

0,35
483K 503K
0,30

0,25

0,20

0,15

0,10
0,0 0,1 0,2 0,3 0,4 0,5 0,6

degree of conversion

Figure 4.35: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/-Al2 O3 mi-
crochannel reactor MMCR8. Reaction conditions: 20% C2 H4 in O2 ,
p= 0.3 MPa, T= 483 / 503 K.
4.1. Epoxidation of ethene in microchannel reactors 99

29.7% conversion degree, 46.3% selectivity were left. At 503 K, the


selectivity is initially 51% and constant for conversion degrees up to
15.7%. Adjusting higher degrees of conversion, the selectivity de-
creased to 15.1.% at 62.4% conversion degree. Again, there was a
crossing of both X/S-curves observed.
In order to determine the optimal ethene concentration for this
highly active catalyst, the selectivity / conversion behavior was ex-
amined by variation of the residence time again applying C2 H4 con-
centrations of 4%, 20%, 40% and 60%. The resulting selectivity /
conversion behavior and the downstream EO concentration as a
function of the residence time for different ethene concentrations
are depicted in figure 4.36 and 4.37.
Surprisingly, the initial selectivity at low degrees of conversion
seemed to be nearly independent from the concentration and was
very close to 50%. The highest selectivities of 51.1% were obtained
0,55

0,50

0,45
selectivity to ethylene oxide

0,40
40% C2H4
0,35
60% C2H4
0,30

0,25

0,20

0,15
20% C2H4
0,10

0,05
4% C2H4
0,00
0,0 0,2 0,4 0,6 0,8 1,0

degree of conversion

Figure 4.36: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/-Al2 O3 mi-
crochannel reactor MMCR8 applying different ethene concentrations in
O2 . Reaction conditions: 4/20/40/60% C2 H4 in O2 , p= 0.3 MPa,
T= 503 K.
100 Chapter 4. Results

concentration of ethylene oxide / % 40% C H


2 4

60% C H
2 4
3

20% C H
2 4

4% C H
2 4

0,0 0,5 1,0 1,5 2,0 2,5

/ s

Figure 4.37: Downstream concentration of ethylene oxide for the Ag/-


Al2 O3 microchannel reactor MMCR8 as a function of the residence time
(STP) at different ethene concentrations in O2 . Reaction conditions:
4/20/40/60% C2 H4 in O2 , p= 0.3 MPa, T= 503 K.

at ethene concentrations of 20% and conversion degrees of approx-


imately 9 to 15%. The lowest selectivity of 5.5% was observed at
4% ethene and the highest conversion degree in the experiment of
97.1%. Generally, the selectivity seemed to be very much the same
and nearly independent from the ethene concentration at low de-
grees of conversion. At 4% and 20% ethene, the selectivity to EO
was initially independent from the conversion degree. Adjusting con-
version degrees exceeding 50% and 15%, respectively, the selectivity
eventually decreased with increasing degree of conversion.
It was noted, that the concentration of ethylene oxide seemed
to be highly a function of the residence time as long as low degrees
of conversion / short residence times were applied. By plotting the
EO concentration for each ethene concentration as a function of the
corresponding residence time (STP, fig. 4.37) this surprising finding
is illustrated. Applying residence times below 250 ms, the EO con-
centration was nearly independent from the C2 H4 concentration.
4.1. Epoxidation of ethene in microchannel reactors 101

Furthermore, at 4% and 20% ethylene the EO concentration showed


a maximum at 250 ms and 750 ms, respectively. Higher residence
times lead to lower downstream EO concentrations, indicating con-
secutive reactions consuming EO.

Influence of the reactor temperature: In order to allow a


calculation of the reaction rates according to the kinetic model of an
ideal differential reactor, the influence of the reactor temperature on
the selectivity and conversion degree was investigated at low degrees
of conversion (<10%) and thus, at high flow rates (Fig. 4.38). The
selectivity seems to decrease nearly linear with increasing reactor
temperature, dropping from 57.8% at 463 K down to 51.0% at 503 K.
Therefore, a selectivity loss of roughly 1.7% per 10 K increase in
reactor temperature was observed. The degree of conversion seemed
to increase exponentially from 1.55% at 463 K to 6.53% at 503 K.
selectivity to ethylene oxide

0,58

0,56

0,54

0,52

0,50
degree of conversion

0,06

0,04

0,02

0,00
460 470 480 490 500 510

reactor temperature / K

Figure 4.38: Selectivity to ethylene oxide and degree of conversion as


a function of the reactor temperature for the Ag/-Al2 O3 microchannel
reactor MMCR8. Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa,
= 24 ms.
102 Chapter 4. Results

Effect of high ethene partial pressures: Within the experi-


ments at high ethene concentration, changes in the selectivity / con-
version behavior of the catalyst had to be noted. In figure 4.39, the
selectivity / conversion behavior of the catalyst directly after prepa-
ration and after an exposition to high (>30%) C2 H4 concentrations
in O2 is shown. Contrary to the sputtered Ag/Al catalysts, this
type of coating showed no noteworthy aging by application of high
conversion degrees but a very sudden and unexpected change after
application of high ethene partial pressures. Initially, selectivities of
55% at a conversion degree of 20% were observed. With increasing
conversion, the selectivity dropped sharply down to 17.5% at 47.1%
conversion degree. After exposition to high C2 H4 concentrations,
the catalyst gained suddenly activity and lost selectivity when data
points at the same residence time are compared. At a residence time
of i.e. 240 ms, the catalyst changed from 19.7% conversion degree

0,60

= 240 ms
0,55
selectivity to ethylene oxide

0,50

0,45 aged, after exposition to high

C H concentrations
2 4

0,40

0,35

0,30

0,25

0,20

fresh,
0,15
after preparation

0,10

0,2 0,3 0,4 0,5 0,6 0,7

degree of conversion

Figure 4.39: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/-Al2 O3 mi-
crochannel reactor MMCR8. Reaction conditions: 20% C2 H4 in O2 ,
T= 503 K, p= 0.3 MPa.
4.1. Epoxidation of ethene in microchannel reactors 103

and 54.9% selectivity to 28.9% conversion degree at a selectivity of


48.9%. The latter state was stable and no further change in activity
or selectivity was observed when high C2 H4 or O2 concentrations
were applied again. In this altered / aged state, higher conversion
degrees and lower selectivities were observed. Applying the same res-
idence times as in the former experiment, conversion degrees from
28.9% to 62.4% were observed and the selectivity decreased from
48.9% to 15.1%, respectively. Applying conversion degrees above
28%, the catalyst exhibits higher selectivities after the C2 H4 treat-
ment than before. A cautious extrapolation of the curves towards
low degrees of conversion suggests, that as long as high selectivi-
ties are favored and low degrees of conversion acceptable, the fresh
state should be preferred. As soon as 28% conversion degree is
exceeded, the aged state shows better results. Unless not otherwise
noted, all experiments in this section refer to the latter stable state
of the catalyst (after initial reduction).

Formation of acetaldehyde The formation of acetaldehyde was


observed at concentrations of about 0.004% only at 503 K and 20%
C2 H4 in O2 applying high degrees of conversion. This suggests a
ratio of rEO /rAcal of at least 1200 to 1500 and selectivities towards
acetaldehyde of 0.03% at the most. Neither at 4%, nor at 40%
or 60% C2 H4 in O2 , the acetaldehyde concentration exceeded the
detection limit of the GC.
104 Chapter 4. Results

4.1.3.4 Usage of a commercial SHELL-800 Series, -


Al2 O3 based EO silver catalyst in microchannel
reactors (MMCR9, MMCR10)

Up to now, it seemed impossible to investigate commercial available


catalysts in microchannel reactors, because this reactor type requires
a catalytic active coating on the walls of the channels. There is no
evidence of a a commercial ready to run catalyst being successfully
applied in a microchannel reactor. Thus, microchannel reactors had
to be catalytically activated by using i.e. sol-gel coatings, anodic
oxidation of aluminum followed by impregnation or coating with
PVD/CVD techniques. The resulting layers have typically a good
adhesion strength, but the catalyst and its deposition method had
to be developed and optimized again in a time consuming manner.
In order to make use of a commercial EO catalyst in a mi-
crochannel reactor, the application of a suitable kind of adhesive in
order to immobilize a fine crushed conventional catalyst at the walls
of the microchannels and sol-gel coatings seems to make sense. Due
to the nature of the silver catalyst and the reaction itself, the use sil-
icon and aluminum based sol-gel coatings as adhesive was excluded,
because:

Silicon based sol-gels exhibit a high surface area, which makes


them inapplicable for isomerization/ combustion-sensitive re-
action products.

Boehmit sol-gel recipes, which form Al2 O3 layers in pres-


ence of Al2 O3 particles [79] contain large amounts of water,
but silver catalysts are according to the manufacturers infor-
mation very sensitive towards H2 O.

Water free aluminum sol-gels tend to form amorphous alumina


or even Al2 O3 layers, which catalyze the isomerization of
ethene oxide to acetaldehyde followed by combustion.

Thus, it is not advised to use silicon or aluminum sol-gels as


glue. Instead, an electrostatic deposition method was utilized in
order to immobilize the fine crushed silver catalyst on the walls of
microstructured aluminum wafers without application of any bind-
ing material. Preliminary investigations showed, the resulting layer
4.1. Epoxidation of ethene in microchannel reactors 105

had a low adhesion strength and mechanical stress such as touch-


ing or wiping should be strictly avoided. But the adhesion was still
sufficient to prevent a blow out or leaching of catalyst during the ex-
periments. Thus, simple microstructured aluminum wafers (AlMg3)
having a metallic surface were coated with the crushed Shell 800
Series Catalyst and used in the modular microchannel reactors.
In these experiments, a different type of microstructure was
used. The material, AlMg3 was the same, but the manufacturing
process of the wafers changed. Instead of a time consuming and
expensive electro discharge machined wafer type, micro milled wafers
having a channel geometry of 300 x 300 m2 and a wafer thickness
of 500 m were used (for details see page 204).
Up to now, the selectivity/conversion curves at low tempera-
tures and high residence times were sometimes lower than at higher
temperatures and lower residence times - an observation being dif-
ficult to understand from a kinetic point of view (see e.g. Fig.4.5,
p.59). This observation was independent from the reactor type and
type of coating. Plain Ag-surfaces in a commercial reactor were af-
fected as well as Ag/Al and Ag/Al2 O3 and Ag/-Al2 O3 -surfaces in
commercial as well as self-made reactors. To test post-catalytic EO
combustion in the diffusers of commercial and mentioned self-made
modular reactors, a new type of a modular reactor was tested hav-
ing an optimized dead volume in the diffusers to minimize residence
time in the post catalytic zone.
In order to allow a comparison of different modular microchan-
nel reactor types, the same wafers were used in the modular mi-
crochannel reactors type I and type II (see page 149 for details),
which are referred as MMCR9 and MMCR10, respectively. Both
reactor types have slightly different geometries, but the catalytic ac-
tive coating and even the wafers in both microchannel reactors were
the same. The geometric parameters of both reactors are listed in
table 4.12.

Activation: The activation of this commercial catalyst was per-


formed similar to the activation of the previously investigated mi-
crochannel reactors under reaction conditions. In contrast to the
activation of all (modular) microchannel reactors investigated until
now and following an advice from the manufacturer, the oxygen par-
106 Chapter 4. Results

Table 4.12: Geometric parameters of the modular microchannel reactors


MMCR9 and MMCR10 coated with the Shell 800 Series Catalyst.

Label MMCR9 MMCR10


Basic reactor type MMCR type I MMCR type II
Channel width 300 m 300 m
Channel height 300 m 300 m
Channel length 50 mm 50 mm
Number of channels per wafer 19 19
Number of wafers 6 4
Wafer height 500 m 500 m
Wafer width 10 mm 10 mm
Wafer length 50 mm 50 mm
Total geometric surface area 68.4 cm2 45.6 cm2
Total channel volume 0.513 cm3 0.342 cm3
Total stack volume 1.5 cm3 1.0 cm3
Coating 250-300 mg 160-200 mg
calc. thickess: 30 m
of a Shell 800 Series Catalyst
Coating method electrostatic deposition

tial pressure was lowered from 20% to 8.8% in order to apply oxygen
conditions, which are comparable to industrially used reaction con-
ditions. It will be shown later (see chapter 4.4) that MMCR10 using
the new modular microchannel reactor type II design having low post
catalytic volumes performed better than the traditional modular mi-
crochannel reactor type I. Therefore, after activation four of the six
wafers were mounted in the new modular reactor type II (MMCR10)
and all further examinations in this part refer to MMCR10 and the
newer construction type of the microchannel reactor.

Selectivity / conversion behavior: Due to the superb heat


transfer abilities of the microchannel reactor, it was possible to ap-
ply high reactor temperatures in order to achieve high degrees of
conversion without risk of explosions or other thermal instabilities.
This is again demonstrated by experiments depicted in figure 4.40
and 4.41.
Having 4% ethene in oxygen, selectivities between 56.7% and
30.5% at conversion degrees between 15.8 and 75.1% were observed
4.1. Epoxidation of ethene in microchannel reactors 107

483K
0,55
selectivity to ethylene oxide
0,50 503K

0,45

523K
0,40

0,35

0,30
543K

0,2 0,3 0,4 0,5 0,6 0,7 0,8

degree of conversion

Figure 4.40: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time vatiation) at different reactor temper-
atures for the Shell 800 Series Catalyst coated modular microchannel
reactor MMCR10. Reaction conditions: 4% C2 H4 in O2 , T= 483, 503,
523 and 543 K, p= 0.3 MPa.

(Fig. 4.40). With increasing temperature, increasing conversion de-


grees and decreasing selectivities were noted at the same residence
time. At conversion degrees below 40-50%, little dependency of the
selectivity on conversion degree was observed. With increasing re-
actor temperature and thus, with increasing degree of conversion,
the selectivity drops from a certain conversion degree on. At 483 K,
nearly unchanged selectivities of 56.7 to 55.6% were obtained at at
conversion degrees between 15.8 to 36.5%. At 503 K, the selectivity
was initially 53.5% and decreased at conversion degrees above 40%
to 50.5%. At 523 K, the selectivity decreased further down to 48%
at conversion degrees between 36 and 50%. Higher conversion de-
grees lead again to decreased selectivities of 41.3% at a conversion
degree of 67.8%. Finally, the lowest selectivities and the highest
conversion degrees were observed at a reactor temperature of 543 K.
In this experiment, conversion degrees between 43.5 and 75.1% were
observed, with decreasing selectivities in the range of 44.3% down
to 30.5%.
108 Chapter 4. Results

0,65

selectivity to ethylene oxide 0,60 483K

0,55

503K
0,50

0,45
523K

0,40

0,35 543K

0,05 0,10 0,15 0,20 0,25 0,30 0,35 0,40 0,45 0,50

degree of conversion

Figure 4.41: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) at different reactor temper-
atures for the Shell 800 Series Catalyst coated modular microchannel
reactor MMCR10. Reaction conditions: 20% C2 H4 in O2 , T= 483, 503,
523 and 543 K, p= 0.3 MPa.

The same investigations were performed for 20% C2 H4 in O2 .


At 483 K, the highest selectivities of 60.8% to 59.9% at conversion
degrees between 4.2% to 14.9% were obtained. With increasing re-
actor temperature and therefore, higher degrees of conversion, the
selectivity became more and more dependent on the conversion de-
gree. At 503 K, the selectivity varied between 54.8% to 51.4% at
conversion degrees of 8.1% to 25.4%, respectively. Initially, at high
flow rates and conversion degrees below 13%, the selectivity seemed
to be constant. A further increase of the reactor temperature up to
543 K led to decreased selectivities in the range of 45.9% to 33.7%
at conversion degree of 18.7% to 46.1%, respectively. Again, the
selectivity seemed to be quite constant at conversion degrees below
27%. It has to be noted, that no crossing of selectivity curves was
observed, neither at 4% nor at 20% C2 H4 in O2 at any temperature.
4.1. Epoxidation of ethene in microchannel reactors 109

Influence of the reactor temperature: The influence of the


reactor temperature is depicted in figure 4.42. At 483 K, the highest
selectivity of 60.8% at a conversion degree of 4.2% was observed.
With increasing reactor temperature, the selectivity decreased to
45.9%, whereas the degree of conversion increased to 18.7%. There-
fore, approximately 2.4% selectivity loss per 10 K increase in reactor
temperature were observed.
selectivity to ethylene oxide

0,6

0,5
degree of conversion

0,15

0,10

0,05

480 490 500 510 520 530 540 550


reactor temperature / K

Figure 4.42: Selectivity to ethylene oxide and degree of conversion for


the Shell 800 Series Catalyst coated modular microchannel reactor
MMCR10 as a function of the reactor temperature. Reaction conditions:
20% C2 H4 in O2 , = 230 ms, p= 0.3 MPa.

Formation of acetaldehyde: Using this Ag/Al microchannel re-


actor MMCR9 / MMCR10, the formation of acetaldehyde was ob-
served only at 543 K, 20% C2 H4 in O2 applying the highest resi-
dence time of 1.7 s. At this point, 0.004-0.006% acetaldehyde was
observed at an EO concentration of 3.3%. At this point, the con-
version degree was 46% and the selectivity 33%. Therefore, the rate
of acetaldehyde formation was approximately 550-800 times slower
than the corresponding EO formation rate at this point. Applying
higher flow rates / shorter residence times of 620 ms and selectivities
110 Chapter 4. Results

above 40% (but comparable EO concentrations), no acetaldehyde


was monitored. Therefore, the ratio rEO /rAcal was normally higher
than 1500, indicating selectivities to acetaldehyde of 0.03% at the
most.

4.1.4 Silver supported on stainless steel surfaces


(MMCR11)
The suitability of stainless steel as a potential support material for
silver was evaluated. Many commercial available microchannel re-
actors / heat exchangers are made of stainless steel instead of alu-
minum, because diffusion bonding of steel is much more simple and
reliable than bonding of aluminum. Therefore, a thin silver layer
supported on steel would be a good activation method for those
devices. In order to investigate the suitability of silver-on-steel, a
stainless steel Betamesh 75 net as support was coated with 300 nm
Ag by sputtering and mounted in the modular microchannel reactor
type I. The porous structure of the stacked steel nets formed some-
thing similar to a microchannel system, but exhibiting a more ir-
regular channel geometry [80]. Nevertheless, this structure is cheap,
easy to handle and suitable for performing preliminary catalytic in-
vestigations. The most important parameters for this structure are
listed in table 4.13. The activation of this Ag/Fe catalytic system
was performed under reaction conditions, using 20% ethene, 20%
oxygen in methane at a reactor temperature of 548 K. Results are
depicted in figure 4.43. Initially, high selectivities to EO of 67 - 69%
were observed and the conversion degree varied between 3.7 - 3.9%.
This level was reached after only 1.5 hours time on stream and no
further activation was observed. Contrary, the selectivity decreased
rapidly with increasing time on stream and after 14 hours, only 45%
selectivity at conversion degree of 3.5% were left. At this time, the
oxygen concentration was increased from 20% to 80%, keeping the
flow rate (100 ml/min) and the reactor temperature on the same
level. The selectivity jumped to 49% and the degree of conversion
to 5.6%. Again, with increasing time on stream, the selectivity de-
creased to 45% at conversion degrees of 4.6% at 26 h time on stream.
Due to the steadily declining selectivities with time on stream, this
catalytic system was abandoned and no further investigations per-
formed.
4.1. Epoxidation of ethene in microchannel reactors 111

Table 4.13: Geometric parameters of the Ag/-Al2 O3 /Al microchannel


reactor MMCR11.

Basic reactor type MMCR type I


Channel width 300 m
Channel height 300 m
Channel length 50 mm
Number of channels per wafer 16
Number of wafers 8
Wafer height 300 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area n.a.
Total channel volume 0.7 cm3
Total stack volume 1.25 cm3
Ag coating thickness 300 nm
Coating method sputtering of Ag
selectivity to ethylene oxide

0,7
20% O 2 80% O 2

0,6

0,5

0,4
0,06

0,05
degree of conversion

0,04

0,03

0,02

0,01
0 10 20 30
time on stream / h

Figure 4.43: Selectivity to ethylene oxide and degree of conversion as


a function of the time on stream for an Ag / stainless steel microchan-
nel reactor. Reaction conditions: 20% C2 H4 in 20% O2 (later 80% O2 ),
T= 548 K, p= 0.3 MPa, 420 ms.
112 Chapter 4. Results

4.1.5 Influence of promotors on Ag/Al coatings


(MMCR12, MMCR13, MCR2Cs)
It is known the selectivity to EO can be increased by 20 to 30%, es-
pecially when 1,2-dichloroethane (DCE) [35, 39], CO2 and NOx [39]
are used as gaseous promoters, and e.g. Cs, Rb, Ba, Re, P, B, S
or Sn [37, 38] as solid promoters. Several patents and publications
from the 1970ies such as [52] describe a process to reactivate a sil-
ver catalyst for ethene epoxidation by passing a Rb and/or Cs salt
containing solution at ambient temperatures through the reactor.
Therefore, the effect of Cs as a classical solid promoter and NO2
as a gaseous promoter was investigated using Ag/Al microchannel
reactors.

4.1.5.1 Influence of NO2 and Cs on an Ag/Al coated mi-


crochannel reactor (MMCR12, MMCR13)

The influence of NO2 on an Ag/Al catalytic surface was investi-


gated in a modular microchannel reactor in order to avoid potential
damage of the expensive commercial made microchannel reactors.
Therefore, microstructured wafers of the FZK-type were coated with
silver and mounted in the modular microchannel reactor type I. The
geometric parameters of this modular microchannel reactor denoted
as MMCR12 are listed in table 4.14. This reactor is similar to MCR3,
regarding the coating, the geometry of the microchannels and the
geometry of the wafer stack. In the moment, MMCR12 was treated
with Cs, its label was changed to MMCR13.
After activation of this catalyst under reaction conditions, four
different sets of experiments were subsequently performed in this
reactor:

1. MMCR12: Unpromoted, unmodified Ag/Al catalyst


Directly after the coating and initial activation procedure,
the selectivity / conversion behavior was determined for the
Ag/Al catalyst applying ethene concentrations of 4% and 20%
in oxygen.

2. MMCR12: NO2 promoted, unmodified Ag/Al catalyst


4.1. Epoxidation of ethene in microchannel reactors 113

Table 4.14: Geometric parameters of the modular microchannel reactor


MMCR12 and MMCR13 having an Ag/Al microstructure as catalyst

Basic reactor type MMCR type I


Channel width 200 m
Channel height 200 m
Channel length 50 mm
Number of channels per wafer 33
Number of wafers 26
Wafer height 300 m
Wafer width 10 mm
Wafer length 50 mm
Total geometric surface area 343 cm2
Total channel volume 1.72 cm3
Total stack volume 3.9 cm3
Ag coating thickness 1200 nm
Coating method 400 nm sputtering of Ag
MMCR12 & MMCR13 followed by PVD, struct.
side 2 times perpendicular
+/-45o , each 800nm.
Back side 1 m Ag by PVD
MMCR13 only Addition of CsCl in CH3 OH

Experiments with NO2 as gaseous promoter were performed


by using a NO2 / O2 mixture, substituting the pure O2
feed. This NO2 -mixture was prepared by filling an evacuated
stainless steel cylinder with a well defined pressure of a 2%
NO2 in He calibration gas mixture followed by a multiple
pressurizing process with pure O2 in order to obtain different
NO2 concentrations. The influence of different NO2 concen-
trations was investigated as well as changes in the selectivity /
conversion behavior caused by use of nitrogen dioxide.

3. MMCR13: Unpromoted, Cs modified Ag/Al catalyst


In order to test a regeneration of the catalyst by immobiliza-
tion of Cs, a 0.1% solution of CsCl in Methanol was pumped
for 40 min through the reactor at room temperature. After
this Cs treatment, the reactor and its enclosed microstructure
114 Chapter 4. Results

was flushed with nitrogen to remove remaining methanol.

4. MMCR13: NO2 promoted, Cs modified Ag/Al catalyst


The influence of NO2 on the now Cs modified Ag/Al catalyst
was investigated by adding NO2 to the feed, as already de-
scribed in experiment (2). Again, the selectivity / conversion
behavior was monitored.

Basic selectivity / conversion behavior of the unpromoted


Ag/Al microchannel reactor MMCR12 - experiment (1):
The selectivity to ethene oxide as a function of the conversion de-
gree using the unpromoted Ag/Al catalyst at ethene concentrations
of 4% and 20% is depicted in figure 4.44. It showed, that the se-
lectivities were within a narrow range of 49.5 to 56% at conversion
degrees of up to 87%. An initial increase as well as a final drop in
0,58

20% C2H4 4% C2H4


0,56
selectivity to ethylene oxide

0,54

0,52

0,50

0,48

0,46
0,0 0,2 0,4 0,6 0,8 1,0

degree of conversion

Figure 4.44: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the unpromoted Ag/Al
microchannel reactor MMCR12. Reaction conditions: 4% and 20% C2 H4
in O2 , T=503 K, p=0.3 MPa.
4.1. Epoxidation of ethene in microchannel reactors 115

selectivity were observed with increasing degree of conversion within


both experiments applying 4% and 20% ethene in oxygen. Using 4%
ethene in oxygen, conversion degrees up to 87% could be attained.
The highest selectivities of 56% were obtained at 50 to 70% degree
of conversion at 4% ethene and at 20 to 30% conversion degree for
a concentration of 20% ethene in oxygen. Having concentrations of
20% ethene, the highest degree of conversion was close to 50%, still
allowing selectivities of about 54%.

Effect of NO2 on the Ag/Al microchannel reactor MMCR13


- experiment (2) The dependence of the selectivity and conver-
sion degree on the NO2 level is depicted in figure 4.45. At NO2 con-
centrations up to 20 ppm, only little influence on selectivity and con-
version degree was observed. At higher concentrations of 130 ppm,
the selectivity increased to 73% accompanied by a remarkably de-
creased conversion degree of about 10% to 11%. A further increase
0,75
selectivity to ethylene oxide

0,70

0,65

0,60

0,20
degree of conversion

0,15

0,10

0 50 100 150 200 250


NO 2 concentration / ppm

Figure 4.45: Selectivity to ethylene oxide and degree of conversion for


the Ag/Al microchannel reactor MMCR12 as a function of the NO2 con-
centration. Reaction conditions: 20% C2 H4 in O2 , T= 503 K, = 850 ms,
p= 0.3 MPa.
116 Chapter 4. Results

in nitrogen dioxide concentration to 240 ppm resulted in decreased


selectivities of 69.8% at a conversion degree of 9.1%.
The dependency of selectivity and conversion degree as a func-
tion of the reactor temperature is depicted in Fig. 4.46 for an NO2
concentration of 130 ppm. The conversion degree of the NO2 pro-
moted Ag/Al catalyst was lower than in absence of NO2 at reactor
temperatures below 510 K. In addition, at a reactor temperature
of 463 K, no conversion and thus, no product formation (neither
CO2 nor EO) was observed and therefore, no selectivity calculated.
In contrast, the conversion degree increased in the presence of NO2
by up to 10% (absolute) at reactor temperatures above 510 K, still
yielding higher selectivities than in the absence of nitrogen dioxide.
The selectivity is generally improved by NO2 at any temperature.
At low reactor temperatures, the highest selectivity of 77% to ethene
oxide was observed at a conversion degree of 5.5%.
0,8
selectivity to ethylene oxide

0,7

0,6

0,5
unpromoted
0,4 130 ppm NO2
degree of conversion

0,3

0,2

0,1

0,0

460 480 500 520 540


reactor temperature / K

Figure 4.46: Selectivity to ethylene oxide as a function of the reac-


tor temperature for the NO2 promoted Ag/Al microchannel reactor
MMCR12. Reaction conditions: 20% C2 H4 in O2 , 130 ppm NO2 ,
= 850 ms, p= 0.3 MPa.
4.1. Epoxidation of ethene in microchannel reactors 117

In order to investigate aging effects as a reason for decreased


selectivities in Fig. 4.45, the selectivity and conversion was moni-
tored at a selected reference point in absence of NO2 (Fig. 4.47).
Initially, a selectivity of 55.3% at a conversion degree of 18.4% was
observed. After the very first experiments using NO2 at low con-
centrations (below 20 ppm, Fig. 4.45), the selectivity at this control
experiment increased in absence of NO2 to 60.1% at a conversion
degree of 19.3%. After the catalyst / reactor had been exposed to
high NO2 concentrations (240 ppm), the selectivity to EO decreased
suddenly to 52.2% at a conversion degree of 22.2% within the con-
trol experiment, again measured in absence of NO2 . Therefore, NO2
exposition lead to a slight initial increase of the selectivity, but high
NO2 concentrations caused a sudden and irreversible loss of selec-
tivity accompanied by increased degrees of conversion.

degree of conversion
selectivity to EO Figure 4.47: Degree of
0,6
conversion and selectivity
to ethylene oxide of the
0,5
Ag/Al microchannel reactor
MMCR12 at NO2 -free con-
0,4
trol experiments. Reaction
conditions: 20% C2 H4 in
0,3 O2 , p= 0.3 MPa, T= 503 K,
= 1.0 s.
0,2

0,1

0,0
inital state after NO 2 treat- after NO 2 treat-
ment (low conc.) ment (high conc.)

Effect of Cs and NO2 + Cs on the Ag/Al microchannel re-


actor MMCR13- experiments (3) and (4) In order to com-
pensate the loss in selectivity in the former experiments, a regen-
eration of the deactivated catalyst was performed. Cs, a classical
solid promoter and regenerator, was immobilized according to a re-
activation / immobilization method described in Alfansreders orig-
inal publication [52]. Thus, CsCl dissolved in CH3 OH was pumped
through the assembled microchannel reactor MMCR12 for half an
118 Chapter 4. Results

hour, followed by drying at room temperature under nitrogen to


remove remaining methanol. This reactor is now denominated as
MMCR13 by this modification with Cs. Directly after the drying,
the reactor was heated to reaction temperature and the reactivation
was performed under reaction conditions. After constant selectivities
and conversion degrees were obtained, the selectivity / conversion
behavior of this Cs modified reactor was monitored. In a follow-
ing experiment, NO2 was added as a promoter to the feed in order
to investigate the combined effect of NO2 and Cs. As soon as this
reactor was exposed to 130 ppm NO2 as in experiment 2, this Cs
modified Ag/Al catalyst did not show signs of activity. Therefore,
the NO2 concentration had to be decreased down to 8 ppm in order
to attain catalytic activity. Thus, the Cs modified Ag/Al catalyst
proofed to be much more sensitive towards NO2 than the unmod-
ified one. The results of these experiments are depicted in figure
4.48. For comparison purposes, the selectivity / conversion behav-

0,76
Exp.1: unprom.
0,74
Exp.2: NO (130 ppm) prom.
2
0,72
Exp.3: Cs modified

0,70 Exp.4: NO (8 ppm) prom.


selectivity to ethene oxide

0,68 & Cs modified

0,66

0,64

0,62

0,60

0,58

0,56

0,54

0,52

0,50

0,48

0,0 0,1 0,2 0,3 0,4 0,5 0,6

degree of conversion

Figure 4.48: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) of the Ag/Al microchannel
reactor MMCR12/13 in presence of NO2 and Cs. Reaction conditions:
20% C2 H4 in O2 , T= 503 K, p= 0.3 MPa.
4.1. Epoxidation of ethene in microchannel reactors 119

ior of the unpromoted catalyst (experiment 1) and of the 130 ppm


NO2 promoted catalyst (experiment 2) are depicted too.
Experiment 3: In absence of NO2 , the Cs modified Ag/Al cat-
alyst showed conversion degrees in the range of 10.7% to 30.0% at
selectivities from 69.8% to 66%. Initially, for conversion degrees
up to 20%, the selectivity was nearly constant and only at higher
degrees of conversion, decreasing selectivities with increasing con-
version degrees were observed.
Experiment 4: In presence of 8 ppm NO2 in the feed, the Cs
modified Ag/Al catalyst showed slightly enhanced selectivities at
low degrees of conversion. Initially, 72.3% selectivity at a conversion
degree of 10% was observed. With increasing conversion degree, the
selectivity decreased steadily and finally, at a degree of conversion
of 30.7%, the selectivity was down to 63.5%.
Generally, the selectivity seems to decrease steadily with in-
creasing degree of conversion, as soon as NO2 is added to the feed. In
absence of NO2 , the selectivity is initially nearly constant as shown
for the Cs modified Ag/Al catalyst or is even increasing slightly as
shown for the totally unpromoted and unmodified catalyst. Due to
the observed sudden and irreversible deactivation in presence of high
NO2 levels and the unpredictable effect of NO2 in presence of other
modifiers as Cs, no further experiments were performed.

REM analysis of the surface: After the catalytic experiments


had been performed, the microstructured wafers were removed from
the reactor and examined by backscatter-REM microscopy (Fig.
4.49). This technique allows to monitor the Cs concentration on
the surface of the catalyst. The bright parts of the pictures refer
to high Cs concentrations. The picture (Fig. 4.49a) shows the top
view of a single channel (running from top left to lower right). Both
webs are bright and therefore modified with Cs, but only half of the
channel seems to be covered with Cs. The lower part of the channel
is dark and thus, nearly Cs free. Therefore, it got either not in con-
tact with Cs or no Cs adsorption took place. The sharp borderline
between the Cs coated and uncoated part is shown enlarged in figure
4.49b.
120 Chapter 4. Results

Figure 4.49: Cs-selective backscatter REM photo of the surface of the


Cs modified MMCR13 after use. (A) overview, (B) enlargement of the
middle of the channel.

4.1.5.2 Regeneration of MCR2 by immobilization of Cs


(MCR2Cs)

Regeneration / catalytic activation: Inspired by the positive


results of the regeneration of MMCR12 with Cs, the Ag/Al2 O3 /Al
microchannel reactor MCR2 was chosen in order to examine the ef-
fect of Cs on the catalytic behavior of an Ag/Al2 O3 /Al catalyst. The
Cs immobilization method was the same as for MMCR13 and again,
the reactivation was performed under normal reaction conditions,
using 20% ethylene in oxygen. This reactor / catalyst combination
is denoted as MCR2Cs. The selectivity and conversion degree as a
function of the time on stream during the reactivation is depicted
in figure 7.16 (see appendix, p. 233ff). Initially, the selectivity was
between 77 to 80% at conversion degrees of 1-2%. With increasing
time on stream, the degree of conversion increased, whereas the se-
lectivity to EO decreased. After approximately 40 hours, the degree
of conversion was quite constant. At this time, the selectivity was
down to 69.5% and the degree of conversion up to 16.5%.

Selectivity / conversion behavior: The selectivity as a func-


tion of the conversion degree was monitored for ethene concentra-
tions of 4% and 20%, each with oxygen as balance and at reactor
4.1. Epoxidation of ethene in microchannel reactors 121

temperatures of 463, 483 and 503 K by variation of the residence


time (Fig. 4.50 and 4.51). At 4% ethene in oxygen and a reactor
temperature of 463 K, conversion degrees of 19.6 to 44.8% were ob-
served at again nearly constant selectivities between 67.8 and 67.2%.
With increasing reactor temperature, lower selectivities but higher
degrees of conversion were observed. At 483 K, selectivities between
64.5 and 66% were observed at conversion degrees between 33 and
65%. Initially, a slight increase in selectivity from 65.4% at a conver-
sion degree of 33% to 66.3% at 45% conversion degree was observed.
With higher conversion degrees, the selectivity decreased slightly to
64.5% at 66% conversion degree. At a reactor temperature of 503 K,
the selectivities decreased to 65.1 - 60.6%, accompanied by increased
conversion degrees of 46 - 77%. The high initial selectivities of 65%
were maintained for conversion degrees up to 60%. Higher conver-
sion degrees up to 77% lead to noticeable decreased selectivities of
60.6%. For comparison, see Fig. 4.20 (p.80).
0,68

463K
selectivity to ethylene oxide

0,66

483K
0,64

0,62

503K
0,60
0,2 0,3 0,4 0,5 0,6 0,7 0,8

degree of conversion

Figure 4.50: Selectivity to ethylene oxide as a function of the conver-


sion degree (attained by residence time variation) for the Cs modified
microchannel reactor MCR2Cs. Reaction conditions: 4% C2 H4 in O2 ,
p= 0.3 MPa, T= 463 / 483 / 503 K.
122 Chapter 4. Results

Increasing the ethene concentration from 4 to 20% in oxygen,


nearly constant selectivities up to 69.9% at conversion degrees be-
tween 5.6 to 18.6% were observed at a reactor temperature of 463 K.
With increasing reactor temperature, the selectivity decreased and
the conversion degree increased. At 483 K and conversion degrees up
to 20%, nearly constant selectivities of 69.1% were observed. Higher
degrees of conversion lead to marginally lower selectivities of 67.7%
at conversion degrees of 30%. Finally, the highest conversion de-
grees of 19.6 to 42.8% were observed at a reactor temperature of
503 K. Within that range, the selectivity varied between 66.2% and
63.3%. At low conversion degrees of up 27%, the selectivity was ini-
tially nearly constant, varying only slightly between 66.2 and 66.8%.
Higher degrees of conversion lead to more and more decreased selec-
tivities.
As it can be clearly seen from a comparison of the unpromoted
MCR2 (Fig. 4.21, p. 80) and the Cs treated MCR2Cs (Fig. 4.51),
0,71

0,70 463K
selectivity to ethylene oxide

0,69
483K
0,68

0,67

0,66

503K
0,65

0,64

0,63
0,05 0,10 0,15 0,20 0,25 0,30 0,35 0,40 0,45

degree of conversion

Figure 4.51: Selectivity to ethylene oxide as a function of the conver-


sion degree (attained by residence time variation) for the Cs modified
microchannel reactor MCR2Cs. Reaction conditions: 20% C2 H4 in O2 ,
p= 0.3 MPa, T= 463 / 483 / 503 K.
4.1. Epoxidation of ethene in microchannel reactors 123

the crossover-effect of the selectivity / conversion curves no longer


observed.

Influence of the reactor temperature: The influence of the


temperature on selectivity and conversion degree of the Cs modi-
fied microchannel reactor MCR2Cs is depicted in figure 4.52. At a
reactor temperature of 433 K, the selectivity is 77.9% at a conver-
sion degree of 1.68%. With increasing reactor temperature, a nearly
linear decrease of the selectivity is observed. The slope of the regres-
sion is -0.160, i.e. the selectivity decreases by approximately 1.6%
with an increase in reactor temperature of 10 K. The conversion de-
gree increased as expected with increasing reactor temperature by
approximately factor 1.5 per 10 K increase in reactor temperature.
selectivity to ethylene oxide

0,80

0,75

0,70

0,65

0,60
degree of conversion

0,20

0,15

0,10

0,05

0,00
430 440 450 460 470 480 490 500 510

reactor temperature / K

Figure 4.52: Selectivity to ethylene oxide and degree of conversion as


a function of the reactor temperature for the Cs modified microchannel
reactor MCR2Cs. Reaction conditions: 20% C2 H4 in O2 , = 0.94s.
124 Chapter 4. Results

4.2 Epoxidation in traditional tube-type


fixed bed reactors
The focus of the ethene epoxidation to EO in the current work was
on microchannel reactors. In order to allow a comparison of the re-
sults obtained by using microchannel reactors, corresponding cata-
lysts were made and used in a traditional tube-type fixed bed reactor
having an irregular packing. In the following sections, the catalytic
behavior of cut silver foils as well as silver coated aluminum foils and
commercial Ag/-Al2 O3 catalyst particles were investigated. The
foil-based catalysts were used in an amount, that the same geomet-
ric surface area as in the corresponding microchannel reactor was
applied, in case of the commercial catalyst, an equivalent catalyst
mass was used.

4.2.1 Bulk silver catalysts (FBR1, similar


MCR1)
The first catalyst type used in the fixed bed reactor was plain silver
foil, which is similar to the bulk-silver microchannel reactor MCR1
(see section 4.1.1). Therefore, silver foil6 (> 99.9% Ag) having a
total surface area of 340 cm2 and a thickness of 0.1 mm was cut
into small pieces to make it usable in a fixed bed reactor, which was
denoted as FBR1a. After the activation was performed in the same
way, the corresponding microchannel reactor MCR1 was activated,
the selectivity / conversion behavior of this catalyst was examined
by varying the residence time at 20% ethene in oxygen. The results
of these experiments are depicted in figure 4.53.
At 523 K, selectivities of initially 70.8% at a conversion degree
of 2.1% were observed. With increasing residence time and thus,
increasing conversion degree, the selectivity increased to 75.1% at
a conversion degree of 6.4%. Higher conversion degrees of 8.2% led
to slightly lower selectivities of 74.3%. At 543 K, higher degrees of
conversion but lower selectivities were obtained. Initially, 65.1% se-
lectivity to EO at a conversion degree of 2.8% were observed. Again,
with increasing residence time and increasing conversion degree, the
6 supplied by Chempur
4.2. Epoxidation in traditional tube-type fixed bed reactors 125

0,75

FB R 1a, 523K

0,70
selectivity to ethylene oxide
FBR 1a, 543K
0,65

0,60
MCR 1, 523K

0,55

0,50 MCR 1, 543K

0,45

0,40

0,00 0,05 0,10 0,15 0,20 0,25 0,30

degree of conversion

Figure 4.53: Selectivity to ethylene oxide as a function of the conversion


degree Reaction conditions (attained by residence time variation) for the
bulk silver fixed bed reactor FBR1a and the bulk silver microchannel
reactor MCR1: 20% C2 H4 in O2 , p= 0.3 MPa, T= 523 / 543 K.

selectivity increased to a maximum of 68.4% at a conversion de-


gree of 7.1%. Higher conversion degrees of up to 12.6% led to lower
selectivities of 67.5%.
In those experiments, the same range of flow rates was applied
as in the microchannel reactor MCR1. At 523 K, MCR1 allowed
much higher conversion degrees of 23.8% at lower selectivities of
59.6 - 49.2%, decreasing with increasing conversion degree. At a
reactor temperature of 543 K, the conversion degree increased to
30.7% at the most, exhibiting selectivities between 57.2 and 42.0%.
Having such a very different catalytic performance, the experi-
ment was repeated with another set of silver foil. This time, silver
foil having the same purity of more than 99.9% and the same thick-
ness of 0.1 mm, but from a different supplier7 was used. Again, the
foil was hackled into small pieces, mounted in the same fixed bed re-
actor (denoted FBR1b), activated and examined applying the same
7 supplied by Goodfellow
126 Chapter 4. Results

selectivity to ethylene oxide


Fixed Bed Reactor FBR1b
0,60
523K

0,55
543K

0,50

0,45

0,40
selectivity to ethylene oxide

Microchannel Reactor MCR1


0,60

0,55

0,50

523K
0,45

543K
0,40
0,05 0,10 0,15 0,20 0,25 0,30 0,35 0,40 0,45

degree of converion

Figure 4.54: Selectivity to ethylene oxide as a function of the conver-


sion degree for the bulk silver fixed bed reactor FBR1b and the bulk
silver microchannel reactor MCR1. Reaction conditions: 20% C2 H4 in
O2 , p= 0.3 MPa, T= 523 / 543 K.

reaction conditions. Results are depicted in figure 4.54. This time,


lower selectivities of 54.4 to 58.0% at 523 K and 49.7 to 53.3% at
543 K, respectively were observed. The conversion degree increased
from 6.7% to 9.7% at 523 K and 14.3 to 39.6% at 543 K, respectively.
Again, the selectivity increased with increasing degree of conversion
by abs. 3-4% in each experiment. In these experiments, the ini-
tial selectivities were lower than in the corresponding microchannel
reactor MCR1.
At 523 K, very similar degrees of conversion were observed for
the fixed bed- and microchannel reactor when similar flow rates were
applied8 . At 543 K, the fixed bed reactor was much more active but
initially less selective than the corresponding microchannel reactor.
Therefore, the trends in the selectivity / conversion curves are very
different and still, the fixed bed reactor showed initially increasing
selectivities with increasing conversion degree, whereas for the mi-
crostructured reactor, decreasing selectivities were observed.
8 From left to right on each curve, the same set of flow rates was applied.
4.2. Epoxidation in traditional tube-type fixed bed reactors 127

This very different catalytic behavior of different silver samples


having a common surface area of 343 cm2 is underlined by an ex-
periment, which was performed with high purity silver wire, having
a diameter of 0.3 mm and a calculated total geometric surface area
of 160 cm2 . After the wire was cut into small pieces and mounted
in the fixed bed reactor, the activation was started at 523 K, us-
ing 20% C2 H4 and 20% O2 . This time, no catalytic activity was
observed, even after the reactor temperature was raised to 573 K.
Similar observations were made with silver wool (0.025 mm thick-
ness) as catalyst. Again, no catalytic activity was observed even at
high reactor temperatures.

4.2.2 Silver supported on Al2 O3 /Al (FBR2, sim-


ilar MCR2)

The sputter and PVD coated aluminum wafers used in the mi-
crochannel reactors MCR2 (Ag/Al2 O3 /Al) and MCR3 (Ag/Al) were
investigated in a fixed bed reactor by coating and treating thin alu-
minum coil material in the same way, the microstructured wafers
were treated. Therefore, two different aluminum supported silver
catalysts were investigated in a fixed bed reactor.
The Ag/Al2 O3 /Al catalytic system used in the microchannel
reactor MCR2 was transferred into a fixed bed reactor by treating a
0.1 mm Al foil (AlMg3) in the same way, the microstructured wafers
were treated. According to the preparation of MCR2, the foils were
anodically oxidized first and then subsequently coated with 800 nm
silver by sputtering. Therefore, the catalytic active surface and the
surface area should be the same. The geometric surface area of this
catalyst denominated FBR2 was 340 cm2 and therefore the same as
in MCR2.
In order to attain better comparability and to exclude aging
effects, an experiment with the fresh microchannel reactor, having
nearly the same initial degree of conversion was chosen as refer-
ence. This time, very similar selectivities and conversion degrees
were obtained, as shown in figure 4.55 and 4.56. Using 20% ethene
in oxygen and a reactor temperature of 503 K, the selectivity was
initially 64.8% at a conversion degree of 11.4%. With increasing de-
gree of conversion, the selectivity decreased nearly steadily to 55.2%
128 Chapter 4. Results

0,66

selectivity to ethylene oxide 0,64

MC R2
0,62

0,60

0,58

0,56

FBR2
0,54

0,10 0,15 0,20 0,25 0,30 0,35

degree of conversion

Figure 4.55: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/Al2 O3 /Al fixed
bed reactor FBR2 and the corresponding microchannel reactor MCR2.
Reaction conditions: 20% C2 H4 in O2 , p= 0.3 MPa, T=503 K.

at a conversion degree of 31.5%. In comparison, the microchannel


reactor MCR2, having the same catalytic active coating, reached
comparable conversion degrees of 15.4% up to 33.7% at selectivities
of initially 63.2%, but the selectivities were different. Despite the
slightly lower selectivities at low, but higher selectivities at high con-
version degrees, the selectivity was nearly constant and close to 63%
for conversion degrees between 15 to 30%. Only at conversion de-
grees above 30%, decreasing selectivities with increasing conversion
degrees were observed.
A similar behavior was observed as the ethene concentration
was lowered to 4% (Fig. 4.56). This time, the microchannel reac-
tor MCR2 was superior to the the corresponding fixed bed reactor
FBR2. In the fixed bed reactor, selectivities of initially 65.8% at a
conversion degree of 31.7% were observed. The selectivity decreased
with increasing conversion degree to 51.3% at a conversion degree
of 68.1%. In contrast, the microchannel reactor exhibited higher
selectivities at the same conversion degree. Initially, the selectivity
4.2. Epoxidation in traditional tube-type fixed bed reactors 129

0,68

0,66
MC R2
selectivity to ethylene oxide

0,64

0,62

0,60
FBR2
0,58

0,56

0,54

0,52

0,50

0,30 0,35 0,40 0,45 0,50 0,55 0,60 0,65 0,70 0,75

degree of conversion

Figure 4.56: Selectivity to ethylene oxide as a function of the conversion


degree (attained by residence time variation) for the Ag/Al2 O3 /Al fixed
bed reactor FBR2 and the corresponding microchannel reactor MCR2.
Reaction conditions: 4% C2 H4 in O2 , p= 0.3 MPa, T=503 K.

was 67% at a conversion degree of 42%. With increasing degree of


conversion, the selectivity decreased initially only slightly to 65.6%
at 60.3% conversion. At higher conversion degrees, notably lower
selectivities of 59.7% at 71.6% conversion degree were observed. By
taking the lowest conversion degree for MCR2 and FBR2 at the same
flow rate, the microchannel reactor MCR2 was clearly more active
than the fixed bed reactor FBR2.

4.2.3 Silver on Al (FBR3, similar to MCR3)


According to the experiments made with FBR1/MCR1 and
FBR2/MCR2, the third catalytic system, namely Ag/Al was
investigated in a fixed bed reactor too. According to the proceeding
performed with the predecessors, aluminum foil having a thickness
of 0.1 mm was treated in the same manner, the microstructured
wafers of MCR3 were treated. Again, the surface area of the
130 Chapter 4. Results

aluminum foil was 340 cm2 and therefore the same as the total
geometric surface area of MCR3. The selectivity / conversion
behavior of this Ag/Al coating in a fixed bed (FBR3), as well as in
the microchannel reactor MCR3 is depicted in figure 4.57.
At low conversion degrees, 60.9% selectivity at a conversion de-
gree of 14.5% was observed for FBR3. With increasing conversion
degree, the selectivity increased slightly with increasing degree of
conversion to 62.9% at a conversion degree of 44.4%. In order to
attain better comparability and to exclude aging effects, an experi-
ment with the fresh microchannel reactor MCR3, having nearly the
same initial degree of conversion was chosen. Initially, a selectivity of
62% at a conversion degree of 14.1% was observed. With increasing
degree of conversion, the selectivity decreased to 57.4% at a conver-
sion degree of 40.1%. Therefore, the microchannel MCR3 showed
initially a higher selectivity, but a very different selectivity / con-

0,64

FB R 3
selectivity to ethylene oxide

0,62

0,60

MCR 3

0,58

0,56

0,10 0,15 0,20 0,25 0,30 0,35 0,40 0,45

degree of conversion

Figure 4.57: Selectivity to ethylene oxide of the Ag/Al fixed bed reac-
tor FBR3 and the corresponding microchannel reactor MCR3 as a func-
tion of the conversion degree. Reaction conditions: 20% C2 H4 in O2 ,
p= 0.3 MPa, T=523 K.
4.2. Epoxidation in traditional tube-type fixed bed reactors 131

version behavior with decreasing selectivities at higher conversion


degrees.
After the experiments at reactor temperatures of 523 K and
below were performed, the reactor temperature was increased to
563 K in order to compare the selectivity / conversion behavior of
both reactor types at high heat production rates. Those experiments
proved to be no problem for the microchannel reactor as shown in
figure 4.26 (p.87) and 4.28 (p.89). In FBR3, the ethene / oxygen
mixture exploded immediately after the feed was switched from O2
flow to 20% ethene in O2 at a total flow rate of each 1100 ml/min
(STP).
After the reactor cooled down to ambient temperatures, the sil-
ver coated aluminum foils were removed from the reactor tube. It
showed that some of the small aluminum foil pieces were molten,
suggesting a severe hot spot having a temperature above the melt-
ing temperature of AlMg3 occurred. A photo of this molten foil is
depicted in figure 4.58. Therefore, no further results at high flow
rates, high reactor temperatures and therefore, at high heat produc-
tion rates are available.

Figure 4.58: Molten


AlMg3 foil based catalyst
after hot spot of the Ag/Al
fixed bed reactor FBR3
after a runaway followed
by an explosion. Reaction
conditions: 20% C2 H4 in
O2 , p= 0.3 MPa, T= 563 K,
v= 1100 ml/min (STP).
132 Chapter 4. Results

4.2.4 Silver supported on -Al2 O3


4.2.4.1 Silver immobilized on -Al2 O3 by impregnation

In the microchannel reactor, two different Ag/-Al2 O3 systems (see


section 4.1.3.2 and 4.1.3.4 had been investigated. According to the
immobilization method performed for MMCR8 (section 4.1.3.2), -
Al2 O3 powder was impregnated with an AgNO3 /lactate solution
to prepare a corresponding fixed bed catalyst. Unfortunately, this
catalyst proved to be practically unusable because of low selectiv-
ities combined with heat transfer problems, which resulted in an
immediate runaway of the reactor (without explosion). Therefore,
no further experiments were performed with this self made catalyst
type due to safety precautions.

4.2.4.2 Use of a commercial SHELL-800 Series, -


Al2 O3 based EO silver catalyst in a fixed bed
reactor (FBR4)

According to the investigations made with the Shell 800 Series


Catalyst catalyst in a microchannel reactor (MMCR9/MMCR10),
similar experiments were performed with the same type of catalyst
in a fixed bed reactor. This catalyst is denoted as FBR4. Due to
a non-analysis and a non-disclosure agreement, there is no data of
this catalyst regarding preparation, surface area, calcination, pre-
treatment, promotion or silver loading available. In the following
experiments, 202 mg of the crushed catalyst (fraction 250 to 500m)
were diluted with 1g of crushed quartz glass having the same corn
fraction and used in the standard fixed bed reactor (see chapter
6.2.1, p. 191).

Side effect of high ethene partial pressures: Due to a mal-


function of the ethene mass flow controller, the catalyst was acci-
dentally exposed to 60% ethene in oxygen at a reactor temperature
of 503 K and 0.3 MPa overnight for approximately 12 hours. The
impact of this malfunction on the selectivity / conversion behavior
is depicted in figure 4.59. The selectivity improved by nearly 4%
at the same degree of conversion, whereas the activity of the cat-
alytic coating decreased by nearly 50%. At the highest flow rate,
4.2. Epoxidation in traditional tube-type fixed bed reactors 133

0,64
catalyst after accidental
treatment with 60% C2H4 in O2
0,63
selectivity to ethylene oxide

0,62

0,61

0,60

0,59 initial catalyst behavior


after the activation process

0,58
0,00 0,02 0,04 0,06 0,08 0,10 0,12 0,14

degree of conversion

Figure 4.59: Selectivity as function of the conversion degree (attained by


residence time variation) for the Shell 800 Series Catalyst in a fixed
bed reactor FBR4 before and after a treatment at high ethene partial
pressures. Reaction conditions: T=483 K, p=0.3 MPa, 20% C2 H4 in O2 .

the conversion degree dropped from 5.3% down to 2.5%. It proved


impossible to reactivate the catalyst even by a reoxidation in O2
at elevated pressures. It showed impossible to reproduce this re-
duction effect in the microchannel reactor, even at elevated reactor
temperatures of up to 543 K. As long as it is not otherwise men-
tioned, all results of the Shell 800 Series Catalyst in the fixed
bed reactor refer to the deactivated / reduced state.

Selectivity / conversion behavior: The selectivity as a function


of the conversion degree was investigated at reactor temperatures
between 483 K and 523 K for ethene partial pressures of 20% in
O2 using the deactivated, but in selectivity improved catalyst. In
order to change the degree of conversion, the GHSV was varied.
Again, the selectivity increased with increasing degree of conversion
for each set of experiments as shown in figure 4.60. At 483 K, the
selectivity improved by 2.8%, at 503 K by 3.7% and at 523 K by
134 Chapter 4. Results

483K 503K
selectivity to ethylene oxide 0,64

0,62
523K

0,60

0,58

0,56

0,54
0,00 0,05 0,10 0,15 0,20

degree of conversion

Figure 4.60: Selectivity as function of the conversion degree (attained


by residence time variation) of the Shell 800 Series Catalyst in a
fixed bed reactor (FBR4) at different temperatures. Reaction conditions:
p=0.3 MPa, 20% C2 H4 in O2 .

5.6% (abs.). After each GHSV variation, the reactor was switched
from minimum throughput (highest conversion degree, lowest heat
production rate) to maximum throughput (lowest conversion degree,
highest heat production rate) to check for potential deactivation. On
this occasion, the reactor had a runaway followed by an explosion of
the C2 H4 / O2 mixture in the cooling traps - directly after switching
between the two flow rates in the last experiment at 523 K. At this
point, the experiments with the fixed bed catalyst were stopped.

Comparison of the fixed bed reactor FBR4 with the mi-


crochannel reactor MMCR10: The selectivity as a function of
the conversion degree had been determined for the Shell 800 Se-
ries Catalyst used in a crushed form within the fixed bed reactor
FBR4 as well as in form of a catalytic active coat on the walls of
four microstructured wafers within the modular microchannel reac-
tor MMCR10 see Fig. 4.61). The flow rates used in these exper-
4.2. Epoxidation in traditional tube-type fixed bed reactors 135

MMCR10, 483K
0,62
FBR4, 483K

MMCR10, 503K
selectivity to ethylene oxide

FBR4, 503K
0,60

0,58

0,56

0,54

0,52

0,05 0,10 0,15 0,20 0,25

degree of conversion

Figure 4.61: Selectivity as function of the conversion degree (attained


by residence time variation) for the Shell 800 Series Catalyst used in
a fixed bed reactor FBR4 and within the modular microchannel reactor
MMCR10. Reaction conditions: p=0.3 MPa, 20% C2 H4 in O2 .

iments varied between 12 and 112 ml/min (STP). For comparison


purposes, the selectivity / conversion behavior of FBR4 was used
before the accidental treatment with high ethene concentrations oc-
curred, which caused the described deactivation effect.
At a reactor temperature of 483 K (Fig. 4.61), the Shell 800
Series Catalyst in the microchannel reactor MMCR10 exhibited
a flat selectivity / conversion behavior with selectivities of initially
61% for conversion degrees up to 11%. At 503 K, the selectivities
were close to 55% for conversion degrees up to 13%. Thus, an
increase in the reactor temperature of 20 K resulted in 6% lower
selectivities. On the other hand, the increased reactor temperature
caused increased conversion degrees. At 483 K, the conversion de-
grees were in a range from 4.2 - 14.9% and increased to 8.1 - 25.3%
at 503 K. The variation of the total flow rate was performed within
the same range in each experiment.
136 Chapter 4. Results

The Shell 800 Series Catalyst (FBR4) showed at both re-


actor temperatures surprisingly increasing selectivities with increas-
ing degree of conversion. At 483 K, selectivities of initially 58.6%
with a nearly linear increase up to 61.8% were obtained. An overall
increase in the reactor temperature to 503 K resulted in lower se-
lectivities of 54.9% up to 57.8%. The conversion degrees were in a
range of 5.4% to 14.1% at 483 K and 6.63% to 17.2% at 503 K.

4.3 Heat management in microchannel


reactors

In the following section, results of observed heat gradients in mi-


crochannel reactors and the influence of the reaction conditions on
those gradients are presented. According to the investigations per-
formed by Westerterp (see section 3.4.1, page 34), stationary temper-
ature gradients and changes of temperature gradients in microchan-
nel reactors were observed by varying the reactor temperature, the
flow rate and the ethene feed concentration. Furthermore, dynamic
temperature gradients were measured by rapid changes of the reac-
tion mixture and the thermal response of the microchannel reactor
and its heating device was monitored as a function of time.
Utilizing the high heat conductivity of the metallic microchan-
nel reactors, low temperature gradients are expected. In order to
observe changes in the temperature profile, high heat production
rates had to be applied. The application of those rates was easily
achieved with the commercially made Ag/Al microchannel reactor
MCR3. This reactor type was equipped with a stainless steel plate
in the middle of the stack, providing 15 drill holes for 0.5 mm ther-
mocouples. Therefore, the temperature gradient along the reactor
axis was accessible. It proved impossible to measure any radial heat
gradient, once a thermocouple was placed in one on the drill holes
(see Fig. 6.4), the temperature was constant.
4.3. Heat management in microchannel reactors 137

4.3.1 Changes of the temperature profile apply-


ing stationary reaction conditions

The stationary temperature profile of the microchannel reactor


MCR3 was monitored at 563 K using 500 ml/min preheated O2 to
get the temperature profiles in absence of a reaction and afterwards
using 500 ml/min, preheated 20% C2 H4 in O2 as feed in presence
of the reaction. The temperature of the preheating tube was
30 K lower than the reactor temperature and not self-tuning, the
temperature in the reactor was adjusted by a fuzzy logic and
self optimizing PID controller9 . Both temperature profiles are
depicted in figure 4.62. In pure O2 and therefore, in absence of

566 O
2
20% 2H4 C in O
2
564
T/K

562
560
1,0
0,5
0,0
T / K

-0,5
-1,0
-1,5
-2,0
0 10 20 30 40 50
pos o
iti n in the wafer stack / mm

inlet = 0 , ou
mm tlet = 50 mm

Figure 4.62: Axial temperature profile within the microchannel reactor


MCR3 with and without heat production by reaction. Reaction condi-
tions: 20% C2 H4 in O2 , T= 563 K, = 200 ms, p= 0.3 MPa, X= 0.22,
S=0.51, Tad = 779 K, P= 11W.

9 A thermocouple mounted in the middle of the stack and attached to the

heating device was used as reference. The heating device always adjusted this
point in the middle of the reactor to the setpoint temperature
138 Chapter 4. Results

any heat production by reaction, the microchannel reactor exhibits


a temperature profile having the highest temperature of 565.4 K
at the inlet and the lowest temperature of 561.6 K directly at the
outlet. Therefore, the temperature profile caused by the heating
device itself is 3.8 K. As soon as the feed was switched to 20%
ethene in oxygen while keeping the flow rate constant, changes
in the temperature profile were observed. At the reactor inlet, a
higher inlet temperature of 565.9 K was observed as well as a lower
outlet temperature of 559.8 K. Therefore, the total temperature
gradient was 6.1 K with chemical heat production present. In
order to compare both gradients and to compensate small drifts of
the total reactor temperature caused by the heating device, each
gradient was normalized to the thermocouple, located in the middle
of the reactor (25 mm from inlet), because the thermocouple used
as the reactor reference temperature for the heating device was
located next to it. Therefore, the difference between both gradients
yields the thermal gradient caused by the heat of reaction. The
results, shown in the lower graph of figure 4.62 indicate, that the
inlet temperature was raised by 0.9 K and the outlet temperature
lowered by 1.6 K due to the exothermic reaction.
In the following contour graphs, this reaction caused temper-
ature gradient was assigned a color in order to illustrate changes
in temperature gradients (reactor inlet on top of Y-axis, outlet at
Y-bottom) as a function of an input function as gas flow rate, tem-
perature or C2 H4 concentration plotted on the X-axis.

4.3.1.1 Temperature gradients caused by changes in the


reactor temperature

Changes in the temperature profile of the microchannel reactor


MCR3 were monitored at a constant ethene concentration of 20%
in O2 , a constant residence time of 200 ms (STP), but varying
the reactor temperature and therefore, the heat generated by the
reaction. For each reactor temperature, the temperature profile
with and without ethene in the feed was measured (at a constant
flow rate), both gradients normalized and the difference of both
gradients calculated as explained before. Therefore, the reaction
caused temperature profile change along the axis was measured and
plotted as a function of the reactor temperature. The corresponding
4.3. Heat management in microchannel reactors 139

800

600
/K

400
ad
T

200

10

5
P/W

460 480 500 520 540 560


T/K

Figure 4.63: Calculated heat production P and adiabatic temperature


rise Tad as a function of the reactor temperature for the microchannel
reactor MCR3. Reaction conditions: 20% C2 H4 in O2 , = 200 ms,
p= 0.3 MPa.

dependence of the calculated heat production by reaction and


the adiabatic temperature rise on the reactor temperature in this
experiment is depicted in figure 4.63.
Applying low reactor temperatures, only little heat was pro-
duced. At the lowest examined reactor temperature of 463 K, the
heat production rate was calculated to be 0.31 W, the adiabatic tem-
perature 4Tad being 23 K. With increasing reactor temperature and
therefore, increasing heat production P and increasing 4Tad , the
temperature profile changed remarkably. At 563 K, a total power of
10.8 W was calculated and the corresponding adiabatic temperature
rise increased to 778 K.
The influence of the heat production on the temperature pro-
file as depicted in Fig. 4.64 was initially at low heat production
rates very small and difficult to measure. At the reactor inlet (top
of Y-axis), the temperature decreased marginally. With increasing
temperature and increasing power such as eventually 563 K, the heat
140 Chapter 4. Results

stack inlet
l=0 mm
-2,0

0,5
-0,5 0 -1,0

-0,5
as flow
ection

T / K
0
0
dir
g

0,5

-0,5 -1,0 1,0

2,0

stack outlet 460 480 500 520 540 560

l=50 mm
T / K

Figure 4.64: Reaction caused changes in the axial temperature profile by


variation of the reactor temperature for the microchannel reactor MCR3.
Reaction conditions: 20% C2 H4 in O2 , = 200 ms, p= 0.3 MPa.

production was 10.8 W and the temperature at the inlet increased


by 0.9 K. At the reactor outlet (bottom of Y-axis), the temperature
decreased by 1.6 K. Therefore, with increasing reactor temperature
and increasing heat production, the inlet temperature increased and
the outlet temperature decreased.
This experiment shows clearly, that increasing heat production
rates by increasing reaction temperature cause increasing tempera-
ture gradients within the microchannel structure.

4.3.1.2 Temperature gradients caused by changes of the


flow rate

In this experiment, the reactor temperature was kept constant, but


the flow rate was varied. Therefore, at low flow rates high conver-
sion degrees and therefore, high adiabatic temperature rises were
4.3. Heat management in microchannel reactors 141

attained, but the total heat production was comparably low. Con-
trary, at high flow rates, low conversion degrees at low adiabatic
temperature rises were attained, but the highest heat production
was observed. The dependence of the calculated adiabatic temper-
ature rise and the heat production on the flow rate is depicted in
figure 4.65.

2500

2000
Tad / K

1500

1000

500

15

10
P/W

0
0 200 400 600 800 1000 1200
flow rate / ml/min (STP)

Figure 4.65: Calculated heat production P and adiabatic temperature


rise Tad as a function of the flow rate for the microchannel reactor MCR3.
Reaction conditions: 20% C2 H4 in O2 , T= 563 K, p= 0.3 MPa.

Applying a flow rate of 80 ml/min, the heat production was


only 5.2 W, but attaining a calculated adiabatic temperature rise
of 2419 K due to the extraordinarily high degree of conversion of
more than 62.3% at a selectivity of 50.1%. With increasing flow
rate of up to 1100 ml/min (STP), the conversion degree decreased
to 16.2%, yielding a selectivity of 51.4%. At this point, the adiabatic
temperature rise decreased to 609 K, but the total heat production
increased to 18.5 W.
Again, changes in the temperature profile were monitored by
comparing the temperature profile of the reactor applying 20%
ethene in oxygen on the one hand and oxygen on the other as feed,
142 Chapter 4. Results

using the same flow rate. After a normalization of the resulting


gradients, changes in the profile were plotted as a function of the
flow rate. The result is depicted in figure 4.66.

stack inlet
l=0 mm
-3,0

-2,0
1,0

-1,0
as flow
ection

T / K
0 0
dir
g

-1,0

1,0

2,0

-2,0

3,0

stack outlet 200 400 600 800 1000

l=50 mm
l
f ow r ate / ml/min (STP)

Figure 4.66: Reaction caused changes in the axial temperature profile by


variation of the flow rate for the microchannel reactor MCR3. Reaction
conditions: 20% C2 H4 in O2 , T= 563 K, p= 0.3 MPa.

At low flow rates, only minor changes in the temperature pro-


file had to be noted. With increasing flow rate and therefore, with
increasing heat production, the temperature at the reactor inlet in-
creased, whereas the temperature at the reactor outlet decreased.
At 1100 ml/min (STP), the inlet temperature increased by 1.95 K,
whereas the outlet temperature decreased by 2.6 K. Therefore, the
reaction induced a temperature gradient of 4.55 K.
This experiment shows clearly, that the gradient is rather re-
lated to the produced power than the adiabatic temperature rise.
4.3. Heat management in microchannel reactors 143

4.3.1.3 Temperature gradients caused by changes in the


ethene feed concentration

Finally, changes in the temperature profile were monitored as a func-


tion of the ethene concentration. This time, the flow rate and the
reactor temperature were kept constant, but the ethene concentra-
tion was varied between 3.8% and 60%, using O2 as balance. The
observed changes in the temperature profile of the reactor are de-
picted in figure 4.68. At the reactor inlet, the highest changes in the
temperature profile were observed as ethene concentrations between
15% and 30% were applied. Within this range, the reactor inlet tem-
perature increased by up to 3.2 K. The application of higher as well
as lower ethene feed concentrations resulted in lower temperature
gradients at the reactor inlet. At the reactor outlet, nearly no influ-
ence on the feed concentration of the temperature profile had to be

stack inlet

l=0 mm

-2,0

-1,0

2,0
0
flow

direction

T / K

1,0
gas

1,0

2,0

-1,0

3,0

4,0
stack outlet 0 10 20 30 40 50 60

l=50 mm
C H concentration / %
2 4

Figure 4.67: Reaction caused changes in the axial temperature profile


by varying ethene concentration for the microchannel reactor MCR3. Re-
action conditions: T= 563 K, = 200 ms, p= 0.3 MPa, balance: O2 .
144 Chapter 4. Results

noted. The adiabatic temperature rise as well as the chemical heat


production as a function of the ethene partial pressure is depicted
in figure ??.
With increasing ethene concentration, the thermal heat pro-
duction increased from 6.3 Watt at 3.8% C2 H4 in O2 to 15.9 W
at 44.5% ethene. A further increase in the ethene concentration to
60% led to a slightly lower heat production of 14.6 W. Similar to
the trend observed with the heat production, the adiabatic temper-
ature rise increased at low ethene concentrations and decreased at
high ethene concentrations with increasing C2 H4 partial pressure.
At 3.8% C2 H4 , the adiabatic temperature rise was 533 K and in-
creased to 1029 K at 30% C2 H4 . Higher ethene concentrations of
up to 60% resulted in decreasing temperature rises of 794 K at 60%
ethene.

stack inlet

l=0 mm

-2,0

-1,0

2,0
0
flow

direction

T / K

1,0
gas

1,0

2,0

-1,0

3,0

4,0
stack outlet 0 10 20 30 40 50 60

l=50 mm
C H concentration / %
2 4

Figure 4.68: Reaction caused changes in the axial temperature profile


by varying ethene concentration for the microchannel reactor MCR3. Re-
action conditions: T= 563 K, = 200 ms, p= 0.3 MPa, balance: O2 .
4.3. Heat management in microchannel reactors 145

4.3.2 Changes of the temperature profile apply-


ing dynamic reaction conditions

Remarkable changes of the reactor temperature profile were also ob-


served, when the microchannel reactor was exposed to rapid changes
of the reaction conditions and therefore, rapid changes of the heat
produced by the reaction. All microchannel reactors used in this
study were built without internal crossflow- or countercurrent cool-
ing. Therefore, the heat had to be removed through the walls of the
reactor to the surrounding metallic structure.
In stationary operation, the sum of heat provided by the heating
device and by reaction is sufficient to compensate the heat losses to
the environment - the temperature stays the same. Rapid changes
of the produced heat implied, that the power of the heating device
had to be adapted quickly in order to prevent overheating of the
device. It was possible to change the flow rates within seconds, but
it proved impossible to adjust the temperature of the heating block
within the same time. Thus, the microchannel reactor was exposed
to temporary higher or lower temperatures before the temperature
controller was able to get back to steady state operation. Having
too little reserves, a sudden start of a reaction at high temperatures
and high throughput could theoretically result in a runaway.
The highest changes in the reactor temperature were monitored,
when the reactor was switched from 100% oxygen flow to 20% C2 H4
in O2 at a flow rate of 1100 ml/min and a temperature of 563 K.
The reactor temperature is defined as the temperature of the PID
controllers thermocouple, which was located in the middle of the
stack right beside the monitoring one. The temperature difference
between the steady state under oxygen flow at t=0 min and the
resulting temperature as a function of time is depicted in figure 4.69
for three different positions of the microchannel reactor, namely at
the reactor inlet, in the middle of the reactor and at the reactor
outlet. By normalizing to the starting temperature at each reading
point, small external temperature gradients caused by the heating
device as shown in Fig. 4.62 were excluded and only the reaction-
caused temperature change left.
At t=1.3 min, the oxygen flow was switched to 20% C2 H4 in
O2 , keeping the flow rate at 1100 ml/min (STP). Due to the high
146 Chapter 4. Results

30

25

20
reactor inlet middle of reactor
T / K

15

10

0
reactor outlet

0 5 10 15 20 25 30 35 40 45 50 55

time / min

Figure 4.69: Temperature difference in the microchannel reactor MCR3


at three different locations as a function of time when switching from 100%
O2 to 20% C2 H4 in O2 . Equilibrium reaction conditions: 20% C2 H4 in
O2 , T= 563 K, v= 1100 ml/min (STP), X=11.4%, S=51.4%, P=13.0W.

heat production of the reactor and the initially constant temperature


of the heated sheathing, the temperature of the reactor increased
rapidly as soon as ethylene was added to the feed. The highest
increase of the reactor temperature to +30 K above reference was
reached at t= 5 min at the reactor inlet. At this point, the temper-
ature rise in the middle of the reactor was 26.9 K and at the end of
the reactor, 20.4 K were observed. This clearly indicates a reaction
caused temperature gradient of nearly 10 K. Ten minutes after the
start of the experiment and five minutes after the peak temperature,
the device began to stabilize approximately 10-15 K above the target
temperature. Minor oscillating effects are seen with a frequency of
approximately 20 minutes10 . Due to the slow adaption of the tem-
perature controller to the new thermal reactor properties, it took
much longer than one hour to adjust the reactor temperature back
10 Half wave time between 15 min and 25 min
4.3. Heat management in microchannel reactors 147

to precisely 563 K.
The opposite effect was observed when the reaction mixture
of 20% C2 H4 was switched to 100% oxygen. Therefore, no more
heat was produced by the reaction and the heated sheathing cooled
down for a short time before the temperature controller increased
the electric power compensating the sudden loss of heat. The results
of this experiment are depicted in figure 4.70. This time, the reactor
temperature decreased rapidly as ethylene was removed from the
feed. At t= 3 min, the reactor temperature at the inlet was down
by 21K, the temperature in the middle of the reactor was down by
19 K and at the end of the reactor, the temperature was down by
14.5 K. After approximately one hour, the reference temperature
was back to 563 K and the temperature controller had successfully
finished its adaption to the new thermal properties of the device.
Applying lower reactor temperatures and therefore, lower heat
production rates, those dynamic temperature gradients of the reac-

reactor outlet
-5
T / K

-10

-15

reactor inlet middle of reactor


-20

0 5 10 15 20 25 30 35

time / min

Figure 4.70: Temperature difference in the microchannel reactor MCR3


as a function of time at three different locations after switching from 20%
C2 H4 in O2 to 100% O2 . Equilibrium reaction conditions: 20% C2 H4 in
O2 , T= 563 K, v= 1100 ml/min (STP), X=11.4%, S=51.4%, P=13.0W.
148 Chapter 4. Results

tor with a rapid change of the reaction conditions decreased. At


523 K and flow rates of 1100 ml/min, the heat production of the
reactor was down to 4.2 Watt, yielding a temperature rise of 7.5 K
at the most (Fig. 4.71). With a decreased flow rate of 110 ml/min,
applying the same ethene concentration and initial reactor temper-
ature, the heat production was down to 1.9 W, yielding a temporal
temperature rise of only 2.6 K at the most (Fig. 4.72).

Figure 4.71: Temperature


8
difference in the microchan-
inlet nel reactor MCR3 as a func-
6 tion of time at three dif-
middle
ferent locations after switch-
outlet ing from 20% C2 H4 in O2
T / K

4 to 100% O2 . Reaction con-


ditions: 20% C2 H4 in O2 ,
2 T= 523 K, v= 1100 ml/min
(STP), X=4.2%, S=56%,
P=4.4 W.
0

0 10 20 30
time / min

3,0 Figure 4.72: Temperature


inlet
difference in the microchan-
2,5
nel reactor MCR3 at three
2,0 middle different locations after
switching from 20% C2 H4
1,5 outlet in O2 to pure O2 . Re-
T / K

action conditions: 20%


1,0 C2 H4 in O2 , T= 523 K,
v= 110 ml/min (STP),
0,5
X=24%, S=57%, P=1.9 W.
0,0

0 10 20
time / min
4.4. Design aspects of modular microchannel reactors 149

4.4 Design aspects of modular mi-


crochannel reactors
In order to compare the two modular microchannel reactor types I
and II (see page 194 for details), the same catalyst was mounted
in two different reactors (MMCR9/MMCR10) and the selectivity /
conversion applying the same reaction conditions was used to eval-
uate both types. It was decided to use the Shell 800 Series Cat-
alyst to compare both reactors.
Therefore, the modular microchannel reactor MMCR9 (type
I) was equipped with six microstructured and Shell 800 Series
Catalyst coated wafers. After the activation procedure, the selec-
tivity / conversion behavior at 483 K, a pressure of 0.3 MPa using
20% C2 H4 in O2 was determined. After this experiment, the reactor
was disassembled, four of the six wafers were reused within the mod-
ular microchannel reactor MMCR10 (type II) and the selectivity /
conversion behavior determined again.
Both curves are depicted in figure 4.73. It is obvious, that the
modular microchannel reactor II exhibits remarkably higher selectiv-
ities at the same degree of conversion than the I type. The reactor
type I containing 6 wafers allowed conversion degrees of 10.3% to
24.5% at selectivities of initially 56.4%, decreasing with increasing
conversion degree to 47.2%. Contrary, the modular microchannel
reactor II, containing 4 microstructured wafers, exhibited lower con-
version degrees ranging between 4.2% to 14.9%. The selectivities
were higher, varying only slightly between 60.0% to 61.1%. Com-
paring the selectivity at the same degree of conversion, the reactor
type II exhibits exactly 5% more selectivity than the older I type.
In order to validate this result, this experiment was repeated us-
ing an Ag/Al catalyst. WEDM made aluminum wafers were catalyt-
ically activated by sputtering, activated under reaction conditions
(same as for MCR3) and investigated subsequently in the modular
microchannel reactors type I and type II (MMCR14/MMCR15, ta-
ble 4.15). The selectivity / conversion behavior of both experiments
is depicted in figure 4.74. Using eight wafers in the microchannel
reactor MMCR14 (type I) as housing resulted in conversion degrees
ranging from 3.35 to 21.4%, showing increasing selectivities of 42.3
to 48.0% with increasing degree of conversion. After mounting 5
150 Chapter 4. Results

0,62

MMCR10
0,60
Type II
selectivity to ethylene oxide
0,58

0,56

0,54

MMCR9
0,52
Type I

0,50

0,48

0,46

0,05 0,10 0,15 0,20 0,25

degree of conver ion s

Figure 4.73: Selectivity as function of the conversion degree (attained by


residence time variation) of the Shell 800 Series Catalyst in the modu-
lar microchannel reactors MMCR9 (type I, using 6 wafers) and MMCR10
(type II, using 4 wafers). Reaction conditions: T=483 K, p=0.3 MPa,
20% C2 H4 in O2 .

0,54

MMCR 15
selectivity to ethylene oxide

Type II
0,52

0,50

MMCR 14
0,48 Type I

0,46

0,44

0,42

0,00 0,05 0,10 0,15 0,20 0,25

degree of conver ion s

Figure 4.74: Selectivity as function of the conversion degree (attained by


residence time variation) for the modular microchannel reactors MMCR14
(type I, 8 wafers) and MMCR15 (type II, using 5 wafers). Reaction con-
ditions: T=523 K, p=0.3 MPa, 20% C2 H4 in O2 .
4.4. Design aspects of modular microchannel reactors 151

Table 4.15: Geometric parameters of the modular microchannel reactors


MMCR14 and MMCR15.

Label MMCR14 MMCR15


Basic reactor type type I type II
Channel width 300 m 300 m
Channel height 700 m 700 m
Channel length 50 mm 50 mm
Number of channels per wafer 14 14
Number of wafers 8 5
Wafer height 1000 m 1000 m
Wafer width 10 mm 10 mm
Wafer length 50 mm 50 mm
Total geometric surface area 112 cm2 70 cm2
Total channel volume 1.17 cm3 0.731 cm3
Total stack volume 4.5 cm3 2.5 cm3
Coating struct. side 2300 nm Ag
unstruct. side 1000nm Ag
Coating method sputtering

of the 8 Ag/Al coated wafers in the modular microchannel reactor


type II, now labeled as MMCR15 and applying the same reaction
conditions, the conversion degree was down to a range of 2.0-12.9%,
showing again increasing selectivities of 45.9% to 54.0%. It is eye-
catching, that the increase in selectivity is higher with type II than
with the old type I reactor. At high flow rates and therefore, low
conversion degrees, the selectivity is initially 42.3% when using the
old type I reactor and 47.2% with the new type II reactor, both
selectivities measured / interpolated at a conversion degree of 3.3%.
Therefore, the selectivity difference is again 5.0%. With decreasing
flow rates and therefore, increasing conversion degrees of 12.1% and
residence time, the selectivity increased to 46.3% in the reactor type
I, whereas 54.0% were observed in the reactor type II. Therefore,
the selectivity difference using one and the same catalyst increased
to 7.7%.
Furthermore, temperature gradients in both reactor types ap-
plying moderate heat production rates of 2-4 Watt were investigated.
In the modular microchannel reactor type I, the temperature was
measured and adjusted in a way, that the product stream leaving the
152 Chapter 4. Results

Table 4.16: Temperature gradients (in K) to reference temperature


(middle, left) within the modular microchannel reactor type II.

heating plate wafer plate cover plate


inlet 1.8 0.3 -6.1
middle left 3.7 0 (Ref) -9
middle right 2.6 0.2 -9.4
outlet 0.5 -2.8 -9.4

wafer stack showed the required temperature. Typically, the temper-


ature between the gas flow emerging the plates and the outer steel
shell, which was girded by electrically heated clamps (see Fig.6.6,
p.197)was between +/- 1.0 K. Therefore, this reactor type exhibited
only minor temperature gradients. The modular microchannel reac-
tor type II consists of a heating plate, the wafer housing plate and
a cover plate. Each plate was equipped with drill holes at the inlet,
in the middle on both sides of the wafers notch and at the outlet
(see experimental section for details). Therefore, the temperature
was measured at 12 different points. The resulting temperature gra-
dients are listed in table 4.16. This table shows clearly, that the
temperature in the heating plate was the highest in the middle close
to the heating cartridge. In the plate containing the wafers, the
temperature was at the inlet and in the middle of the plate nearly
constant, but the outlet temperature with -2.8 K remarkably lower
than the temperature in the middle. The cover plates temperature
was with deviations of -6 to -9.4 K at the most much lower than the
temperature in the middle of the stack. This strong deviation was
lessened to -3 K to -6 K as the whole reactor was thermally insulated
with glass fiber tape. Nevertheless, it showed impossible to achieve
a uniform temperature with gradients below 3 K in this device.
Chapter 5
Discussion

This chapter is divided into four parts. In the first part as a phase-
in for the experimental results, selectivity vs. conversion curves are
computed from published kinetic parameters to allow a re-evaluation
of the obtained results and their consistency with published data.
In the second part, the general catalytic performance of different
silver coatings is summarized.
In the third part, the influence of the reaction conditions on
the catalytic performance of the best coating methods is discussed.
The performance and catalytic properties of different coatings will
be compared, applying ethylene/oxygen concentrations within the
explosion range to reach high degrees of conversion and therefore,
challenging reaction conditions.
In the fourth part of the discussion, reaction engineering aspects
are addressed, with a main focus on reactor and wafer construction.

153
154 Chapter 5. Discussion

5.1 Computation of selectivity vs. con-


version behavior

5.1.1 Computation based on published kinetic


data

According to the kinetic model, constants and equations given by


Schouten [29, 31] (see p. 29), the dependence of selectivity and
conversion degree as a function of the ethene and oxygen partial
pressure was computed for the Berty and tubular reactor. These
results are shown in figure 5.1-5.2. Unfortunately, no clear trend
emerged when both computed selectivity/conversion curves are com-
pared. Depending on the reactor model, the selectivity increases
with increasing ethene concentration when the Berty reactor model
is applied, whereas the selectivity decreases as soon as the tubular
reactor model is used (Fig. 5.1). The selectivities calculated for
the tubular reactor of approximately 15-30% are well below those
calculated for the Berty reactor exhibiting selectivities up to 64%.
In both calculations, the conversion degree decreased with increas-
ing ethene concentration. Again contradictory results are obtained,
when the influence of the oxygen concentration is calculated (Fig.
5.2). With increasing oxygen concentration, the selectivity decreases

0,05 0,35

0,014 0,65

0,012 0,60 0,04 0,30


selectivity to ethene oxide

selectivity to ethene oxide


conversion degree

conversion degree

0,010 0,55

0,03 0,25

0,008 0,50

0,006 0,45 0,02 0,20

0,004 0,40

0,01 0,15

0,002 0,35

Berty-reactor tubular-reactor
0,000 0,30 0,00 0,10
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35

ethene concentration / % ethene concentration / %

Figure 5.1: Calculated influence of the ethene concentration on selectiv-


ity and conversion for a Berty- and tubular-reactor according to a kinetic
model developed by [31]. Reaction conditions: T=503 K, p=0.3 MPa,
balance O2 .
5.1. Computation of selectivity vs. conversion behavior 155

0,10 0,12

0,32
0,09 0,55

0,10
0,08

selectivity to ethene oxide


selectivity to ethene oxide
0,28
0,07 0,50
conversion degree

conversion degree
0,08

0,06
0,24

0,05 0,45 0,06

0,20
0,04

0,04
0,03 0,40
0,16
0,02
0,02

0,01 0,35
tubular-reactor 0,12
Berty-reactor
0,00 0,00
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70

oxygen concentration / % oxygen concentration / %

Figure 5.2: Calculated influence of the oxygen concentration on selectiv-


ity and conversion for a Berty- and tubular-reactor according to a kinetic
model developed by [31]. Reaction conditions: T=503 K, c(C2 H4 ) = 1%,
p=0.3 MPa.

in the Berty reactor, whereas an increasing selectivity is obtained


for the tubular reactor. In both calculations, the conversion degree
shows a maximum close to 10% oxygen. Higher oxygen concentra-
tions lead to decreased conversion degrees in both models. Obvi-
ously, the surface of the catalyst seems to be predominantly covered
with oxygen, causing a decreased ethene coverage and therefore, a
decreased overall reaction rate.
In order to calculate the selectivity / conversion behavior of
the given kinetic equations and parameters, a reactor was simulated
and its selectivity-conversion behavior calculated using a simple fi-
nite element approach. The effect of the reaction products carbon
dioxide, water and ethylene oxide were taken into account as long
as constants for absorption were provided (see table 3.1, p. 29).
The results are depicted in figure 5.3. Again, different trends have
to be noted. In the Berty-reactor, the selectivity decreases from a
reasonably high level with increasing degree of conversion, whereas
in the tubular reactor, the selectivity increases starting from a very
low level. The different trends may be ascribed to the fact that ethy-
lene oxide adsorption was not taken into account within the tubular
reactor model. The low selectivity of the ethylene epoxidation even
at low conversion degrees is in agreement with the published kinetic
parameters, although the set of experiments yielding these kinetic
156 Chapter 5. Discussion

0,55

0,50
selectivity to ethene oxide

0,45 Berty reactor

0,40

0,35

0,30

0,25

tubular reactor
0,20

0,0 0,1 0,2 0,3 0,4 0,5 0,6

s
conver ion degree

Figure 5.3: Calculated selectivity / conversion behavior for a Berty- and


tubular-reactor according to a kinetic model developed by [31]. Reaction
conditions: T=503 K, c(C2 H4 )=2%, c(O2 )=16%, p=0.3 MPa.

constants parameters look doubtful, especially as both curves are


based on the same catalyst but a different reactor type.
Recapitulating the contradictorily influence of basic reaction
parameters like oxygen and ethene partial pressure on selectivity
and conversion it can be concluded, the reactor design seems to
have a high influence on the selectivity and conversion behavior.
This also indicates that the provided kinetic constants are strongly
affected by heat- and/or mass-transfer effects. Thus, a microchannel
reactor having a uniform temperature profile accompanied with a
superior mass-transfer should give a good insight into the ethylene
epoxidation even applying severe and industrially impossible to
handle reaction conditions.
5.1. Computation of selectivity vs. conversion behavior 157

5.1.2 Computation based on a triangular reaction


scheme
Another approach to estimate a likely selectivity / conversion curve
is to take a closer look at the very basic reaction scheme of ethylene
oxide synthesis as depicted in figure 5.4. This approach will help to
estimate potential selectivity vs. conversion curves for high degrees
of conversion.

C2H4O
k2 k3

k1
C2H4 CO2 + H2O

Figure 5.4: Basic reaction scheme of ethylene oxidation [84].

Having this simple scheme of two parallel and one consecutive


reaction, an exemplary selectivity / conversion curve can be cal-
culated. At low conversion degrees, the consecutive combustion is
negligible and the selectivity is directly driven by the rates of re-
action 1 and 2. With initial selectivities of 50%, the initial ratio
of reaction rate one and two must be 1. With increasing degree
of conversion, the consecutive combustion of ethylene oxide (reac-
tion 3) takes place and the selectivity to ethylene oxide decreases.
For simplification purposes, the rate equations used in the following
computations are assumed to follow a first order power law equation.
This approximation is based on three assumptions / observations:

1. Having a large excess of oxygen available, even a second order


reaction turns into a pseudo first order one.

2. The extensive investigation of the reaction kinetics (i.e. [31])


did not allow a really clear discrimination between different
kinetic approaches.

3. The Dutch group [31, 32] assumed a first order reaction for the
consecutive combustion of ethylene oxide
158 Chapter 5. Discussion

k2/k3=10.0
0,50

5.0
0,45

0,40 3.3

0,35
2.0
Selectivity to EO

0,30

1.0
0,25

0.5
0,20

0,15

0,10

0,05

0,00

0,0 0,2 0,4 0,6 0,8 1,0

D egree of Conversion

Figure 5.5: Selectivity vs. conversion plot for the reaction scheme de-
rived from Fig. 5.4 assuming a first order power law kinetic, a ratio of
k1
k2
= 1 and k2
k3
of 10, 5, 3.3, 2, 1 and 0.5 for the rates of the consecutive
combustion.

A theoretical reaction network can be described by the following


three equations, the species A, B and C as a substitute for reactant,
an intermediate and a follow-up product.

cA
= k1 cA k2 cA (5.1)
t
cB
= k2 cA k3 cB (5.2)
t
cC
= k1 cA + k3 cB (5.3)
t
It can be clearly seen (Fig. 5.5), even high rates for the subsequent
combustion of ethylene oxide still allow high selectivities for the in-
termediate product ethylene oxide. As a matter of fact, this simple
model can neither explain any increase in selectivity with increas-
ing degree of conversion, nor it allows any change in the selectivity
/ conversion curve caused by different concentrations of a reactant
A. All reaction rates follow a first order and thus, their conversion
degree is only a function of the residence time. A dependence on
5.2. Catalytic performance of different silver coatings 159

reagent concentration requires at least a second order (power law)


approach. As soon as the selectivity / conversion curve changes
with different ethene concentration, a more complex mechanistic as-
sumption like a Langmuir-Hinshelwood approach, inhibition terms
for certain reaction pathways by reaction products or even different
reaction orders for some reactions are required.

5.2 Catalytic performance of different


silver coatings
The majority of the prepared catalysts were suitable for the partial
oxidation of ethene and allowed high selectivities and yields of ethy-
lene oxide. Nevertheless, silver supported on steel proved to be not
suitable (see chapter 4.1.4), because its initial high selectivity to-
ward ethylene oxide declined rapidly and nearly steadily within the
very first hours of operation. Surprisingly, this coating was initially
highly active and selective for the ethene epoxidation - contrary
to the results observed for other, on the longer term well working
coatings. Those coatings required several dozen hours of activation
under reaction conditions showing increasing conversion degrees be-
fore constant and high conversion degrees and selectivities could be
observed.
Another unsuitable type of coating was silver impregnated alu-
minum oxide (regardless of the immobilization method), prepared by
anodic oxidation of aluminum. Details have been published [100].
Ethylene oxide was observed only at very low conversion degrees in
small amounts. With increasing conversion degree, those coatings
exhibited rapidly decreasing selectivities. It is known, those highly
porous alumina coatings exhibit a comparably high surface area on
the one hand and they are acidic on the other hand [87]. Both fea-
tures are not favorable as long as the isomerization-sensitive ethylene
oxide is present.
However, microchannel reactors with bulk silver, silver de-
posited by PVD, silver impregnated on -Al2 O3 or an immobilized
commercial ethylene oxide catalyst proved to be suitable for ethy-
lene oxide production and showed very similar catalytic properties.
Details will be discussed in the next section.
160 Chapter 5. Discussion

5.3 Influence of the reaction conditions


on catalytic performance
At first, the performance of the different catalytic silver coatings
is summarized. Later, the effect of the reactant concentrations on
selectivity, conversion degree and space-time-yield (STY) will be dis-
cussed. The space-time-yield will be defined as an integral reaction
rate (moles ethylene oxide per second), which is normalized on the
total channel volume of the reactor (in ml) in order to avoid an
influence of the wafers micromachining method and therefore, the
wafers bulk material on this parameter. Thus, the space-time-yield
was calculated according to the following equation:
nC2 H4 X SC2 H4 O
ST Y = (5.4)
Vchannels

5.3.1 Impact of different oxygen concentrations


Increased oxygen concentrations proved their highly positive effect
on the conversion degree with every single coating. Generally, the
oxygen concentration was varied between 5-7% (which is close to
oxygen concentrations used in industrial reactors) up to 80% result-
ing in binary ethylene/oxygen mixtures. As shown in i.e. figure 4.2
for the bulk-silver microchannel reactor MCR1 at low reactor tem-
peratures (503 K), the conversion degree raised from 0.4% to 4% by
factor 10. Similar observations were made with every other well per-
forming catalytic active coating. The microchannel reactor MCR3
as an example for an Ag/Al coating showed an increase in the con-
version degree by factor 3-4, depending on the reactor temperature
(Fig. 7.6). Even the sol-gel prepared and silver impregnated Ag/-
Al2 O3 /Al microchannel reactor MMCR8 behaved similar, yielding
improved conversion degrees by factor 2.5 to 3.4 (Fig. 7.9), depend-
ing slightly on the reactor temperature. Eventually, the promoted
catalysts and coatings like the Shell 800 Series Catalyst and
the Cs promoted microchannel reactor MCR2Cs performed in the
same way. With the Shell 800 Series Catalyst, improved con-
version degrees of factor 3.2 were observed and the MCR2Cs showed
a comparable improvement of the conversion degree by factor 3.3.
The influence of the oxygen concentration on the selectivity to
5.3. Influence of the reaction conditions on catalytic performance161

ethylene oxide is more complicated. At low reactor temperatures


and therefore, low degrees of conversion, only little influence of the
oxygen concentration had to be noted. With increasing reactor tem-
perature, the selectivity became more and more a function of the
oxygen concentration, as shown in Fig. 4.2 (Ag), 7.6 (Ag/Al), 7.17
(Cs promoted Ag/Al2 O3 /Al) and 7.13 (SHELL Ag/-Al2 O3 cata-
lyst). Therefore, high oxygen concentrations are essential as soon as
ethene conversion degrees with favorable high selectivities to ethy-
lene oxide are required. The space-time-yield of the catalyst at a
given flow rate is improved by several hundred percent when the
oxygen concentration is increased. Thus, three conclusions have to
be drawn:

1. The higher the reactor temperature, the higher the gain in con-
version degree as soon as the oxygen concentration is increased
(e.g. Fig. 4.2, 7.2, 7.6, 7.9, 7.13). Generally, improvements of
the conversion degree by several hundred percent in compari-
son to the typical 8-9% O2 have been observed.

2. High oxygen concentrations eventually resulting in binary


C2 H4 /O2 mixtures result in higher selectivities, especially at
high degrees of conversion.

3. Those positive effects of the oxygen concentration were ob-


served with every single type of coating and seemed to be un-
affected by commonly used promoters.

In these investigations, the oxygen partial pressure was typically


0.24 MPa (80% O2 , 0.3 MPa total pressure) and therefore, similar
oxygen partial pressures are obtained at 3 MPa and 8% oxygen.
Thus, in the vast majority of all experiments, the oxygen partial
pressure applied in a microchannel reactor was comparable to the
available oxygen partial pressure in industrial ethylene oxide plants.
The applied peak partial pressures were much higher than those
in commercial reactors. These high oxygen partial pressures of up
to 1.9 MPa (96% O2 at 20 MPa total pressure, Fig. 4.7) are not
available in commercial ethylene oxide plants. Given, the oxygen
concentration in a production plant is approximately 8%, a total
pressure of 23 MPa would be required to attain a similar oxygen
partial pressure.
162 Chapter 5. Discussion

5.3.2 Impact of different ethene concentrations


The influence of the ethene concentration on selectivity and conver-
sion degree is quite simple compared to the more important influ-
ence on the space-time-yield (STY). With every single catalytic ac-
tive coating, the conversion degree decreased with increasing ethene
concentration, whereas the selectivity increased (see Fig. 4.3, 7.3,
7.7, 7.10, 7.14). In order to calculate the influence of the ethene con-
centration on the space time yield, this parameter was calculated for
the Ag/Al microchannel reactor MCR3 (data from figure 4.26, T=
563 K) and the Ag/-Al2 O3 -microchannel reactor MMCR8 (data
from Fig. 7.10, T=483 K) as a function of the ethene concentration.
The results are depicted in figure 5.6. It is obvious, that both types
of coatings exhibit their best catalytic performance in the range of
20 to 45% ethene. The Ag/Al microchannel reactor showed the
highest yields at 40%, whereas the Ag/-Al2 O3 microchannel reac-
tor MMCR8 showed the best performance at about 30% ethene in
oxygen. Both reactors were able to produce close to 5.5 moles ethy-
lene oxide per m3 channel volume. The latter one even despite the
comparably low reactor temperature of 483 K for MMCR8 instead
of 563 K as for MCR3. The much higher activity of MMCR8 in
6,0 6,0

5,0 5,0
STY / mol/ms

STY / mol/ms

4,0 4,0

3,0 3,0

Ag/-Al2O3 (MMCR8) Ag/Al (MCR3)


2,0 2,0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
ethene concentration / % ethene concentration / %

Figure 5.6: Space time yield (STY) of WO as a function of the ethene


concentration for the Ag/Al microchannel reactor MCR3 and the Ag/-
Al2 O3 -microchannel reactor MMCR8. Reaction conditions: MCR3:
T=563 K, v= 500 ml/min, p=0.3 MPa; MMCR8: T=483 K, v= 500
ml/min, p=0.3 MPa; balance O2 .
5.3. Influence of the reaction conditions on catalytic performance163

comparison to the Ag/Al microchannel reactor MCR3 may be par-


tially ascribed to the slightly porous structure of the sol-gel support
material, resulting in an enhanced surface area of the coating. This
ethylene oxide production rate corresponds to an annular production
of approximately 7000 tons ethylene oxide per m3 channel volume.
This demonstrates the potential performance of microchannel reac-
tors and the advantage of being able to handle high heat produc-
tion rates and ethylene / oxygen concentrations within the explosion
range.

5.3.3 Impact of the total pressure

The influence of the total pressure on selectivity and conversion de-


gree, keeping the mass flow rate constant, was investigated for nearly
every single catalyst. In every experiment, initially increasing selec-
tivities accompanied by increasing degrees of conversion were ob-
served. At a certain point between 0.7 to 1.5 MPa, the selectivity
stopped increasing and kept on a nearly constant level (see e.g. Fig.
7.4) or even decreased slightly (see e.g. Fig. 4.7). Two potential
explanations for this behavior will be discussed.
It is generally assumed, that the surface reaction between ab-
sorbed ethene and absorbed oxygen follows a Langmuir-Hinshelwood
mechanism. At very high total pressures and therefore, at high
ethene and oxygen partial pressures, the selectivity was nearly con-
stant but with still increasing degrees of conversion. Caused by
strong absorption of oxygen on silver surfaces and very little ab-
sorption of ethene, this behavior may be ascribed to the oxygen
saturation of silver particles. According to e.g. Gleaves [22] inves-
tigations, this saturation is a crucial factor for the selectivity of the
epoxidation. The assumption, high oxygen partial pressures improve
the saturation of silver particles with surface and subsurface oxygen,
is in agreement with the experimental observations. At higher oxy-
gen partial pressures, the selectivities were constant or even slightly
decreasing. Assuming, that the subsurface of the silver particles is
saturated at high oxygen partial pressure (1.2 MPa O2 partial pres-
sure at ptotal of 1.5 MPa), the surface gets more and more saturated
with oxygen, eventually accelerating the total oxidation in favor of
the selective epoxidation reaction.
164 Chapter 5. Discussion

5.3.4 Selectivity and conversion behavior of dif-


ferent silver catalysts
5.3.4.1 Experimental findings

In order to evaluate the different coating methods applied for the mi-
crochannel reactors, the selectivity / conversion behavior of the fresh
catalysts was used as criterion, applying a temperature of 503 K,
20% ethene in oxygen and a pressure of 0.3 MPa. The results are
shown in figure 5.7, with the data points originating from figures
4.5, 4.21, 4.25, 4.35, 4.41 and 4.51. The selectivities were in a range
of nearly 70% down to 15% and conversion degrees up to 62%. The
highest selectivities were observed for the fresh Ag/Al microchannel
reactor MCR3 and the Cs modified reactor MCR2Cs. The lowest
selectivities were observed for the Ag/-Al2 O3 -microchannel reactor
MMCR8 having selectivities slightly above 50%. The lowest conver-
sion degrees were observed for the bulk-silver reactor MCR1 and the

0,7

0,6
selectivity to ethylene oxide

0,5

0,4

MCR1 (Bulk-Ag)
0,3 MCR2 (Ag/Al O /Al)
2 3

MCR2Cs (Cs modified Ag/Al O /Al)


2 3

0,2 MCR3 (Ag/Al)

MMCR8 (Ag/ -Al O )


2 3

MMCR10 (SHELL Catalyst)


0,1

0,0 0,1 0,2 0,3 0,4 0,5 0,6

degree of conversion

Figure 5.7: Selectivity as a function of the conversion degree for the


microchannel reactors MCR1, MCR2, MCR2Cs, MCR3, MMCR8 and
MMCR10. Reaction conditions: T=503 K, p=0.3 MPa, 20% C2 H4 in O2 .
5.3. Influence of the reaction conditions on catalytic performance165

Ag/Al reactor MCR3, although the latter gained a some activity and
lost selectivity with increasing time on stream (see Fig. 4.31). This
aging effect makes it difficult to compare the Ag/Al based catalysts
MCR2 and MCR3 with each other, because the catalytic proper-
ties depend on the time on stream and treatment of the reactor.
It showed, that the selectivity as well as the conversion degree was
initially constant for hours, as long as the flow rate was kept con-
stant. As soon as the flow rate was decreased and thus, the degree of
conversion increased, selectivity was irreversibly lost (see Fig. 4.12,
page 68). Probably, the reaction by-product water is responsible for
this effect. It is known, that aluminum surfaces are sensitive toward
steam, especially at higher temperatures.
All microchannel reactors had in common, that the selectivity
was nearly independent from the degree of conversion as long as
the degree of conversion was kept low. This initially flat selectiv-
ity/ conversion curve indicates an initially similar reaction order of
the partial and the total oxidation reaction as long as simple tri-
angular scheme (Fig. 5.4) is assumed. Therefore, with increasing
reaction progress and decreasing ethylene concentrations, the ratio
of both rates stays similar. With further progress, remarkable selec-
tivity losses are observed. One of the highest degrees of conversion
applying 20% C2 H4 in O2 of 62.4% was observed for the unpromoted
and unmodified Ag/-Al2 O3 catalyst MMCR8 as depicted in figure
4.35.
It can be clearly seen, the self made Ag/-Al2 O3 microchannel
reactor MMCR8 exhibited slightly lower selectivities at the same
degree of conversion than the commercial made Shell 800 Series
Catalyst in the microchannel reactor MMCR10. Concerning the
space time yield, this commercial catalyst is even inferior. At a
conversion degree of 8.1%, a reactor temperature of 503 K and a
pressure of 0.3 MPa, the Shell 800 Series Catalyst in the mi-
crochannel reactor MMCR10 yielded 2.12 mol/m3 s EO, whereas a
space-time-yield of 12.45 mol/m3 s EO was calculated for MMCR8,
adjusting the same pressure, temperature and conversion degree. It
cannot be excluded, that the SHELL catalyst performs better in
presence of chlorine containing modifiers, for which it was intention-
ally designed.
Decreasing selectivities with increasing degree of conversion
166 Chapter 5. Discussion

cannot be ascribed to a lack of oxygen or oxygen partial pressure.


Having 62.4% conversion degree and 15.1% selectivity left (MMCR8,
Fig. 5.7), about 2 vol% ethylene were formed, consuming 1 vol%
O2 . The dominant total oxidation consumed 10.8% of the available
ethylene, requiring close to 33 vol% oxygen. With 80% oxygen in
the feed, only 41% of the available O2 was converted. Therefore,
lack of oxygen is an unlikely explanation for this remarkable loss of
selectivity, especially since O2 concentrations of 40 to 50% resulted
in still good selectivities (see Fig. 7.9). Very similar calculations can
be performed for Fig. 4.36. This graph shows clearly, that losses
in selectivity (with the same degree of conversion) are increasing
with ethene concentration having still sufficient amounts of oxygen
available.

5.3.4.2 Computations

Some very basic approaches for a better kinetic understanding of the


reaction mechanism will be tested for their validity. The selectivity
vs. conversion plots looking like a simple triangular reaction scheme
as computed (Fig. 5.3) may be capable to describe the kinetics
of this reaction. In the following, the reagents C2 H4 and O2 are
denominated as A1 and A2, the intermediate product EO as B and
the total combustion products H2 O and CO2 as each C (equimolar).
The direct oxidation of C2 H4 will be described by reaction 1 (r1), the
EO formation by reaction 2 (r2) and the consecutive EO combustion
by reaction 3 (r3).
The observation of an initially constant selectivity / conversion
behavior is in agreement with the assumption of having the same re-
action orders for the selective and total oxidation of ethylene. This
is experimentally confirmed with practically close to constant se-
lectivities at low conversion degrees. The calculations for oxygen
consumption (see above) suggest a large excess of oxygen. Thus,
a pseudo first order kinetic for the selective and total oxidation is
again in agreement with the experimental observations, although a
different kinetic approach is likely and cannot be excluded.
Contrarily to a plain first-order approach for all three reactions,
a dependency of the selectivity / conversion curve on the initial
ethene concentration was found. The selectivity to EO as inter-
5.3. Influence of the reaction conditions on catalytic performance167

mediate product B (keeping the same degree of conversion) was the


lower, the higher the reactant concentration (A1 and A2) was chosen
and thus, the higher the concentration of the resulting intermediate
product got1 . Therefore, one of the reaction products seems to cat-
alyze or support the follow up reaction of B (combustion of ethylene
oxide) and a textbook first order kinetic for the total combustion
reaction path has to be discarded. Theoretically, there are three
species formed:

1. Ethylene oxide (B). An impact of this species on the selectivity


would be possible, if adsorbed EO would react and combust
with gaseous EO similar to an Eley-Rideal mechanism of one
absorbed and one gaseous species or if two absorbed ethylene
oxide species would react with each other. This option would
be described by the following equations, using a plain second
order reaction scheme for this approach (Model 1):

cA1
= k1 cA1 cA2 k2 cA1 cA2 (5.5)
t
cB
= k2 cA1 cA2 k3 cA2 cB 2 (5.6)
t
cC
= k1 cA1 cA2 + k3 cA2 cB 2 (5.7)
t
In this approach the decomposition rate is proportional the
second power of the ethylene oxide concentration.

2. Water (C). Water may absorb on alumina, forming Lewis-


acidic sites and these catalyze an isomerization of ethylene
oxide to acetaldehyde in a rate determining step. This compo-
nent is oxidation sensitive and will rapidly subsequently com-
bust (Model 2):

cA1
= k1 cA1 cA2 k2 cA1 cA2 (5.8)
t
cB
= k2 cA1 cA2 k3 cB cC cA2 (5.9)
t
cC
= k1 cA1 cA2 + k3 cB cC cA2 (5.10)
t
1 see e.g. Fig. 4.36, p.99
168 Chapter 5. Discussion

in this approach, the decomposition rate is proportional to the


ethylene oxide concentration and proportional to the concen-
tration of water.

3. Carbon dioxide (C). The concentration of carbon dioxide is


the same as the concentration of water. This is much more a
theoretical option than a practical one. As already mentioned,
CO2 is considered to suppress the total oxidation rather than
to boost the consecutive combustion.

The estimation of the initial ratio for k1 and k2 is straightfor-


ward, because the ratio must reflect the selectivity at low conversion
degrees. k3 was computed by using the experimental data from fig-
ure 4.36 and aiming for a good fitting quality. Therefore, the rate
constant of the consecutive reaction k3 was chosen to match at least
one of these curves as good as possible. If such a simple kinetic ap-
proach model is assumed to be valid, a good fitting quality for every
single concentration has to be expected. The results of this compu-
tation are depicted in figure 5.8 and 5.9. It is obvious, that the fitting
quality for both schemes is good at low ethene concentrations of 4%
/ 20% and the predicted conversion/ selectivity pairs get much less
precise the higher the ethene concentration and the higher the de-
gree of conversion gets. Applying ethene concentrations of 20% and
a conversion degree of 55%, 30% selectivity to ethylene oxide were
observed. Model 1 suggests a computed selectivity of 38%, whereas
model 2 suggests 34%. With even higher conversion degrees, the de-
viation between calculation and experiment increases, both models
predict much more EO than observed in the experiments. Neverthe-
less, model 2 seems to be more suitable than model 1, although both
models practically fail to predict the selectivity/ conversion curve at
40% and 60% ethene properly.
In order to improve the prediction quality and to take deviations
at high ethylene concentrations as well as high degrees of conversion
into account, model 2 was modified. It may be assumed, the rate
for the total combustion of ethylene oxide is not necessarily a func-
tion of the oxygen concentration. This can be explained as soon
as the adsorption and isomerization of ethylene oxide on an acidic
site formed by water is considered to be the rate determining step
for this reaction. Therefore, the modified approach denominated as
5.3. Influence of the reaction conditions on catalytic performance169

0,6
0,5
selectivity to ethylene oxide

0,4
0,3
calc. / exp. data

0,2 / 4% 2H4
C

/ 20% 2H4 C

40% 2H4
0,1
/ C

/ 60% 2H4 C

bal. O
2
0,0
0,0 0,2 0,4 0,6 0,8 1,0
s
degree of conver ion

Figure 5.8: Model1: Calculated and measured selectivity as a function


of the conversion degree at different C2 H4 conc. in O2 . Calculation per-
formed for model 1 with k1=1.1, k2=45, k3=1; experimental data taken
from figure 4.36.

0,6
0,5
selectivity to ethylene oxide

0,4
0,3
calc. / exp. data

0,2 / 4% 2H4
C

/ 20% 2H4
C

0,1 / 40% 2H4


C

/ 60% 2H4
C

0,0 2
bal. O

0,0 0,2 0,4 0,6 0,8 1,0


s
degree of conver ion

Figure 5.9: Model 2: Calculated and measured selectivity as a function


of the conversion degree for different C2 H4 conc. in O2 . Calculation
performed for model 2 with k1=1.1, k2=23, k3=1; experimental data
taken from figure 4.36.
170 Chapter 5. Discussion

model 3 has to be described as follows:


cA1
= k1 cA1 cA2 k2 cA1 cA2 (5.11)
t
cB
= k2 cA1 cA2 k3 cB cC (5.12)
t
cC
= k1 cA1 cA2 + k3 cB cC (5.13)
t
This approach will favor the total oxidation as soon as the oxy-
gen concentration gets low, because the total oxidation of EO is now
independent from the oxygen concentration. The lower the amount
of oxygen, the slower will be the EO production and the more water
produced, the faster will be the ethylene oxide consumption. This
effect will decrease the selectivity at high degrees of conversion and
high C2 H4 partial pressures.
As expected, model 3 shows a better prediction of the selectivity
vs. conversion curves for high ethene partial pressures and high

0,6 calc. / exp. data

/ 4% 2H4 C / 40% 2H4C

20% 2H4 60% 2H4


0,5
/ C / C
selectivity to ethylene oxide

b 2
al. O

0,4
0,3
0,2
0,1
0,0
0,0 0,2 0,4 0,6 0,8 1,0
degree of conver ion s

Figure 5.10: Model 3: Calculated and measured selectivity as a function


of the conversion degree for different C2 H4 conc. in O2 . Calculation
performed for model 3 with k1=1.1, k2=18, k3=1; experimental data
taken from figure 4.36.
5.3. Influence of the reaction conditions on catalytic performance171

conversion degrees (see Fig. 5.10). Applying ethene concentrations


of 4% in O2 , the prediction is close to a perfect match. At 20%
and conversion degrees up to 62%, the prediction is at the most 5%
higher than the experimental data and it follows the experimental
data points with only a few percent error. The fitting quality for 40%
and especially 60% is still not satisfying, but remarkably better than
for model 1 and 2. Therefore, this approach seems to describe the
observations made for MMCR8, but still showing some mismatch as
soon as the C2 H4 concentration as well as the degree of conversion
gets high. The simulated selectivities are higher than the measured
ones.
In order check oxygen being the source of mismatch, the selec-
tivity and conversion degrees were calculated for a constant residence
time and a constant ethene feed concentration, but with decreasing
oxygen concentration. The experimental data for this set is depicted
in figure 7.9. If this kinetic model 3 and the reaction scheme is really
applicable for this reaction, a high quality of predicted and measured
selectivities and conversion degrees has to be expected.
Unfortunately, the simple kinetic model fails as soon as the
oxygen concentration decreases as shown in figure 5.11. Using this
model, decreasing degrees of conversion cause lower concentrations
of water and therefore, the selectivity is expected to increase with
decreasing oxygen concentration as it can be seen in Fig. 5.11. As
known from every single investigated ethylene oxide catalyst, re-
markable selectivity losses are expected as soon as the initial oxygen
concentration is lowered. Therefore, the fitted data to the modified
triangular scheme looks like a good description and prediction of the
experimental data, but the prediction of selectivity and conversion
degrees relies more on a coincidence rather than on a proven model.
As already pointed out, subsurface oxygen plays an important
role in the selective oxidation of ethene. In every single (simple)
model following a triangular scheme, the formation of subsurface
oxygen is not of importance. With increasing ethylene concentra-
tion and high conversion degrees, the demand for reactive and se-
lective oxygen species increases. Therefore, a potential explanation
for this lack of fitting quality may be a lack of reactive and selective
(subsurface) oxygen species. With increasing demand, the diffusion
rate of oxygen into the silver particles regenerating subsurface oxy-
172 Chapter 5. Discussion

calculated
selectivity to ethylene oxide
0,5

0,4

experimental data
0,3

0,2

0,3
degree of conversion

experimental data

0,2

calculated
0,1

0,0
0 10 20 30 40 50 60 70 80 90
oxygen concentration / %

Figure 5.11: Calculated and measured selectivity as a function of the


conversion degree. Calculation performed for model 3 with k1=1.1,
k2=18, k3=1; experimental data taken from figure 4.36.

gen may be too low to compensate the oxygen consumption from


the silver surface by total oxidation. Therefore, the concentration
of subsurface oxygen decreases and without this species, the selec-
tivity decreases faster than expected. An experimental evidence for
this theory may be seen again in figure 4.37. In this graph, the
ethylene oxide concentration is plotted versus the residence time. It
could be observed, that the concentration was initially nearly inde-
pendent from the concentration of ethene at ethene concentrations
of 20% and higher. This is partially a contradiction to the former
made assumptions with the reaction rate of the the total oxidation
r1 and selective oxidation r2 being proportional to the ethene con-
centration as stated in [31]. If this assumption would be correct, the
concentration of EO should be directly proportional to the ethylene
concentration. This observation indicates, a more complex model
has to be applied as soon as high degrees of conversion are adjusted.
Nevertheless, this observation suggests a crucial species seems to
limit the formation of ethylene oxide and the availability of subsur-
face oxygen may be restricted.
5.3. Influence of the reaction conditions on catalytic performance173

5.3.5 Calculation of activation energies

Activation energies of unpromoted silver surfaces Exper-


iments with varying reactor temperatures, keeping the concentra-
tions and flow rates of the feed constant, were performed in order
to allow a calculation of the activation energy of this reaction. As
long as the conversion degree did not exceed 10-15% at the most,
kinetically differential conditions may be assumed. Therefore, the
reaction rate (moles/sec) for the ethylene oxide and carbon dioxide
formation was calculated, assuming a simple parallel reaction scheme
without subsequent total oxidation of ethylene oxide. The results
of this calculation are depicted in figure 5.12 for the unpromoted
microchannel reactors MCR1, MCR2, MCR3 and MMCR8. It is
evident, plotting the logarithmic reaction rates versus 1/T resulted
in very good correlation coefficient of typically 0.99 and better (not
shown). The activation energy of the epoxidation was calculated ac-
cording to the Arrhenius-equation and stayed within a narrow range
of 58.8 kJ/mol to 63.6 kJ/mol. The activation energy for the to-
tal oxidation of ethene was calculated to 73.3 up to 76.8 kJ/mol
and again, no remarkable dependence of this energy on the coating
method was found.

Activation energy of promoted silver surfaces In order to


evaluate the influence of promoters on the activation energy of the
selective and total oxidation of ethene in microchannel reactors, both
parameters were calculated for the Cs modified MCR2 (MCR2Cs)
and the highly, but with unknown metals promoted Shell 800 Se-
ries Catalyst used in the microchannel reactor MMCR10 (Fig.
5.13). The activation energy for the selective oxidation of ethene
was 59.6 kJ/mol for the MCR2Cs and thus, nearly exactly the same
as for MCR2. The activation energy for the total combustion was
71.4 kJ/mol and thus, approx. 5 kJ/mol lower than before the
CS treatment. Therefore, Cs seems to decrease the reaction rate
of the total oxidation rather than enhancing the selective oxida-
tion. The commercial Shell 800 Series Catalyst exhibited with
49.1 kJ/mol clearly the lowest activation energy of all investigated
catalytic coatings (nearly 10 kJ/mol lower). Contrary, the activa-
tion energy for the total oxidation was with 73 kJ/mol comparable
with other coatings.
174 Chapter 5. Discussion

-13
Epoxidation -13,0 Epoxidation
Ea = 60.3 kJ/mol Ea = 58.8 kJ/mol
+/- 1.4 kJ/mol (95%)
-14 -13,5 +/- 1.1 kJ/mol (95%)

-14,0
ln(r) / a.u.

ln(r) / a.u.
-15
-14,5
Total Combustion
Ea = 76.5 kJ/mol -15,0 Total Combustion
+/- 1.7 kJ/mol (95%)
-16 Ea = 73.4 kJ/mol
-15,5 +/- 1.2 kJ/mol (95%)

MCR2 - Ag/Al O /Al

-17 -16,0
2 3 MCR3 - Ag/Al

0,0020 0,0021 0,0022 0,0023 0,0019 0,0020 0,0021 0,0022


1/T / K
- 1 1/T / K
1
-

-14,0 Epoxidation -14,5 Epoxidation


Ea = 61.4 kJ/mol Ea = 63.63 kJ/mol
+/- 1.5 kJ/mol (95%)
+/- 1.0 kJ/mol (95%)

-14,5 -15,0
ln(r) / a.u.

ln(r) / a.u.

-15,0
-15,5
Total Combustion Total Combustion
Ea = 76.8 kJ/mol
-15,5 Ea = 73.3 kJ/mol +/- 0.8 kJ/mol (95%)
+/- 1.5 kJ/mol (95%)
-16,0
MCR1 - bulk Ag MMCR8 - Ag/ -Al O
-16,0 2 3

0,0019 0,0020 0,0021 0,00200 0,00205 0,00210 0,00215


1/T / K
-1
1/T /
-1
K

Figure 5.12: Logarithmic reaction rates as a function of the (inverse)


reactor temperature as well as calculated activation energies for the
ethene epoxidation and total combustion in the microchannel reactors
MCR1 (Ag), MCR2 (Ag/Al2 O3 /Al), MCR3 (Ag/Al) and MMCR8 (Ag/-
Al2 O3 ). Reaction conditions: See Fig. 4.8, 4.22, 4.30 and 4.38.

The decrease of the reaction rate for total oxidation seems to


originate from a suppression of acidic sites resulting from bare alu-
mina surfaces. An indication for this assumption is the selectivity
gain after Cs-treatment of MMCR12 as depicted in Fig. 4.49. The
non uniform distribution of Cs in this channel suggests a not homo-
geneous silver distribution. It is likely, one of the +/- 45o C silver
PVD coatings failed (see table 4.14) leaving roughly 50% of the
channel surface coated with only small amounts of silver originating
5.3. Influence of the reaction conditions on catalytic performance175

-13 -12,0

Epoxidation
Epoxidation
E = 49.1 kJ/mol
a
E = 59.6 kJ/mol
a
+/- 4.0 kJ/mol (95%)
+/- 0.8 kJ/mol (95%)
-14

-12,5

ln(r) / a.u.
ln(r) / a.u.

-15

-13,0

Total Combustion
-16 Total Combustion
E = 73.0 kJ/mol
E = 71.4 kJ/mol a
a

+/- 4.6 kJ/mol


+/- 1.3 kJ/mol (95%)
-13,5
(95%)

MCR2Cs MMCR10 (SHELL)


-17

0,0020 0,0021 0,0022 0,0023 0,0019 0,0020 0,0021

-1
1/T / K
-1
1/T / K

Figure 5.13: Logarithmic reaction rates and calculated activation ener-


gies for the ethene epoxidation and total combustion in the microchan-
nel reactors MCR2Cs (Ag/Al2 O3 /Al) and MMCR10 (Shell 800 Series
Catalyst) as a function of the (inverse) reactor temperature. Reaction
conditions: See Fig. 4.52 and Fig. 4.42.

from the initial silver sputtering. Due to the color of bare aluminum
and silver, such an incident would be unnoticed during preparation.
Such an issue could furthermore explain the lower than anticipated
selectivity of this catalyst (MMCR12, only 55% at 503 K instead of
approx. 60% as anticipated and suggested by the MCR2 and MCR3
results). Exposing the reaction mixture to large areas of bare alu-
mina, selectivity losses caused by isomerization / combustion are
likely. With Cs decreasing the acidity and number of Al2 O3 -sites,
a selectivity gain of 18% and peak selectivities close to 70% were
observed (Fig. 4.48). Contrary, the treatment of MCR2 with Cs
resulted in selectivities of 66% in comparison of initially observed
60% (Fig. 4.21). Therefore, the MCR2 catalyst gained only 6% by
this Cs-treatment but both catalysts reached nearly the same level.
Therefore, an alkali-treatment seems to lower the total oxidation ac-
tivity of Lewis acidic sites on Al2 O3 such that both types of reactors
and catalysts reached a similar performance level despite MMCR12
had a (likely) defect of its coating. The degree of conversion was
176 Chapter 5. Discussion

nearly unaffected by this treatment. The aged MCR2 showed 10.6%


conversion at a selectivity 60.3% (483 K, 20% C2 H4 in O2 , Fig. 4.23,
p.82). Applying the same flow rate and reaction conditions, the se-
lectivity was up by 9% to to 69.3% after Cs treatment at a nearly
unchanged conversion degree of 10.1% (Fig. 4.51, p.122). There-
fore, the influence of Cs on the catalytic properties of silver particles
seems to be limited.
A very different result was obtained for the commercial Shell
800 Series Catalyst. The activation energy for the total oxi-
dation was calculated to 73.0 kJ/mol and therefore comparable to
every other investigated silver catalyst. The activation energy for
the selective oxidation was calculated to 49.1 kJ/mol and approxi-
mately 10 kJ/mol lower than for every other investigated catalytic
surface. Therefore, the good selectivity of this commercial catalyst
especially at low temperatures is based on an enhancement of the
selective oxidation, whereas the activation energy of the competing
total combustion is similar to conventional catalysts.
These findings for the activation energies of the commercial cat-
alyst are consistent with results reported for a similar commercial
catalyst used by [31]. Their kinetic experiments applying very low
conversion degrees suggested activation energies of 48.1 kJ/mol for
the epoxidation and 76.5 kJ/mol for the total combustion. Con-
trary, Hall and Neubauer [21] found a lower activation energy of 35
kJ/mol for total oxidation and epoxidation single crystal silver cata-
lysts and applying vacuum conditions. Generally, activation energies
of 35 kJ/mol up to 90 kJ/mol were reported for the epoxidation and
a range of 35 kJ/mol to 122 kJ/mol for the total oxidation [85].

5.4 Heat effects and reaction engineering


aspects

5.4.1 Temperature profiles and heat effects in mi-


crochannel reactors
Temperature profiles of an industrial multi-tube type reactor were
already described in the theory section (see p.34ff). The influence
of the wall temperature, flow rate and ethene concentration had
5.4. Heat effects and reaction engineering aspects 177

been investigated. The most important finding was, that the hot-
spot temperature is strongly correlated with the conversion degree
and not with absolute heat production of the reactor. Increased wall
temperature as well as increased hydrocarbon concentration resulted
in higher hot-spot temperatures.
As shown in chapter 4.3 (see p.136ff), similar investigations were
performed for the microchannel reactor MCR3.
Applying only O2 gas flow in absence of C2 H4 , a temperature
profile had to be noted along the reactor axis as shown in figure
4.62. This is likely caused by the outer electric heating device, which
seems to show a non uniform heat distribution although it was made
of aluminum because of its high heat conductivity. Nevertheless,
there must be a contact area between the reactors housing and the
heating device with its cartridges and minor differences in local heat
conductivities caused by non uniform clearance between the two
devices may result in small but measurable heat transfer issues.
After switching to an ethylene/oxygen mixture, the tempera-
ture profile changed slightly. The difference between the maximum
and minimum temperature in absence of a reaction was approx-
imately 3.8 K, with this highly exothermic reaction the gradient
increased to 6.1 K. Furthermore, it was observed that in presence
of an exothermic reaction, the temperature at the reactor inlet in-
creased and the temperature at the outlet decreased. This can be
understood, as soon as the behavior of electrical heating support
is taken into account. The reactor temperature is measured by a
thermocouple located in the middle of the reactor and a self opti-
mizing PID controller adjusts the electrical energy in the heating
cartridges in a way that the temperature at this point is exactly as
adjusted. In presence of an exothermic reaction, fewer electrical en-
ergy is required to keep the reactor on temperature and the gradient
caused by the exothermic chemistry evolves. The reactors inlet with
the highest reaction rate may be therefore slightly hotter than be-
fore and the reactors outlet with a lower reaction rate gets slightly
cooler. Nevertheless, this gradient of 6K at the most is despite the
high heat production per reactor volume by far lower than that of
the industrial reactor tube, having gradients of up to 50 K.
In order to compare microchannel with fixed bed reactors, the
gradient across the reactor is therefore used to characterize the mi-
178 Chapter 5. Discussion

crochannel reactor thermally whereas the hot spot temperature is


used for thermal characterization of the tube reactor. It has to be
kept in mind, that roughly 50% of the microchannel reactors overall
gradient is caused by an inhomogeneous heating.
The influence of the hydrocarbon concentration on the tempera-
ture profile of each device is depicted in Fig. 3.8 (p.40) and Fig. 4.68
(p.144). It can be clearly seen, that the higher the ethene concentra-
tion, the higher the hot spot temperature in the industrial tube type
reactor. Using a wall temperature of 465 K, the hot spot tempera-
ture was roughly 530 K at 0.9% C2 H4 (in air). The microchannel
reactor MCR3 showed only minor changes of the temperature profile,
having an extra gradient of roughly 1K as hot spot at the reactors
inlet. Therefore, this reactor type is practically insensitive towards
changes of the feed composition as expected.
Similar observations were made for changes of the mass flow
rate. The influence of this parameter on the temperature profile was
investigated for both reactor types and again, interesting differences
can be easily worked out. Having a look at the axial temperature
profile of the tube type reactor (Fig. 3.6, p.38), the hot spot gets the
higher the lower the adjusted flow rate is. At the lowest investigated
flow rate, the hot spot is roughly 35 K above the wall temperature
compared with 10 K at a 10 times higher flow rate. A complete
contrary behavior was observed for the microchannel reactor MCR3
as depicted in Fig. 4.66 (p.142). With increasing flow rate, the tem-
perature of the reactors inlet is practically constant and only the
temperature of the reactors outlet decreases with increasing flow
rate (and therefore heat production rate). Thus, the hot spot tem-
perature of a conventional tube type reactor is very strong correlated
with the adiabatic temperature rise (see Fig.3.5, p.37 for details),
whereas the temperature gradient across a microchannel reactor is
only a function of the heat production rate of the device. Further-
more, the tube type reactor showed a shift of the hot spot from the
reactor inlet to the middle of the reactor with increasing flow rate.
This phenomenon was unnoticed in the microchannel reactor, which
exhibited only some minor changes to the temperature profile.
Changes in the reactor temperature have little effect on the
temperature profile of the microchannel reactor as shown in Fig. 4.64
(p.140). With an increase of the reactor temperature from 463 K
5.4. Heat effects and reaction engineering aspects 179

to 563 K, the reaction caused temperature gradient increased from


nearly +/-0 K to 2.6 K. The corresponding experiment in the tube
type reactor showed a hot spot temperature of approximately 8 K
for wall temperatures of 435 K up to 50 K for wall temperatures of
475 K (Fig.3.4, p.36). Again, there are very large gradients within
the tube reactor and negligible gradients within the microchannel
reactor.
As a consequence, every single microchannel reactor was insen-
sitive towards sudden changes of flow rates and/or reagent concen-
trations. Every single incident in tube type reactors was observed as
soon as the flow rate was changed, preferably when changing from
low flow rates back to high flow rates. An explanation for this in-
stability may be result from a conclusion, drawn from the industrial
fixed bed reactor. Due to the comparably poor heat transfer between
the catalyst particle and the heat exchanger/reactor wall, these re-
actors exhibit a partially adiabatic behavior as shown in figure 3.5.
At low flow rates and high conversion degrees, the overall heat pro-
duction rate is low, but the spot temperature is very high. As soon
as the flow rate is increased, a still hot particle may encounter in-
creased flow rates and therefore, increased overall heat production.
This may result in such a runaway.

5.4.2 Reactor construction

Many microchannel reactors showed an unexpected selectiv-


ity/conversion behavior. With increasing degree of conversion
and increasing residence time, the selectivity/conversion curves
for different temperatures had crossovers (e.g. Fig. 4.5, p. 59),
which can not be explained by textbook kinetic approaches. This
unexpected behavior is most likely caused by a total oxidation
activity of the reactor material (especially within the diffusors)
towards the oxidation sensitive ethylene oxide. Therefore, at high
residence times ethylene oxide was combusted in the post catalytic
zone and some selectivity lost. At higher reactor temperatures and
higher total flow rate (keeping the degree of conversion constant),
the residence time in the post catalytic zone was therefore lower,
resulting in less combustion and higher overall selectivities. This
theory is supported by observations made with the modular mi-
180 Chapter 5. Discussion

crochannel reactors MMCR type I and type II (see Fig. 4.73). The
reactor type II is designed for a low diffusor volume, providing very
low residence times in the post catalytic zone. This type proved
to be more selective (by 5%) when compared to the older MMCR
type I with its large diffusors using the same flow rate in both
devices. Therefore, post catalytic total oxidation of ethylene oxide
has to be assumed in every single reactor having large diffusors.
Similar observations were found by co-workers [86]. Cs salts seem
to suppress even total oxidation activity of stainless steel diffusors.
When comparing the selectivity/conversion curves for MCR2 and
MCR2Cs (Fig. 4.21 and 4.51), the previously observed crossover
between curves for 483 K and 503 K is surprisingly missing after Cs
treatment.
The modular reactor type II showed much higher temperature
gradients due to its construction and especially at the bottom of
the device with its attached heating cartridge. In comparison to the
MMCR type I with its heating device enclosing the whole reactor
housing, temperature gradients of several Kelvin are unnecessarily
high. In order to decrease the temperature gradients it would be
advisable to attach a second heating device to the reactor, preferably
on the top side. Therefore, the wafers would be heated from the top
and bottom simultaneously. Such an improvement would result in
a reactor having low temperature gradients and a low post-catalytic
volume at the same time. Thus, the advantages of both modular
reactor designs could be combined in a single device if these reactors
types are used in further investigations.

5.4.3 Performance parameters of fixed-bed and


microchannel reactors

In the present work, several tube type and microchannel reactors


have been tested using the same or similar prepared catalysts. Thus,
the only major difference between both kinds of reactors is the re-
actor itself, the catalyst size, form and arrangement.
5.4. Heat effects and reaction engineering aspects 181

5.4.3.1 Comparison of tube type and corresponding mi-


crochannel reactors (FBR1-3, MCR1-3)

MCR1/FBR1: At the beginning of the present work, the likely


most simple catalyst was used - bulk silver. Such a catalyst was
used in the fixed bed reactors FBR1a and FBR1b as well as in the
microchannel reactor MCR1. Unfortunately, the comparison of this
most simple type of catalyst failed, because the catalytic results
and especially the selectivity/conversion curves were too different.
Obviously, the catalytic performance of silver is not really constant
and selectivity as well as conversion degree depend strongly upon
the history of the silver and its manufacturing process. As seen
in Fig. 4.53 (p.125), the selectivity of silver foils used in the tube
type reactor is remarkably higher than that of silver used in the
microchannel reactor MCR1, but the overall activity of the last is
way lower. Exactly the opposite behavior is observed as soon as
silver foils from a different supplier (Goodfellow instead of Chempur)
were used (FBR1b). As seen in Fig. 4.54 (p.126), the selectivity of
the tube type reactor is lower than that of the microchannel, but
the obtained maximum degrees of conversion using the same test
procedure were higher. Taking the completely negative experiments
with silver wire and silver wool into account (p.126), the assumption
of the manufacturing process of silver being crucial for the selectivity
and activity of bulk silver is supported.
Therefore, the experiments using bulk silver as catalyst cannot
be used to proof any advantage or disadvantage of a reactor con-
struction. In this experiment, no thermal instabilities have been
observed for the fixed bed reactor - likely supported by the high
heat conductivity of silver allowing a sufficient heat transport from
the catalyst to the walls.

MCR2/MCR3 and FBR2/FBR3: Both reactor types showed


similar performance as shown in figure 4.55 to 4.57 (p.128 to 130).
Although different trends of the selectivity as a function of the con-
version degree have been observed, the catalytic performance of both
reactor types seems to be similar. This has to be expected as long
as the same catalyst is used and the same surface area is exposed
to the same flow rates. The differences in selectivity and conversion
curves of the fixed bed reactor FBR2 and FBR3 may be ascribed
182 Chapter 5. Discussion

to evolving hot spot effects. Although both catalyst beds consist


of Aluminum pieces and are supposed to have a much higher inner
heat transfer than a traditional tube type reactor, a hot spot can
still emerge. Small changes in the thermal environment (such as the
homogeneity of the heating coils as well as the exact location of the
thermocouple in the pile of catalytic activated aluminum foils) will
have a major influence on the temperature profile. Such a tempera-
ture profile is required if a runaway as depicted in Fig. 4.58 (p.131)
is encountered.
The most important observation within the present work was
the operational safety of the investigated microchannel reactors. As
long as similar catalysts were used and the overall reaction condi-
tions were not demanding, any catalyst showed very similar perfor-
mance regardless of the reactor type. This has to be expected as
long as heat and/or mass transfer issues do not affect the reaction.
The difference between both types can be clearly seen as soon as
the reaction conditions are chosen to challenge the catalyst/ reac-
tor combination. With its inferior heat transfer capacity, hot spots
resulting in an explosion of the C2 H4 /O2 mixture were likely in the
fixed bed reactor, whereas even the most severe reaction conditions
did not affect any of the microchannel reactors.

5.4.3.2 Comparison of reactor performance

In order to allow a comparison between different reactor types, some


parameters such as heat production rate per volume, ethylene oxide
production rate and adiabatic temperature rise were calculated. In
order to allow a comparison between microchannel reactors and their
conventional counterpart, a normalization onto a single parameter
has to be performed.
The most appropriate method seems to be a normalization on
the reactor volume. For the conventional reactor, the tube volume
is used whereas the stack volume excluding heat exchangers is taken
for the microchannel reactors. The advantage of this method is,
that a comparison will be performed on the basis of the required
reactor volume including the support material (which is important
for the microchannel structure). The disadvantage is, that thinner
channel walls (i.e. optimized structuring method) will affect the
5.4. Heat effects and reaction engineering aspects 183

Table 5.1: Benchmark conditions and performance parameters for differ-


ent ethylene oxide catalysts and reactor types

Pilot Plant MCR3 MMCR10 MMCR8


T/K 463 [63] 523 503 503
C2 H4 /% 0,5 20 20 20
Flowrate/ml/min 84000 290 74 111
Degree of conv./% 48 10 10,4 32
Selectivity/% 59 56 55 47
Reactor Vol./cm3 1130 3,9 1 1,5
flow rate/ ml/cm3 min 74 74 74 74
Heat/W 80,0 2,7 0,8 4,0
Heat per vol./ kW/m3 71 703 750 2667
EO prod. rate/ kg/m3 h 12,4 98,2 99,8 262,3
4Tad / K 46 344 365 1296
4THotSpot / K 28 <2 <2 <2

overall spreadsheet performance of the microchannel reactors.


The data attained from a commercial catalyst [31] will be used
as a benchmark. Furthermore, MCR3 having a commercial mi-
crostructure will be compared to this catalyst/reactor as well as
the modular microchannel reactor MMCR10 using the Shell 800
Series Catalyst. Finally, the self made Ag/-Al2 O3 coated mi-
crochannel reactor MMCR8 will be benchmarked, because this re-
actor exhibited the highest overall activity.
The reaction conditions were chosen in a way, that all catalysts
were compared by applying the same flow rate per gross reactor
volume. The temperature was chosen such that a good performance
was obtained. It is practically impossible to compare all four reactors
using the same temperature, because of the very different kind of
catalyst (i.e. Ag/ Al for MCR3 instead of Ag/-Al2 O3 for MMCR10
and MMCR8) and their different catalytic properties. Results are
given in table 5.1.
The pilot plant sized reactor tube exhibited a hot spot of ap-
proximately 28 K, resulting in a peak temperature of 491 K instead
of 463 K wall temperature. At these low temperatures, the Ag/Al
coated MCR3 was hardly active, therefore a higher temperature of
523 K was chosen for this catalyst/reactor combination. The tem-
184 Chapter 5. Discussion

perature for the Ag/-Al2 O3 reactors MMCR10 and MMCR8 were


503 K and chosen to get as close to the peak temperature of the pi-
lot plant sized reactor as possible. Both reactors proved experimen-
tally, that even higher temperatures and therefore, higher space-time
yields would be possible without compromising safety.
The results of this benchmark prove, that the pilot plant sized
reactor has clearly the lowest overall performance and the highest
internal temperature gradient. It is obvious, that this tube reactor
having a diameter of approximately 53 mm and a surface to volume
ratio of 75 m2 /m3 cannot manage the heat released by the reac-
tion. Even having 2000 W/m2 K at the walls of the tube available
(typical number for phase transition heat transfer), the reactor has
only 150 kW/m3 K available to remove the heat. Considering the
internal heat transfer problems proven by the published tempera-
ture gradients [32, 62], the reactor suffers clearly of heat exchange
problems. It is likely, that the poor thermal contact of the catalyst
pellets (source of heat) by itself and with the wall (sink of heat) is
expected to be the reason for this underperformance. Therefore it
is no surprise, that the investigated microchannel reactors with the
catalyst coating attached directly to the walls of the structure are
superior by far.
In a cooled cross flow micro heat exchanger, approximately 5 to
20 kW/m2 K heat transfer capacity are potentially available. Hav-
ing a surface to volume ratio of 8000 m2 /m3 (i.e. 500 x 500 m3
channels), the potential heat transfer capacity per channel may get
close to 80 MW/m3 K even if the channel structure is not optimized
for heat transfer.
Therefore it is no surprise, that the ignition safety test as de-
picted in figure 4.13 (p.70) passed with success, even in a not cooled
modular designed reactor type. Having 20% C2 H4 in the feed and
conversion degrees of 75% with a selectivity of 47%, the adiabatic
temperature rise is calculated to be 3065 K. The pilot plant sized
reactor tube with its hot-spot temperature measured to be approx-
imately 60% of the adiabatic temperature rise, (i.e. 62%, see Fig.
3.5, p.37 and 61%, see table 5.1), the hot spot of the microchannel
reactor is estimated to be below 0.5% of the adiabatic temperature
rise. Having one thermocouple right below the heating device and
one directly behind the microstructure, 10 K was the highest ever
5.4. Heat effects and reaction engineering aspects 185

measured deviation between both thermocouples. Therefore, these


structures allow increased reaction rates by application of high oxy-
gen partial pressures without hot spot formation. Thus, even clearly
explosive binary C2 H4 / O2 mixtures could be handled safely regard-
less of the applied reaction conditions.
Chapter 6
Experimental

6.1 Flow apparatus


The apparatus as depicted in figure 6.1 is divided into three parts,
namely the gas-supply with pressure- and flow-controllers, the reac-
tor unit equipped with a back-pressure controller and the analytic
devices, consisting of an online-GC, an online-IR and some cooling
traps for collection and offline-analysis of volatile compounds.

6.1.1 Flow control

The reactants ethene, oxygen and methane (only if required) are


fed into the reactor by PIC 1..3 and FIC 1..3 via a long and thin
1/8 stainless steel tube to allow diffusional mixing of the reagents.
All flow controllers (type Brooks 5850S) are controlled by a PC
allowing automated operation of the whole setup. After a short
preheating zone (TIC2), the reagents are passed through the reac-
tor. The pressure within this device is controlled by the following
backpressure-regulator PIC4 allowing a maximum pressure of up to
25 bar. The product mixture leaving the regulator is passed via the
safety valve V5 (to avoid over-pressures) and split into two sepa-
rate streams for analysis purposes. All tubing behind the reactor
is heated to approximately 80o C to avoid condensation of products,
especially water.

187
188 Chapter 6. Experimental

C2H4

CH4

O2
1
K3

K2

K1
V4

V5
0 C
- 78 C

- 78 C

V3

V1
V2
Q

Q
GC
IR
3

2
off-gas

off-gas

Figure 6.1: Scheme of the experimental setup. 1: reactor, 2: GC, 3: IR,


FIC1-3: Massflow controller, PIC1-3,5: pressure regulator, PIC4: pres-
sure controller, TIC 1-4: temperature controller, V1-3: stopcocks, V4:
needle valve, V5: pressure relief valve, K1-3: cooling traps
6.1. Flow apparatus 189

6.1.2 Analytics

The first stream passed some cooling traps to remove volatile com-
pounds such as water and is analyzed online by a Fisher Rosemount
IR regarding its CO2 level. This instrument was also capable of an-
alyzing CO by IR but CO was practically never observed. Instead,
a minor cross sensitivity towards C2 H4 had to be noted.
The condensed (liquid) products collected in the ice traps were
analyzed by a Shimadzu QP-5000 GC/MS and consisted mostly of
water. Besides the main product ethylene oxide, others as acetalde-
hyde, 1.4-dioxane (cyclic dimer of EO), ethylene glycol (EO+H2 O),
acetic acid, formic acid, formalin and formic acid methyl/ethyl ester
have been found in traces. It is likely that most of these components
were formed by side-reactions within the trap.
The second stream was directly fed by an automatic 10-port-
valve having two independent injection loops into a Chrompack CP-
9001 GC equipped with two columns. Both were operated using
Helium as carrier gas. The first column is a Chrompack Silicaplot
column1 and performs the separation of the hydrocarbons including
the separation of methane, ethylene, ethylene oxide and acetalde-
hyde. This column is equipped with a FID2 . The second column
is a Chrompack Poraplot-Q column3 equipped with a TCD4 . This
column is used for the separation of ethene, oxygen and methane.
Water was hardly detectable, because the very broad peak shape did
not allow a sufficient integration / accuracy. Furthermore, its pos-
sible to use this column for the additional quantification of carbon
dioxide to double check this result with the attached IR-photometer.
This method allowed the quantification even at high CO2 concen-
trations above 5%, leaving the metering range of the IR. Therefore,
ethene and methane are analyzed on both columns, CO2 on the
Poraplot- and ethylene oxide and acetaldehyde on the Silicaplot-
column. An example chromatogram is depicted in figure 6.2. The
ability to measure the concentration of the most abundant species
on two independent columns avoids problems caused by statistical
analysis effects. Furthermore, the ability to quantify ethene, ethy-
1 30 m x 0.32 mm, dF= 4 m, column pressure 120 kPa, splitflow 8.4 ml/min
2 Flame ionization Detector
3 30 m x 0.32 mm, dF= 10 m, column pressure 83 kPa, splitflow 15.1 ml/min
4 Thermal Conductivity Detector
190 Chapter 6. Experimental

lene oxide and carbon dioxide as the main carbon containing species
allows to cross-check the results for a good mass balance. This check
was done automatically by a small program, analyzing the logbook
of the GC.
Both GC-detectors as well as the IR were calibrated using
ethene (2.7), oxygen (3.5) and methane (4.0). The calibration for
ethylene oxide and CO (IR only) as well as CO2 was performed
using certified gas-mixtures supplied by Messer-Griesheim.
Selectivities and conversion degrees were automatically calcu-
lated from the GC results making use of the additional CO2 level
measured by IR. It has to be noted, that the calculation of the
selectivity as well as conversion degree are based on the amount

120
1
100 Silicaplot [FID]
Signal / a.u.

80

60
2
40

20

0 2 4 6 8 10 12 14

t/ min
3
400 Poraplot Q [TCD]

300
Signal / a.u.

200

4 5 6
100

0 2 4 6 8 10 12 14
t/min

Figure 6.2: Online GC of gaseous product streams within an experiment


using the microchannel reactor MCR3 (data see table 4.8) Reaction con-
ditions: T=563 K, p=3 bar, 20% C2 H4 in O2 , =700 ms, X=38.1%,
S=52.2%; signal denomination: 1+5 C2 H4 , 2+6 ethylene oxide, 3 O2 , 4
CO2 , CH4 1.51 min on Silicaplot and 2.26 min on Poraplot (if applicable).
Temperature program: 5.5 min @ 60o C, with 20 K/min to 230o C, 1 min
@ 230o C
6.2. Reactor design 191

of ethylene and not on the concentration. Due to the reaction


2 C2 H4 + O2 2 C2 H4 O the STP volume of reagents and prod-
ucts are not the same and therefore, it is not advised to calculate
the selectivity and/or the conversion degree directly on the product
concentrations. This volume deficit caused by the above men-
tioned reaction causing slightly higher concentrations than expected
by mass-balance was always taken into account.

6.2 Reactor design

In this section, the reactor design of the four used types of labo-
ratory reactors is described. Generally, all reactors were made of
stainless steel to allow a guaranteed pressure resistance of 2.5 MPa.
All sealings were made of copper, the reactor material was stainless
steel type 1.4571.

6.2.1 Pressure resistant laboratory tube type re-


actor (FBR1-4)

The pressure resistant laboratory reactor as shown in figure 6.3 was


used for experiments with conventional, traditional fixed bed cata-
lysts as well as any bulk material not suitable for a microchannel
reactor. In order to avoid potential side-reactions on the metal sur-
face, the steel surface was shielded with a glass tube. In order to use
powders or crushed catalysts, this reactor was equipped with a plug
of DMCS deactivated quartz glass wool (supplied by Chrompack,
normally used in GC injectors). It was experimentally verified, that
this wool had neither blind activity nor negative effects on product
selectivity by product decomposition. The heating was performed
by resistance wire on the reactors surface and thermally as well as
electrically insulated by glass fiber tape. The estimated maximum
heating power was approximately 100 Watt and limited by thyris-
tors, the length of the wire approximately 10 m.
192 Chapter 6. Experimental

Figure 6.3: Design of the pressure resistant laboratory reactor

6.2.2 Commercial microchannel reactors


(MCR1-3)

The microchannel reactors labelled MCR1 to MCR3 were made by


the Forschungszentrum Karlsruhe. The design of those reactors
was a result of an earlier cooperation [81]. The microstructured
parts encapsulated in this reactor were made by use of mechanical
micromachining followed by a stacking and sealing process. Such
a microchannel reactor consists of a stack of 26 microstructured
metallic wafers of 10 x 50 x 0,3 mm3 , each having 33 rectangular
microchannels of e.g. 0.2 x 0.2 x 50 mm3 . Two diffusors for reactant
inlet and product outlet (length approx 60 mm, volume each 2.0 ml)
6.2. Reactor design 193

a b

Figure 6.4: Example of a sealed microchannel reactor (a), with attached


diffusors containing a stack of packed wafers (b); 26 wafers (300 m x 10
mm x 50 mm), channel geometry 200 m x 200 m, 33 channels / wafer
(858 channels total).

are attached to the stack on each side. A wafer in the middle of the
stack with drillings for thermocouples is provided. Such a reactor is
depicted in figure 6.4.
The heating was performed by resistance wire coiled around the
reactor and insulated by glass fiber tissue (early experiments), later
with an aluminum shell encapsulating the cylinder and providing
sufficient wall thickness of 10mm for four 6.35 mm heating cartridges
having each 50 W electrical power.
This type of reactor has several advantages:

pressure tightness: this device is absolutely pressure tight and


has no sealings. Therefore, there are no potential issues with
leakages at elevated pressures, which is an important point
when handling toxic or corrosive materials.

low weight: due to the absence of any fittings, the wall thick-
ness of the housing-material is low compared to the construc-
tions described in the following sections and therefore, the
heat-up and cool down cycle times are very short.

Due to the nature of this construction, there are also some disad-
vantages:
194 Chapter 6. Experimental

fixed housing: it is impossible to change the catalytic material


enclosed in the reactor or even to get access for later analysis.
The development or improvement of a certain catalytic system
requires access to the coated wavers and the ability to change
the catalytic material quickly is greatly appreciated.

the pricing: as mentioned above, access to the enclosed wavers


either for analysis with surface-science techniques or exchange
requires destruction of the reactor. Therefore, this reactor is
very suitable for completely developed systems, but during the
development of a catalyst and/or process, other constructions
are more favorable and cheaper.

6.2.3 Modular microchannel reactors


Due to the above mentioned disadvantages of commercial available
microchannel reactors and the urgent requirement to have nearly
instant access to the wafers during the development of suitable cat-
alytic coatings, two types of modular microchannel reactors have
been developed and evaluated.

6.2.3.1 Modular Microchannel Reactor Type I

The first modular microchannel reactor was developed to overcome


the restrictions by the commercial made reactors without serious
changes to fluidics of this reactor type. Therefore, the construction
was made in a way, that the wafer stack was enclosed in a special
cylindric housing with two diffusors attached to each side.
The modular microchannel reactor type I consists of five dif-
ferent assembly groups (fig. 6.5). The microstructured wafers, the
inner housing in which the wafers are mounted, the two diffusors
for a proper flow distribution and the outer shell with the flange.
The microstructured wafers are mounted within in a cylindrical in-
ner housing, which has a rectangular millcut of 12 x 10 x 50 mm3
with a cover plate on its top.
Two identical diffusors (1.15 cm3 volume each) are attached on
each side of the housing providing an accurate, distributed gas-flow
from the cylindric 1/8 tube connectors at the reactor inlet/outlet to
6.2. Reactor design 195

Figure 6.5: modular microchannel reactor type I: flexible inner housing


for microstructured wafers (up to 10 x 10 x 50 mm3 )
196 Chapter 6. Experimental

the rectangular millcut. As shown in the sketch, this part is placed


in a cylindrical outer shell having a flange, which is bolted with six
5 mm screws and tightened using a sealing gasket made of copper to
guarantee gas-tightness even towards hydrogen at pressures up to 5
MPa. All parts (except of the copper gasket) are made of stainless
steel SS316Ti (1.4571) and manufactured with an accuracy of better
than 1/10 mm in order to diminish bypass flow between the inner
housing enclosing the microstructured parts and the outer pressure
tight shell. This system has proven its usability in laboratory for
roughly four years and no severe problems occurred even at tem-
peratures up to 600 Kelvin. Specific heat production rates of 10 to
20 Watts were controlled without heat transfer problems. At higher
temperatures, the oxidation of the steel surface caused some trou-
ble and required higher tolerances between the inner housing and
outer shell. Using this modular microchannel reactor it is possible
to exchange the microstructured wafers within 15 to 30 minutes after
cooling down to ambient temperatures.
Heating was performed by two nozzle band heaters type DGM
made by Keller, Ihne & Tesch GmbH, having a total heating power
of more than 500 W and restricted by thyristors to approximately
100 W (Fig. 6.6)
The advantage of this construction is its modularity, allowing
various numbers of microstructured wafers and the low area with its
sealing gasket, allowing the application of large pressures onto the
gasket by thoroughly tightening the 5 mm screws.
The disadvantage of this construction is the potential bypass
flow not only between the wafers and the housing, but between the
inner housing and outer shell. Using very small channel diameters
and a low number of channels, the pressure drop above the mi-
crostructure increases and therefore, the amount of gas bypassing
the channels increases.
Therefore, a small amount of gas passes the catalyst/wafer hous-
ing and causes certain problems. With gas bypassing, the maximum
conversion degree is always below 100%, because the stainless steel
is practically inert for the reagents at reaction temperature. This
causes trouble when comparing two catalysts, because the appar-
ent conversion degree observed at the outlet of the reactor is always
below the one observed directly above the catalyst. On the other
6.2. Reactor design 197

Figure 6.6: Modular microchannel reactor type I with nozzle heating


bands installed.
198 Chapter 6. Experimental

hand, the residence time on the catalytic surface for the stream still
passing the catalyst section is higher than estimated.
In order to improve the situation, the application of additional
aluminum foils as an additional gasket was tested. These gaskets
were applied to increase the pressure drop in the bypass pathway
(i.e. between the inner housing and the diffusors). These gaskets
decreased the bypass-flow a bit, but it proved impossible to suppress
it completely.
This drawback combined with the problem of total oxidation
(most likely in the diffusors, see p. 179), a different design of a
modular microchannel reactor was built and tested.

6.2.3.2 Modular Microchannel Reactor Type II

The design of this improved modular microchannel reactor type II


is depicted in figure 6.7 and it consists typically of four parts:

1. The cover plate. This plate is equipped with two 1/8


Swagelok connectors on the top to allow a pressure tight
adaption of tubing to this device. Between this and the next
plate, a thin copper sheet is used as sealing material. In order
to apply high pressures to this sealing, this cover-plate has a
sharp 0.1 x 0.1 mm2 nose on its bottom side. The tightness
of this construction was successfully tested with nitrogen at
pressures up to 5 MPa. An early model (with fewer screws
& without nose) was hardly usable at pressures exceeding 0.2
MPa.

2. The copper gasket. This gasket was in combination with the


nosing required to achieve pressure tightness.

3. The reactor plate housing. This housing has a special millcut


for microstructured wafers having a width of 10 mm and a
length of up to 50 mm. The depth of this millcut restricts the
number of wafers. Three different reactor plates with depths
of 1,4 and 10 mm were available.

4. The heating plate. This plate located at the bottom of the de-
vice contained a heating cartridge having a maximum power of
6.2. Reactor design 199

Figure 6.7: Modular microchannel reactor MMCR type II. (1) cover
plate, (2) copper gasket, (3) reactor plate, (4) heating plate. Photo with-
out insulation.
200 Chapter 6. Experimental

approximately 100 W. In order to get a good heat distribution,


this plate was made of either aluminum or copper. Both ma-
terials have high heat conductivity coefficients. Unfortunately,
both materials proved to be not long term stable because of
either oxidation (Cu) or deformation (Al), especially at high
temperatures exceeding 350o C/623 K.

A scheme of the modular reactor is depicted in figure 6.8. This


type of reactor showed several major improvements:

its simplicity. The construction of the modular microchannel


reactor Mk1 took roughly 20 man-hours, whereas this type II
was build within half a day. Furthermore, this reactor requires
only one millcut, some welding for the gas-connectors and little
drilling for the screws and the 1/4 heating cartridge.
decreased bypass flow. Having only a single housing, there are

Figure 6.8: Scheme of the modular microchannel reactor MMCR type II


6.2. Reactor design 201

no more opportunities for the gas-flow to bypass between the


inner housing and outer shell. With an improved manufactur-
ing accuracy of 0.05 mm, the gap between the wafers and the
housing itself was decreased. Therefore, overall bypass flow
was decreased.
lower thermal mass. This reactor has roughly 1/3 of the weight
of the Mk1 type and therefore, faster heating up and cooling
down cycles were possible.
improved reliability. With time on stream, the type I reactor
failed by oxidation of the surface causing friction when trying
to remove the inner housing from the outer shell. Due to sur-
face defects (especially at temperatures above 623 K/350o C)
by oxidation, the inner housing was sometimes struck in the
outer shell and quite difficult to remove.
lower dead volume. With the very different flow distribution,
the dead volume within the diffusors was reduced and there-
fore, the residence time within the diffusors decreased - poten-
tial decomposition of products resulting in lower selectivities
was suppressed. The diffusor volume was not constant but
varied with the number of wafers installed. Each millimeter
wafer height required a corresponding diffusor volume of ap-
proximately 2 x 70 l. Thus a small stack of four wafers (height
4 mm) with 280 l post catalytic diffusor volume had only a
quarter of the contact time as the MMCR type I with 1.15 ml
per diffusor, independent from the number of installed wafers.

This design revealed also some practical disadvantages:

pressure tightness. Having a larger sealing area, it proved to be


more difficult to get this reactor really pressure tight. It proved
that a kind of nose on the stainless-steel top cover was required
to decrease the sealing area and increasing the overall pressure
applied to the sealing gasket. Without that nose, it proved im-
possible to get this reactor virtually leak-free. Therefore, only
the later model was used in catalytic tests. It was also tried
to make use of a silicon based high-tech glue5 to get the reac-
tor tight, but the silver catalyst was irreversibly deactivated
5 Delo-Gum, www.delo.com.
202 Chapter 6. Experimental

and destroyed by chlorine components in this glue. For cata-


lysts, which are not affected by chlorine, a bit of silicone based
high temperature glue as sealing (instead of copper) seems to
be a very practical and reliable method to get such a device
leak free. Experiments with glue as sealing proved pressure
tightness of more than 5 MPa without failure.

a higher surface to volume ratio. Although it was easily pos-


sible to heat this reactor up to 573 K with the single heating
cartridge build in the bottom plate of the reactor, the temper-
ature gradient across the reactor was much larger than that of
the MK1 type. This problem was partially solved by a thick
insulation with glass-wool. In order to really overcome this
issue, the design should change a bit and a second heating
cartridge on top of the device should be installed.

6.3 Design and manufacturing of mi-


crostructured wafers
The microstructured wafers are the heart of the reactor and
offered by several manufacturers like IMM Mainz and the
Forschungszentrum Karlsruhe with different product manufac-
turing technologies like punching, etching or micromachining. At
the beginning of the present work, it proved that a lot of catalytic
tests were necessary in order to develop a good ethylene oxide
catalyst. With every single catalyst and its supporting wafer
being inseparably integrated and forming one piece, every catalyst
requires its own wafer. Due to the huge number of tests and the
limited availability of commercially available structures, it showed
that a comparably cheap manufacturing process on-site utilizing
available machinery was inescapable.

6.3.1 Wire electro discharge machining


At the beginning of this work, microstructured wafers were manu-
factured by the wire electro discharge machining (fig. 6.9). This
method is very suitable for aluminum or aluminum alloys such as
6.3. Design and manufacturing of microstructured wafers 203

Figure 6.9: WEDM made wafer and scheme of its manufacturing


method. 14 channels/cm, each 300 m x 700 m x 50 mm. Machin-
ing time per wafer: 1h
204 Chapter 6. Experimental

Dural, AlMg3 or AlMg4.5Mn. Other conducting materials such as


steel and titan are also suitable for this kind of machining. After
some preliminary tests and solving some problems with deformation
of wafers by thermal stress caused by the production process, a ge-
ometry evolved which was easy to manufacture, had channel sizes
of 300 m x 700 m x 50 mm, and a high yield with nearly no re-
jects of structures caused by manufacturing defects. Furthermore,
one wafer took roughly one hour of machine time and therefore, the
operating expense was acceptable. Generally, an aluminum block
having a geometry of 10 x 50 mm was used for production. In the
front wall of this block, microchannels with a depth of 700 m were
machined. The width is mainly controlled by the wire diameter.
Using a 250 m wire, the channel width was approximately 300 m.
After the last channel was made, the wire was used to cut off the
wafer from this block, resulting in the wafer to drop down. After
that, the next wafer was made using a recursion programming tech-
nique in the machine control. Therefore, the machine was able to
manufacture a sufficient number of wafers overnight without need of
readjustments.
Each wafer made by this method has a thickness of 1 mm, a
width of 10 mm and a length of up to 50 mm. 14 parallel microchan-
nels, each 300 m wide and 700 m deep are machined in each wafer
using a commonly available 250 m copper wire. Channels with a
different width are also producible using wires with different diam-
eters. Due to the nature of this production process, the surface of
the microchannels was very rough. This roughness is very suitable
and desired for high adhesion strength of sol-gel or other coatings as
catalytic active layers.

6.3.2 Parallel multiple milling method

This method was developed in order to achieve three major goals.


The most important drawback of the WEDM machined wafers was,
that the outer Al surface contains a lot of copper caused by the
spark erosion of the WEDM machine. In different applications of
co-workers, this copper contamination of the Al-surface caused un-
expected catalytic trouble [80]. Second, WEDM machining is a quite
expensive method and therefore, the production cost should be low-
6.3. Design and manufacturing of microstructured wafers 205

Figure 6.10: Parallel multiple milling method to produce microstruc-


tured wafers.

ered. And finally, the availability of the WEDM machine was limited
and therefore, a different method had to be tested.
One of the less expensive production methods is the milling
process and CNC machines are widely spread. Therefore, 19 par-
allel side milling cutters were precisely mounted on the same arbor
(fig. 6.10). Each cutter had a width of 250 m and was separated
from its neighbors by a thin ring washer of approximately 200 m.
Now, a thin aluminum sheet of 250 x 100 x 0.5 mm was fixed on the
CNC machines work space by a vacuum table. The CNC machine
was now operated in a way, that the parallel mill cuts had a depth of
250-300 m (fig. 6.11). Without further optimization, one 200 mm
millcut took approximately 5 minutes and a whole sheet with four
to five of these millcuts was made within an hour. Using precision
206 Chapter 6. Experimental

Figure 6.11: Sheet with the parallel mill cuts of each 19 channels by use
of the parallel multiple milling method. Resulting channel geometry: 300
x 300 m2 , channel length up to 200 mm possible.

laser cutting, single wafers were made from this sheet. One of these
sheets yielded up to 20 single wafers. After degreasing in acetone
followed by chemical deburring for 30 s in 10 wt% NaOH, the wafers
were ready for use.

6.4 Catalyst preparation and coating


procedures
The main challenge of performing heterogeneous catalysis within a
microchannel reactor is the immobilization of catalytic active mate-
6.4. Catalyst preparation and coating procedures 207

rial on the walls of the microchannels. As soon as the reactor cannot


be made out of catalytic active material, a suitable coating method
is crucial for success. Unfortunately, most of the coating methods re-
quire the re-invention of the catalyst and its manufacturing process.
This is typically a very time-consuming process.

6.4.1 Physical immobilization methods


One of the most convenient methods to immobilize a catalytic active
material on an inert surface is the application of physical vapor
deposition or sputtering. Using one of these methods, a thin layer
of metallic and catalytic active material is deposited on the wafer
surface, hoping that the catalytic activity of the material is high
enough to have sufficient degrees of conversion. Unfortunately, there
is no surface enlargement possible and the surface to volume ratio
is the same as for the raw wafer material. Furthermore, the dense
metallic layer will cause problems for all reactions which depend not
only on the material, but also on the particle size of the metal. Due
to surface diffusion and agglomeration of silver particles, the dense
silver surface roughened and small particles were formed during the
initial activation period of the catalysts / wafers (see fig.4.11 on
page 66). Within the present work, a Polaron SC 7640 sputtering
machine was used for coating microstructured wafers with silver.
Larger amounts of wafers and layers having a thickness of more
than 800 nm were coated with a more powerful magnetron sputter
machine type P30, made by Vakuum Technik Dresden.

6.4.2 Anodic oxidation & impregnation


Another method to immobilize catalytic active species on the walls
of microchannels is the application of an anodic oxidation process as
long as aluminum is suitable as wafer material. During the oxidation
(scheme of equipment see Fig. 6.12), an outer layer of alumina
having small pores is formed (Fig. 6.13). After a calcination process
simple impregnation methods may be applied to immobilize catalytic
active species within the pores.
As already indicated and reported elsewhere [100], the selec-
tivities towards ethylene oxide for silver impregnated anodically oxi-
208 Chapter 6. Experimental

Figure 6.12: Scheme of the anodization apparatus. a: voltage supply


(constant voltage), b: heat exchanger, c:electrolyte pump, anode: mi-
crostructured aluminum wafer / aluminum wire, cathode: aluminum plate

Figure 6.13: A schematic diagram of the oxide layer. X : electrolyte


anion,U: voltage, T: temperature, c:conc.,i: anodization current, t: an-
odization time [99]

dized aluminum was poor and hard to measure due to absence of the
target product ethylene oxide (< 0.002Vol%). Using H2 SO4 as elec-
trolyte, in most cases only CO2 was observed. Using H3 PO4 , nearly
no activity was left accompanied by a not detectable ethylene oxide
formation. Only using 1.5% Oxalic Acid as electrolyte, noteworthy
amounts of ethylene oxide were observed only when large amounts of
silver were immobilized. The immobilization of these amounts was
6.4. Catalyst preparation and coating procedures 209

performed by quadruple impregnation using a saturated AgNO3 in


CH3 CN solution followed by a calcination in air at 450o C for 30
minutes. The final calcination was performed in a muffle furnace at
450o C for 8 hours.
The overall performance of this catalytic system was poor and
therefore it was abandoned. Most likely, the acidic properties of the
anodically made alumina layer caused an isomerization followed by
total combustion of ethylene oxide to CO2 and H2 O.
Thus, the anodic oxidation was used only to enhance the oxide
layer of selected wafers, still keeping the layer thickness thin. This
enhancement was performed by oxidizing a microstructured wafer in
1.5 wt% oxalic acid at 285 K, 50 V for 20 min, resulting in approxi-
mately 1 m oxide layer thickness. Within the present work, every
anodic oxidation was performed in the same experimental setup us-
ing the same types and concentrations of electrolytes as described
by Wiessmeier, Dietzsch and eventually Kah [87, 80, 86].

6.4.3 Sol-gel coatings


In principal, the sol-gel method may be a suitable way to immobilize
catalytic active material on the walls of microchannels. It is easily
possible to produce a gel with any documented method and to dip-
coat the wafers in the gel. After a calcination, the catalyst would
be ready to use.
Unfortunately, aluminum based sol gel recipes result typically
in the formation of the thermodynamically stable Al2 O3 when
the calcination temperature may not exceed 450o C. Al2 O3 is due
to its acidity unsuitable for ethylene oxidation.
In order to get a working Ag/-Al2 O3 catalyst, a different coat-
ing method was used. The sol-gel coating was performed by the
HITK6 . This process is based on a sol-gel method with nano dis-
persed -Al2 O3 within the gel as a seed crystal. The exact recipe
of this sol/gel is not available. The gel was immobilized by using
spin coating using a temperature of 450o C. According to the man-
ufacturer, the resulting layer is slightly porous, having an -Al2 O3
content of more than 99%. The so prepared -Al2 O3 layer was acti-
6 Hermsdorfer Institut fur Technische Keramik e.V.
210 Chapter 6. Experimental

vated by using a conventional silver-lactate immobilization method


as described by Minahan [78] and was used for MMCR8, see p. 96ff.

6.4.4 Electrostatic powder deposition

The disadvantage of all these methods is the necessity of re-inventing


a catalytic system rather than making use of an already (commer-
cially) available one. The particular problem is, that one needs a
kind of binding material or glue in order to immobilize the finely
crushed catalyst on the walls of the microchannels. Its certainly a
risk, that the binding material may affect the catalytic properties of
the device by either decreasing the surface area of the catalyst by
simple encapsulation of catalytic particles within the binder or that
one of the reagents / products may undergo (catalytic) side reac-
tions. It is known, that ethylene oxide is very sensitive towards acidic
surface sides and therefore, commonly used materials like alumina
or silica based gels should not be used n presence of this epoxide.


 

Figure 6.14: REM photo of an electrostatically Shell 800 Series Cat-


alyst coated microstructured wafer. (a) overview, (b) wall of a single
microchannel, (c) enlarged view of that wall, (d) further (highest) en-
largement.
6.4. Catalyst preparation and coating procedures 211

In order to overcome this problem, it was tried to immobilize


a commercial catalyst on the walls of a microchannel without mak-
ing use of any binding material such as water-glass. Therefore, a
different bonding method had to be used. Aluminum has typically
a protective oxidic layer on its surface. As soon as this surface is
treated with 10 wt% NaOH solution at room temperature for a short
time (roughly 90 sec), a less stable surface consisting of fresh and
very reactive Al-OH groups emerges. These functional groups allow
to immobilize a thin layer of powder. Its sticky enough to handle
the wafer with its coating, but definitely neither long term stable
nor scratch-proof.
Directly after the etching process by NaOH, the aluminum wafer
must be subsequently cleaned with distilled water (to remove NaOH)
and acetone (to remove water) and immediately placed for coating in
a stirred vessel. This vessel contains a suspension of 600 mg crushed
& ball-milled catalyst powder dispensed in 100 ml acetone. The
wafer is located in the middle of two electrodes, which are electrically
used as cathode (-). The wafer itself is electrically used as anode
(+), the applied voltage7 was 100 V with a distance between anode
and cathode of roughly 17 mm. The current was typically 6 mA
or below and it decreased during the coating process - which took
approximately 90-120 seconds per wafer.
Using this setup, a homogeneous coating quality was obtained.
REM pictures of this coating are depicted in figure 6.14. Unfortu-
nately, the adhesion strength of the coating is far away from being
scratch resistant, but the quality is good enough to be used in gas
phase catalysis for testing purposes for some weeks.

7 Standard laboratory DC power supply, adjustable voltage and current lim-


iter
Chapter 7
Appendix

7.1 Chemical properties of ethylene ox-


ide
Ethylene oxide is a hazardous, highly flammable and, under certain
conditions, a violently explosive chemical. Due to its exothermic dis-
proportionation, polymerization and isomerization ability combined
with a low flammability limit and an exothermic decomposition and
combustion, there is always a risk of a severe incident even in ab-
sence of oxygen. Pure ethylene oxide is involved in several basic
reactions, which affect the handling of this chemical:

Combustion
5
C2 H 4 O + O2 2 CO2 + 2 H2 O
2
The combustion is highly exothermic and yields -1210 kJ/mol
or -27469 kJ/kg. The lowest flammability limit is 2.6% in Air.
Even solutions of ethylene oxide in water support combustion
for concentrations above 5 wt%. Therefore, a dilution of 1:100
is necessary if ethylene oxide has to be removed by water [12].

Decomposition

213
214 Chapter 7. Appendix

C2 H4 O CO + CH4
The reaction yields 3051 kJ/kg (134 kJ/mol), causing a
pressure rise in closed systems. This decomposition reaction
is known to proceed explosively under certain circumstances.
TNT for example has a maximum decomposition energy of
4600 kJ/kg [96].

Disproportionation

6 C 2 H4 O 5 C2 H4 + 2 CO2 + 2 H2 O
C2 H 4 O CO + CH4
2 C 2 H4 O 2 CO + C2 H4 + 2 H2
2 C 2 H4 O 2 CO + C2 H6 + H2

The disproportionation of ethene oxide can occur in presence


of high surface area metal oxides such as Fe2 O3 causing
a strong local temperature rise even above the decompo-
sition temperature of ethylene oxide. Thus, rust might
cause severe incidents and should be avoided when handling
ethylene oxide. It was found, that in presence of rust temper-
atures of 140o C are high enough to initiate an EO ignition [88].

Isomerisation

C2 H4 O CH3 CHO
Ethylene oxide can be isomerized to acetaldehyde, yielding a
reaction enthalpy of 2621 kJ/kg (115 kJ/mol). This reaction
is normally performed by use of acidic and / or high surface
materials, such as Al2 O3 , SiO2 , iron oxides or phosphoric
acid [92].

Polymerisation

n C2 H4 O (CH2 CH2 O)n


7.2. Environmental effects of ethylene oxide 215

Ethylene oxide can be polymerized to polyethyleneglycol (with


traces of water) using catalysts such as strong alkali or even
rust. The latter causes clogging of equipment if improper steels
are used. This exothermic reaction takes place at temperatures
of approx. 200o C even under non catalytic conditions. The
reaction enthalpy is approximately -2093 kJ/kg (-92 kJ/mol).

Insertion into polarized bonds

R NH2 + 2 C2 H4 O R N(C2 H5 OH)2


The insertion of ethylene oxide into polarized bonds such as
N-H, R-OH or R-X is the most important technical application
of this chemical. More examples are discussed in section 1.3.

7.2 Environmental effects of ethylene ox-


ide
Because of its high reactivity, ethylene oxide is a highly toxic chem-
ical. According to US-OSHA1 regulations, the maximum exposure
limit is 0.5 ppm as time-weighted average and has excursion limit
of 5ppm at the most. In Germany, the TA-Luft concentration
is 5 mg/m3 (2.8 ppm) [93]. According to the German regulation
TRGS-900, the maximum tolerable concentration at a workplace is
2 mg/m3 (1.1 ppm) [93, 94].
The high odor threshold of about 250 ppm prohibits nor-
mal protection by sense of smell. The LC50Rat concentration is
5000 ppm /1h [12] and 1460 ppm for 4h exposition time [12, 95].
Skin contact should be avoided, because the evaporative cool-
ing can cause freezing. Ethylene oxide is highly irritating to eyes
and skin and even diluted ethylene oxide solutions are dangerous
and may cause severe injuries. The LC50 2 value for fish is only 84
mg/l (exposition time: 4d). Due to hydrolysis, biodegradation and
1 United States, Occupational Safety and Hygiene Act
2 LD50 is the amount of a material, given all at once (solid or liquid), which
causes the death of 50% (one half) of a group of test animals. The LC50 refers
to a concentration of a chemical in air that kills 50% of the test animals in a
given time (usually four hours)
216 Chapter 7. Appendix

evaporation, the ethylene oxide concentration in free flowing water


is reduced by 95% within 4h (at 25o C) [5]. Furthermore, ethylene
oxide penetrates cloth, leather or rubber easily and thus, appropri-
ate protective clothing should be used and contaminated clothing
removed. Ethylene oxide is known to

be harmful to embryos (teratogenic)


increase incidents of cancer (mutagenic, especially leukemia,
stomach, brain)

poisoning, especially of the central nerve system, liver, kidneys


damage the lungs (pneumonia).

Thus, due to its toxicity, ethylene oxide should be handled with


extreme care and emission of ethylene oxide should be avoided.
7.3. Physical properties of ethylene oxide 217

7.3 Physical properties of ethylene oxide


Ethylene oxide has an etheric, sweet smell. At room temperature, it
is a colorless gas with a melting point of 161.5 K and boiling point
of 283.6 K. It is highly flammable with flammability limits between
2.6 and 100% and a flash point below 255 K make it dangerous and
it should be handled with extreme care. Its autoignition tempera-
ture is 702K, being decomposited above 833 K. The most important
molecular constants and properties are listed in table 7.1.
Ethylene oxide is completely miscible with water in any pro-
portion. Due to the formation of hydrates, ethylene oxide / water
solutions are highly non-ideal and do not follow Raoults law [12].

Table 7.1: Physical properties of ethylene oxide [12].

Molecular mass 44.053 g/mol


Critical Temperature 469.15 K
Critical Pressure 7.191 kPa
Melting Point 161.46 K
Triple Point 161.46 K
Triple Point Pressure 0.0078kPa
Boiling Point 283.6 K
Heat of Formation (Gas) -1194.8 kJ/kg
Gibbs Energy of Formation (Gas) -300.3 kJ/kg
Standard Entropy (Gas) 5.52 kJ/kg K
Standard Heat of Combustion -27649 kJ/kg
Standard Heat of Decomposition -3051 kJ/kg
Standard Heat of Isomerization -2621 kJ/kg
Standard Heat of Polymerisation -2093 kJ/kg
Standard Heat of Hydrolysis -2081 kJ/kg
Heat of Solution in Water -142.7 kJ/kg
Flash Point < 255.16 K
Flammability Limit (Air) 2.6-100%
Autoignition Temp. 702 K
Decomposition Temp. 833.2 K
218 Chapter 7. Appendix

7.4 Additional figures and measure-


ments for MCR2
Activation: The selectivity and conversion degree of this reactor
are depicted in figure 7.1. Initially, the conversion degree is close
to 0.018% at selectivities of approximately 68.5%. With increasing
time on stream, the conversion degree increased to 9.7-10.2% at
decreasing selectivities of 59.7%. The highest selectivities of 73%
were observed after about 60h time on stream at a conversion degree
of 2.3%. Unfortunately, the analysis stopped at this point due to a
system malfunction, but the catalyst did not further activate as well
as no further decline of the selectivity was observed in subsequently
performed control experiments.
0,8
selectivity to ethylene oxide

0,7

0,6

0,10
degree of conversion

0,08

0,06

0,04

0,02

0,00
00 24 48 72 96 120 144 168
time on stream / h

Figure 7.1: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the initial time on stream for the Ag/Al2 O3 /Al mi-
crochannel reactor MCR2. Reaction conditions: 20% C2 H4 , 20% O2 ,
60% CH4 , = 1.03 s p= 0.3 MPa, T= 503 K (80 data points per 24h)

Influence of the oxygen concentration: The influence of the


oxygen partial pressure on selectivity and the degree of conversion
7.4. Additional figures and measurements for MCR2 219

is depicted in figure 7.2.


Keeping the ethene concentration at a constant level of 20%,
the oxygen concentration was varied between 6% and 80% using
methane as balance. At a reactor temperature of 463 K, the selec-
tivity seemed to be nearly independent from the oxygen concentra-
tion, varying only slightly between 62.7% at low and 63.9% at high
oxygen concentrations. However, the degree of conversion depended
strongly on the oxygen concentration. At 6% O2 , the conversion
degree was 1.7%, whereas an O2 concentration of 80% resulted in a
conversion degree of 6.2%.
With increasing reactor temperature and therefore, with in-
creasing degree of conversion, the oxygen concentrations became
more important for the selectivity. At 503 K and conversion de-
grees between 7.5% (at 6% O2 ) and 20.8% (at 80% O2 ), the selec-
tivity increased with increasing oxygen concentration from 45.4% to
57.8%.
463K
selectivity to ethylene oxide

483K
0,6
503K

0,5

0,20 503K
degree of conversion

0,15

483K
0,10

0,05 463K

0,00
0 10 20 30 40 50 60 70 80 90
oxygen concentration / %

Figure 7.2: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the oxygen concentration for the Ag/Al2 O3 /Al mi-
crochannel reactor MCR2. Reaction conditions: 20% C2 H4 , balance CH4 ,
=1.1s (STP), p= 0.3 MPa, T= 463 K / 483 K / 503 K.
220 Chapter 7. Appendix

Influence of the ethene concentration: The influence of the


ethene concentration on selectivity and conversion degree was inves-
tigated by a variation of the ethene concentration at reactor tem-
peratures of 463 and 503 K, using a constant residence time and
oxygen as balance in order to attain the best performance of the
catalyst. The results of both experiments are depicted in figure 7.3.
At a reactor temperature of 463 K, the selectivity increased slightly
with increasing ethene concentration, whereas the degree of conver-
sion decreased. Using 0.7% ethene in oxygen, the conversion degree
was 49.8% at a selectivity of 57.9%. With increasing C2 H4 concen-
trations of up to 26.1%, the degree of conversion decreased to 4.8%,
yielding a selectivity of 64.6%. With increasing temperature, the de-
gree of conversion increased as expected. Furthermore, the ethene
concentration had less influence on the selectivity. At 503 K, the
selectivity between 4% and 32.6% ethene was within a narrow range
of 60.7-61.1%, whereas at lower reactor temperatures of 463 K (still
applying the same conditions), an increase in selectivity from 60.5%
to 64.8% was observed.
0,66
selectivity to ethylene oxide

463K
0,64

0,62
503K
0,60

0,58
0,8

0,6
degree of conversion

0,4

0,2
503K
463K
0,0
0 10 20 30 40
ethene concentration / %

Figure 7.3: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the ethene concentration for the Ag/Al2 O3 /Al microchan-
nel reactor MCR2. Reaction conditions: balance O2 , =1.1s, p= 0.3 MPa,
T= 463 and 503 K.
7.4. Additional figures and measurements for MCR2 221

Influence of the total pressure: The influence of the total pres-


sure on the reactors performance is depicted in figure 7.4. Applying
a total pressure of 0.1 MPa, the selectivity is 53.0% at a conversion
degree of 13.7%. With increasing pressure of up to 0.5 MPa, the se-
lectivity increased to 61.3% at a conversion degree of 17.0%. Higher
pressures of 0.7 and 0.9 MPa did not affect selectivity and conversion
degree. Applying pressures above 0.9 MPa lead to condensation of
water and malfunctions of the pressure controller.

residence time / s (ST)

0 2 4 6 8
selectivity to ethylene oxide

0,65

0,60

0,55

0,50

0,18
degree of conversion

0,16

0,14

0,12

0,10
0,0 0,2 0,4 0,6 0,8 1,0

pressure / MPa

Figure 7.4: Selectivity to ethylene oxide and degree of conversion as a


function of the total pressure for the microchannel reactor MCR2. Reac-
tion conditions: 20% C2 H4 in O2 , p= 0.3 MPa, T= 503 K.
222 Chapter 7. Appendix

7.5 Additional figures and measure-


ments for MCR3
Activation: The activation was performed using a mixture of
20% ethene, 20% oxygen with methane as balance (Fig. 7.5), apply-
ing a reactor temperature of 503 K (same conditions as for MCR2).
Initially, conversion degrees below 1% and selectivities of approx-
imately 80% were observed. With increasing time on stream, the
conversion degree increased slowly to approximately 2.6% after 48h.
Regaring the slow activation process and low conversion degrees, the
reactor temperature was increased to 523 K in order to accelerate
the expected reconstruction of the silver surface. Therefore, higher
and still increasing conversion degrees of 5.0% (at 50 h TOS) to 7.0%
(at 90 h TOS) were observed. The selectivity seemed to decrease
nearly linear from 70% (at 50 h TOS) down to 64% (at 90h TOS).
At this point, the oxygen concentration was increased from 20% to
80%, keeping the reactor temperature at 523 K, yielding constant
selectivities of 66.4% at conversion degrees of 12.2%.
1,0
selectivity to ethylene oxide

0,8

0,6

503K 523K
20% C 2 H4 in 20% O 2
degree of conversion

0,10 20% C 2 H4 in 20% O 2

0,05
523K
20% C 2 H4
0,00 in 80% O 2

0 10 20 30 40 50 60 70 80 90 100 110
TOS / h

Figure 7.5: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the initial time on stream for the Ag/Al microchannel
reactor MCR3. Reaction conditions: p= 0.3 MPa, T= 503 K and 523 K,
20% C2 H4 in 20% (balance CH4 ) and finally 80% O2 .
7.5. Additional figures and measurements for MCR3 223

Influence of the oxygen concentration: The influence of the


oxygen concentration on selectivity and conversion degree is depicted
in figure 7.6. Keeping the ethene concentration at a constant level
of 20%, the oxygen concentration was varied between 6% and 80%
using methane as balance. At a reactor temperature of 483 K, the
selectivity showed moderate dependence on the oxygen concentra-
tion. In the whole range of oxygen concentrations, the selectivity
increased from 58.4 to 62.5%. As already observed for many cata-
lysts before, the degree of conversion depended strongly on the oxy-
gen concentration. At 6.8% O2 , the conversion degree was 1.6% and
increased to 6.2% at 80% oxygen. A similar behavior is observed for
higher reactor temperatures. At 503 K, the curve is shifted down-
wards by 2.5% to lower selectivities (54.8 - 59.6%). Simultaneously,
the degrees of conversion roughly doubled and varied between 3.5%
and 12.9% for oxygen concentrations of 6.2% and 80%, respectively.
At 523 K, the selectivity decreased to 42.3 -56.1% at conversion de-
grees of 8.5 -23.4%. Therefore, the dependence of selectivity and
conversion degree increased with increasing reactor temperature.
0,65
selectivity to ethylene oxide

483K
0,60
503K

0,55 523K

0,50

0,45

523K
0,20
degree of conversion

0,15
503K
0,10

0,05 483K

0,00
0 10 20 30 40 50 60 70 80 90
oxygen concentration / %

Figure 7.6: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the oxygen concentration for the Ag/Al microchannel
reactor MCR3. Reaction conditions: 20% C2 H4 , balance CH4 , =1.1 s
(STP), p= 0.3 MPa, T= 483 K / 503 K / 523 K.
224 Chapter 7. Appendix

Influence of the ethene concentration: The influence of the


ethene concentration on selectivity and conversion degree of MCR3
was investigated by a variation of the ethene concentration at re-
actor temperatures of 483 K and 523 K, using a constant residence
time of 1.1 s and oxygen as balance. The results of both experiments
are depicted in figure 7.7. At a reactor temperature of 483 K, the
selectivity increased slightly with increasing ethene concentration,
whereas the degree of conversion decreased. At 1.0% ethene in oxy-
gen, the conversion degree was 27.7% at a selectivity of 67.5%. With
increasing C2 H4 concentrations of up to 34.5%, the degree of conver-
sion decreased to 2.4%, yielding a higher selectivity of 73.0%. The
selectivity seemed to increase nearly linear with increasing ethene
concentration.
An increased reactor temperature of 523 K lead to higher de-
grees of conversion and lower selectivities. At 523 K, the selectivity
varied between 57.0 and 67.3% at ethene concentrations of 0.9% to
34%, respectively. The corresponding degree of conversion decreased
0,75
selectivity to ethylene oxide

483K

0,70

523K
0,65

0,60

0,5
degree of conversion

0,4

0,3

0,2

0,1 523K
483K
0,0
0 10 20 30 40
ethene concentration / %

Figure 7.7: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the ethene concentration of the Ag/Al microchannel re-
actor MCR3. Reaction conditions: balance O2 , =1.1 s, p= 0.3 MPa,
T= 483 and 523 K.
7.5. Additional figures and measurements for MCR3 225

with increasing ethene concentration from 48.0% (at 0.9% C2 H4 ) to


8.9% (at 34.4% C2 H4 ). Obviously, the selectivity showed, contrary
to the experiment at 483 K, initially a sharp increase in selectivity
followed by only minor improvements. At C2 H4 concentrations be-
tween 1.0% and 10%, the selectivity increased sharply from 57% to
66.4%. A further increase of the ethene concentration up to 33.4%
lead to slightly enhanced selectivities of up to 67.3%.

Influence of the total pressure: The influence of the reactors


total pressure on selectivity and conversion degree is depicted in
figure 7.8. The degree of conversion increased in the pressure range
from 0.1 to 1.5 MPa from 7.1% to 14.1%. In parallel, the selectivity
increased from 55.9% at 0.1 MPa to a maximum of 63.6% at 0.9 MPa.
The application of higher reactor pressures lead to marginal lower
selectivities of 62.9% at a reactor pressure of 1.5 MPa.
selectivity to ethylene oxide

0,64

0,62

0,60

0,58

0,56
degree of conversion

0,150

0,125

0,100

0,075

0,0 0,2 0,4 0,6 0,8 1,0 1,2 1,4 1,6

pressure / MPa

Figure 7.8: Selectivity to ethylene oxide and degree of conversion as a


function of the total pressure for the microchannel reactor MCR3. Reac-
tion conditions: 20% C2 H4 , each in O2 , = 1.1 s, T= 503 K.
226 Chapter 7. Appendix

7.6 Additional figures and measure-


ments for MMCR8

Influence of the oxygen concentration: The influence of the


oxygen concentration on the selectivity and conversion degree of
this catalyst at two different reactor temperatures is depicted in
figure 7.9. At 483 K, the selectivities were between 50.3% at 80%
oxygen and 41% at 11.6% oxygen. The corresponding conversion
degrees were lower at lower oxygen concentrations. At 80% O2 21.2%
conversion was observed, whereas at 11.6% O2 only 8.8% could be
measured. At higher reactor temperatures of 503 K and thus, higher
degrees of conversion, lower selectivities were observed. At 80%
oxygen, the selectivity was 45.5% at a conversion degree of 36.7%.
Applying lower oxygen concentrations, the selectivity as well as the
conversion degree were lower. At 6.8% oxygen, 10.9% conversion
degree at a selectivity to EO of 24.7% was measured.
selectivity to ethylene oxide

0,5 483K

503K
0,4

0,3

0,35 503K
degree of conversion

0,30

0,25

0,20 483K

0,15

0,10

0 10 20 30 40 50 60 70 80 90
oxygen concentration / %

Figure 7.9: Selectivity to ethylene oxide and conversion degree of ethene


as a function of the oxygen concentration for the Ag/-Al2 O3 microchan-
nel reactor MMCR8 (impregnated three times). Reaction conditions:
20% C2 H4 , balance CH4 , =1.1 s (STP), p= 0.3 MPa, T= 483 K and
503 K.
7.6. Additional figures and measurements for MMCR8 227

Influence of the ethene concentration: The impact of the


ethene concentration on this Ag/-Al2 O3 microchannel reactor
MMCR8 was investigated using oxygen as balance. Due to the
high activity of this catalyst, higher flow rates and higher ethene
concentrations were applied. Thus, a different experimental setup
of the flow control had to be used. Results are shown in figure
7.10. At a reactor temperature of 483 K, increasing selectivities and
decreasing conversion degrees were observed with increasing ethene
concentration. Using 3.5% ethene in oxygen, selectivities of 51.7%
at a conversion degree of 21.3% were observed. With increasing
ethene concentrations of up to 60%, the conversion degree decreased
to 1.55% and the selectivity improved slightly to 54.6%. Applying a
higher reactor temperature of 503 K, higher conversion degrees and
lower selectivities were observed as expected. Using 3.6% ethene in
oxygen, 49.1% selectivity at a conversion degree of 33.0% had to be
noted. With increasing ethene concentration of up to 60%, the de-
gree of conversion decreased to 3.21% and the selectivity decreased
very slightly to 48.0%. Thus, slightly increasing selectivities were
observed at 483 K and slightly decreasing ones at 503 K.
selectivity to ethylene oxide

0,54
483K

0,52

0,50

503K
0,48

503K
0,3
degree of conversion

0,2
483K

0,1

0,0

0 10 20 30 40 50 60
ethene concentration / %

Figure 7.10: Selectivity to ethylene oxide and conversion degree of


ethene as a function of the ethene concentration for the Ag/-Al2 O3 mi-
crochannel reactor MMCR8. Reaction conditions: balance O2 , = 53 ms,
p= 0.3 MPa, T= 483 / 503 K.
228 Chapter 7. Appendix

Influence of the total pressure: The influence of the total


pressure on selectivity and conversion degree was investigated at
483 K and 503 K at 20% C2 H4 in O2 . The residence time was
adjusted to 150 ms to take the high activity of this catalyst into
account. Results are depicted in figure 7.11. At 483 K, the selectivity
increases from 48.4% with increasing total pressure to 58.5% at 0.9
to 1.2 MPa. Higher total pressures led to decreased selectivities
of 55.4% at 2 MPa. Contrary, the conversion degree increases in
the entire range of pressures from 9.4% at 0.1 MPa to 14.6% at
2 MPa. The application of higher reactor temperatures led to higher
degrees of conversion and slightly lower selectivities. Initially, 17.1%
conversion degree and 43.7% selectivity were observed at 0.1 MPa.
With increasing pressure, the conversion degree increased steadily
to 24.1% at 2 MPa, whereas the selectivity showed a maximum of
50.9% at 0.6 to 0.9 MPa before dropping down to 45.6% at 2 MPa.
Therefore, the selectivity maximum is shifted to lower pressures with
increasing reactor temperature.
0.60
selectivity to ethylene oxide

0.55 483K

0.50

503K
0.45

0.25
503K
degree of conversion

0.20

0.15
483K

0.10

0.0 0.5 1.0 1.5 2.0 2.5

pressure / MPa

Figure 7.11: Selectivity to ethylene oxide and degree of conversion as a


function of the total pressure for the Ag/-Al2 O3 microchannel reactor
MMCR8. Reaction conditions: 20% C2 H4 in O2 , = 120 ms, T= 483 /
503 K.
7.7. Additional figures and measurements for MMCR9/10 229

7.7 Additional figures and measure-


ments for MMCR9/10
Activation The selectivity to EO and the conversion degree as a
function of the time on stream are depicted in fig. 7.12. This com-
mercial catalyst showed catalytic activity from the very first moment
on, suggesting, this catalyst was delivered by the manufacturer in an
activated state. Initially, about 54% selectivity at conversion degree
of 2.8% was observed. With increasing time on stream, the selec-
tivity decreased to 51% at a slightly higher degree of conversion of
approximately 3.2%.

0,035

0,030 0,60

selectivity to ethene oxide


degree of conversion

0,025

0,020 0,55

0,015

0,010 0,50

0,005

0,000 0,45

0 5 10 15 20 25

time on stream / h

Figure 7.12: Selectivity to ethylene oxide and conversion degree of


ethene as a function of the time on stream for the Shell 800 Series Cat-
alyst microchannel reactor MMCR9. Reaction conditions: 20% C2 H4 ,
8.8% O2 , balance CH4 , p= 0.3 MPa, T= 483 K.

Influence of the oxygen concentration: The influence of the


oxygen concentration on selectivity and conversion in the tempera-
ture range of 483 K to 543 K on this commercial catalyst is depicted
in figure 7.13. At a reactor temperature of 483 K, avoiding high
conversion degrees, the selectivity increased with increasing oxygen
concentration by 6.2% from 52.9% to 59.1% as well as the degree of
230 Chapter 7. Appendix

0,6
483K
selectivity to ethylene oxide 503K
0,5 523K
543K

0,4

0,3

543K

0,15
degree of conversion

523K

0,10
503K

0,05 483K

0,00
0 10 20 30 40 50 60 70 80 90
oxygen concentration / %

Figure 7.13: Selectivity to ethylene oxide and conversion degree of


ethene as a function of the oxygen concentration for the Shell 800 Se-
ries Catalyst coated modular microchannel reactor MMCR10. Reaction
conditions: T= 483, 503, 523 and 543 K, p= 0.3 MPa, = 230 ms, 20%
C2 H4 , balance CH4 .

conversion, which increased from 1.6% to 5.1%. At higher reactor


temperatures and thus, higher conversion degrees, the increase in
selectivity with increasing O2 concentration raised sharply. Using a
reactor temperature of 503 K, conversion degrees from 3.1% to 9.1%
were observed yielding selectivities between 41.7% to 54.0%. At the
highest investigated temperature of 543 K, the selectivity was 28.1%
at 6.9% O2 and increased to 46.6% when 80% oxygen were applied.
The corresponding conversion degree increased from 6.9% to 18.3%,
respectively.

Influence of the ethene concentration: Figure 7.14 shows


clearly, that high ethene partial pressures are favorable. At a
reactor temperature of 483 K, the selectivity increased from 51.2%
to 58.5% with increasing C2 H4 partial pressure, whereas the degree
of conversion decreased from 38.8% down to 5.1%. As the reactor
7.7. Additional figures and measurements for MMCR9/10 231

0,60
483K

selectivity to ethylene oxide


0,55
503K

0,50

0,45

0,4
degree of conversion

0,3

0,2

0,1 503K

483K
0,0
0 5 10 15 20
ethene concentration / %

Figure 7.14: Selectivity to ethylene oxide and conversion degree of


ethene as a function of the ethene concentration for the Shell 800 Se-
ries Catalyst coated modular microchannel reactor MMCR10. Reaction
conditions: balance O2 , T= 483 / 503 K, p= 0.3 MPa, = 230 ms.

temperature was increased to 503 K, lower selectivities and slightly


higher conversion degrees were obtained. The selectivity varied
between 45.2 - 53.0% at ethene concentrations between 0.7% and
20%. The corresponding conversion degree decreased from 42.7%
to 7.0%, respectively.

Influence of the total pressure: The impact of the total pres-


sure on selectivity and conversion degree was investigated at a re-
actor temperature of 503 K, using 20% C2 H4 in O2 - results are
depicted in figure 7.15. At atmospheric pressure (pressure controller
completely open to atmosphere), the selectivity was 48.6% at a con-
version degree of 5.3%. With increasing total pressure, the selectiv-
ity increased to a maximum of 58.6% at 0.9 MPa. The application
of higher pressures resulted in lower selectivities of about 54.8% at
2 MPa. Again, the degree of conversion increased with increasing
total pressure. At the highest selectivity, the conversion degree was
7.1%. With further increasing pressure, there was no more increase
in conversion degree observed. At 2 MPa, the conversion degree was
232 Chapter 7. Appendix

0,60

selectivity to ethylene oxide


0,55

0,50

0,070
degree of conversion

0,065

0,060

0,055

0,050
0,0 0,5 1,0 1,5 2,0
total pressure / MPa

Figure 7.15: Selectivity to ethylene oxide and degree of conversion as a


function of the total pressure for the Shell 800 Series Catalyst coated
modular microchannel reactor MMCR10. Reaction conditions: 20% C2 H4
in O2 , T= 503 K, = 230 ms.

still around 7.2%. It has to be noted, this experiment was performed


at the end of the experimental series and after a malfunction of the
ethylene mass flow controller resulting in an uncontrolled oscillation
overnight. As a result, some selectivity and conversion were lost. Be-
fore this incident, a reference point at 503 K, 0.3 MPa, = 190 ms)
was measured with 8% conversion and a selectivity of 55%. After
this incident, the conversion was 6% and the selectivity 51.5%. It
is unclear wether its a reduction effect or if small amounts of the
catalytic coating have been blown out by the pressure oscillations.
7.8. Additional figures and measurements for MCR2Cs 233

7.8 Additional figures and measure-


ments for MCR2Cs
Reactivation of MCR2Cs after Cs-impregnation: See Fig.
7.16. It took nearly two days on stream before a steady operation
was observed. The peak selectivities were initially nearly 80% and
stable selectivities close to 70%.
0,85
selectivity to ethylene oxide

0,80

0,75

0,70

0,15
degree of conversion

0,10

0,05

0,00
0 10 20 30 40 50
time on stream / h

Figure 7.16: Selectivity to ethylene oxide and degree of conversion as a


function of the time on stream for the Cs modified Ag/Al2 O3 /Al mi-
crochannel reactor MCR2Cs. Reaction conditions: 20% C2 H4 in O2 ,
T= 503 K, p= 0.3 MPa, = 1.0 s.

Influence of the oxygen concentration: The influence of the


oxygen partial pressure on this regenerated and Cs modified cata-
lyst / reactor was investigated at different reactor temperatures by
varying the oxygen concentration, keeping the ethene concentration
constant. Methane was used as balance. The results of those experi-
ments are depicted in figure 7.17. At 463 K, the conversion increased
with increasing oxygen partial pressure from 2.0% at an O2 concen-
tration of 5.8% to 6.5% when 80% O2 were applied. In the same
time, the selectivity increased only slightly from 68.1% to 70.0%, re-
spectively. With increasing reactor temperature, increasing degrees
234 Chapter 7. Appendix

0,70
selectivity to ethylene oxide
463K

0,65
483K

0,60

503K
0,55
503K
degree of conversion

0,15

483K
0,10

463K
0,05

0 10 20 30 40 50 60 70 80 90
oxygen concentration / %

Figure 7.17: Selectivity to ethylene oxide and conversion degree of


ethene as a function of the oxygen concentration for the Cs modi-
fied Ag/Al2 O3 /Al microchannel reactor MCR2Cs. Reaction conditions:
20% C2 H4 , balance CH4 , =1.1s (STP), p= 0.3 MPa, T= 463 K / 483 K /
503 K.

of conversion were observed. At 483 K, the conversion increased from


4.3 to 11.8%, and the selectivity increased from 62.9 to 69.1%. At
503 K, the conversion degree was varied between 7.6 to 18.2%, result-
ing in selectivities of 57.2 and 67.7%, respectively. Therefore, with
increasing reactor temperature and increasing degree of conversion,
the increase in selectivity from low to high oxygen partial pressures
increases. For comparison with MCR2, see Fig. 7.2 (p.219).

Influence of the ethene concentration: In order to investigate


the influence of the ethene concentration on the promoted and re-
generated microchannel reactor, the ethene concentration was varied
between 0.7 and 26% at reactor temperatures of 463 K and 503 K, us-
ing O2 as balance. The results are depicted in figure 7.18. In general,
the conversion degree decreased and the selectivity increased with
increasing ethene concentration. At 463 K, the selectivity increased
from 61.7% at 0.7% ethene concentration to 69.4% at 26% ethene
7.8. Additional figures and measurements for MCR2Cs 235

concentration. The corresponding conversion degree decreased from


44.8% to 5.05%, respectively. At higher reactor temperatures of
503 K, the overall observed selectivities decreased and the conver-
sion degrees increased. At 0.8% ethene, the selectivity was 55.7% at
a conversion degree of 71.9%. At higher ethene concentrations up
to 26%, the selectivity increased to 67.5% at conversion degrees of
16.1%. For comparison with MCR2, see Fig. 7.3 (p. 220).
selectivity to ethylene oxide

0,70 463K
503K
0,65

0,60

0,55

0,7
0,6
degree of conversion

0,5
0,4
0,3
0,2
503K
0,1
463K
0,0
0 10 20 30
ethene concentration / %

Figure 7.18: Selectivity to ethylene oxide and conversion degree of


ethene as a function of the ethene concentration for the Cs modified
Ag/Al2 O3 /Al microchannel reactor MCR2Cs. Reaction conditions: bal-
ance O2 , =1.1s, p= 0.3 MPa, T= 463 and 503 K.
236 Chapter 7. Appendix

Influence of the total pressure: The influence of the total pres-


sure on the Cs modified reactors performance is depicted in figure
7.19. Applying a total pressure of 0.1 MPa, the selectivity was 58.0%
at a conversion degree of 16%. With increasing pressure of up to 0.6 -
0.9 MPa, the selectivity increases to 67.8% at a conversion degree
of 19.8%. Higher total pressures lead to lower selectivities of 64.4%
and decreased conversion degrees of 17.3%.
selectivity to ethylene oxide

0,80

0,75

0,70

0,65

0,60
degree of conversion

0,20

0,15

0,10

0,05

0,00
430 440 450 460 470 480 490 500 510

reactor temperature / K

Figure 7.19: Selectivity to ethylene oxide and degree of conversion as


a function of the total pressure for the Cs modified Ag/Al2 O3 /Al mi-
crochannel reactor MCR2Cs. Reaction conditions: 20% C2 H4 in O2 ,
T= 503 K.
7.9. Additional figures and measurements for FBR4 237

7.9 Additional figures and measure-


ments for FBR4

Influence of the oxygen concentration: After the initial acti-


vation, performed in the same way as for MMCR9/10, the influence
of the O2 partial pressure was investigated. Selectivity and conver-
sion degree as a function of the O2 concentration at a constant space
velocity of this commercial catalyst are depicted in figure 7.20. With
increasing O2 partial pressure (from 6.5% up to 80%), the selectivity
increased from 50.3% to 61.3% and the conversion degree increased
simultaneously from 2% to 5.1%. Thus, the oxygen partial pressure
had a strong positive effect on selectivity and conversion degree.
selectivity to ethylene oxide

0,60

0,55

0,50

0,06
degree of conversion

0,04

0,02

0,00
0 10 20 30 40 50 60 70 80 90
oxygen concentration / %

Figure 7.20: Selectivity and conversion degrees as a function of the O2


partial pressure for the Shell 800 Series Catalyst in a fixed bed reac-
tor FBR4. Reaction conditions: T=503 K, p=3 bar, GHSV = 27 l/gh,
c(C2 H4 ) = 20%, balance CH4 .
238 Chapter 7. Appendix

Influence of the C2 H4 concentration: The influence of the


ethene partial pressure on selectivity and conversion (at a constant
residence time) is depicted in figure 7.21. Considering the positive
effect of a high oxygen concentration, it was decided to apply binary
CH4 / O2 mixtures in order to keep the oxygen partial pressure as
high as possible. With increasing ethene partial pressure, the selec-
tivity increased from 49.2% at 0.7% C2 H4 to 60.5% at 19.3% C2 H4
showing a strong initial increase in selectivity for ethene concentra-
tions up to 4%. The degree of conversion decreased simultaneously
from 28.3% to 5.12%.
selectivity to ethylene oxide

0,60

0,55

0,50

0,30

0,25
degree of conversion

0,20

0,15

0,10

0,05

0 5 10 15 20
ethene concentration / %

Figure 7.21: Selectivity and conversion degrees as a function of the C2 H4


partial pressure for the Shell 800 Series Catalyst in a fixed bed reac-
tor FBR4. Reaction conditions: T=503 K, p=3 bar, GHSV = 27 l/gh,
balance O2
Bibliography

[1] A.Wurtz, Justus Liebigs Ann. Chem., 110 (1859), p.125128

[2] Occupational Exposure to Ethylene Oxide, US Federal Reg-


ister, June 22, 49 (1984), 25734809; Information provided by
the Occupational Safety & Health Adimistration, US. Depar-
tent of Labor.

[3] Ullmanns Encyclopaedia of Industrial Chemistry, Release 2002,


6th Edition (CDROM).

[4] Francis Whellan, The History, Topography, and Directory of


the County Palatine of Durham, 2nd. Edition (1894) (partial
reprint, pictures and summary found at the Durham Mining
Museum)

[5] Ullmanns Encyclopaedia of Industrial Chemistry, 5th Edition,


Volume A10 (1987), VCH Weinheim , ISBN 3527201106,
p.117135

[6] S.A. Miller, Ethylene and its Industrial Derivatives, Ernest


Brenn Limited, London (1969), p.524562

[7] T.E. Lefort, French patent 729952, 1931, to Societe Francaise


de Catalyse Generale

[8] R.A. Van Santen, H.P.C.E. Kuipers, Adv. Catal. 35 (1987),


p.265321

239
240 Bibliography

[9] A.M.Lauritzen, US patent 4766105, 1986 and EP 0 226 015,


1987, to Shell Oil Company

[10] J.C.Zomerdijk, M.W.Hall, Catal. Rev.Sci. Eng. 23 (1981),


p.163185

[11] Report by M.T. Devanney, SRI Consulting / Access Intelligence


LLC (4/2007), http://www.sriconsulting.com

[12] Ethyleneoxide Users Guide, 2nd Edition, published by


C.Buckles, P.Chipman, M.Cubillas et al. for Shell Chemicals,
August 1999.

[13] Ullmanns Encyclopaedia of Industrial Chemistry, 5th Edition,


Volume A10 (1987), VCH Weinheim , ISBN 3527201106,
p.101115

[14] Sterigenics Food Safety, 344 Bonnie Circle, Corona, CA: Online:
http://www.sterigenics.com

[15] P.A. Kilty, W.M.H.Sachtler, Cat. Rev.Sci.Eng. 10 (1)


(1974), p.116

[16] A. Michaelides, M.L. Bocquet, P. Sautet, A. Alavi, D.A. King,


Chemical Physics Letters 367 (2003), p.344350

[17] V.I.Bukhtiyarov, V.V.Kaichev, Journal of Molecular Catalysis


A 158 1 (2000), p.167172

[18] W.Herzog, Ber.Bunsenges. Phys. Chem. 74 (1974), p.216220

[19] R.A. van Santen, C.P.M. De Groot, J. Catal. 98 (1986),


p.530539

[20] A.V.Khasin, Kinetics and Catalysis 34 No. 1 (1993), p.4253


(ISSN: 00231584/93/34010042)

[21] R.Hall, G.Neubauer, J. Catal. 105 (1987), p.3954

[22] N.C.Rigas, G.D.Svoboda, J.T.Gleaves, ACS Symp. Series 523


(1993), ISBN 0841226377, p.183203

[23] R.B.Grant, R.M.Lambert, J.Chem.Soc. Chem. Comm. 1983,


p.662663
Bibliography 241

[24] C.J.Bertole, C.A.Mims, J.Catal. 184 (1999), p.224235

[25] M.R.Salzar, J.D.Kress, A.Redondo, Surf. Sci. 469 (2000),


p.8090

[26] M.V.Badani, M.A.Vannice, App. Cat. A 204 (2000), p.129142

[27] L.M.Akella, H.H.Lee, J.Catal. 86 (1984), p.465472

[28] N.W.Cant, W.K.Hall, J.Catal. 52 (1978), p.8194

[29] P.C.Borman, K.R.Westerterp, Ind. Eng. Chem. Res. 34 (1995),


p.4958

[30] S.N.Goncharova, A.V.Khasin, S.N.Filimonova, D.A.Bulushev,


Kinetics and Catalysis 32 4 (1991), p.852859 (ISSN
00231584/91/32040768)

[31] E.P.S.Schouten, P.C.Borman, K.R.Westerterp, Chem. Eng.


Proc. 35 (1996), p.107120

[32] P.C.Borman, K.R.Westerterp, Chem. Eng. Sci. 47 (1992) 911,


p.25412546

[33] G.H.Law, S.Charleston, H.C.Chitwood, US patent 2279470,


1942, to Carbide and Carbon Chemicals Corporation, USA.

[34] G.L.Montrasi, G.C.Battiston, Ox. Comm. 3 (1983), p.259267

[35] K.L.Yeung, A.Gavriilidis, A.Varma, M.M.Bhasin, J. Catal. 174


(1998), p. 112

[36] D.Lafarga, A.Varma, Chem. Eng. Sci 55 (2000), p.749758

[37] R.A.Kemp, US patent 5545603, 1996, to Shell Oil Company

[38] N.Rizkalla, US patent 5945551, 1999, to Scientific Design Com-


pany

[39] M.Nakajima, US patent 4831162, 1989, to Mitsui Toatsu Chem-


icals, Inc., Japan

[40] C.T.Campbell, J.Catal. 99 (1986), p.2838

[41] C.F.Mao, M.A.Vannice, App. Cat. A 122 (1995), p.6176


242 Bibliography

[42] C.T.Campbell, B.E.Koel, J.Catal. 92 (1985), p.272283


[43] C.T.Campbell, M.T.Paffett, Appl. Surf. Sci. 19 (1984), p.2841
[44] S.A.Tan, R.B.Grant, R.M.Lambert, J.Catal. 100 (1986),
p.383391
[45] C.F.Mao, M.A.Vannice, App. Cat. A 122(1995), p.4159
[46] G.Boxhoorn, O.M. Velthuis, A.H.Klazinga, US patent 4731350,
1987, to Shell Oil Company
[47] J.R.Lockemeyer, US Patent 5929259, 1997, to Shell Oil Com-
pany
[48] H. Nakatsuji, H.Nakai, K.Ikeda, Y.Yanamoto, Surf. Sci. 384
(1997), p.315333
[49] M.C.N.A. Carvalho, C.A.Perez, R.Simao, F.B. Passos and
M.Schmal, Anais da Academia Brasileira de Ciencias 76(1)
(2004) p.1927
[50] Y. Jun, D.Jingfa, App. Cat. A 92 (1995), p.7380
[51] N.Rizkalla, US patent 5736483, 1198, to Scientific Design Com-
pany
[52] J.Alfansreder, US Patent 4335014, 1982, to Hoechst Aktienge-
sellschaft, Germany
[53] E.Vangermain, C.D.Mengler, R.Elm, H.Ueberschaer,
W.Brauckmann, US patent 4409394, 1983, to Chemische
Werke Huls AG, Germany
[54] G.B.Hoflund, D.M.Minahan, J. Catal. 162 (1996), p.4853
[55] H.V.Holler, US patent 3892679, 1975, to Shell Oil Company
[56] N.Nojiri, Y.Sakai, US patent 4786624, 1988, to Mitsubishi
Petrochemical Co, Japan
[57] V.M.Mastikhin, S.N.Goncharova, V.M.Taplilin, V.V.Terskikh,
B.S.Balzhiminaev, J.Mol.Cat. A 96 (1995), p.175179
[58] D.A.Bulushev, E.A.Paukshtis, Y.N.Nogin, B.S.Balzhinimaev,
App. Cat. A 123 (1995), p.301322
Bibliography 243

[59] D.A.Bulushev, S.N.Goncharova, E.A.Paukshtis,


B.S.Balzhinimaev, App. Cat. A 126 (1995), p.6784
[60] R.Baratti, A.Gavriilidis, M.Morbidelli, A.Varma, Chem. Eng.
Sci. 49 (1994), p.19251936
[61] M.Morbidelli, A.Servida, R.Paludetto, S.Carra, J.Catal. 87
(1984), p.116125
[62] E.P.S.Schouten, P.C.Borman, K.R.Westerterp, Chem. Eng. Sci.
No. 24a 49 (1994), p. 47254747
[63] P.C.Borman, K.R.Westerterp, Chem. Eng. Sci. 47 (1992)
p.25412546
[64] P.C.Borman, K.R.Westerterp, Chem. Eng. & Chem. Res. 34
(1995) p.4958
[65] J.Brandner, M.Fichtner, U.Schygulla, K.Schubert, IMRET 4
Proceedings, AIChE National Spring Meeting, Atlanta 2000,
p.244249 (ISBN 0816998825)
[66] H.O.Kuchling, Taschenbuch der Physik (13. Auflage), Fach-
buchverlag Leipzig, 1991, ISBN 3343007595, p.612-613
[67] M.T.Janicke, H.Kestenbaum, U.Hagendorf, F.Schuth,
M.Fichtner, K.Schubert, J. Catal. 191 (2000), p.282293
[68] T.Stief, PhD thesis, UniDortmund (2001), Shaker Verlag
Aachen, ISBN 3-8265-9334-0
[69] T.Zech, D.Honicke, IMRET 3 Proceedings (Ed: W.Ehrfeld),
Springer Verlag 2000, ISBN 3540669647, 2000, p.260266
[70] L.J.Rogers, US patent 3840341, 1974, to Ionics Inc.
[71] J.W.Parce, A.R.KopfSill, L.J.Bousse, US patent 6150180,
2000, to Caliper Tech Corp.
[72] S.H.DeWitt, Current Opinion in Chemical Biology 3 (1999),
p.350356.
[73] Buchi technical information data sheet, Pilot plant with Buchi
Chemreactors, document ID 40.9P211.E004, p.9. Reprint
http://www.buchiglas.ch/pdf-files/glass/pilot plant e.pdf
244 Bibliography

[74] T. Schwalbe, V. Autze, G. Wille, Chemical Synthesis in Mi-


croreactors, CHIMICA 56 (2002) 636646

[75] T.Schwalbe, V.Autze, M.Hohmann, W.Stirner, Organic Process


Research and Development 8 (2004), p.440454

[76] S.TaghaviMoghadam, A.Kleemann, K.Golbig, Organic Pro-


cess Research & Development 5 (2001), p.652658

[77] L.Shen, S.Kiang, G.Zhenrong, US Patent Application


20040230045 for BristolMyers Squibb

[78] David M. Minahan, Gar. B. Hoflund , J. Catal. 158 (1996),


p.109115

[79] A.Kursawe, R.Pilz, H.Durr, D.Honicke, IMRET 4 Proceedings,


AIChE National Spring Meeting, Atlanta 2000, p. 227235
(ISBN 0816998825)

[80] E.Dietzsch, PhD thesis, TUChemnitz (2004), online available


http://http://archiv.tuchemnitz.de/pub/2005/0095

[81] G.Wiessmeier, K.Schubert, D.Honicke, IMRET 1 Proceedings


(Ed. W.Ehrfeld), Springer Verlag 1998, p.2026

[82] A.Kursawe, E.Dietzsch, S.Kah, D.Honicke, M.Fichtner,


K.Schubert, G.Wiessmeier, IMRET 3 Proceedings (Ed.
W.Ehrfeld), Springer Verlag 2000, ISBN 3540669647,
p.213223

[83] G.Patermarakis, C.Pavlidou, J. Catal. 147 (1994), p.140155

[84] A.Gavriilidis, A.Varma, ACS Symp. Series 523 (1992), p.


410415

[85] R.B.Grant, R.M.Lambert, J.Catal. 92 (1985), p.364375

[86] S.Kah, PhD Thesis, TUChemnitz, 2004, online available


http://archiv.tuchemnitz.de/pub/2004/0098/

[87] G.Wiessmeier, D.Honicke, J.Micromech. Microeng. 6 (1996),


p.285289

[88] L.G.Britton, Plant/Operations Progress 9 (1990), p.7586


Bibliography 245

[89] Photo courtsey of CPCSystems GmbH, Germany. Device in-


stalled at Clariant facilities in FrankfurtHoechst.
[90] Forschungszentrum Karlsruhe GmbH, Mitglied der Hermann
von HelmholtzGemeinschaft Deutscher Forschungszen-
tren (HGF), HermannvonHelmholtzPlatz 1, 76344
EggensteinLeopoldshafen. http://www.fzk.de/
[91] K.H.Dittrich, W.Krysmann, P.Kurze, H.G.Schneider, Crystal
Research and Technology 19 1 (1984), p. 9399
[92] K. Fischer, K. Vester, US Patent 3067256, 1959.
[93] GESTIS Stoffdatenbank Database of the Berufsgenossen-
schaftliches Institut fur Arbeitssicherheit, Sankt Augustin /
Germany (2001)
[94] PS Gefahrstoffe 2.4 Database published by WEKA Fachver-
lag fur technische Fuhrungskrafte, Augsburg / Germany
(1997)
[95] Chemikalieninformationssystem fur verbraucherrelevante
Stoffe (CIVS) of the German Bundesinstitut fur gesund-
heitlichen Verbraucherschutz und Veterinarmedizin (BgVV)
(2001)
[96] SRAG Chemical Warehouses Version 6, 26 June 2002, p.37
http://www.hse.gov.uk/comah/sragcwh/hazards/images/hazards.pdf
[97] A.Kursawe, D.Honicke, IMRET 4 Proceedings, AIChE Na-
tional Spring Meeting, Atlanta 2000, ISBN 0816998825,
p.153166
[98] A.Kursawe, D.Honicke in Microreaction Technology, IM-
RET 5 Conference Proceedings (Ed: Matlosz, Ehrfeld, Baselt),
p.240251 ISBN 3540424989, Springer Verlag 2001.
[99] F.Keller, M.S.Hunter, D.L.Robinson, J. Electrochem. Soc. 11
(1953), S. 411419
[100] S.Kah, A.Kursawe, D.Honicke, Final Report for the AiF
Project 11148B/1 heterogen katalysierte Gasphasenoxidatio-
nen in Mikroreaktoren (http://www.aif.de)
[101] Y.S.Yong, N.W.Cant, Applied Catalysis 48 (1989), p.3750
Lebenslauf
Name Ansgar Kursawe
Adresse Am Schafersberg 32
65527 Niedernhausen

Geburtstag 25.4.1971
Geburtsort Essen
Eltern Hildegard und Hans-Theo Kursawe
Familienstand ledig
Nationalitat deutsch

Berufserfahrung
Siemens AG seit 05/2007
CPC-Systems GmbH 02/2002 bis 05/2007

Ausbildung
Promotion Experimenteller Teil 10/1996 - 01/2002
Technische Universitat Chemnitz

Studium 10/1991 - 7/1996


Chemie, Ruhr-Universitat Bochum
Abschluss Diplom
Vertiefungsausbildung Technische Chemie

Grundwehrdienst 7/1990 - 7/1991


Gymnasium 1981 - 1990
Grundschule 1977 - 1981
Selbstandigkeitserklarung

Ich erklare hiermit, da ich die vorliegende Arbeit selbstandig und


nur unter Verwendung der angegebenen Literatur und Hilfsmittel
verfasst habe.

Niedernhausen, den 27. September 2008 Ansgar Kursawe


Veroffentlichungen
A.Kursawe, E.Dietzsch, S.Kah, D.Honicke, M.Fichtner, K.Schubert,
G.Wiessmeier: Selective Reactions in Microchannel Reactors,
IMRET 3 Proceedings, W.Ehrfeld (Ed.), Springer Verlag 2000,
ISBN 3-540-66964-7, p. 213-223

A.Kursawe, D.Honicke: Epoxidation of Ethene with pure Oxygen,


IMRET 4 Proceedings, AIChE Spring National Meeting, Atlanta
2000, ISBN 0-8169-9882-5, p. 153-166

A.Kursawe, R.Pilz, H.Durr, D.Honicke: Development and Design


of a Modular Microchannel Reactor for Laboratory use, IMRET 4
Proceedings, AIChE Spring National Meeting, Atlanta 2000, ISBN
0-8169-9882-5, p. 227-235

A.Kursawe, D.Honicke: Comparison of Ag/Al and Ag/-Al2 O3


Catalytic Surfaces for the Partial Oxidation of Ethene in Mi-
crochannel Reactors, IMRET 5 proceedings, Springer Verlag 2005,
ISBN 3-540-42498-9, p. 240-251

A.Kursawe, D.Honicke, Ethene epoxidation in Ag/Al microchannel


reactors: Effects of NO2 and Cs, Catalysis Communications 2
(2001), p. 347-351

R.Fodisch, A.Kursawe, D.Honicke: Immobilizing heterogeneous


catalysts in microchannel reactors, IMRET 6 Proceedings, AIChE
Spring Meeting, New Orleans 2002, ISBN 0-8169-9779-9, p. 140-146

T.Schwalbe, V.Autze, K.Sahin, S.Oberbeck, A.Kursawe: A New


Technology for Accelerating Scale-Up from Bench to Pilot Plant
by Continuous Reaction, 2003 AIChE Annual Meeting Conference
Proceedings, New York, AIChE (ISBN 0-8169-0941-5)
K.Golbig, M.Hohmann, A.Kursawe, T.Schwalbe: Verweilzeitver-
halten in Mikrokanalen als Vorraussetzung fur den Bau sequenziell
arbeitender Syntheseautomaten, Chemie Ingenieur Technik 76
(2004), No. 5, p. 598-603

K.Golbig, A.Kursawe, M.Hohmann: Designing Microreactors in


Chemical Synthesis Residence-time Distribution of Microchannel
Devices, Chemical Engineering Communications 192 No.5 (2005),
p. 620-629

T.Schwalbe, A.Kursawe, J.Sommer: Application Report on


Operating Cellular Process Chemistry Plants in Fine Chemical
and Contract Manufacturing Industries, Chemical Engineering
Technology 28 No.4 (2005), p. 408-419

US Patent 7101515, T.Schwalbe, V.Autze, S.Oberbeck, A.Kursawe,


K.H.Sahin: System and method for determining optimal reaction pa-
rameters using continuously running process, see also EP1618372A2

Das könnte Ihnen auch gefallen