Keuler Optimising 2000 PDF
Keuler Optimising 2000 PDF
by
I N ENGINEERING
(Chemical Engineering)
Promoters:
PROF. L LORENZEN
PROF. RD SANDERSON
STELLENBOSCB
September 2000
DECLARATION
I hereby certifY that this dissertation is my own original work, except where specifically
acknowledged in the text. Neither the present dissertation, nor any part thereof, has previously
been submitted for a degree at any University
JNKEULER
September 2000
SUMMARY
Stricter government regulations and higher energy costs have forced the chemical industry to
focus more on environmentally friendly processes and to reduce energy consumption. The
main goals of chemical companies are to obtain a high product yield and selectivity, and to .
reduce unwanted side products. Furthermore, if reactions can be performed at lower
temperature, while maintaining the reaction conversion, it will result in large energy savings.
Low temperature dehydrogenation reactions (below 300°C) are very selective and do not
produce many by-products, but conversion is limited by the reaction equilibrium. The
conversion limitations have resulted in the development of alternative processes in recent
years for producing alkenes from alkanes and aldehydes or ketones from alcohols. Advances
in membrane technology have created the possibility of using a new type of reactor, called a
catalytic membrane reactor, in which separation and reaction occurs simultaneously. A
catalytic membrane reactor, of the palladium composite type, can selectively remove hydrogen
and manipulate the reaction equilibrium in dehydrogenation reactions. The possibility exists
to save energy, obtain high conversions and perform very selective reactions in the catalytic
membrane reactor.
This dissertation describes a thorough investigation carried out into the design, optimisation,
operation and modelling of a catalytic membrane reactor. The two components of the
membrane reactor, i.e. the catalyst packing and the membrane structure, were optimised
individually for the dehydrogenation of ethanol and 2-butanol. The optimised catalyst and
optimised membrane were combined and their combined performance compared to a
conventional plug flow reactor. A fundamental model was developed for the catalytic
membrane reactor and a full sensitivity analysis was conducted to test the effects of membrane
parameters, reaction rate parameters and process variables on reaction conversion.
11
dehydrogenation results and an optimum copper loading of 15% gave the highest
dehydrogenation conversion for both the ethanol and the 2-butanol reactions. The copper-
silica catalyst was stable at 280°C and below for ethanol dehydrogenation, and at 250 °C and
below for 2-butanol dehydrogenation. At higher temperatures in ethanol dehydrogenation, the
catalyst deactivated due to both sintering and coking. Kinetic data from 200°C to 300 °C for
ethanol dehydrogenation, and from 190°C to 280 °C for 2-butanol dehydrogenation, indicated
that both reactions could be well described by the dual site, surface reaction, controlling
mechanism.
Significant advances were made in the production of very thin Pd films (1.0 to 1.5 1IDl) on the
inside of 200 run a-alumina membrane tubes (from Societe des Ceramiques Techniques). A
modified electro less plating technique was used for producing the Pd films. Hydrogen
permeances through the films varied between about 8 and 15 1IDl01lm2.Pa.s for temperatures
from 330°C to 450 °C and palladium films from 1.0 to 1.5 1IDl. Hydrogen to nitrogen
selectivity was greater than 100 for all membranes tested and greater than 400 for all but two
membranes (thickness 1.0 to 1.5 1IDl). These values are a significant improvement over other
published results. Pd membranes can only be used above 300°C and since the catalyst was
unstable in that temperature region, Pd-Ag membranes had to be prepared. Pd-Ag films of
thickness less than 2.2 Ilm were successfully synthesised and tested. Good high temperature
(500°C) and low temperature (below 300°C) stability was obtained for the Pd-Ag
membranes.
III
OPSOMMING
Strenger wetgewing en hoilr energiekoste dwing die chemiese industrie om te kyk na meer
omgewingsvriendelike prosesse en om energieverbruik te minimeer. Die belangrikste
doelwitte van chemiese maatskappye is om hoe opbrengste en hoe selektiwiteite te verkry van
hul verlangde produkte. Ongewenste byprodukte moet so ver as moontlik uitgeskakel word.
Reaksies wat hoe omsettings gee by laer temperature sal lei tot groot energie besparings.
Lae temperatuur dehidrogeneringsreaksies (onder 300°C) is baie selektief met min newe
produkte, maar omsetting word beperk deur die reaksie ewewig. Ewewigsbeperkings in
konvensionele prosesse het gelei tot die ontwikkeling van nuwe prosesse in die laaste dekade
vir die produksie van alkene vanaf alkane en vir aldehiede en ketone vanaf alkohole.
Membraanontwikkeling het die moontlikheid geskep vir 'n nuwe generasie reaktore, die
katalitiese membraanreaktore, waarin skeiding en reaksie gelyktydig plaasvind. 'n Palladium-
tipe reaktor kan selektief waterstof skei en die ewewig verskuif in dehidrogeneringsreaksies.
Die moontlikheid bestaan om energie te bespaar en om hoil omsetting sowel as hoe
selektiwiteit te verkry in die katalitiese membraanreaktor.
Hierdie proefskrif beskryf 'n omvattende ondersoek van die ontwerp, bedryf, optimering en
modellering van 'n katalitiese membraanreaktor. Die twee komponente van die
membraanreaktor, nl. die katalis pakking en die membraanstruktuur, is individueel geoptimeer
vir die dehidrogenering van etanol en 2-butanol. Die optimum katalis en die optimum
membraan is gekombineer en hul gekombineerde werking is vergelyk met die werking van 'n
propvloeireaktor. 'n Fundamentele model is gefonnuleer vir die membraanreaktor en 'n
volledige sensitiwiteitsanalise is uitgevoer op die model. Die effek van die reaksietempo
parameters, membraan parameters en die proses veranderlikes op die reaksie omsetting is
ondersoek.
Koper kataliste is berei deur die impregnasie van alumina, silika en magnesIum oksied
partikels. Die lae oppervlak area van MgO het gelei tot lae omsettings terwyl alumina meer
dehidrasieprodukte as dehidrogeneringsprdukte opgelewer het. Koper op silika het die beste
IV
dehidrogeneringsresultate getoon met 'n optimum koperkonsentrasie van 15 massa % op
silika. Vir laasgenoemde koperkonsentrasie is die hoogste dehidrogeneringsomsettings verkry
vir beide die etanol en 2-butanol reaksies. Die koper-silika katalis was stabiel by 280°C en
laer vir etanol dehidrogenering en by 250°C en laer vir 2-butanol dehidrogenering. By hoer
temperature (etanol dehidrogenering) het die katalis begin deaktiveer weens sintering en
koolstofvorrning. Kinetiese data vanaf 200°C tot 300 °C vir etanol dehidrogenering en vanaf
190°C tot 280 °C vir 2-butanol dehidrogenering het daarop gedui dat die reaksies goed
beskryfkan word deur die dubbel posisie, oppervlak reaksie, beherende meganisme.
Goeie vordering is gemaak met die vervaardiging van baie dun Pd films (1.0 tot 1.5 /lm) op
die binnekant van 200 om a-alumina membraanbuise (verskaf deur Societe des Ceramiques
Techniques). 'n Gemodifiseerde elektrodelose plateringstegniek is gebruik vir die
vervaardiging van die films. Waterstof perrneasies deur die films het gewissel van 8 tot
15 /lmol/m 2.Pa.s vir temperature vanaf330 °C tot 450°C en palladium filmdiktes vanaf 1.0 tot
1.5/lm. Waterstoftot stikstofselektiwiteit was meer as 100 vir al die getoetste membrane en
meer as 400 vir almal behalwe twee membrane (dikte 1.0 tot 1.5 1J.IIl). Hierdie waardes is
beduidend beter as ander gepubliseerde resultate. Pd membrane kan slegs gebruik word bo
300°C en aangesien die kataliste onstabiel was in daardie temperatuurgebied, is Pd-Ag
membrane berei. Pd-Ag membrane met totale filmdikte kleiner as 2.2 IJ.IIl is suksesvol berei
en getoets. Die Pd-Ag membrane was stabiel tussen 200°C en 500 0c.
Die werking van 'n geoptimeerde membraan, gepak met 'n geoptimeerde 14.4 massa % koper-
silika katalis (die katalitiese membraanreaktor), is vergelyk met 'n propvloeireaktor. Die beste
resultate verkry vir etanol dehidrogenering was by 275 °e. By daardie temperatuur is die
etanol omsetting verhoog vanaf 45% vir die propvloeireaktor tot 60% vir die membraanreaktor
by lae etanol vloeitempo's en vanaf 36% tot 46% by hoe etanol vloeitempo's. Die maksimum
2-butanol omsetting vir die propvloeireaktor was 80% by 240°C en dit is verhoog tot bo 90%
vir die membraanreaktor.
v
TABLE OF CONTENTS
CHAPTER 1: INTRODUCTION 1
VI
2.5.4. Palladium and rare earth elements, uickel or others 22
2.6. PREPARING PALLADIUM MEMBRANES 23
2.6.1. Wet impregnation 24
2.6.2. Sol gel process 24
2.6.3. Vapour deposition techniques 25
2.6.3.1. Physical vapour deposition 25
2.6.3.2. Spray pyrolysis 26
2.6.3.3. Chemical vapour deposition 26
2.6.4. Electroplating 27
2.6.5. Electroless plating 27
2.6.5.1. Substrate pretreabnent 28
2.6.5.2. Electroless plating solution composition 29
2.6.5.3. Recent advances in electroless palladium plating 30
2.6.5.4. Electroless palladium-silver coatings 32
2.6.6. Palladium membrane temperature stability 33
2.6.7. Palladium-silver alloying 33
2.6.S. Deactivation of palladium membranes 34
2.7. APPLICATIONS OF INORGANIC MEMBRANES 35
2.S. CATALYTIC MEMBRANE REACTORS 36
2.S.1. General advantages of catalytic membrane reactors 37
2.S.2. Catalytic membrane applications 37
2.8.2.1. Membranes as distributors 38
2.8.2.2. Membranes for dehydrogenation reactions 38
2.8.2.3. Other applications 39
2.S.3. Problems with catalytic membrane reactors 39
2.9. DEHYDROGENATION OF ALKANES 40
2.9.1. General principles 41
2.9.2. Alkane dehydrogenation catalysts 41
2.9.3. Commercial dehydrogenation processes 42
2.9.3.1. FBD Process (Snamprogetti-Yarsintez) 43
2.9.3.2. UOP Oleflex process 44
Vll
2.9.3.3. Lummus Catofin process 44
2.9.3.4. STAR process (Phillips steam active reforming) 45
2.9.3.5. Linde-BASF process 45
2.9.4. Alternative dehydrogenation processes 45
2.9.4.1. Coupled dehydrogenation and hydrogen oxidation 46
2.9.4.2. Oxidative dehydrogenation 46
2.9.4.3. Dehydrogenation in a membrane reactor 46
2.9.5. Methane steam reforming reaction 46
2.9.6. Water gas shift reaction 48
2.9.7. Ethane dehydrogenation 49
2.9.8. Dehydrogenation of propane 50
2.9.9. Dehydrogenation of butanes and butenes 52
2.9.9.1. Iso-butane dehydrogenation 52
2.9.9.2. n-Butane and butene dehydrogenation 53
2.9.10. Dehydrogenation of ethylbenzene to styrene 55
2.10. ALCOHOL DEHYDROGENATION CATALYSTS 56
2.10.1. Effect of copper percentage on catalyst performance 57
2.10.2. Catalyst preparation techniques 58
2.1 0.2.1. Precipitation 59
2.10.2.2. Urea hydrolysis 59
2.10.2.3. Electroless plating 60
2.10.2.4. Impregnation 60
2.10.3. Catalyst reduction 60
2.10.4. Catalyst deactivation 62
2.10.4.1. Sintering 62
2.10.4.2. Coking and poisoning 63
2.11. DEHYDROGENATION KINETICS 64
2.12. DEHYDROGENATION OF ALCOHOLS 65
2.12.1. Dehydrogenation of methanol to yield formaldehyde 65
2.12.2. Dehydrogenation of iso-propanol to acetone 66
2.12.3. Dehydrogenation of cyclohexanol to cyclohexanone 67
viii
2.12.4. Dehydrogenation of C4 alcohols 69
2.12.5. Dehydrogenation of ethanol 69
2.12.5.1. Catalyst selectivity 70
2.12.5.2. Catalyst activity 72
2.12.5.3. Ethanol dehydrogenation kinetics 74
2.12.5.4. Oxidative dehydrogenation of ethanol 76
2.12.5.5. Dehydrogenation of ethanol in a membrane reactor 77
2.13. ALCOHOL DEHYDRATION REACTIONS 77
2.14. SUMMARY 78
3. EXPERIMENTAL PROCEDURES 80
IX
3.3.2.2. Electroless Pd plating 97
3.3.2.3. Membrane cleaning 100
3.3.2.4. Electroless silver plating 101
3.3.3. Determining palladium film thickness 102
3.3.4. Membrane testing 102
3.3.4.1. Reactor temperature profiles .107
3.3.5. Membrane heat treatment and reduction 108
3.3.6. Palladium-silver alloying 109
3.3.7. Membrane characterisation 110
3.4. CATALYTIC MEMBRANE REACTOR EXPERIMENTS 111
3.4.1. Ethanol dehydrogenation 112
3.4.2. 2-Butanol dehydrogenation _ 112
3.5. SUMMARY 113
x
4.1.3.4. Surface area data for silica-based catalysts 132
4.1.3.4.1. TEM and XRD data/or copper on silica catalysts 134
4.2. DEHYDROGENATION OF 2-BUTANOL 136
4.2.1. MgO catalysts 136
4.2.2. Silica catalysts 139
4.2.2 .1. Effect of particle size on 2-butanol conversion . 141
4.3. SUMMARY 142
4.3.1. Ethanol reaction 142
4.3.2. 2-Butanol reaction 143
Xl
6.3. THE ELECTROLESS Pd PLATING PROCESS 176
6.3.1. Solution feeding to membrane tube 177
6.3.2. Effect of plating rate on membrane performance 179
6.3.3. Pd membrane thickness measurement 181
6.3.4. Membrane post plating treatment 181
6.4. THE EFFECT OF SUPPORT STRUCTURE ON Pd FILMS .183
6.5. STRUCTURAL CHARACTERISATION OF Pd MEMBRANES 184
6.6. PERMEANCE TESTING OF Pd MEMBRANES 186
6.6.1. The effect of.iP on H2 and N2 permeance 187
6.6.1.1. Nitrogen experiments 187
6.6.1.2. Hydrogen experiments 188
6.6.2. The effect oftemperature. on H2 and N2 permeance 190
6.6.3. The effect of film thickness on permeance 192
6.6.3.1. Hydrogen penneance 192
6.6.3.2. Nitrogen penneance 193
6.6.3.3. Membrane selectivity 194
6.6.3.4. Arrhenius parameters for hydrogen penneance 194
6.6.4. Comparison with literature data 195
6.6.4.1. Hydrogen penneances in the present study 196
6.6.5. H2 permeance employing a sweep gas 197
6.7. SUMMARY 204
7. Pd-Ag MEMBRANE ALLOYING AND STABILITY TESTING 206
7.1. Pd-Ag MEMBRANE PREPARATION 206
7.2. UNALLOYED Pd-Ag MEMBRANES 207
7.2.1. Characterising unalloyed Pd-Ag membranes 209
7.2.1.1. Unalloyed Pd-Ag membranes tested with a sweep gas 211
7.3. LITERATURE DATA ON ALLOYING Pd-Ag MEMBRANES 214
7.3.1. Alloying co-deposited Pd-Ag films 214
7.3.2. Alloying successive Pd-Ag films 214
7.3.3. Diffusion kinetics 215
7.4. ALLOYING RESULTS FOR Pd-Ag MEMBRANES 216
Xli
7.4.1. Alloying in a tube furnace 217
7.4.1.1. Alloying at a temperature of 545°C 217
7.4.1.2. Alloying at a temperature of 530 °C 219
7.4.1.3. Alloying at a temperature of 590°C 220
7.4.2. Temperature stability of Pd-Ag membranes 220
7.4.2.1. Literature data on Pd membrane stability .220
7.4.2.2. Pd-Ag membrane stability 221
7.4.2.3. Structural changes ofPd-Ag membranes during heating 224
7.4.3. Alloying in situ in the reactor 227
7.4.3.1. Alloying in a hydrogen environment 227
7.4.3.2. Alloying in nitrogen and argon environments 229
7.4.3.3. Further treatment ofpd-Ag films 231
7.4.3.4. Suggested heat treatment procedure for Pd-Ag films 232
7.5. FULL CHARACTERISATION OF MEMBRANES N4x AND N8x 233
7.6. SUMMARY 234
XIll
8.4. EXPERIMENTAL DATA FOR A PLUG FLOW REACTOR 253
8.4.1. Ethanol dehydrogenation in a plug flow reactor 253
8.4.2. 2-butanol dehydrogenation in a plug flow reactor 256
8.4.3. Summary of model for ethanol dehydrogenation 257
8.4.4. Summary of model for 2-butanol dehydrogenation 262
8.5. SELECTIVITY ANALYSIS OF THE PROCESS MODEL 266
8.5.1. The effect ofk'-values on model performance 267
8.5.1.1. Back diffusion 268
8.5.2. The effect of K. c • t on model performance 268
8.5.3. The effect of the membrane selectivity on model performance 269
8.5.4. The effect of the reaction effectiveness factor on model performance 270
8.5.5. The effect ofthe ethanol feed rate on model performance 270
8.5.6. The effect of the sweep gas on model performance 271
XIV
9.2.4. Further discussion of 2-butanol dehydrogenation 300
9.3. SUMMARY 302
APPENDIX D1: Hydrogen and nitrogen permeance data for Pd films 358
APPENDIX D2: Hydrogen and nitrogen permeance data for Pd-Ag films 378
xv
APPENDIX Fl: Plug flow reactor data and membrane reactor data for
ethanol dehydrogenation 402
APPENDIX F2: Plug flow reactor data and membrane reactor data for
2-butanol dehydrogenation 410
APPENDIXG: Sensitivity analysis for the catalytic membrane reactor model 420
XVI
LIST OF FIGURES
CHAPTER 1:
Figure 1.1: Project flow diagram 5
CHAPTER 2:
Figure 2.1: Separation by a semi-penneable membrane 6
Figure 2.2: Knudsen diffusion 15
Figure 2.3: Surface diffusion 15
Figure 2.4: Capillary condensation 15
Figure 2.5: Molecular sieve separation 15
Figure 2.6: Hydrogen and oxygen flow through a non-porous membrane 17
Figure 2.7: Equilibrium solubility isothenns ofPdHn for bulk Pd (Shu et aI., 1991) 20
Figure 2.8: Penneability ofH2 through Pd alloy membranes at 350°C and 2.2 Mpa 23
Figure 2.9: Pore blocking by deposited copper 58
Figure 2.10: The effect of temperature on copper surface structure 61
CHAPTER 3:
Figure 3.1: Set-up used for testing the kinetics of the catalyst at the CNRS, France 84
Figure 3.2: SCT membrane structure 94
Figure 3.3: Curing process for enamelled membranes 94
Figure 3.4: Pre-treatment set-up 96
Figure 3.5: Electroless Pd plating set-up 99
Figure 3.6: Membrane reactor used to test the membrane penneance 103
Figure 3.7: Set-up used for membrane testing at room temperature 103
Figure 3.8: Set-up used for high temperature (> 300°C) hydrogen and nitrogen
penneance testing 105
Figure 3.9: Set-up used for testing ethanol penneance 106
Figure 3.10: Reactor temperature profiles at different oven temperatures 108
Figure 3.11: Membrane packed with catalyst III
XVll
CHAPTER 4:
Figure 4.1: Total ethanol conversion for Cu on MgO catalysts 115
Figure 4.2: Ethanol to acetaldehyde yield (Cu on MgO catalysts) 115
Figure 4.3: Acetaldehyde selectivity for Cu on MgO catalysts 115
Figure 4.4: Total ethanol conversion (Cu on alumina catalysts) 118
Figure 4.5: Ethanol to ethene yield (Cu on alwnina catalysts) 118
Figure 4.6: Ethanol to di-ethyl ether yield (Cu on alwnina catalysts) 118
Figure 4.7: Ethanol conversion to di-ethyl ether (0% Cu on alumina) at
different W IF ratios 118
Figure 4.8: Acetaldehyde yields for Cu on alumina catalysts (WIF = 393 kg.s/mol) 119
Figure 4.9: Butene yields for Cu on alumina catalysts (average WIF) 119
Figure 4.10: Effect of Cu% on catalyst surface area for Cu on alumina catalysts 120
Figure 4.11: Effect of Cu% on Cu surface area for alumina-based catalysts 120
Figure 4.12: Pore size distribution for Cu on alwnina catalysts 121
Figure 4.13: Effect ofCr, Ni and Co (on alumina) on total ethanol conversion 125
Figure 4.14: Effect ofCr, Ni and Co (on alumina) on ether yield 125
Figure 4.15: Effect ofCr, Ni and Co (on alwnina) on ethene yield 126
Figure 4.16: Total ethanol conversion (Cu on silica catalysts) 128
Figure 4.17: Ethanol to acetaldehyde yield (Cu on silica catalysts) 128
Figure 4.18: Ethanol to acetaldehyde selectivity (Cu on silica catalysts) 128
Figure 4.19: Total ethanol conversion (15 wt % Cu on silica catalyst) 129
Figure 4.20: Ethanol converted to acetaldehyde (IS wt % Cu on silica) 129
Figure 4.21: Effect of Co, Cr orNi additives on total ethanol conversion (Cu on silica) 130
Figure 4.22: Effect of Co, Cr or Ni additives on acetaldehyde yield (Cu on silica) 130
Figure 4.23: Effect of Co, Cr or Ni additives on total ethanol conversion (Cu on silica) 131
Figure 4.24: Effect of Co, Cr or Ni additives on acetaldehyde yield (Cu on silica) 131
Figure 4.25: Effect of catalyst particle size on total ethanol conversion 131
Figure 4.26: Effect of catalyst particle size on acetaldehyde yield percentage 131
Figure 4.27: Effect ofCu% on catalyst surface area for Cu on silica catalysts 133
Figure 4.28: Effect of Cu% on Cu surface area for silica-based catalysts 133
Figure 4.29: Pore size distribution for copper on silica catalysts 133
XVI11
Figure 4.30: TEM images of the unused 14.4 wt % copper on silica catalyst 135
Fig 4.30a: TEM (20 nrn) 135
Fig 4.30b: TEM (20 nrn) 135
Fig 4.30c: TEM (20 nrn) 135
Fig 4.30d: TEM (20 nrn) 135
Fig 4.30e: TEM (20 nrn) 135
Fig 4.30f: TEM (20 nrn) 135
Fig 4.30g: TEM (50 nrn) 135
Fig 4.30h: TEM (200 nrn) 135
Fig 4.30i: TEM (200 nrn) 135
Figure 4.31 : Total2-butanol conversion for Cu on MgO catalysts 137
Figure 4.32: Total butene yield for Cu on MgO catalysts 137
Figure 4.33: MEK yield for Cu on MgO catalysts 138
Figure 4.34: MEK yield for a 16.9 wt % Cu on MgO catalyst 138
Figure 4.35: MEK selectivity for a 16.9 wt % Cu on MgO catalyst 138
Figure 4.36: Total2-butanol conversion for Cu on silica catalysts 139
Figure 4.37: MEK yield for Cu on silica catalysts 139
Figure 4.38: Butene yields for Cu on silica catalysts 140
Figure 4.39: MEK yield for a 15 wt % Cu on silica catalyst 141
Figure 4.40: MEK selectivity for a 15 wt % Cu on silica catalyst 141
Figure 4.41 : Effect of catalyst particle size on total 2-butanol conversion 141
Figure 4.42: Effect of catalyst particle size on MEK yield percentage 141
Figure 4.43: Equilibrium ethanol conversion vs. measured values for a 15 wt %
Cu on silica catalyst 143
Figure 4.44: Equilibrium 2-butanol conversion vs. measured values for a 15 wt %
Cu on silica catalyst 144
CHAPTERS:
Figure 5.1: Effect of reduction T on acetaldehyde production rate over time
(14.4 wt % Cu on silica) 147
XIX
Figure 5.2: Effect of reduction T on acetaldehyde production rate over time
(14.4 wt % Cu on silica) 147
Figure 5.3: Acetaldehyde production rate as a function of time
and temperature (14.4 wt % Cu on silica) 148
Figure 5.4: Effect of additives (Cr and Co) on acetaldehyde production rate 149
Figure 5.5: Effect of additives (Cr and Co) on acetaldehyde production rate at 280°C 149
Figure 5.6: TEM images of a 14.4 wt % copper on silica catalyst after
being in use 400°C lSI
Figure 5.6a: TEM (20nm) 150
Figure 5.6b: TEM (20nm) 150
Figure 5.6c: TEM (50 nm) 150
Figure 5.6d: TEM(50nm) 150
Figure 5.6e: . TEM(100nm) 150
Figure 5.6f: TEM(100nm) 150
Figure 5.6g: TEM (100 nm) lSI
Figure 5.6h: TEM(200nm) 151
Figure 5.6i: TEM (1 fUll) 151
Figure 5.7: Catalyst activity (14.4 wt % Cu on silica) after re-oxidation 152
Figure 5.8: MEK production rate as a function of time for a 14.4 wt %
Cu on silica catalyst 154
Figure 5.9: The effect of feed flow rate on acetaldehyde production rate ISS
Figure 5.10: The effect of 2-butanol feed flow rate on the MEK production rate 156
Figure 5.11: Linear fits of reaction rate data at I atm. total pressure and 225°C 158
Figure 5.12: Linear fits of reaction rate data at I atm. total pressure and 300 °C 158
Figure 5.13: % Catalyst deactivation after 24 hours of operation 160
Figure 5.14: Parameters for ethanol reaction equation as a function of temperature 161
Figure 5.15: Difference between model 1 rates and actual rates at different
temperatures 164
Figure 5.16: Difference between model 2 rates and actual rates at different
temperatures 164
Figure 5.17: Percentage deviation between model 1 rates and measured rates 165
xx
Figure 5.18: Percentage deviation between model 2 rates and measured rates 165
Figure 5.19: Modell rates vs. actual rates at different rate values 165
Figure 5.20: Model 2 rates vs. actual rates at different rate values 165
Figure 5.21: Percentage deviation between model values and measured values
at 573 K 166
Figure 5.22: Linear fits of reaction rate data at I atm. total pressure and 220°C 168
Figure 5.23: Linear fits of reaction rate data at 1 atm. total pressure and 280 °C 168
Figure 5.24: Parameters for 2-butanol reaction equation as a function oftemperature 170
Figure 5.25 Comparison between measured reaction rates and model reaction rates 172
CHAPTER 6:
Figure 6.1: Cross section view of a three layer SCT membrane 175
Figure 6.2: Top view (20 OOOx) of a three layer SCT membrane 175
Figure 6.3: Top view (5 OOOx) of a three layer SCT membrane 175
Figure 6.4: Pd concentration in solution after repeated plating sessions 178
Figure 6.5: Cross section of membrane (a) (10 OOOx) 179
Figure 6.6: Top view of membrane (a) (25 OOOx) 179
Figure 6.7: Top view of membrane (a) (50 OOOx) 180
Figure 6.8: Cross section of membrane (b) (10 OOOx) 180
Figure 6.9: Top view of membrane (b) (25 OOOx) 180
Figure 6.10: Top view of membrane (b) (5000x) 181
Figure 6.11: Top view of membrane (b) (25000x) 181
Figure 6.12: Effect of oxidation post treatment on H2 permeance
(membrane 2, 1.43 jl111 Pd) 182
Figure 6.13: Effect of oxidation post treatment on selectivity
(membrane 2, 1.43 jl111 Pd) 182
Figure 6.14: Cross section of membrane (3b) (10 OOOx) 184
Figure 6.15: Top view of membrane (3b) (5000 x) 185
Figure 6.16: Top view of membrane (3b) (25 OOOx) 185
Figure 6.17: Cross section of membrane (11) (10 OOOx) 185
Figure 6.18: Top view of membrane (11) (5000 x) 186
XXI
Figure 6.19: Top view of membrane (11) (25 OOOx) 186
Figure 6.20: Effect of pressure on N2 permeance for a 1.47 micron Pd film (6) 188
Figure 6.21: Effect of pressure on N2 permeance for a 2.4 micron Pd film (3a) 188
Figure 6.22: Effect of pressure on N2 permeance for a 3.08 micron Pd film (la) 188
Figure 6.23: Effect of pressure on N2 permeance for a 4.43 micron Pd film (3b) 188
Figure 6.24: Effect of pressure on H2 permeance for a 1.47 micron Pd film (6) 189
Figure 6.25: Effect of pressure on H2 permeance for a 2.4 micron Pd film (3a) 189
Figure 6.26: Effect of pressure on Hz permeance for a 3.08 micron Pd film (la) 189
Figure 6.27: Effect of pressure on Hz permeance for a 4.43 micron Pd film (3b) 189
Figure 6.28: Selectivity data for membrane (3a) 190
Figure 6.29: Selectivity data for membrane (N8) 190
Figure 6.30: Selectivity data for membrane (N2) 191
Figure 6.31: Hydrogen permeance in J.lmollmz.Pa.s (membrane 2a) 191
Figure 6.32: Hydrogen permeance in J.lmollm 2.Pa.s (membrane N7) 191
Figure 6.33: Hydrogen permeance for Pd films from 1 to 6.5 micron thickness 192
Figure 6.34: Hydrogen permeance for Pd films from 1 to 1.5 micron thickness 193
Figure 6.35: Nitrogen permeance for Pd films from 1 to 6.5 micron thickness 193
Figure 6.36: Nitrogen penneance for Pd films from 1 to 1.5 micron thickness 193
Figure 6.37: Hz to N z selectivity for Pd films from 1 to 6.5 micron thickness 194
Figure 6.38: Hz to Nz selectivity for Pd films from 1 to 1.5 micron thickness 194
Figure 6.39: Arrhenius parameters for hydrogen permeance (1 to 6.5 micron Pd films) 195
Figure 6.40: Arrhenius parameters for hydrogen permeance (1 to 1.5 micron Pd films) 195
Figure 6.41: % Hz permeated with N2 sweep gas and space time = 2.37 s 201
Figure 6.42: % Hz permeated with N z sweep gas and space time = 1.19 s 201
Figure 6.43: Hydrogen flow in tube (cmJ/min) as predicted by model for
N7 (t = 2.37 s) 202
Figure 6.44: Hydrogen flow in tube (cmJ/min) as predicted by model for
N7 (t = 2.37 s) 202
Figure 6.45: Hydrogen flow in tube (cml/min) as predicted by model for
N7 (t = 1.19 s) 202
XXll
Figure 6.46: Hydrogen flow in tube (cm3/min) as predicted by model for
N7 (,t = 1.19 s) 202
Figure 6.47: % H2 permeated with N z sweep gas and space time = 2.37 s 203
Figure 6.48: % H2 permeated with N2 sweep gas and space time = 1.19 s 203
Figure 6.49: Hydrogen flow in tube (cm3/min) as predicted by model for
N4 (1: = 2.37 s) 203
Figure 6.50: Hydrogen flow in tube (cm3/min) as predicted by model for
N4 (1: = 2.37 s) 203
Figure 6.51: Hydrogen flow in tube (cm 3/min) as predicted by model for
N4 (1:= 1.19 s) 204
Figure 6.52: Hydrogen flow in tube (cm3/min) as predicted by model for
N4 (1:= 1.19 s) 204
CHAPTER 7:
Figure 7.1: Hydrogen permeances forPd and Pd-Ag membranes (8, NI, 8b, Nib) 209
Figure 7.2: Hydrogen permeances for Pd and Pd-Ag membranes (N3, N4, N3b, N4b) 210
Figure 7.3: Measured vs. calculated % H2 permeance for 8b 212
Figure 7.4: Measured vs. calculated % H2 permeance for NI b 212
Figure 7.5: Measured vs. calculated % H2 permeance for N3b 212
Figure 7.6: Measured vs. calculated % H2 permeance for N4b 212
Figure 7.7: % H2 permeated with N2 sweep gas and space time = 2.37 s 213
Figure 7.8: % H2 permeated with N2 sweep gas and space time = 1.19 s 213
Figure 7.9: Hydrogen flow in tube (cm3/min) as predicted by model for
N3b (1: = 2.37 s) 213
Figure 7.10: Hydrogen flow in tube (cm3/min) as predicted by model for
N3b (,t = 2.37 s) 213
Figure 7.11: Hydrogen flow in tube (cm3/min) as predicted by model for
N3b (F 1.19 s) 213
Figure 7.12: Hydrogen flow in tube (cm3/min) as predicted by model for
N3b (1: = 1.19 s) 213
XX111
Figure 7.13: Heating times required to obtain similar Pd-Ag diffusion at different
temperatures 216
Figure 7.14: Effect of heating time in argon on hydrogen permeance through a
Pd-Ag film (membranes N3b to N3e) 208
Figure 7.15: Effect of heating in argon on Hz permeance for N4 220
Figure 7.16: Effect of heating in argon on N z permeance for N4 222
Figure 7.17: Effect of heating in argon on nitrogen permeance (N3) 223
Figure 7.18: Effect of heating in argon on selectivity (N3) 223
Figure 7.19: Membrane dl (25 OOOx) 225
Figure 7.20: Membrane d2 (25 OOOx) 225
Figure 7.21: Membrane d3 (25 OOOx) 225
Figure 7.22: Membrane dl (50000x) 225
Figure 7.23: Membrane d2 (50 OOOx) 225
Figure 7.24: Membrane d3 (50 OOOx) 225
Figure 7.25: Membrane dl (5 OOOx) 226
Figure 7.26: Membrane d2 (5 OOOx) 226
Figure 7.27: Membrane d3 (5 OOOx) 226
Figure 7.28: Membrane d3 (2000x) 226
Figure 7.29: Side view of d2 226
Figure 7.30: Side view of d3 226
Figure 7.31: Effect of heating time in Hz on Hz and N2 permeance for N4x 228
Figure 7.32: Effect of heating time in H2 at 590°C on Hz permeance for NIx 228
Figure 7.33: Effect of heating time in Hz at 590°C on Nz permeance for NIx 228
Figure 7.34: Effect of temperature on Hz and N z permeance for NIx 229
Figure 7.35: Effect of heating time in N2 and Ar at 500 and 550°C on Hz
permeance for N2x 230
Figure 7.36: Effect of heating time in N2 at 500°C on N2 permeance for N2x 230
Figure 7.37: Effect of heating time in Ar at 500 and 550°C on Ar permeance for N2x 231
Figure 7.38: Effect of temperature on H2 permeance for N2x 231
Figure 7.39: Hz and N z permeances for N4x 233
Figure 7.40: Hz and N2 permeances for N8x 233
XXIV
Figure 7.41: Measured vs. calculated % Hz permeance for N4x 234
Figure 7.42: Measured vs. calculated % Hz permeance for N8x 234
CHAPTER 8:
Figure 8.1: Theoretical effect of feed rate on equilibrium conversion 236
Figure 8.2: Comparison between Hz to N z and Hz to ethanol selectivities
for a 1.5 flm Pd film 238
Figure 8.3: Ethanol to nitrogen permeance ratio for a 1.5 flm Pd film 238
Figure 8.4: Description of catalytic membrane reactor process 241
Figure 8.5: Effect of Reynolds number on effectiveness factor 251
Figure 8.6: C! vs. inverse Re at different temperatures 252
Figure 8.7: C! vs. inverse ReO. 8 ! at different temperatures 252
Figure 8.8: Plug flow reactor data for ethanol dehydrogenation at 250°C 254
Figure 8.9: Plug flow reactor data for ethanol dehydrogenation at 275 °C 254
Figure 8.10: Plug flow reactor data for ethanol dehydrogenation at 300°C 254
Figure 8.11: Selectivity towards acetaldehyde production for a plug flow reactor 254
Figure 8.12: Re as a function of temperature and conversion
(feed rate = 4.77*10. 5 molls) 255
Figure 8.13: Re as a function oftemperature and conversion
(feed rate = 9.54*10. 5 molls) 255
Figure 8.14: Re as a function of temperature and conversion
(feed rate = 1.43*10.4 molls) 255
Figure 8.15: Plug flow reactor data for 2-butanol dehydrogenation at 190°C 256
Figure 8.16: Plug flow reactor data for 2-butanol dehydrogenation at 215°C 256
Figure 8.17: Plug flow reactor data for 2-butanol dehydrogenation at 240°C 256
Figure 8.18: Selectivity towards MEK production for a plug flow reactor 256
Figure 8.19: Effect of sweep gas to standard feed molar ratio on ethanol conversion
and ethanol losses 275
xxv
CHAPTER 9:
Figure 9.1: Ethanol conversion at 250°C vs. sweep gas flow rate
(Fe' = 2.39*10.5 molls) 280
Figure 9.2: Ethanol conversion at 250°C vs. sweep gas flow rate
(Fe' = 4.77*10-5 molls) 280
Figure 9.3: Measured acetaldehyde selectivity (250°C) 282
Figure 9.4: Ethanol conversion at 275°C vs. sweep gas flow rate
(Fe' = 4.77*10"5 molls) 283
Figure 9.5: Ethanol conversion at 275°C vs. sweep gas flow rate
(Fe' = 9.54*10- 5 molls) 283
Figure 9.6: Ethanol conversion at 275°C vs. sweep gas flow rate
(Fe' = 1.43*10-4 molls) 284
Figure 9.7: Acetaldehyde yield at 275°C for a membrane reactor 284
Figure 9.8: Measured acetaldehyde selectivity (275°C) 286
Figure 9.9: Ethanol conversion at 300°C vs. sweep gas flow rate
(Fet = 4.77*10- 5 molls) 287
Figure 9.10: Ethanol conversion at 300°C vs. sweep gas flow rate
(Fe' = 9.54*10-5 molls) 287
Figure 9.11: Measured acetaldehyde selectivity (300°C) 288
Figure 9.12: Measured acetaldehyde selectivity at a constant feed rate
(Fe' = 9.54*10-5 molls) 290
Figure 9.13: 2-Butanol conversion at 190°C vs. sweep gas flow rate
(FZBu,= 1.52*10-5 molls) 292
Figure 9.14: 2-Butanol conversion at 190°C vs. sweep gas flow rate
(FzBu ,= 3.04*10-5 molls) 292
Figure 9.15: 2-Butanol conversion at 190°C vs. sweep gas flow rate
(F2Bu,= 4.56* 10-5 molls) 292
Figure 9.16: MEK yield at 190°C 292
Figure 9.17: Measured MEK selectivity at 190°C 294
Figure 9.18: 2-Butanol conversion at 215°C vs. sweep gas flow rate
(FzBu,= 1.52*10-5 molls) 295
XXVI
Figure 9.19: 2-Butanol conversion at 215°C vs. sweep gas flow rate
(F 2But = 3.04*10-5 moUs) 295
Figure 9.20: 2-Butanol conversion at 215°C vs. sweep gas flow rate
(F 2But = 6.08*10-5 moUs) 296
Figure 9.21: MEK yield at 215°C 296
Figure 9.22: Measured MEK selectivity at 215°C 297
Figure 9.23: 2-Butanol conversion at 240°C vs. sweep gas flow rate
(F 2But = 1.52*10-5 moUs) 298
Figure 9.24: 2-Butanol conversion at 240°C vs. sweep gas flow rate
(F 2But = 3.04*10.5 moUs) 298
Figure 9.25: 2-Butanol conversion at 240°C vs. sweep gas flow rate
(F 2But = 6.08*10. 5 moUs) 299
Figure 9.26: MEK yield at 240°C 299
Figure 9.27: Measured MEK selectivity at 240°C 300
Figure 9.28: MEK yield % VS. flow rate and temperature 301
Figure 9.29: Effect of sweep ratio on MEK yield % 301
XXVll
LIST OF TABLES
CHAPTER 2:
Table 2.1: The effect of repairing of electro less Pd plated coatings (Li et aI., 1999) 31
Table 2.2: Operating conditions for various dehydrogenation processes 42
Table 2.3: Process efficiency for propane and iso-butane dehydrogenation 43
Table 2.4: Catalyst supports (Carrizosa and Munuera, 1977; Rosynek et aI., 1990) 57
Table 2.5: Possible rate equations for solid catalysed dehydrogenation reactions 64
Table 2.6: Common supports for copper catalysts 67
Table 2.7: Selectivities for ethanol dehydrogenation (Iwasa and Takezawa, 1991) 71
Table 2.8: Effect of Cr and Co addition to Cu on ethanol dehydrogenation
(Church et aI., 1951) 71
Table 2.9: Reaction rate parameters for ethanol dehydrogenation (Tu et aI., 1994b) 72
Table 2.10: Activities and TOFs for ethanol dehydrogenation at 190°C
Kanoun et al. (1991, 1993) 73
CHAPTER 3:
Table 3.1: Characteristics of alumina and silica supports used 81
Table 3.2: Different copper loadings investigated for copper-supported catalysts 83
Table 3.3: Other copper supported catalysts 83
Table 3.4: Reactor conditions used to test Cu on Si02 and MgO supports in
the ethanol dehydrogenation reaction 85
Table 3.5: Reactor conditions used to test Cu on Al20 3 supports in the ethanol
dehydrogenation reaction 86
Table 3.6: Reactor conditions used to test Cu on SiOz and MgO supports in
the ethanol dehydrogenation reaction 86
Table 3.7: Different combinations of reduction and reaction temperatures studied 89
Table 3.8: Summary of catalyst deactivation tests for ethanol dehydrogenation 89
Table 3.9: Determining the region free of interphase mass transfer resistance 91
Table 3.10: Conditions investigated for determining interphase mass transfer
limited regime for 2-butanol dehydrogenation 93
XXVlll
Table 3.11: Membrane layer characteristics of a SCT membrane 94
Table 3.12: Compositions of pre-treatment solutions 96
Table 3.13: Stirring sequence and times used in pre-treatment 97
Table 3.14: Composition of the Pd plating solution per litre (for 2.00 g/litre Pd
in solution) 98
Table 3.15: Plating procedure used for producing Pd films 100
Table 3.16: Composition of silver plating bath per litre of plating solution 101
Table 3.17: Plating procedure used for producing Ag films 102
CHAPTER 4:
Table 4.1: BET surface areas for Cu on MgO catalysts in m2/g 116
Table 4.2: Yield matrix indicating main products and by-products for Cu/MgO 117
Table 4.3: Surface areas of unsupported copper catalysts 123
Table 4.4: Surface areas of supported copper catalysts 124
Table 4.5: BET and copper surface areas for Cu/alumina catalysts with additives 126
CHAPTERS:
Table 5.1: Summary of catalyst deactivation at different reaction temperatures 148
Table 5.2: Performance ofCU/Cr/Co on silica catalyst 149
Table 5.3: XRD and TOC results of tested catalysts 150
Table 5.4: Copper surface areas for 14.4 wt % Cu on silica catalysts 152
Table 5.5: Reaction rate parameters for ethanol dehydrogenation 159
Table 5.6: Kinetic model parameters for ethanol dehydrogenation 164
Table 5.7: Reaction rate parameters for 2-butanol dehydrogenation 168
Table 5.8: Kinetic model parameters for 2-butanol dehydrogenation 171
CHAPTER 6:
Table 6.1: Composition (per litre) of plating solutions for membrane plating 177
Table 6.2: Experimental H2 permeances vs. calculated values for membrane (N7) 199
Table 6.3: Experimental H2 permeances vs. calculated values for membrane (N4) 200
XXIX
",!,
CHAPTER 7:
Table 7.1: Alloying procedures used for Pd-Ag membranes 208
Table 7.2: N2 permeance and selectivity data for 8b, N1 b, N3b and N4b 211
Table 7.3: Ag-Pd diffusion coefficients 215
Table 7.4: Effect of heating on hydrogen permeances for membrane N3 219
Table 7.5: Nitrogen permeances of membranes (8c) and (N6) after heating at 600°C 223
Table 7.6: Description of different membrane stages (d l-d3) 224
Table 7.7: Hydrogen permeances at 410°C 231
Table 7.8: Hydrogen and nitrogen permeances at 410°C after oxidation 232
CHAPTER 8:
Table 8.1: Different liquid hourly space volumes (LHSV) used for
dehydrogenation reactions in a membrane reactor 237
Table 8.2: Surface reaction rates at different temperatures 251
Table 8.3: Ethanol reaction rate parameters 261
Table 8.4: Permeance data for N8x at 250, 275 and 300°C 261
Table 8.5: Parameters for solving ethanol dehydrogenation model 261
Table 8.6: Reaction rate coefficients for 2-butanol dehydrogenation 265
Table 8.7: Permeance data for N8x at 190, 215 and 240°C 265
Table 8.8: Parameters for solving 2-butanol dehydrogenation model 265
Table 8.9: Parameters for ethanol dehydrogenation at 275°C 266
CHAPTER 9:
Table 9.1: Model differences for ethanol conversion at 250 DC 281
Table 9.2: Improvements in total Xet for the membrane reactor at 250°C 282
Table 9.3: Model differences for ethanol conversion at 275 °C 285
Table 9.4: Improvements in total Xet for the membrane reactor at 275°C 285
Table 9.5: Model differences for ethanol conversion at 300 °C 287
Table 9.6: Improvements in total Xet for the membrane reactor at 300 DC 288
Table 9.7: Improvements in total X2But for the membrane reactor at 190°C 293
Table 9.8: Model differences for 2-butanol conversion at 190 DC 293
xxx
Table 9.9: Improvements in total X28ut for the membrane reactor at 215°C 296
Table 9.10: Model differences for 2-butanol conversion at 215°C 296
Table 9.11: Improvements in total X 2But for the membrane reactor at 240°C 299
Table 9.12: Model differences for 2-butanol conversion at 240°C 299
XXXI
ACKNOWLEDGEMENTS
And most of all, I want to thank the LORD for the gift of life.
XXXll
1. INTRODUCTION
Traditional dehydrogenation reactions have become less favourable due to high energy
costs and have been replaced by alkene oxidation processes to yield the same products.
An example is the oxidation of ethylene (Wacker process) to produce acetaldehyde. By
exploiting advantages offered by membranes, dehydrogenation in a catalytic membrane
reactor becomes a very competitive alternative to alkene oxidation. This process is more
selective and provides high purity separated hydrogen as a by-product.
1
• To model the membrane separation process with a sweep gas and a pressure
differential,
• To compare the performance of a membrane reactor consisting of the optimised
catalyst and optimised membrane with a plug flow reactor, and
• To model the membrane reactor.
A flow diagram for the project is shown in Figure 1.1. The project can be divided into
three sections: firstly, membrane development and modelling; secondly, some catalyst
development and modelling and thirdly, a final section in which catalysis and separation
are combined in the membrane reactor. For Pd plating (block 1), the composition of the
plating solution, the plating conditions and the solution feeding mechanism to the inside
of the membrane tube were investigated. The thickness of the Pd film was minimised to
obtain films with very high hydrogen permeances, while retaining good hydrogen to
nitrogen selectivities. Films were tested from 330°C to 450 °C (block 2) under positive
pressure and by using a sweep gas. Palladium thin films were used as supports for
depositing silver (block 3). Pd-Ag films were heat treated in different gas environments
and at different temperatures. The stability of the films was determined as a function of
time.
Catalysts were prepared (blocks 7 and 8) using the impregnation technique. The effects
of support type, copper loading, added stabilisers and catalyst particle size were studied
for the dehydrogenation of ethanol and 2-butanol. For each catalyst, experiments were
performed over a wide range of operating conditions and both the conversion and
selectivity towards the desired product were optimised. The optimised catalyst was then
2
used for determining the reaction kinetics of the dehydrogenation of ethanol and the
dehydrogenation of 2-butanol.
The optimised catalyst and optimised membrane were combined to form a catalytic
membrane reactor. Experiments in the membrane reactor were performed at different
temperatures, feed flow rates and sweep gas to feed molar ratios. The results were
compared to those obtained in a conventional plug flow reactor.
In this dissertation, the following contributions were made to existing work from other
researchers:
Catalyst development: The contribution made in this area was not large. The
deactivation of a copper-based catalyst during ethanol dehydrogenation was studied in
depth and a more accurate deactivation mechanism was developed. Reaction rate
mechanisms and kinetic parameters were determined for ethanol and 2-butanol
dehydrogenation with the optimised catalyst. Kinetic data was used for modelling
purposes.
Membrane development: Significant advances were made in this field. Much thinner,
yet very selective, Pd films were prepared on the inside of alumina membrane tubes with
a modified electroless plating technique. The Pd film tickness in this study was less than
half of the lowest thickness previously reported (see Appendix E) for the same
membrane configuration. Thin Pd-Ag films were successfully prepared on the inside of
membrane tubes with thickness less than 2.2 microns. The stability and alloying of Pd-
Ag membranes were investigated thoroughly, something that has not been done
previously. A method was developed for improving the hydrogen permeance through Pd-
Ag films.
3
dehydrogenation reaction at higher ethanol feed flow rates, yielding very good results. A
thorough investigation was conducted into 2-butanol dehydrogenation in a membrane
reactor. This, to our knowledge, has not been investigated previously. A full sensitivity
analysis on the membrane reactor model gave more insight into the effects of different
parameters on the dehydrogenation process.
4
Developing 1 Characterising Pd plated 2
Pd electroless
Plating process
membranes
330-450°C ~ 6
Membrane separation
modelling with positive
3 4 AP and a sweep gas
Developing Alloying and characterising
Ag electroless ~
Pd/Ag plated membranes V~ 15
Plating process 250-450°C 14
Performing catalytic Developing and
~ 5 membrane reactor testing cataly-
Testing stability of experiments with
~
. tic membrane
PdAg membranes optimised catalyst reactor model
and membrane
.~
j
16 Performing a
Developing suitable 7 Determine kinetic 11 sensitivity
catalyst for the dehy- / parameters for etha- analysis on
drogenation of ethanol noldehydrogenation the model
9 ...,.
~
Testing stability of Modelling of 13
optimised catalysts pure kinetic
~
data
8 , 10 ..------'
Developing suitable Further catalyst Determine kinetic 12
catalyst for the dehy- deactivation parameters for 2-buta
drogenation of 2-butanol studies noldehydrogenation
•••
o· 0 • .0. •
.0
-
o •
o
0
••
- ..
.0
0 •
0
0 ••
.-0
••
00.0. ••
Figure 2.1: Separation by a semi-permeable membrane (right)
Membrane processes are classified according to the separation method and the size of the
separated species. Basic processes are micro filtration, ultrafiltration, reverse osmosis,
dialysis, electrodialysis and gas separation.
6
membrane consists of two or more layers, prepared in consecutive steps, it is called a
composite membrane. For composite membranes, the initial layer usually provides
mechanical strength and acts as a support on which further layers are deposited on. The
second layer and subsequent layers, determine the membrane's separation properties.
7
(RbAg415), simple or complex oxides (~-aluminas) and oxide solid solutions (ZrOz-Y20],
Zr02-CaO, Th0 2-Y 20]).
Silica deposited on porous Vycor glass gives high hydrogen separation factors.
Hydrogen passes through defects in the silica network. Gavalas et al. (1989) obtained
Knudsen separation values for hydrogen and nitrogen at room temperature, but the value
increased to over 2000 at 450 DC. The main problem with silica deposited membranes is
that they have a very poor stability.
8
niobium (150 Ilm thickness) and tantalum (75 Ilm thickness) tubes coated with palladium
by electro less plating. The membranes were very stable over time.
Some general membrane applications will be mentioned later (see Section 2.7).
2.2.2.1. Porous glass
Macroporous Vycor glass membranes became available in the 1940s. They are made by
acid leaching one of the phases in the glass. Currently these membranes can be prepared
with pores as small as 4 nm. Kameyama et al. (1981) claimed to produce porous Vycor
glass membranes with 86% of their pore diameters within 1 nm of the 4.5 nm average
pore diameter. The brittleness and loss of microstructure upon heating for long periods at
elevated temperatures (> 300°C), limit their application.
10
2.2.2.4. Porous ceramic and composite membranes
Ceramics have several properties that make them the superior choice for inorganic
membranes. Ab03 remains stable up to 800°C without degradation of the pore structure,
(Kameyama et a!., 1983), it is resistant to corrosive environments, it is mechanically
stable and can withstand pressure drops of up to 1.5 MPa. Metals and oxides can easily
be dispersed on the membrane surface and into the pores to add catalytic properties. The
acidity of the support must be taken into account and modified if it catalyses undesirable
reactions.
Ceramics are mainly used as composite membranes, where several layers with decreasing
pore sizes are deposited on one another. The final or permselective layer is typically a
few microns thick and allows for high fluxes. A common example is one or more a-
alumina support layers with a final y-alumina separation layer, yielding a membrane with
4 to 5 nm pores. The top layer determines the characteristics (permeance and selectivity)
and the pore size of the membrane. Top layers that have been deposited and studied
include y-alumina, zirconia, titania, oxide mixtures, zeolites, silica, metals and metal
alloys. Each of these top layers will result in different pore sizes, with the aim being to
make the membrane very selective (very small pore sizes in the Angstrom range) and
allowing for a high flux to pass through the membrane (very thin selective layers in the
nanometer range).
II
Crystals can be grown parallel or perpendicular to the support. Ishikawa et al. (1989)
deposited zeolite coatings on porous Vycor glass and obtained water to butanol
separation factors of 2700. Suzuki (1987) prepared A, X, Y, L, FU-I, ZSM5 and
silicalite zeolites on porous Vycor glass, stainless steel, nickel and alumina. Jia et al.
(1993) prepared silicalite zeolite membranes in situ on a porous ceramic support, while
Sano et al. (1994) prepared a similar zeolite on stainless steel.
In recent years, special attention has been given to porous stainless steel membranes.
Their larger pore size (0.2-0.5 J.Ul1), compared to ceramic membranes, and the wider pore
size distribution are, however, some drawbacks for depositing thin films on them. A
further problem is the diffusion of palladium or its alloy into the stainless steel and a
gradual decline in hydrogen flux upon use. Nam et al. (1999) modified a porous stainless
steel support with a nickel deposit to reduce pore size. Jemaa et al. (1996) used a shot
peening technique to reduce the pore size of stainless steel membranes. During shot
peening, iron particles of less than 125 ~m were fired onto the support. Jemaa et al.
(1996) obtained favourable results and observed a reduction in pore size.
To reduce inter metal diffusion, Gryaznov et al. (1993) introduced a diffusion barrier
between the stainless steel and the palladium layer. It was mainly iron that diffused into
the palladium. They achieved a constant hydrogen permeability for 1000h at 800°C with
an intermediate layer of tungsten (0.8 J.Ul1), tantalum oxide (0.1 J.Ul1), magnesia (0.5 J.Ul1)
and zirconia (1.0 ~m), each applied individually. Shu et al. (1996a) applied a 0.1 J.Ul1
titanium nitride layer onto porous stainless steel. Their Pd-AglTiN/stainless steel
membrane was thermally stable up to 700°C.
12
Edlund and McCarthy (1995b) observed a very rapid hydrogen flux decline in their
palladium-vanadium composite metal membranes at 700°C. Vanadium diffused into
palladium and the process was accelerated in a hydrogen atmosphere. They introduced a
250 !-1m thick porous alumina diffusion barrier between palladium and vanadium.
Thereafter, the hydrogen flux remained constant at 700 °C for the duration of the testing
time (76h).
When composite membranes are prepared, additional thin film formation techniques are
required to modify the membrane support. Thin film formation techniques will be
discussed in greater detail later.
13
Membranes can also be characterised by the transport mechanism through the membrane
pores. In the next section, separation mechanisms, which are determined by pore size and
pore structure, will be discussed.
In the literature the following terms are used for flow through a membrane.
2
• Permeability, in mol.m1(m .s.Pa),
2
• Permeance, in mol/(m .s.Pa),
• Flux, permeation flux or permeation rate, in mol/(m2.s), and
• Flow rate, in mol/so
The selectivity of two components is the ratio between their permeation rate. The
separation factor (a) is a similar parameter and defined as:
(2.1)
with F the flow rate, x the high pressure side, y the low pressure side and i and j are the
components being separated. When the pressure drop across the membrane is small, back
diffusion will take place and it will reduce the separation factor. The separation factor
further depends on pore size distribution, temperature, and interaction between the gases
being separated and the membrane surfaces.
14
r-r-" ~"'~ ~~ ""'Jr"''''' 1'~~'1
" 1 o
'.
...
~ ____
.• . . . .
'" , _ '" ~"~~ ~"-d
o
.0
.~.7 '.~.T. 0
o 0
o
o 0
o
o
o
o o
o
with G r the geometric factor accounting for porosity and tortuosity, t.P the pressure
difference across the membrane, M the molecular weight and Sc the Sievert's constant.
The separation factor for an equimolar gas mixture diffusing by Knudsen diffusion is the
square root ofthe ratio of the molar masses:
(2.3)
15
Separation by Knudsen diffusion is limited in membrane reactors, since a lot of the feed
is lost through the membrane's pores, which reduces the product yield. The best
separation is obtained for light components like hydrogen.
16
200°C. In another study by Suzuki (1987), an equimolar ratio of methane, ethane and
propane at 15 bar yielded a mixture of73.5% methane, 26% ethane and 0.5% propane as
the permeate in a porous stainless steel membrane with a zeolite layer.
YoO
11
2e"+O ..
,--- 01"
adsorption
1
The permeation flux (1) can be expressed using Fick's law (Buxbaum and Kinney, 1996):
17
J = D (C;,] - C;,,)
(2.4)
I
(2.5)
The hydrogen surface concentration (C) is the product of the Sievert's constant (Sc) and
the hydrogen pressure ( PH, ):
(2.6)
When Sievert's law applies, n = Yo. The conditions for Sievert's law have been discussed
by Shu et al. (1991) and Ward and Dao (1999). In general, as films get thicker (above
10 flm) they approach Sievert's law and n = Yo. Diffusion becomes the rate limiting step
in hydrpgen permeation. For very thin films, in the order of a few microns, the value of n
approaches one. Hydrogen chemisorption on the palladium surface becomes the rate
limiting step (Nam et aI., 1999; Yan et aI., 1994). Surface poisoning, grain boundaries
and external mass transfer will cause further deviations from Sievert's law. The limiting
transport mechanism is very temperature dependent. Ward and Dao (1999) concluded the
following after an intensive investigation into hydrogen transport:
• Diffusion is likely to be rate limiting above 300°C, even for thin membranes
(approaching 1 flm).
• Desorption is likely to be rate limiting at lower temperatures.
• Adsorption is likely to be rate limiting for low hydrogen partial pressure and high
surface contamination.
• For thin films (much less than 10 flill), external mass transfer becomes important,
especially on the low pressure side.
• The membrane fabrication technique plays a significant role in permeation, which is
probably related to the microstructure.
18
Furthennore, the penneability (Per) expressed in mol.mJ(m 2.Pa.s) is defined as:
The flux equation can now be expressed in tenns of pressure difference and penneability.
Substituting equations (2.5) and (2.6) into (2.4), and then (2.7) into the result gives:
(2.8)
P = Pee (2.9)
m I
The hydrogen flux is very high through palladium and palladium alloys, mainly because
palladium has a high hydrogen solubility. Do and ED values for the different palladium
phases and at different temperatures have been given by Shu et ai. (1991).
Oxygen penneance through silver is similar to that of hydrogen through palladium. The
value of n can be taken as Y, (Gryaznov et ai., 1986a). Competitive adsorption by other
gases in a gas mixture on silver, reduces the oxygen penneability. For nonporous silica
glass, the activation energy for hydrogen penneance is significantly higher than for
palladium. For palladium it is in the order of 20-25 kJ/mol (Shu et aI., 1991) and for
silica about 35 kJ/mol (Gavalas et ai., 1989). The flux of oxygen through solid
electrolytes has been reported to be proportional to p\\ (Itoh, 1990).
19
2.5. PALLADIUM AND PALLADIUM ALLOYS
Palladium and certain palladium alloys have a high hydrogen permeance. The
characteristics and performance of palladium and palladium alloys have been studied for
many years.
"
a.
::;:
.
:I:
a.
'"
Figure 2.7: Equilibrium solubility isotherms ofPdHn for bulk Pd (Shu et aI., 1991)
20
De Ninno et a!. (1997) discussed the stress fields that are created when hydrogen
dissolves in palladium. The results were hardening, embrittlement and distortion of the
film, which led to cracks in the membrane after a few hydrogenation-dehydrogenation
cycles. To avoid these negative effects, the palladium must be kept in the a phase above
300°C at all time. Lewis (1994) studied the irreversible effects that took place near the
phase transition in a hydrogen atmosphere.
Alternatively, the palladium can be alloyed to suppress a to ~ phase transitions and avoid
distortion. The permeability of the alloy should be comparable to or better than that of
the pure palladium, have high mechanical strength and be resistant to poisoning.
Yan et a!. (1994) and Aoki et a!. (1996), performed temperature cycling tests on
palladium. Thin Pd films « 1 J.lm) prepared by chemical vapour deposition remained
stable for many temperature cycles between 100°C and 300 °C.
21
The morphological changes of the Pd-Ag system upon hydrogen permeation was
described by Shu et al., 1997a. They used SEM images to study changes in a 50 Ilm Pd-
Ag (75:25 wt %) foil before and after hydrogen permeation. Several defects such as
trans granular cracks, intergranular cracks and hydrogen blisters were identified under
different conditions. Lattice strains caused minor cracks, which later developed into
large trans granular cracks.
22
1991) have been investigated as well as ternary alloys ofPd-Ag with Ni, Rh, Pt and Au
(Knapton, 1977).
6
5.5
5 I
... Y
U!
~
4.5 I
E 4 I
~" 3.5 I
.l!. 3 :
c:
.,
0
.'" 2.5
2
.'
I
Ce
.... -.)." .. .' .
.-' .....AD
"
,
..
E
Q.
1.5
;,~:,"." \
,, ,
Au
". .. ....
Cu
,,.,
\
1
" .. ' - (
'.
0.5
0
'
--- -- -- ...... " , ~ ~
~ --
0 10 20 30 40 50 60
% of other component
Figure 2.8: Permeability ofH2 through Pd alloy membranes at 350°C and 2.2 Mpa
The addition of elements to palladium not only suppresses the phase transition, but can
also improve strength. Rodina et aI. (1968) tested the strength implications of the
addition ofNi and Au to palladium and found Ni to be very effective. The improvements
in catalytic properties of palladium with Ni (Smirnov et aI., 1978; Bulenkova et aI.,
1978), Ru (Gryaznov, 1986b; Skakunova et aI., 1988), Rh (parfenova et aI., 1983;
Gryaznov et aI., 1986c), Sn, Sb, Cu and Mo have been studied.
23
1"'"
Prior to any deposition, the membrane support needs to be thoroughly cleaned. Keuler
(1997a) discussed substrate cleaning. Different cleaning methods for porous glass,
porous stainless steel and porous alumina were presented. Different thin film deposition
techniques will be discussed briefly in the next few paragraphs. More detail on preparing
thin films can be found in Keuler (1997a).
24
Several attempts have been made to prepare Pd or Pt composite membranes by the sol gel
method. Hongbin et al. (1995) prepared Pdly-alimuna membranes of which the pore size
varied between 5.5 and 6.5 nm for the different Pd concentrations employed, indicating
that there was little success in reducing the pore size. Vitulli et al. (1995) prepared a
PtlSi0 2 layer by the sol gel process on a SCT multi-layer alumina support. The resultant
membrane showed Knudsen
, diffusion properties with a hydrogen to nitrogen selectivity
. ofless than 3. Zhao (1997) was able to obtain a hydrogen to nitrogen selectivity of23 for
his composite membrane. He used an AlOOH sol to prepare Pdly-alumina membranes.
Evaporation can be performed with resistive heating, but it is far less common than
sputtering. During sputtering, atoms from the target are dislodged through ion
bombardment by an inert gas and deposited on the substrate. Argon is most frequently
used. A magnetron sputtering setup is depicted in Xomeritakis and Lin (1997) and
sputtering conditions for palladium (Xomeritakis and Lin, 1997) and palladium-silver
(Gobina and Hughes, 1996) have been reported. Key parameters during sputtering are
the sputtering time, plasma power, substrate temperature and target to substrate distance.
Jayaraman et al. (1995a, b) investigated the effects of some of those parameters on the
quality and permeance of the sputtered film. Gryaznov et a1. (1993) prepared complex
25
alloys of palladium and one or more ofRu, Co, Pb, Mn and In on porous metal discs. All
examples encountered in literature for palladium sputtering on porous supports used discs
as the substrate.
26
2.6.4. ELECTROPLATING
Metals and alloys can be plated on a conducting substrate that acts as a cathode.
Ceramics and plastics need to be treated before they can be electroplated. The metal
cations are suspended in solution and reduced by an external current passing through the
electrolyte. The cation concentration, bath temperature and current density determine the
deposition rate. Even deposition on large surfaces is difficult due to a variance in current
density and a declining metal ion concentration in the plating bath. Kikuchi (1988)
electroplated Pd-Cu alloys on a porous support. The Pd content was varied between
71 and 94 wt % and the Cu content between 6 and 29 wt %. A method for plating Pd and
its alloys on porous supports was also described by Itoh and Govind (I 989a).
In a more recent study, Nam et al. (1999) used a vacuum electroplating technique to
deposit palladium on a modified porous stainless steel support. A submicron Ni layer
was dispersed on the surface of the porous stainless steel support (0.5 f.tlll pore size)
under low vacuum and then sintered at 800°C for 5h under high vacuum. A thin copper
layer was deposited on the Ni and finally a Pd layer was electroplated on the copper
under vacuum. The resultant film was about 1 f.tlll thick with 78 wt % Pd and 22 wt %
Ni. Hydrogen to nitrogen selectivities varied between 500 and 5000 at temperatures over
350°C.
27
• Dense, non-porous films of even thickness can be prepared on any shape.
• There is good metal to ceramic adhesion.
Palladium ions are reduced and Pd nuclei are deposited on the substrate. Models for
nuclei growth on the substrate have been developed by Cohen et al. (1971, 1973).
Several pretreatment solutions are listed (Osaka and Takematsu, 1980) in literature and
have been tested and evaluated (Horkans, 1983). The two step process deposits more
28
metal than the exchange process does and there is a higher Pd content (Horkans, 1983) in
the deposit. This is favourable for preparing high purity deposits.
Not all metals can be electroless plated, but metals that form good hydrogenation-
dehydrogenation catalysts can be plated. A universal plating mechanism is described by
Van den Meerakker (1981). Ethylene di-amine tetra acetate (EDTA) is most commonly
used as stabiliser with hydrazine or sodium hypophosphite as the reducing agent. Ohno
et al. (1985) lists five reducing agents that can be used for various metal depositions. The
amine complex of palladium is used for electro less plating: (NH3)4PdX, with X = Ch or
N03 .
Rhoda (1959) and Athavale and Totiani (1989) reported the chemical reactions involved
in electroless palladium plating based on hydrazine and sodium hypophosphite as
reducing agents, respectively. Various factors need to be taken into consideration to
ensure a stable plating bath, an even thickness coating and an adequate plating rate.
These are:
• bath temperature,
• solution pH,
• Pd ion concentration,
• reducing agent concentration and stabiliser concentration, and
• solution volume to plated area ratio.
29
Furthermore, different plating characteristics are observed when plating the inside and the
outside of porous tubes. Pearlstein and Weightman (1969) studied some of the plating
variables. Keuler et a!. (1997b) investigated the interaction between various plating
variables and their effect on solution stability.
Yeung et a!. (1995a,b) studied the application of osmotic pressure during electroless
plating. They used porous Vycor membranes as well as aly-alumina membranes as
supports. They found that modifying the plating solution with between 1 and 2 ml
formaldehyde per litre made the films denser, with smaller grain sizes and higher scratch
resistance. A higher formaldehyde concentration inhibited plating, while a lower one
showed no improvement. The plating solution was pumped through the membrane tube
with air, distilled water, sucrose solution or CaCh solution on the other side of the
membrane. By having plating solution on the inside of the tube and either sucrose
solution or CaCh solution on the other side, an osmotic pressure was created from the
inside to the outside. The pressure strength was dependent on the solute concentration.
Yeung and co-workers found that the osmotic pressure made the Pd coatings more dense,
nonporous, thinner and with a smoother surface morphology.
Li et a!. (1997, 1999) used a similar approach to repair defects in their electro less plated
Pd films. Porous stainless steel (0.1 ).tID pore size) and a-alumina (0.16 ).tm pore size)
membranes were used as supports. An initial Pd coating was applied and then one or
more coatings were added under osmotic pressure with NaCI as solute. This resulted in
film densification and defect minimisation. Table 2.1 (Li et a!., 1999) shows how the
selectivity of the Pd film improved after it was repaired.
30
Zhao et al. (1998) used a different pretreatment process to the traditional Pd/Sn activation
and sensitisation process. The porous alumina substrate was activated by a Pd(II)
modified boehmite sol. The gel-coated substrate was dried, calcinated at 600°C and then
reduced in hydrogen at 500 °C. Electroless plating was performed on those activated
substrates which they claimed had a smoother surface and more uniform distribution than
those prepared by conventional pretreatment. After using very high hydrazine
concentrations, they observed that the electro less Pd coating consisted of much finer
particles and this resulted in a more compact film.
Table 2.1: The effect of repairing of electroless Pd plated coatings (Li et aI., 1999)
Plating Film thickness (Jlm) Total Hz permeance HzIN z ratio
Jlmollmz.s.Pa
Original 7.6 6.24 10.3
I" Repair 9.2 3.24 96.5
2"u Repair 10.3 2.68 970
Paglieri et a1. (1999) investigated the effects of pretreatment on hydrogen permeance for
coatings prepared by electroless plating. They tested tin sensitisers of different
concentration as well as a new approach based of palladium acetate without any tin.
They found that high tin concentrations deposited in the pretreatment step led to poor
membrane stability at temperatures exceeding 500°C. Hydrogen to nitrogen selectivity
declined quite rapidly with time. One possible explanation is that tin, with its low
melting point (232°C), could be enhancing metal diffusion along the grain boundaries.
This creates wider channels for diffusion of gases. A pretreatment method based on
palladium acetate gave a better high temperature stability. Pretreatment consisted of
dipping the membrane in a Pd acetate, chloroform solution, drying the membrane by
heating it up to 500°C and reducing it in hydrogen at 500 °C before cooling it down to
room temperature.
31
2.6.5.4. Electro less palladium-silver coatings
Palladium-silver coatings can either be made by consecutive palladium and silver coating
steps or by a single co-deposition technique. Some attempts have been made on Pd-Ag
co-deposition (Shu et aI., 1993; Yeung and Varma, 1995b). The co-deposition process is
very complex and the success has been limited. Silver is much less stable than palladium
in solution, thus very dilute solutions have to be used. Furthermore, silver passivates
palladium deposition, but not vice versa. During the plating process, more and more
silver will be deposited with the net result that the final coating composition varies
significantly from the initial plating solution composition (Pd to Ag ratio). Shu et al.
(1993) investigated the electrochemistry of palladium-silver co-deposition, but many
questions remained unanswered.
An easier approach is to deposit palladium and silver separately. Keuler (1997a) has
given the composition of a typical silver plating bath. The textural differences after
changing the deposition order (pd on Ag or Ag on Pd) were discussed by Keuler et al.
(1999a, b). Results indicated that better metal to ceramic adhesion could be obtained by
depositing silver on palladium rather that palladium on silver. Concentration profiles
across the thickness of the Pd-Ag films, after initial alloying attempts, were also
discussed.
Cheng and Yeung (1999) attempted to model palladium, silver and palladium-silver
electroless plating on porous Vycor tubes (5 nm pores). Separate palladium and silver
plating was well modelled by their equations. Only the metal ion concentration and
hydrazine concentration were, however, treated as variables. Ammonia and EDTA
concentrations and the temperature were fixed. They were unable to properly describe
palladium-silver co-deposition. In general, higher temperatures, lower ammonia
concentrations and higher hydrazine concentrations favoured palladium deposition from a
palladium-silver plating bath. These conditions may result in poor bath stability. Silver
inhibited palladium plating and influenced the film microstructure.
32
2.6.6. PALLADIUM MEMBRANE TEMPERATURE STABILITY
There has been some work done on the long term stability of metal composite membranes
(Buxbaum and Kinney, 1996), where palladium is coated onto other refractory metals.
Very little has been published on the long term stability of palladium or palladium alloys
deposited on porous supports and on alloying procedures.
Paglieri et aI. (1999) recently studied the high temperature stability ofPd composite films
prepared by electro less plating. Plating was performed on the inside of a 200 nm
a-alumina support. At temperatures of 550°C and above, the membranes failed after a
few days and the separation factors dropped to the Knudsen level. Removing tin from the
pretreatment procedure in electroless plating reduced the problem of membrane failure,
but substantial selectivity decline still occurred above 550°C. At 450 °C and 500 DC, the
membranes remained fairly stable for a number of weeks, and the time of stability
depended on the Pd film thickness. It was found that the amount of time to fail was
proportional to the Pd film thickness and that the same failing mechanism prevailed for
ail thicknesses. Possible reasons for failing were:
• Impurities might be trapped at the Pd-alumina interface during pretreatment and
plating, which later result in pore formation.
• Differences in thermal expansion ofPd and alumina can cause cracking.
• Residual porosity in the Pd film can transform into pores.
33
Uemiya et aI., (199Ia) coated the outside of a 200 run porous a-alumina membrane with
palladium and silver. They heat treated it at 900°C in argon for 12h and observed the
formation of a Pd-Ag alloy. Hydrogen permeance was 2.7 times higher after heat
treatment than before heat treatment. They reported no tearing or loosening of the alloy
film from the support. Kikuchi (1995) used the same process for heat treatin& their Pd-
Ag coating on the outside of a porous alumina tube. In both instances the heating process
was only vaguely described; no heating and cooling rates were supplied and no long term
stability test results were reported.
No data could be found on alloying processes for coatings on the inside of porous
ceramic tubes.
34
CO adsorption and loss of hydrogen flux. Jorgensen et al. (1997) reported that CO
adsorption could be used favourably during Pd membrane start-up and shutdown. By
covering the Pd surface with CO at low temperatures (during Pd membrane start-up and
shutdown), the solution of hydrogen in Pd was substantially reduced and the risk of
hydrogen embrittlement restricted.
Jung et al. (2000) studied hydrogen permeance through palladium in the presence of
steam, methane, propane and propylene. Propane and methane had a negligible effect on
the hydrogen flux through the palladium film. Propylene caused severe flux decline,
which dropped further with time. A carbonaceous layer was formed on the Pd due to the
dehydrogenation of propylene. Steam had both a positive and a negative effect. Steam
adsorbed strongly on palladium to decrease the available surface for hydrogen adsorption
and thus the hydrogen flux through the film. On the other hand, steam volatilised carbon
species on the palladium surface to reduce coking and improve the hydrogen flux. The
findings by Li et al. (2000) on palladium deactivation by steam and CO were in line with
those of other studies. Steam adsorbed more strongly on palladium than CO did and
caused greater reduction in hydrogen permeance than CO did.
Inorganic membrane technology has been fully commercialised and is used in many
different industries. These include food processing, processing of beverages, drinking
water, biotechnology, pharmaceuticals, waste oil treatment and petrochemical processing.
Hsieh (1989) and Hsieh (1996) discussed these and many other specialised applications.
Pd alloy metal membranes are commercially used as hydrogen purifiers (Hsiung et aI.,
1999). The presen~ study will concentrate on a class of modified inorganic membrane
35
called catalytic membranes, and more specifically on palladium-based catalytic
membranes.
The catalyst can either be deposited into the membrane pores (Cannon et aI., 1992) or the
membrane can be packed (Tsotsis et aI., 1992; Yeung et aI., 1994) with a catalyst. A
third alternative is to deposit the catalyst only on the inner or outer membrane· surface
next to, or as part of, the separation layer. In the latter case the catalytic surface area is
very small and not effective unless the catalyst is on the inside of a membrane with a very
small inner diameter (hollow fibre). Membrane tubes need to be packed to provide
sufficient catalyst surface area. There are two flow possibilities. One or more reactants
can enter the membrane reactor on the same side (either both shell side or both tube side)
and one or more of the products are separated by the membrane. The second alternative
is to feed reactants into both the shell side and the tube side of the reactor. One of the
reactants moves through the membrane, which acts as a distributor, to react with the other
one(s) on the opposite side of the membrane. This is often used in hydrogenation and
oxidation reactions and in such a case the catalyst is usually in the membrane pores.
To create a driving force for the components being separated by the membrane, either a
sweep gas or a pressure difference is used. The sweep gas enters the shell and tube
reactor on the opposite side of the membrane than the reactant( s), and the sweep gas can
be inert or active. In the case of an active sweep gas, the sweep gas will react with the
36
component permeating through the membrane, for example a oxygen sweep (Itoh and
Govind, I 989b) in dehydrogenation reactions. Furthermore, the sweep gas can be co- or
counter current. The reactor can be adiabatic or isothermal. The effects of different flow
patterns and reactor configurations on reaction conversions have been well documented
(Shu et aI., 1991; Itoh, 1995a; Ross and Xue, 1995).
37
feeding control. The latter is mainly used for oxidation and hydrogenation reactions,
while the first application is for dehydrogenation reactions.
Gryaznov et al. (1986a) studied the oxidation of alcohols with oxygen fed through a
100 ~m silver film. Zaspalis et al. (1989) used microporous membranes for feeding
oxygen and observed an improved catalyst life in some cases. Many more oxygen
reactions are given by Noble and Stem (1995).
38
,.,,
Some niche applications of palladium membranes that have been investigated include: the
production of hydrogen from methanol for fuel cell powered vehicles (Jorgensen et aI.,
1997) and for low CO 2 power generation (Weyton et aI., 1997).
39
,l'i
• Methods have been developed for sealing ceramic tubes (10 mm outer diameter) at
high temperatures. Sealing capillaries or hollow fibres at temperatures above 400°C
will pose problems and replacing broken membranes will be time consuming.
• High temperature operation of Pd composite, ceramic membranes might reqUire
special reactor construction materials. The difference in thermal expansion of the
ceramic membranes and, for instance, a steel reactor will cause damage to the
palladium film deposited on the ceramic supports.
• Films are sensitive to start-up and shutdown procedures. It should be performed
slowly and proper purging procedures must be followed.
Besides these mechanical problems, there are numerous other membrane structural and
related problems that still need to be addressed.
• The membrane cost is very high and needs to be reduced by making the supports
cheaper, as well as producing thin very selective Pd alloy films.
• Membrane poisoning and competitive adsorption on palladium sites reduce hydrogen
permeance.
• Pd alloying and the long term performance of palladium alloy membranes, need to be
studied further.
• Migration of elements in the palladium alloy matrix takes place, resulting in loss of
uniformity.
• Membrane regeneration procedures need to be improved.
40
derivatives. Dehydrogenation not only exploits the cheap and abundant natural gas
liquids, but it also offers improved selectivity towards the desired products.
Unwanted side reactions produce heavy aromatics and coke during dehydrogenation.
These compounds are irreversibly adsorbed on the catalyst and can cause very significant
catalyst deactivation, even though they make up only a fraction of the reaction products
(1 % or less). Catalysts are frequently exposed to an air atmosphere to bum off coke.
41
The higher thermal stability and lower acidity of Zr02 compared to alumina has
resulted in further investigation into this possibility (Ertl et a!., 1997).
LHSV is the liquid hourly space volume, i.e. the time it takes to process one reactor
volume of feed liquid
42
Table 2.3: Process efficiency for propane and iso-butane dehydrogenation
Process Feed Conversion (%) Selectivity (mol%)
FBD Propane 40 89
Iso-butane 50 91
Oleflex Propane 25 89-91
Iso-butane 35 91-93
Catofin Propane 48-65 82-87
Iso-butane 60-65 93
Star Propane 30-40 80-90
Iso-butane 40-55 92-98
Linde Propane 30 90
The catalyst recycle rate is 5-15 kg catalyst per kg iso-butane. The process is quite safe,
since the catalyst regeneration zone (oxidation) and reaction zone are separated. The
absence of fired heaters and the presence of highly effective dust filters make the process
environmentally friendly. The well controlled temperature profile allows for the highest
temperature and maximum conversion at the reactor outlet.
43
2.9.3.2. UOP OIeflex process
This process uses separate units for reaction and regeneration to dehydrogenate mainly
C3 and C4 paraffins. The reaction is performed in moving bed reactors (stacked and
adiabatic) with external heating between stages (Pujado and Vora, 1990). Hydrogen is
used as a diluent. The catalyst properties are:
• spherical pellets,
• a y-alurnina support with platinum (0.1-1 wt %), tin (0.1-4.0 wt %) and alkali metals
(0.1-4.0 wt %), with a surface area of between 50 and 120 m2/g, and
• the operational lifespan for the catalyst is between 1 and 3 years.
Tne hydrogen to hydrocarbon ratio is between I and 10. The dehydrogenation process is
continuous and the catalyst retains its activity for a long time period, giving a constant
yield. The positive reaction pressure improves safety by reducing air leakage and allows
for smaller equipment to be used. The separated reaction and regeneration units also
ensure safe operation.
Due to the low operating pressure, this process gives a high selectivity and a high yield
per pass of the required dehydrogenated product. Heat is supplied by the exothermic
regeneration reactions as well as by external heating. The catalyst is not only thermally
very stable, but has a high resistance to breaking and good tolerance to poisons.
44
2.9.3.4. STAR process (Phillips steam active reforming)
This process is suitable for both the dehydrocyclization of C6 and C7 alkanes and the
dehydrogenation of shorter alkanes. A multitude of fixed beds operate in fired furnaces
with steam added as a diluent. The catalyst properties can be summarised as:
• a support ofzinc or magnesium aluminate with a calcium aluminate binder,
• platinum (0.01-5 wt %) as active component with tin (0.1-5 wt %) as promoter (alkali
metals are optional), and
• the catalyst is very stable in a steam atmosphere and its life is between 1 and 2 years
(Dunn et a!., 1992).
The steam to hydrocarbon ratio is 4 to 5. Steam reduces the hydrogen and hydrocarbon
partial pressures, which shifts the reaction equilibrium towards the product side.
Furthermore, it supplies heat to the reaction zone and reduces coking through the steam
reforming reaction. Coke formation is also limited by the use of fired reactors and the
absence of high temperature pre-heating sections. The positive pressure reduces leakage
risks.
45
2.9.4.1. Coupled dehydrogenation and hydrogen oxidation
The oxidation of hydrogen is an exothermic reaction, which can be used to supply heat to
the endothermic dehydrogenation reaction. This process can take place in two separate
steps or in a single step. The latter is referred to as oxidative dehydrogenation (see
Section 2.9.4.2.). Hydrogen must first be removed in the dehydrogenation step and then
oxidised in the oxidation step. The catalyst must be selective towards hydrogen oxidation
without oxidising the hydrocarbon. The catalyst must also be very stable in steam and air
or molecular oxygen.
Single multi-functional (Imai and Schmidt, 1989) and dual catalysts (Imai and Schmidt,
1989; Imai, 1983) have been proposed, but without commercial application.
46
,'i
is the number one choice for desulphurising fuels and as environmental considerations
increase, more hydrogen is required (Armor, 1998).
Methane steam reforming is one of the most widely applied commercial processes for the
production of hydrogen from synthesis gas. The reactions (2.10 and 2.11) are very
endothermic and require operating temperatures up to 850°C.
The reaction requires a nickel supported catalyst operating at 800-850 °C under pressures
ranging from 1.6 to 4.1 MPa and with steam to methane ratios of between 2 and 4 (Shu et
aI., 1995). Under these conditions conversion is around 78%. Methane steam reforming
kinetics and carbon deposition reactions for methane steam reforming are discussed
extensively by Shu et al. (1994).
Oertel et al. (1987) obtained 90% methane conversion below 850°C in a palladium
membrane reactor (Pd thickness was 50 Iffil). Shu et al. (1995) indicated, through
mathematical modelling, that hydrogen removal through the membrane could best be
exploited at temperatures between 500 and 600°C. Using porous stainless steel
supported Pd and Pd-Ag membranes, packed with Ni/Alz0 3 catalyst, significant
improvements with the membrane reactor were obtained by Shu's group. Methane
conversion was 1.4 times higher in the membrane reactor at 500°C, 136 kPa and a steam
to methane ratio of 3. Uemiya and co-workers (1991c) achieved 80% conversion at
500°C and 100 kPa with a Ni/Alz03 catalyst and a steam to methane ratio of3 compared
to an equilibrium value of 42%. They used a membrane consisting of a 80 Ilm Pd77 -Ag23
alloy, coated on a porous glass tube. In similar work a team from Haldor Topsoe
(Jorgensen et aI., 1995; Armor IN, 1998) achieved 51% conversion (equilibrium = 21%)
at 500°C and 6 atm pressure with a 100 Iffil Pd-Ag alloy tube and commercial NilMgO
catalyst.
47
The group from Haldor Topsoe performed an economic analysis of this process and
indicated that the membrane process was not yet competitive with the conventional
process. This was due to a low hydrogen flux through the membrane, poor membrane
selectivity, membrane sealing problems and the absence of data regarding carbon
formation on the membrane during reaction.
48
have received attention are Fe-Cr catalysts (high temperature), Cu-Zn (low temperature)
and the more expensive Pt on ZnO catalysts (Bracht et aI., 1997). The latter one shows
higher sulphur resistance, higher activity and very good selectivity compared to the
"
Basile et al. (1996) reported conversions below equilibrium values with their Pd coated
alumina membranes. The 0.2 J.lIll Pd coating was too thin and resulted in poor membrane
selectivity and low reaction yield. With a 20 ).1m Pd coating on porous glass, Uemiya and
co-workers (1991d) achieved a conversion of 88%, compared to a 78% equilibrium
value. They used a Fe-Cr catalyst at 400°C, 1 bar and a steam to CO ratio of 2: I. The
catalyst unit and membrane separation unit can be separated to simplifY catalyst and
membrane regeneration (Ross and Xue, 1995). The units can then be operated at
different temperatures and membrane sealing is simplified. The additional cost of a
separate unit and the smaller driving force in the reaction are some disadvantages to be
considered.
Van Veen and co-workers (Burggraaf and Cot, 1996) calculated that the optimum H2 to
C02 membrane selectivity required for this process, is about 40. The high pressure, high
temperature membrane sealing and the large membrane surface areas required (1500 m2
for a 300 MWe power plant), remain problematic.
49
<ii:
,;!\:,., "
I~.y'.
illl,
Mi~" = + 137 kJ I mol (2.13)
Platinum (Champagnie et ai., 1992) and palladium (Gobina and Hughes, 1994) supported
on alumina catalysts can be used for ethane dehydrogenation. Champagnie and co-
workers performed ethane dehydrogenation experiments in a multi-layer alumina
membrane (MEMBRALOX) impregnated with Pt. At 550°C, a trans-membrane.
pressure of 1 bar and a sweep gas to feed ratio of 2, they obtained about 19% conversion
compared to the equilibrium value of just under 10%. A good fit was obtained with a
model developed by them. Gobina and co-workers (1994, 1995a, 1995b) used a 6 )lm
Pd77 -Ag23 deposited on a porous Vycour glass membrane, for studying ethane
dehydrogenation. The membrane was housed inside a stainless steel reactor and packed
with a 0.5 wt % Pd on y-alumina catalyst. They studied different parameters including
the type of sweep gas used, sweep gas to feed flow ratio and feed contact time, to draw
up a mathematical model for the process.
50
The latter is favoured from a thermodynamic point of view, but should be suppressed by
using a very selective catalyst. Thermal cracking at high temperature should also be
avoided. Different types of catalysts can be used depending on the desired result. For the
oxidative dehydrogenation of propane to propene in a conventional reactor a V -Mg-O
catalyst gave 65% selectivity and 10% conversion at 540 °C (Chaar et a!., 1987). Cadus
et a!. (1996) studied Mg-Mo-O catalysts for the same reaction, obtaining very good
selectivity, but poor conversion. Saracco et a!. (1996) applied the process of dosing to
the oxidative dehydrogenation of propane. Air was fed in a controlled manner through a
Pt-impregnated tubular alumina membrane to react with propane on the opposite side of
the membrane.
Burggraaf and Cot (1996) performed modeling calculations and did an economic analysis
on the use of membranes for the dehydrogenation of propane. Membrane selectivity
should be significantly greater than Knudsen diffusion selective membranes to prevent
51
reactant and product loss through the membrane pores. High hydrogen diffusion rates
must be obtained to reduce the membrane surface area. The use of Knudsen diffusion
membranes improve the return on investment for both the Catofin and Oleflex processes
by about 3% and I %, respectively. However, the process seems to be uneconomical with
a price difference ofless than $250-300 between a ton of propylene and a ton of propane.
Iso-butene demand is increasing due to changes in the Clean Air Act passed by the US
Congress in 1990 (Udomsak and Anthony, 1996). Iso-butene is one of the reactants for
producing the octane booster, methyl tert-butyl ether (MTBE).
52
the reaction. The best yield obtained was 76% at 500°C towards total butenes (68% iso-
butene) with a reaction conversion of 81%. Increasing the flowrate reduced the
conversion to that of the equilibrium value (32.4%), due to poor hydrogen permeance
through the relatively thick Pd membrane. A hydrogen feed concentration of about 2%
gave better results due to less catalyst deactivation. Matsuda et al. (1993) obtained iso-
butane conversions exceeding the equilibrium value at temperatures between 350°C and
450 °C with a Pd (deposited by electroless plating) on alumina membrane. Membrane
deactivation was reduced by co-feeding iso-butane with a small percentage of hydrogen.
Shu et al. (1997b) improved iso-butane conversion from 12% to 32.7% and iso-butene
yield from 8.8% to 30.8% at 450°C with a membrane reactor. Their membrane consisted
of a 2 11m Pd coating on a multi-layer (a-alumina support and y-alumina toplayer) SCT
membrane.
Zhu et al. (1993) impregnated a 4 nm y-alumina membrane with Cr203 and used it to test
iso-butane dehydrogenation, while Ioannides and Gavalas (1993) used a membrane
consisting of a CVD deposited Si02 layer on porous Vycor glass for the same reaction.
Zeolite-coated membranes, for iso-butane dehydrogenation, have also been studied
(Casanave et aI., 1995).
53
Dehydrogenation has been studied widely and modelled in fixed bed reactors (Acharya
and Hughes, 1990). Chromialalumina is the most common catalyst used for butane
dehydrogenation (Happel et a!., 1966; Hakuli et a!., 1996) and the kinetics of the reaction
have been documented (Noda et a!., 1967). Since catalyst coking is such an important
phenomenon, many researchers have investigated chromialalumina coking during butene
dehydrogenation (Hughes and Koon, 1994; Brito et a!., 1995, 1996). Coking kinetics has
been formulated (Romero et a!., 1996; Pena et a!., 1993). Other types of catalysts that
have also received attention include iron oxides supported on titania (Boot et a!., 1994,
1996), calcium-nickel-phosphate catalysts (Arnold, 1961; Swift et a!., 1976) and platinum
on alumina catalysts (Loc LC et a!., 1993, 1996).
Oxidative dehydrogenation of butenes takes place mainly on ferrites (Xu et a!., 1992;
Yang et a!., 1984) or magnesium ferrites (Gibson and Hightower, 1976; Yang et a!.,
1991). Iron-zinc oxides (Armendariz et a!., 1992) and vanadium-magnesium oxides
(Bhattacharyya et a!., 1992) are just some of the other catalysts that have also been
studied for this reaction.
54
2.9.10. DEHYDROGENATION OF ETHYLBENZENE TO STYRENE
The worldwide demand for styrene reached 18.2 million tons in 1992. This figure is
estimated to grow at 3-5% per annum, to reach 23.9 million tons in 2000 (Burggraaf and
Cot, 1996). Styrene is converted to polystyrene (> 65%) and co-polymerised with
butadiene (13%) for rubber elastomer production. A fraction is used for styrene-
acrylonitrile production (9%) and mixed with unsaturated polyester resins (James and
Castor, 1994). The endothermic reaction is usually performed at temperatures between
550°C and 650 °C, 0.5-1.0 aim pressure and with steam as a diluent. Ertl et aL (1997)
lists many reasons for adding steam. The main by-products of the dehydrogenation
reaction are toluene and benzene, with smaller amounts of ethylene, methane and coking
products.
Potassium-promoted iron oxide catalysts are mostly used for this reaction and have been
studied extensively (Muhler et aI., 1990, 1992; Addiego et aI., 1994). Catalyst activity
decays due to changes in the surface concentration of the potassium promoter (Matsui et
aI., 1989, 1991). The kinetics of the dehydrogenation reaction (Abdalla et aI., 1994a,
1995; Coulter et aI., 1995) and modelling of commercial fixed bed reactors (Abdalla and
Elnashaie, 1995; Elnashaie et aI., 1993) have received significant attention. Oxidative
dehydrogenation allows for lower reaction temperatures (around 450°C), to be used but
catalysts usually suffer from poor selectivity. Magnesium vanadates (Chang et aI., 1995),
Zn-Fe-Cr catalysts (Jebarathinam et aI., 1996) and carbonaceous catalysts (Drago et aI.,
1994) are some that have been investigated for use in oxidative dehydrogenation.
55
membrane. Becker et al. (1993) reported improvements as high as 20%. Wu and Liu
(1992) as well as Abdalla and Elnashaie (1994b) have modelled the membrane reactor for
ethylbenzene dehydrogenation. Gobina et al. (1995c) compared the effects of the
separation factor of different microporous membranes and dense Pd-Ag membranes on
styrene yield. There seems to be an optimum membrane thickness for microporous
membranes, which depends on the porosity of the separation layer. Pd-Ag membranes of
less than 50 IJ.l11 thickness outperform microporous membranes.
56
Table 2.4: Catalyst supports (Carrizosa and Munuera, 1977; Rosynek et a!., 1990)
Basic oxides Acidic oxides
Pure silica Alumina
MgO Silica-alumina (zeolites)
ZnO Ti0 2
Zr02
Alkiline earth oxides
It is not only the acidity of the support that is important. Other factors which are also
very important are:
• surface area,
• abrasion resistance and crushing strength,
• temperature stability and resistance to fouling, and
• toxicity.
Silica and alumina have very high surface areas compared to the other oxides (typically
in the hundreds of m 2/g area). High copper surface areas can be obtained by depositing
copper on these supports. The activity of the catalyst is usually proportional to the
surface area of the active sites and thus a large copper surface area will yield a more
active catalyst.
57
(image (c». The total copper surface area available for reaction declines as no reactant
can enter the blocked pores to react.
+Cu + Cu
) )
58
2.10.2.1. Precipitation
This technique is suitable for preparing copper-supported or unsupported catalysts (Opitz
and Urbanski, 1958). A copper salt (usually copper nitrate) is dissolved in distilled water
in a stirring vessel. The support is added to the vessel while stirring and sodium
carbonate is added as a precipitant. The rate of addition is controlled to keep the solution
at a constant pH, usually pH=8.0. The precipitation temperature and precipitation time
both have an effect on catalyst performance. Increasing the precipitation temperature
from 20 °C to 90°C (Jeon and Chung, 1994) resulted in a catalyst that gave a much
higher cyc1ohexanol conversion at the higher precipitation temperature. From TEM
measurements, Jeon and Chung (1994) observed that copper particles precipitated at
20°C were about 7 nm to 15 nm in size, while those precipitated at 90°C were about
1.7 nm to 7 nm in size. Conversion was a weak function of precipitation time, with about
4 hours precipitation time giving the best conversion.
Unsupported copper catalysts are prepared in a similar way, with other metals (for
example chromium and/or cobalt) added in the correct mass ratio. The precipitate is
filtered, washed with ample amounts of water and then dried overnight at about 100 °C.
Thereafter, it is calcinated at temperatures exceeding 450°C. Kanoun et al. (1991,1993)
described methods for preparing vanadium-copper-zinc and copper-chromium-aluminum
catalysts via precipitation. Ammonia was added to precipitate the hydroxides. Further
ammonia was added to redissolve the hydroxides as amine complexes. Ammonia was
then removed by heat to yield the precipitate with the desired structure.
59
2.10.2.3. Electroless plating
Electroless copper plating has been used on occasion to deposit copper on the support
(Chang and Saleque, 1993, 1994). The support is first pretreated with palladium and tin
chloride solutions before copper plating. Palladium seeds on the support provide
catalytic centres for copper to plate on. Copper plating requires an alkaline solution and
is performed at high temperatures (typically> 70°C). Either formaldehyde or hydrazine .
is used as reducing agent and EDTA as a stabiliser. The concentration of the plating
solution or the volume of the plating solution can be altered to give different copper
loadings on the support. After plating the catalyst is filtered, washed, dried and
calcinated.
2.10.2.4. Impregnation
This is a quick and easy way to prepare copper-supported catalysts with or without
additional elements like chromium, cobalt or rhodium. The support particles are dried at
about 200°C overnight to remove water vapour from the pores. The particles are then
cooled in a desiccator. Either a copper solution or copper mixed with additives in the
desired mass ratio are added to the support particles. The mixture is well stirred while
adding the solution. The paste is then dried at 90°C to 110°C for several hours and later
calcinated at above 400°C in air for at least 5 hours. Iwasa and Takezawa (1991) used
this technique to impregnate Ah03, SiOz, MgO, zrOz and ZnO with copper. Shiau and
Liaw (1992) prepared Ba-copperlSi0 2 catalysts and Mendes and Schmal (1997) prepar.ed
Rh-copperly-alumina catalysts by impregnation for alcohol dehydrogenation.
CuO (s) + H2 --- Cu (s) + H20 (g) LlH~20K = -85 kJ I mol (2.16)
60
The temperature inside the catalyst particle can rise considerably, causing sintering and
reducing the surface area of the copper. On the other hand, poor reduction will occur if
the reduction temperature is not high enough. From literature (Bart and Sneeden, 1987),
it appears as if standard prescribed reduction methods for CU/ZnO catalysts result in poor
reduction. Reduction temperatures in excess of 300°C give a better reduction percentage
fr9m Cu2+ to CuD, but give rise to more sintering of copper clusters. Lower temperatures
cause less sintering, but also poor reduction.
Furthermore, the copper concentration on the support influences the reduction ability. At
low Cu concentrations, the Cu-ion to support bond is very strong. Bart and Sneeden
(1987) reported that for 5-10 wt % CuO on ZnO, the surface content of copper was not
altered by hydrogen reduction for 2h at 250°C. Cu2+ was converted to both Cu 1+ and
CUD. For copper concentrations up to 35 wt % CuO, Cu 1+ was stilI detected.
Tohji et al. (1985) developed a structural model for copper particles in hydrogen during
reduction (Figure 2.10). In the first step, below 127°C, a two dimensional copper layer
develops over the ZnO support. Between 127°C and 277 °C copper metal clusters are
formed reversibly on the support. Above 327°C, large copper clusters are formed
irreversibly due to sintering.
C2CX:XJC)
127°C
~e------~ fmD 400°C ) @
Zno Zno Zno
Mendes and Schmal (1997) reported that there is a strong interaction between alumina
and CuO, for low CuO percentages (0.5 wt % in his study). They could only obtain 57%
CuO reduction even if the reduction temperature was increased up to 500°C. Tu et al.
(1994a) reduced their unsupported Cr/Cu catalysts at 200°C for 6 hours in a
hydrogen/argon mixture. Some sintering occurred during the reduction process. A
61
proper dispersion of Cr203 promoter slowed down sintering during reduction. At high Cr
to Cu ratios (20:40), the reducibility of the catalyst at 200°C was very poor. Kanoun et
al. (1991) reduced their vanadium-copper-zinc catalyst for 3 hours at 300°C. Chung et
al. (1994) reduced their silica-supported copper catalyst for 4 hours at 250°C with
reduction percentages of 65% and higher. Iwasa and Takezawa (1991) reduced their
copper-supported catalyst for 18 hours at 250°C.
2.10.4.1. Sintering
Irreversible sintering of copper catalysts at temperatures exceeding 250°C is a major
cause of catalyst deactivation. Sintering will increase gradually as the temperature is
raised (see Section 2.10.3) and the degree of sintering depends on the metal to support
physical and chemical bond as well as the copper crystallite size. Bart and Sneeden
(1987) reported that copper sintering for CU/ZnO/Ah03 catalysts is less than expected on
the basis of Tamman temperatures, due to strong bonding between Cu and ZnO, which
reduces copper mobility.
Church et al. (1951) found that Cr and Co addition improve long term copper catalyst
stability. After 100 hours of ethanol dehydrogenation, the activity of a Cu-asbestos
catalyst was 60% of the original activity, for a Cu(95%)-Co(5%)-asbestos catalyst the
activity was 70% of the original and for a Cu(93%)-Co(5%)-Cr(2%)-asbestos catalysts
the activity was 83% of the original. The support also proved to have a large effect on
catalyst life. Under identical conditions a Cu-Cr-Co/asbestos catalyst retained 80%
activity and a Cu-Cr-Co-pumice catalyst retained only 32% of initial activity after
reaction at 275°C for 48 hours.
62
i'
I
,
Tu et al. (1994a, 1994b ) detennined copper surface areas of unsupported Cr/Cu catalysts
before and after eight hours of use at 310°C (ethanol dehydrogenation reaction). For
catalysts both before and after use, the Cu surface area passed through a maximum at a Cr
to Cu ratio of 4:40. There was a significant decline in surface area after use. For pure
copper, the decline was from 9 to 6 m2/g copper and for Cr:Cu = 4:40, the decline was
from 19 to 17 m2/g copper. They found the deactivation kinetics to be second order and
concentration independent, which is similar to common sintering kinetics (Fogler, 1992).
Chinchen et al. (1988) concluded that in the absence of catalyst poisoning, sintering is the
major deactivation process for copper-based methanol synthesis catalysts.
63
Copper-zinc-alumina was very resistant to sulphur poisoning, with 2% sulphur in the feed
only causing a 20% decline in catalyst activity. For copper-alumina catalysts, 0.2%
sulphur in the feed caused total deactivation of the catalyst.
64
5. diffusion of the products from inside the pores and close to the external surface into
the bulk fluid.
(2.23)
with R2 = H for primary alcohols and R2 = alkyl or aryl for secondary alcohols. Alcohol
dehydrogenation reactions are endothermic and are usually performed at temperatures
between 250 DC and 450 DC. Reactants are fed in the gas phase with conversion being
favoured by low operating pressures. The principal side reaction is dehydration of the
alcohol to yield the alkene and the di-ether. Oxidative dehydrogenation is an exothermic
reaction and requires strict temperature control to prevent total combustion and the
formation of carbon oxides.
Very few studies have dealt with membrane assisted dehydrogenation of alcohols. Most
membrane reactor experiments have thus far focussed on alkane dehydrogenation.
65
,
i
CH3-0H (g) ..... HCHO + H2 L1H~98 = +84 kJ I mol (2.24)
CH3-0H (g) + O2 ---> HCHO + H20 (g) L1H~98 = -159 kJ Imol (2.25)
Reaction temperatures vary between 647°C and 717 DC, giving a formaldehyde yield of
about 87% (Ertl et a!., 1997) for the older process. For the newer process, reaction
temperatures vary between 350°C and 450 DC, with a formaldehyde yield of about 90%.
Zaspalis et al. (1991) used an a-alumina membrane with a y-alumina toplayer (4-5 nm
pore size) and a commercial ZnO catalyst for methanol dehydrogenation. At 500°C, the
membrane reactor gave a 19% formaldehyde yield compared to the 14.4% yield of the
conventional reactor. The disadvantage of this Knudsen separation membrane IS,
however, the loss of reactant through the pores due to poor separation characteristics.
Many different catalysts have been tested for use in this reaction, the most common being
copper and/or zinc-based. For ZnO the reaction is performed at 300-400 °c with a
selectivity of 90% and a conversion of 98%. Cunningham et al. (1986) studied
unsupported CuO, CU20 and Cu metal, while Szabo et aI. (1975) determined activation
66
energies and Arrhenius parameters for MgO, CaO and srO. Yashima et al. (1974)
studied iso-propanol dehydration and dehydrogenation on alkali exchanged zeolites and
found the dehydration behaviour to be proportional to the acidity of the catalyst. Other
catalysts tested include lead, manganese oxide, tin-iridium complexes (Matsubara et aI.,
1991) and supported platinum and rhodium catalysts. Oxidative dehydrogenation of iso-
propanol uses similar catalysts (Gil et aI., 1996a, 1996b).
The reaction becomes pseudo first order when surface adsorption is weak. The use of
several catalysts has been studied, usually in the temperature range between 200°C and
400 0c. The most common catalysts are copper-supported or unsupported catalysts, as
indicated in Table 2.6.
Other catalysts that have also been studied for cyclohexanol dehydrogenation include
alumina-supported Pt-Co catalysts (Reddy et aI., 1997), carbon-supported cobalt catalysts
67
(Uemichi et aI., 1995), tin oxides (Hino and Arata, 1990) and zinc phosphate complexes
(Aramendia et aI., 1995).
The activity and selectivity of copper-based catalysts depends on several factors. The
more important ones are:
• acidity ofthe support,
• preparation technique,
• copper loading,
• calcination temperature, and
• reduction temperature.
Chang and Saleque (1993) studied three preparation methods viz. electroless plating,
impregnation and deposition precipitation to determine the effects of the preparation
method on cyclohexanol conversion. Each method showed an optimum copper
percentage for maximum cyc1ohexanol conversion. Chang and Saleque (1993) listed.
activation energies and Arrhenius frequency factors for various catalysts. Catalysts
prepared by electro less plating gave the highest BET surface area and the best
conversion. Chang and Saleque (1994) also compared the activities of electro less plated
Cu on a-Ah03, and y-Ah03 supports, while Sivaraj et al. (1990) determined the
relationship between dehydrogenation selectivity and catalyst acidity.
68
2.12.4. DEHYDROGENATION OF C4 ALCOHOLS
The dehydrogenation of butanol yields butyraldehyde, an important intermediate in the
manufacture of solvents, plasticisers, synthetic resins and rubber vulcanisation
accelerators. This process is not performed on an industrial scale, but several researchers
have investigated the reaction. Copper, copper-chromium and zinc-based catalysts have
been studied (Rao et aI., 1969; Purohit and Gandhi, 1975).
Shiau and Liaw (1992) used copper-barium supported on silica to investigate the kinetics
of the oxidative dehydrogenation of n-butanol between 230°C and 290 0c. Sintering
occurred and the addition of oxygen increased catalyst coking. Raizada et al. (1993)
determined reaction rate parameters for the dehydrogenation of n-butanol over a zinc
oxide catalyst at temperatures between 350°C and 450 0c. Copper, zinc or copper-zinc
catalysts were most widely used.
69
,!,
Combining membrane technology with ethanol dehydrogenation results in advantages
that could not be exploited previously. The added value of separated hydrogen as a co-
product and a dehydrogenation processes that operates at a lower temperatures, give more
credibility to ethanol dehydrogenation.
There are two main catalyst classes for ethanol dehydrogenation; zinc or zinc.oxides and
copper-based catalysts. With lnO and Cu catalysts, the ethanol dehydrogenation reaction
proceeds via different mechanisms on the catalyst surface (Chung et ai., 1993). Copper is
more active and reaction temperatures are usually in the region of 250°C to 300 0c.
Higher temperatures cause catalyst deactivation and promoters like cobalt and chromium
are often added to copper to improve catalyst stability. The lower activity of zinc-based
catalysts lead to higher operating temperatures, typically 350°C to 450 °C. Although
zinc-based catalysts are thermally more stable, they have other disadvantages. Higher
temperatures decrease reaction selectivity and cause thermal cracking of reactants and
reaction products.
70
fonnation, while Ah03 supports resulted in C4 and di-ethyl ether fonnation at 220°C.
Furthennore, they found that high copper percentages suppress ethyl acetate fonnation.
Table 2.7: Selectivities for ethanol dehydrogenation (Iwasa and Takezawa, 1991)
Support Active T P Conver- Se1ecti-
Materials (oq (kPa) sion vity
AIz 0 3 30%Cu 220 20.5 50* 54.1
Si0 2 30%Cu 220 20.5 50* 77.9
MgO' 30%Cu 220 20.5 50* 74.2
Zr02 30%Cu 220 20.5 50* 57.3
ZnO 30%Cu 220 20.5 50* 67.3
- Cu 220 20.5 50* 70.6
...
*ConverslOn was fixed at 50% to test selectIvItIes.
Church et al. (1951) found an improvement in both conversion and selectivity when Co
and Cr were added to Cu supported on asbestos (see Table 2.8).
Asbestos Cu 328 - 79 90
Asbestos Cu, 5% Co 337 - 94 89
Asbestos Cu, 5% Co, 2% Cr 330 - 93 83
71
2.12.5.2. Catalyst activity
To test catalyst activity, the reaction is usually performed under differential reactor
conditions, i.e. very low reactant conversion. Under such conditions, the selectivity
towards acetaldehyde production is high and usually not considered.
Tu et a!. (l994b) tested the effects of chromium addition to unsupported copper catalysts
for reaction temperatures ranging from 250 DC to 310 DC. There appeared to be an
optimum Cr to Cu ratio of 4 to 40. That composition gave the highest activity over time
with the smallest deactivation rate. The high activity was related to the highest copper
surface area at that Cr to Cu ratio. Table 2.9 (Tu et a!., 1994b) gives the reaction rate
parameters for the first order dehydrogenation reaction of ethanol over Cr:Cu = 0:40 and
Cr:Cu = 4:40 catalysts. There was a significant improvement in the reaction rate for the
optimum Cr to Cu ratio compared to that for pure copper.
Table 2.9: Reaction rate parameters for ethanol dehydrogenation (Tu et aI., 1994b)
Catalyst T (oq k (dm'gcaf'.h"') A (dm'gcaf' ,h"') E. (cal/mol)
Cr:Cu= 0:40 250 14.39 1.12*10° 12100
280 23.36 " "
310 51.97 " "
The catalyst activity for dehydrogenation is also strongly dependent on the copper
loading. Sivaraj and Kantarao (1988a) investigated the effects of the copper percentage
on y-alumina for 4 to 34 wt % Cu. They found an optimum copper surface area in the
region of 20-25 wt % Cu. The overall reaction rate dropped sharply, to about 16 wt %
copper for temperatures from 250 DC to 300 DC, thereafter it dropped only marginally and
became almost constant above 27 wt % Cu. Below 20 wt % copper, the acidic
72
y-alumina support caused a sharp shift in product distribution at 275°C. Dehydration
products increl).sed sharply and dehydrogenation products decreased sharply when the
copper content dropped below 20 wt %.
Kanoun et ai. (1991, 1993) determined the activities of V-Cu-Zn and Cu-Cr-Al catalysts
for the dehydrogenation of ethanol at 190°C. Changes in total catalysts surface area and
copper surface' area, with an increase in Cu percentage, were discussed. They also
calculated the molecules of acetaldehyde produced per surface copper atom per second
(turnover frequency or TOP) as indicated in Table 2.10. For the Cu-Cr-Al systems the
activation energy varied between 20 and 22 kcallmol, which was much higher than values
ofll to 12 kcallmol obtained by Tu et ai. (1994a).
Table 2.10: Activities and TOPs for ethanol dehydrogenation at 190°C Kanoun et al.
(1991, 1993)
Catalysts A (mol.kg TOF Catalysts A (mol.kg TOF
carl.h- I) x 10
3
cat-I.h- I) x 10 3
Cu 5.2 45 Cu-Cr 4.34 1.8
V-Cu 2.89 2.8 Cu-CrO.9AlO.l 4.32 1.9
Zn-Cu 1.59 6.5 Cu-CrO.7AlO.3 4.16 1.9
V-Zn 0 - Cu-CrO.5AIO.54.34 4.07 1.6
V-CuO.9ZnO.l 3.93 2.4 Cu4.25-CrO.3AIO.7 4.34 1.9
V-CuO.8ZnO.2 3.84 2.1 Cu-CrO.lAlO.9 4.25 1.5
V-CuO.6ZnO.4 3.98 2.1 Cu-Al 4.43 4.9
V-CuO.5ZnO.5 4 2.1
V -CuO.4ZnO.6 3.84 2.2
V-CuO.2ZnO.8 2.84 2.1
V -CuO.lZnO.9 2.43 1.8
Por the catalysts in mentioned inTable 2.10 the following two criteria apply:
• Cu/(Cr+Al) = 1
• V/(Cu+Zn) = 1
73
Cu-Al and Cu-Cr were the best performing binary systems where the ratio of Cu to other
metal is 1 to I. 'Varying the Cr to Al ratio did not have a significant effect on catalyst
activity. For V -Cu-Zn catalysts, an equal amount of copper and zinc combined with
vanadium gave the highest dehydrogenation activity.
(2.30)
The water term was added for when water might be present in the feed, but for modelling
purposes it was excluded. The calculated values for the constants (equations 2.31 to
2.34) between 225°C and 285 °C were:
(2.31)
(2.32)
(2.33)
74
InK =6850_ 7.18 (2.34)
S R oT
Peloso et al. (1979) also studied dehydrogenation kinetics between 225°C and 280 °C
with the following catalyst: 41.2% CuO, 33.4% Cr203, 9.3% Si02, 3.3% Na20 and 12.8%
binder. Of the mechanisms investigated the dual site reaction mechanism was once again
rate controlling (see equation 2.19).
(2.35)
with,
k [mol/kg cat.h.atm]
75
Tu et al. (1994b) found the dehydrogenation reaction to be pseudo first order and of the
form:
(2.40)
Cullis and Newitt (1956) investigated ethanol oxidation between 270°C and 370 °C.
Initially acetaldehyde was formed, but then further oxidation occurred, yielding
methanol, formaldehyde, methane and carbon oxides. Rao et al. (1991) tested a medium
pore, titanium silicate molecular sieve catalyst for the oxidative dehydrogenation of
ethanol. At 300°C the selectivity towards acetaldehyde production was over 90%, but
the conversion was only about 20%. At 400°C the conversion increased to above 90%,
but the selectivity dropped to below 70%. Quaranta et al. (1994) investigated the
possibility of using vanadium complexes for oxidative dehydrogenation.
76
2.12.5.5. Dehydrogenation of ethanol in a membrane reactor
Deng et al. (l995) modified alwnina membranes (500 nm pore size) with a y-alumina
layer containing Pd, Pt, Cu or Ni. The net pore diameter varied between 3 nm and 9 nm.
Hydrogen-argon selectivities for the Pt and Pd-based membranes were higher than the
Knudsen values and for the others lower than the Knudsen selectivities. Ethanol
dehydrogenation was studied in the temperature range from 250°C to 310 cC, employing
a Cu-P/Si02 catalyst. The acetaldehyde yield for the conventional reactor was slightly
below the equilibriwn value, while the values for the alwnina membrane were higher
than the equilibrium value. Cu and Ni-modified alwnina yielded results similar to
alumina membranes. The best results were obtained with the Pd and the Pt-modified
alumina membranes. Acetaldehyde yields were further improved by increasing the space
time and/or the sweep gas flow rate.
Raich and Foley (1998) studied ethanol dehydrogenation in a Pd tube with a wall
thickness of 76 11m. The operating temperature varied between 175°C and 225 °C and
the tube was packed with Cu or Pt on silica catalysts. The best results were obtained with
a copper on silica catalyst prepared by ion exchange followed by copper on silica
prepared by impregnation. The latter catalyst gave higher selectivity, but lower activity
and lower overall yield. Ethyl acetate was the main by-product at lower temperatures.
They compared a palladium reactor packed with copper on silica catalyst (prepared by
ion exchange) with a conventional reactor and obtained the following results: conversion
increased from 60% to 90% and selectivity from 35% to 70%.
77
temperatures, while for secondary and tertiary alcohols, olefin formation dominates also
at the lower temperatures.
De Boer and Visseren (1971) observed that the rate constant for ethanol dehydration is
proportional to the amount of aluminium (acidic sites) on the catalyst's surface. For
ethanol the following was found:
Several papers describe the dehydration of ethanol over zeolitic and non-zeoli tic
molecular sieves (De las Pozas et aI., 1993; Teo and Ti, 1990). For n-butanol
dehydration, the alumina percentage in silica-alumina catalysts has a significant effect on
the dehydration products, with a higher alumina percentage favouring di-butyl ether
formation (Berteau et aI., 1991). C4 dehydratio~ has been studied over various types of
zeolites (Makarova et aI., 1994; Williams et aI., 1991) and over y-aIumina (Lu et aI.,
1995).
2.14. SUMMARY
This chapter has described the various types of inorganic membranes that are currently
available. The main advantages and disadvantages of using inorganic membranes as well
as. the areas of application have been mentioned. Inorganic membranes were separated
into three classes i.e. dense, porous and composite membranes. The different separation
mechanisms through porous membranes were discussed. Palladium and palladium alloy
membranes were discussed in more detail, with specific attention being given to the
palladium-hydrogen system and the effects of alloying on palladium stability in
hydrogen. Different methods by which to prepare composite palladium membranes were
described. The more important methods were listed and compared. Electroless plating
was discussed in detail, covering substrate cleaning, pretreatment and the actual plating
process. Recent advances in electroless plating were mentioned. Techniques available
78
for alloying palladium-silver membranes and what deactivation processes occur during
operation have also been described.
The main advantages of using catalytic membranes are for manipulating the reaction rate
in dehydrogenation reactions and as gas distributors to improve mass transfer. The
various dehydrogenation reactions that have been studied in membrane reactors,
including both alkane and alcohol dehydrogenation, have been discussed. For alcohol
dehydrogenation, catalyst preparation techniques, catalyst reduction and catalyst
deactivation due to sintering and coking, were discussed. Dehydrogenation kinetics was
listed and the kinetic parameters for ethanol dehydrogenation, obtained by previous
researchers, were also summarised.
79
3. EXPERIMENTAL PROCEDURES
The chapter on experimental procedures will be divided into four separate sections. The
first section (3.1) will focus on catalyst preparation and testing. The second section (3.2)
discusses the procedures followed for determining reaction kinetics. The third section
(3.3) will focus on Pd composite membrane preparation using a modified electro less .
plating technique and permeance testing of the membranes with hydrogen and nitrogen.
In the last section (3.4), the optimised catalyst and a suitable membrane are combined to
perform catalytic membrane reactor experiments. The contributions made to the field by
performing these experiments were discussed in the introduction (Chapter 1).
80
« 10%). The difference was more significant for the silica support (about 50%). The 'as
received' pellets were crushed and then sieved to obtain different particle size fractions.
A commercial magnesium oxide (MgO) powder (surface area = 27.4 m 2/g) was mixed
with a binder and pressed into extrusions. The extrusions were heated to 1200 °C to
agglomerate the powder particles. The extrusions were then crushed and sieved into only
a 300 to 850 Ilm fraction. The surface area of the particles was 16.7 m 2/g. Most of the
experimental work focussed on the silica and the alumina supports. Only a few
experiments were conducted with MgO and thus not much time was spent on trying to
prepare a catalyst with a larger surface area.
The MgO support was introduced into a flask containing a copper nitrate solution of a
specific concentration. The flask was placed on a magnetic stirrer and the solution stirred
81
for two hours. Thereafter the Cu-MgO particles were filtered, washed and dried at 90°C.
The catalyst was calcinated at 500°C and then reduced in hydrogen, in situ, at 350 °C for
two hours.
The silica and alumina supports were dried at 200°C for at least two to three hours and
then stored in a desicator. The dried supports were then placed in heated copper solutions
of different concentrations. The amount of solution required to impregnate a fixed mass
of either silica or alumina was experimentally determined. An excess of about 10%
solution was prepared in each case. The copper solution was kept warm on a hotplate
while the support was added. The support-solution mixture was stirred throughout while
adding the support particles. The hotplate was kept at about 80°C to evaporate the
remaining solution. The paste was stirred every few minutes. When all the water had
evaporated, the catalyst was dried in an oven at 120°C for at least four hours. The
catalyst was then ca1cinated at 500°C and reduced in situ in hydrogen at 350 °C for two
hours. When preparing Cu-Cr, Cu~Co and Cu-Ni-supported catalysts, copper nitrate was
mixed with the nitrate of the other metal in the same Cu to other metal ratio as required
on the support.
The catalysts used for determining catalyst stability and dehydrogenation kinetics were
prepared in a similar manner. The only difference was the reduction temperature.
Different reduction temperatures were investigated to determine the effect of sintering
during reduction.
82
The copper percentage and percentage of the other metal were determined by atomic
adsorption (AA). The catalyst (0.100 g) was dissolved in warm aqua regia. When all
metal had dissolved, distilled water was added to yield exactly 1000 ml solution. The
ppm Cu reading ofthe solution gave the Cu percentage.
83
Thennocouple
4mmid
quartz
U-tube
Liquid
~ I )I~:r'e
collector
i and gas
vent Quartz wool
Infors AG Basel
Perfusion pump Tube furnace
Catalyst bed
Enlargement
Quartz wool
Figure 3.1: Set-up used for testing the kinetics of the catalyst at the CNRS, France
84
and keep the products in the gas phase. A gas sample was extracted at the sample point
with a heated syringe. The syringe was kept inside a stainless steel tube and the
temperature of the syringe was controlled at about 100 DC. The first set of experimental
tests was performed to determine the optimal catalyst composition and was conducted at
the laboratories of the University of Stellenbosch (Stellenbosch, South Africa). The
kinetic testing was conducted at the laboratories of the IRC-CNRS (Institut de
Recherches sur la Catalyse, Centre National de la Recherche Scientifique) in Lyon,
France. Figure 3.1 is the set-up used at the CNRS. The set-up at Stellenbosch
University differed in the following way: Hastings flow controllers (HFC 202C) were
used in stead of Brooks, the inner diameter of the quartz tube was 8mm (10mm outer
diameter) and a Braun perfusion pump was used.
II
'Ii
The use of Ah03, Si02 and MgO supports were investigated for the ethanol reaction. All
the catalysts listed in Tables 3.2 and 3.3 were tested. Tables 3.4 and 3.5 list the matrix
of flow rates and temperatures used to test for Si02 and MgO (Table 3.4) and for Ah03
(Table 3.5) supported catalysts. The flow rates indicated in the tables are the liquid feed
flow rates of only the ethanol. In both cases the ethanol was diluted with nitrogen in the
molar ratio ethanol:N2 = 1:4 and then passed over the catalyst bed in the quartz tube. For
Si02 and MgO catalysts, 1 g of catalyst was used for each set of runs (6 temperatures at 4
flow rates = 24 runs). For AI20 3 catalysts, 3 g of catalyst was used for each set of runs (6
temperatures at 3 flow rates = 18 runs).
Table 3.4: Reactor conditions used to test Cu on Si0 2 and MgO supports in the ethanol
dehydrogenation reaction
0.6 mllh; 240 DC 1.6 mllh; 240 DC 3.2 mllh; 240 DC 6.4 mllh; 240 DC
0.6 mllh; 280 DC 1.6 mllh; 280 DC 3.2 mllh; 280 DC 6.4 mllh; 280 DC
0.6 mllh; 320 DC 1.6 mllh; 320 DC 3.2 mllh; 320 DC 6.4 mllh; 320 DC
0.6 mill); 360 DC 1.6 mllh; 360 DC 3.2 mllh; 360 DC 6.4 mllh; 360 DC
0.6 mllh; 400 DC 1.6 mllh; 400 DC 3.2 mllh; 400 DC 6.4 mllh; 400 DC
0.6 mllh; 440 DC 1.6 mllh; 440 DC 3.2 mllh; 440 DC 6.4 mllh; 440 DC
85
Table 3.5: Reactor conditions used to test Cu on AhO) supports III the ethanol
dehydrogenation reaction
1.6 mllh; 180°C 3.2 mllh; 180°C 6.4 mllh; 180°C
1.6 mllh; 220°C 3.2 mllh; 220°C 6.4 mllh; 220°C
1.6 mllh; 260°C 3.2 mllh; 260°C 6.4 mllh; 260°C
1.6 mllh; 300°C 3.2 mllh; 300°C 6.4 mllh; 300°C
1.6 mllh; 340°C 3.2 mllh; 340°C 6.4 mllh; 340°C
1.6 mllh; 380°C 3.2 mllh; 380°C 6.4 mllh; 380°C
Table 3.6: Reactor conditions used to test Cu on Si02 and MgO supports in the ethanol
dehydrogenation reaction
1.6 mllh; 240°C 3.2 mllh; 240°C 6.4 mllh; 240°C
1.6 mllh; 270°C 3.2 mllh; 270°C 6.4 mllh; 270°C
1.6 mllh; 300°C 3.2 mllh; 300°C 6.4 mllh; 300°C
1.6 mllh; 330°C 3.2 mllh; 330°C 6.4 mllh; 330°C
1.6 mllh; 360°C 3.2 mllh; 360°C 6.4 mllh; 360°C
1.6 mllh; 390°C 3.2 mllh; 390°C 6.4 mllh; 390°C
86
gas. Response factors of the main products were calculated after inj ecting numerous
liquid mixtures and constructing response factor curves. Response factors were not taken
as linear functions, but were determined by fitting data of many (at least 6) different
compositions of each binary mixtures. The ethanol and 2-butanol response factors are
listed in Appendix B1 and B2. A short description is also given in Appendix B1 on how
the response factors were determined experimentally.
The following system was used for determining reaction rate kinetics at the CNRS in
Lyon, France. Carbon-containing products were analysed on a HP 5850 gas
chromatograph equipped with a FID detector. Two capillary columns, a 30 m HP
Innowax column and a 30 m HP Plot/Ah03 column, were used in series. Hydrogen
analysis was done on a similar GC with a TCD detector. A Porapak Q column and a
molecular sieve column were operated in series.
87
3.2. KINETIC TESTING
The kinetics of the dehydrogenation of ethanol and the dehydrogenation of 2-butanol
were studied using a 14.4 wt % copper on silica catalyst. This was found to be the
.optimal catalyst from experiments conducted in the catalyst screening phase. All catalyst
testing was performed with the set-up shown in Figure 3.l.
To determine accurate kinetic data, interphase and intraparticle mass transfer resistance
needed to be eliminated or minimised. The data must preferably be gathered in the flow
regime, free of interphase mass transfer resistance. Intraparticle mass transfer resistance
was minimised by using small catalyst particles. A 300-425 Ilm particle size was used
for all experiments.
Several other important factors had to be taken into account when designing an
experiment for determining the kinetic parameters of a reaction. They were:
• The reactor had to be operated as a differential reactor (see Chapter 5) to be able to
accurately calculate reaction rates. A differential reactor is similar to a plug flow
reactor, but the overall conversion is kept low (typically less than 10%) by using very
small amounts of catalyst or very high feed flow rates.
• The catalyst bed had to be at a constant temperature.
• If the reaction rate is independent of the feed flow rate, but the conversion is still
high, then the catalyst mass can be changed. The amount of catalyst can be reduced,
while maintaining the same flow rate, thus lowering the conversion.
88
and the conversion documented with time increase. An undiluted 99.8% ethanol (from
Prolabo) feed was used.
Table 3.7 indicates the different reduction temperatures evaluated. After these initial
experiments, all further catalyst reductions were perfonned at 255°C for two hours in
hydrogen (25 cm3/min). The catalysts were heated up at 8 °C/min in 50 cm3/min of
nitrogen before being reduced at the reduction temperature.
Table 3.8 summarises all the stability tests perfonned on the dehydrogenation of ethanol.
Catalyst (a) is the standard 14.4 wt % Cu on silica catalyst, while catalyst (b) is a
13.5 wt % Cu, 1.0 wt % Cr and 0.5 wt % Co on silica catalyst. Cr and Co are often added
to copper to reduce sintering at a high temperature. Both catalysts were prepared in
South Africa from copper nitrate as source. The support used was a silica support from
Engelhard. Co and Cr were obtained from their nitrate salts.
Catalyst ~}
Reduction T ceC) 400 340 280 -
Reaction T ceC) 255 255 255 -
The 14.4 wt % Cu on silica catalyst (catalyst (a» was reactivated after 24 hours and
89
48 hours of operation at 400°C. This was done to determine whether coking, sintering or
both were the dominating deactivation mechanism. Reactivation consisted of the
following:
• Oxidising the catalyst at 400°C in oxygen for 4 hours,
• Cooling the catalyst down to 255 °C in oxygen,
• Reduction for 2 hours in hydrogen, and
• Heating up to 400°C in nitrogen.
The reaction was then continued at 400 °C for the next 24 hours.
90
Table 3.9: Conditions investigated for detennining interphase mass transfer limited
regime for ethanol dehydrogenation
Reaction T (OC) Catalyst mass (g) Min feed flow (mllh) Max feed flow (mllh)
200 0.35 1 14
225 0.35 1 16
250 0.20 2 20
275 0.10 2 20
300 0.10 2 25
When co-feeding ethanol with hydrogen or nitrogen, the products were analysed with a
FID detector and reaction rates calculated from acetaldehyde production. A TCD
detector was used to determine the amount of hydrogen produced when ethanol was co-
fed with acetaldehyde. In the latter case, hydrogen production was used to detennine
reaction rates.
The availability of equipment limited the tests which could be performed at feed rates
higher than 14 mllh, mainly due to the size of the mass flow controller and difficulty with
temperature unifonnity in the catalyst bed at higher flow rates. All experiments were
thus done at feed rates of 14 mllh and reaction rates had to be adapted, using curves
constructed from data in Table 3.9, to compensate for mass transfer resistance. For the
ethanol dehydrogenation reaction, the reaction rate was determined at 70 different
conditions. At least three injections were made into the GC at each condition and the
91
average value was used for modelling. The conditions for determining the kinetic
parameters were as follows:
• T = 200, 225, 250, 275, 300°C, and
• at each temperature:
• N2 molar % in feed = 15, 30, 45, 57 mole %,
• H2 molar % in feed = 15, 30, 45, 57 mole %, and
• Acetaldehyde mass % in feed = 4.4, 11.6,21.3,29.8,50 mass %.
Every day a new catalyst sample was used for experiments. The reaction rates for the
catalyst (at the operating temperature) were determined using pure ethanol as feed. All
data was determined relative to a set of reference values. Reference values at the same
temperature were averaged and each data set normalised, relative to the global average at
each temperature. This yielded more accurate and consistent data for modelling.
A feed flow rate of 12 mllh was used for all further experiments with 2-butanol. The
following mixtures were fed to the catalyst at each of the reaction temperatures:
• N2 molar % in feed = 10, 25, 40, 55, 70 mole %,
• H2 molar % in feed = 10,25,40,55,70 mole %, and
• Methyl ethel ketone mass % in feed = 6.7,10.6,20.1,40.3,62.8 mass %.
92
Table 3.10: Conditions investigated for determining interphase mass transfer limited
regime for 2-butanol dehydrogenation
Reaction T (0C) Catalyst mass (g) Min feed flow (ml!h) Max feed flow (ml!h)
190 0.15 2 12
220 0.15 2 16
250 0.075 2 16
280 0.075 2 16
Data was gathered in a similar marmer to that for ethanol dehydrogenation. Reaction
rates were determined relative to a set of reference values, which was then used to
calculate more accurate and consistent data for modelling. A GC, with a TCD detector
and packed columns, was used for determining hydrogen production rates when MEK
was co-fed with 2-butanol. For all other experiments, products were analysed with a FID
detector and capillary columns.
The membranes had a length of 250 mm, an outside diameter of 10 mm and an inside
diameter of 7 mm. The final layer had a pore size of 200 nm. Further characteristics
regarding pore diameter, thickness and porosity of each layer are presented in Table 3.11.
93
4: Pd rmallayer
constant 30 min
12200 C - - - - - - - - - - - - - _..- : - - - - - - - - ,
llOO ·C
oven swi1l:hed
off
20 0 C/min
25°C
94
The outside membrane ends had to be sealed with an enamel supplied by SCT prior to
plating in order to achieve proper membrane reactor sealing. This process fiHed the
outside membrane pores with a non-porous material, thus preventing gas leaks along the
outside membrane surface during the testing stage. Enamel was applied along a length of
10 mm at the membrane ends by dipping the membranes in the enamel slurry. Enamelled
membranes were then placed in a high temperature furnace (from Vecstar Furnaces) .and
cured according to the procedure suggested by SCT and indicated in Figure 3.3. Two or
three layers of enamel were applied on the membrane ends to obtain good sealing.
95
composition of the pre-treatment solution was similar to that of Shu, Grandjean et al.
(1993).
The outside surface of the membrane tubes were wrapped with PTFE tape so that only the
insides of the tubes would be catalysed. The set-up for pre-treatment is shown in Figure
3.4. The stirrer was a RW 11 basic stirrer from IKA Labortechnik (0 to 2000 rpm) and
the stirring speed was set at about 1200 rpm.
\-=+__ Rotating
device
_ _ _ Membrane
_ _ _ Pretreatment
solution
96
The procedures and their sequence for pre-treatment are tabulated in Table 3.13.
Approximately 270 ml of each solution was used. Fresh tin solution was prepared for
every membrane, while the Pd solution was changed after every three membranes
catalysed.
After pre-treatment, the teflon tape was removed and the membrane stirred in clean
distilled water for an additional half an hour to remove any solution in the membrane
pores. The membrane was then placed in an oven at 200°C overnight and the mass
recorded the next morning after cooling: The mass increase varied between 13 and 16
mg for different membranes prepared using a 0.45 g per litre SnCh sensitising solution.
Some membranes were pre-treated with a 0.25 g per litre SnCh sensitising solution, in
which case the mass increase was about 8 to 10 mg.
97
Table 3.14: Composition of the Pd plating solution per litre (for 2.00 gllitre Pd in
solution)
(NH3)4PdCIz.H20 (g) 4.94
28 wt % Ammonia (ml) 400
EDTA(g) 80
35 wt % hydrazine (ml) 0.65 (hydrazine:Pd - 0.35:1)
increased with time
Temperature (oq 71-73
Three different methods for feeding plating solution to the inside of the membrane tube
were tested:
I. Plating solution was pumped continuously through the inside of the tube at flow rates
varying between 90 and 120 ml per hour.
2. A membrane was covered on the outside with teflon tape and stirred in the plating
solution.
3. A batch process was used in which between 10 and 12 ml plating solution was
introduced into the sealed tube at a time and allowed to react for a fixed period of
time.
The third process was chosen for preparing all further membranes (see discussion in
Chapter 6). The plating set-up is shown in Figure 3.5. The membrane was sealed in a
teflon reactor with O-rings. The reactor had a single shell side outlet allowing for a
vacuum to be pulled on the shell side. A defect-plugging technique was developed to
produce thin films « 2 microns). An initial I micron Pd base was deposited on the inside
of the membrane tube using 35 ml plating solution without any vacuum applied. The
membrane was closed off at the bottom with a 10 mm silicon tube which was closed at
one side. A silicon tube of about 15 cm was also placed over the membrane at the top.
Between 11 and 12 ml plating solution was introduced into the membrane tube fixed in
the reactor at a time.
"'i
98
Control volve
ShellsUie
vacuum --<>1<:1-----, ,------,
Liquid collector
Sealed membrane
Teflon plaililg
reacior
Water bath
(71.73 ·C)
Closed off
membrane
Three to four plating sessions were performed with the same plating solution. Repeated
plating sessions ensured that all the Pd in solution was deposited on the membrane. The
hydrazine concentration was increased after each plating session to compensate for
thermal decomposition of hydrazine. The plating procedure as outlined in Table 3.15, is
essentially a batch process repeated several times. When preparing a 1 micron film, each
session in Table 3.15 was repeated 3 times (11.5 ml *3= 34.5 ml solution). The first
0.15 ml of 1.75 wt % hydrazine that was added to the 1l.5 ml plating solution (see Table
3.15) is equivalent to 0.65 ml of 35 wt % hydrazine per litre of plating solution, as
mentioned in Table 3.14.
Membranes were then cleaned and dried overnight before the next layers were applied.
The effect of Pd film thickness on hydrogen and nitrogen permeances was investigated.
For thicker films (> 2 microns), a thicker initial base was applied, but for membranes of
between 1 and 2 microns total thickness, the first layer was always about 0.9 to 1.1
microns. Membrane cleaning or post-treatment will be discussed in the next section.
99
Table 3.15: Plating procedure used for producing Pd films
Plating session Reaction time for Vol. 1.75 wt %
11.5 ml plating hydrazine added for
solution (min) 11.5 ml solution (ml)
I 20 0.15
2 20 3 drops
3 20 0.5
After the initial Pd layer was deposited, an additional one, two or three layers were
deposited to obtain the final product. The thickness of each layer depended on the
required thickness of the final Pd film: for example a film with a final thickness of
1.5 !lm required two extra layers of 0.25 !lm each. After the second layer, the membrane
was once again cleaned and dried overnight before applying the third layer.
For the second and third Pd layers, a vacuum was applied on the shell side of the teflon
reactor. Pd solution will concentrate in the more permeable or defected areas in the
membrane. More plating will occur in weakly plated (defected) areas, .film defects will
be covered with palladium and a film with less defects will result.
Pd membrane preparation focussed on preparing a thin film of thickness less than 2 !lm.
Some films of between 2 and 5 microns were also prepared to study the effect of film
thickness on hydrogen and nitrogen permeance through the film.
100
work, the membranes were dried at 100°C and again (two or three times) stirred in
ammonia solution and water before drying them at 240°C. This was an attempt to try
and extract more EDTA trapped in the membrane pores.
Table 3.16: Composition of silver plating bath per litre of plating solution
Ag (ppm) 1000
AgN03 (g) 1.576
EDTA(g) 40
Ammonia (28%) (ml) 200
Hydrazine (3.5 wt %) (ml) 8.50
The hydrazine concentration (for Ag plating) was increased with time, similar to the
procedure for Pd plating (see Table 3.17). For every 11 ml fresh plating solution, the
procedure in Table 3.17 was performed. Generally, a maximum of about 75% of silver
in solution was deposited on the palladium. An excess of 30% of the required amount of
silver that needed to be deposited on the membrane was used in solution for plating.
Silver plating was performed and if the deposited silver was not sufficient, a second layer
was deposited. For the second layer, the initial silver plating solution was diluted to
ensure a volume of 11 ml plating solution. This was necessary because if the silver
plating solution (original) was less than 10 ml it would not wet the entire membrane
surface and thus not plate over the entire surface.
101
Table 3.17: Plating procedure used for producing Ag films
Plating session Reaction time for Vol. 0.35 wt %
11 ml plating hydrazine added for
solution (min) 11.5 ml solution (ml)
I 20 0.4
2 20 0.4
3 20 0.2 ml, 1.75 wt %
hydrazine
ICP (with a Spectroflame Modula from Analytical Instruments) analysis of the plating
solution was performed after plating to determine the amount of palladium deposited.
Since the initial mass of the plating solution was known, the mass of the deposited Pd
could be calculated. The average of this mass and the measured mass was used to
calculate Pd film thickness. The calculated film thickness is an average thickness value.
102
against th~ enamelled membrane endings. If the nuts were sufficiently tightened, a very
good seal was obtained. Over tightening could result in either breaking or cracking of the
membrane.
r
Tube side
--t
exit
Tube side
Graphite s ea1 Shell side Enamelle d endings
entry
Figure 3.6: Membrane reactor used to test the membrane permeance
Some difficulty was, however, still experienced when trying to obtain good reactor to
membrane seals. The main reason was that the enamel on the outside membrane surface
was not always of uniform thickness. The equipment shown in Figure 3.7 was used to
minimise the leak resulting from membrane sealing.
Nitrogen
Exit side
closed off
Membrane sealed
at both ends
3 way valve
103
The membrane was placed inside the reactor, with graphite rings at the edges (see
Figure 3.6) and the nuts tightened moderately. The reactor (with the exit side closed off)
was then connected to a pressure controller and two mass flow meters as indicated in
Figure 3.7. The pressure controller was set at 400.0 mbar and the value of the mass flow
meters monitored while reactor nuts were tightened. Tightening was stopped when, upon
further tightening of the nuts, there was no further decline in the reading of the mass flow
meters. The nuts on opposing sides were not necessarily turned or tightened equally, to
obtain the best seal. Sometimes the reactor had to be re-opened, the membrane shifted
slightly and nuts re-tightened to obtain a good seal.
Figures 3.8 and 3.9 show the. equipment used for high temperature membrane testing
with hydrogen, nitrogen and ethanol as feeds, respectively. In both instances the reactor
was wrapped with heating wire and insulated. A thermocouple was placed in the centre
ofthe membrane tube to record the temperature.
For hydrogen and nitrogen testing (Figure 3.8), one of the two shell side tubes of the
reactor was closed. The reactor was operated in the dead end mode, in other words, the
exit tube side was closed and the feed gas forced through the Pd film. The temperature
inside the reactor was varied between 330°C and 450 °C, using a temperature controller.
The flow rate of the permeated gas was measured using two bubble flow meters. A 0 to
100 ml flow meter was used for hydrogen measurements and a 0 to 4 ml flow meter for
nitrogen. The effect of differential pressure on hydrogen and nitrogen permeance was
studied. The mass flow controllers were set on different flow rates and the differential
pressure recorded. Initial testing was conducted at the laboratories of the CNRS (Lyon,
France). For hydrogen permeance, the maximum differential pressure that could be
tested was limited by the mass flow controller (0 to 600 cm'/min). For films in the order
of 1.5 flm, this maximum differential pressure was less than 100 mbar at 450°C. The
pressure probe had a maximum measuring ability of 2000 mbar, which was the limit for
testing nitrogen permeance.
104
Closed
Hydrogen r+_i~~:-~T:cl
I I I
Nitrogen
I ,
L.. ______ ..J
Tube side closed
Heated zone
with sealed Shell side
membrane exit
MFC: Brooks mass flow controller (5850 TR) for hydrogen; 0-600 cm3/min
MFC: Brooks mass flow controller (5850 E) for nitrogen; 0-100 cm3/min
P: Keller tube side pressure probe (0-2000 mbar)
with WEST (8100) displayer
T: Thermocouple
TC: Temperature controller for heated zone
BFM: Bubble flow meter for nitrogen (0-2 ml)
BFM: Bubble flow meter for hydrogen (O-IDO ml)
...... Ball valve
~ Needle valve
Figure 3.8: Set-up used for high temperature (> 300°C) hydrogen and nitrogen permeation testing
Ethane sweep gas
r-----iT,ci
_1 __ ,
I
Safety
pressure
gauge
L.. _ _ _ _ _ _ l.-_--I~ Tube side closed
Autoclave with
30 mI ethanol Heated zone
with sealed Exit to vent
temp erature membrane '-------ll GC
control
MFC: Brooks mass flow controller (5850 TR) for ethane; 0-50 cm3lmin
P: Keller tube side pressure probe (0-2000 mbar) with WEST (8100) displayer
T: Thermocouples
TC: Temperature controller for heated zone
..... Ball valve •
r:ki Needle valve
** All tubing containing ethanol is heated to 130°C using heating wire .*
106
Further testing on both pure palladium films and Pd-Ag alloy films was conducted at the
University of Stellenbosch, South Africa. At Stellenbosch, Hastings flow controllers
(HFC 202C) were used instead of Brooks and they had a maximum capacity of 1000
cm'/min. The pressure probe also had a maximum of2000 mbar.
The apparatus used for ethanol permeance testing is shown in Figure 3.9. An ethane
sweep gas was used and fed to a gas chromatograph after moving pass the membrane. By
measuring the ethanol content in the sweep gas, the ethanol permeance could be
calculated. All lines containing ethanol were wrapped with heating wire and insulated.
The temperature of the heated lines was kept between I 10°C and 130°C to prevent any
condensation in the lines. An autoclave was used to feed the ethanol to the reactor. The
autoclave was well insulated and heated on a temperature controlled hotplate from lKA
Labortechnik. The desired tube side pressure could be obtained by correctly setting the
temperature of the hotplate.
The autoclave was filled with 30 ml 99.8% ethanol. The testing pressure was stable and
fluctuated by less than I %. The slow ethanol permeance ensured a constant tube side
pressure.
After testing, the Pd membrane was kept in nitrogen at 330°C for 2 hours and then
cooled at a rate of2 °Clmin to room temperature (also in nitrogen).
107
000,--------------------------------,
· ................ . ....
Avenlge." 301 "C
. ... - ........... - '--""'"'' .....
Awnlge • 21M ·c
_crv.n: 100'C . - - - - - - - - - - - . - - - - - - - - - - - ... _____ _
- . - o..n._200"C
...... o..n_300'C
100 -"... 0\I..,=400'Cr+--------+--------+-_ _ _ _ -l
_o.toro:440'C
o~-- __----~----~--__----~----~
o 2 4 6 8 10 12
Distance from reactor centre (em)
There were two thermocouples for measuring temperature. One was situated inside 'the
membrane tube and one was situated on the outside of the reactor, next to the reactor
wall. The temperature controller was connected to the thermocouple on the outside of the
reactor wall, since that temperature provided for more stable temperature control than the
thermocouple inside the membrane tube. The temperature inside the membrane tube was
measured, but not used for control. The temperature difference between the two
thermocouples was taken into consideration when setting the reaction temperature. In
other words, to control the reaction temperature at for example 300°C, the oven was
adjusted by the difference and set at 291 °C. At 300°C and below, the temperature
profiles were excellent from the centre of the reactor up to a distance of 9 cm from the
centre. The variance from the average temperature was less than 5°C. In the last 3.5 em,
the temperature dropped by between 10 and 13 °C. The total membrane length was
25 cm. Catalytic membrane reactor experiments were performed below 300°C and the
variance in reactor temperature was small enough to assume isothermal conditions. At
400 °c and above, the difference between the oven and reactor temperatures increased.
The variance around the average was also larger and increased to between 10 and 15°C.
108
450°C for one and a half hours. A very high EDT A concentration (80 g per litre) was
used for plating. The cleaning process and overnight oxidation at 240°C did not remove
all the precursor. It was found that even after reduction not all the carbon in the
membrane pores was removed either by oxidation or by thermal decomposition. The
heating process was then changed and it produced membranes with higher hydrogen
permeances and selectivities. The process was changed as follows:
• Heat the membrane in nitrogen from room temperature to 320°C, at 2.5 °C/min.
• Switch from using nitrogen to oxygen and force 10 cm3 /min oxygen from the tube
side through the membrane pores to the shell side. Cary out oxidation for 2 hours.
• Switch back to nitrogen and heat from 320°C to 450 °C, at 2.5 °C/min.
• Reduce at 450°C in hydrogen for 1.5 hours.
The alloying experiments were performed on membranes with metal coatings of similar
thicknesses. This was important, to be able to compare results of different membranes.
Both heating methods were investigated. Membranes were placed either in a constant
109
temperature tube furnace or in the reactor (Figure 3.6) and heated at a rate of 1.5 DC/min
to the required temperature. Initial experiments were performed with the membranes in
the free mode under the following conditions:
• argon and hydrogen atmospheres,
• alloying temperatures ranging from 520 DC to 600 DC, and
• alloying times up to 150 hours.
The amount of data obtained from membranes heated in the tube furnace, was limited.
The oven had to be cooled down (at 1.5 DC/min) to room temperature after each
experiment, and the membrane placed in the reactor for testing, before data could be
gathered. Not only did the thermal cycling weaken the coating, but the data was also of a
discontinuous nature.
Most of the alloying was performed in situ in the reactor. That allowed for continuous
monitoring of the process. Permeance and selectivity data could be obtained at any time.
The following conditions were tested during alloying in the reactor:
• argon, nitrogen and hydrogen atmospheres,
• alloying temperatures from 500 DC to 600 °C, and
• alloying times from 10 hours to more than 100 hours.
110
3.4. CATALYTIC MEMBRANE REACTOR EXPERIMENTS
Catalytic membrane reactor experiments were performed in a set-up similar to that shown
in Figure 3.1. The quartz tube in Figure 3.1 was replaced by the reactor in Figure 3.6,
with temperature profiles as in Figure 3.10. The alcohol feed passed through a 1 meter
coil, which was placed in a pre-heating oven and connected to the reactor inlet. The pre-
heating oven ensured that the feed entered the reactor at the reaction temperature. Exit
lines were heated with heating wire to keep the products in the gas phase. A heated
syringe (120°C) was used for taking gas samples and injecting them into a Gc. The
products were analysed with a HP G 1800A gas chromatograph, equipped with a mass
spectrometer and a flame ionisation detector (for more details see section 3.1.5).
For both the ethanol and 2-butanol reactions, the membrane was packed with a 14.4 wt %
Cu on silica catalyst. The catalyst particle size fraction was 500 to 850 microns. The
catalyst was kept in position with quartz wool at the edges of the membrane (see
Figure 3.11).
Quartz
wool
Figure 3.11: Membrane packed with catalyst
During the start-up procedure, the membrane reactor was heated in nitrogen at 2 °C/min
from room temperature up to 275°C. The membrane and the copper catalyst were then
reduced in a hydrogen atmosphere (flow of 50 cm3/min) for 1.5 hours. After 1.5 hours,
hydrogen was replaced with nitrogen for a further 10 minutes and then the reactor was
either heated or cooled at 2 °C/min to the required reaction temperature. At the required
reaction temperature, the alcohol was introduced at 10 mllh for 1.5 hours before analysis.
111
3.4.1. ETHANOL DEHYDROGENATION
The ethanol dehydrogenation reaction in the catalytic membrane reactor was investigated
from 250°C to 300 °c. The feed rate depended on the reaction temperature, but varied
from 5 mIlh to 30 mllh. The membrane was packed with 3.00 g catalyst. The following
conditions were used for testing ethanol dehydrogenation in a membrane reactor:
• 250°C: 5,10 mIlh feed rate,
• 275.o C: 5, 10, 20, 30 mIlh feed rate, and
• 300°C: 10, 20 mIlh feed rate.
At each feed flow rate, four or five different sweep gas flow rates were tested and the
dehydrogenation results compared to results obtained when no sweep gas was used (i.e. a
plug flow reator). Nitrogen was used as the sweep gas and the flow was co-current for all
experiments.
112
3.5. SUMMARY
This chapter covered all experimental work performed. There were three basic sections:
• Catalyst optimisation and kinetic testing,
• Membrane optimisation and permeance testing, and
• Catalytic membrane reactor experiments.
The' equipment and procedures used in each section were described in detail. The
variables that were investigated, were listed. The values at which each variable were
tested, were also indicated.
113
4. OPTIMISING CATALYST COMPOSITION
Both the dehydrogenation of ethanol and 2-butanol were investigated. For each reaction,
different catalysts were investigated at a range of operating conditions. The composition of
the catalysts and the conditions at which they were tested were discussed in Chapter 3. All
the data presented in this chapter is for average WfF values at specific temperatures and
copper .loadings. The WfF ratio is the catalyst mass divided by the feed flow rate. A low
WfF value indicates a fast feed rate and a high WfF value indicates a slow feed rate for a
constant mass of catalyst. Each experiment was performed at either three or four different
feed flow rates. The conversions, yields and selectivities discussed are average values, unless
otherwise stated. The following definitions were used:
114
I-butanone, 2-butanone and I-butanol. Below 360°C there was very little reaction of
ethanol (Figure 4.1). At low temperatures, acetaldehyde production was low (Figure 4.2),
but the selectivity towards acetaldehyde was high (Figure 4.3). When the temperature
increased, both the total ethanol conversion and the conversion to acetaldehyde (this is called
the acetaldehyde yield) increased sharply.
Figure 4.1: Total ethanol conversion for Figure 4.2: Ethanol to acetaldehyde yield
Cu on MgO catalysts (Cu on MgO catalysts)
The difference between the values in Figure 4.1 and Figure 4.2 is the ethanol converted to
other products and this difference is expressed via the selectivity curve. At 440°C the
115
selectivity dropped to between 25% and 35% for the vanous copper percentages. The
majority of ethanol was not converted to acetaldehyde, but to I-butanol. Higher temperatures
not only favoured acetaldehyde formation, but also C4 formation.
J'he Cu content on the catalyst has an effect on the BET surface area of the catalyst. When
the Cu content increased, the BET surface area decreased and less support sites were
available for reaction (Table 4.1). With no copper on the catalyst, the main product was
ethene and acetaldehyde was the main by-product. As soon as copper was deposited on the
support, ethene production ceased and acetaldehyde production increased. Figure 4.2 shows
an increase in acetaldehyde yield with an increase in Cu content. The selectivity towards
acetaldehyde formation (Figure 4.3) remained fairly constant with an increase in copper.
The ethene that was formed in the absence of copper, was replaced by I-butanol in the
presence of copper. Table 4.2 gives a summary of the main product and by-product at
different conditions.
Using MgO catalysts, both total ethanol converSIOn and acetaldehyde yield were poor,
because of the low BET surface area of the catalysts (see Table 4.1). Iwasa and Takezawa
(1991) studied ethanol dehydrogenation with a 30 wt % Cu on MgO catalyst at 220°C. The
selectivity towards acetaldehyde production was 74% with ethyl acetate and other C4 species
the 1!lain by-products. Takezawa et aL (1975) concluded that acetaldehyde was the main
product when ethanol reacted over pure MgO between 340°C and 360 °C. The selectivity
under differential conditions was above 90% at those temperatures. A reaction mechanism for
116
pure MgO catalysts was discussed by Takezawa et al. (1975). Results in this dissertation
differ from the results obtained by Takezawa et al. (1975). For a pure MgO catalyst the main
product was ethene at 340°C and the selectivity towards acetaldehyde poor.
Table 4.2: Yield matrix indicating main products and by-products for CulMgO
320/360°C 440°C
O%Cu Ethene (1.6%) Ethene (17.8%)
Acetaldehyde (0.8%) Acetaldehyde (9.1 %)
117
Figure 4.7 shows the effect of flow rate and temperature on di-ethyl ether yield for a 0"10 eu
on alumina catalyst. The curves have similar profiles. With a decrease in WIF (due to an
increase in F), the optimum ether yield shifts towards the higher temperatures.
10 0
" eo
~ eo
t
<! ~o
~
.0
Figure 4.4: Total ethanol conversion Figure 4.5: Ethanol to ethene yield
(eu on alumina catalysts) (eu on alumina catalysts)
60.0
...-
+----,;'::,-:''-.-------------i
~ SO.O
A"\
+--1/,,~:,-:".~\'-:>'\--"---l~_=w;;;;IF:::,-;;;393;;:..:::.-~r--i,
~ 40.0 .y/'-i-
"i----\-4--1 _WIF::I 197 kg.shnaI H
10.0
\ ,',
t-------\:--';'.~=--...;:::_--i
_____
'---"""
0.0 +--_--~----'_-~.__-""'"
1eo 220 260 300 380
Reaction T (·C)
Figure 4.6: Ethanol to di-ethyl ether yield Figure 4.7: Ethanol conversion to di-ethyl
(eu on alumina catalysts) ether (0% eu on alumina) at different WIF
ratios
The copper concentration on the support did not have much of an effect on the total ethanol
conversion, but had a definite effect on the product distribution, Pure alumina gave the
highest ethene and di-ethyl ether yields. The reason was that pure alumina represented the
118
maximum available acidic sites, resulting in the maximum dehydration products. The ethene
and the di-ethyl ether yields dropped when the catalyst was impregnated with copper. For
10% Cu loading and above, the yields remained within a narrow band, with not more than 5%
variation from the average.
The main by-products were butenes, acetaldehyde and a small amount of hexenes. Figures
4.8 and 4.9 show acetaldehyde and butene yields with temperature increasing from left to
right. For acetaldehyde production there was an optimum copper loading on the alumina
(13.2 wt %), which resulted in the maximum acetaldehyde production. The butene
production increased with temperature and was the highest for 8.8 wt % Cu and 13.2 wt %
Cu on alumina. Higher and lower Cu loadings caused a decline in butene production.
T'" - .. -
YIoId(%1
Figure 4.8: Acetaldehyde yields for Cu on Figure 4.9: Butene yields for Cu on alumina
alumina catalysts (W/F = 197 kg.s/mol) catalysts (average W/F)
De Boer and Visseren (1971) observed a proportionality between the dehydration rate
constants and the amount of aluminum on the surface. Iwasa and Takezawa (1991) tested an
alumina-based catalyst with a high copper loading (30 wt %) at 220°C. They reported 54%
selectivity towards acetaldehyde production, 19% towards di-etheyl ether and 22% towards
C4 production.
Copper alumina catalysts have often been used for cyclohexanol dehydrogenation. The
selectivity towards the dehydrogenation reaction was higher for cyciohexanol than for
ethanol on alumina-based catalysts. The effects of copper concentration on the selectivity
119
and activity towards cyclohexanone fonnation have been investigated (Chang and Saleque,
1993, 1994; Sivaraj et a1., 1990).
x ~
1000
".
.
16.00
g ..
:!f 1.60 , ,,
"'-
14.00
~
155 E 1.50 ,, U
:
""'.. -" ""'- ""'-
12.00
,
". i ,!!I 1.40 , 10.00
i•
'"
"E
'".. l '
---- 8.00
<!•
o ---
-,. "'.
-;; 1.30
. ~ , .i '.20
I
'.00
". '.00
• • '" 1.10
2.00
'00
• ". 1.00 '.00
10
Cu%
15
" " • 12
"
Cu%
20
"
Figure 4.10: Effect ofCu% on catalyst Figure 4.11: Effect ofCu% on Cu surface
surface area for Cu on alumina catalysts area for alumina-based catalysts
The total surface area (Figure 4.10) declined from 215 ml/g at 0 wt % Cu to 120 ml/g at
24 wt % Cu as the deposited copper filled the micro- and mesopores in the support. Ethene
and di-ethyl ether fonned mainly on the acidic sites, because ethene and di-ethyl ether yields
dropped (Figures 4.5 and 4.6) when copper was added to alumina, to reduce the number of
acidic sites. Even though the catalysts' surface area was almost halved for 0 to 24 wt % Cu
loadings, the ethanol converted remained constant. The deposited copper sites provided
additional catalytic activity and resulted in more products being formed (mainly butenes and
acetaldehyde) to compensate for the loss of dehydration products (ethene and di-ethyl ether).
The added copper sites were much more active than the alumina sites, since the small
increase in copper surface area (Figure 4.11) fully compensated for the large decline in total
surface area (Figure 4.10).
120
Copper surfaces areas were detennined as described in Sections 3.1.6 and 4.1.2.2.2. The
copper surface area (in ml/g sample) had an optimum value at 13.2 wt % Cu on alumina. At
that value, the maximum acetaldehyde and butene yields were observed (Figures 4.8 and
4.9). From this it could be concluded that copper on alumina catalyses the formation of
acetaldehyde and butenes.
35.00
"iii 30.00
Copper % on support
!. 25.00 increases for lines from
tDpm boIlcIm at 100A
..i!i!! 20.00
&.
"ii 15.00
1
.E
10 00
.
5.00
10 100 1000
Awrage pore alze {AI
Figure 4.12 clearly shows how the pore size distribution changed when the copper loading
on the support increased. The y-axis is the incremental pore area for every pore fraction from
very large to very small pores. The curve with the highest differential area values is the 0%
Cu line. As the Cu % increases the differential areas decrease (lines drop on the y-axis for
example at 100 A pore size). The line with the lowest differential area values is the 24 wt %
Cu one. Figure 4.12 shows both a shift in the pore size distribution to higher pore sizes and a
decline in the values on the differential area axis with an increase in Cu loading. The average
pore size changes from 132 A for a 0 wt % Cu on alumina catalyst to 166 A for a 24 wt % Cu
on alumina catalyst.
Result were in line with those of other researchers (see Tables 4.3 and 4.4) who also obtained
a optimum copper percentage (on alumina) giving a maximum copper surface area on their
catalysts.
121
4.1.2.2. Literature data on surface areas of copper catalysts
There are two methods for determining copper surface areas. One method is based on N 20
titration and the other method on either CO chemisorption or H2 chemisorption. For both
methods, the catalyst must be reduced in hydrogen prior to analysis. The catalyst reduction
step has been discussed in Chapter 2 (2.10.3). Bonding between copper species and the
support sites, interactions between different copper species and the formation of amorphous
phases are only a few of the problems encountered when reducing the copper ions to metallic
copper. The result is very poor reduction. Better reduction requires higher reduction
temperatures, which result in sintering (see 2.10.4.1) and further deviation from a true copper
surface area. Experimentally-determined copper surface areas thus have a significant margin
of error.
4.1.2.2.1. N 20 titration
Sivaraj and Kantarao (1988a) described the N 20 titration method. The nitrous oxide
decomposes according to the following equation:
(4.6)
where nm is the total amount of nitrous oxide molecules that decompose, Xm is the
chemisorption stoichiometry and ns is the amount of copper metal atoms per unit surface area
(1.47*10 19 m,2). The catalyst is first reduced in hydrogen for 5 hours at 250°C, evacuated (to
10'6 Torr) for 2 hours and then reacted with N20 at 200 Torr and 90°C. The reaction does
not cause a pressure change. After 5 hours of reaction, the remaining N 20 is frozen out in a
nitrogen trap and the pressure difference between the initial pressure and final pressure is
used to calculate the amount ofN20 that reacted.
122
4.1.2.2.2. Chemisorption
A detailed discussion on this process can be found in Jeon and Chung (1994). The same
apparatus used for measuring BET surface areas is used for chemisorption experiments. The
catalyst is first reduced and then either CO or H2 is used as the analysis gas. The method
works on the principle that these two gases adsorb reversibly on CuD, but not on Cu l +.
Chemisorption is carried out and the first adsorption isotherm constructed. This value
represents adsorption onto both Cuoand Cu l +. The catalyst is then evacuated and the gas
desorbs from CUD, but not from Cu l +. The experiment is repeated, giving a second adsorption
isotherm. The second isotherm represents CuD, while the difference between the isotherms
represents CUI +.
123
Many authors have reported either BET or chemisorption data, but not both. In most cases
the optimum copper area that was observed, was in accordance with what was discussed in
section 2.10.1. An exception was Jeon and Chung (1994), who prepared silica-supported
catalysts.
124
Chang and Saleque (1993) compared the use of electroless plating, impregnation and
precipitation for preparing copper on alumina catalysts. The total BET area declined in that
same order, but the BET values showed only a small variance. Similar preparation methods
resulted in significant differences in the copper surface areas. Compare values from:
• Kanoun et al (1993) and Tu et al. (1994a),
• Sivaraj and Kantarao (1988a) and Sivaraj et al. (1990), and
• Pure copper values from Kanoun et al. (1991) and Tu et al. (1994a).
The data presented from Jeon and Chung (1994) in Table 4.4, show similar Cu percentages
on Si02 . Cu surface areas were obtained for different pH-values of the precipitate.
'80
Tempo_.. rC) "" TemperatuN (OC)
380
Figure 4.13: Effect ofCr, Ni and Co (on Figure 4.14: Effect ofCr, Ni and Co
alumina) on total ethanol conversion (on alumina) on ether yield
125
70
618.5%CU
!~r----------__--
~ 20 +------------IM--
Co and especially Cr addition made the catalyst more selective for di-ethyl ether formation
(Figure 4.14), with significant improvements at 220°C and 260 °C. For ethene formation
(Figure 4.15), pure copper on alumina resulted in the best performance at all temperatures,
followed by Cu-Ni and Cu-Cr. The addition of additives did not improve the acetaldehyde
yields for copper on alumina catalysts. The acetaldehyde yields were below 8% for the
different catalysts tested in Figures 4.13 to 4.15.
There was little difference in the BET surface areas for the different catalysts as indicated in
Table 4.5. The larger difference in copper surface areas can only be due to a difference in
reduction percentages. Chromium addition makes it more difficult to reduce the eu2+ ions
with hydrogen. Some copper will be present as Cu2+ and not Cuo Less Cuo sites will be
detected, which results in a lower copper surface area.
..
Table 4 S· BET and copper surface areas for copper/alumina catalysts with additives
Cu% Additive BET area Cuarea Cu area
(m2jg) (m2/g sample) (m1/gCu)
18.5 - 140 1.58 8.54
17.7 2.09% Co 151 1.67 8.42
17.8 1.73% Cr 157 1.10 5.65
16.0 2.0%Ni 137 1.40 8.72
126
4.1.2.4. Summary for using Cu on alumina catalysts
This type of catalyst is not suitable for acetaldehyde production via the dehydrogenation
reaction. Very good yields of ethene and di-ethyl ether were obtained by suitable choice of
the copper loading, addition of the correct additives and performing the reaction at the correct
operating conditions. Alumina-based catalysts favoured the dehydration reaction. At higher
temperatures (see Figure 4.9), butene production became significant for catalysts with the
highest copper surface areas.
The different catalysts investigated were listed in Tables 3.2 and 3.3. The operating
conditions were listed in Table 3.4.
127
The amount of acetaldehyde formed on a pure silica catalyst was negligible. For pure silica,
ethene was the main product with no significant amount of by-products. Ethanol conversion
to acetaldehyde (acetaldehyde selectivity) was not very dependant on the copper loading
(Figure 4.18) for copper loadings from 4 to 34 wt %. At 440°C the selectivity varied
between a maximum of 88% (for a 18.6 wt % Cu on silica catalyst) and a minimum of 79%
(for a 4.2 wt % Cu on silica catalyst). The selectivity was more dependent on temperature.
Up to 320°C the selectivity remained in the mid 90% level, but it dropped to about 80"10 at
440 DC. The main by-products were: ethene, di-ethyl ether, etoxy ethane, I-butanone and
ethyl acetate.
\()J sO
~g: : ~
~.\O
(jJ
....t~
40
~o
~ ~o
Figure 4.16: Total ethanol conversion Figure 4.17: Ethanol to acetaldehyde yield
(Cu on silica catalysts) (Cu on silica catalysts)
128
Figures 4.19 and 4.20 show the effect of the feed flow rate on total ethanol conversion and
acetaldehyde yield for the optimum copper loading. The total ethanol conversion
(Figure 4.19) remained high, with a decrease in WfF (increase in F) to between 131 kg.sJmol
and 66 kg.sJmol. At higher flow rates, i.e. lower WfF values, the total ethanol conversion
dropped due to a shorter catalyst contact time. For the shorter contact time, the reaction
could not go to completion. There was an optimum W fF ratio for a maximum acetaldehyde
yield (Figure 4.20). At the highest W fF values (slowest feed rates) the acetaldehyde yield
decreased (Figure 4.20), due to a very long residence time that increased the formation of
propene, di-ethyl ether, I-butanone and ethyl acetate. For high flow rates (WfF < 66
kg.sJmol) the lower total ethanol conversion caused a lower acetaldehyde yield.
129
lower than for pure copper catalysts (Figures 4.21 and 4.23). The addition of Co resulted in
the poorest conversions at all temperatures. A higher Ni content (Figure 4.21) resulted in
better conversions than a lower nickel content (Figure 4.23). At both ratios, the Cu-Cr
catalysts performed similarly, with Cu:Cr = 9: 1 being slightly better. At 400 DC and 440 DC,
Cu-Cr catalysts and the one Cu-Ni catalyst (13.5 wt % Cu, 1.5 wt % Ni) gave conversions
which differed only marginally from those obtained with a pure copper catalyst.
The addition of additives made the dehydrogenation reaction far less selective. Acetaldehyde
yields decreased sharply, even at high temperatures, as indicated in Figures 4.22 and 4.24.
The by-products formed on each of the three catalyst were:
• Cu-Cr: ethene, propene, butenes, di-ethyl ether, I-butanone, MEK, ethyl acetate,
• Cu-Ni: ethene, propene, di-ethyl ether, I-butanone, and
• Cu-Co: ethene, propene, butenes, di-ethyl ether.
Kanoun et aI. (1993) calculated catalyst activities and turnover frequencies (TOFs) for Cu-
Cr-AI catalysts prepared by precipitation. Variations in the Cr content didn't have a
significant effect on the TOFs or activities. Tu et aI. (1994a) prepared unsupported Cu-Cr
catalysts via precipitation. Their results indicated an optimum ethanol dehydrogenation
activity for a Cr:Cu = 4:40 (on mol basis) catalyst. Selectivity data was not included
100
90
,,-I---------i'iIo~l!II_ so
eJ 15.0% cu
E'l15.O%Cu 70
!!6 .. +---~.-~~~
• 13,S% Cu: lao
.• ,,+----lli~_li\. 1.5% Co
"so -I-----§---I'!1--.Jii!-,=--lil---I 1113,5% Cu:
Reaction T I"C)
-400 4040 2<, ,., 32'
""
Reaction T (-C)
400 ...
Figure 4.21: Effect of Co, Cr or Ni additives Figure 4.22: Effect of Co, Cr or Ni additives
on total ethanol conversion (Cu on silica) on acetaldehyde yield (Cu on silica)
130
!II'
100
"'"90
eo
"eo
~ 70 III 15.I)%Cu
70 1!I1S.O%CU
ceo
.2 .1425%CU; l
~" 0.75% Co 50
11 t----;::----tI---1H!--li-1a--I-fa-i1
•c
'" ~14.25%CU:
O.75%Cr >"" iii 14.26% Cu;
D.7a%er
8" • 14.25'1' CU; .14.~Cu:
2D Q,15%Nl 20 O.75'l1oNi
10 10
Figure 4.23: Effect of Co, Cr or Ni additives Figure 4.24: Effect of Co, Cr or Ni additives
on total ethanol conversion (Cu on silica) on acetaldehyde yield (Cu on silica)
100 100
"so
so 1lI1&J.300
.,""'"
.- .-
micrmlll; mlaOllIll;
14.1%C\I 70 14.1% OJ
l
j mlCf1)l'lS; ~eo mia-ens;
1S.Q%CU 15.0%0"
~50 :l!"
• 1i1850-1180
C
iJ85D-1180 ):40 IIIlerm.;
microns:
ll.ncu
8" .>
13.~CU
3000
30
.> 3000
20 mlCfons;
",Iaoo.;
13.7%Cu 10
13.7% eu
240 280 320 360 "'" .., 280 320 3110 400 «0
Figure 4.25: Effect of catalyst particle size Figure 4.26: Effect of catalyst particle size
on total ethanol conversion on acetaldehyde yield percentage
Two effects are responsible for the decrease in ethanol conversion using the larger particles:
channelling of the feed gas and intra-particle mass transfer resistance. Channelling usually
131
starts occurring when the particle diameter is less than one tenth of the reactor diameter.
Some channelling occurred for the 3 mm catalyst spheres, because the inside diameter of the
quarts tube that housed the catalyst was only 8 mm. For the smaller particles (1.2 mm and
less) channelling was less dominant. The decrease in ethanol conversion with the larger
catalyst particles confirmed that those two effect were present during the ethanol reactions.
Figure 4.26 indicates that at 360°C and above, the acetaldehyde yield was similar for all
particles up to a size of 1180 I-lm. The mass transfer limitation for ethanol dehydrogenation
to acetaldehyde at 360°C and above was not significant for particles less than 1180 I-lm. A
further observation that could be made when comparing Figures 4.25 and 4.26, was that the
selectivity towards acetaldehyde decreased at high temperatures as the catalyst particles
became smaller.
132
!'"
-i'<---I
r _BET.r.. I1------"'--1 4.5
'-
430
.- \ I • A........ pare size I • 93 4.0
i 380"'--.... 91 g
3.5
S'
'k
t '-=,-""'-..o--------i 89~
2.5
.. 2lIO j------ L
2.0
J
~
1.5
230 j -_ _ _ _ _ _ __ _---=:""'"____I 87
Figure 4.27: Effect ofCu% on catalyst Figure 4.28: Effect of Cu% on Cu surface
surface area for Cu on silica catalysts area for silica-based catalysts
.0,-------------------,
Copper % on support
Increases for linea from
top to bottom at 90 A
0~~~------_ _----~~~--_4
10 1(,) 1000
Average pore size (A)
The acetaldehyde yield data (Figure 4.17) and copper surface area data (Figure 4.28) do not
correlate well. Many studies have shown that dehydrogenation activity is proportional to the
copper surface area (Tu et al., 1994b; Jeon and Chung, 1994), but in the present study there
was a difference between the two. The best explanation for this is the unreliability in copper
surface area data. Jeon and Chung (1994) did an in depth study on copper on silica catalyst
characterisation. They experienced similar problems with determining copper surface areas
on silica supports and drew the following general conclusions:
• Copper reduction ranges from 80% in the best cases to as low as 40%, depending on the
copper percentage on the silica support. For low copper content catalysts, the strong
silica-copper bond makes reduction very difficult.
133
• Silica solubility increases with solution temperature. An amorphous copper silicate phase
forms or can form. This contaminates copper sites, which can inhibit the
dehydrogenation reaction and prevent chemisorption.
• Copper surface areas determined by chemisorption, X-ray line broadening and N 20
titration varied by up to a factor of 3, depending on the method employed.
• Copper particles range from small to very large. The catalyst prepared by precipitation is
not very homogeneous.
In the present study silica was impregnated with copper solution at about 80°C, which could
lead to the formation of a copper-silicate phase. Since both copper-silica and silica are
amorphous, it is difficult to detect the different phases with X-ray diffraction analysis. The
existence of a copper-silicate phase and the poor reduction at the lower copper loadings could
possibly be responsible for the low copper surface areas and the difference in acetaldehyde
yields and copper surface areas. Further TEM and X-ray analyses were performed on a 14.5
wt % Cu on silica catalyst to investigate copper crystallite size and crystallite distribution.
134
Fig 4.30a: TEM (20 nm) Fig 4.30b: TEM (20 nm) Fig 4.3Oc: TEM (20 nm)
Fig 4.30d: TEM (20 nm) Fig 4.30e: TEM (20 nm) Fig 4.3Of: TEM (20 nm)
Fig 4.30g: TEM (50 nm) Fig 4.30h: TEM (200 nm) Fig 4.30i: TEM (200 nm)
Figure 4.30: TEM images of the unused 14.4 wt % copper on silica catalyst
135
X-ray line broadening can be used to calculate average crystallite size, employing Schearer's
equation (equation 4.7). The crystallite diameter, D in A, can be expressed as (Tu et aI.,
1994b):
D _ 1.542
hkI- (4.7)
xp cos(2B)
with x the peak width in radians at 2/3 of the peak height and 29 the angle in degrees at which
the peak is recorded. The peak width was 0.20 degrees at a 29 angle of 43.3 degrees. This
yielded an average crystallite size of about 60 nm. For such large crystals the Schearer's
equation was not very accurate, but still gave a fair estimate of crystallite size.
136
MgO sites, with the net result of only a slight decrease in the total 2-butanol reaction rate or
conversion (Figure 4.31).
The 2-butanol to MEK reaction took place on both the MgO and the copper sites. With no
Cu on MgO (Figure 4.33), similar amounts of butenes and MEK were formed at all
temperatures. MEK formation had to be on the MgO sites. When copper was added to the
support, the MEK yield increased to an optimum yield at 16.9 wt % Cu on MgO. For that
catalyst, the total surface area was about a third of the value of the pure MgO catalyst (see
Table 4.1). Further copper addition to the support caused both a decrease in available copper
area and available MgO area and resulted in a decline in MEK yield.
The copper sites have much higher catalytic activity for MEK production than the MgO sites
do. A simple calculation can verify this:
At 390 °C: 0% Cu on MgO 35% MEK yield
16.9% CuonMgO --+ 47% MEK yield
but MgO surface area is about 1/3 of pure MgO
thus 12% MEK yield from MgO sites
35% MEK yield from Cu sites.
10
(fJ
~ 50
~ 40
~ il
~ ZO
-r ",,,
10
137
Figure 4.33: MEK yield for Cu on MgO catalysts
Figures 4.34 and 4.35 show the effect of flow rate on MEK yield and selectivity for the best
performing catalyst (16.9 wt % Cu on MgO). The selectivity declined with temperature as
more butenes were formed. The longer residence time at the low flow rates (high WIF
values) improved the MEK yield (Figure 4.34), but resulted in a sharp decline in selectivity.
The total2-butanol conversion dropped from 65% to 35% when WIF decreased from 206 to
51 kg. slmol. The higher flow rates meant that insufficient time was allowed for the 2-butanol
to fully react, while the slower rates did. This is also clear in Figure 4.34 where the MEK
yield dropped as WIF decreased.
(fJ
,0
~40
~'jO
l ~o
!- 10
~
Figure 4.34: MEK yield for a 16.9 wt % Figure 4.35: MEK selectivity for a
Cu on MgO catalyst 16.9 wt % Cu on MgO catalyst
138
4.2.2. SILICA CATALYSTS
The silica catalysts that were employed for 2-butanol dehydrogenation were similar to those
used for ethanol dehydrogenation. The operating conditions have been listed in Table 3.6.
Figure 4.36 shows total 2-butanol conversion as a function of temperature and copper
loading. The main product was MEK (Figure 4.37) and the only significant by-products
were a mixture ofbutenes (Figure 4.38). Total 2-butanol conversion increased sharply with
temperature, but started levelling off at about 360°C. For MEK production, there was an
optimum temperature. This temperature was between 300°C and 360 °C for average WIF
values. The optimum temperature did not vary with copper loading. The maximum MEK
yield was very dependent on the copper content on the support. At low copper loadings the
MEK yield was lower due to butene formation (Figure 4.38). As the copper loading
increased, butene formation decreased. Butene formation also increased with temperature.
100
() sO
~ (fJ
!It
if- AO
op
Figure 4.36: Total 2-butanol conversion Figure 4.37: MEK yield for Cu
for Cu on silica catalysts on silica catalysts
Butene formation mainly took place on the silica sites, while MEK formation took place on
the copper sites. The optimum MEK yield occurred at 15 wt % copper on silica. This was in
good agreement with the results obtained for the dehydrogenation of ethanol (see 4.1.3.1)
139
MEK yield data (Figure 4.37) suggested that the maximum copper surface area should be at
15 wt % Cu on silica.
The optimum catalyst for 2-butanol dehydrogenation was the same as for ethanol
dehydrogenation. For a discussion on the characteristics of this catalyst see 4.1.2.4.
\~
"
fIJ
i
~
6<1
aO
1jI: ].0
1 .."p
Figures 4.39 and 4.40 show the MEK yield and selectivity for the best performing catalyst
(15 wt % Cu on silica) as a function of temperature and 2-butanol feed flow rate. As
mentioned previously the yield increased with temperature. The optimum yield shifted to
higher temperatures when the flowrate increased (yVIF decreased). For example:
• WIF = 206 kg.s/mo\: 93-91% yield at 270-300 °C,
• WIF = 103 kg.slmol: 93-92% yield at 300-330 °e, and
• WIF = 51 kg.slmol: 89-87% yield at 330-360 °C.
The selectivity towards MEK production (Figure 4.40) decreased with an increase in
temperature and decreased very marginally with an increase in WIF. Longer residence times
(slower feed rates) stimulated butene formation. At 390°C the selectivity varied between
83% and 86% for the different feed flow rates.
140
Figure 4.39: MEK yield for a 15 wt % Figure 4.40: MEK selectivity for a 15 wt %
eu on silica catalyst eu on silica catalyst
100 100
"
80
!!iJ1!5G-300
mlQ-ons;
13.",,"Cu
~70
.~ eo
.300-850
mlc:rone; .,.,.".
mlerOM;
1!5.O%CU
!!50 15.O%Cu
!'I6lIO-1180
~40 milJ"DnI;
13.9'!(.CII
rl85tJ..1180
mierCII.;
830 13.9'\I.Cu
20 .'''''''
mia>ons: .'3000
microns;
13.~Cu
13.7" C..
10
240 270 300 330 360 390 2-40 710 300 330 380 390
Reaction T ("C) Reaction T (OC)
Figure 4.41: Effect of catalyst particle size Figure 4.42: Effect of catalyst particle size
on total 2-butanol conversion on MEK yield percentage
141
The MEK yields (Figure 4.42) were the highest for the 300-S50 flll1 fraction. Smaller
particles (150-300 11m) and particles ranging from 850-IISO 11m gave high butene yields,
which decreased the MEK yield.
4.3. SUMMARY
Copper-impregnated silica catalysts gave very high acetaldehyde yields. The selectivity
towards acetaldehyde formation was high (above 7S% in the worst case) for temperatures up
to 440°C. Characterisation of the catalysts indicated that impregnation leads to the
formation of a very non-homogeneous catalyst. The copper surface areas for catalysts
prepared by impregnation were lower than for catalysts prepared by precipitation or urea
hydrolysis. For copper on silica catalysts, the addition ofNi, Co or Cr had a negative effect
on acetaldehyde yield. In contrast with this observation, additives to copper on alumina
catalysts had a positive effect on di-ethyl ether formation. An increase in particle size of
silica-based catalysts caused a decrease in total ethanol conversion. This indicated that mass
transfer resistance was important for the reaction of ethanol over Cu-silica. The acetaldehyde
yield, however, varied little at 320°C and above for catalyst particles up to 11S0 11m. The
mass transfer resistance was more important for other ethanol reactions (for example
dehydration and recombination), than for dehydrogenation.
142
The best dehydrogenation catalyst for ethanol dehydrogenation was a 15 wt % Cu on silica
catalyst. Figure 4.43 compares the equilibrium ethanol conversion for the dehydrogenation
reaction with results obtained at different ethanol feed flow rates. For the equilibrium curve
the ethanol partial pressure was taken as 0.2 bar, since the feed (1 bar pressure) consisted of
nitrogen and ethanol in a molar ratio of 4 to J (see section 3.1.4.1). From 320°C to 400 °C
the acetaldehyde yields were close to the theoretical values for WIF = 131 and 66 kg. slmo!.
The difference between the theoretical equilibrium value and the measured values was due to
either incomplete conversion of ethanol (for high flow rates, WIF = 33 kg.slmo!) or a drop in
acetaldehyde selectivity below 100% (for low flow rates, WIF = 349 kg.slmol).
! 20 ~'"'-------.,....."-------
10t----~~·~~·--------
o ~.--.--.- ..
240 280 320 360 440
Reaction T (OC)
Figure 4.43: Equilibrium ethanol conversion vs. measured values for a 15 wt % Cu on silica
catalyst (Pdhanol = 0.2 bar for equilibrium curve)
143
... --:: .......... .:.,··,······-···a._ .
•
I
, .'
, _equi~bfium
For silica catalysts there were an optimum copper concentration on the support (15 wt %),
which gave the highest MEK yields. Catalyst particles in the range of 300-850 !l11l gave the
highest MEK yields. Smaller or larger particles produced increasing amounts of butenes.
For a 15 wt % Cu on silica catalyst the selectivity towards MEK production was close to
100% at 240°C, but declined to between 83% and 86% at 390 °C for various WfF ratios.
Figure 4.44 compares the equilibrium conversion (theoretical) of 2-butanol to MEK, at a
partial pressure of 0.2 bar, with experimental values. For WfF = 51 kg.slmol, the MEK yield
differed considerably from the equilibrium value below 330°C, while for
WfF = 103 kg.s/mol the deviation was below 270°C. Above these temperatures the
experimental values approached the equilibrium values, but were lower due to less than 100%
selectivity towards acetaldehyde production. The reaction conversion will be similar to the
equilibrium value if the reaction goes to completion at a specific temperature. If the flow rate
is too high, however, the reaction will not go to completion, since the reaction time will be
too short.
144
5. REACTION KINETICS
The conditions required for catalyst testing were described in section 3.2. The reactions must
be performed in the steady state regime. This implies that there should be little or no
catalysts deactivation during testing, hence the catalyst deactivation conditions must be
determined. The reactor must be isothermal and the catalyst particles free from interphase
and intra-particle mass transfer resistance. The reactor must be operated as a differential
reactor. For a plug flow reactor, the relationship between the reaction rate and the conversion
is given by the following equation.
w
F
= r~
-rA
(5.1)
Under differential conditions the conversion is kept low (typically below 10%). Equation
5.1 can then be simplified to equation 5.2.
w= X (5.2)
The difficulty with the integration term is removed, while still maintaining very high
accuracy. The following sequence was performed to determine the reaction rate kinetics of
both ethanol and 2-butanol dehydrogenation.
1. Determine the temperature range which results in an acceptably low deactivation rate.
2. Determine the flow rate which results in low conversion (below 10%) and which
minimises interphase mass transfer resistance.
3. Determine reaction rate kinetics at the above temperatures and flow rates.
Small catalyst particles in the range of 300 to 425 !lm were prepared for all experiments to
minimise intra-particle mass transfer resistance. The mass of catalyst for testing was varied
according to the reaction temperature. At high temperatures, a smaller amount was used to
145
reduce the total conversion and keep it below 10%. At low temperatures, the catalyst mass
was increased to ensure a detectable conversion.
For all the data presented in Chapter 5, the following units were used for the reaction rate
parameters:
k': mol/kg.cat.h.kPa
K.q : kPa'l
Ki: kPa'l
r: mollkg.cat.h
Initial kinetic model fitting was performed with the above units and then the final step was to
convert the parameters back to the standard units. The standard units were used for process
modelling in Chapters 8 and 9.
Reduction at 340 DC and 400 DC resulted in low initial reaction rates. Since the initial
reaction rates were much lower than the initial reaction rates for catalyst reduced at 255 DC, it
was concluded that reduction at 340 DC and 400 DC caused changes in the copper particle
size. Figure 5.1 compares the curves for reduction at 400°C and 255 DC. The initial activity
of the catalyst reduced at the higher temperature was similar to the activity of the catalyst
146
reduced at the lower temperature after about 12 hours reaction. This was considerably longer
than the reduction time (2 hours) of the catalyst reduced at the higher temperature.
100 70
';:C 70 ..;-...... .~
~ 1;1 60 +-------'."'"''-''",---------
~ ~ 50 +------',::,'''''.~------ ..
~o ~~ 40 11,C-
•• -----'-,.-',,"'-"'=------
f\ '.
--""~
It - 30 "
'\
•_ _ \
20+-~.~~.-------------2-
~
10+----~----------
o+---~--~-~--~-~
o 5 10
Time on stream (h)
15 20 25
10
,
, ----
10 15
n,. on stream (h)
20 30
The reducing temperature, the reducing environment and the effect of water formation from
the reduction reaction all have an effect on the change of copper particles during reduction.
Although there was initially no coking, it did occur over time. From Figures 5.1 and 5.2 it is
clear that sintering of copper occurred at reduction temperatures of 340°C and 400 DC. To
avoid the loss in catalyst activity all further reduction was performed at temperatures between
250°C and 260 DC.
147
90.0 '.'.'.''.
'{. ~ Reaction T =400 ·C
80,0 ~
• Reaction T =340 ~C
70.0 .~
.a .. .. Reac:tion T =280 'C
1!5 60 ·0
c i t _ .......
•
_: ~ • Reaction T =22O'C
-
'~2 i~ 4.0t~:=.~.!.~.~.~~"~;;=;:::~:~
50.0
_____ ...
CL - 30.0 t-:-:-----:--;---------------=::~~.t::;::=-l
20.0 t-1~~~·i·~~~======~======~~~.L4
10.0 •••• ••••• ••
0,0 -I-'---~--_--~--~--___j
.., ..
o 5 10 15 20 25
Time on stream (h)
Figure 5.3: Acetaldehyde production rate as a function of time and temperature (14.4 wt %
Cu on silica)
..
Table 5 1· Summary of catalyst deactivation at different reaction temperatures
Reaction T (0C) Reaction rate (O.5h) Reaction rate (24h) % decline
moIlkg cat. h moIlkg cat. h
400 87.7 22.6 74
340 54.5 37,0 32
280 23.3 20.5 12
220 7.0 6.3 10
At 220°C and 280 °C, the catalyst remained stable throughout the 24 hour testing period, In
absolute terms there was very little change in reaction rate.
Chromium and cobalt were added to copper to try and improve catalyst stability. Stability
tests were repeated with a 13.5 wt % Cu, 1.0 wt % Cr and 0.5 wt % Co on silica catalyst. The
results are shown in Figures 5.4 and 5.5. Similar stabilities to that of a pure copper catalyst
were achieved. The catalysts with additives did, however, show poorer activity towards the
dehydrogenation of ethanol. At each temperature tested, the acetaldehyde production rate
was lower over time, on average, as indicated in Table 5.2. The CuiCr/Co on silica catalyst
was also less selective towards acetaldehyde production. At 400°C, the main by-products
148
were ethylene, MEK and butyraldehyde (similarly at 340°C and 280 0c). Selectivities are
indicated in Table 5.2.
100 30
90 .'. • 14.4 'JI.CuCI'llllca, r_400"C
80
~
.13.5%01, 1%Cr, O.5%Coon wilk:II; T" 4OO"C
..1<t4%CUontU.... T=340OC f- .,..,.. .
~~~~-.----------------~
.s - 70
! ~ 80 !~. . .13.5%01,1% Cr, O.5%Coon liIica; T" 3IIO"C
i~" I,".~ ••
.
'>..
..
'....... . .
gu
il ,. 50 :..... .:~- ..........
gu I_~'~::~==========~~~
11 ~0 -'"
.t.e. 30
1'(
.
-......... :---- ,..
.. il~15+-
• 0
.
£.510 r-~=.=;::~=;:::;:;===;-----j
----....
-- I I
20 !d4.4"AoCUonsllC<l,T_280·C
10 5
o~ r-l.
13.S%Cu, '-.Cr,o.ncoondlca;r_280'C/!----i
______________________ ~
o 4 , 12 16 20 24
o 8 12 16 ,.
TIme on stream (h) Time on stram (h) "
Figure 5.4: Effect ofadditives (Cr and Co) Figure 5.5: Effect of additives (Cr and Co)
on acetaldehyde production rate on acetaldehyde production rate at 280°C
..
Table 5 2· Performance ofCU/Cr/Co on silica catalyst
Temperature (OC) Average decline in reaction Selectivity (%) towards
rate over time (%) acetaldehyde
400 22 70-80
340 35 > 88
280 23 70-90
149
..
Table 5 3· XRD and Toe results of tested catalysts
Temperature Peak width from Crystallite size Carbonwt %
(OC) XRD (degrees) (nm)
Reduced, but 0.20° 60 0.065
unused
220 0.15° 80 3.213
280 0.15° 80 3.386
340 0.14° 80 3.693
400 0.15° 80 3.737
150
f"
Figure 5.6: TEM images of 14.4 wt % copper on silica catalyst after being in use at 400°C
TEM studies showed a very wide size distribution of copper particles in the utiused catalyst.
Some areas had copper agglomerates in the 1 micron range, down to 40 to 50 nm. Other
areas had small copper crystallites of 4 to 5 nm, evenly distributed over the support (see
Figures 4.30a to 4.30i). After reaction at 400°C, there was definite agglomeration of small
crystals (4 to 5nm) to yield bigger polycrystalline particles (Figures 5.6a to 5.6i).
Some areas had no copper (Figure 5.6a), similar to observations made of the catalyst before
reaction. The small crystallites, in the order of 4 to 5 nm (Figures 4.30b to 4.30d),
agglomerated to form particles in the 20 nm range (Figures 5.6b to 5.6d). Larger
agglomerates were more frequently observed after reaction than before reaction (see
Figures 5.6f, 5.6h and 5.6i). This confirmed that some sintering occurred during reaction at
400°C.
Chemisorption results are shown in Table 5.4. In each case the surface area was the value
after 24 hours at the reaction temperature. The values indicated a decline in copper surface
area with an increase in reaction temperature, confirming TEM results. Sintering occurred
even at the low temperatures.
151
Table 5.4'. Copper surface areas for 144 wt % Cu on silica catalysts
Reaction T (DC) Copper area Copper area % Decrease
(m1/g sample) (m1/g Cu)
Catalyst reactivation studies were performed to obtain more information regarding the
deactivation mechanism. The 14.4 wt % Cu on silica catalyst was reactivated after 24 hours
and 48 hours of operation at 400°C. This was done to determine whether coking, sintering,
or both, were the dominating deactivation mechanism. Figure 5.7 indicates a decline in
initial acetaldehyde production rate (about 25% to 30"/0) after the first re-oxidation step,
compared to the first run. This is in good agreement with results in Table 5.4. After the
second re-oxidation step, the same initial activity as after the first re-oxidation step was
obtained. From this experiment it could be concluded that some irreversible sintering
occurred during the first 24 hour reaction period. Deactivation that occurred in the second 24
hour period could be recovered by oxidation and hydrogen reduction.
120
i
• fresh catalyst
100 \ • after 1st oxidation ~
.\
,
... after 2nd oxidation
!
\
"". •
~
!
;
;
20
~• •
• ,,
o
,;
o 12 16 20 24
nnw on stream (h)
152
5.1.3.1. Deactivation mechanism
The following mechanism for catalyst deactivation is proposed: At all temperatures above'
220°C there is some initial sintering within the first 24 hours of use and the sintering rate
increases with temperature. At 400°C, the reduction in copper surface area after 24 hours of
use is twice the amount of the catalyst used at 220 °C for 24 hours. Results of re-oxidation
experiments indicate that most of the sintering occurs within the first 24 hours of use.
Sintering and carbon deposition on the catalyst takes place simultaneously. At the higher
temperatures, mainly coke-like carbon is deposited, which reduces catalyst activity and
prevents ethanol molecules from reaching the active copper sites. At the lower temperatures
(280°C and lower) carbon is present in molecular form, probably as oligomers and polymers
that do not prevent ethanol from reaching the active sites.
Coking contributes more to catalyst deactivation than sintering. Figure 5.7 indicates a
decrease in acetaldehyde production rate of 80% for the second 24 hour run at 400°C after
the first re-oxidation step. This value is considerably higher than the decrease in
acetaldehyde production due to sintering.
153
.
160
..
~-
140
120
.~
~~
\
- .T=310·C
_T=310·C
J.T=250"C
r---
I--
c 5 100 ..
~::s :1at
~:l:0
,..s
e.
80
60
•
, .
•
~
W 40
:IE
20
o
o 5 10 15 20 25
Time on stream (hI
Figure 5.8: MEK production rate as a function of time for a 14.4 wt % Cu on silica catalyst
There was significant interphase mass transfer resistance for ethanol dehydrogenation at the
higher temperatures. At higher temperatures, the difference between the reaction rate
limitations and the mass transfer limitations increased, because the reaction rate became
faster (less resistance). A small increase in the feed flow rate (which determines interphase
154
mass transfer resistance), would thus result in a larger increase in acetaldehyde production
rate at high temperatures as compared to lower temperatures.
350
.T-200"C
300 r-- -T-225°C ,
,.
_T_250·C
,•
.
I--
-T-275"C
I ~
I-- "T-300·C
,
,
.t---~
.. . •
··V • •
50
V: // ...
!-_.-- -
:
--'.
•
~""t-~~~",=-.- . .;:::-::r
o
o 5 10 15 20 25 30
Feed flow rate (mllll)
Figure 5.9: The effect of feed flow rate on acetaldehyde production rate
At temperatures of250 °C and below, the reaction rate was fairly constant above a feed flow
rate of 14 mVh. At 275°C and 300 °C it appeared, from the data, that the region free of mass
transfer resistance commenced at a feed flow rate of30 mVh. Due to limitations of the feed
pump, it was not possible to do full kinetic studies at such high flow rates. All data for
determining kinetic parameters was obtained at a feed rate of 14 mllh. In the modelling
calculations the final rate data at the higher temperatures was to be fitted to the curves in
Figure 5.9, to compensate for rate reduction from mass transfer resistance.
155
dehydrogenation of n-butanol over zinc oxide. They concluded that bulk diffusion was not
significant.
250
,
~
200 •
-------- .. ..-- ..• --=1
;,
--~-.-."
•
.'H'" • !
"- ,,I +19O"C
.. ..J 150
~
,,
Ii
•
"
.. !
,
.220 o e
A250"C
100
'U
... -
i
<lIIo 280 "C
50
• •
• • .. •
i
!
0
0 2 , 6 8 10 12 l' 16
Feed now rate (mllh)
Figure 5.10: The effect of2-butanol feed flow rate on the MEK production rate
(5.3)
Numerical values for the rate and adsorption coefficients could be calculated using least
square analysis. Equation (5.3) must be linearised in the unknown coefficients. Rearranging
equation (5.3) yielded:
156
,<
or (5.4)
y= and (5.6)
(5.7)
K -~a
A - (5.9)
The constants a and b were determined first (using a pure ethanol feed at different pressures)
from the y-axis intercept and the gradient of the best linear fit to the data. The ethanol partial
pressure was varied by feeding it with an inert gas (nitrogen). Thereafter the ethanol was fed
with hydrogen and acetaldehyde separately, while keeping the total pressure constant at one
atmosphere. Different molar feed ratios were used. Equation (5.5) can be expressed as:
where d and c are used depending on whether acetaldehyde or hydrogen is used. From the y-
axis intercept and the gradient of the best linear fit at each temperature the remaining
constants could be determined (in 5.7). The process was repeated at each temperature, to
determine the Arrhenius dependence of each adsorption constant. The Arrhenius equation is:
157
(5.11)
Figure 5.11: Linear fits of reaction rate data Figure 5.12: Linear fits of reaction rate
at I atm. total pressure and 225°C data at 1 atm. total pressure and 300 °C
The y-value was proportional to the inverse square root of the reaction rate (equation 5.6). A
decrease in the reaction rate would thus cause an increase in the y-values. For the ethanol-
acetaldehyde feed there was a sharp increase in the y-values (decrease in reaction rate) with
an increase in acetaldehyde concentration. This indicated that the reaction rate was most
sensitive to the acetaldehyde concentration. Acetaldehyde adsorbed strongly on the catalytic
sites and reduced the dehydrogenation reaction rate. Hydrogen adsorption on the surface was
not very strong and ha only a small effect on the reaction rate.
158
Good linear fits for the acetaldehyde-ethanol feed in Figures 5.11 and 5.12 indicated that the
reaction rate was not dependant on the acetaldehyde pressure, but on the square of the
acetaldehyde pressure. Table 2.5 listed other possible reaction rate mechanisms. Only
equations. (2.18) and (2.19) had the acetaldehyde pressure square term. Furthermore,
equation (2.19) did not have adsorption coefficients for ethanol, hydrogen and acetaldehyde,
but only for either ethanol and hydrogen or ethanol and acetaldehyde. The data confirms that
equation (2.18) is most suitable for describing the rate mechanism.
The value of each parameter at the different temperatures is given in Table 5.5.
The adsorption coefficients (KR and Ks) decreased with an increase in temperature from 200
°e to 275°C. The value of KA should theoretically also decrease and the calculated values
were in reasonable agreement with the theory, excluding the value at 300°C. The reaction
rate coefficient (k') increased with temperature (as expected). This was typical for a reaction
which follows the dual site, surface reaction mechanism (Peloso et aI., 1979). At 300°C, the
adsorption coefficients deviated from the downward trend. The values were higher than at
275 °e. To comply with the dual site, surface reaction controlling mechanism, the adsorption
coefficients must decrease with temperature, and they must be positive (Peloso et aI., 1979).
The increase in the calculated adsorption coefficients at 300 oe, compared to the values at
275°C, is illustrated in Figure 5.13.
159
.'""
.. '
75
65 /
l 55 L
=
~
~
g 45 /
"~35
~
/
-5'
<0;
25 /
15
....- /:
5
220 240 260 280 300 320 340 360 380 400
T ('C)
There was little change in the deactivation rate of the catalyst at temperatures from 200°C to
275 °C (from 10 % to about 11.2 % over a 24 hour period). At 300°C, the deactivation rate
dropped to 16.5 % after 24 hours of operation. During the testing period at 300 °C (about
two days), some deactivation would thus have taken place. This would result in the reaction
rate values being lower than the true values, due to catalyst deactivation. These lower
'falsified' reaction rate values are reflected in the rate mechanism as higher adsorption
coefficients, Stronger adsorption or higher adsorption coefficients lead to slower rates,
Nonetheless, the model still holds at higher temperatures, even though the changes in catalyst
activity gave rise to conflicting parameter values.
The Arrhenius parameters for k', KA, KR and Ks were determined only at temperatures from
200°C to 275 °C (thus excluding values at 300°C) using equation (5.12). AU adsorption
1
coefficients (KA, KR and Ks) are given in kPa- • The foUowing expressions were obtained
(with Tin K):
160
The data is shown in Figure 5.14.
,,---------------------------------,
•
l-, __========~======~y;·~~~1.'~x+~"~.~:'~~~
O.
o
18 0.0019
•
0.002
R1 • 0.8253
0.002
Keq is a constant and could be found in the literature as an empirical correlation (peloso et al.,
1979) or calculated using firsts principles via the Van't Hoff's equation (Fogler, 1992).
Many of the correlations were very old and their accuracy doubtful, with the result that an
expression was reduced from first principles.
AG=AH-TAS (5.18)
K =ex
p p
[-AG~(T)]
RT (5.19)
o
Kp in equation (5.19) is defined in atmospheres. Using thermodynamic data for ethanol and
acetaldehyde, Kp (in atmospheres) could be calculated at different temperatures. Kp values
were changed to K.q values in kPa (to standardise all K-values) and then the Arrhenius
equation could be fitted through the different Keq-values to obtain the necessary Arrhenius
parameters (similar to 5.11 and 5.12). The equilibrium constant is then:
161
Ln(Keq) = 19.014 - 8503.3/T (5.20)
The reaction rate from 473 K to 548 K and at 1 atmosphere feed pressure could now be
expressed as:
. _ 8.014xlO xe
4 -S49I. lf ( P -P P /(1.810X10 xe
IJ -8503.~)
)
A R S
- r _ - - -_ _ _ _ _ _ _----'--.:..:._-'..:--"--_ _ _ _ _ _.L-_ _ _--::- (5.21)
A (
1+3.557xlO·'xe
13Sl.){
xP +3.941xlO·'xe
3586.%
xP +9.683 x lO·'xe
228~ xP ) 2
A S R
with pressures in kPa and temperatures in K. The accuracy of equation (5.21) was tested for
all the data from 200 °C to 275 °C (Figures 5.15 and 5.17).
The structure of the error function was of great importance. Either the model percentage
deviation could be used or the actual model value difference. In mathematical format:
• 100*(model value - measured value)/measured value or
• model value - measured value.
The reaction rate increased from 33 (mol/kg cat.h) at 200°C to 227 (mol/kg cat.h) at 275°C.
At high temperatures, a percentage deviation of 20 %, for example, would result in a large
difference in measured reaction rates, while the difference in measured reaction rates at low
temperatures and at the same percentage deviation, would be small. It was more important to
correctly predict reaction rates at the higher temperatures than at the lower temperatures, thus
the actual deviation and not the percentage deviation was used in the error function. Since
the deviations in actual rates at the higher temperatures were so much larger than at the lower
temperatures, different weights were applied at each temperature to the error function. The
error function is as follows:
162
!i'
EF = ~(CV
L.. _ MV)'.
T_472K +
(CV - MV»)' + (CV - MV»)'
1 1.125 T.498K 1.25 T.S2JK
with EF = error function, CV = calculated values, MV = measured values and n the number
of experimental points.
The model deviation (MD) from the measured data is simply the following:
There were 62 experimental reaction rate values. The values in equation (5.21) were used as
the starting values for further optimisation. The parameters were varied using an 8
dimensional matrix (varying each one of the eight parameters in equations 5.13 to 5.16). The
smaller the error function, the closer the calculated values were to the measured values and
the better the fit. As the error function approached the minimum value, the step size for each
parameter was decreased to obtain more accurate values. The final step size employed was
0.2 % steps for each of the 8 parameters. A Turbo Pascal program was compiled to
determine the error function and calculate the parameter values. The best fit was called
model 2. Both model I and model 2 were only for data obtained from 200 °c to 275 °c. The
parameter values are listed in Table 5.6 and the models are compared in Figures 5.15 to
5.20.
There was a significant improvement in the model deviation and correlation coefficient
(R 2-value) of model 2 compared to model 1. When comparing Figures 5.15 and 5.16, it
could be seen that model 2 gave a much better prediction of values at 548 K (275 0c) than
model 1. For model I, the model values at 548 K were much lower than the measured
values. At 523 K, both models gave similar results. At 498 K, model 2 was slightly better,
while at 473 K (200 0c) model I gave better predictions that model 2. The error function
163
was structured in such a way as to minimise the difference between the model values and the
measured values. The main reason for the improvement obtained by model 2 over model I,
was due to better fitting of measured data at the higher temperatures.
As 9.683*10-" 6.1854*10-
MD 8606.49 4914.21
R> 0.928 0.966
..
A IS In [mollkg cat.h.kPa]; -EfRo is in K
30
•
30
· -----.
.!
20
10
. i l!
"
10
·•
I -----!"
1i
~
0
I * ~
1i 0 I I
·30
•; .."'& ...,
~o
.
~ •
-40
• '" ~o _._---
•
Temperature (K)
... Temperature (K)
Figure 5.15: Difference between model 1 Figure 5.16: Difference between model 2
rates and actual rates at different rates and actual rates at different
temperatures temperatures
164
In Figures 5.17 and 5.18, the percentage deviation between model values and actual values
are shown. For model 2 there are a few points at 498 K and above with very large percentage
deviations (greater than 50%). Those are reaction rates at very high acetaldehyde
concentrations. The reactions rates at high acetaldehyde concentrations are very low and a
small difference between the predicted and measured values cause huge percentage
deviations.
OJ 250
....
:; · :• ...,°41-0·'--""'---490---l~f.-OO--5-"---52-0~-53-0
"
540"10
Figure 5.17: Percentage deviation between Figure 5.18: Percentage deviation between
model 1 rates and measured rates model 2 rates and measured rates
Figures 5.19 and 5.20 compare the measured reaction rates with the modelled reaction rates.
The solid diagonal line on both graphs indicates a perfect fit.
250
""
200 /. 200 /.
... -
..
~t
'8 ~
;;-
';;;:'" 150
~
100
/....
/. •
•
.h
!
N
.~
5'~
•
150
'00
/..
/" ,
- - ' - ..
'.
•
"- .i:/ • . ? ---
" • 5O~ ....
0 1/ . 0
0 '00 'so 200 250
0 50 '00 '50
Mea.ured rate (mollkg.cat.h)
200
"" 50
Measured rate (mollkg.cat.h)
Figure 5.19: Modell rates vs. actual Figure 5.20: Model 2 rates vs. actual
rates at different rate values rates at different rate values
165
I'
Figure 5.19 indicates that model I fits data well when the reaction rate is below about SO
(mol/kg cat.h), which is the reaction rate at the lower temperatures. At high reaction rates
(high temperatures), the model values are significantly lower than the measured values. For
model 2, the calculated values show more variance (compared to model I) for reaction rates
below 50 (mol/kg cat. h). For reaction rates above SO (mol/kg cat.h), the model values are,
however, much closer to the measured values than for model 1.
It must be stressed that the models were formulated for data from 200°C to 275 °C (473 K to
548 K). Measured reaction rates at 300°C (573 K) were compared to predictions from both
model I and model 2. The difference between model values and measured values are
summarised in Figure 5.21. For both models, the model values were generally higher than
the measured values (see Figure 5.13 and the discussion thereof). Model I predicted
measured reaction rates at 573 K far better than model 2 did. The correlation coefficients are:
• Modell: R2 = 0.953 and
• Model 2: R2 =0.918.
160
• •
140 - • Model 1
~
~ 120 • Model 2
~
c:
0 100
:;:;
''>" 80
•
"
'tI
60 · •
"
C>
~
'"
c: 40 • -.- • • • •
"I:! 20 • • •
a.."
• • •
• • •
0 • • •
·20
2
• • 6 a 10 1"2 14 16 1
Runs
Figure 5.21: Percentage deviation between model values and measured values at 573 K
The reason why model 1 predicted reaction rates at 300°C better than model 2 did, was due
to the fact that catalyst deactivation occurred at 300°C. The measured reaction rates were
much lower than the true reaction rates at 300 °C. Since model I generally under predicted
166
reaction rates, it thus gave a better prediction of the 'lower' reaction rates at 300 °e that
resulted due to catalyst deactivation.
For equation 2.19 to be applicable, the initial reaction rates for different hydrogen-2-butanol
feeds had to be independent of 2-butanol pressure (Thaller and Thodos, 1960). That was not
true and equation 2.19 could not be used. Equation 2.18 provided a reasonable fit to the data
obtained in this study. The parameters in equation (5.4) were determined at temperatures
from 190 °e to 280°C for the dual site, surface reaction controlling mechanism. The linear
fits to the data, which were used to determine the reaction rate parameters, are given in
Appendix C2. Figures 5.22 and 5.23 show typical linear fits at 220°C and 280 °e,
respectively.
167
The reaction rate was most sensitive to the MEK concentration. The large negative slope
indicates strong MEK adsorption: The adsorption coefficients at the different temperatures
are listed in Table 5.7 .
• , 2-8ut_I.MEK
y .. -e.8llll&02lo.+1.126EotOD 3
,.., RI.9.844&01
y· .... l11E-02Jc +4.71BE..oQ
R' .. I.MlS-Ot
2 y" 2.21SS03I<+ 1.138E+OO
rf .. 5.747&01 y .. , .7ME-03k + 6A18E-01
2-aut.wN. NIIfogIwI R' .. 8.7406-01
'I .. 3.574&0310. + 11._6-01
P;'-7.9IIEI&Ol
~ • • • '" ro '"
2-Butanol partial pressure (kPa)
'" .00 •• o 20
• eo eo
24*1nol partial pressure (kPa)
". ".
Figure 5.22: Linear fits of reaction rate data Figure S.23: Linear fits of reaction rate
at 1 atm. total pressure and 220°C data at I atm. total pressure and 280 °C
The trends in the k' and KR values were in line with the theory (peloso et al., 1979), except at
280°C where KR showed an increase instead of a decrease. From initial experiments, the
same conclusion could be drawn for this reaction as for the ethanol dehydrogenation reaction
(see 5.3_1). Very little catalyst deactivation took place at 250°C and below (see Figure 5.8),
but at 310°C (Figure 5.8) significant deactivation took place. Both sintering and coking
played an important role in catalyst deactivation at the higher temperatures. The observed
168
reaction rate would be lower due to changes in catalyst activity and not due to stronger MEK
adsorption.
Both KA and Ks (the adsorption coefficients for 2-butanol and hydrogen) were negligible
compared to the adsorption coefficient of MEK (KRJ When adsorption took place, the
reaction rate slowed down. This was because diffusion resistance of the feed molecules to
the active sites increased. Negative adsorption coefficients indicated an increase in reaction
rates. The negative hydrogen adsorption coefficients (Ks-values) contradicted the theory of
the dual site, surface reaction controlling mechanism, but the values were so small that the
equation still fitted the data very well.
The reasons for the increase in reaction rate with hydrogen in the feed have been documented
for other dehydrogenation reactions and was not unexpected where coking tended to
deactivate the catalysts. Sheintuch and Dessau (1996) cited many references where hydrogen
was co-fed with either an alcohol or an alkane and where improved dehydrogenation activity
was reported. Hydrogen in the feedstream reduced coking (Sheintuch and Dessau, 1996) and
it reduces the partial pressure of the alkane or the alcohol, which is favourable for higher
conversions (Ertl et aI., 1997).
KA and Ks were taken as zero for a first approximation. The following Arrhenius expressions
were formulated for data obtained at temperatures from 190°C to 250 °C (463 K to 523 K)
by using equation (5.12), (temperature is in K). Figure 5.24 represents the data graphically.
The equilibrium constant for 2-butanol dehydrogenation could be expressed by the following
equation (Kolb and Burwell, 1945):
169
2
1 ...
~ y - -6903, + 13.628
R2 =0.9441
0
>-
0.0 17 0.0018 0.0019
~21 0.0 22
·1
--.....:
·2.
Y- 3298, . 9 . 3 3 7 7 _ _ . _ _ _ _ _
R2 .. 0.9047
·3 •
-4
1fT (11K)
Kp, in atmospheres, was solved as a function of temperature and changed to ~q in kPa. The
function was then expressed in the exponential form (see equation 5.11) and substituted into
the rate equation (5.3) to yield equation (5.26).
The reaction equation from 190°C (463 K) to 250°C (523 K) could now be expressed as:
(5.26)
170
EF = I(cv - MV)~=463K + (CV - MV)]' +(CV - MV)]' (5.27)
I 1.125 T=493K 1.25 T=",K
The different model parameters are listed in Table 5.8, where model 1 is the values
determined with multiple linear regression and model 2 is the optimised values. Both models
were only valid from 190°C to 250 °C and at a 2-butanol feed pressure of 1 atmosphere.
MD 12724 3162
R- 0.9478 0.9430
..
A IS III [moVkg cat.h.kPa], and
-EIRQ is in K
The most important indicator of model performance, relative to measured data was the model
deviation (MD value). The lower the deviation, the better the model. The optimised values
of model 2, compare to modell, gave a more accurate prediction of reaction rates. The
exclusion of adsorption coefficients for 2-butanol and hydrogen in the reaction rate equation
was an acceptable simplification. Predictions remained accurate without these parameters in
the rate equation. Figure 5.25 compares the performances of model 2 and model 1. Model 2
is more accurate at the higher reaction rate values.
171
160 ...........-- -- -"-.-... --.. -.---. -_........ _................... ----_........ -_•.....•......•.....
•
'40 • •
:c
oJ
120
3'00
~
'" 80
g
Ii 60
~ 40
20
o ~ 40 80 80 '00 ,~ ~ '50
Measured value. (mollkg c:ath)
Figure 5.25: Comparison between measured reaction rates and model reaction rates
5.4. SUMMARY
Catalysts reduced at 340°C and 400 °C showed a sharp decline in activity after reduction
compared to catalysts reduced at 255°C, mainly due to sintering. At 280°C and below, a
14.4 wt % copper on silica catalyst remained stable over a 24 hour period. At
340°C and higher, the catalyst deactivated by more than 30"10 over a 24 hour period. The
addition ofCr and Co to Cu did not improve the stability of the 14.4 wt % copper on silica
catalyst. The activity of the CulCr/Co on silica catalyst was similar to that of pure copper on
silica, but the selectivity towards acetaldehyde production was significantly lower at all
temperatures tested. Further deactivation testing indicated that during the first 24 hour
period, both sintering and coking occurred during the dehydrogenation of ethanol.
Thereafter, coking was the main deactivation mechanism.
For the dehydrogenation of ethanol, there was strong interphase mass transfer resistance,
while for 2-butanol dehydrogenation there was no clear indications of interphase mass
transfer resistance.
The controlling reaction mechanisms and reaction rate parameters for both the
dehydrogenation of ethanol and 2-butanol were determined. Ethanol dehydrogenation was
studied from 200°C to 300 °c and 2-butanol dehydrogenation from 190 ° to 280°C. Both
172
reactions could be well described by the dual site, surface reaction controlling mechanism.
The reaction rate coefficients and the adsoption parameters for each reaction were
determined. In both reactions the organic product (either acetaldehyde or MEK) had a
dominant adsorption coefficient.
173
6. Pd MEMBRANE PREPARATION AND
CHARACTERISATION
This chapter will present results on electroless Pd plating. Comments will be made on the
various plating steps and the composition of the plating solution. Membrane characterisation
results will be presented and discussed. This will include surface characterisation with SEM
and permeance testing with both hydrogen and nitrogen. For permeance testing, a positive
feed pressure or a sweep gas was employed alternatively.
Permeance results were compared to literature data and the effect of film thickness on
permeance parameters will be discussed. Arrhenius parameters were determined by
performing experiments at different temperatures.
4: Pd rmallayer
174
Figure 6.1: Cross section view of a three layer SCT membrane
In the cross section view there were the three, clearly visible, layers (Figure 6.1). Figures
6.2 and 6.3 show that the surface had a smooth structure, with plenty of pores. The latter
were very suitable for electroless plating.
Figure 6.2: Top view (20 OOOx) of a Figure 6.3: Top view (S OOOx) of a
three layer SCT membrane three layer SCT membrane
175
Shu et aI., 1993) deposited a small amount of tin. Keuler (1997a) found, by PIXE analysis,
between 0 and I % tin in Pd films of 5 microns. Such low tin percentages would increase by
about 3 times when the film thickness is reduced to 1.5 microns, since the pretreatment
procedure remained the same. The result is that in this study a tin to palladium molar ratio
that was about eight times less than the conventional ratio was employed. This was done by
both increasing the palladium concentration and reducing the tin concentration in the
pretreatment procedure.
The tin in the film should not have a significant effect on the hydrogen permeance
parameters. The main reason for reducing the tin was to try and improve its high temperature
membrane stability. In the time of this study, another research group had similar ideas to
improve the membrane stability. Paglieri et al. (1999) started experimenting with
pretreatment procedures without tin. They speculated that tin, with its low melting point (505
K), could enhance metallic diffusion at the grain boundaries and lead to an increase in
defects. They concluded that the presence of tin at the alumina-palladium interface
contributed to selectivity decline.
Keuler (l997a) used 27.5 g of 10 wt % solution (NH3)J'd(N03)2 per litre of plating solution
for plating the outside of membrane tubes (selective layer on the outside). In the present
study, 4.96 g of (NH3)4PdCh.H20 per litre of solution was used for plating the inside surface
of membrane tubes (selective layer on the inside). When plating the inside of tubes, a much
higher Pd concentration in the plating solution can be used as compared to plating on the
outside. Initially the hydrazine:Pd molar ratio was I to I at the start for plating the inside of
176
tubes. Later it was reduced to the value given in Table 6.1 to slow down the plating rate. In
no experiment did decomposition of the plating solution occur while plating the inside of the
membrane tube, even though a higher Pd concentration was used, in addition to a much
higher hydrazine concentration, initially (hydrazine:Pd of I: 1). The reasons are that the
catalysed membrane surface is the only one available for plating and the volume to available
plating area ratio is much smaller for tubes on the inside. When plating on the outside of the
membrane, plating solution is also in contact with the plating reactor, thus increasing the
available area for Pd deposition.
Table 6.1: Composition (per litre) of plating solutions for membrane plating
Coml!onents Outside of tube Inside of tube
Pd (g) 1.47 2.00
Ammonia (28 wt %) (ml) 200 400
EDTA(g) 100 80
Buffer pH - 10 (ml) 100 -
Hydrazine:Pd molar ratio about 0.7 0.3 5 at start
Temperature (0C) 72 72
The second solution feeding method tested was continuous pumping of solution through the
tube fixed in the reactor and placed in the water bath. A flow rate of 120 mVh was used and
the same solution was pumped through the tube several times, increasing the hydrazine
concentration after each run. The sharp decline in Pd concentration with time, made this
method unsuitable. At a flow rate of 120 mVh, it took the solution 5 minutes to pass from
end to end in the membrane tube. During the first run the Pd concentration dropped by about
30% during that time, resulting in a film of non-uniform thickness over the length of the
177
membrane. The film was thicker at the entry point than at the exit point, because of the
declining Pd concentration and decreasing reaction rate. For this method to be successful ,
much higher feed flow rates have to be used and the feed direction reversed every few
minutes. It was decided to use a batch process to produce coatings of even thickness along
the full length of the tube.
Figure 6.4 indicates the decrease in Pd concentration in solution as a function of the number
of plating session. Values were obtained using rep analysis. Hydrazine was added after each
plating session. The initial hydrazine:Pd molar ratio was I to 1. Half the initial volume of
hydrazine was added after 10 minutes of plating. After a further 15 minutes of plating
(25 minutes total plating), the original volume ofhydrazine was added and then after a further
20 minutes (45 minutes total plating), a few drops of35 wt % hydrazine was added. Reaction
continued for an additional 30 minutes (75 minutes in total). The results are presented in
Figure 6.4.
2000
1800
1600
1400
1200
"E
Il.
1000
Co
c- eoo
eoo
80 to 91%
400
200
• 95 to 97% 98 to 99.5% 99 to 99.5% 99.5%
0
• :>
0 234 5 6
Plating session
The amount ofPd deposited after the first 10 minute session varied between 80 and 91 % of
the available amount in the 11.5 ml solution. Four different membranes were tested this way.
After 4 plating sessions, 99% of the available Pd was extracted from solution and deposited
on the membrane. Poor plating was observed in two cases and the hydrazine concentration
reduced to the values listed in Table 3.14. The plating procedure was changed to that
178
tabulated in Table 3.15. After the first 20 minute session, between 50 and 60% of the Pd was
deposited on the membrane. After three 20 minute sessions, more than 98% of the palladium
was deposited.
Figure 6.5: Cross section of Figure 6.6: Top view of membrane (a)
membrane (a) (10 OOOX) (25000x)
179
Membranes (a) and (b) in Figures 6.5 to 6.11, had two different surface structures and in
both cases there were clear defects. The cross section view of membrane (a) (Figure 6.5)
indicates poor adhesion of the metal film to the alumina support. The Pd film is the thin layer
on top of the alumina base. The dense layer higher up is part of the resin. The surface is not
smooth (Figure 6.6), but consists of small metal clusters scattered over the surface area.
Under high magnification (Figure 6.7), there are tiny pores visible in the metal particles. It is
those pores or defects that caused poor selectivity.
Figure 6.8: Cross section of membrane Figure 6.9: Top view of membrane (b)
(b) (10 OOOx) (25000x)
180
Figure 6.10: Top view of membrane (b) Figure 6.11: Top view of membrane (b)
(SOOOx) (2S000x)
The surface of membrane (b) seems dense both on the cross section view (Figure 6.8) and on
the top view (Figure 6.9) images. There were no continuous defects in the structure. Upon
further investigation some areas in the coating were identified where the coating was clearly
porous (Figure 6.10). These defected areas were spread out over the surface. Furthermore,
the defects did not seem to be inside metal clusters as in the case of membrane (a), but
between metal particles (Figure 6.11).
181
carbon frQm the membrane pores. BrQwn spots were visible in some areas on the outside
membrane surface, indicating the presence Qf carbon. Two. PQs~ibilities exist: either 240 DC
was too Iowa temperature for full QxidatiQn to. take place Qr the Qxygen to carbQn CQntact in
the PQres behind the dense palladium layer was very PQQr.
Hydrogen and nitrQgen permeance tests were perfQrmed Qn a membrane withQut and with
additiQnal QxidatiQn treatment. The results are shQwn in Figures 6.12 and 6.13.
1.4E.Q5
I
H
• Without oxidation
1.2E.Q5
I
------
• Witt! OXidation
~ 1.0E.Q5
~
1 1..-
I!!II ..... 8.OE.(]6
~
a.N·
51 ~ 6.0E-06 ./
es
J.
:z:
4.0E·06
2.0E-06
-------
D,OE+OO
330 350 370 390 410 430 460
Reactor temperature (OC)
1200
• Without oxidation
1000
_With oodation
~--------------~
.~
II
800
i N
600
Z
S '"'0
£
200
0
330 350 370 390 410 430 450
Reactor temperature ("C)
182
The characteristics of membranes (2a) and (2b) are listed in Appendix Dl. There was a
significant increase in hydrogen permeance after oxidation at 320°C. These results
confirmed the presence of either an EDTA or carbon layer in the pores behind the Pd film.
Reduction in hydrogen at up to 500°C did not thermally decompose the layer. When oxygen
was forced through the defects in the Pd film and the membrane support pores under pressure
(between 1.0 and 2.5 bar, depending on the membrane selectivity) at 320°C, most of the
remaining precursor was removed. That resulted in higher hydrogen permeances through the
Pd membrane and improved selectivity (see Figures 6.12 and 6.13).
In certain cases (membrane N7 and N8, see Appendix Dl), even after oxidation there was
some carbon present in the fihn. A good indicator of the presence of carbon in the film was
the rate at which steady state was obtained after the membrane was reduced and switching
from nitrogen to hydrogen during analysis. If steady state was obtained quickly (in less than
3 to 5 minutes) it indicated a pure fihn. When carbon was present in the film it could take 10
to 15 minutes (or even longer) for the hydrogen flux to stabilise (especially at the lower
hydrogen feed pressures).
When carbon was present in the fihn it could be expected that the film would show poorer
stability over time at high temperatures compared to pure films. This assumption was,
however, not further investigated.
183
For those experiments, the higher hydrazine concentration was used (a hydrazine:Pd molar
ratio of 1: 1). Electroless plating with a lower hydrazine concentrations was not tested on the
y-a1umina membranes. A lower hydrazine concentration (a hydrazine:Pd molar ratio =
Data for all tested (permeances) Pd membranes are listed in Appendix Dl. Membrane (11)
broke in the reactor before permeance testing was performed and there is therefor no
permeance data in Appendix Dl for membrane (11).
The side view image of (3b) (Figure 6.14) clearly shows two Pd layers on the alumina
support. Since no alloying was performed after the second coating, a single layer did not
form at that stage. After application of the fIrst layer (membrane 3a), the membrane was
184
tested. A second layer was applied and the membrane (membrane 3b) tested again. The
combined thickness of the two layers should be about 4.4 microns. Figure 6.14 shows,
however, a total thickness of closer to 7 microns. This was the only membrane that showed
deviation between the calculated Pd thickness and the SEM determined Pd thickness. For
other membranes tested by SEM (a, b, c, d and 11), the results of the calculated thicknesses
and the SEM determined thicknesses, were a good agreement. Top view images of (3b)
(Figures 6.15 and 6.16) show a dense structure without any pores or defects. Selectivity data
on (3b) confirmed a dense and compact film with very little defects (see Appendix Dl).
Figure 6.15: Top view of membrane (3b) Figure 6.16: Top view of membrane (3b)
(5000 x) (2500Ox)
185
The calculated thickness of membrane (11) (1.45 microns) was in good agreement with the
SEM determined thickness (from 1.0 and 1.7 microns). Figure 6.17 shows a very dense
layer on top of the 200 urn a-ruumina support. Top view images (Figures 6.18 and 6.19)
confirmed this. Under high magnification (Figure 6.19) it appears as if there are more grain
boundaries for this thinner film compared to the thicker film of (3b) (Figure 6.16).
Figure 6.18: Top view of membrane (II) Figure 6.19: Top view of membrane (II)
(5000 x) (2500Ox)
The hydrogen permeance process was described mathematically in Chapter 2. The main
equations were:
(2.7)
(2.8)
186
Pressure data is necessary to calculate the value of n, which is an indication of the flow
process through the film. Temperature data is necessary to calculate the Arrhenius
parameters (Po and ED) in equation (2.7).
The parameters of each tested membrane were calculated and are listed in Appendix Dl.
The values will be discussed in the followings sections.
The contribution of the final two factors cannot be quantified, but from experience it is
known that there is at least some leakage between the membrane and the reactor seal. The
measured nitrogen permeance represents the worst case scenario or the maximum value.
187
'16.0 •••.••••.•.••••••••••.•.••••.• ---- .• -------------.-----.-------------------.-------.----- '.0
\
•.•--------,~-
~~ot-------;-------,~ . . 7.0 t-----------------------------;;---
l ~ot_----------------~--------~ 1iii
8.0
i
t---:---le-----~-----------=-'
~ ~
__
l! ~0t_------------r============]
2!.ot_---------------j .T .. 450"C _T=410"C f--
i~ 4··00t======::J~~~~~~~
. f--- .T=4&O"C .T=410"C
1~O~============~.T~.~m;~~~';T.~~~~~~t=
8
I
+- 15 0
:
t-----------------------------
10..0
I
i ,o~----========-
I---
3.0 t_-------------/ • T = 370"C xT. 330·C
.ItO
, 1.' ,
"i!
~
35.0
, ~
•
,
.
• :
'iii
:. 1.2
1.4
,
.
,
.
30.0
] • • • •
o
E
25.0 0
E
1.0
c c
- ~o ;- 0.8 .T=450"C .T=.10"C -
I 150 .. T=450"C .T=410·C ---;
~
m 0.' .T"370"C ><T"'330"C -
i 100 .T=370"C xT=330"C
---j ~ 0
1 .•
Z '.0 Z 0.2
0.0 1150 1200
001100 1500 1700 11000 ,,,.
1060 1100
AVllrage pressu... (mbar)
"" 13'"
'''' Average pressure (mbar)
188
the differential pressures are small, changes in the value of n have little effect on the quality
ofthe fit.
4.0
.T=4&lOC
.T"370"C
,T:410"<:
"'T=330"C
~
~! II .0
3.0
t+-::::::::::::::::::::::::::::::::::::-_::::--Ir======;-"1
2.0 + - - - - - - - - - - j • T= 370"C
.T:450"C H .T~410"C
xT. 33O"C ~
i: 2 . 0 t - - - - - - - - - - - - - - j ;
i: 1.0j-------"======"'---i
a.o+----_--_--_-_---.--J 0.0+----_-_--_-_-_--1
o 2Q 40 60 8tI 100 o 20 40 8Q 80 100
Dll'ferentlal pressure (mbar) DIra,ential prliSsurelmbat)
1
"E ! '.0 • ,
·• !
a
6.0
5.0t------------::........_I
0
E
at •
·• • ,i
I .0 t---"-------'~-----"--~-_I c
I '.0
3.0 •
The hydrogen permeance (Figure 6.24 to 6.27) did not vary considerably with pressure.
Differential pressures below 40 mbar caused a slight deviation (see Figures 6.25 and 6.27)
from the trend. The reason for this was that the error of the pressure probe was between 2
and 4 mbar and thus, at low pressures, there was some error in the measured values. A n-
value of 1 indicates that hydrogen chemisorption on the palladium surface is the rate limiting
step (Nam et aI., 1999; Yan et aI., 1994). Sievert's law, where n = \1" is not applicable to the
thin films synthesised in this study. Diffusion is not the rate limiting step.
189
The film thickness and the permeance temperature have a very significant effect on the
hydrogen permeance as will be discussed in the next sections.
In section 6.6.1.1 pressure data for nitrogen permeance suggested that Knudsen flow might
have been the mechanism of nitrogen transport through the defects in the Pd film. This
indicated that the defects were in the lower nanometer range. For each of the fifteen Pd films
tested, the nitrogen permeance was plotted as a function of temperature (see Appendix Dl).
For ten of the films, the nitrogen permeance declined with an increase in temperature (see
Figure 6.28 for a typical example). In four cases there were no clear permeance trend with
temperature change (Figure 6.29) and in one case the permeance increased with increasing
temperature (Figure 6.30).
. "" ""
..-
..t------------I
.,• . ,. ,
__ NIIrogeo pennance
.... t
;;-
N
~
E
,
. ." HycftlgllnJNItrog.n
",,"
"
"'"
ft'
~I
2.8
ut--~~-~------I
". """"
-
6600
I
.I
0 1.3 U
~,"'" /" -il
... f
E
~
sc "
• U ',. ~.
E
18
'
:i'-.. ".
".
- .--
/
.,/ ;j
I ... "
", ~-
--~-'"
,..
.x. HI
i!
t=====1'~"~"~~~~-~":"':-::::'tj ....
U
1/'" z 1.4
u
I,'
." ". T('C)
.., .'"""" ...
~---_--_----4~
." T(OC) '" ...
Figure 6.28: Selectivity data for Figure 6.29: Selectivity data for
membrane (3a) membrane (N8)
190
22 '00
....• .•...
~
-E
21
/,>- 730
~ 20
E ....... ..... .' 710 ~
S
'
..• - •
8
I•
"
"
- .'
.. /
'" I
i 870
z I --:---:~:=-~_I
17
"330 370
TC'C)
.10 ...'"
Figure 6.30: Selectivity data for membrane (N2)
The temperature data confirmed that some kind of Knudsen flow dominated when nitrogen
passed through the palladium film defects. The reason for the decline in permeance was that
the greater vibrational energy of the N2 molecules at the higher temperature resulted in more
resistance to flow through tiny pores and thus a decrease in permeance.
Hydrogen temperature data was fitted to the Arrhenius equation (2.7). Arrhenius parameters
2
for each film are listed in Appendix Dl. The high R -values of the Arrhenius fits indicate
that the data fitted the equation well. The hydrogen permeance increased with temperature,
as predicted by equation (2.7). Figures 6.31 and 6.32 show typical increases in hydrogen
permeance with an increase in temperature.
8
1
'8 6
i 6
~. ,
4
...
~
...01>
-<>
~ . ..t>
':::- ol>
~.gP
191
6.6.3. THE EFFECT OF Fll..M 1IDCKNESS ON PERMEANCE
The effect of film thickness on hydrogen permeance, nitrogen permeance, membrane
selectivity and Arrhenius parameters are depicted in Figures 6.33 to 6.40. The hydrogen
permeance should be inversely proportional to the Pd film thickness (equation 2.9).
P =p.. (2.9)
m I
16.0
.T=450°C .T::410·C
N "O~~~----~============~
!~ 10.0 t--,:;J~"",.___---------____i
~
~ 8.0 t-~~~~""""~-------____i
Figure 6.33: Hydrogen permeance for Pd films from 1 to 6.5 micron thickness
192
18.0 ------------.-------------------------------------------------.-.--. ___ •______________________________ _
• •
ii' 14.0 t-----""'--S.=:i.[--;----~--------___1
~
f'I' 12.0 t-~fL--~-~-~·~~~~~-~~~"'-~~--._-___1
~_§. ~=:"--:.:.=. ~~'~i• '~'='.:..~.~~.-~..;-~-:~:-:~~-~-::.~~~.
10.0 ='
-.... -- ~
.--
8.0 • -.
~N •.0
:
•E .T=450~C .T=410°C
•
1-_______---1
"~ '.0
N
eT=370°C :r:T=330"C
:t 2.0
0.0
0.95 1.05 1.15 1.25 1.35 1.46
Pd film thickness (microns)
Figure 6.34: Hydrogen permeance for Pd films from I to 1.5 micron thickness
~
100
.:. . I
·T=450.C
·T=370·C
1E
o I'
1 E
o I'
c ~
! 10t-------~~----~ S 10t-------------~
••
I
£
:
i
£
0.50 1.50 2.50 3.50 4.50 5.50 6.50 0.95 1.05 1.15 1.25 1.35 1.45
Pel tim thlc:kness tmlc:rons) Pel film thick,.... (microns)
Figure 6.35: Nitrogen permeance for Pd Figure 6.36: Nitrogen permeance for Pd
films from I to 6.5 micron thickness films from I to 1. 5 micron thickness
For the first membranes that were prepared (membranes Ia, Ib and Ic), the plating process
was not well refined and the selectivity was poor as shown in Figure 6.35 at thicknesses of
3.1, 3.9 and 6.2 microns. If those values were excluded then the decline in nitrogen
permeance with an increase in film thickness would be clear. The worst membrane (1.0 !J1Il)
had a nitrogen leak rate of 70 nmoVm2 .Pa.s, while the best ones had values close to
193
1 nmollm2Pa.s. Figure 6.36 indicates that the majority of the Pd films with thicknesses
ranging from 1.0 to 1.5 fUll had a nitrogen leak rate of between 20 and 2 nmol/m2 .Pa.s.
10000 10000
"'.
j
~ 1000 J-------'-~-_:_---------
;: ~t :
.,
'T:450"C ·T"'410"C I
I ·T=3roOC .T;330"C I
100 +--_-_---.::.........--~-_--
0,50 1.50 2,50 3.50 4.50 5.50 6.50 1.05 1.15 1.25 1.35 1.45
Pd film thic:knen (microns) Pd film thickness (microns)
194
"""'" . 2.0£-<>< 20000 2.~
",m
• .--- 2.0E-04 10100
W·Loft ..·.1
• Rightaxi.
•
2.0E-04
•
. ~ .•
;;-
•
. ~
I·loft ... ,l
i"12000 1.5E-G4 i"12000
. .. ..
1.5&04
'E
.8_
~ ----
'$'
. . . • R1d1tultl
1.0E-04
'6
oS
~
.8_ · ~
!
.....
. ~ . . .. ·
1.0E-04
~
4000
. 5.0E-05 4000
· 5.0E-05
0 O.OE+OO 0 O.OE.OO
0.50 1.50 2.50 3.50 4.00 '.50 ...0 0." 1.05 1.16 1." 1." 1."
Pd film thickness (microns) Pd film thkltnass (merons)
Figure 6.39: Arrhenius parameters for H2 Figure 6.40: Arrhenius parameters for H2
permeance (1 to 6.5 micron Pd films) permeance (1 to 1.5 micron Pd films)
The following general remarks summarised in Appendix E can be made regarding hydrogen
permeance through Pd films:
1. Unsupported Pd films or foils are thick (typically 24 J.UIl and thicker) and hydrogen
permeance poor. The best permeance value from literature was 1.2 J.UIlollm2 .Pa.s at 350
°C (Hurlbert and Konecny, 1961), but usually the values were lower than I J.UIlollm2.Pa.s
even at much higher temperatures.
2
2. Pd-Ta-Pd (15 J.UIl) foils had permeance values of up to 1.76 J.UIlollm .Pa.s at 340°C
(peachey et aI., 1996).
3. CVD and wet impregnation were successfully used to produce Pd films ofless than I J.UIl.
4. The surface structure of Pd films prepared by CVD appeared to be unsuitable for high
hydrogen permeance. Typically, films had a hydrogen permeance of less than
195
2 .
1 flmol/m .Pa.s, except for those prepared by Yan et al. (1994) which had values of up to
4 flmol/m 2.Pa.s.
5. The best values for hydrogen permeance were obtained with alumina substrates. Most
hydrogen permeances of films on porous glass were less than I flmol/m 2 .Pa.s. For films
on porous stainless steel permeances were less than 1 flmol/m 2.Pa.s and for films on
refractory metals permeances were less than 2 flffiol/m 2.Pa.s.
6. The best values obtained with alumina supports and electroless plating were:
2
• 9.75 flffiol/m .Pa.s at 450°C for plating on a disc by Zhao et al. (1998), but the H2 to
N2 selectivity of the film was only 23.
2
• 2.86 flmol/m .Pa.s at 400°C for plating on the outside of a tube by Kikuchi (1995).
The selectivity was not mentioned.
• 5.27 flffiol/m 2.Pa.s at 500°C for plating on the inside of a tube by Shu et al. (1997b).
The selectivity was not mentioned.
The best overall hydrogen permeance values that have been published were for Pd films on
modified porous stainless steel discs, prepared by wet impregnation. Film thickness varied
from 0.5 to 0.8 flm. The hydrogen permeance was 15.8 flmol/m 2 Pa.s at 450°C (Iun and Lee,
1999) and 17.8 flffio!/m 2.Pa.s at 550°C (Nam et aI., 1999). In both cases the H2 to N2
selectivity was above 1000.
196
During the present study some important advances have been made in the preparation of Pd
membranes. Pd film thickness on the inside of alumina tubes has been reduced to between
1.0 and 1.5 /lm, while maintaining H2 to N2 selectivities exceeding 400 for the majority of the
films. An important cost advantage is that the cheaper 200 nm a-alumina support was used
successfully. The more expensive y-a-alumina support was not necessary. The thinnest Pd
films prepared previously by electro less plating on the inside of membrane tubes were:
• a 2.0 /lm film on the inside of an assymetric aly-alumina membrane with a 5 nm pore
size (from SeT) by Shu et al. (1996b), and
• a 2.1 /lm film on the inside of an assymetric aly-alumina membrane with a 3-4 nm pore
size (from SeT) by Shu et a!. (1997b).
The highest permeance obtained for a Pd film on the inside of a tube (see section 6.6.4),
excluding the results from Zhoa et al. (1998), due to poor selectivity, and results from Shu et
al. (1997b) and others, where selectivity was not mentioned, is 2.68 /lmolJm2 .Pa.s at 467 °e
(Li A et a!., 1999). In the present study, hydrogen permeances ofPd films from 1.0 to 1.5 /lm
varied between about 8 and 15 /lmol/m2 .Pa.s for temperatures from 330 °e to 450 °e and
(Figures 6.33 and 6.34). These values are a significant improvement over other published
results.
2
Only the results of Jun and Lee (1999): 15.8 JlIDol/m .Pa.s at 450 °e, and Nam et al. (1999):
17.8 JlIDol/m2.Pa.s at 550 °e, are comparable to values in this study, but they used the
unfavourable disc membranes.
197
(6.1)
where L is the axial position along the membrane. The axial position is made dimensionless
by dividing both sides of (6.1) by the reactor length, Lo. Substituting parameters yields:
(6.2)
For hydrogen, the flow was from the tube side (high H2 pressure) to the sweep side (low H2
pressure), while for nitrogen the flow was from the sweep side (high N2 pressure) to the tube
side (low N2 pressure). Partial pressures was expressed in terms of flow rates. Substituting
molar flow rates in (6.2) yielded two coupled differential equations:
(6.3)
(6.4)
F H(O) is the hydrogen molar feed rate on the tube side and FN(o) is the nitrogen molar feed rate
on the shell side. The atmospheric pressure was taken as 100 000 Pa. The boundary
conditions for (6.3) and (6.4) were:
198
Equations (6.3) and (6.4) were solved using average hydrogen and nitrogen permeances
(found in Appendix 01) and employing Euler's method (Fogler, 1992) for first order
differential equations. A very small step length of 0.001 was chosen.
The amounts of hydrogen that permeated through the Pd films at different sweep gas ratios,
hydrogen space times and different temperatures are given in graphical format in Appendix
Dl. Hydrogen flow profiles along the axis of the membrane were calculated for membranes
(N4) and (N7), to compare the experimental data with the calculated data. To solve equations
(6.3) and (6.4) the following assumptions were made:
• The reactor was isothermal, and
• Hydrogen and nitrogen permeances were independent of pressure, even at low pressures.
Table 6.2 shows that the error between the experimental values and calculated values was
typically less than 1.5%, indicating very high accuracy. It was concluded that the model
assumptions were valid and that the model predicted hydrogen permeance very well.
Table 6.2: Experimental H2 permeances vs. calculated values for membrane (N7)
Space time = 2.37 seconds (200 cm'/min Hz feed)
Temperature eC)
Sweep gas mol 450°C 410°C 370 °C 330°C
% of Hz feed % Hz permeated
5 96.8 96.7 96.3 96.0
(98.2) (98.0) (97.6) (97.0)
10 97.7 97.5 97.2 96.9
(99.0) (98.9) (98.7) (98.4)
20 98.5 98.2 98.1 97.8
(99.5) (99.4) (99.3) (99.1)
199
Space time - 1.19 seconds (400 cm'/min Hz feed)
Temperature ("C)
Sweep gas mol 450°C 410°C 370°C 330 °C
% of Hz feed % Hz permeated
5 96.7 95.9 93.3 87.9
(98.1) (97.6) (93.9) (85.7)
10 98.3 98.1 97.9 97.6
(99.0) (98.9) (98.7) (98.4)
20 98.9 98.7 98.6 98.6
(99.5) (99.4) (99.3) (99.1)
Values III brackets are the model values or calculated values
Table 6.3 tabulates a similar set of data for membrane (N4). Only at the very low sweep gas
flow rates (2.5% of H2 feed) did the experimental and calculated values deviate by more than
2%.
Table 6.3: Experimental H2 permeances vs. calculated values for membrane (N4)
Space time = 2.37 seconds (200 cm'/min H2 feed)
Temperature ("C)
Sweep gas mol 450°C 410°C 370 °C 330°C
% ofH2 feed % H2 permeated
5 97.8 97.5 97.7 97.4
(97.0) (96.8) (96.6) (95.7)
10 98.6 98.4 98.6 98.3
(98.4) (98.3) (98.2) (97.7)
20 99.2 99.0 99.1 99.0
(99.1) (99.1) (99.0) (98.7)
200
Space time -1.19 seconds (400 cm'/min Hl feed)
.
Temperature (0C)
Sweep gas mol 450°C 410°C 370°C 330°C
% ofH1 feed % Hl permeated
2.5 94.7 92.8 91.4 87.2
(92.0) (90.3) (87.4) (81.2)
5 98.0 97.7 97.6 96.8
(97.0) (96.8) (96.6) (95.6)
10 98.9 98.8 98.8 98.6
(98.4) (98.3) (98.2) (97.7)
Values In brackets are the model values or calculated values
Figures 6.41 and 6.42 show the effects of temperature and sweep gas flow rates on hydrogen
permeance through the Pd film of membrane (N7). Both an increase in temperature and
sweep gas flow caused an increase in hydrogen permeance. The increase was more sensitive
to the sweep gas flow rate (see Figure 6.42).
99.0 100.0
i
98.• t-------------""'"
~; ------:::: 98.0
-:?</
... . ~ .'
\
1i 97.' +-----/"""""?/""-:,.~~,.....,
/-<~.~~·:;. >·t 1
98 0 ".0
I ;
.
...~:~. ~.~.~'"'-=.-,..-.".~; ~
".0
, ,I
8.
-£ 97.0 +----,r-:7..'.p.. "-
.. 7,""','-o.rr"'-----------j :£
92.0
,
~T·"·~H
I--T=«iO"C
96.5i----::-..'/,-.L--;::::======:-i ~
<I-
".0 +-
__----.o'<-'_·_--i1 - -r :4&O"Cr;;410-C:
__
8M
! I.".· T=370"C ..... ·T=330"C
The effects of space time, temperature and sweep gas flow rate are clearer in Figures 6.43 to
6.46, where hydrogen flow along the membrane axis is plotted. By comparing Figures 6.43
and 6.44 it can be seen that increasing the sweep gas rate was more effective for removing
201
hydrogen than increasing the temperature was. For all experiments with a hydrogen feed rate
3
of 200 cm /min (space time = 2.37 s), equilibrium was achieved at some point along the
membrane axis. For a feed flow rate of 400 cm3/min (Figures 6.45 and 6.46), the permeance
rate at low temperatures and low sweep gas flow rates was too slow to allow all the hydrogen
too permeate through the Pd film. The large effect of sweep gas flow rate on hydrogen
permeance can be seen in Figure 6.46. Hydrogen permeance was incomplete at a sweep gas
flow rate of 20 cm3/min, but increased to full completion at four tenths of the membrane
length for a sweep gas flow rate of80 cm3/min.
200
.. \,
"
0
0 0.2 0.4 0.8 0,8 o 0.2 0.4 0.8 0.8
DifTW!nslonllss .......ne J.ngth Dlrmnstonllss . .mbrane length
Figure 6.43: H2 flow in tube (cm3/min) as Figure 6.44: H2 flow in tube (cm3/min) as
predicted by model for N7 (T = 2.37 s) predicted by model for N7 (T = 2.3 7 s)
40J .---,'----.---------.---- .-.•.. ---.---.--.• ------,---------------_ •. -•. ----.•.• - 40J ------.,---.------.. -._-.-----•. ------.----.--------•. ----.. ---•.... ---.-----.-------- ----l
T-330°C
320
T_460"C,41D"C,370 -e.nd".,'C . . . . . . . _·80,40 ... 20
1 ,.. \
:5 I1'IO'oW18 from I.n: to rlghI: ;; cnI'/mkt ft!KMnO from IIIItD right
E ~
i 240
e
~
o ". ! ,0>
'"£ £'"
" '~
'-'.~.-".~.. ::~~
"
0 0
0 0.2 0.4 0,8 0.8 0 0.2 0.4 0,8 o.a
Dlnwnstonlus membrane length DI....n5kx11nli n.rnbrane a.ngth
Figure 6.45: H2 flow in tube (cm3/min) as Figure 6.46: H2 flow in tube (cm3/min) as
predicted by model for N7 (T = 1.19 s) predicted by model for N7 (T = 1.19 s)
202
Figures 6.47 to 6.52 are a repetition of Figures 6.41 to 6.46, but for membrane (N4).
Membrane (N4) had significantly higher hydrogen permeance values than (N7) (see
Appendix Dl).
From Figures 6.49 to 6.52 it can be seen how, when compared to data of (N7), all the curves
were shifted to the left. Complete hydrogen permeance for membrane (N4) was achieved
much quicker along the membrane axis i.e. at lower dimensionless reactor lengths, than for
membrane (N7).
100.0
'
-.---- --
.... ~'l
J ,., 89.0
I .... Ii" I: I
,.
/"
i I 920
J .... !
%:
". 8B.O
%:
". ,
& "
I---T=450"C __ T"'410-C
~;' r 90.0
i17.S t-_'-----j
_T=460"C -+-T=.10-C
, I
1···T.moe .... ·T=330-C
g7.0+-~-_-~-~~-_-~_ ....
5 10 15 ~ ~
Hz: sweep rate as molar" f1l H,; feed
M ~ ~ 0
• 10
N,; swap me as molar % of H,; fHd " "
Figure 6.47: % Hz permeated with N z Figure 6.48: % Hz permeated with N z
sweep gas and space time = 2.37 s sweep gas and space time = 1.19 s
200 ,, ""
IwMP a- t.ecI_ 10 I;ftlllmln T_moe
'0]
!
, '0]
nI_."
:e
~ ".
1\ T-450"'C,41D"C, 3'ftJ '"CandlSO"C
mo'otng from ... to light
!,
,
:eE
11>' \\\
:!!weep u- 20 md 10
cm·..... 1NnIng from "'10 right
e
~ so
~ !
Ii
~ eo \\\
~
'"%:
%:
40
!
! . \\\
\~ ,~~
i
i \\\ ~ -- --- ----,. . _., ......
•• ...
_.,-,
, •• ...
,
'.2
Dlrmnstonless lM~n. length
••• ••• 0.2 0.'
DkNnslonless nw:....n. t.ngth
0.' 0.'
Figure 6.49: Hz flow in tube (cm3/min) as Figure 6.50: Hz flow in tube (cm3/min) as
predicted by model for N4 (I: = 2.37 s) predicted by model for N4 (I: = 2.37 s)
203
... ...
s..., .. fMcI_ 20 cm'hnln T.3SO"C
1\ l-
>20
\ T. 45D "C, 410 "Co '70 "C lOCI !SO "C "" \~ aw...,.,.. .... -10, 4O.,d 20
:5 :5
~ 240
mollWlU frvm , ... 10 right
.!i.
~ 100 ~ ! S\~
\\\, ~
1GO
'"-£ '~ -£
80 eo
'~~ "
~
0
.... :-::.~
, 0
\ "'" " ,
Figure 6.51: H2 flow in tube (cm3/min) as Figure 6.52: H2 flow in tube (cm3/min) as
predicted by model for N4 (. = 1.19 s) predicted by model for N4 (. = 1.19 s)
6.7. SUMMARY
Each step in the e1ectroless plating process must be carefully optimised to produce thin,
highly selective, Pd films. For pretreatment, a low Sn to Pd ratio was employed to limit Sn
deposition and increase thermal stability of the film. The quality of the Pd film was very
dependent on the plating rate. A high plating rate, due to: a high hydrazine concentration, a
low EDTA concentration and/or a high plating temperature, must be avoided to produce
selective thin films. In some cases the Pd films had defects due to high plating rates. The
last critical step in the membrane production process was the post plating cleaning. After
plating, membranes were stirred in ammonia solution for several hours, dried and then further
oxidised at 320°C in pure oxygen before reduction.
Pd films, of thicknesses down to 1 fllll, were deposited on the inside of asymmetric SCT
a-alumina membranes (200 nm pore size). Hydrogen permeances ofPd films from 1.0 to
1.5 fllll, varied between about 8 and 15 fllllol/m2.Pa.s for temperatures from 330°C to
450 °c. Hydrogen to nitrogen selectivity was> 100 for all membranes tested and > 400 for
all but two membranes (thickness 1.0 to 1.5 fllll) tested. These values are a significant
improvement over other published results.
204
Hydrogen permeance fitted the flux equation well, with the permeance constant at different
differential pressures. This implied a n-value of I in the flux equation. Temperature data
fitted the Arrhenius equation with high accuracy. For the majority of the membranes,
nitrogen flow through defects in the films showed signs of Knudsen flow. Finally, hydrogen
permeance was tested using an inert sweep gas. Two coupled differential equations were
formulated for modelling the membrane as a plug flow reactor with permeance, but without
reaction. The equations provided very good fits for the experimental data.
205
7. Pd-Ag MEMBRANE ALLOYING AND STABILITY
TESTING
Pd membranes can only be used above 300°C. Pd-Ag membranes can, however, be used at
lower temperatures since to the a to ~ phase transition in the Pd crystals is suppressed by the
silver addition. This a to ~ phase transition in the Pd crystals is what causes defects and loss
of selectivity in Pd films below 300°C.
In Chapter 7 the preparation and testing of Pd-Ag membranes will be described. They will
also be compared to pure Pd membranes. The membranes were heat treated at high
temperatures in an effort to form a homogeneous alloy between the Pd and Ag. The stability
of the membranes was monitored while the alloying process was in progress. The Pd-Ag
membranes were also tested at temperatures below 300°C.
The main advantage of sputtering is that complex alloys can be prepared. Metals can be
deposited in any ratio and with high accuracy. The disadvantage is that the technique is not
suitable for depositing metal on the inside of a membrane tube. Even depositing metal on the
outside of tubes may cause problems with thickness control and composition uniformity.
206
Only electroless plating can be used for co-depositing Pd and Ag on the inside of a membrane
tube, but the process control is poor. It is very difficult to obtain the desired Pd to Ag ratio on
the membrane surface (Shu et aI., 1993; Cheng and Yeung, 1999).
The plating kinetics for Pd and Ag in solution differs. Their stabilities also differ, with Pd
being more stable in solution than Ag. This resulted in very diluted concentrations having to
be used for co-deposition. The Pd to Ag ratio in the starting solution differs considerable
from the deposited composition on the membrane. Co-deposition is discussed in more detail
in Keuler (1997a), Cheng and Yeung (1999) and Shu et al. (1993).
In the present study, Pd-Ag membranes were prepared by successive Pd and Ag plating. The
Pd layer was deposited first, followed by the Ag layer. The plating procedure was discussed
earlier in section 3.3.2.4. The composition of the plating solution was given in Table 3.16
and the plating procedure given in Table 3.17.
Before the silver layer was deposited, the membrane,. covered with teflon tape on the outside,
was dipped twice in both the pretreatment solutions. Post plating cleaning was similar to that
after Pd deposition.
In the table, when two temperatures and two times are given in a row, the first heating time is
for the first temperature and the second heating time is for the second temperature.
207
Table 7.1: Alloying procedures used for Pd-Ag membranes
Name Elements Alloying Heating Gas Temp ("C) Time (b)
system
8 Pd No
8b Pd+Ag No
8e Pd+Ag Yes Tube oven Ar 550,600 15,25
NI Pd No
NIb Pd+Ag No
NIx NewPd+Ag Yes Reactor H2 590 10
N3 Pd No
N3b Pd+Ag No
N3e Pd+Ag Yes Tube oven Ar 545 50
N3d Pd+Ag Yes Tube oven Ar 545 100 (total)
N3e Pd+Ag Yes Tube oven Ar 545 150 (total)
N4 Pd No
N4b Pd+Ag No
N4e Pd+Ag Yes Tube oven Ar 530 30
N4x NewPd+Ag Yes Reactor H2 540 30
208
7.2.1. CHARACTERISING UNALLOYED Pd-Ag MEMBRANES
Four Pd-Ag membranes were tested prior to alloying. They were: 8b, NIb, N3b and N4b
(see Table 7.1). The metal layer, at that stage, consisted of a Pd film and a second Ag film.
The hydrogen permeances of the membranes are depicted in Figures 7.1 and 7.2. The
addition of silver resulted in a sharp decline in hydrogen permeance at all temperatures. This
confirmed that silver has a very poor hydrogen permeance and that separated Pd and Ag
layers in the film should be avoided.
15 ........................................................-.......................................-.... -j
t..
12 ~
__-- --------
_- .. "'" - .. ~i
N
~
E
.
__ ., .• - - -
-+-1.1 "mPd(8)
'
IE 6~-------------------------I
~ 1.4 ~m Pd-Ag(22%) (N1b)
!. ~.",.~..;--.=.t
-£ t::. :." :;:.;:." ~
3+---------~------~--------~
330 370 410 450
T (.C)
Figure 7.1: Hydrogen permeances for Pd and Pd-Ag membranes (8, NI, 8b, NIb)
Figure 7.2 provides a good example of how silver influences the hydrogen permeance.
Metal films of membranes (N3) and (N4b) were of similar thickness. Membrane (N3)
consisted of pure Pel, while N4b had a Pd layer of 1.18 J.Un and a silver layer of
approximately 0.25 J.Un. The hydrogen permeance for membrane (N4b) was more than three
times lower than that of (N3), even though this silver layer was very thin.
Pd and Ag can be considered as two resistances in series. The purpose of alloying was to
obtain a homogeneous alloy. The two resistances in series would change to two resistances
in parallel and the hydrogen permeance would increase.
209
15 .................................................................................................................. .,
~
In
oj - ..,a..- - - -
~ 12+---------~r_~~----------------~
1ia 9~~~-=-~-=-~4~====;::;~-~~======~
~ --+-1.471J1Tl Pd (N3)
3 ---2.141'"1 Pd--Ag(24%) (Nab) .
m 6 + - - - - - - - - - - - - - - - - - 1 - .... 1.181'"1 Pd (N4) r-j
~ 3 -+-1.43"m Pd·Ag(23%) (N4b) ,
~- .--::.
o+---------~----------.---------~
----
330 370 410 450
T (OC)
Figure 7.2: Hydrogen permeances for Pd and Pd-Ag membranes (N3, N4, N3b, N4b)
Table 7.2 lists nitrogen permeance and selectivity data for membranes 8b, Nib, N3b and
N4b. For all the membranes except (NIb), the nitrogen permeance decreased after the silver
layer was deposited. That was expected, since silver plugged defects that were present in the
Pd layer. Thicker films tend to have less defects, as discussed in Chapter 6. Membrane
(NI) had a very low nitrogen permeance and high selectivity. After silver deposition, the
film showed poor characteristics. That was the only silver plating experiment that resulted in
this behaviour. Other unalloyed Pd-Ag membranes in Appendix D2 (8b, N3b and N4b)
showed good selectivity after silver deposition. Two possible explanations are offered for
this behaviour:
• Some kind of physical damaged to the film might have taken place.
• There might have been hydrogen left in the system during cooling after testing (NI). The
hydrogen would then have caused some embrittlement of the pure Pd film on which the
silver was deposited.
The selectivities of the Pd-Ag films were lower than those of pure Pd films. The loss of
hydrogen permeance was far greater than the gain in reducing defects or the decline in
nitrogen permeance. The net result was a decline in the H2 to N2 ratio.
210
Table 7.2: N2 permeance and selectivity data for 8b, NI b, N3b and N4b
Temp. (0C) N2 permeance Selectivity Nz permeance Selectivity
(nmollm 2.Pa.s) (nmollm2.Pa.s)
MembraneS MembraneSb
450 21.42 671
410 21.85 627 11.48 472
370 22.42 553 11.90 394
330 23.07 457 12.24 336
211
100 100
•0
0
••
95 A ,
B
"~
•
95 r-,o--------_____ _
90 ' -_ _ _ _ _ _ __
E
•a. 90 / &Mt-------------
1
;E
1f.
""~ "
/ ,
;E
1f.
!
80
75
t----------
t--------- •
"
;;
;!
u•
80 / ,
]
~
10 t--------:,.-L- .--~-.- -, _._ .. -
/
65 /'
8O~
" 75 so
" 90 95 100
Measured V. Hl parmeance Measured % Hz penneance
Figure 7.3: Measured vs. calculated % Hz Figure 7.4: Measured vs. calculated % Hz
permeance for 8b permeance for Nl b
100
100
•
0
:"
90 /- ,
0 •
••"
90 ~
E
• /- E
• /._---
/,
0. 80 a. OJ
;E
1f.
70
;E
1f.
70 /-. --- --
""-!i ~•
.
;;
0
eo / ;!
•
80 /'. --- .- --
~
u
V
U
so 50
so 60 70 60 90 100 so eo 70 eo 90 100
Measured %Hz penneanee Measured % Hz penneance
Figure 7.5: Measured vs. calculated % Hz Figure 7.6: Measured vs. calculated % Hz
permeance for N3b permeance for N4b
The model equations (listed in section 6.6.5) predicted Hz permeances well, as long as the
nitrogen permeances were low (less than 20 nmol/mz.Pa.s). For membrane (Nlb),the error
between measured values and calculated values was large due to very poor selectivity of the
membrane (high nitrogen permeances). For the other membranes (N3b, N4b and to a lesser
extent 8b), the error between the actual and measured values was small. Most of the time the
calculated percentage of hydrogen permeated was less than the measured percentage.
Figures 7.7 and 7.8 show the measured hydrogen permeances as a function of temperature
and sweep gas flow rate for membrane (N3b). Figures 7.9 to 7.12 show hydrogen flow
profiles along the axis of the membrane for (N3b).
212
100.0 -------.---------------- ------------------------------------- -.----------.---------------..
~--....--.---------~-~
100.0
~ -,:-..--.""1"1""----_.-..-..._-----
-:'
".0 ".0
90.0 ,-
".0
: . r •rJ
I .
~
".0
I
: I I :
94.0 BO.O
,I r
1
-£
920
90.0
t-__f-/'_ _ _ _---il-+-T:::.UO-c ."110- T:370"C ~ ;£ ro.O
75.0
I:
I: i
: r
;I. I ;I.
".0 . I 1---'T=330"C ".0
80.0
.' j_+_T;;410"C
__ 'T"330'C
-"110.- T=370"C ~
'6.0 55.0
".0
0 10 15 ~ ~ ~ ~ ~
50.0
0 , 10
" " 25 30 35
"
NI; SWeep rata IS molar % of ~ fHd NI; SWHP rate as molar % or HI; fIIltd
'80HIr-----------------i '0)
~
1•
0
120
80
e
.,•
0
1aJ
80
\0>'~
.\
\
'
...........
";£ ,£ \,~""'"
,\ '...... ~
., ~
......
'"
\\,~
'--
............................
0
0 0.2 0.4 0.11 0.8
0
0 0.2 0.' 0.' 0.' ,
DImensionless membrane length Dimensionless membrana length
Figure 7.9: Hz flow in tube (cm3/min) as Figure 7.10: Hz flow in tube (cm3/min) as
predicted by model for N3b (1: = 2,37 s) predicted by model for N3b (1: = 2.37 s)
400 ----------------------.---------------------.------.----.-------------------------------------
.00
'20
T .. 410 '"C, 310 "C MIl S30 "C
'" Sw. . 0-"" .1,,10, 4O..-.d 20
C moWlt from ,... to right c anI""", -*'Ibm'" to righI:
~ "0 ~ 240
e e
~ 160 •
~ 160
"'
;£ ;£
80 80
0 0
0 0.2 0.4 o.e 0.8 0 0.2 0.4 0.8 0.8
Dimensionless membrana length DlnansfonlHs memtnnaltnglh
Figure 7.11: Hz flow in tube (cm3/min) as Figure 7.12: Hz flow in tube (cm3/min) as
predicted by model for N3b (1: = 1.19 s) predicted by model for N3b (1: = 1,19 s)
213
Figures 7.7 to 7.12 show that, in most cases, at 330°C the hydrogen penneance did not go to
completion along the length of the membrane tube. When Figures 7.9 to 7.12 are compared
with Figures 6.43 to 6.46 and Figures 6.49 to 6.52 (pure Pd films), it can be seen how far the
curves shift to the right when silver was added. Much higher sweep gas feed rates were
needed to extract a high percentage of hydrogen through the Pd-Ag films.
Other references in Appendix E do not specifically state if and how they annealed their Pd-
Ag membranes.
214
from 11 to 30.5 wt %. Membranes were heat treated at 500°C, but the double layer structure
was not changed to a single alloy layer. Membranes were then treated in argon for 12 hours
at 800°C to 900 °C to yield a miscible palladium-silver alloy. They claimed to have 100%
selective membranes with a hydrogen permeance varying between 1.45 and
2
2.24 ).tmol/m .Pa.s at 400°C. The highest value was obtained for a 23 wt % silver film. The
reactor they used had alSO °C temperature difference between the inlet and the centre.
Shu et al. (1996a) prepared Pd-Ag membranes on porous stainless steel disks (200 nm pore
size) with successive electroless plating steps. The film consisted of a 2.8 ).tm Pd-Ag layer.
The Pdso -Ag20 film was annealed in hydrogen at 500°C for 5 hours. The Tamman
temperatures were 630°C for Pd and 344°C for Ag. They claimed that annealing at 500°C
would cause significant vibration and migration of silver atoms into the palladium lattice,
resulting in interdiffusion. Silver penetrated into the Pd layer to 1, 1.3 and 1.46 ).tm after
annealing 5 hours at 400°C, 500 °C and 600°C, respectively. Results were obtained by
Auger electron depth profiling. No hydrogen permeance or selectivity data was provided.
215
The Arrhenius equation is applicable for diffusion data:
D=Doexp(- ED)
RoT
(7.1)
Figure 7.13 was constructed to compare diffusion rates (of Table 7.3) at different
temperatures by applying equation (7.1). The figure must be interpreted in the following
way: The times on the y-axis are indications of the time required to obtain the same amounts
of diffusion as after 100h at 500 °e, but at different temperatures. There is a significant
difference in the values from the three different sets of diffusion coefficients.
100 .................................................................................................................. .
g ~+---~~~~~----------------~
..
E
~ ~+-----~~----~.---~~~----~
20+-------------~~------~~~~
o+-----.-----~----~----~----~
500 520 540 560 sao 600
Temperature 1°C)
Figure 7.13: Heating times required to obtain similar Pd-Ag diffusion at different
temperatures
216
In the present study the temperature range used for alloying was from 500°C to 600 °C.
Different alloying times and gas environments were employed. No literature data could be
found on alloying films on the inside of a membrane tube. A major problem associated with
online diffusion analysis is that the membrane has to be broken to reach the inner surface for
determining alloying (XRD analysis) and performing a depth analysis (Auger depth profiling,
PIXE analysis or EDX). Once a membrane is broken, it cannot be used again. A different
approach was taken in the present study. The permeance and the selectivity of the Pd-Ag
films were monitored to determine the effect of annealing on the membrane performance.
From the changes in hydrogen permeance and membrane selectivity, certain conclusions
could then be drawn.
Section 3.3.6 described the advantages of both methods and the different variables
investigated.
217
7.0 .---------------... ------------------------------------_.-.. _._ .. -----------------.-------------------
.. 6.0 /
~ //
~ 5.0 //-/
! W/ 4.0
~ V 3.0
~ 2.0 t------------f'----:-+---;Tc::=-;:41~0o;;:C-,---!
~ 1.0 +------------1 ~T=370°C
........ T=33Q°C
-
0.0 +-----~----~-------1
o 50 100 150
TIme (hI
Figure 7.14: Effect of heating time in argon on hydrogen permeance through Pd-Ag
(membranes N3b to N3e)
Figure 7014 indicates a sharp increase in hydrogen permeance after the first 50 hours of heat
treatment. The values were more that double those for the unalloyed Pd-Ag film. After 100
hours, the hydrogen permeance showed a slight decline, but values changed little from 100
hours to 150 hours. The increase in hydrogen permeance confirmed that silver diffused into
the palladium matrix to yield a structure with less resistance than the separate silver and
palladium layers. The decline in hydrogen permeance at 100 hours and 150 hours might be
due to one of two reasons:
• The surface structure and morphology of the film changed after prolonged heating,
causing a decline in hydrogen permeance, and/or
• Silver diffused through the Pd layer and started accumulating next to the alumina
interphase. The structure then started approaching separate layers again, but with a high
silver concentration next to the membrane support.
Table 7.4 compares hydrogen permeabilities for the pure Pd film with those of the Pd-Ag
films. Permeability (Pm*l) was used rather than permeance to compare values, since
permeance did not allow direct comparison of films with different thicknesses. The Pd film
had a thickness of 1.47 !J.IIl and the Pd-Ag films a thickness of2.14!J.1Il. The silver content of
the film was 24%.
218
Table 7.4: Effect of heating on hydrogen permeances for membrane N3
Membrane Heating T - 410°C T-370°C T - 330°C
time (h)
Permeability (llffiol.J.ll11/m".Pa.s)
N3 0 15.89 14.19 12.27
N3b 0 6.87 6.03 4.96
N3c 50 14.10 12.82 11.09
N3d 100 12.54 11.24 9.18
N3e 150 12.90 11.58 9.82
Values in Table 7.4 indicate that after silver was added to palladium, the hydrogen
permeance did not return to the values obtained from the pure palladium. At all heating times
(from 0-150 hours), the hydrogen permeances were lower than for the pure palladium. This
contradicted results obtained by Cheng and Yeung (1999) and Uemiya et al. (1991a). They
recorded improvements for Pd-Ag films compared to Pd films.
Reasons for this contradiction could not be obtained from permeance and selectivity data
alone and could not be fully explained without further investigation. It must be noted,
however, that the Pd-Ag film produced on membrane (N3) had a hydrogen permeance that
was three times higher than that ofUemiya et al. (1991a) and more than 60 times higher than
that of Cheng and Yeung (1999). From this initial experiments, it could be concluded that the
surface structure and morphology of the metal film had a greater impact on hydrogen
permeance than the film composition when the hydrogen permeance was very high and the
film very thin (less than 2.5 microns).
219
14
"iii' 12
:.
~ 10
~ 8
-=.
I:! •
ill
E 4
1N
::c 2
o
410 370 330
Permeance temperature (OC)
220
film cracking and pore formation, leading to an increase in the amount of defects in the film
and an increase in the nitrogen permeance. Paglieri et al. (1999) performed the first
comprehensive study on the effects of film temperature on the film stability of pure Pd film
deposited on the inside of alumina membrane tubes. The majority of the films had
thicknesses between 8 and 27 micron.
Paglieri et al. (1999) tested membranes from 400°C to 600 DC, and some even up to
temperatures of 800 DC. They showed that the same deactivation occured in films of all
thicknesses. The thicker the film, the longer it took for the selectivity to fall at 550 DC.
Initial hydrogen to nitrogen selectivities were only about 30 to 80 for most membranes. After
several days at 550 DC, selectivity declined to a Knudsen value of about 4. They obtained
significant improvements in stability by omitting tin from the pretreatment process and
replacing it with Pd acetate. They concluded that the stability of the film was influenced by:
• Tin deposition during pretreatment,
• Components of the electroless plating solution (mainly EDTA) trapped between the
alumina surface and the Pd film, causing pore formation upon heating, and/or
• A difference in the thermal expansion coefficients of alumina and Pd.
After plating, the membranes were soaked overnight in water at 70°C. There was a very
good probability that there was EDTA trapped in the 200 nm pores of their membranes. In
the present study, plated membranes were rotated at high speed in an ammonia solution for
several hours. (EDTA has a high solubility in an ammonia solution, but a very low solubility
in water.) Thereafter, membranes were dried and oxidised at 320 DC in pure oxygen. Even
then, carbon was still present in some samples.
221
strength sensitiser tested by Paglieri et al. (1999). Furthermore, a lot of attention was given
to post plating cleaning (see sections 3.3.2.3 and 3.3.5).
Figure 7.16 shows the increase in N2 permeance (at the permeance temperatures) of
membrane (N4) after 30 hours of alloying in argon at 530°C. The permeance remained
below that of pure Pd, because of the extra thickness of the film. The decrease in permeance
was about 50"10. The membrane retained good selectivity at 530°C.
2STr====================~----------'
iii I~ Pure Pd ED Pd-Ag Ell Pd-Ag, after 30h at 530"C I
oj
D. 20
1E~ 1S
I:
~
g 10
~
l
N
S
Z
o
410 370 330
Permeance temperature rCJ
Figures 7.17 and 7.18 show nitrogen permeance and selectivity data for membranes (N3b) to
(N3e). The heating temperature was 545°C in argon. Figure 7.17 shows a steady increase
in nitrogen permeance as the amount of defects in the film increased. Nitrogen flow
increased with decreasing temperature, indicating that the defects that formed in the film
were of Knudsen dimension (5 to 30 nm range). The selectivity of the membrane decreased
to values between 110 and 75 for temperatures from 410°C to 330 °C. A disadvantage of the
heating system was that it required cooling before the membrane could be tested. After each
set of tests, the membrane had to be reheated to 545°C to be alloyed further. This repeated
temperature cycling, between room temperature and 545 °C, could contribute significantly to
the decrease in membrane stability. The complete temperature cycling for membrane (N3)
was:
• 25°C to 450 ° to 25 °C for testing the Pd film,
222
• 25 °C to 410 to 25 °C for testing the Pd-Ag film,
• 25 °C to 545 to 25 °C for alloying the Pd-Ag film (50 hours at 545 0C),
• 25 °C to 410 to 25 °C for testing the Pd-Ag film,
• 25 °C to 545 to 25 °C for alloying the Pd-Ag film (50h; total = 100 hours at 545 0C),
• 25 °C to 410 to 25 °C for testing the Pd-Ag film,
• 25 °C to 545 to 25 °C for alloying the Pd-Ag film (50h; total = 150 hours at 545 0C), and
• 25 °C to 410 0 to 25 °C for testing the Pd-Ag film.
I :+-------#"'-------1 8001-~-----_1~~~r_~
r --T:4,O.C
~> ,OOf~~~----~~~T~.3~70~.C~
i:t-----;&:2:::.--i-~;::Tr;-.41i370;:;.cOl
-T=33Q"C
) <00 t---~,;:---------j
z 10~~~-----~~~T~.~~~.c~
o~---_---_---~
100 150 o 50 100 150
Time (hI Time (hI
Figure 7.17: Effect of heating in argon Figure 7.18: Effect of heating in argon
on nitrogen permeance (N3) on selectivity (N3)
Membranes (8c) (1.54 J.UI1 Pd75-A~5) and (N6) (2.05 J.UI1 Pd7"Ag21 ) were heat treated in
argon for 15 hours at 550 °C and then for a further 25 hours at 600 °C. The results are shown
in Table 7.S. A temperature of 600 °C is too high to tolerate for films of thickness in the
order of 2 J.UI1. Selectivity will decrease rapidly, making the membrane unsuitable for
separation at those high temperatures.
..
Table 7 S' Nitrogen permeance of membranes (8c) and (N6) after heating at 600 °C
Sample Nl permeance before Nl permeance after
heating (nmoVm1.Pa.s) heating (nmoVm1.Pa.s)
8c 12.2 (330 0c) 114 (room)
N6 8.5 (room) 151.4 (room)
223
7.4.2.3. Structural changes of Pd-Ag membranes during heating
A scanning electron microscope was used to analyse the surface structure of a Pd-Ag
membrane (membrane d) after various heating times. Top view images were taken to detect
pore formation and possible clustering with time. Side view images were taken to detect
peeling. The film had a theoretical thickness of 1.45 ).1m, which correlated well with the SEM
determined thickness. The silver content was 20%. Table 7.6 describes the different
membrane stages.
Figures 7.19 to 7.30 compare membrane (d) at different heating stages and different
magnifications. When comparing Figures 7.19 to 7.24, it can be seen that there was an
increase in cluster size with an increase in heating time. Small clusters tended to agglomerate
to form bigger clusters. Some holes were visible in (d3) (Figures 7.21 and 7.24). As
agglomeration and segregation took place during prolonged heating at high temperatures, the
grain boundaries decreased. To compensate for this decrease, small pores or holes formed.
Figures 7.25 to 7.28 give very clear pictures of membrane changes during heating. After
50 hours of heating, there were no pores visible on the membrane surface. Some pores
started to develop after 100 hours of heating. In the middle of the left side of Figure 7.26
some holes can be seen. Membrane (d3), heated for 135 hours at 545°C and a further 25
hours at 600 °C, showed two different deterioration mechanisms. There were many pores
scattered all over the membrane surface (Figure 7.27). Furthermore, some porous clusters
started to develop. Two clusters can be seen in the middle left and middle right sides of
Figure 7.28. What is interesting is that those clusters had a large height dimension, many
times the thickness of the film. For the film to have had such an increase in height, the
increase in pore area had to be very large.
224
Figure 7.19: Membrane dl (25 OOOx) Figure 7.20: Membrane d2 (25 OOOX) Figure 7.21: Membrane d3 (25 OOOX)
Figure 7.22: Membrane dl (5000Ox) Figure 7.23: Membrane d2 (50 OOOX) Figure 7.24: Membrane d3 (50 OOOX)
225
Figure 7.25: Membrane dl (5 OOOX) Figure 7.26: Membrane d2 (5 OOOX) Figure 7.27: Membrane d3 (5 OOOx)
Figure 7.28: Membrane d3 (200Ox) Figure 7.29: Side view of d2 Figure 7.30: Side view of d3
226
Side view images (Figure 7.29 and 7.30) were taken to determine whether thermal expansion
caused the film to detach from the support. After 100 hours and after 170 hours, the figures
showed that film adhesion to the ceramic support remained strong.
Figures 7.32 and 7.33 show hydrogen and nitrogen permeances for membrane (NIx) heated
in hydrogen at 590°C, as a function of time. (NIx) had a thickness of 2.13 11m and a silver
content of22 wt % Ag. Unlike with (N4x), the hydrogen permeance decreased with time at
590°C (Figure 7.32).
227
6 ................................................................................................. 10
-~ 5.8t===~~~.==l··
. J .. ' ,.'
5.6 .'
9
8
Ui'
.;
~
[ ..",0::--/ - r - - - - - - - - - - - - t 6 E
5.2 i----,/,-;;
i 5+7,+'--~-----------------------+ 5 S-
f!r::
= 4.8 ttt--------;====;-----t 4
III
E 4.6 -1'-1---------1
1 ~h"'rog"" 11----+ 3 E
..1>.4.4 -r--I---------L1 .-...:.:nitroge=n..J-1_--t 2 8.
~
----=-
J: 4.2
N
+-+_______________+ 2
N
4+--·--r_--~--~----r_--._--_+ 0
o 5 10 15 20 25 30
Time (hI
~.5t--~~---------~
..
2S .'
240
'l 4 t--------'=-,-==~===l
,,r' .... >~. .---,.... .---~- ..--
~
."
'
~ 20
c ,; i
~ • IL ! 210 ]
= ~ N'
~
EEls
K"
! iI '. = N
1
,
. 180 Z
-2 10 . .2N
r:
2
t----------------i S L~' __ H
". . .
.. 150
l:
1
,·, ...-'1
0
o 2 4 , 10 0 2 4
• • 10
120
Figure 7.32: Effect of heating time in H2 Figure 7.33: Effect of heating time in H2
at 590 °C on H2 permeance for Nix at 590 °C on N2 permeance for Nix
The nitrogen permeance increased sharply (Figure 7.33), indicating that defects developed
quickly at 590 °C in a hydrogen atmosphere. More defects should favour an increase in
hydrogen permeance rather that a decrease in hydrogen permeance. The decrease in
hydrogen permeance confirmed that structural and morphological changes took place on the
metal alloy surface. There was a decrease in grain boundaries, which resulted in a decrease
in hydrogen permeance. Figure 7.34 shows the effect of temperature on the Pd-Ag film after
heat treatment in hydrogen at 590 0c. Both nitrogen and hydrogen permeances were poor
compared to the permeances for pure palladium films.
228
4.50 40
4.00
38
/
.. .,,; . .,
3.50
U
3.00 ---- /' --
36
U
,;
"E"' "-. 2.50 ~---- /_------ "
34 CIS CL
/
"
~ .§
N
E N'E
"'" -0E
N
2.00
/'
32:;
'"
N
0
::::
E
:I: .a 1.50
,/' 30 Z .s
1.00
1 ....... hydrogen I
0.50 I ...- nitrogen I
• 28
0.00 28
250 300 350 400 450 500 550 600
Permeance temperature (Oe)
With the introduction of argon, the hydrogen permeance started increasing at 500°C. Argon
did not cause passivation of the Pd-Ag surface, as was the case with nitrogen. Not many
229
defects fonned in the film at 500°C from 25 to 50 hours. There was only a very slight
increase in argon penneance during this period (Figure 7.37).
From 50 to 75 hours, there was a sharp increase in hydrogen penneance at 550°C in argon
(see Figure 7.35). Values increased from 3.0 jlmol/m2.Pa.s to 5.3 jlmol/m 2 Pa.s. Many
defects were introduced into the film during this time (see Figure 7.37). The stepwise
increase in argon flux indicates that heating in hydrogen was the main reason for the increase
in defects. Hydrogen was used for reduction at the following times and the argon penneances
were measured before and after hydrogen introduction:
• 50-51 hours,
• 55.5-56.5 hours,
• 63.5-64.5 hours, and
• 73.5-75 hours.
Figure 7.38 shows the effect of temperature on the membrane after 75 hours of heat
treatment.
..
6.0 25.0
•• 5.0
---- C. 20.0 { .-_.. -
0,
le 4.0 / .'"e
~ I °e 15.0
E
2.0
••
E
! 1.0 !l 5.0
,;: 2:
0.0
0 15 30
Time (h)
45 .. 75
0.0
0 5 10
Time (h)
15 20 25
Figure 7.35: Effect of heating time in N2 Figure 7.36: Effect of heating time in N2
and AI at 500 and 550°C on H2 at 500°C on N2 penneance for N2x
penneance for N2x
230
30.0 6.0
or,;
0,
26.0 / or
,; 5.0 t----'--------------,A
/
}:-T
~0
22.0
E
~
oS
~
0 16.0 ------------
••E ~
-
Co 14.0
«
.- .. ---~--------- ----_._-- --- k : ------_____ . _
8.
r."
1.0
Values in Table 7.7 were similar to those of (N3b) and (N4b), where no heat treatment was
perfonned. The exception was (N3c), which had a hydrogen penneance about twice that of
the other membranes. The difference between (N3c) and the other membranes was that after
50 hours of heat treatment in argon, (N3c) was oxidised at 320 °e and then reduced. The
other membranes were not re-oxidised after heat treatment.
231
This phenomenon was investigated further. Membrane (N4x) was oxidised after heat
treatment (see 7.4.3.1). (N8x) was prepared, heat treated for 10 hours in argon and then
oxidised. The effects of further oxidation on hydrogen permeance for those two membranes
are summarised in Table 7.8.
After oxidising and reducing the membranes agam (after heat treatment), the hydro,gen
permeances increased. In the case of (N8x), the H2 permeance more than doubled. On the
negative side, additional oxidation weakened the film and created defects. Two different
oxidation temperatures and times were tested. Oxidation at 350°C was too severe.
Oxidation for 1 hour at 310 °C yielded much better results. Further oxidation and reduction
changed the surface morphology and structure of the film. These changes promoted
hydrogen movement through the film.
N2 permeance Nl permeance
Before oxidation After oxidation
(nmoIlml.Pa.s) (nmoIlm2.Pa.s)
N4x 350 3 9.6 24.5
N8x 310 1 13.1 18.2
232
• Heat up to 400°C in argon, and
• Reduce in hydrogen for 2 hours.
The stability of the Pd-Ag membranes at temperatures from 200°C to 300 °C were monitored
during the catalytic membrane reactor experiments and will be mentioned in Chapter 9.
5.'
__ HyG'ogen perm-.c:e
28 '.5
"
..- ;;
.•
-+--Hylt'ogM pIIrmunDII
if OJ '.0 ,
l, "
•.,
- tI- Nltragen pBIIINnCl!
.'
, ..-
27 &: ,
- tI- Nllrvslen plIITTI_nce
.// l
".'"••
/'/'
1; ,
1; 5.5
'"E•
,
,,
'"~
"
" -_.¥'. 21 ~
" / ......- 5.0
"
~
.; "
'" s. .; /' oS
3.5
" •uc ~- ... .:~--
3
c
I 3.'
.. "
./
:<
"~
"
" ,
,
3c
•• =
~
4.5
, ~/
.'
1.
:
=
•
~
7 E
I
"
4.0
1 2.5
,
, ,
25
•... &
:i! 3.5 ./
..-
:£
%: -£
3.0 1.5
2.0
250 290 370 " 250 290 330 370 410
'"
TrC)
410
T I'C)
Figure 7.39: H2 and Nl permeances for N4x Figure 7.40: Hl and Nl permeances for N8x
Calculated values correlated well with measured values when a sweep gas was employed
(Figures 7.41 and 7.42). For membrane (N8x), the error was typically less than 5%, but for
N4x it was sometimes slightly larger. This was due to the higher N2 permeance of (N4x).
233
100 100
~ 95
c 1-- ~ •c "
0 .---.-~
:"
•• •• 90
i 90 '.
A'" E / .. •
.-.---.--.-~
/. .
eo
~ ~
L .
. '"
!l
•
i
1;1
75
/.
/. .. ~ L
"5 70 ----
] 0
1i
~ ~
U•
75
70 V . U
60
60 65 70 75 80
----" --
85 90 95 100
70 75 80 85 go 95 100
Measured V. H3 permeance Measured % H2 permeance
Figure 7.41: Measured vs. calculated Figure 7.42: Measured vs. calculated
% H2 permeance for N4x % H2 permeance for N8x
7.6. SUMMARY
Pd-Ag membranes of less than 2.2 IJlIl thickness were prepared by successive electroless
plating steps. The permeances of those films at 410°C ranged from 3.0 to 5.5 lJlIlol/m2.Pa.s
before heat treatment. That was significantly lower than the values for pure Pd films.
Different heat treatment methods were employed to improve the hydrogen permeance of the
Pd-Ag film. In all the gas environments tested, the films weakened at and above 590°C.
Hydrogen created defects in the film at a moderate rate at 550 °C and at a fast rate at 590°C.
Furthermore, hydrogen passivated the film at 590 °C but not at 540°C. Film passivation
resulted in a decline in hydrogen permeance. Continuous thermal cycling contributed
towards film defection.
Some defects formed in argon at 545°C after 100 hours of heating, but very few defects
formed at 500°C in either nitrogen and argon. Nitrogen also passivated the film and should
not be used for heat treatment. The best conditions for heat treatment were:
• an argon environment,
• at a temperature of 550 °C, and
• a heating time of about 10 to 15 hours.
234
After heat treatment, the film has to be oxidised at 310 °C for not more than 1 hour and then
reduced to obtain the best combination of hydrogen permeance and selectivity. This
additional oxidation-reduction step resulted in a great improvement in hydrogen permeance.
This phenomenon has not been observed previously by any other research group.
235
8. CATALYTIC MEMBRANE REACTOR MODELLING
This chapter will focus on the development of a catalytic membrane reactor model. All
aspects of the model will be investigated and a complete sensitivity analysis will be
performed to determine the effect of each parameter on the performance of the reactor. The
effect of the feed rate on reaction conversion is discussed. Experimental data for a plug flow
reactor at different feed flow rates are presented.
~
equilibrium
L--
• •
1 2
•
•
Feed rate ---4
+ - - - Space time
In the present study high feed rates were employed to determine the practical importance of a
catalytic membrane reactor. Experiments were performed in the region marked by block 2.
236
If a catalytic membrane reactor can significantly improve reaction conversion when the feed
rate is fast enough not to limit the reaction by the equilibrium conversion, then the reactor
becomes of practical importance. Table 8.1 summarises some experimental conditions
employed by other researchers for dehydrogenation reactions with liquid feeds in a
membrane reactor. The feed rates used in the present study are included, to illustrate the
where the present study 'fits in'.
Table 8.1: Different liquid hourly space volumes (LHSV) used for dehydrogenation
reactions in a membrane reactor
Dehydrogenation Temperature range LHSV (h) Reference
reaction (0 C)
237
To simplify calculations, an average of 4 was assumed for modelling. This value was
different from what would be expected using the Knudsen theory, which suggests a nitrogen
to ethanol selectivity of about 1.3. Other transport effects through the defects in the
membrane occurred simultaneously with Knudsen flow. Since the measured nitrogen to
ethanol selectivity was much larger than the Knudsen value it shows that many defects were
smaller than the Knudsen defects. The presence of a large amount of molecular sieving
defects would explain the difference between the nitrogen and ethanol permeances. The
ethanol molecules were too large to pass through many defects where nitrogen molecules
could pass through.
The permeances of other organic species were not measured. The worst case scenario was
used for determining permeances for modelling. The selectivities were, taken as the ratio
between the molecular masses of each species (Knudsen theory) relative to ethanol. This
could be expressed as:
The permeance for 2-butanol will thus be 1.27 times lower than that of ethanol.
~
2"'"
2300
r
+-_________f • "2,N2
~
L
~ '"
• "2j-_~\-_--
!,.+-------\\c-------
J
1800
I- H2:.tlwloI r-
1300
! ,.+---------''r-\----
iii ,.,+-_______~..._---_
'---...
80IJ
300
,.,+--------------
320
"" 360 380 400
Temperature ('C)
420 440 460
". "" .20
238
8.3. DEVELOPING A PROCESS MODEL
A preliminary model was constructed to mathematically describe the dehydrogenation of
alcohols in a catalytic membrane reactor. The process used a nitrogen sweep gas to create a
partial pressure differential across the membrane. The flow of the sweep gas was co-current
to that of the feed. A practical sweep gas in industry could be steam, which would supply
heat to the endothermic dehydrogenation reaction and create the hydrogen concentration
gradient across the membrane. Alternatively, the shell side of the reactor could operate under
vacuum conditions.
Both the shell and the tube sides were operated under atmospheric conditions. The pressure
drop across the length of the packed membrane tube (length of catalyst bed = 20 cm) was less
than 15 mbar. Isobaric conditions along the membrane's axis was therefore a good
assumption. To incorporate an axial pressure drop, the Ergun equation (8.2) could be used
(Fogler, 1992).
(8.2)
where
P = pressure [Pal
= void fraction of packed bed
239
gc = conversion factor (1 for SI units)
Dp = particle diameter [m]
Jl = gas mixture viscosity [kg/ms]
L = length down the tube [m]
U = empty column velocity [mls]
p = gas density [kg/m3 ]
G = empty column mass velocity = pU [kg/m2s]
For porous membranes (non palladium based) exhibiting Knudsen diffusion, the penneance
equation cannot be used and a radial pressure drop model must be employed (Deckman et a!.,
1995).
Radial concentration models have been studied for a catalytic membrane reactor (Becker et
a!., 1993; Gobina et aI., 1995a-c). Gobina et al. (1995b) found that the radial concentration
change was negligible for a membrane with a 7.8 mm inner diameter.
The final assumption of steady state was not very accurate, but was made to simplify
calculations. The data was gathered within a 24 hour period to give a better approximation of
steady state. Thereafter, both the catalyst and the membrane were re-activated before the
next set of tests. Non-steady state is due to both catalyst and membrane deactivation with
time, which influences both the kinetics and the separation aspects of the process.
240
(8.4)
mainly
nitrogen sweep hydrogen and
penneation nitrogen
alcohol - - ---
-- .
_ reaction - _
_
reaction _
alcohol +
- --_--» products
) mainly
nitrogen sweep penneation hydrogen and
nitrogen
dQ; =V[~J]
L
d-
R'
m
(8.5)
La
In (8.4), i = alcohol feed, reaction products and hydrogen, while in (8.5), i = nitrogen sweep
gas. The boundary conditions for co-current flow are:
At L= 0, F; = F;o. (8.6)
At L= 0, Qi = Q;o. (8.7)
The flux equations were derived in section 6.6.5. For equations (6.1) to (6.5), the right hand
side of the equations represents the penneance component, expressed in tenns of feed flow
rates.
241
A kinetic expression for ethanol dehydrogenation was derived in equation (5.21), with
optimised parameters in Table 5.6. The 2-butanol dehydrogenation expression was given in
equation (5.26) with optimised parameters in Table 5.8. The permeance and kinetic
expressions will be summarised in a later section.
In most instances, correlations do not exist to predict properties of ternary and higher order
mixtures. Furthermore, the available correlations may contain significant errors.
(8.8)
242
Cb = void fraction of packed bed
y = shape factor (external area divided by 7rD/)
JJ. = viscosity [kg/m.s]
p. = fluid density [kg/m 3 ]
u = JJ./p = kinematic viscosity [m 2/s]
DA,m = gas phase diffusivity of A in bulk fluid (mixture) [m 2 /s]
U = Empty column velocity [rnis]
The equation is valid for: 0.25 < Cb < 0.5; 40 < Re' < 4000 and 1 < Sc < 4000, with
DpU
Re=-- (8.9)
v
v
Sc=-- (8.10)
DAm
Re
Re'= (8.11)
(1- Gb)Y
ShGb
Sh'= (8.13)
(I-Gb)Y
243
mixtures (Toor, 1957; Duncan and Toor, 1962) and multi-component gas mixtures (Curtiss
and Hirschfelder, 1949; Wilke, 1950; Fairbanks and Wilke, 1950) have been performed. In
the simplest form, the diffusion coefficient can be estimated by the following:
1- Y A
D A •m = Y Y Y (8.14)
_B_ + _C_ + _D_ + ...
DAB D AC DAD
For laboratory reactors with low flow rates and thus low Reynolds numbers, the Thoenes -
Kramers equation is invalid. Satterfield (1980) listed some alternative correlations for
determining the mass transfer coefficient.
(8.15)
(8.16)
The value of the constant, C 1, differs depending on the literature source. Satterfield (1980)
suggested a value of 0.07, while Collins (1993b) used a value of 0.015, determined from
experimental data.
With kc known, the Mears' criterion can be used to determine whether external diffusion is
rate limiting (Fogler, 1992). When
(8.17)
244
where
n = reaction order
Dp = particle diameter [m]
Pb = bulk density of catalyst bed [kg/m l ]
CA,b = bulk concentration of A [kmoVml ]
ko = mass transfer coefficient [mls]
then external mass transfer effects can be neglected. It is clear from (8.17) that external mass
transfer will become more important as the feed concentration drops along the membrane
axis and CA,b decreases.
To link the reaction rate coefficient and the mass transfer coefficient, a mass balance can be
performed around the outside of the catalyst surface.
(8.18)
where
Ap = external surface area of the catalyst
= 6m(1- eb)fPbD p [m2]
m = mass of catalyst [kg]
r' A,surf the surface reaction rate [moVkg.catalyst.s)
= the reaction rate in the absence of any mass transfer limitations
1] = effectiveness factor
The bulk concentration of the reactant is known, but not the surface concentration or the
surface rate. Equation (8.18) or (8.19) must be used to determine the surface concentration or
surface pressure in terms of the bulk concentration or bulk pressure. The difference between
245
the two is due to the external mass transfer resistance. When no external mass transfer
resistance is present, the surface concentration is equal to the bulk concentration. It is thus
very important to obtain kinetic data free from mass transfer limitations, because if mass
transfer limitations are present, the measured rate will be different from the surface rate and
calculations will be faulty.
In the cases of ethanol and 2-butanol dehydrogenation, the reaction rate on the right hand side
of (8.18) or (8.19) must be replaced by either equation (5.21) or (5.26): The surface
concentration or pressure of the alcohol must then be detennined in tenns of the overall
effectiveness factor and the bulk concentrations or bulk pressures of all species. The
dehydrogenation rate equations include surface partial pressures for the feed as well as for the
products. As a first approximation, it can be assumed that the interphase mass transfer
resistance of all the species to reach the surface is the same, and then the following
simplification can be made:
p
p _ A,surf P
i.surf - P i.b
(8.20)
A.b
Equation (8.20) is substituted into the rate equations (5.21 or 5.26) and the rate equation is
then substituted into (8.19). The result is an equation with three unknowns: kc, 11 and PA,sun'
Mathematically PA,surf can be expressed by:
(8.21)
To solve equation (8.21), k, kc and 11 must first be detennined. All the parameters can be
obtained from experimental data by perfonning experiments in the absence of intraparticle
mass transfer resistance (see 8.3.4).
246
reaction rate and the rate that exists on the catalyst surface. Both the effects of interphase and
intraparticle mass transfer resistance are incorporated into TJ.
The catalyst particle is modelled as a sphere. The concentration of the reactant declines from
the outside surface into the pores of the particle, The following expression can be used to
describe the radial concentration profile in the particle (Fogler, 1992).
d[-D, ~(r')2]
(8.23)
dr'
For non-reversible nth order reactions, the equation can be written in dimensionless form and
solved algebraically (Fogler, 1992). For complex reactions, the problem must be solved
247
munerically. Collins (1993b) formulated equations for solving a reversible reaction. In
dimensionless form the following equation was obtained:
(8.24)
and
(8.26)
Equations (8.24) to (8.26) apply for both the dehydrogenation of ethanol and 2-butanol.
Boundary conditions and material balances for each speCIes In the reaction have been
summarised by Collins (1993b). The effectiveness factor, 11, is the actual overall rate of
reaction divided by the rate of reaction at bulk conditions (CA,b)' Some authors define 11 as
the actual reaction rate divided by the rate that would result if the entire interior surface was
exposed to the surface conditions (Fogler, 1992, calls it an internal effectiveness factor). This
definition was, however, not used by Collins (1993b). He defined 11 as an overall
effectiveness factor. The effectiveness factor at a specific axial position along the
membrane's axis can be expressed by:
(8.27)
248
Equation (8.27) must be evaluated numerically. Bulk conditions are used for solving
equation (8.27). With the value of 11 known, the surface concentrations or partial pressures
can be calculated with equation 8.21. With the surface properties known, the surface reaction
rate can be calculated by using equation (8.18) or (8.19). Finally the reaction rate at bulk
conditions can be calculated:
The main goal of the modelling exercise was to calculate the value of 11. The surface reaction
rate could be measured under special conditions and once 11 was known, the reaction rate at
any bulk conditions could be calculated.
The Weisz-Prater criterion (Fogler, 1992) for internal diffusion can be used to detennine
whether there are significant internal diffusion limitations. The Weisz-Prater parameter is
define by:
(8.28)
249
1. Detennine the dimensionless parameters, Re and Sc.
2. Calculate the mass transfer coefficient, kc, using one of the equations (8.8), (8.15) or
(8.16). In the case of (8.16), C1 will have to be detennined experimentally.
3. Detennine whether the external mass transfer is rate limiting by testing Mears' criterion
(equation 8.17).
4. Calculate the effective diffusion coefficient defined by Satterfield (1980).
5. If external mass transfer is not rate limiting, then surface conditions can be taken as bulk
conditions and the Weisz-Prater criterion can be tested to detennine if internal mass
transfer is limiting. If both the interphase and itraparticle mass transfer resistances are
negligible, then the effectiveness factor equals one.
6. If external mass transfer is rate limiting then the Weisz-Prater criterion cannot be tested
since the surface concentrations are not known. The reactant surface concentration or
pressure must be calculated in tenns ofT] by using equation (8.18) or (8.19).
7. The effectiveness factor must be detennined from equation (8.27).
8. With kc (after point 3), PA,surf (from point 6) and T] (from point 7) known, the surface
reaction in equation (8.19) can be calculated.
9. The reaction rate using the bulk conditions can now be detennine from equation (8.22),
since both the effectiveness factor and the surface reaction rate are known.
250
determined by Wilke-Lee's (Millat et aI., 1996) method and the ethanol diffusion coefficient
in the mixture was determined by Wilke's method (Fairbanks and Wilke, 1950). A
simplification was made for determining mixture properties. The mixture was assumed to
only consist of the three main components, ethanol, acetaldehyde and hydrogen. The ethanol
selectivity towards acetaldehyde production was close to 100% for kinetic experiments
(Chapter 5) and thus the assumption was very good. Figure 5.9 was taken from Chapter 5
to show the effects of interphase mass transfer resistance. Figure 8.5 converts data from
Figure 5.9 to a plot of the effectiveness factor vs. Reynolds number. The surface reaction
rates (r'A,,) used for calculating the effectiveness factors (see Figure 5.9) are listed in Table
8.2.
1.20
1.00 +------,--.,----:;=""""------1
'ff?
ooo+__------~~~-----------
)_o~+------~~~~·-/-··-·--T-=-~-O-.c-w-~-o-.c-----1
/
, .., +__--i#"-------------------
If
O~+__~---------------------
,~~~;~~.~:~.---------.~----------
, ,.
000+----__- -__- -__- -__- -__- - _
5 10 15
_flow_I....)
2D 30 O.OOQ 0.020 0.040 0.080
~Id.
0.080
ftlilllHr
0.100
'12'
Figure 5.9: The effect offeed flow rate on Figure 8.5: Effect of Reynolds number on
acetaldehyde production rate effectiveness factor
251
For the low Reynolds numbers in Figure 8.5, the format of equation (8.16) is best suited to
determine the mass transfer coefficient as a function of feed rate, but C 1 is unknown.
Equation (8.19) is also valid, but in this case Cs is unknown and kc cannot be determined.
D
k ,=C~R
I e (8.16)
Dp
(8.19)
(8.29)
In the region of strong interphase mass transfer resistance, CA.surf «< CA.b. As the Reynolds
number increases, the surface concentration will start increasing. An assumption was made
that CA,,,," «< CA,b in the region where TJ < 0.8. Equation (8.19) could be solved at different
temperatures and different Reynolds numbers. In Figure 8.6, the constant in equation (8.16)
is plotted against inverse Reynolds number at different temperatures.
0.045 0.02
0.04
0.035
7A--=-....-::~
~ - 0,016
0.03
..... 111"'0'---, 0.012
0.025
U T = 200 ·Cto 3OO·C U T =200 ·C to 300 ·C
0.02
0.'"
0.015
0,01 0,004
0.005
0
0 20 40 60 60 100 120
0
, 10 15 ~ ~ ~ ~ ~ ~
Inverse Re Inwrs.Ra"'1
Figure 8.6: C l vs. inverse Re at different Figure 8.7: C l vs. inverse Reo. 8l at different
temperatures temperatures
252
The value of C I is not supposed to change when the temperature and/or the feed flow changes
(C I is a constant). Figure 8.6 indicates that CI was not constant at different inverse Reynolds
numbers. A much better correlation was obtained when C I was plotted against the inverse of
the Reynolds number to the power of 0.81 (Figure 8.7). Equation (8.16) reduces to:
(8.30)
253
34 ....••.........•••••.•.•.•.••..•...•.•...•..•.•.•...•....••.•.....•.•....•.•..•. _ ......••.• 50 ............................................................... __ •.....•......•......•...•..
~t_--------~.~---------------
"
... . ... . ... .•. . . . -.......• .. t_------~.--------~--------
•
1
30
~~t_--------~~---------------
•
8 •
f ,. t1rI=:H;::;...:;:...=""'==:=~="=••::.~===:;---------
i 24 H .EthII\OI conwnlonto .~ •
• ~lIbrium cortYWIion to -mldahydll
~~============~---------
~~--------------------------~ O.OE1-00 3.0E-05 8.0E-05 1.2E-D4 1.5E-04
O.OE+OO 2.0E-D5 4.0E-D5
F(moIIs) F(moVs)
Figure 8.8: Plug flow reactor data for Figure 8.9: Plug flow reactor data for
ethanol dehydrogenation at 250°C ethanol dehydrogenation at 275°C
eo 100
" •
j
~"
•
8
-_ .. _. - --.....• -_ .. _ .. _ ... -_
•
..
•0 ..
'I! •
! .. r-- • TalaIlIIh.,GI con......lon
• Btl.,'" ~onto~ya.
•
..
O.OE+OO
.t.~llbrIumCQ"""""tJI'Ito~
Figure 8.10: Plug flow reactor data for Figure 8.11: Selectivity towards acetaldehyde
ethanol dehydrogenation at 300°C production for a plug flow reactor
Figure 8.11 shows selectivity towards acetaldehyde production at different temperatures and
feed rates. The selectivity dropped at the lower feed rates as more unwanted products were
formed. Selectivity was less dependent on temperatures than on reaction time. Side products
were formed due to the reaction between acetaldehyde and ethanol. Acetic acid ethyl ester
was one of the main by products. Further products that formed were due to acetal and
hemiacetal formation (Streitwieser et aI., 1992). The acetal, 1,1'-diethoxyethane and the
dehydrated acetal, ethoxyethene were present in small quantities in most experiments.
Butanals were also present in low concentrations. These products were similar to those
observed by Raich and Foley (1998).
254
At higher feed rates or shorter catalyst contact times, the reaction time was too short for the
C4 by products to form and the selectivity towards acetaldehyde increased. Under
differential conditions, the selectivity of ethanol dehydrogenation towards acetaldehyde was
close to 100"10 (Chapter 5).
0.0800
0.0700t====:~~==
F=-=
11.0Il00
----- - ------
0.0500+---------------=
~ O.IMOO +---IEI'"h"'an"'oIr;;
.....
;;;;;;;;.iiiiOftii•.,.~I.""
...... ~._
.• "' .
, .,.----
= _____
+-___---""''":::.m;;;:.DP'''to'''b=......
D.....,
D."""'+--------------
0.0100+--------------
o.ooool--_--_--_--_-_
200 280 300
Figure 8.14: Re as a function of temperature and conversion (feed rate = 1.43*10-4 moVs)
255
At all tested feed flow rates, the Reynolds numbers were less than 0.07 and thus in the region
of strong interphase mass transfer resistance according to Figure 8.5.
•• ~ _. M ~ __________ • _. -.-A
~t_--~----_r-----------------
a 7Or---~--------------~----y_-
•
II~t_--------------------------- j"r----"----"-----"----.-----------
• Tat:ll2-butlnol CIOIlWrIIon
.2-Bubrlol conVlQlon 10 MEK
" r 1r=;.T;...;......
•
55
;;;.;~;n.;.!III;oo=::;-------------
.2..eutmoI CIIInWI-.Ian to MEK
" EquHlbriwn COIIYIIrSkrI to MEK
.. E~ CCIIIWI'SICf'I to MEK
0.0&00 2l1E-05 ...-os 0..... O.OEtOO 2.CE-05 •. IE-05 e.Df-05 a,CEo05 1./JE-04
F(molls) F(molls)
Figure 8.15: Plug flow reactor data for Figure 8.16: Plug flow reactor data for
2-butanol dehydrogenation at 190°C 2-butanol dehydrogenation at 215°C
!
82
" ''''' • • •
" .............
............ _-_ !
iao
c
8
• a
.i
~
> • " •
r
•
,.
•
•
.. Tahll2-butw101 ~on
.2-1kitmoI currvw-!lon to MEK
.. ~IbI'I\rnCCll'l\ltlnlontoMEK
:l
.. " !
I
.T= 190'C
.T=215"C
... T"240"C
9.
"O.OE+OO 2JlE-OS ""DE-OS &'oE-OS 0.06-05 1.0E-G4 '.2!;<4 ,...... O_OOE+OO 4_00E-05 B.OOE-05 1.Z0E-04 1.60E-04
F(molls) F (maUs)
Figure 8.17: Plug flow reactor data for Figure 8.18: Selectivity towards MEK
2-butanol dehydrogenation at 240°C production for a plug flow reactor
The measured conversions at 190°C (Figure 8.15) and at 215 ac (Figure 8.16) declined with
an increase in the feed flow rate. The reaction time became too short at the higher flow rates
256
for full conversion to take place. At 240°C, the reaction rate was very fast compared to the
rate at the lower temperatures. Even at the maximum feed flow rates tested, the 2-butanol
conversion was still close to that of the equilibrium value and did not decline significantly.
Selectivity for this reaction was very high. Figure 8.18 indicates that the selectivity was
more than 96% for all experiments conducted. The main by-product was 3-octanol.
At 300°C:
11 = -224.00*Re3 - 69.20*Re2 + 18.60*Re + 0.053 for 0.01 < Re:S 0.08 (8.31)
11 = 1 for Re > 0.08
At 275°C:
11 = -241.18*Re 3 - 36.65*Re2 + 15.89*Re + 0.039 for 0.01 < Re:S 0.085 (8.32)
11 = 1 for Re > 0.085
At 250°C:
11 = 363.16*Re3 - 189.62*Re2 + 26.34*Re - 0.080 for 0.01 < Re:S 0.08 (8.33)
11 = 1 for Re > 0.08
It should be emphasised that these effectiveness factors will only be true for small particles
without intraparticle mass transfer resistance. For larger particles, equation (8.27) must be
solved using the kc value determined in equation (8.30). The Reynolds number is:
(8.34)
257
u = 8.314T(FEt.t + FH"t + FA"t + FN"t)
(8.35)
Pt(nR~)
with
with (8.37)
(8.38)
(8.39)
For hydrogen:
with (8.40)
(8.41)
258
For acetaldehyde:
with (8.42)
(8.43)
For nitrogen:
(8.44)
dFN", =2nR L P [P -P
d( ~o) m 0 m,N, N,,' N,,!
(8.45)
For ethanol:
(8.46)
For hydrogen:
d(dF 2
H "
~)
2nR LP
m 0 m.H,
[
P 1
-P
H,.! H 2 .,
(8.47)
For acetaldehyde:
(8.48)
259
The initial conditions were:
AtL=O: FE.,. = FE',t(O) FE." = 0
FH ,.! = 0 FH , = 0
"
F Ac ,. = 0 FAc ,. = 0
FN,.t = 0
(8.49)
(8.50)
(8.51)
(8.52)
(8.53)
(8.54)
P =p FH", ] (8.55)
H,.. !,' [ F
Et,s
+F
Ac,s
+F
H1,s
+F
N 1,5
P =p FN". ] (8.56)
N,.. !,' [ F
Et,s
+F
Ac.s
+F
H 2 .S
+F
N 2 ,s
260
To solve the model, kinetic coefficients and gas permeances must be known. Membrane
(N8x) was used for catalytic membrane reactor experiments. Permeance properties of the
membrane are listed in Appendix D2. Measured reaction rate coefficients will be used
instead of fitted values (from Chapter 5) at 250 °e, 275 °e and 300 °e, respectively. Table
8.3 lists reaction rate coefficients. Reaction rate coefficients were taken from Table 5.5 and
changed to standard units i.e. [Pa], [Pa'!] and [mol/kg.cat.s.Pa]. The permeance data is listed
in Table 8.4 and further data necessary to solve the model is given in Table 8.5.
261
8.4.4. SUMMARY OF MODEL FOR 2-BUTANOL DEHYDROGENATION
The model fOf 2-butanol dehydrogenation is similar to that of the ethanol dehydrogenation
model. Figure 5.10 shows that the dehydrogenation rate changes very little with feed flow
rate, indicating that the interphase mass transfer resistance is small compared to the reaction
rate. The effectiveness factor can be taken as 1 and Reynolds numbers do not have to be
calculated. The model can be expressed as:
dF'Bu,
d(~') =
'v[·
- f
Pb 'Bu"b77 - Rm2 J 'Bu' ] with (8.57)
Lo .
(8.58)
(8.59)
For hydrogen:
with (8.60)
(8.61)
ForMEK:
(8.62)
262
(8.63)
For nitrogen:
(8.64)
(8.65)
For 2-butanol:
(8.66)
For hydrogen:
d(elFtH",J=2nR L P
m 0 m,H,
[
H"I
l
P -P
H",
(8.67)
ForMEK:
(8.68)
FH "I =0
263
FMEK,t = 0 FMEK,t = 0
FN , ,I =0
(8.69)
(8.70)
(8.71)
P =P FN"I ] (8.72)
N"t ![ (F'BUI,! + FMEK,! + FH,,') + (FN"s(O) - FN,,,)
(8.75)
P
Nz's
=P
t,s
FN,,' ] (8.76)
[ F2But" + FMEK" + FH", + FN",
Table 8.6 lists reaction rate coefficients. Values were taken from Table 5.7 and changed to
standard units i,e. [Pal. [Pa· l ] and [mol/kg.cat.s.Pa]. Membrane (N8x) was used for
experiments. The permeances were determined from data in Appendix D2. Values were
264
estimated at 190°C, 215 °C and 240°C, respectively, by fitting an equation through the
hydrogen and nitrogen permeance data for membrane (N8x) .
Table 8.7: Permeance data for N8x at 190, 215 and 240°C
T (0C) H2 permeance N2 permeance 2-Butanol MEK permeance
(Ilmollm 2.Pa.s) (nmollm2.Pa.s) permeance (nmollm2.Pa.s)
(nmollm 2.Pa.s)
190 2.40 26.61 5.25 5.25
215 2.81 24.80 4.89 4.89
240 3.22 23.21 4.57 4.57
The other parameters necessary to solve the model are listed in Table 8.8.
265
8.5. SELECTIVITY ANALYSIS OF THE PROCESS MODEL
The standard values are the measured reaction rate and membrane parameters for ethanol
dehydrogenation at 275°C (Table 8.9).
Ke q [Pal 24080
1
"
Pm.hydrogcn [J.UlloIlm" .Pa.s] 3.78
Pm,nitrogen [nmoIlm".Pa.s] 21.29
266
The performance of the membrane reactor can be judged by several factors. The most
important factors are the exit alcohol conversion, the fraction of ethanol feed lost through the
membrane, the fraction of produced hydrogen permeated to the shell side and the purity of
the hydrogen on the shell side. Process conditions, reaction parameters and membrane
parameters that yield high alcohol conversion and pure hydrogen on the shell side, represents
optimum conditions. The effect of each parameter on the dehydrogenation of ethanol at
275°C is presented graphically in Appendix G. It should be kept in mind that the
equilibrium ethanol conversion for a plug flow reactor is 40% at 275°C and 1 atmosphere
total pressure.
Model equations (8.37) to (8.56) were solved numerically with Euler's method. A very small
step length of 0.001 (LILo)was employed to obtain results of high accuracy.
In the very first part of the reactor, the hydrogen production rate is much faster than the
permeance rate. The driving force shoots up from zero to a maximum value. The faster the
rate (larger the k' -value), the higher the maximum hydrogen driving force. The hydrogen
driving force then drops quickly to very low values further down the axis of the membrane.
At different k' -values there is almost no change in the percentage of produced hydrogen that
permeates to the shell side. The percentage of produced hydrogen that permeates to the shell
side, increases sharply in the first part of the reactor and then declines slowly as back
267
diffusion takes place. The ethanol flow rate on the tube side also drops quicker along the
membrane's length as the k'-value increases.
An optimum k' -value would be about four times the standard rate. Higher values will not
improve the ethanol conversion significantly.
The sweep gas permeance (nitrogen) is much higher than the permeances for ethanol and
acetaldehyde. A fraction of the nitrogen sweep gas will permeate from the shell side to the
tube side and this fraction is much larger than the amount of ethanol and acetaldehyde that
will permeate from the tube side to the shell side. The nitrogen partial pressure will drop on
the shell side and increase on the tube side. Once the produced hydrogen reaches a maximum
amount permeated, no more hydrogen goes from the tube side to the shell side and back
diffusion of hydrogen will commence. The hydrogen partial pressure on the shell side will
increase due to a decline in nitrogen partial pressure, which is larger than the combined
increase in ethanol and acetaldehyde partial pressures on the shell side. Similarly, the
hydrogen partial pressure on the tube side will decrease due to an increase in nitrogen partial
pressure. The hydrogen driving force is reversed and hydrogen will permeate back from the
shell side to the tube side.
268
a lower Kacer, but remains at acceptable levels for all Kacet-values. The variance in ethanol
conversion for changes in K acet is much smaller than the variance when k is changed. For
Kaceevalues ofless than five times the standard value, the improvement in ethanol conversion
becomes very small for a further decline in Kacet.
Hydrogen purity on the shell side varies between 85% and 88%. On the tube side the
nitrogen content is constant at 10%. The initial hydrogen driving force is larger for small
Kacet-values. The driving force either goes to very low values or to below zero along the axis
of the membrane. The lack of a strong driving force results in a low hydrogen percentage
permeated to the shell side. The value is 45% and changes little with changes in the value of
K acet . This result is similar to that obtained for changing the k' -value.
An optimum Kacecvalue would be at least 5 times smaller than the standard value.
The nitrogen fraction on the tube side increases from 5% at a selectivity of 320 to over 40%
at a selectivity of 10. With such an increase in nitrogen content on the tube side, a significant
amount of back diffusion can be expected. This is indeed the case and the hydrogen driving
force becomes negative at low dimensionless length values for low selectivities. The low
percentage of produced hydrogen that permeates to the shell side, indicates very strong back
269
diffusion. At a selectivity of 10, only 20% of the produced hydrogen permeates to the shell
side and this percentage increases to 47% at a selectivity of320.
A selectivity of at least 150 must be used to ensure that not more than 3% of the ethanol feed
is lost through the defects in the Pd-Ag film.
The best results will be obtained with the maximum value of the effectiveness factor, i.e.
when the effectiveness factor approaches one.
The ethanol fraction at the tube side exit increases with an increase in feed flow rate due to a
decrease in conversion. At the higher flow rates, a small hydrogen driving force remains
over the full length of the membrane and a maximum percentage (47%) of produced
hydrogen permeates to the shell side. At a low flow rate (0.5 multiple), back diffusion takes
270
place in the second half of the reactor. The hydrogen driving force shifts along the axis of the
membrane when the feed rate increases.
The value of the feed rate should be determined by all other parameters. The quality of the
membrane and the speed of the reaction rate will determine which feed rate is best suited for
the model. For example, if the reaction rate is very slow, it is of no use to employ a very fast
feed rate, since the benefits of the membrane reactor will be cancelled.
Ethanol conversion increases from 42% to almost 50% for sweep gas to feed ratios of 0.25 to
8. The ethanol feed lost through the membrane increases slightly, because of a higher ethanol
driving force, but it remains at acceptably low values. A hydrogen purity of above 90% is
obtained on the shell side for sweep to feed ratios greater than 4. There is a sharp decline in
hydrogen content (from 25% to 5%) on the tube side with an increase in sweep rate. The
acetaldehyde fraction increases on the tube side as more ethanol is converted and the nitrogen
fraction increases as more nitrogen permeates through the membrane at higher sweep rates.
A higher sweep rate reduces the hydrogen partial pressure on the shell side. The result is a
higher hydrogen driving force and a higher percentage of produced hydrogen that permeates
the membrane. A high sweep rate is very effective for extracting both pure, and a very high
percentage of, produced hydrogen. The percentage of produced hydrogen permeated,
increases from less than 15% at a sweep to feed rate of 0.25 to over 85% for a sweep to feed
rate of8.
271
A sweep to feed ratio of at least 4 should be used, but this depends to a large extent on the
feed rate and must be chosen in conjunction with the feed rate.
When comparing the data in Appendix G.3 and Appendix G.7, it is clear that the membrane
selectivity plays a more important role in the membrane reactor's performance than the
hydrogen permeance does. This statement is true under standard conditions and in the ranges
that each parameter was investigated. With the standard selectivity, the hydrogen permeance
should not move outside the range of 0.5 to 1.5 times that of the standard permeance.
272
(Appendix G.8). The overall ethanol conversion in Appendix G.8 is similar to that in
Appendix G.7, but there is very little change in ethanol feed lost through the membrane. The
hydrogen composition on the shell side exit remains constant at just above 88%. The
composition of the tube side exit mixture also varies little with changes in the hydrogen
permeance. The hydrogen driving forces in Appendix G.8 are similar to those in Appendix
G.7, but less back diffusion of hydrogen occurs, since less nitrogen permeates from the shell
side to the tube side.
The percentage of produced hydrogen that permeates from the tube side to the shell side
starts declining at hydrogen permeances of less than 0.5 times the standard value. At lower
hydrogen permeances, the reactor length is too short to reach equilibrium for hydrogen
permeance. The result is a decline in hydrogen permeance as compared to when equilibrium
is reached.
At standard selectivity, the hydrogen permeance should be at least one half or more of the
standard hydrogen permeance for optimum membrane performance.
273
In both cases the ethanol conversIOn reaches a mronmum of 64% when the hydrogen
permeance is increased. The gain in ethanol conversion is small for hydrogen permeances
larger that 0.5 times the standard value. At smaller values the ethanol conversion will drop
sharply towards the equilibrium value or lower in the case of higher feed rates. The
equilibrium ethanol conversion at 275 °C is 40%. Hence the membrane reactor offers a
significant improvement in ethanol conversion. This improvement is more than 50%.
The second set of conditions represents a slight improvement in catalyst activity. Membrane
parameters are similar to those of the optimised membrane. For the standard conditions, the
ethanol conversion is 44.5% and the ethanol feed lost through the membrane is 2.33%. For
the second set of conditions, the ethanol conversion is given as a function of sweep gas to
feed ratio in Figure 8.19. Conversion increases sharply with an increase in sweep gas flow
rate. The membrane reactor is very effective if a large hydrogen driving force across the
membrane can be maintained along the length of the membrane.
To optimise the membrane reactor's performance, two main conditions must be satisfied.
1. The dehydrogenation rate and the hydrogen permeance through the membrane should be
similar.
2. A high hydrogen driving force must be maintained along the full length of the membrane.
274
80 -------------------------------.. ---------------------_._----------.-.-------...---------------------.-------.. _- 2
75 =:__: :_~_~_~__~__:-:_j I
~rr-<O::--'i·::::--:-:--::::---:;;...;::-:::-~:::--:-::--:::-~--:.:-::-;:::_;::;_::;::__~_~__=__::=__ 1.6
i
70
J
i I
: /
65 I 1.2
i..u'ot~~----------~============~__j
55
fI I ~COn I .
I -0-
....lon(IoII ....'
LosttMd(rlght ....,
0.8 ~
.;.
i 50 I
0.' ~
45 t---------------------------1 ..
40~- __--~-__- __- - _ -__- __- _ 4 0
o 4 8 12 16 20 24 28 32
Sweep gas to standard feed molar ratio
Figure 8.19: Effect of sweep gas to standard feed molar ratio on ethanol conversion and
ethanol losses
For the membrane parameters and reaction rate parameters measured in this study, the
hydrogen permeance was much faster than the reaction rate for the ethanol dehydrogenation
reaction. A more active catalyst will improve the overall ethanol conversion significantly,
since the ethanol conversion in the membrane reactor was mainly limited by the reaction rate.
The hydrogen driving force may be improved by using a vacuum on the shell side or a
positive feed pressure on the tube side instead of a sweep gas. The larger the pressure
difference between the shell and the tube sides, the better the overall reaction conversion will
be if the reaction rate stays constant (Figure 8.19). Higher feed pressures will, however,
have a negative impact on the reaction rate and cause a decline in the reaction rate.
8.6. SUMMARY
The effects of feed flow rate on reaction conversion were discussed and the low flow rates
that have been investigated in membrane reactors by previous authors, mentioned.
Membrane permeance data for ethanol was presented. It was found that the nitrogen to
ethanol selectivity of the Pd film membrane was about four to one.
A process model was developed for the catalytic membrane reactor. All model conditions
were covered, including the effects of pressure drop along the packed membrane and
275
interphase and intraparticle mass transfer resistance. The reaction temperature was assumed
to be constant and there were no radial concentration profiles in the catalysts bed. Equations
were formulated to calculate the mass transfer coefficient for ethanol dehydrogenation at
different Reynolds numbers. Experimental plug flow reactor data was compared to
equilibrium data at different feed flow rates. At low flow rates, the alcohol conversion was
similar to the equilibrium conversion at the tested temperatures. Flow rates were determined
where the conversion started decreasing to below the equilibrium values. For the catalytic
membrane reactor to be applicable, an improvement in reaction conversion must be obtained
in the higher flow rate region.
276
sweep rates, the membrane parameters play very important roles and can result in large
improvements in alcohol conversion.
• The two process parameters, the feed rate and the sweep rate, must be chosen in
conjunction with one another. In both a plug flow reactor and a membrane reactor, the
alcohol conversion will drop with an increase in feed flow rate, because the reaction time
becomes shorter. The feed rate has a effect on the percentage of produced hydrogen that
permeates to the shell side. At low feed rate, back diffusion of hydrogen will occur. At
high feed rates, the purity of hydrogen on the shell side is high. The sweep to feed molar
ratio is very important for extracting hydrogen from the tube side to the shell side. At
high sweep rates, most of the produced hydrogen can be extracted from the reaction side
(tube side) to the shell side. The reaction equilibrium is shifted and the reaction
conversion increases.
277
9. CATALYTIC MEMBRANE REACTOR DATA
The experimental procedures for catalytic membrane reactor testing were discussed in
section 3.4. The Pd-Ag membrane (N8x) was packed with the optimised catalyst. The
optimised catalyst consisted of 14.4 wt % copper on silica, with a particle fraction ranging
from 500 to 800 microns. The performance of the reactor with and without an inert sweep
gas was compared. The effect of the sweep gas flow rate on reaction conversion and
selectivity was studied at different temperatures. Results are listed in Appendix F.
Experimental data was compared to model predictions for both the dehydrogenation of
ethanol and the dehydrogenation on 2-butanol. All daia presented was at the exit conditions
of the reactor. Experiments conducted to optimise the catalyst composition (Chapter 4) were
also performed in a plug flow reactor, but the values in Chapter 4 were not directly
comparable to the plug flow reactor data in Chapter 9. In Chapter 4 a diluted feed (feed to
nitrogen molar ratio of 1 to 4) was employed, but in this chapter an undiluted feed was used.
Conversions for the plug flow reactor will thus be higher in Chapter 9 than in Chapter 4,
due to higher feed concentrations.
For ethanol dehydrogenation, two models were constructed. Model 1 used the measured
reaction rate and membrane parameters for the reaction at the specific temperatures. In
model 2, the reaction rate parameters were changed to better fit the experimental data. Model
2 does not have significant physical importance, but indicates how the parameters must be
altered for the model to better predict the experimental data. For 2-butanol dehydrogenation,
only one model was developed at every temperature. The model did not use measured kinetic
data, since the model then under predicted the experimental 2-butanol conversion
significantly. Instead the model used altered parameters to improve the prediction of
experimental values (similar to model 2 for ethanol dehydrogenation).
It should also be emphasised that the reaction rate models developed in Chapter 5 are
independent of the membrane reactor models in this chapter. This implies, for example, that
model 2 (ethanol dehydrogenation reaction) in this chapter does not use the reaction rate
parameters of model 2 in Chapter 5, but the reaction rate parameters defined in this chapter.
278
9.1. ETHANOL DEHYDROGENATION
The results of ethanol dehydrogenation in a membrane reactor are listed in Appendix Fl.
The results at each temperature will be discussed separately (sections 9.1.1 to 9.1.3). The
following definitions were mentioned in Chapter 4 and will also be used in this chapter.
Standard parameters (measured reaction rate and membrane parameters) for modelling were
listed in Tables 8.3 to 8.5. Included in the standard parameters is an effectiveness factor of 1.
279
membrane reactor provided a good increase in ethanol conversion even under the higher feed
flow rate conditions.
~r_--------------------------.
-- - - .• 36
• - --
lfl.
~r_-----+--~-=-~
•.~---------------~.
.... ...•.. ,," ,..... •..... . ..• .,. • --- -
-.-
· .. -.•
~."~ ... ,__: . I
32
•.. . ..,
:.!; 40 •
, . •
~
'~'
•
., ......
••
8~~~~~----------------------
,:'~i··
+ .
28
8
..
"f=:=~~~~~~~ Bhlnoi to aCllblld!ltyda
H
• Tatalllthlnol ODIWIII'tIon
ett..01 tv.~
H
• Tut.!.tn.IoI oonv.llon
25 ." Ethtnol conv.-lIiIon (mcrMl1) -+. EthmolllOllVW$lon (m~ 2) 24 .,."' Elhnol con""ulon (modal 1) - ... EthlnollXln\oWtlon (model 2)
-Eq.IiibriLftl ecnveriecn -~lbrilm conv.illQ1
20~~~~:::=:::==:=~ t...=:::::~~::::::::=::=::::==:=='--, 20
o 2 3 456 7 o M 1 1~ 2 U 3 ~ 4
Sweep gas to teed molar ratio SWMp gas to fMd molar ratio
Figure 9.1: Ethanol conversion at 250°C vs. Figure 9.2: Ethanol conversion at 250°C vs.
sweep gas flow rate (F.. = 2.39*10-5 mol/s) sweep gas flow rate (F.. = 4.77*10-5 mol/s)
Two models were formulated to estimate total ethanol conversion for the membrane reactor.
Both models only take the dehydrogenation of ethanol into account. They do not account for
side reactions. They predict the total amount of ethanol converted and not the acetaldehyde
yield. The acetaldehyde yield will be lower than the total ethanol conversion due to the
formation of side products.
According to data for a differential reactor (see 8.3.4.1), the feed flow rates in both these
figures (Figures 9.1 and 9.2) are in the mass transfer limited regime, resulting in a
effectiveness factor of less than one. Modelling calculations with an effectiveness factor of
less than 1 and model parameters from Chapter 8, totally under predicted the ethanol
conversion. The reaction rate for the packed bed, plug flow reactor was much higher than the
values obtained from kinetic experiments with a differential reactor. The higher reaction rate
could be due to a higher k' -value, lower Kac<t-value or a higher TJ-value.
The individual changes in each parameter could not be obtained from plug flow reactor
experiments in this study, but the combined effect of the three parameters could be predicted
fairly accurately. The hydrogen permeance in the reaction mixture also varied significantly
from pure component data. Hydrogen permeance through the Pd-Ag film decreased due to
280
competitive adsorption on Pd sites. Other components adsorb on Pd and fewer sites are
available for hydrogen adsorption. The result was a decrease in hydrogen permeance.
For model 1 in Figures 9.1 and 9.2, standard membrane and kinetic parameters were used,
but with an effectiveness factor of 1 (see Tables 8.3 to 8.5). For model 2, the following
values were used for modelling:
• k'-value: 1.5 times the standard value,
• hydrogen permeance: 0.5 times the standard value,
• TJ: 1,
• other permeances: 0.5 times standard values to maintain standard selectivities, and
• all other data: similar to standard values.
Model 2 gIVes a much better prediction of the measured data. Table 9.1 indicates the
difference in total ethanol conversion between the model values and the measured values.
The average absolute error is defined as the difference between the measured total ethanol
conversion and the predicted values, averaged over all the sweep gas flow rates.
Figure 9.3 shows the selectivity of acetaldehyde as a function of the sweep gas flow rate. At
250°C, there was no improvement in selectivity with an increase in sweep gas flow rate.
Selectivity was between 80% and 95%. The acetaldehyde selectivity was most sensitive to
the flow rate and a higher flow rate resulted in a higher acetaldehyde selectivity. By-products
(see section 8.4.1) required a longer reaction time to form than the acetaldehyde did, and thus
decreased with an increase in feed flow rate.
281
98
96
94
• -'"
..
...... .... . .. ......• __ F=2.39E-5 movs
~
..•.. F-4. T7E-5 molfs !
92
~ 90 ,
~ •\.~ i
> sa
~
1J
III
.
8.
I'\.
\ -= •
!
;,
82
\ / ,I
80
V
78
0 1 2 3 5
•
Sweep gas to feed flow molar ratio
• 7 8
The absolute and relative improvements obtained in total ethanol conversions by the
membrane reactor are summarised in Table 9.2. Absolute and relative improvements are
defined by:
.
Table 9.2· Improvements in total x.. for the membrane reactor at 250°C
Flow rate Sweep to feed Absolute Relative
(molls) molar ratio improvement in improvement in
conversion (%) conversion (%)
2.39*10·' 8 19.7 60.8
4.77*10·' 4 5.7 19.8
282
9.1.2. MEMBRANE REACTOR EXPERIMENTS AT 275°C
Figures 9.4 to 9.6 show measured and model data for ethanol dehydrogenation at 275 °C.
The feed flow rate in Figure 9.4 resulted in the equilibrium acetaldehyde yield for the plug
flow reactor. The equilibrium ethanol conversion and the measured acetaldehyde yield
differed by only about 2%. As the feed flow rate increased, the acetaldehyde yield decreased
and the difference between the equilibrium ethanol conversion and the measured
acetaldehyde yield increased (see Figures 9.5 and 9.6). In Figure 9.6 the exit ethanol
conversion dropped to below the equilibrium value due to the higher feed flow rate. The
same principles that were discussed in section 9.1.1 apply to the experiments performed at
275°C.
65 -' "
•
60
.. - - -
5Or-----~.--------------------~ •
55 - - - -- - - • • • - -! - .. ___ - - - - .. _- - - .. - -t
• -.- - - -'--"- -
,
'if. ,-
, 'if. .. r------_~.~-~==i.c__-----------
c
f!•
50
~ .. J.-.--
• ....
." ..;0,"
•
u
.. ~~
• .- ........
o·
Figure 9.4: Ethanol conversion at 275°C vs. Figure 9.5: Ethanol conversion at 275°C vs.
sweep gas flow rate (F", = 4.77*1O-s moVs) sweep gas flow rate (F", = 9.54*1O-s moVs)
Model I in Figures 9.4 to 9.6 used the standard parameters with an effectiveness factor of 1.
Model 2 had the following parameters:
• Tt: 1,
• other permeances: 0.5 times standard values to maintain standard selectivities, and
283
For the highest tested feed flow rate at 275 °C (Figure 9.6), model 2 started under predicting
values at the higher sweep rates. For lower feed rates (Figures 9.1,9.2,9.4 and 9.5), model 2
was accurate in predicting total ethanol conversion at the higher sweep rates. A very likely
explanation for this is the presence of interphase mass transfer resistance. If interphase mass
transfer resistance was present, the value of lJ would increase with an increase in feed flow
rate. Since the lJ-value was fixed at 1 for all calculations, the model would under predict total
ethanol conversion at high feed flow rates in the presence of interphase mass transfer
resistance. Not enough data was available to determine the k' -value, K.... and lJ,
individually, and therefore values were fixed at each temperature and not treated as variables.
The model cannot account for changes in reaction rate parameters with changes in feed flow
rate.
50
30 j...~~·"'··C· ':c.:'~----.C.~--.-~
.. . . . -"'~"~~~~~~---1
Figure 9.6: Ethanol conversion at 275 °C vs. Figure 9.7: Acetaldehyde yield at 275 °C
sweep gas flow rate (Fa = 1.43*10-4 mol/s) for a membrane reactor
Figure 9.7 shows the combined effect of feed flow rate and sweep gas to feed molar ratio on
total ethanol conversion. The feed multiple is the multiple of the standard feed rate
5
(4.77*10- mol/s). Total ethanol conversion increased more sharply at the lower feed rates
with an increase in the sweep gas flow rate. Towards the higher feed rates, the total ethanol
conversion for the plug flow reactor (sweep gas to feed molar ratio = 0) dropped slightly.
284
Model 2 gave accurate predictions of the exit membrane conditions. Table 9.3 compares the
model values with measured values for data at 275 DC.
Table 9.4 indicates the measured improvements obtained by using the membrane reactor at
275 DC. The higher the equilibrium conversion, the more difficult it becomes to push the
reaction's equilibrium further towards the product side. The improvements in total ethanol
conversion with a membrane reactor will become srnaller at higher temperatures.
Table 9.4: Improvements in total Xet for the membrane reactor at 275 DC
Flow rate Sweep to feed Absolute Relative
(moVs) molar ratio improvement in improvement in
conversion (%) conversion (%)
4.77*10. 0
4 13.8 29.9
9.54*10. 0
4 7.0 16.0
1.43*10· 3 10.4 28.9
Figure 9.8 indicates an improvement in acetaldehyde selectivity at the lower feed rates with
an increase in sweep rates. Selectivity ranged between 80% and 95%, with the highest feed
rates giving the best selectivities at the lower sweep gas ratios. At the higher sweep gas
ratios, selectivity improved to above 90% for all feed rates.
285
96 ------------------- ------------------------------------------------________________________ ·---..-..------------0-'
""..---.... ~-... ~ ................
94 --A_
~-- ... ----- ·········/i
92
~OO
•
:f 88 r.~ ....•.
~ 86
jiB4 -+-F=4.77E-5 molls
..•.. F=9.54E-5 molls H
82 t-------------------~~-.~-£~~1~.4~~4~mo~~~~,
80
78
0 0.5 1 1.5 2 2.5 3 3.5 4
Sweep gas to feed now molar ratio
• T]: 1,
• other permeances: 0.5 times standard values to maintain standard selectivities, and
• all other data: similar to standard values.
At the lower ethanol flow rate (Figure 9.9), the exit ethanol conversion was the equilibrium
conversion, but at the higher flow rate the exit conversion dropped to below the equilibrium
5
value. At a flow rate of 9.54*10- moVs and a reaction temperature of 275 °C (Figure 9.5),
the total ethanol conversion was above the equilibrium value. To be consistent with data at
275°C, the total ethanol conversion should not have been below the equilibrium value at
300°C for a plug flow reactor and a feed rate of 9.54*10-5 moVs (Figure 9.10), due to the
higher reaction rate. The conversion was, however, below the equilibrium value. This
suggested that the reaction rate was slowed down at 300°C. This finding is consistent with
the catalyst stability data in Table 5.1. Between 280 °C and 340°C there was a sharp
decrease in catalyst stability, mainly due to coking. In the membrane reactor at 300 DC,
286
coking must have occurred during the testing period. Total testing took about 16 hours.
After testing at a flow rate of 9.54*10.5 molls (20 mVh), the feed rate was increased to 30
mllh, but the plug flow reactor results became very inconsistent.
60
- - - -- •
"r---------------._~._=_=_=~~----~ __
· r
l
an.,01 to 1e«lldllhydlo • Tobit eth...111 converllion I 40 ,," ........ _£_.
45 ., .'. Ettn!lnol CO!WI.rllion (model 1) -+,' Shanol conv.llioo (model 2)
- EIJoIIIbr11.1Tt c::onverlson
'" L.:=;:~~=::::=:==::==.::==:j
o 0.5 1 1.!I 2 2.5 3 3.5 " 1.5 2 2.5 3 3.5 "
SWeap gas to feed molar ratio SWeep gas to 1Hd mollr ratio
Figure 9.9: Ethanol conversion at 300 °e vs. Figure 9.10: Ethanol conversion at 300 °e vs.
sweep gas flow rate (Fe< = 4.77*10. 5 molls) sweep gas flow rate (Fe< = 9.54*10-5 molls)
Table 9.5 tabulates the accuracy of models 1 and 2 at 300 °e. Once again model 2 is very
accurate at the tested sweep flow rates, but model 1 under predicts the measured values due
to a reaction rate that is too slow in model 1.
The improvements in total ethanol conversion using a membrane reactor at 300 °e are
indicated in Table 9.6.
Acetaldehyde selectivity varied between 85% and 96% (Figure 9.11). For the plug flow
reactor, the selectivity was the highest at the highest feed flow rate. At low sweep rates, the
selectivity dropped for both feed flow rates tested and then increased again towards the
287
higher sweep gas flow rates. In both cases the selectivity was above 90% at the higher sweep
gas flow rates.
Table 9.6: Improvements in total X.. for the membrane reactor at 300°C
Flow rate Sweep to feed .. Absolute Relative
(mol/s) molar ratio improvement in improvement in
conversion (%) conversion (%)
4.77*10·) 4 8.1 14.3
9.54*10·) 4 7.7 16.3
9. ,
,!
98
• ,,
94
!
i'-
~92
__-----1
····················T
.~ ................ ~
......
¥
11
90
!
til
88 /' :,
88 1\ I V
I
-+-F=4.77E-5 rnoVs
..•.. F 9,54E-5 moVs
!
,
,!
84 , , , ,!,
0 0.5 1 1.5 2 2.5 3 3.5 4
Sweep gas to feed now molar ratio
288
I
I
I
At 300°C, coking began to take place and it lead to inaccurate data. From an operating point
of view, there seemed to be an optimum working temperature. If the temperature was too
low, the reaction kinetics was slow and the conversion poor, even with a membrane reactor.
Organic molecules adsorb more strongly onto the Pd-Ag membrane film at low temperatures
than at higher temperatures and reduce hydrogen permeance more substantially at lower
temperatures. Higher temperatures improve hydrogen permeance in two ways: by increasing
the hydrogen permeance kinetics and by reducing the organic molecule adsorption onto the
Pd-alloy surface.
On the negative side: at high temperatures, catalyst deactivation occurs quickly, which
reduces the activity of the copper based catalyst and leads to poor conversion. Coking will
most probably also occur on the membrane surface and thus reduce hydrogen permeance
severely. The sensitivity analysis in Chapter 8 indicated that the performance of the
membrane reactor was more sensitive to reaction rate parameters than to membrane
parameters. To obtain large improvements in ethanol conversion over the conventional plug
flow reactor, the catalyst activity must remain high.
A temperature of 275°C gave the best ethanol dehydrogenation results. The selectivity
towards acetaldehyde formation was above 80% at all times. Selectivity was more dependent
on feed flow rate than on temperature or sweep gas flow rate. The selectivity increased with
an increase in feed flow rate and also, to a lesser extent, with an increase in sweep gas flow
rate. Selectivity was above 90% for all experiments at the highest tested sweep rates, except
for the very low feed flow rate of 2.39* 10.5 molls. Figure 9.12 indicates some selectivity
profiles at different temperatures. There was a lot of variance in the data at the lower sweep
rates, but the final selectivities at the higher sweep rates were not very dependent on
temperature.
Total ethanol conversions were modelled with two different models. The first model used
standard conditions (as defined in Tables 8.3 to 8.5) and an effectiveness factor of 1. In most
instances the model under predicted the measured data. This indicated that the reaction rates
289
determined in kinetic experiments in a differential reactor were much lower than the values in
the plug flow reactor. Higher k' -values were employed in the second model, keeping all
adsorption coefficients constant The effectiveness factor was fixed at 1 and the permeances
for each component in the reaction was halved. The net result was a model that predicted
total ethanol conversion to within 4.5%.
96 ....................................................................................................................... .
94f---I-~-----=~'======::::!
--,:,:::
/ ,--- --'
........
-T-250°C -
Deng et aI. (1995), Liu et aI. (1997) and Raich and Foley (1998) studied ethanol
dehydrogenation in a membrane reactor, under conditions that differed greatly from those
used in the present study. The feed rates that they employed were more than an order of
magnitude lower than the feed rates used in this study (see Table 8.1). Liu et aI. (1997)
tested ethanol dehydrogenation as a function of feed rate. At a similar WIF ratio of
1000 g.min/mol to Liu et aI. (1997), this study achieved a maximum of 65% conversion at
300°C compared to their 30%.
Acetaldehyde selectivity in this study was significantly higher than the values obtained by
Raich and Foley (1998). This is to be expected at the higher feed rates. They obtained higher
conversions because they used a very low feed rate and a very diluted feed. When the partial
pressure of ethanol was reduced (by using a diluted feed), the equilibrium conversion for the
ethanol dehydrogenation reaction increased.
290
9.2. 2-BUTANOL DEHYDROGENATION
Results of 2-butanol dehydrogenation in a membrane reactor are listed in Appendix F2.
Catalytic membrane reactor experiments were performed at temperatures ranging from
190°C to 240 dc. Standard reaction rate parameters and membrane parameters used for
modelling were listed in Tables 8.6 to 8.8. Catalytic membrane performance at 190 DC,
215 °C and 240°C will be discussed separately. For the three dimensional figures shown in
this section, the standard feed rate was 3.04*10-5 molls and the feed multiple was the multiple
of the standard feed rate.
· ,.,:
• permeances:
fixed at 1 for all flow rates, and
the same as in Table 8.7 at 190°C.
Increasing the model's k- value for 2-butanol dehydrogenation alone was insufficient to
obtain predicted conversions close to the measured conversions for the plug flow reactor.
The MEK adsorption coefficient was also reduced by a factor 10 for modelling. This may
seem exaggerated, but it should be noted that the individual adsorption coefficients were
determined for pure component data. For example, the MEK adsorption coefficient was
determined when only 2-butanol and MEK were passed over the catalyst bed in the
differential reactor. The conversion was kept very low and very little hydrogen formed. In
291
the plug flow reactor, 2-butanoi was converted to MEK and hydrogen in equal molar
amounts. The MEK was essentially in a MEK-hydrogen environment and not in a MEK-2-
butanol environment. The MEK adsorption behaviour could thus be expected to be
significantly different from the behaviour in a MEK-2-butanol mixture.
80
"
..• "
•
"'70
.. . .. .
, .. -'
~t-----------------
•
•= • • ~ .~
='" j----l!-_---~==-~=:c::-..::.-.---
~"
> •
• 2-.... ~. ~ MEK • T""2-IM""""""", ~ j ... .
860
1- .' -2-eut.nol COfIva'elon (1) -EquAlbrium o::onv.,.;$Q'I
8~~~==========~
/. . . . :. . . I • 2-BubnoiIoMEK
"+7"---1I .•
" 2-BWnoIllIIfl'.'WnIon(1) -E4II~~
50 ,,~----------"------, , ,
0 2 4 6
SWeep gas to teed molar ratio
6 10 12 14 o 2 4
Sweep gas to fad molar ratio
•
Figure 9.13: 2-Butanol conversion at Figure 9.14: 2-Butanol conversion at
190°C vs. sweep gas flow rate 190°C vs. sweep gas flow rate
(F2But= 1.52*10-5 moVs) (F2But= 3.04*10-5 moVs)
eo ,
. ,
•
'j" 52
.. ....•
~
•>
~ .. ! .
....... ... '" ."
. I '~_ ~
U
.. I
•
·
2-Bl.Mnailo MEK
Figure 9.15: 2-Butanol conversion at 190°C vs. Figure 9.16: MEK yield at 190°C
sweep gas flow rate (FzBut= 4.56*10-5 moVs)
The 2-butanol conversion was sensitive towards the feed rate. There was a drop in 2-butanol
conversions at 190°C with an increase in feed rate. Figure 9.16 illustrates this very clearly.
292
Optimum conversions were obtained at low flow rates and high sweep ratios as predicted by
the membrane reactor model.
At 190°C, the model predicted the measured values well. There was little difference
between the total 2-butanol conversion and the MEK yield, indicating very high selectivity.
The best 2-butanol conversion in the membrane reactor at a feed flow rate of 1.52 * 10. 5 molls
was 79% and it dropped to 53% at a feed rate of 4.56*10. 5 molls.
Table 9.7: Improvements in total X2But for the membrane reactor at 190°C
Flow rate Sweep to feed Absolute Relative
(molls) molar ratio improvement in improvement in
conversion (%) conversion (%)
1.52*10"0 12.6 24.3 44.3
3.04*10"' 6.3 8.8 16.1
4.56*10"' 6.3 8.0 17.7
As mentioned previously (section 8.4.2), the selectivity towards MEK production at 190°C
was very high. Figure 9.17 plots selectivity as a function of feed rate and sweep gas to feed
293
i .
ILl.
molar ratio. Selectivity was above 99% for all experiments and increased towards 100"10 at
the higher flow rates. The only by-product observed in some experiments was 3-octanol.
100
..
99.6
•• • •
;!.
~992
•
~
OJ 98.B
'" .F=1.52E·5 molls
• f=3.04E-5 moVs -
98.4
.. F=4.56E-5 molfs
9B
0 1 2 3 4 5 6 7
Sweep gas to _ !low molar ratio
The model did not take mass transfer resistance into account, since the effectiveness factor
was fixed at I. Kinetic data obtained in the differential reactor indicated a reaction free from
interphase mass transfer resistance (see discussion in section 5.2.2). The plug flow reactor
data contradicted the differential reactor data regarding interphase mass transfer resistance.
One possible reason is the difference in the Reynolds numbers of the two situations. For
differential reactor experiments, a 4 mm inside diameter tube was used, resulting in much
larger gas velocities as compared to the 7 mm inside diameter tube used for plug flow and
membrane reactor experiments. By assuming 11 as a variable dependent on feed flow rate, a
better model fit might be obtained, but as a first approximation this was not investigated.
294
If 1] decreases with a decrease in feed flow rate, the k' -value must be even higher than the
value used for this model. In such a case, the model should be fitted to the measured data at
the maximum feed rate with a high 1]-value to determine the k' -value and KMEK. The 1]-value
must then be reduced at the lower flow rates to yield lower overall reaction rates and lower
model conversions at the lower feed rates.
00
, ........ . •
•
• 85~------------------------~'~~
.• .
i: .' ./
;1.85 ;I. ...•...... ' .. ' .' ....
e
!.~_d~MEK • Tm.~~~_ ~
.2 ••
~80
•
•0 " ~--.·~-II • :z..Butanoi to MEK
~
U
• L""'" conversion (1) -
2-Bllt.nol
• Tobd 2-8ut (:anv.1'$Ian
Equllbftum convtlrison n 8 . . .> . I. ...
2-BubI'IOI converlion (1) -Eq.JIlbriIm converitol'l !i
70 t--~======~;l
1
70
;
,
At a feed rate of 1.52*1O-s molls and 3.04*10-5 molls, the exit 2-butanol conversion for the
plug flow reactor was similar to the equilibrium conversion. The model values were larger
5
than the equilibrium values. At the highest feed rate tested (6.08*10- molls), the 2-butanol
exit conversion in the plug flow reactor dropped below the equilibrium value and the reaction
moved into block 2 as described in Figure 8.1. Figure 9.21 indicates the difference in 2-
butanol conversion at different feed rates in the plug flow reactor (at zero sweep gas). At
high sweep rates, the membrane reactor improved 2-butanol conversion for all feed flow rates
tested. The highest conversions were in the region of 90%, with the equilibrium value being
70%.
295
15
90
.. ..
, ...
'
6O+-t-_-_-_-_-_-_-~
o 234 5
Sweep gas to feed molar raHo
•
Fil!Ure 9.20: 2-Butanol conversion at 215°C vs. Figure 9.21: MEK yield at 215°C
sweep gas flow rate (F2But= 6.08*10.5 mol/s)
The performance of the membrane reactor at 215°C is listed in Table 9.9 and the accuracy of
the model is summarised in Table 9.10.
Table 9.9: Improvements in total X2But for the membrane reactor at 215°C
Flow rate Sweep to feed Absolute Relative
(molls) molar ratio improvement in improvement in
conversion (0/0) conversion (%)
1.52*10·> 12.6 21.2 30.5
3.04*10·' 6.3 IS.2 26.6
6.0S*10·> 6.3 7.4 1l.S
.
Table 910·. Model differences for 2-butanol conversion at 215°C
Feed (molls) Average absolute error in total
conversion for model 1 (%)
l.52*10·> 6.6
3.04*10·> 1.9
6.08*10·> 2.5
For model differences in Table 9.10 and in all other tables, the absolute differences are listed.
There is no distinction between model values higher or lower than the measured values, only
296
the difference between them is stated. The model loses accuracy at the lower feed rates, with
the difference between measured values and model values increasing to above 6%.
The MEK selectivity at 215°C (Figure 9.22) was high (above 96%), but slightly lower than
the values obtained at 190 °C. The selectivity improved with an increase in feed flow rate
and at the highest feed flow rate tested, the selectivity reached 100%.
100
•
99 •
.... 98
• •
~ •
.a:
j
.,.. 97
•
• • • F=1.52E-5 molls
95
0 2 3 4 5 6 7
Sweep gas to feed flow molar ratlo
• 1]:
fixed at 1 for all flow rates, and
• permeances: the same as in Table 8.7 for 240°C.
At 240°C, the exit 2-butanol conversion remained at the equilibrium value for all flow rates
4
tested in the plug flow reactor, up to a feed flow rate of 40 mllh (1.22*10- molls).
Membrane reactor experiments were only performed at feed flow rates of up to 20 mllh
(6.08*10-5 molls) due to limitations of the sweep gas flow meters. At the low flow rates
(Figure 9.23 and 9.24) the model over predicts the measured values and at the higher flow
297
rates, the model under predicts the measured values. The same explanation supplied in
section 9.2.2 is applicable at 240°C. Initial modelling results indicated the presence of
interphase mass transfer resistance at low flow rates for 2-butanol dehydrogenation. The
effectiveness factor was not constant with feed flow rate, but declined as the feed flow rate
declined. Measured 2-butanol conversions were at or above 90% for all flow rates tested.
'00 '00
os .. ' .
--, ....... "
ae
~t-----------------~~~-----
............................
....
"
'"
c
.2
00
••
.
5 90 t-----.
..•.,-'."'----------------------~-
;> ./
0
c
0
"
• i8 "
Mt---,--~---------------------
•
80
o 2 4 5 6 7
2 4 II e
SWeep gas to fAd molar ratio
Sweep gas to fHd molar ratio
For the plug flow reactor in Figure 9.26, the measured data confirmed, to some extent, the
model prediction of interphase mass transfer resistance. At a zero sweep gas to feed molar
ratio, the total 2-butanol conversion increased very slightly with an increase in feed flow rate.
The improvements in 2-butanol conversion in the membrane reactor for high sweep rates
were similar at all feed multiples. If higher feed rates were to be employed, the benefits of
the membrane reactor would start declining.
The improvements obtained by the membrane reactor are listed in Table 9.11. As mentioned
previously, improvements become smaller at the higher temperatures, because the
equilibrium conversion becomes very high. The model predicts the measured values at lower
feed rates with less accuracy than at the higher feed rates (see Table 9.12). The difference
between the model predictions and the measured data increases above 5% at the lower feed
rates.
298
1'l~!II!' ',
t-============~-
!
I •
r
. .... . ... .. . ......•
............
.............
....•', ....
8D~=============--
o 2 3 4 5 ,
SWftp gas to fMcI motIr ratio
Figure 9.25: 2-Butanol conversion at 240°C vs. Figure 9.26: MEK yield at 240°C
sweep gas flow rate (F2But= 6.08*10.5 molls)
.
Table 9 11'. Improvements in total X2But for the membrane reactor at 240 °C
Flow rate Sweep to feed Absolute Relative
(molls) molar ratio improvement in improvement in
conversion (%) conversion (%)
1.52*10·' 12.6 16.8 21.4
3.04*10·' 6.3 11.9 15.1
6.08*10·' 6.3 8.2 10.0
Selectivity didnot decrease with an increase in temperature up to 240 °C, it remained above
97% (Figure 9.27). At higher temperature, butenes will start forming and the selectivity will
299
III!!
drop. Figure 4.41. shows the MEK selectivity as a function of temperature up to 390°C,
Selectivity at 240°C was very good but then gradually declined towards the higher
temperatures.
•
99,5 •
• •
•
'#-
99
•
~
>
~ 98.5
•
'iI
U)
98 • F 1.52E-5 molls -
97.5
• • F=3.04E~5
97
0 2 3 4 5 6 7
Sweep gas to feed now molar rstlo
Catalytic membrane reactor experiments were performed with feed flow rates resulting in
both equilibrium restricted and non-equilibrium restricted 2-butanol conversion (at 190°C
and 215 0c). For high flow rates resulting in the non-equilibrium restricted conversions, the
theoretical benefits of the membrane reactor were confirmed by experimental data. At
240°C, membrane experiments were only performed for feed flows resulting in the
equilibrium restricted conversion.
300
Figure 9.28: MEK yield % vs. flow rate and temperature
(feed mUltiple of one equals 3.04*10.5 molls (lOml/h»
Figure 9.29 shows the improvements in total 2-butanol conversion obtained from a
membrane reactor. Figure 9.29 illustrates results for a feed rate of3.04*10· 5 molls (10 m1/h).
To achieve a high MEK yield, a high operating temperature must be used and the temperature
should be kept at or below 250°C to improve catalyst life. Figure 5.8 indicates catalyst
deactivation over a 24 hour period. at different operating temperatures. At 250°C, the
catalyst was stable, but at higher temperatures the activity of the catalyst declined over time.
At 240°C, the membrane reactor gave an increase in MEK yield of about 10"10.
301
No data on 2-butanol dehydrogenation in a membrane reactor could be found in literature. It
was not possible to compare results on 2-butanol dehydrogenation with data obtained from
other researchers.
9.3. SUMMARY
Theoretical benefits of a catalytic membrane reactor as compared to a plug flow reactor were
realised for the dehydrogenation of ethanol and 2-butanol. For ethanol dehydrogenation,
catalyst deactivation above 280°C limits the upper operating temperature of copper based
catalysts. For 2-butanol dehydrogenation, the copper based catalyst was stable at 250°C and
lower. Coking at higher temperatures will deactivate the copper on silica catalysts and result
in a rapid decrease in catalyst activity.
The optimum working temperature for ethanol dehydrogenation with copper on silica
catalysts is between 270 °C to 280°C. The equilibrium conversion in this temperature range
is just over 40% for an undiluted feed at atmospheric pressure. The membrane reactor
improved the total ethanol exit conversion from 45% (plug flow reactor) to 60% at low feed
flow rates and from 36% to 46% at high feed flow rates. The acetaldehyde selectivity of the
reaction was above 80% for all experiments performed. The selectivity was most sensitive to
the feed flow rate and the selectivity improved with an increase in feed flow rate. The
selectivity also improved with an increase in sweep gas flow rate, but changed little with
temperature from 250°C to 300 °C. At a high sweep gas flow rate, the acetaldehyde
selectivity was above 90% for the majority of the experiments.
The membrane reactor is less suited for 2-butanol dehydrogenation, because there are less
temperature limitations for this reaction. Improved conversion of 2-butanol can be obtained
by increasing the operating temperature up to 250°C, without any significant deactivation of
the catalyst. It is more economical to increase the reaction temperature than to use a
membrane reactor at a lower temperature to obtain the same results. The membrane reactor
does have some benefits in that it can obtain a near complete conversion at the higher
temperatures. Maximum plug flow reactor conversion at 240 °C is 80%. This can be
302
increased to above 90% at 240"C by employing a membrane reactor. The exit 2-butanol
conversIOn in the membrane reactor was also modelled with equations developed in
Chapter 8. A proper model could only be formulated by using a k'-value twice (or more) the
value obtained from the differential reactor experiments and a MEK adsorption coefficient
one tenth of the value obtained from the differential reactor data. The model predicted the
experimental data to within 6.5% of the measured exit 2-butanol conversion. Selectivity
towards MEK production was high under all conditions (> 96%) and increased towards 100%
at the higher feed flow rates.
The Pd-Ag membrane was stable between 190 °C and 300 "C in both an ethanol and a 2-
butanol environment for about 30 working days (450 hours) in total. The same membrane
was used for both reactions and the membrane reactor maintained its superior performance as
compared to the plug flow reactor for the full duration of testing.
303
10. CONCLUSIONS
In Chapterl, the six main goals of the project were stated. They were:
• To optimise the performance of an alcohol dehydrogenation catalyst,
• To model. the kinetics of an alcohol dehydrogenation reaction,
• To optimise the composition and thickness of Pd and Pd-alloy films for hydrogen
separation,
• To model the membrane separation process with a sweep gas and a pressure
differential,
• To compare the performance of a membrane reactor consisting of the optimised
catalyst and optimised membrane with a plug flow reactor, and
• To model the membrane reactor.
The main conclusions to the work performed to reach each goal are summarised in the
following few pages.
For silica based catalysts, there were an optimum copper percentage of 15 wt % on silica
to obtain the highest conversion of the alcohol towards the desired product. The addition
of cobalt, chromium or nickel stabilisers to copper on silica catalysts had negative effects
304
on acetaldehyde yield. More by products fonned and the yield of acetaldehyde decreased
due to a decline in acetaldehyde selectivity. When the same stabilisers were added to
copper-alumina catalysts, the effect was positive and the production of di-ethyl ether (the
desired product) increased. A catalyst particle size of up to 1180 microns had a very
small effect on product yield, indicating that intraparticle mass transfer resistance for both
ethanol and 2-butanol dehydrogenation were negligible. The production of methyl ethyl
ketone from 2-butanol and acetaldehyde from ethanol, approached the equilibrium
conversion at low feed flow rates.
305
10.3. MEMBRANE OPTIMISATION
A palladium membrane is sensitive towards the operating conditions of the membrane. If
the membrane is to be operated above 300°C, a pure palladium membrane can be used iri
a hydrogen environment, but if the membrane is to be operated below 300°C, a
palladium alloy must be used. Pure palladium embrittles quickly below 300°C in
hydrogen, and thus becomes unusable.
Palladium films were prepared be electro less plating. To produce a high quality and very
thin palladium film, supported on a membrane, each production step must be performed
with care. Prior to plating, the asymmetric alumina membrane support must be pretreated
with palladium and tin solutions. A high palladium to tin ratio gives a more stable film.
Alternatively, new tin free pretreatment solutions can be employed. The initial plating
rate should be slow to prevent the formation of defects in the film. The plating rate was
controlled by the addition of hydrazine at a constant plating temperature of 71°C.
Thorough cleaning of the membrane after plating is crucial to remove EDTA (from the
plating solution) trapped in the membrane pores. Membranes were stirred in ammonia
solution, dried at 240°C and further oxidised at 320°C to remove organic components in
the membrane pores.
Palladium films of thicknesses down to 1 J.l.m were deposited on the inside of asymmetric
SCT a-alumina membranes (200 nm pore size). Hydrogen permeances varied between
about 8 and 15 J.l.mollm2.Pa.s for temperatures from 330°C to 450 °C and palladium films
from 1.0 to 1.5 !lm. Hydrogen to nitrogen selectivity was above 100 for all membranes
tested and above 400 for all but two membranes (thickness 1.0 to 1.5 ).Lm). These values
are a significant improvement over other published results.
Silver was deposited on palladium to form palladium-silver films of less than 2.2 !lm
thick. The separated layers caused a sharp drop in hydrogen permeance through the film.
The bi-Iayer was heat treated to improve hydrogen permeance and form a homogeneous
alloy. In a hydrogen atmosphere at 590°C, defects formed in the Pd-Ag film within
306
10 hours. The hydrogen also passivated the film at 590°C. At 550 °e, defects still
fonned in hydrogen, but at a much slower rate. Some defect fonned in an argon
atmosphere at 550°C after more than 100 hours. Few defects fonned in both argon and
nitrogen at 500°C. Nitrogen passivated the film and thus argon was chosen for
perfonning alloying experiments.
307
For ethanol dehydrogenation in a membrane reactor, the best results were obtained at
275°C. At 250 °C, the reaction rate was too slow to obtain large improvements in
conversion for the membrane reactor compared to the plug flow reactor. At 300°C,
coking of the copper on silica catalyst took place, which deactivated both the membrane
and the catalyst and which led to poor results. At 275 °C, the membrane reactor
improved the total ethanol exit conversion from 45% (Plug flow reactor) to 60% at low
feed flow rates and from 36% to 46% at high feed flow rates. Modelling the process with
kinetic data obtained from differential reactor experimentss, under predicted the
experimental data. By increasing the reaction rate (k' -value and effectiveness factor), a
very good fit for the experimental data was obtained.
A sensitivity analysis of the membrane reactor model (for ethanol dehydrogenation) was
performed to test the effects of membrane parameters, reaction rate parameters and the
operating variables on the overall ethanol conversion. Membrane data from the
308
optimised membrane and reaction rate data from the optimised catalyst were used as
standard inputs for the model.
The reaction rate parameters have a significant effect on the overall conversion, but have
no effect on the percentage of produced hydrogen that penneates to the shell side of the
reactor. A high k' -value, a low Kace,-value and/or a high effectiveness factor will increase
the reaction rate and improve the reaction conversion, while a slow reaction will increase
the ethanol losses through the Pd-alloy film defects. The membrane selectivity (H 2 :N2)
should be at least 150 to prevent large losses of ethanol feed through the film defects, and
to prevent severe back diffusion of hydrogen from the shell side to the tube side. The
feed flow rate and the sweep gas to feed ratio must be chosen in conjunction with one
another. The higher the pressure difference (i.e. the higher the sweep gas flow rate), the
better the hydrogen driving force and the higher the overall conversion.
309
11. FUTURE WORK
')
The performance of palladium and palladium alloy membranes have been optimised to
such an extent that little further work in this area is necessary. Model calculations have
shown that little will be gained by further improving membrane selectivity and/or
hydrogen permeance. The main challenge now is in improving catalyst activity and
stability. Work performed in this dissertation has proved that Pd-based membranes can
be used successfully from 190°C and upwards, without deterioration of the film. From a
thermodynamic point of view, higher temperature reactions are more favourable, since a
high temperature will improve hydrogen permeance through Pd or its alloy.
The future lies in identifying high value added alkane or alcohol dehydrogenation
reactions. Such reactions will typically be found in the pharmaceutical industry and the
perfume industry, in the production of pesticides and insecticides and other speciality
chemicals. Higher temperatures are beneficial for endothermic reactions to improve
conversions and also to improve hydrogen permeance, but result in a decrease of
selectivity. Cracking and polymerisation side reactions must be avoided to limit catalyst
coking.
After identifying suitable reactions, the catalysts for those reactions should be optimised
to maintain high selectivity and high activity over long periods of time. The performance
of the membrane, packed with the most suitable catalyst, should then be monitored over a
long time period. This step was not part of this dissertation. The final step would then be
to scale up the technology.
310
12. REFERENCES
Abdalla BK, Elnashaie SS, Alkhowaiter S et aI., Applied Catalysis A: General, 113
(1994a) 89-102
Abdalla BK and E1nashaie SS, Journal ofMembrane Science, 101 (1995) 31-42
Addiego WP, Estrada CA, Goodman DW et aI., Journal of Catalysis, 146 (1994) 407-414
Ali JK, Newson EJ and Rippin DWT, Journal ofMembrane Science, 89 (1994a) 171-184
Ali JK, Newson EJ and Rippin DWT, Chemical Engineering Science, 49 (1994b)
2129-2134
Antoniazzi AB, Haasz AA and Strangeby PC, Journal of Nuclear Materials, 162-164
(1989) 1065
Aramendia MA, Borau V, Jimenez C et aI., Journal of Catalysis, 151 (1995) 44-49
311
Annor IN, Applied Catalysis, 49 (1989) 1-25
Basile A, Drioli E, Santella F et a!., Gas Separation and Purification, 10 (1996) 53-61
Becker YL, Dixon AG, Moser WR et aI., Journal of Membrane Science, 77 (1993)
233-244
Bhattacharyya D, Bej SK and Rao MS, Applied Catalysis A: General, 87 (1992) 29-43
Bolt M and Zimmermann H, AIChE Spring Meeting, Houston, (1991), Paper 26b
Boot LA, Van der Linde SC, Van Dillen AJ et aI., Studies in Surface Science and
Catalysis, 88 (1994) 491-498
Boot LA, Van Dillen AJ, Geus JW et aI., Journal of Catalysis, 163 (1996) 195-203
312
Boudart M and Hwang HS, Journal of Catalysis, 39 (1975) 44-52
S159-S164
Brandes EA and Brook GB, Smithells Metals Reference Book, 7th Edition, Butterworth-
Heinemann Ltd., London (1992), Chapter 13, 1-99
Brinker CJ and Scherer GW, Sol-gel Science, Academic Press, New York (1990) 1-40
Brito A, Arvelo R, Villarroel R et aI., React. Kinet. Catal. Lett., 54 (1995) 77-83
Bulenkova LF, Shimanskaya MV, Gryaznov VM et al., Chemical Abstracts, 92: 180393m
(1978)
Cadus LE, Abello MC, Gomez MF et al., Industrial and Engineering Chemistry
Research, 35 (1996) 14-18
313
Canizosa I and Munuera G, Journal of Catalysis, 49 (1977) 189-200
Chaar MA, Patel D, Kung MC et a!., Journal of Catalysis, 105 (1987) 483
Champagnie AM, Tsotsis TT, Minet RG et a!., Journal of Catalysis, 134 (1992) 713-730
Chang Hand Saleque MA, Applied Catalysis A: General, 103 (1993) 233-242
Chang WS, Chen YZ and Yang BL, Applied Catalysis A: General, 124 (1995) 221-243
Cheng YS and Yeung KL, Journal ofMembrane Science, 158 (1999) 127-141
Chinchen GC, Denny PJ, Jennings JR et a!., Applied Catalysis, 36 (1988) 1-65
Church 1M and Joshi HK, Industrial and Engineering Chemistry, 43 (1951) 1804-1811
314
Cohen RL, D' Amico JF et aI., Journal Electrochemical Society: Electrochemical Science
and Technology, 118 (1971) 2042
Cohen RL and West KW, Journal Electrochemical Society: Electrochemical Science and
Technology, 120 (1973) 502
Collins JP and Way JD, Industrial and Engineering Chemistry Research, 32 (1993a)
3006-3013
Cullis CF and Newitt EJ, Procedures Royal Society A, 237 (1956) 530-542
315
Cunningham J, AI-Sayyed GH, Cronin JA et ai., Journal of Catalysis, 102 (1986)
160-171
Curtiss CF and Hirschfelder JO, The Journal of Chemical Physics, 17 (1949) 550-555
Deckman HW, Corcoran EW, McHenry JA et aI., Catalysis Today, 25 (1995) 357-363
Deng J, Cao Z and Zhou B, Applied Catalysis A: General, 132 (1995) 9-20
316
r:
I
EgloffG and HulJa G, The Oil and Gas Journal, 41 (1942) 36-51
Fairbanks DF and Wilke CR, Industrial and Engineering Chemistry, 42 (1950) 471-475
Faith WL, Keyes DB and Clark RL, Industrial Chemicals, 2nd Edition, John Wiley &
Sons Inc., New York (1957) 2-3
nd
Fogler HS, Elements of Chemical Reaction Engineering, 2 Edition, Prentice-Hall, New
Jersey (1992) 1-837
Gallaher GR, Gerdes TE and Liu PKT, Separation Science and Technology, 28 (1993)
309-326
317
'lr=
Gavalas GR, Megiris CE and Narn SW, Chemical Engineering Science, 44 (1989)
1829-1835
L
l Gibson MA and Hightower JW, Journal of Catalysis 41 (1976) 420-430
Gobina E, Hou K and Hughes R, Journal ofMembrane Science, 105 (1995c) 163-175
GrashoffGJ, Pilkington CE and Corti CW, Platinum Metals Review, 27 (1983) 157-168
Gryaznov VM, Vedemikov VI and Gul'yanova SG, Trans. Kinet. Katal., 27 (1986a) 142
Gryaznov VM, Mishchenko AP and Sarylova ME, French Patent 2,595,093 (1986c) 1
318
Haag WO and Tsikoyiarmis JG, US Patent 5,019,263 (1991) 1
Hao Y, Tao Land Zheng L, Applied Catalysis A: General, 115 (1994) 219-228
HearIe JWS, Sparrow JT et aI., The use 0/ the scanning electron microscope, Pergamon
Press, Oxford, Great Britain, (1974) 1-278
Hsieh HP, Bhave RR and Fleming HL, Journal o/Membrane Science, 39 (1988) 221-241
Hsieh HP, Inorganic membranes for separation and reaction, Elsevier Science Publishing
Company, New York, USA (1996) 1-591
Hsiung TH, Christman DD, Hunter EJ et aI., AIChE Journal, 45 (1999) 204-208
319
!'!'
I
Huang L, Chen CS, He ZD et aI., Thin Solid Films, 302 (1997) 98-101
Hughes Rand Koon CL, Applied Catalysis A: General, 119 (1994) 153-162
Hurlbert RC and Konecny 10, The Journal of Chemical Physics, 34 (1961) 655-658
Husen PC, Deel KR and Peters WD, The Oil and Gas Journal, 69 (Aug 1971) 60-61
320
Itoh N and Govind R, AIChE Symposium Series, 85 (1989a) 10-17
Itoh N and Sathe AM, Journal ofMembrane Science, 137 (1997) 251-259
James DH and Castor WM, Ullmann's Encyclopaedia Industrial Chemistry, 5th Edition,
25 (1994) 329-344
Jansen JC, Koegler JH, Van Bekkum H, et a!., Microporous and Mesoporous Materials,
21 (1998) 213-226
Jayaraman V, Lin YS, Pakala Met a!., Journal ofMembrane Science, 99 (1995a) 89-100
Jayaraman V and Lin YS, Journal ofMembrane Science, 104 (1995b) 251-262
Jeon GS and Chung JS, Applied Catalysis A: General, 115 (1994) 29-44
321
Jewett DN and Makrides AC, Trans Faraday Soc., 61 (1965) 932-939
Jia MD, Peinemann KV and Behling RD, Journal of Membrane Science, 82 (1993) 15-26
Johansson SAB, Campbell JL et aI., PIXE: A novel technique for elemental analysis, John
Wiley and Sons, Great Britain (1988) 1-347
Jung SH, Kusakabe K, Morooka S et aI., Journal ofMembrane Science, 170 (2000) 53-60
322
Kanoun N, Astier MP and Pajonk GM, Journal of Molecular Catalysis, 79 (1993)
217-228
Keizer K, Leenaars AFM and Burggraaf AJ, Eur. Coli. on Ceramics in Adv. Energy Tech.
(1982) 367
Keuler IN, Lorenzen L and Sanderson RD, Plating and Surface Finishing, 34 (1997b)
34-40
Kolb HJ and Burwell RL, Journal of the American Chemical Society, 67 (1945) 1084
323
Li A, Liang Wand Hughes R, Workshop: 'Applications and foture possibilities of
catalytic membrane reactors', Turnhout (16-17 Oct 1997) 106-109
Linkov VM, Preparation and application of hollow fibre carbon membranes, PhD
Thesis, University of Stellenbosch, Stellenbosch, South Africa (1994) 1-126
Liu B, Cao Y and Deng J, Separation Science and Technology, 32 (1997) 1683-1697
Loc LC, Gaidai NA, Kipennan SL et aI., Kinetics and Catalysis, 34 (1993) 451-455
Loc LC, Gaidai NA, Kipennan SL et a!., Kinetics and Catalysis, 36 (1995) 504-510
Loc LC, Gaidai NA, Kipennan SL et a!., Kinetics and Catalysis, 37 (1996) 790-796
Loweheim FA, Modern Electroplating, John Wiley & Sons, New York (1974) 342-357,
739-747
324
Lundin J, Holmlid L, Menon PG et ai., Industrial and Engineering Chemistry Research,
32 (1993) 2500-2505
Maestas S and Flanagan TB, The Journal of Physical Chemistry, 77 (1973) 850-854
Makarova MA, Paukshtis EA, Thomas JM et ai., Journal of Catalysis, 149 (1994) 36-51
Mendes FMT and Schmal M, Applied Catalysis A: General, 151 (1997) 393-408
McCool B, Xomeritakis G and Lin YS, Journal of Membrane Science, 161 (1999) 67-76
325
McKinley DL, US Patent 3,350,845 (1967) 1
326
Nourbakhsh N, Champagnie A and Tsotsis TT, AIChE Symposium Series, 85 (1989)
75-84
Ohashi H, Ohya H, Aihara Met ai., Journal ofMembrane Science, 146 (1998) 39-51
Paglieri SN, Foo KY, Way JD et aI., Industrial and Engineering Chemistry Research, 38
(1999) 1925-1936
Peachey NM, Snow RC and Dye RC, Journal of Membrane Science, 111 (1996) 123-133
Perra JA, Monzon A and Santamaria J, Journal of Catalysis, 142 (1993) 59-69
327
Perona JJ and Thodos G, AIChE Journal, 3 (1957) 230-235
Perry RH and Chilton CH, Chemical Engineers' Handbook, 5th Edition, McGraw Hill,
Tokyo, Japan, 1973, Chapter 3,1-250; Chapter 4,1-75
Quaranta NE, Martino R, Gambaro L et aI., Studies in Surface Science and Catalysis, 82
(1994) 811-818
Raich BA and Foley HC, Industrial and Engineering Chemistry Research, 37 (1998)
3888-3895
Raizada VK, Tripathi VS, Lal D et aI., Journal of Chemical Technology and
Biotechnology, 56 (1993) 265-270
Raman NK and Brinker CJ, Journal ofMembrane Science, 105 (1995) 273-279
Rao PR, Thangaraj A and Ramaswamy AV, Journal Chemical Society, Chemical
Commun., (1991) 1139-1140
Rao UR, Kumar R and Kuloor NR, Industrial and Engineering Chemistry Proc. Des.
Dev., 8 (1969) 9
Rezac ME, Koros WI and Miller SJ, Journal ofMembrane Science, 93 (1994),193-201
328
Rezac ME, Koros WJ and Miller SJ, Industrial and Engineering Chemistry Research, 34
(1995) 862-868
Rodina AA, Gurevich MA and Doronicheva NI, Russian Journal Physical Chemistry, 42
(1968) 959-960
Rosynek MP, Koprowski RJ and DeliiSante GN, Journal o/Catalysis, 122 (1990) 80-94
Sakamoto Y, Chen FL, Kinari Yet aI., International Journal 0/ Hydrogen Energy, 210
(1996) 1017-1024
Saracco G, Veldsink JW, Versteeg GF et aI., Chemical Engineering Comm, 147 (1996)
29-42
329
Sathe AM, Itoh Nand Tominaga M, Indian Chemical Engineering, Section A, 36 (1994)
131-135
Shirai M, Pu Y and Arai M et aI., Applied Surface Science, 126 (1998) 99-106
Shu J, Grandjean BPA, Van Neste A et aI., Canadian Journal of Chemical Engineering,
69 (1991) 1036-1060
Shu J, Grandjean BPA, Ghali E et aI., Journal ofMembrane Science, 77 (1993) 181-195
Shu J, Grandjean BPA and KaJiaguine S, Applied catalysis A: General, 119 (1994)
305-325
Shu J, Adnot A, Grandjean BPA et aI., Thin Solid Films, 286 (1996a) 72-79
Shu J, Grandjean BPA, Kaliaguine S et aI., Journal Chem. Society Faraday Trans., 92
(1996b) 2745-2751
330
Shu J, Bongondo BEW, Grandjean BPA et a!., Journal of Materials Science Letters, 16
(1997a) 294-297
Sivaraj CH, Reddy BM and Rao PK, Applied Catalysis, 45 (1988b) Lll
Skakunova EV, Ermi!ova MM and Gryaznov VM, Izv. Akad. Nauk SSSR. Ser. Khim., 5
(1988) 986-991
Smimov VS, Gryaznov VM, Miropo!'skaya MA et a!., Soviet Patent 437,743 (1978) 1
Swift HE, Beuther H and Rennard RJ, Industrial and Engineering Chemistry Prod. Res.
Dev., 15 (1976) 131-136
331
r
I
Takezawa N, Hanamaki C and Kobayashi H, Journal of Catalysis, 38 (1975) 101-109
Teo WK and Ti HC, Applied Biochemistry and Biotechnology, 24/25 (1990) 521-532
Udomsak S and Anthony RG, Industrial and Engineering Chemistry Research, 35 (1996)
47-53
332
Uemiya S, Matsuda T and Kikuchi E, Journal ofMembrane Science, 56 (1991a) 315-325
Uhlhorn RJR, Keizer K, Burggraaf AJ, Journal of Membrane Science, 46 (1989) 225-241
Uhlhorn RJR, Keizer K and Burggraaf AJ, Presented at the 1990 International Congress
on Membranes and Membrane Processes (ICOM '90), Chicago, lL, Session 25, 451
Van den Tillaart JAA, Kuster BFM and Marin GB, Applied Catalysis A: General, 120
(1994) 127-145
Van den Tillaart JAA, Kuster BFM and Marin GB, Catalysis Letters, 36 (1996) 31-36
333
Welch LM, Groce LJ and Christmann HF, Hydrocarbon processing, 57 (1978) 131-136
Wu JCS and Liu PKT, Industrial and Engineering Chemistry Research, 31 (1992)
322-327
Xomeritakis G and Lin YS, Journal ofMembrane Science, 120 (1996) 261-272
Xomeritakis G and Lin YS, Journal afMembrane Science, 133 (1997) 217-230
Yang BL, Kung MC and Kung RH, Journal of Catalysis, 89 (1984) 172-176
Yang BL, Cheng DS and Lee SB, Applied Catalysis, 70 (1991) 161-173
334
Yashima T, Suzuki H and Hara N, Journal of Catalysis, 33 (1974) 486-492
Yeung KL, Aravind R, Zawada RJX et aI., Chemical Engineering Science, 49 (1994)
4823-4838
Zaspalis VT, Keizer K, Van Praag W et aI., Brit. Ceram. Soc. Proc., 43 (1989) 103
Zaspalis VT, Van Praag W, Keizer K et aI., Applied Catalysis, 74 (1991) 223-234
Zhao HB and Baron GV, Workshop: 'Applications and future possibilities of catalytic
membrane reactors', Turnhout (16-17 Oct 1997) 52-58
Zhao HB, Pflanz K, Gil JH et aI., Journal of Membrane Science, 142 (1998) 147-157
Ziaka ZD, Minet RG and Tsotsis TT, Journal of Membrane Science, 77 (1993) 221-232
335
LIST OF SYMBOLS
336
I = film thickness [m]
L = distance from reactor inlet [m]
Lo = reactor length [m]
m = catalyst mass [kg]
M = molecular mass [g/mol]
n = pressure exponent or reaction order
nm = number ofN20 molecules that decomposes
ns = eu metal atoms per surface area [1.47*10- 19 m- 2]
P = pressure [Pal
Po = pre-exponential factor of permeability coefficient [moLmlm2 Pa,s]
Per = permeability coefficient [moLmlm 2 ,Pa,s]
Pm = permeance [mol/m 2 ,Pa,s]
P, = total pressure [Pal
337
I"~
Greek symbols
a = separation factor
eb = void fraction of packed bed
SubscriI!ts
0 = inlet conditions
2But = 2-butanol
A = feed (ethanol or 2-butanol depending on reaction)
Ac = acetaldehyde
b = at bulk conditions
Et = ethanol
Hl = hydrogen
= component i
338
J = componentj
m = mixture properties
MEK = methyl ethyl ketone
Nz = nitrogen
obs = observed
R = main products (acetaldehyde or MEK depending on reaction)
s = on the shell side ofthe reactor
surf = conditions at the catalyst surface
S = hydrogen
t = on the tube side of the reactor
W = water
x = high pressure side
y = low pressure side
Note:
• The reaction rates and reaction rate parameters (k, K and r) described in Chapter 2 were
taken from published work in their original fonnat and not transfonned to standard units.
In each case the published units were stated.
• For kinetic testing in Chapter 5, the reaction rates were expressed in (mollkg.cat.h) and
the parameters used kPa instead of Pa for their units. Once the parameters were
detennined in kPa, it was changed to Pa for modeling in Chapters 8 and 9.
339
/
I
I LIST OF ABBREVIATIONS
i AA atomic adsorption
BET Brunauer-Emmett-Teller
CNRS Centre National de la Recherche Scientifique
CVD chemical vapour deposition
dc direct current
ECN Energy Corporation of the Netherlands
FID flame ionisation detector
GC gas chromatograph
HP Hewlett Packard
ICP inductively coupled plasma
IRC Institut de Recherches sur la Catalyse
MEK methyl ethyl ketone
PIXE particle induced X-ray emission
SCT Societe des Ceramiques Techniques
SEM scanning electron microscope
TCD thermal conductivity detector
TEM transmission electron microscope
TOC total organic carbon
XRD X-ray diffraction
340
r
APPENDIX A
341
r
I
I
342
APPENDIXBl
343
Apparatus: HP GI800A GC and
50 m capillary column (50QGI.5IBPI PONA from SGE)
This apparatus was used for all experiments where the catalyst composition was
optimised (Chapter 4) and for membrane reactor experiments (Chapters 8 and 9). The
response factors discussed in Appendix Bl are only for the work in those two chapters.
344
Ethanol response factors:
...
~
70
60
~
•E 60
,
~ 40
~ 30
20
10
0
0 ~ • 80 80 100
GC area % of other component
...".. 80
Ethane
.... 50
40
E
-
~
30. y = -2.72E.07x 3 • 4.61E-03x 2 -+ 1.46E+OOx -+ 7.87E-01
..." ~ . R2 = 9.gsE.Ol
10
0
0 20 40 80 80 100
GC area % ethene
Excluding acetaldehyde, di-ethyl ether and ethene, the majority of other by-products that
formed for all ethanol catalyst optimisation experiments (Chapter 4) were below 5 GC
area %. Exceptions were for MgO catalysts, where the following maximums were
detected in different experiments:
propene: max. 10 GC area %
butenes: max. 15 GC area %
I-butanol: max. 16 GC area %
345
For some Cu on silica optimisation experiments (for ethanol dehydrogenation), up to 12
GC area % propene and 9 GC area % I-butanone were detected. For Si0 2 and MgO
supports, those high percentages were detected at and above 400°C and at very low
ethanol feed flow rates.
For area percentages below 5%, the response factors were determined or estimated. The
measured response factors correlated reasonably well with the molecular mass ratio
between ethanol and the other component. This observation was used as a very rough
approach to estimate the other response factors as indicated.
346
The error that resulted from this simplification will have little if any effect on the
membrane reactor work and model development. The error was restricted to the catalyst
optimisation stage, where the same calculation procedures were used for all catalysts
tested. Data obtained during the catalyst optimisation stage were not used for
fundamental calculations and model development, but only used to compare catalysts.
Since the error will be similar for the different catalysts tested, the simplifications in the
response factors will not have a significant effect in identifying the optimum catalysts.
For membrane reactor work, the acetaldehyde selectivities were above 80% and no
alkanes or heavy products formed. Ethyl acetate was the only component of more than 2
GC area %, which did not have a measured response factor. Even a very large difference
between the estimated response factor and the true response factor for ethyl acetate, will
have a negligible effect on the total ethanol conversion and the acetaldehyde yield.
60
...
W 60 MEK
~
~ 40
2
I-
20
o 20 40 60 eo 100
Gearea %
The 2-butanol-butene response factors were determined in the same way as that of the
ethanol-ethene response factors. MEK and butenes were the only products that formed in
the catalyst optimisation experiments. For membrane experiments, 3-octanol was the
only by-product that formed, with a maximum value of up to 3.5 GC area %. The
response factor was taken as 0.5.
347
APPENDIXB2
348
Apparatus: HP 5850 GC equipped with a FID detector and
30m HP Innowax and a 30m HP Plot/AhO) columns in series
This apparatus was used for al experiment where the kinetics of the optimised catalysts
were investigated (Chapter 5).
Acetaldehyde area%
..
.s::.
"C
50
40
-..
ii
~
30
20
y = -1.21 E-Q4x' + 1.29E-02x' + S.72E-01x + 3.96E+OO
R' = 9.93E-01
10
0
10 20 30 40 50 60 70 SO 90 100
Acetaldehyde area%
349
For the kinetic experiments, the selectivity of the reaction towards acetaldehyde
formation was above 97%. The ethanol-acetone response factor was measured as I. All
response factors excluding ethanol-acetaldehyde were taken as I; in other words, the true
mass % was taken as the GC area %.
Note:
For all gas samples, i.e. for determining response factors as well as for all reactor
experiments, a heated syringe was used for injections into the GC. A heated syringe was
used to prevent condensation of the products in the syringe. A very small tube furnace
was made by wrapping heating wire around a ceramic tube and insulating it. The syringe
was placed inside the tube prior to taking a gas sample and the temperature inside the
tube was controlled at 110°C. Only the bottom halve of the syringe was heated to still be
able to handle it at the top end. A very small gas sample was extracted (about one
fifteenth of the syringe volume). This procedure gave very reproducible results when the
same gas mixture was injected into the Gc. When the syringe was not properly heated,
there was significant variation in the results due to condensation of some of the products
in the syringe.
For ethanol-acetaldehyde mixtures (liquid), the syringe was not cooled, but used at room
temperature when the response factors were detennined. At high acetaldehyde
concentrations (> 80%) some flashing of acetaldehyde from the syringe might have
occlUTed and the true response factors might differ marginally from the measured values.
350
APPENDIXCI
351
The linear equations determined by least square analysis at each temperature are:
At T = 200 °c:
For an ethanol-N2 feed:
y= 7.21 *10-3x + 1.02
AtT = 250°C:
For an ethanol-N2 feed:
y= 3.76*1O-3x+0.73l
352
For an ethanol-acetaldehyde feed:
y = -4.76*1O-zx + 5.78
AtT = 275°C:
For an ethanol-Nz feed:
y = 2.02* I 0-3 X + 0.468
K-values
The K-values were detennined using both the gradients and the y-axis intercepts of the
fitted lines. The values are summarised in Table Cl.
353
Table Cl: K-values for ethanol dehydrogenation
T (0C) Ks(I) Ks(2) Ks KR(I) K R(2) KR
1 1 1 1 1
kPa- kPa- kPa- kPa- kPa- kPa- 1
200 0.00772 0.00732 0_00752 0.1182 0.1212 0_1197
225 0.00660 0.00715 0.00688 0.1014 0.1061 0.1038
250 0.00314 0.00334 0.00324 0.0684 0.0703 0.0694
275 0.00298 0.00283 0.00291 0.0652 0.0661 0.0657
300 0.00401 0.00334 0.00368 0.0811 0.0825 0.0818
354
APPENDIXC2
355
The linear equations detennined by least square analysis at each temperature are:
AtT = 190°C:
For an 2-butanol-Nz feed:
y = 2.461 *10-3x + 1.717 (RZ = 0.452)
AtT = 220°C:
For an 2-butanol-Nz feed:
y = 2.213*10-3x + 1.138 (R Z = 0.575)
AtT = 250°C:
For an 2-butanol-Nz feed:
y = 2.282*10-3x + 0.764 (RZ = 0.636)
356
For an 2-butanol-MEK feed (0-41.0 wt% MEK):
y = -3.824*10·zx + 4.745 (Rz = 0.955)
AtT = 280°C:
For an 2-butanol-Nz feed:
y = 1.764*10·3x + 0.547 (RZ = 0.874)
K-values
The K-values were determined using both the gradients and the y-axis intercepts of the
fitted lines. The values are summarised in Table C2.
357
APPENDIXDI
358
Membrane Thickness T Avg.H2 Avg. N2 Selectivity Arrhenius parameters for R value for n-I
(micron) ('C) penneance penneance hydrogen penneance in permeance
(J.Ullollm1pa.s) (nmollm2.Pa.s) equation
359
Membrane Thickness T Avg. H, Avg.N, Selectivity Arrhenius parameters for IV value for n-I
(micron) (0C) penneanee permeance hydrogen permeance in penneance
2
(f1II1oUm' .Pa.s) (nmoUm Pa.s) equation
360
Membrane Thickness T Avg. H, Avg. N, Selectivity Arrhenius parameters for R value for n-I
(nticron) (0C) penneance permeance hydrogen permeance in permeance
(l1fIloVm' .Pa.s) (nmoVm'.Pa.s) equation
N2 1.08 450 15.11 20.80 728 Po - 5.267E-II 0.998
410 14.37 19.50 737 En- 6984 0.9994
370 13.38 18.10 738 R'· 0.9857 1
330 11.99 17.20 698 0.9999
361
Name: la (3.08 pm)
~ .' - -" --
-
35.0
---
-----
220
~
C
0:
•
34,0
~ --
:: ...D.:
-,,- --
EE
~i5
E
330
---.-
---- - ------- ~
__ Nftrogen permsanco
Z S 32.0
,,- -
"
--- -.-. Hydrogen/Nitrogen
''''
--
31.0
selectivity
30.0 100
330 370 410 450
T('C)
362
Name: Ib (3.86 Jim)
10.6~
,,,~~ 650
t--~--=::".-=:::::::::===-7'---'-'-1:·"
t:e:I ~
; 10.2
/
:t::
>
ii
z.s
9.8
/
---------------~
-. .,
J __ . :
363
Name: Ie (6.19 Jim)
'.3
BOO
.n'."
u
u
c
•
•
~
'.1
q 5."
zE.,
600
5.7
••-. HydrogenINitrogen
r-- 550
selectivity
5.5 l-------------__============~~~o
370 410
T('C)
...
364
Name: 2a (1.43 Jim)
10.1 750
__ Nitrogen permeance
--
10.0
."
. •••
: .,
8c ..
~
------
- .... Hydrogenn-litrogen
----...".
selectivity
-
----
- ... ~>
n '.8 '. '. ~.~ ...
-- II
•
it • " , / ...,;lj
" E
zs. 9.7
_t
- --- .' ..,
".
. J5()
9.'
---
9.'
330 370 '10 ...,250
TI"C)
365
Name: 2b (1.43 J.lm)
13.1 "-- .,
, . -- -'
900
~ ,--- ,, -'
BOO
,
~
/~ 700 jj
, " - -- --
,-
..
- -
,, -
.-
-+- Nitrogen penneance ~ .eo
.!!
.ll
11.9
-- - __ . Hydrogen/Nitrogen 500
selectivity
11.5 400
330 370 410
T('C)
366
Name: 3a (2.40 fll11)
6.5 1500
,,
.
6.3
--- ,,
, 1400
penn~ ,, 1300
,,
I- ~
><
-+--Nitroge"
1200
l-
-.-, Hydrogen/Nitrogen
/' 1100 IJl
selectivity
., , ,
~
,
/
., " 1000
5.7
, ,
, ,-- , ,
, , ,
, ,
" 900
5.' 800
330 370 410 450
T('C)
367
Name: 3b (4.43 11m)
6,5
5
11 4,5
i A
. ,
~ ,,5
~ - . - Nitrogen permeanca ~
selectMty
3. / /'
Ii ~ 1.3 ~ w" 3500 t
ll,.2 t--------'.=__=====--=-=o:==----t J 3000
z .s. ._-
~~~
_______
----
1.1 -I---:~"-'----------------t 2500
1.0 --- .-
+-_____----_-------+2000
330 370 410 450
T("C)
368
Name: 5 (1.00 pm)
16
15
114
11:1
;a 1~
lIS 11
.,p
<1P
..., of>
',;;-'!P -,»
...""#
369
Name: 6 (1.47 11m)
- • . Hydrogen/Nitrogen
selectMty
36.0 L-----_-===:::;:==='----1. 200
330 370 410 450
TI'C)
370
Name: 8 (1.10 11m)
j 1~·6
~~ -----------_ 800 ~
~
11.5
10.6
=.
11
225
no
..
. - .---
~~
--- 550¥
~
j
4> z! 21.5
-- --
,4>. .
__ Nitrogen permellnce 500
...
4> 21.0 ,-
...
,;>
...~ ...4>
20.5
20.0
"0 370
..o-
-.- HydrogenJNl1fogen
r-
'10 ""
T('C)
% H2 permeated with N2 sweepgas and space % H2 permeated with N2 sweepgas and space
lime = 2.885 lime = 1.335
99.8 100.0
;j ---
- ..
99.6 .. ,_. ';:-'~~'''~--'';:- :--
-~
... .... ~:.>-.-
99.0
A>""-'
"",'-"
i
E
i
99.2
/~ .........
':/:' .- !
; IE 98.0
97.0
~""
J.'
I
.I
".0
../.~/' ! t I
£ 98.8 ; :£ 'B.O I
... ."F
. ; ... __ T= 45Q·C _ _ T=410°C
--+-T = 450°C _T",410·C
, I
1
98.S
".2
0 , 10 15 20
94.0
0 5 10 15 20
N2 sweep rata as molar % of H2o feed N2 sweep rate as molar % of HI! fead
371
Name: Nl (1.12 Jim)
'"' __'""----------------t""
u + - - - - - - - _ - -_ _ _ _ _-+""
330 370 410
Trel
% H, permeated with N, sweepgas and space % H, permeated with N, sweepgas and space
time =2.375 time 1.195 =
".. --.--------------------...--------- .--------------------------~=~=-~-;t 102.0
1I -~ ".0
99.' -I-----;--:y~<'"'.9""---------
I
~
96.0
: I
.'/1 ==F~~~
94.0
:'1
l ::1=4'.'/:='
:£
iJt. 98.8 -T-45Q"C _T_410·C
i!.
:£
...
92.0
90.0
• I
+-_-t--____--jI--T=450°C __ T-410·C [
...... t=="~"/=:=:=~1.~.=.T=-~37=O=.C~-=.T=.330::=.=C~
88.0
I· ". T=370·C - -T"m"C L
88.0
84.0
•
o 5 10 15 ~ ~ ~ ~ ~ 0 5 10 15 20
NI sweep rata as molar % of HI feed N2 sweep rate as molar % of HJ; feed
372
Name: N2 (1.08 I'm)
22 700
370 410
T("C)
% H, permeated with N, 5weepga5 and space % H, permeated with No sweepgas and space
time =2.375 time -1.195
. . ~~~=======
100,0 -.------------- 100.0
J
99.'
99.4
j __ T = 450'C -+-T = 410'C l
,11-------
~•
".'
99.' f-- _T::z450"C -+-T-410'C
~
98.2±===::!=:'======:::====,
".0
o 5 10 15
Hz swa.p rate as molar % of Hz feed
20 o
£
5 10 15
N z sweep ra.. as molar '% of Hz feed
20
373
Name: N3 (1.47 Jim)
2000
"""
5~ ______________ ~ ______________ ~1~
374
Name: N4 (1.18 J'm)
l'
13
22
21.5
,
'00
'"
Is 10
1~
\1 .
8
Ii .
11E
~
t;' 21
2{l5
\
.-
" -.'
'.
--"
"
--- ---,--_..---
---.- ./"/-'
,.,
"'~
i
~
...01'> £ So .........
20 ---. Nltrogeo permeance
""
4> - ...
-
"-'"
01>
...... .,;?
19.5
"
HI-
- .-. HydrtJgen'Nllrogen
.10 .....,
""
37<)
TrC)
% H,. permeated with N2 sweepgas and space % H2 permeated with N2 sweepgas and space
Ume = 2.375 time =1.19s
~ -:>I"';==~~~=:::::::===:'
100.0
100.0 __
98.0
~.
I I
96.0
99.0
¥
no'·
I: I
94.0
8.
98,5
&. fI:' .'
92.0 [--r:..so'c __ T 410'C
.,.
-£ •i
-£
"-
98.0
-j:(
// J-r 450'C - T 410 'C~ ~ 90.0 ·""·T-370·C "·T-330'C
97.5 t--f'-.-----il-..t..- T:370 "C -- -T=33O·C 88.0
97.0 88.0
0 5 10 15 W ~ ~ ~ ~ 0 5 10 15 20
Nl sweep rate as mohIr % of H, feed N, sweep rata as molar '% of H2 feed
375
Name: N7 (1023 11m)
'2~~~------------~,oo
, .• + - - - - - - - _ - - - - - - - - 1 0 0 0
330 )70 .10
TI°C)
% H2 permeated with N. sweepgas and space % H2 permeated with N2 sweepgas and space
Ume= 2.37s Ume = 1.195
99,0 100.0
98.5 96.0
i ... ; ,
I
98.0 96.0
•~ 97.5
~~~ ..O.·:>-·-·- 94.0 .. .'
I 97.0
""'. .....- it 92.0
%:
'it 96.5
......
.' .-;
- ,..- l_r=450'C _1=410':f %:
~
90.0 , __ T-450'C
1.•.. Te370·c
__ T .. 410·C
_·T_330"C
96.0 +---~"'"------1I_·'IIo· T=370'C _·T .. m"e 88.0
95.5 86.0
0 5 10 15 20 0 5 10 15 20
Nt sw.ep rate as molar % of Hz feed Nt SWIHlP rate as molar % of HI: feed
376
Name: N8 (lo19J1m)
I 10 2.'
2.' , ",
5500
I1~
0
I 8
tI ii' 2.4
22
-,
/
,
,
'/<"
5000~
• ~
..
2 ..00
i'-, /
~ 1.8 ,
.-------- 4000 JJ
..tP
..".
..<§>
z .s i.e
/
--I'a'ogen perm.... c:e
<:P~-6> .,;....
'.2 91111:1....1'1
"
"" 370
T(°C)
410 ... >DOD
% H2 permeated with N2 sweepgas and space % H, permeated with N, sweepgas and space
time = 2.375 time 1.195 =
....
99.7 102.0
100.0
_ T = 450 "C __ T .. 410·C /."/
:g 99.'
I
98.0
......;~/
II 99.5 - -,.,·T-370·C _"T=33Q'C
., 98.0
i 99.5
........."
"..
~
~
~
.;Y
./
./ ~
!.
94.0
•
'%,.. 99.4
;jt. 99.4
99.3
...
~
/' .
;£ 92.0
90.0
88.0
_ T _450·C __ T:410·C
377
APPENDIXD2
378
Alloyin2 procedures
NI Pd No
Nib Pd+Ag No
Nix NewPd+Ag Yes Reactor H2 590 10
N3 Pd No
N3b Pd+Ag No
N3e Pd+Ag Yes Tube oven AI 545 50
N3d Pd+Ag Yes Tube oven AI 545 100 (total)
N3e Pd+Ag Yes Tube oven AI 545 150 (total)
N4 Pd No
N4b Pd+Ag No
N4c Pd+Ag Yes Tube oven AI 530 30
N4x NewPd+Ag Yes Reactor H2 540 30
When two temperatures and two times are given in a row, the ftrst heating time is for the ftrst temperature
and the second heating time is for the second temperature.
379
Membrane Thickness T Avg.H, Avg. N, Selectivity Arrhenius parameters for R" value for n I
(micron) (0C) permeance permeance hydrogen permeance in penneance
(fllIIoVm'.Pa.s) (nmoVm'.Pa.s) equation
380
¥+,~-
Membrane Thickness T Avg. H, Avg. N, Selectivity Arrhenius parameters for R< value for n I
(micron) (DC) permeance penneance hydrogen perrneance m penneance
(!illIoVm' .Pa.s) (nmoVm'.Pa.s) equation
381
Name: 8b (1.54 J,1m)
11.3+-_ _ _ _ _ _ _ _ _ _ _ _---1
11.0+-------_--------1300
330 370 410
% H2 permeated with N2 sweepgas and space % H2 permeated with N2 sweepgas and space
Hme =2.865 Hme = 1.33s
100.0 100.0
.... ~~------- ...
~j 99.'
,,":--,
99.' .-. ,
.,.-..... --
....~:"' r-: ;:-.:;.::-- 96.0
.. I, 94.' .'
I 99.' .-;,..-:..
.... "; ...
,
i I
~
92.'
/
i 98.5
J' .17
i
!. 90.'
.
;£
980
• _T::410°C - •. T"370OC
i
H .. '"'.
;£ 88• •
;
,. ~T::410GC .•. T-370°C
~
... ·T-330·C
.... ·T_330°C 84.'
97.5
97.0
i
i
82.0
80.'
" ,
0 5 10 15 20 0 • 10 15 20
Nt sweep rate as molar % of H2 feed N2 sweep rate as molar % of H2 feed
382
Name: NIb (1.40 11m)
B i i 520+---~"--------7"'/---foo
~t. / /.... f
i="'E:w 48.0
--.- ~
/
. ,/ 85 t;
•
; ! ~.oh-'-/-''------------------''''~-··-,.,,----r-j80 i
,/ - - NI\rogm ptnneanco
'k H2 permeated with N2 sweepgas and space 'k H, permeated with N, sweepgas and space
time 2.375 = time -1.195
100.0
--':-~''::;i?-:;;~;;-~;;;;;;;';;~'~=:=;.o;-=--
100.0 -.--.----.-.. ------.-.--.-.--~-;:-;;-::-:::;;.:--.::-.::-::~~:~..:~~.-,:=-.;.'1
95.0
95.0 ,.
,r 90.0
I• 90.0
~ 85.0
80.0
~
i!.
65.0
l 75,0
/
r
..
:£ BO.O
75.0
•
1_.
_T-410'C
T ",330 C
0
-,.. T .. 370·C
..
-£
70.0
05.0
60.0
55.0
J __ -T:::330'C
+-------,.'"'----i
•
__ T .. 410·C - .... T=370'C
[
70,0 SO.O
0 5 10 15 ~ ~ 30 ~ ~ 5 10 H
Nt sweep rat. as molar % or HI teed Nz sweep"" as molar" of ~ feed
383
Name: N3b (2.14 11m)
% Hz permeated with Nz sweepgas and space % Hz permeated with Nz sweepgas and space
=
----
time 2.375 time = 1.195
100.0 100.0
95.0
"
..
.,' - .;'
96.0 90.0
~ f1".' • I.: f
~
•~
!.
92.0
88.0
:
i
1
I
!.
85.0
80.0
75.0
70.0
I:
/:
I: .'
.'
,
-£
~
I-T=410°C ..... T=370'C
-£
~
65.0
:
_ T 410'C .... T=370'C t-
84.0 +--------1,_ -T=3JO'C 60.0
55.0
+--ri------IJ -- -T = 330'C t
80.0 50.0
0 10 15 ~ ~ ~ ~ ~ 0 5 10 15 E ~ E ~ ~
Nt SWHP rate as malar % of Hz feed Nz swe.p nita as molar % of HI feed
384
Name: N3c (2.14 11m)
6.&
G,A
Selecltivity data for N3c
..
i 5.:
~ 5~
".+----------------:=~_=J
..' '"
..~""
17.0+-----------------1
1 1 1 . 0 + - - - - - - - _ - - - - - - -.. 200
330 370 410
T("CI
% H, permeated with N, sweepgas and space % H, permeated with N, sweepga5 and space
time = 2.375 time = 1.195
100.0,.----------------- 100.0,.-----------------
911.0+----------------- 95.0
i• 96.0 I 90.0 _ :f
~
!. ".0
i 85.0
:'.'
:£ -£
"-
i· , J_T=410°C ••" T=370"C
"- 80.0 t - - ; i - - - - - - - j
92.0
• !__ 'T_330°C
90.0 75.0
0 5 10 15 ~ ~ E ~ ~ 0 5 10 15 ~ ~ E ~ ~
Nt sweep rata as molar % of Hi feed Nt sweep rata as molar % of Hi teed
385
Name: N3d (2.14 Jim)
, ..
% H, permeated wtth N, sweepgas and space % H, permeated wtth N, sweepgas and space
time s2.37s time -1.195
--
100.0 100.0 _.• -----------. ---.--_._---------.'------------.. _----.'---.-.--------- ..--... _.---.------.-.
95.0 +-___.... -::-:'-="=~:_'-':"-"'-;':'-"'.;,;.~
.•",
.•;,:.-",.-",.-':..-:.=.
-::-::-.-:.-.~: ~:.:.-.-.
-'
.. -:':.·-"",-1
96.0
IE " ..
90.0
I~
92.0
• iI.
80.'
Go
-£
88.0
.....
.: ,,'
, -£ 75.0 __ T-410·C - .... ··T=370·C
;II
84.0
I
t--~-.T'------I! __ -r ..
-T;410°C
33O'C
.... T=370'C
;II 70.0 1 __ .T 330"C
65.0
80.0 60.0
0 5 10 1S ~ ~ ~ ~ ~ 0 5 10 15 ~ ~ ~ ~ ~
Nz sw.ep rate as molar % of Hz feed Nt sw..p rate as molar % or Hz feed
386
Name: N3e (2.14 Jim)
\
, -~-' ''"
5 .......
I( 6rA
-"'''',
'"' ~
••6 ~"e
&. ~
:f!
6B.D
.
·~,:x::.a'
~
.I
... ..
"'" ,--- ---....--- .
__ Ntogen p.-me.'lCCl
,
. "'"
- .... HyctogaVtlltrogen
" - ...? seledl\o1ty
70
'" 4>
oS> ."."...
'"
T ("CI '"
% H. permeated with N. 5weepgas and space '10 H. permeated with N, sweepgas and space
time s 2.375 time =1.195
100.0 , - - - - - - - - - - - - - - - - 100,0
I
II.
92.0
III.
'M
80.0
...
.
-£
".0
64.0
..
~~"~.' J--T=410-C . '110- T=370·C
.
-£ 75.0
70.0
'5.0
I__ r
_T=410·C
33O"C
.... T=370·C
80.0 80.0
0 5 10 15 ~ ~ ~ u ~ 0 5 10 15 ~ ~ ~ H ~
Nz sweep rata as molar % of HI teed Nz sweep rata as molar 'I. of Hz faed
387
Name: N4b (1.43 11m)
i
10' 1
'"
% H, permeated with N, sweepgas and space % H, permeated with N, sweepgas and space
time 2.375 = time = 1.198
--··-K:;':~:·:::'::;;~:'=;=:::=-;':;':-==:';':---------~ 100.0 .•. ----.-.----.---------.----.---~:~~~?~!~:~:::-:::-::.-=':=:~~-""II
"'~.
i 70.0 l--~."':-'," " - - - - - - - - - - - - -
... " 1"
~
~ 80.0 l-~:['-----f~~~~~~~
... l--T."10~C T_370'C -'II._ ,;:
'if!. "" +~.'
-.0 I--...::,,:·.i/-----r::=:;=;::;:--:-::-;;~;;:]
l--r=410 c - 'IL- T=370'C
0
75.0 t--~"-----L....
~·T~-..:330~.C'__ _..._J 50,0 j--=------I..::....~.T.:.~~330~·::.C_ _ _..J
7O.0j......-_-~-~_-~-~-~-~ ...ol--_-~-_ _-~-_-_-~
o 10 15 ~ ~ m ~ ~ o 10 15 ~ ~ m ~ ~
N2 SWHP rata as molar % of HI feed Nz SWHP rata as molar % of H,: feed
388
Name: N4c (1.43 11m)
- .-. HydrogenlNltrogen
selectivity
14.0 L ___--.-'==~;::::===;='...----to
n "" _ ~ m m
389
Name: N4x (2.13 Jim)
Ii 2"
27.0
...... ",.
~.
- ... ' l-/yItogenlNlrogB'l
9IId\4ty
.. ...'
J'"
.., f
~ '"E
.'
2t.O ". Ii
a.'"!.
%' 2U
215.0
.", ..><:....
------ ""
li
..- ---~
" .• '00
% H, permeated with N, sweepgas and space % H, permeated with N, sweepgas and space
arne = 2.375 time = 1.19 5
100.0
95.0
[==::::;;;;~.;_;._~~~~.~.~.~.~
~>.-'
. ".0
~ 90.0 . ,.,
fI":-" .rI
~ 90.0
~
II.
".0 , ~
3. ".0
.'
.'
,
..
,£ .<10
75.0
__ T:370'C _T_330"C
r · ..··T:290·C _'T=25Q"C
,£
;I-
80.0
75.0
+-------,"'-------1 _ T " 370·C
• .' J '6' T=290"C
__ T '" 33O'C
_'T=250'C
70.0 70.0
0 10 15 E ~ ~ ~ ~ 0 5 10 15 ~ ~ ~ ~ ~
N2 sweep rata as molar % 01 HI feed H2 SWHP nrte as molar % of HI feed
390
Name: N8x (2.16 p.m)
§ ~D.
til 21.0+
I __ ··...'.\¢==H=-===~'"".,::::/-=--.-'-----t
,
280 ~
'S:
•
E :§
N'
iE
90.0
".0
.-:.' i 85.0
f/;' .I
.
• 0
• I E
1 0
• ".0
~
,
I '(lO
: I
_ T = 370°C -...-T""330·C-
N
l: 92.0 _T-370·C ___ T _330°C I- -£ : I
o 0
__ ·T= 250·C -
~ 75.0 - .. ·T .. 290"C
~ I
-,.,- T:290·C ___ ·T"Z50·C i- ;'
90.0
• 70.0
65.0
~
88,0
0 5 10 15 20 25 30 35 40 0 5 10 15 20
Nt sweep rate as molar % of Hz feed
25 30 35 40
,
Nz sweep rate as molar % of ~ feed
391
For some membranes a full characterisation was not done:
392
Name: N2x (2.02 Jim)
Mass (g) (mm)
Clean 28.5978 Plated length 250
After pretreatment 28.6115 Permeable length 205
393
APPENDIXE
394
"'"
----------------------------------------------------------~.'
-'-:.-~ -
CVD in pores Porous Vycor Si02 450, 100 0.0191 25 Tsapatsis (1991)
of membrane 4 nm (Coming) Ah 0 3 450, 100 0.0020 210
tube B2 0 3 450, 100 0.0068 20
Ti02 450, 100
-_. -
0.0041
'----. _.
63
395
I
l
Preparation
method_ _
J
Support
description
Coating
composition
Thickness
(I'm)
Temperature
(0C); OP (kPa)
Permeance
Mol/m2/s/Pa
Selectivity
H2/gas
Reference
396
I
Preparation
method
JSupport
description
Coating
composition
Thickness
(Jim)
Temperature
(0C); 1)P (kPa)
Permeance
Mol/mz/s/Pa
Selectivity
Hz/gas
Reference
397
Preparation Support Coating Thickness Temperature Permeance Selectivity Reference
method description composition (JIm) (0C); liP (kPa) Mol/mz/s/Pa Hz/gas
CVD (In pores a-alumina Pd 4 300; 100 4.0 10000 (N2) Van (1994)
+ outside tube) l50nm(NOK) 500; 100 4.0 6700 (N2)
CVD (In pores a-alumina Pd 4.4 400;200 0.38 3000 (N2) Aoki (1996)
+ outside tube) l50nm(NOK) 500;200 0.60 1900 (N2)
CVD (In pores a-alumina Pd 4.4 400 0.80 1330 (N2) Kusakabe (1996)
+ outside tube) 150 nm (NOK)
CVD (In pores a-alumina Pd 4.4 400 0.7 1160 (N2) MOTOoka (1995)
+ outside tube) 150nm(NOK)
398
Preparation Support Coating Thickness Temperature Permeance Selectivity Reference
method description composition (/lIn) (oq; OP (kPa) Mol/m2/s/Pa H2/gas
Plating (electr) Porous Vycor Pdgg -Ag 12 1.2 500; 170 anneal 0.10 Cheng (1999)
Outside tube 5 nm (Coming) 400;170 0.074
Plating (electr) Porous glass (8) Pd93 -Ag7 21.6 400; 196 0.32 - Uemiya (1991b)
Outside tube 300 nm (lSI)
Plating (electr) Porous glass (s) Pd 94 -C U 6 18.9 400; 196 0.21 - Uemiya (1991b)
Outside tube
~ -
300 nm (lSI) ----
399
Preparation Support Coating Thickness Temperature Permeance Selectivity Reference
method descriptioD_ composition (11m ) (0C); OP (kPa) Mollm2/s/Pa H~gas
Magnetron air-alumina (a) Pd7SAgz s 0.4 300 0.15 80 (He) Xomeritakis (1997)
sputtering 4.5nm
(disc)
Dc sputtering air-alumina (a) PdnAgs 0.33 300 0.071 63 (He) McCool (1999)
(disc) 4nm
Spray pyrolysis air-alumina (a) Pd7~gz4 1.75 500 0.080 24 (N2) Li ZY (1993)
Outside tube 5 nm(NOK)
Sputtering MgOss-Ytrials Pd~lI6 10 700; 1000 1.23 Gryaznov (1993)
300 nm (s)
Sputtering Porous glass Pdn Ag23 6.0 380; 125 0.056 Gobina (1994) i
Sputtering Poly(dimethy]si Pd76A g24 0.05 25 0.003] 100 (CO 2) Athayde (1994)
loxane) coated
Membrane (a)
Sputtering Polyarilyde (a) Pd94RlI6 0.4 200; 1000 0.16 ._-
Gryaznov (1993)
400
Comments
1. All permeances were changed from the original units reported to j.Ullolf(m2 .s.Pa)
2. Open cells in the data pages represent data not supplied in the specific reference.
3. Selectivities were only included if permeation data for the other gas was listed as a function of
pressure and/or temperature.
4. In one reference (!lias 1997), the graphs and values mentioned in the paper contradict each other
and was thus not included.
Abbreviations
ECN: Energy Corporation of The Netherlands
GFM: Goodfellow Metals, Berwyn, P A
GTC: Golden Technology Company, Colarado
lSI: Ise Chemical Industry Company
MM: Mott Metallurgical
MPV: Metallurgical plant, Yyksa, Russia
NOK: NOK Corporation, Japan
SCT: Societe des Ceramiques Techniques
SS: Stainless steel
TCC: Toshiba Ceramics Co.
TK: Tanaka Kikinzoku, Japan
USF: US Filter Corporation, Warrendale, Pennsylvania
YCMC: Velterop Ceramic Membrane Company of The Netherlands
(a): asymmetric
(s): symmetric
401
APPENDIXFl
402
Membrane name: N8x Catalyst: 3.00 g (14.4 wt% Cu on silica)
Reaction temperature: 250°C Catalyst particle size: 500-800 microns
Feed Sweep to Total ethanol Ethanol conversion Ethanol Ethanol to acetic Reaction selec-
rate feed molar conversion to acetaldehyde conversion to MEK acid ethyl ester tivity towards
mlfh ratio acetaldehyde
Average Average Average Average Average
5 0.00 35.7 27.8 0.7 2.8 78.0
27.1 26.2 0.5 0.4 96.7
34.5 32.4 30.3 28.1 0.7 0.6 1.4 1.5 87.7 87.4
403
Membrane name: N8x Catalyst: 3.00 g (14.4 wt% Cu on silica)
Reaction temperature: 250°C Catalyst particle size: 500-800 microns
Feed Sweep to Total ethanol Ethanol conversion Ethanol Ethanol to acetic Reaction selec-
rate feed molar conversion to acetaldehyde conversion to MEK acid ethyl ester tivity towards
mIlh ratio acetaldehyde
Average Average Average Average Average
10 0 27.4 26.4 0.5 0.5 96.3
28.5 27.2 0.5 0.5 95.2
30.3 28.8 27.7 27.1 0.7 0.6 1.0 0.7 91.4 94.3
404
Membrane name: N8x Catalyst: 3.00 g (14.4 wt% Cu on silica)
Reaction temperatnre: 275 DC Catalyst particle size: 500-800 microns
Feed Sweep to Total ethanol Ethanol conversion Ethanol Ethanol to acetic Reaction selec-
rate feed molar conversion to acetaldehyde conversion to MEK acid ethyl ester tivity towards
mlfh ratio acetaldehyde
Average Average Average Average Average
10 no sweep 47.0 38.2 0.9 2.7 81.3
45.2 46.1 38.2 38.2 0.9 0.9 2.0 2.4 84.5 82.9
405
"
:;;;
Feed Sweep to Total ethanol Ethanol conversion Ethanol Ethanol to acetic Reaction selec-
rate feed molar conversion to acetaldehyde conversion to MEK acid ethyl ester tivity towards
mllh ratio acetaldehyde
Average Average Average Average Average
20 0.00 41.8 33.3 0.6 2.5 79.8
45.9 43.8 37A 35A 0.8 0.7 2,1 2.3 81.5 80.6
406
F.,!~,,",O-
Feed Sweep to Total ethanol Ethanol conversion Ethanol Ethanol to acetic Reaction selec-
rate feed molar conversion to acetaldehyde conversion to MEK acid ethyl ester tivity towards
mllh ratio acetaldehyde
Average Average Average Average Average
30 0.00 34.3 31.9 0.5 0.7 92.9
36.2 34.2 0.6 0.6 94.6
31.5 34.0 30.0 32.0 0.4 0.5 0.5 0.6 95.2 94.2
3:1 -
46.4 46.4 42.7 42.7 1.4 1.4 0.8 0.8 91.9 91.9
407
i';'<t'0-~-.-
Feed Sweep to Total ethanol Ethanol conversion Ethanol Ethanol to acetic Reaction selec-
rate feed molar conversion to acetaldehyde conversion to MEK acid ethyl ester tivity towards
mllh ratio acetaldehyde
Average Average Average Average Average
10 no sweep 59.4 50.4 2.9 2.3 84.8
58.2 50.3 2.2 2.1 86.4
I 54.5 51.2 1.4 0.3 94.1
53.7 56.5 50.0 50.5 1.7 2.1 0.4 1.3 93.2 89.6
408
Membrane name: N8x Catalyst: 3.00 g (14.4 wt% Cu on silica)
Reaction temperature: 300°C Catalyst particle size: 500-800 microns
Feed Sweep to Total ethanol Ethanol conversion Ethanol Ethanol to acetic Reaction selec-
,
rate feed molar conversion to acetaldehyde conversion to MEK acid ethyl ester tivity towards
mllh ratio acetaldehyde
Average Average Average Average Average
20 0.00 47.8 45.1 1.2 0.3 94.3
47.8 45.8 1.3 0.2 95.8
46.2 47.2 44.7 45.2 1.2 1.2 0.2 0.2 96.9 95.7
409
APPENDIXF2
410
$
411
Membrane name: N8x
Catalyst: 3.50 g (14.4 wt% Cu on silica)
Reaction temperature: 190 DC
Catalyst particle size: 500-800 microns
412
,.
,
Membrane name: N8x
Catalyst: 3.50 g (14.4 wt% Cu on silica)
Reaction temperature: 190°C
Catalyst particle size: 500-800 microns
413
Membrane name: N8x
Catalyst: 3.50 g (14.4 wt% Cu on silica)
Reaction temperature: 215°C
Catalyst particle size: 500-800 microns
414
Membrane name: N8x
Catalyst: 3.50 g (14.4 wt% Cu on silica)
Reaction temperature: 215°C
Catalyst particle size: 500-800 microns
415
Membrane name: NSx
Catalyst: 3.50 g (14.4 wt% Cu on silica)
Reaction temperature: 215 DC
Catalyst particle size: 500-S00 microns
416
Membrane name: N8x
Catalyst: 3.50 g (14.4 wt% Cll on silica)
Reaction temperature: 240 DC
Catalyst particle size: 500-800 microns
417
Membrane name: N8x
Catalyst: 3.S0 g (14.4 wt% Cu on silica)
Reaction temperature: 240°C
Catalyst particle size: SOO-800 microns
418
Membrane name: N8x
Catalyst: 3.50 g (14.4 wt% Cu on silica)
Reaction temperature: 240°C
Catalyst particle size: 500-800 microns
419
APPENDIXG
420
The following conditions were studied:
Note:
Each of these ten investigations consists of six different figures. For the first three figures, the
conditions at the membrane tube exit are used. For the last three figures, data is plotted along
the axis of the membrane.
The ethanol conversion in the first figure on each page is defined as follows:
X = Fet(O),t -F-F
et,t ct,!
(G.1)
ethanol F
et(O),t
421
Standard conditions for ethanol dehydrogenation
Rm[m] 0.0035
Ph [kg/mj] 430
m [kg] 0.0030
Eb 0.4
1] I
422
Shell and tube flow profiles for standard conditions
'"
i
~
-~
4.0E-05
3.0E-05
""'"----- Ethanol
II
...
I;:
/- ---
.!!! 2. DE-05
o Acetaldehyde
Hydrogen
== 1.0E-05 I
NHrogen I
O.DE+OO
o 0.2 0.4 0.6 0.8
Dimensionless length
-
1.0E-05
Hydrogen
_ 8.0E-06
~
o
§. 6.0E-OO
~
~
I;: 4.0E-06
/
So
== 2.DE-OO i =
Ethanol lop
=
Aclltaldehyde bottom
-
O.OE+OO
Il
o 0.2 0.4 0.6 0.8
Dimensionless length
423
1. The etTect of reaction rate k' -values at standard conditions.
Effect of reaction rate on ethanol conversion and Errect of reacUon rate on shell side exit
percentage of ethanol feed lost through membrane composlUon (excluding sweep gas)
--
.."
,,------.--.~-----------.------ 90
,
--- ..
• / ..
e-
7
l """ ,
'-- -
------ .. ~-- -------- - ------ --- ---
-.- Ethanol
·· .. ··AcetaId.~
--Hydrogen
.. J ~
Errect of reacUon rate on tube side exit H. driving force (PrP.) vs. axial position
i"
so·
50
\
, -~-
composlUon
I--Ethanol
I·· ..·· Hydrogen
'> _.. ------------4--------------------------
-'-_'d_~1
-." Nitrogen
~
;;
12000
10000
8000
._--------------
0._
Reaction rate cCMffIcIenW;
O.6k, k, 2k, 4k, '"
for linn from bottom to top
----,----
~
8000
•-
30
10
...: 2000 ~
Ir--·~
o
0
0 1 , 3 4
• 6 7 8 o 0.' 0.4 0.5 0.'
Multiple of standard k-value Dimen.ionle•• length
V
f
'\ 0.'
lL 0.4
0.6
" ----'''----. = ..__.,.-
- -~ f
l Reaction ,.. coeft'lclentl::
o.21k for bottom and 8k for tap line
I
0.3 20
Reaction rate coftflcl.nt.:
0.' 0.261e, 0.11e, k, 211, 4k, SIr;
for lin•• from top to bottom 10
0.1
o o
, 0.' 0.4 0.6 0.8 o 0.2 0.4 0.5 0.9
Dimensionless length Dlmen.~nles. length
424
2. Effect of acetaldehyde adsorption coefficient at standard conditions.
Effect of K.c.t on ethanol conversion and percentage E1'fect of K...t on shell side exit composilion
ethanol feed 100t through membrane (excluding sweep gas)
• -- - ---_._- u 11 - - - - - - - - - - - - - -.-_ _ _ _ _ _ _ _ _ _ _ _ _ 90
---- ~. - ---- --
- 2.3 ., ~
•,
10
--- .
...e
-_ .. -
r'~·-
2.1 li·
-~
!• ~
------ .. •• "l-
-.- --
-- -- -
~ 1.9
1~
i I ii 7
._.a--
--- -.- ElhanCII
--- .. ~ ~
-- -'- ~ ..
i ··.··AcNld~
Ii •
".-
~- ~
Ii
I --+-Converslon (left axis) 1
" ..................... ...............
__ Hydrogen
-.- Lost feMi (right axle)
1.7
5 "
40
o U U U U ~ ..
fAlltlpla of K....
u .. U
1.5 4
0.1 02 0.3 0.' 0.5
Multlpl. rif K.at
.. 0.7 ,. 0.' 1
eo
E1'fect of K...t on tube side exit compos Ilion Hz driving force (PrP.I vs_ axial position
6000 ,-------- ._-
4500
4000
t\
';" 3600
\~ ,----
···l\~ IEffwt of acetaldehyd. adIIorptk'ln:
I!:. 3000 O.UK, 0.1K, UK, D.I5K, G.70K, K -
~ 2500
\\\\ for linn from top to bottom
-
~ 2000 \\~
:£ 1500 \\.'\.
1000
""~
"--~
~
500
' ....
o
~1 U U U M MUM U o 0.2 0.4 0.6 0.'
Multiple of K"" Dimensionles. length
,,.
0.'
I\, 50
0.8
~ 45
0.7 ~ 40
~ 0.6
'~ -----.::: 35
J
"'-----:::::-- 30
' \ 0.5
lL 0.4
~-
.-
",. 20
I EffKt of ac.tlIldehyde adsorption:
O.OI5K, 0.1K, 0.2K, 0.6K, 0.75K, K
0.3 r--- Effect of ac.taldshyde ad.orptlon: I.
0.2 r--- Q,OOK, o.1K, O.2K, o.OK, 0.76K, K
for lin.. fro m bottom to top
ID
I
0.1 5
o o I
o 0.2 0.4 0.6 0.8 o 0.2 0.4 0.6 0.8
Dimenslonle.s length Dlmenstonless length
425
3. Effect of Hz to N z selectivity at standard conditions (Pm.hydrogen is constant).
Effect of membrane selectivity on ethanol conversion Effect of membrane selecUvlty on shell side
and percentage ethanol feed lost through membrane
exit composlUon (excluding sweep gas)
....• .. •
45 - - - - - -_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ .. 14
50 --.--- 100
1!•
! .....6
,
•
10
• 1"
-J
2
l' }"
;30
,
/
../"
--- EthanDi
..• ". Acetaldehyde
~
f--- eo
70
..
" t
1.. . . II/
g /'\
h ~" -+-Hydrogen
ii I M J
4 ,
•, ,
~20
•li15 / . " I
• !III.
"
30
~
'" 44.2
.. --------------
---.
iii 10 --.........
~.,---.
------- ---. 20
.. '"
0 0
100 10>
. ................. 10
0
0 50 100 150
~N, ..Iec:tlvtty "" "" '" 50 200
H,:N2 ·elacttvity "" 300
'"
E1Tect of membrane selecUvlty on tube side Ii, driving force (PrP.1 vs. axial position
exit composlUon
45 - r - - - -, 3600 . , . - - - - - - - - - --
40 3000 A
&30
~,.
35 '.
/,----_.--
---
./' ----
_____ -----------------------a ~
~2000
2500
N
HycIrogan to nItrOgen MttctlVlties:
10. 20,. 40., ID, 1eo, 320 _I
1'-:'--- , .; 1500 (\ for On" from bottom to top
500
~
..•.. Hydrogen
~
5 -." Nitrogen
0 0
0
"" 100
'" 200
H,:N, seloctivily
250
"'" 350 0 0.2 0.'
Dimensionless length
0.' 0.' 1
i 0.6
-...::. -
tt-=------
40
~----" .
---~,,-
lit 0.5
IL 0.4
0.3
'.-
~ I
10, 20, 40, 80, 110, 320 Hydrogen to nitrogen ..1.ct1¥1t...:
0.2
tor Inn from bottom to top 10 10, 20, 40, 10, 110, 120
f
0.1
o o
for Un.. from bottom to top
,
o 0.2 0.4 0.6 0.9 o 0.2 0.4 0.6 0.8
Dimensionless length Dlmansionless length
426
--
Effect of effectivene•• factor on ethanol conversion and Effect of etrectlveness factor on shell side exit
. percentage ethanol feed lost through membrane
" "
composition (excluding sweep gas)
..
~ ,. •
------ ..
•" , ~
m i'20 "
""".,'/ 2
l'
'Ii • L -.- Ethanol
....... AcMaIdehyde
/
--V . . _-----.---___
"
82
i'
•tIi
" ,
2.8 1~ ~ . ---Hydrogen
---a _______ '" r
I
7r
,
Ii
/
-,
- '-, 78
I
__ COnversion (left Dis) 2.3 iti
~
_-
5
-.- Loetfeed(rlghtuls)
1" 78
o ~ u u U U M
err.ctlYene•• factor
U U U
2 0
0 0.1 02 0.3
.........
0.4 .. 0.' 0.7 . O. 1 "
Effect of etrectlveness factor on tube side exit Hz driving force (P,.P.) vs. axial position
composition ,.----
I
.-
70
r__ Ethanol
-.- Ac.u.ldehyde I 3000
60 -I------~
~ T·.·· ......... -e-. Nitrogen r 11\\
ett.ctlvt'n'" hlctor:
8. " ~ 2000
11\ 0.2, 0.4, O.e. 0.8, 1.0
for lin.. frum bottom to top
J!40
c
• --------- ~ IA~
~ 1500
'\..~
30
20 £ 1000
10 500
~
0 o
0 0.2 0.4 0.6 0.8 o 0.2 0.4 0.6 0.'
Effectiveness factor Dimensionless length
i
0.8
0.7
0,6
""". 40
35
--, I
#'
EffKtlYe"_ fador.
~ 0.5
... "
30
I 0.2 and 1.0 r
IL 0.4 l-- Eff.ctlvenen factor. 20
0.3 l-- 0.2, 0.4, 0.1, O.B, 1.0
15
for lin.. from top to bottOm
0.2 10
0.1 5
o o
o 0.2 0.4 0.6 0.8 o 0.2 0.4 0.6 0.6
Dimensionless length Dimensionless length
427
5. Effect of ethanol feed flow rate at standard conditions.
Effect of ethanol feed rate on ethanol conversion and Eflect of eflecllveness factor on shell side exit
perc:entage ethanol feed lost through membrane
eo .. _ _ _ _ _ _ _ _ _._____ • composilion (excluding sweep gas)
--.~--.'------
25 - - - - . - - - - - _ _ _ _ _ _ ____.______ ..
",. ---------- ..
J ~.:~.:;::I.hyd. ~,:.:-..",,~e:...----------- 82 "-
i ~Hyd J7 -------.------___ ._______ ~ t
15 ....n
~1O+_------~L-------------~~~=-~
/
L l---------'------,~==
;;
.. ...... =
= .......'F=
.... =
.... ---4
"
f
1'·"" 'l<=
....
"
~+----~-----~-----~----+
2 2.' ~ u u u u u u u
- ....-
,+------------------------------------+~
u
Eflect of eflectlveness factor on tube side exit H. driving force (PrP.) ys, axial position
composition
"'00 ._---_._._----
ro.-----------~==============-~-
1--- Ethanol -.- AHlaldehyOe. ~
oo+----~-----~
~ I·· .. ·· Hydrogen -." Nitrogen 3000+~~~~~
m f""
5Ot======~~
..... 2500 Mullp" of ethanol rat.:
J.I 0.1, 1, UI, 2,. 2.15 - for linn from
:l. ~ 2000 -II---'lA~"---I bClttcm to tap III r1ght side of graph
J!4O
~
5
~ ~+----------------------=~~----~~~~~
~----~----- --- ~ 1600.JI~--~~~~~======================~--~
e;:
£ 1000 -1-----'._--=>.,,::,...,;"'--.;;;:"---==-________________--1
20+-------~~~~----------------- . ..~
.•~..~~~c{
",,_. on::::' -:;.:.~ ... :::::..:: ~.'~ '_" '.'~ ".:.":
10+_----~~~~~~~~~~--~------~
o~----~----~----~----
o 0.2 0.' 0.6
__----~
0.8 o 0.2 ,.• 0.8 0.8
Effectivenes. factor Dimensionless length
..,,, m
36
i O.S ~
"i. 0.5
LL OA f-
0.' f -
Multi" of ethanol feed rate:
0.5, 1, 1.0, 2, 2.11 ~ for II,," from
bottom to top at right .1_ of graph
.,. l Multiple of IItllanal fMd rate:
0.6, 1, 1.6,2, 2.11- for """ from
bClttam to top ill: left sdgs at graph
428
6. Effect of N2 sweep gas to ethanol molar feed ratio at standard conditions.
Effect of sweep:feed molar ratio on ethanol conversion Effect of sweep 10 feed molar ratio on shell
and percentage ethanol feed lost through membrane
30
side exit composition (excluding sweep gas)
- - . -___ •_ _ _ _ _ _ _ _ •_ _ _ _. _ _ _ _ _ _ _ _ _ "
.. --------.-------.------------ 3
~
55 +----~===~_--------_..,.,
.. ---------
.~~.~-~~~ .... e~ ~~ ,.11'\"-----:/:::::---==~~~=1 ..
j ~ 2 ~ •
~. 20 +-"-.,L-----------...j .. ,. .. Acetmdehyde I---
-.- Eth.nol 85
....
i~
" WVL_____~~=~=.... :E"=='_____.: i
o ____
S!
-!
f-- "U•
80
i",
~,,_
,./
.---______---1 '1
15
"I-
11O~~~~----------------------~
1 - - COnwrwlon (left axlfl) \
t
i .. , 1
-.- ' iII~
m '....... '~~~-- .. ____________ .--______________________
..
Lost feed (rlgtd axle) ';It. 70
" +--------------------+ •. , .. ,...........................
M+--__--_--_----_--_--__--__---+. 'I •
O+---~ __~_____--~--~----~--__---480
, 2 , o 2345678
s-p g •• to.thanol f ..d molar ~o Sw..p gas fa ..... anolfHd molar ratio
Effect of sweep to feed molar ratio on tube H. driving force (PrP.) vs. axial position
side exit composition 5000,-----------------·--------·---
46 - - - - - . - - . - - - - - - - - - - - - - - - - -
~O~~Ar\----------------------------~
~~~~--------~~~--~--~--~--~-~=---~
---------- ~----
.... ~~~~ ..
., 30 ._ 35t:Z'-=:' 1-- Acetald.hyd't Ethanot ---
'j;' @o
e,.
3500 W\ \\\'=:r=:=====~
tf.:-,,:t=
,---i
#>Tri,cn~,\-
Nltrogens....pto.thenolhed-
f:t:S··~;.:=~============~·=··=..=~~~=~==n====-.==.N=.=Mg==.n==~~
lOOO molar nIHil); 0.26,0.5, 1,2, 4, a -
e::~ j=~'\~\"~~==~==~f~
..~H~n~n~f~~~~~.~m~w~~~p~==t-==j
'I- 2500 ~
. 2000 \
~ 15+---~·..~··~ __------------------~ \"-->."::''''~~''''''''''--------__1
.
~ :~!::"'~".:;:.:' .-_._+ '-'-' -" _._.-.-. _. _.-' _.-._._' _._.
:£ 1500 -t\-"....-->"
'~t~\,,~··~"~~~~~~~--~~--~~~~~~~~
10~~~~~~~~~~~~--------_1 ................. .
5+----------------~~~=t
0+-___- __---__--__--__----__--__---4
o 2 3 4 5 e 7 8 o 0.2 0.4 0.• 0.8
Sweep gas to ethanol feed molar ratio Dimensionless length
0.1 10
o 0
,
o 0.2 0.4 0.6 0.' o 0.2 0.4 0,6 0.8
Dimensionless length Dimensionl••• length
429
7. Effect of hydrogen permeance at constant selectivity and standard conditions.
For selectivity to be constant; N 2, ethanol and acetaldehyde permeance have to change with H2 perrneance
Effect of Pm,H2 (con,t. selectivity) on ethanol conversion Effect of Pm,H2 (at constant selectivity) on shell
. ------------_.
and percentage ethanol feed lost through membrane
-- 8
30 I"
side exit composition (excluding sweep gas)
-- 100
#- 44.8
-- -
,.,.-,,-- 7
-.- Ethanol
-- 00
1 ~ -' ,
,
--
i: 44.2
.- -- ---
..i•8 7 - ..- --
43,8
iT ~~~~'
, --
t
0( 43.4 - I --+- Converalon (I'" axis) !
,
.-•
~.- LCIat fHd (r1ght axla)
/-'
o
0.5 1.5 '_5
Multiple ofP• ..., .......
3.5 4 o 1.5 2
Multiple ~ P..".,.....
2.5 3 3.5
•
.
EITect of Pm,H2 (at constant selectivity) on tube H. driving force (PrP.) vs_ axial position
side exit composition 12000 , - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - -
--------------..:.--------~
50
Hydrogln perm•• nee WIIh cormant
.5+.~~~---------------~ 10000 r----.. seteutlvlty: D.2lPm, O.6Pm, 1Pm, 2Pm, 4Pm-
____
.o+--~~~~- - - - - - - - - - - - - i
I
"'" ~
for lin.. from tap to bottom
.
E~i::::~~~~~~~::::~~======;;~::===i
··a _____ _
30
825+----------------"'_-,~==-~~-
'ii'
I!,.
~
8000
8000
~
!£ ~ ~
1 : 1:=:--':-:.--:-':.=:-=. ':.:---~--~-:::--~.-'~-'~,.~-~-~--.~---;---:-:- .:- :- ~-.-;- ~- ~- -;- ~- ~-.-;,- ~- .;,-.~- -
10
5
t /
-I-_~.c::_____-I .. " .. Hydrogtn
o .....
1 --+- Ethanol -.- AcftI1del¥l;r
-.' Nitrogen r
:£ 4000
2000
o "'-
~
~ "'"
~"'( ____
---- - .
--------
o 0.5 1.5 2
Multiple of Pm,hrlll'Ogln
2.5 3 3.5
• 0 0.2 0.'
Dimensionless length
0.8 0_8 1
0.8
~ (77/' .--
i
0.7
0.3
0.2
B Hyd_n ..mo..,•• wHh
Mlectlvity. O.26Pm, O.I5Pm, 1Pm, 2Pm, 4Pm-
fOr Iinetl from top to bottom
""""ont
.- ;1<"
20
15
II
,-
7
/
/"
-I
/"
430
8. Effect of Hz permeance at constant permeances for Nz, ethanol and acetaldehyde.
This implies that the selectivities will increase with an increase in Hz permeance
Etfectof p .... H2 (conal other p ..·.lon ethanol conversion Effect of Pm,H2 (at constant other Pm's) on shell
and percentage ethanol feed lost through membrane side exit composition (excluding sweep gas)
45 r--------..- - - - - - . - - - - - - - - - . - -
".. 44.7 + - - - - - - - - - - - - - - - - - - - - \ 2 . 4
2.5
1m
.,
12,-------------------------------
~ 10~~~--~~------~~-----------------i~
00
~i --.---------- ----------- ~- •
iI i
... ---
~44.4·- 2.3.~
8 ( .iii
] 441 2.2 H 8 ..."'.=. =
.•.=
t{,.,", ....=
..."'•.=...=...=. =...=...=...=...='.•=========t:
...
l
t I -+- COnv.....on (left axle, I ~
... 4t----------------------r=_.=_=a.==~==m====,---~
00( 43.8 _._ Lo.,ueod (right axil;, f---+2.1 ~ 2 ~---------__I ..•.. AcftaIdehyde _ 82
Multiple of p........
Effect of Pm,H2 (at constant other Pm's) on tube H. driving force (PrP.) vs, axial position
side exit composition 12000
45 . - - - - - - - - - - - - - . - - - - - Hydrogen permeanee WIth conatant eU.nol,
acetaldehyde WId nitrogen permeIUIee:
.
40 10000
t: .-.--..------.-------------.--------------------------
35
~
-1--------11............. J 8000
~20 - ... NHl'OSIen
61 ........................... ...................................... __ •••••••••••••••••••• _
:. 1. £ 4000
10
2000
• -."-~--
0 0
0 D•• 1.' 2 2.' 3 3.' 4 0 0.2 0.4 0.8 0.8
431
9. Effect of hydrogen permeance at varying selectivity and non-standard conditions.
Four times standard sweep gas flow rate and four times standard k-value
Effed of P"",M2 (varying selactivity)on ethanol conver- Effect of P",H2 (varying selectivity) on shell
sion and lost feed through membrane (4k, 4 x sweep)
------ .. ,_._--------
side exit composillon (excl. N2 ; 4k, 4 x sweep)
-;==::===;--,"
. . •\
05 ~-.---
1.8
-.- Ethanol
~83
V",
L .
#.2t'·----------4
}
'
\
.. " .. Ac.tllldehyde
-+- Hydrogen
1---+"
-3 ,. l-I-\".=!=:::!:::===========l03 ~
: 62
c
~0 --- Y'
e
i
i Iii
61
t~.. ~
c
•
tOO
.
.t !l9
I . . . . . . COnv....lon (left axis)
/ -.- ~Ifeed(rlghtax")
~
ii 3.3
3.8 ...
'-•
It.• ....... ' It···
__ _____ __
i2
91
57
0 0.5 1.5 2.5 3.5 , o 0.' 1.5 2 2.5 3
•
Multiple of P......-- Multiple ~ P""IIYdN9...
Effect of P",H2 (at varying selectlvlty) on tube H. driving force (P.-P.) vs. axial position
side exit composition (4 x k, 4 X sweep) ~oo --------;=~~~~-=========il
00,---- Hydrogen perme.nee wtth verytng ••Ieo-
tlv1ty (rate • 4k and 4 tim .. SWMP rats):
~t~~-=---~--~-~---~~========~1
o.2IPm, o.8Pm, 1Pm, 2Pm,04Pm
•
".-- .1 ........... Ethanol
for tines from top to battom
Ii'
& ".)-=--------I ..•.. Hydl'f;Jgln
!
-.- Acetaldehyde/
L-~~~_~~~
-"'Nltrogen r ~ '~IJO HP"-----"''''''-----------4
~30~~~==~======~==============~ ~
~ 10000 tf:-"'--",,------"""---:::--------t
~wt--------------~ £
o~--_-- __----_--_--_--_----_--~
o 0.5 1.5 2 2,5 3 3.5 4 o 0.2 0.4 0.6 0.'
Multiple of PIIl,hylhgen Dimensionless length
----------
O,2IPm, O.ePm,1Pm, 2Pm, 4Pm
0.7 '\. for 11M. fram top to bottom - 70 //
......... 00 '// ~
; 0.6
~ '/ / ----~
~-----
' \ 0.'
#" /
LL 0.4
0.3
0.2
30
20
! ..
/"
/ ...---
Hydrogen permeanee wtth varylnll_.I.e-
tlvtty (..... 411 and" tim" sweep me):
o.2t5Pm, o.8Pm, 1Pm, 2Pm, o4Pm
0.' 10 / for IInH from bottom to top on ten: edge
o 0
o 0.2 0.4 0.6 0.' 0 0.2 0_4 0.6 0.'
Dimensionless length Dimensionless length
432
10. Effect of hydrogen permeance at constant selectivity and non-standard conditions.
Four times standard sweep gas flow rate and four times standard k-value
Effect of PIII,M2 (consl selectivity-)on ethanol conversion E1'Iect Or Pm,H2 (at const selectivity) on shell
and lost feed through membrane (4 x k, 4 x sweep) side exit composition (exel. N.; 4k, 4 x sweep)
" -.-.-.-----.---.----------,,-~
--100
"
o 1.5 2.5
MulUpi. rif p ....h.,.......
3
•
E1'Iect Or Pm,H' (at constant selectivity) on tube H2 driving force (PrP.l vs, axial position
side exit composition (4 x k, 4 x sweep) 2~00r-.------r_=-~~-=-==_=_=============~
60
Acatald.hyd~
Hydrogln perm•• nc. with COMtIInt ....c-
I--+-EthanOI -.- UYIy (rata. 4k and .. times swap rate):
50 • ,--.",--. ...... Hydrogen - ... Nitrogen 20000 +---,.--.~-----! O.2fPm, O.6Pm, 1Pm, 2Pm,"Pm
for linn from top to bottom
------------- ';'
~ lrooo+4~~--~~~--------~
-------
0.8
"
G.2I5Pm, o."Pm,1Pm, 2Pm,"Pm
0.7 for liMa from top to bctI:om f--I 70
~ Ill/'
i 0,6 ..
80
17 / .' ~
"i. 0.' '
~.",-: . " ... ~
40 / ,/
,,
lL 0.4
0.3
-----:.•. .. -.. ". ." 30
1/ .' /,/
H)'dfDPn penn8anctl with con.tant ....c-
20
. /' tlvlty ,fIlM -
o4k and 4 tim.. - p rabJ;
0.2 D.21SPm, O.ISPm, 1Pm, 2Pm. o4Pm
0.1 10 / for liMa from bottOm totop on left eda-
V/
o 0
o 0.2 0.. 0.6 0.8 0 0.2 0.4 0.6 0.8
433