Yu Zhengnan 201406 MSC PDF
Yu Zhengnan 201406 MSC PDF
Yu Zhengnan 201406 MSC PDF
by
Zhengnan Yu
Master of Science
in
Chemical Engineering
production and is a highly energy intensive process. The increasing demand for
ethylene has stimulated substantial research into the development of new process
Lower temperatures can thus be used to generate comparable ethylene yields. Using
area to reactor volume ratio of 0.16 m-1 improved ethylene yield by 15.6%
catalyst for ethane dehydrogenation was also developed. With this catalyst, unlike
the Pt/Al2O3, side reactions could be completely suppressed and the ethylene
i
Acknowledgements
directions, guidance, and support. Without his help and supports, this thesis would
I appreciate all of the help I have received from Dr. Kuznicki’s group.
Specifically, I wish to thank Dr. James Sawada for his guidance and support of my
research as well as his help in the revision of my thesis. I also wish to thank Dr.
Adolfo Avila for valuable discussions, assistance with data analysis, and help in
for her insights, suggestions, and varieties of help in the lab. I also wish to thank
Lan Wu and Tong Qiu for their support of my research efforts in the lab. I would
like to extend my appreciation to Albana Zeko and Dr. Tetyana Kuznicki for their
Jing Shen, PhD candidate in Dr. Natalia Semagina’s group, conducted the
TPR tests for me and helped guide the XPS data analysis. Dr. Lianhui Ding
graciously offered to help revise my thesis and provide valuable suggestions about
assistance.
I also thank Lily Laser, Graduate Administrative Assistant, for her kind
ii
Table of Contents
Table of Contents
Chapter 1. Introduction........................................................................................ 1
2.1.2 Pore Size Distribution, Specific Surface Area, and Pt Particle Dispersion
.............................................................................................................. 24
iii
Table of Contents
iv
Table of Contents
v
Table of Contents
Dehydrogenation .................................................................................. 95
Dehydrogenation .................................................................................. 97
Dehydrogenation .................................................................................. 98
vi
Table of Contents
List of Tables
Table 4-1 Dimensionless numbers defined for the comparison between reactor and
Table 4-2 Experimental values of the dimensionless numbers defined in Table 4-1
Table 4-3 Ethylene yield enhancements achieved with the membrane reactor at two
Table 5-1 A product yield comparison using different feedstocks [13] .................. 73
vii
Table of Contents
List of Figures
.............................................................................................................. 7
Figure 1-7 Global Ethylene Cash Cost Index: 2010 [12] .......................................... 9
viii
Table of Contents
Figure 3-9 Refined natural mordenite structure viewed along [001] at different
temperatures: (a) 25 °C, (b) 200 °C, (c) 450 °C, (d) 830 °C .............. 41
Figure 3-10 BET surface area of natural mordenite treated at different temperatures
............................................................................................................ 42
............................................................................................................ 44
Figure 3-12 Pore size distribution of natural mordenite calculated by BJH method
............................................................................................................ 46
Figure 4-5 Dynamics of molar fractions of H2, C2H4, and C2H6 in the reactor outlet
Figure 4-6 Dynamics of outlet molar fractions of C2H6, H2, and C2H4 between
membrane reactor and reactor mode at 500 °C, WHSV = 0.4 h-1...... 62
Figure 4-7 Dynamics of outlet molar fractions of H2 and C2H4 between membrane
reactor and reactor mode at 500 °C. , WHSV = 0.4 h-1 ..................... 63
ix
Table of Contents
Figure 4-8 Dynamics of outlet molar fractions of C2H6, H2, and C2H4 between
reactor and membrane reactor mode at 500 °C, WHSV = 0.4 h-1 with
Figure 4-9 Dynamics of outlet molar fractions of C2H6, H2, and C2H4 between
reactor and membrane reactor mode at 550 °C, WHSV = 0.4 h-1...... 68
Figure 5-1 The scheme of packed-bed reactor for ethane dehydrogenation ......... 81
Figure 5-2 Pt mean particle size and dispersion change with Zn content ............. 83
Figure 5-4 XRD patterns of ETS-2 dried at 150 °C and calcined at 600 °C for four
hours ................................................................................................... 85
Figure 5-5 XRD patterns of: (a) ETS-2 dried at 150 °C , (b) ETS-2 treated at 600 °C,
Figure 5-6 Industrial ETS-2 (a) dried at 150°for 4 hours; (b) calcined at 600 °C for
4 hours ................................................................................................ 87
Figure 5-7 Catalyst 5%Zn, 1%Pt/ETS-2 (a) calcined at 500 °C for 4 hours; (b) tested
at 500 °C............................................................................................. 87
Figure 5-8 Catalyst 5%Zn, 1%Pt/ETS-2 (a) calcined at 600 °C for 4 hours; (b) tested
at 650 °C............................................................................................. 88
x
Table of Contents
Figure 5-10 TEM images of the catalyst 1%Pt, 5%Zn/ETS-2: (a) calcined at 500 °C,
(b) spent (run at 500 °C), (c) calcined at 600 °C, (d) spent (run at 600 °C)
............................................................................................................ 89
Figure 5-11 XPS of the Zn 2p3/2 of catalyst 5%Zn, 1%Pt/ETS-2 calcined at 600 °C
............................................................................................................ 91
Figure 5-12 XPS of the Zn 2p3/2 of catalyst 5%Zn, 1%Pt/ETS-2 calcined at 500 °C
............................................................................................................ 92
Figure 5-14 TG-MS analysis of catalyst 5%Zn, 1%Pt/ETS-2 tested at 650 °C ... 94
Figure 5-15 TG-MS analysis of catalyst 1%Zn, 1%Pt/Al2O3 tested at 500 °C .... 95
Figure 5-16 Product spectrum obtained with the 1%Pd/ETS-2 and 1%Pt/ETS-2
Figure 5-17 Product yield change with space velocity at 509 °C over catalysts
Figure 5-18 Product spectrum obtained with the 1%Pt/ETS-2, 1%Pt/Al2O3, and
............................................................................................................ 98
Figure 5-19 Product concentration change with WHSV at 509 °C over catalysts
Figure 5-20 Product concentration changes with time over 1%Pt/ETS-2 at 509 °C,
xi
Table of Contents
Figure 5-21 Product concentration changes with time over 0.5%Zn, 1%Pt/ETS-2 at
Figure 5-22 Product concentration changes with time over 1%Zn, 1%Pt/ETS-2 at
Figure 5-23 Product concentration changes with time over 1%Zn, 1%Pt/ETS-2 at
Figure 5-24 Product yield changes with time over 2%Zn, 1%Pt/ETS-2 at 509 °C,
Figure 5-25 Product yield changes with time over 2%Zn, 1%Pt/ETS-2 at 509 °C,
Figure 5-26 Product yield changes with time over 5%Zn, 1%Pt/ETS-2 at 509 °C,
Figure 5-27 Product yield changes with time over 7.5%Zn, 1%Pt/ETS-2 at 509 °C,
Figure 5-29 C2H4, H2, and CH4 yield change with space velocity on different
Figure 5-31 C2H4, H2, and CH4 yields change with WHSV at 509 °C over
xii
Table of Contents
Figure 5-32 C2H4, H2, and CH4 yields change with space velocities at 509 °C over
Figure 5-33 Product spectrum obtained with the 1%Pd, 1%Sn/Al2O3 and 1%Pd,
xiii
Table of Contents
Abbreviations
Symbol Description
BET Brunauer-Emmett-Teller
BJH Barret-Joyner-Halenda
EB Ethylbenzene
EO Ethylene oxide
GC Gas chromatography
HK Horvath-Kawazoe
ID Inner diameter, mm
OD Outer diameter, mm
PE Polyethylene
xiv
Table of Contents
Symbols
xv
Table of Contents
𝛼𝐴 Polarizability of adsorbate
𝛼𝑠 Polarizability of adsorbent
𝜀 Porosity
𝜆 Mean free path 𝑚
𝜏 Tortuosity
𝜒𝐴 Magnetic susceptibility of adsorbate
𝜒𝑠 Magnetic susceptibility of adsorbent
xvi
Chapter 1. Introduction
Chapter 1. Introduction
modern society. As Figure 1-1 shows, a wide series of daily items such as plastic
bags, milk and food cartons, washing-up liquids, paints, antifreeze, toys, pipes,
window frames, car components, and many other industrial materials are produced
from ethylene. Based on Chemical Market Associates Inc. (CMAI) data (Figure
1-2), the majority of global ethylene is for the production of polyethylene (PE),
ethylbenzene (EB)/styrene are also important ethylene derivatives. The figure also
According to Oil & Gas Journal’s survey, global ethylene capacity has been
increasing since 1995 or earlier [1], as shown in Figure 1-3. Asia-Pacific and Middle
East were the two fastest increasing regions. The ethylene capacity in North
America and West Europe has been almost unchanged in recent years. Survey data
[2]
revealed that as of January 1, 2013 the global ethylene production was more than
143 million tons versus 141 million tons in 2012. Formosa Petrochemical Corp. had
the largest ethylene complex in Mailiao, Taiwan with the capacity of 2.94 million
tons. Nova Chemicals Corp. had the second largest complex in Joffre, Alberta with
1
Chapter 1. Introduction
the capacity of 2.81 million tons. Complex capacity varied yearly, but the rank did
packaging film,
LDPE,
toys, diapers,
LLDPE
housewears
polymerization
bottles, food
HDPE containers, drums,
crates, housewears
ethylene-
co-polymerization propylene rubbers, paint
rubber
surfactant
ethylene ethylene
oxide glycol
oxidation
ethanol-
Natural gas amines
Ethylene
/Crude oil
acetic fibers,
acet-
polyster resin,
aldehyde acid misc.
pipes,
ethylene vinyl
chlorination PVC window
dichloride chloride
frames
polystyrene, styrene
ethyl-
alkylation styrene acrylonitrile, styrene
benzene butadiene rubber
hydration ethanol
linear
detergents
alcohols
adhesives, coatings,
vinyl acetate textile finishing,
flooring
Saudi Basic Industries Corp. (SABIC) replaced Dow Chemical Co. in 2010 to
become the largest ethylene producer with the total capacity of 23.67 million tons
2
Chapter 1. Introduction
per year. Dow Chemical Co. became the second largest one worldwide with the
total capacity of 23.57 million tons per year. The capacity rank of the two
companies did not change since 2010. As of January 2013, the United States was
the largest producer, with the capacity of 28.12 million tons. China and Saudi
Arabia were the second and third ones, with the capacity of 13.78 and 13.16 million
tons respectively. Canada was the seventh with the capacity of 5.53 million tons.
160
Ethylene demand, million tons
140
120 14.3 14.3
15.1 15.1
13.6 21.1 22.6
100 13.6 13.6 21.1
14.3 11.3 18.1 19.6
80 14.3 13.6 16.6 17.4
14.3 14.3 14.3 16.6
60
86.8 92.1 95.8
40 68.7 74.0 75.5 79.2 82.3
64.2 64.2 67.2
20
0
2006 2007 2008 2009 2010 2011 2012 2013 2014 2015 2016
Year
As shown in Figure 1-3, global ethylene capacity reached 147.7 million tons
in 2011, and was predicted at over 170 million tons in 2016 at 3.4% annual growth
[4]
. Relying on its huge market advantage and rapid growth of demand, Asia-Pacific
becomes a region attractive to investors for ethylene production. Even though the
newly added production capacity was small, the advantage of low production costs
3
Chapter 1. Introduction
could make, Middle East the world’s most investment-intensive region in the
4
Chapter 1. Introduction
160
21.2 23.5
19.4
Production capacity, million tons
0
1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011 2012 2013 2014 2015
Year
5
Chapter 1. Introduction
Natural gas (ethane) and petroleum distillates (naphtha) are the main
feedstocks for ethylene production. The trend of their ratios over the years are
illustrated in Figure 1-4. The figure indicates naphtha is the major feedstock for
increased. The cost of feedstock accounts for 60% - 80% of ethylene production
costs. The main factors driving feedstock changes for petrochemical plants are the
200
180
Feedstock, million tons
160
140
79.0
120
67.7 69.7
100 62.1 64.1 65.1
61.0 61.5 60.5
80 60.5 55.4 55.9 13.8
60 11.3 11.3 11.8
10.8 10.8 11.3 10.8
40 8.7 9.7
8.2 8.2 71.3
45.1 47.2 49.7 52.8 55.4 57.9
20 33.8 33.3 36.4 40.0 44.1
0
2007 2008 2009 2010 2011 2012 2013 2014 2015 2016 2017 2022
Year
In natural gas-rich regions, like North America, ethane was the main feedstock
for ethylene production in recent years [4, 7, 8]. As showed in Figure 1-5, ethane was
the main feedstock for ethylene production in the United States from 2005 to 2011.
6
Chapter 1. Introduction
Regional cash cost and global cash cost of ethylene production are illustrated in
Figure 1-6 and Figure 1-7 respectively. As revealed from the two figures, ethane as
ethylene feedstock was more profitable than LPG and naphtha. The higher margins
capacity to take advantage of the low-cost and increased availability of ethane from
natural gas. A vast number of chemicals and products are made from ethylene. It is
the most produced organic chemical worldwide. In 2012, global ethylene sales
100
90 16.1 17.0 14.7
28.7 28.0 24.1
32.6 3.7 3.2 3.2
80
Ethylene Feedstock, wt%
7
Chapter 1. Introduction
Ethane is the most competitive feedstock for ethylene production due to its
cost and the highest ethylene yield among all other feedstocks. Generally, ethylene
yields decrease with the increase in the density of feedstocks. Heavier feedstocks
produce more low-value by-products. To produce the same amount of ethylene, the
use of AGO feed was about three times that of ethane. Furthermore, the utilities
1200
1006
951 967
1000
883
799 824 830
Cash Cost, USD/ton
800
600
400 318
200 103
0
MDE NAM NEA NAM MDE MDE NEA SEA WEP
Ethane Ethane LPG Naphtha LPG Naphtha Naphtha Naphtha Naphtha
8
Chapter 1. Introduction
350
313 319
303 307
300 277 285
Index, Canada Cost = 100
250 232
200
150
100
100
50 28
0
MDE Canada NAM MDE MDE WEP NEA NAM SEA
Ethane Ethane Ethane Naphtha LPG Naphtha Naphtha Naphtha Naphtha
MDE: Middle East, NAM: North America, NEA: Northeast Asia, SEA: Southeast
Asia, WEP: Western Europe.
Figure 1-7 Global Ethylene Cash Cost Index: 2010 [12]
Driven by the pursuit of lower investment cost the tube furnace pyrolysis
technology has continually improved over the course of several decades. Currently,
About 99% of global ethylene production uses tube furnace pyrolysis method. A
9
Chapter 1. Introduction
Convection
Feedstock Zone
BFW
Steam
Superheated
Steam
~850°C Steam
Drum
< 0.1s
Radiation
Zone
350~600°C
To Quench
Steam Cracker Transfer-Line
Exchanger
dilution steam in the convection zone and then quickly discharged to the radiation
zone (750 to 875 °C) where the feed is cracked to produce ethylene, other small
olefins, di-olefins, and by-products. The method is also known as steam cracking
process. The residence time is 0.1 – 0.5 seconds. Due to the high reactivity of the
products, the high temperature effluent has to be quenched within 0.02 – 0.1
seconds in the transfer-line exchanger to avoid further unwanted side reactions. The
with olefin yields of around 50% [13]. Single-pass conversion and yields are lower
in naphtha crackers.
10
Chapter 1. Introduction
ethylene and feed hydrocarbons are very reactive at high temperature, which means
means that lower product partial pressure is favored for the higher conversion. In
the steam cracking process, steam is used as a diluent to lower reaction partial
suppress coking in radiant coils. The design constraints for increasing ethylene
product distribution are residence time, the partial pressures of the feedstock and
The main six technology licensors for ethylene production and their main
11
Chapter 1. Introduction
1.4.1 Coking
Coking is one of the challenges to radiant coils and quench exchangers. There
[14]
are two types of coking, tube metal catalyzed coking and pyrolysis coking .
Carburized metals are problematic as they can boost the catalyzed coking process.
Catalytically formed coke is about 80% of the total coke when using gas feedstocks,
like ethane, propane, and butane. For liquid feedstock, catalytically formed coke is
about 30% to 40% of the total coke. Pyrolysis coking is the lesser coke product and
12
Chapter 1. Introduction
reactions.
Coke layer could reach 10 mm or thicker depending on the type of feedstock and
operation severity. Coking in furnace tubes can reduce product yields, decrease
production capacity, increase energy consumption, shorten tube service life, and
These coating agents can be applied on-line to make a stable layer coated on the
inner surface of the tubes in the furnace. The coated layer can reduce coking
effectively, protect the tube metal base from oxidation, and improve tube life and
/Kubota’s ANK 400 coating agents can be applied on the inner surfaces of new
furnace tubes and transfer line exchangers. Coking inhibitors can also reduce
coking. The agents are the chemicals containing S, P, Sn, Sb, and Cr among others.
and chromium phosphate. Both coating agents and inhibitors are effective in
reducing coking and extending operation cycles. However, these technologies only
13
Chapter 1. Introduction
Steam cracking for light olefins such as ethylene and propylene production is
the most energy consuming process in the chemical industry. During ethylene
steam cracker consumes about 65% of the total process energy and approximately
[15]
75% of the total energy loss . It is a challenge to increase ethylene production
higher temperatures also accelerate the relative rate of the primary reaction
coking because these secondary reactions are more thermodynamic favorable. This
tube and furnace temperatures are required, which increases the energy
industrial process for ethylene manufacture, it has its disadvantages. First, steam
cracking is a highly energy intensive process. When using ethane as feedstock, the
14
Chapter 1. Introduction
[13]
total energy consumption is 16 GJ/ton versus 23 GJ/ton when using naphtha .
Second, fired furnace produces CO2 and NOx emissions. approximately 1 – 1.6 tons
of CO2 are produced per ton of ethylene. In addition, coking and high temperature
be selectively removed from the reaction system, it can break the thermodynamic
barrier and will increase the conversion even at lower temperature and achieve more
Membrane Reactor
According to the IUPAC classification of pore diameter [16], there are three general
larger than 50 nm, mesoporous membranes with average pore diameter in the range
between 2 nm and 50 nm, and microporous membranes with average pore diameter
less than 2 nm. The micropores can be further classified into subgroups by the
diameter: supermicropores (1.4 to 3.2 nm), micropores (0.5 to 1.4 nm), and
15
Chapter 1. Introduction
Pore size and distribution, openness (volume fraction) and the nature of the
thus generally leading to higher permeation rates for the same pore size. The nature
of the pores determines the mechanisms of gas stream transport through a porous
simultaneously.
From the table, it can be seen that viscous flow, molecular diffusion, and
Knudsen diffusion occur in macro- and meso-pores, but viscous flow and molecular
diffusion do not separate species. However they can affect the total flow resistance
capillary condensation, and molecular sieving are responsible for gas separations.
16
Chapter 1. Introduction
between molecules and the pore surface. When the mean free path length of the
molecules is larger than the characteristic pore diameter, the permeating species
collide with the pore surfaces more frequently than with other molecules. As shown
in equation (1-1), smaller molecules have longer mean free path length because the
𝑘𝑇
𝜆= 2𝑃 (1-1)
√2𝜋𝑑𝑝
depends on the square root of the molecular weight ratio of the gases separated,
where F and M represent the permeation fluxes and molecular weights of the gas
species A and B.
𝐹𝐴 𝑀𝐵
𝛼∗ = =√ (1-2)
𝐹𝐵 𝑀𝐴
2𝜆
𝐾𝑛 = (1-3)
𝑑𝑝
gas diffusion (also refer to Table 1-2): (1) for viscous flow, 𝐾𝑛 ≪ 1 ; (2) for
17
Chapter 1. Introduction
Knudsen diffusion, 𝐾𝑛 ≫ 1; and (3) for transition flow, 0.01 ≤ 𝐾𝑛 ≤ 10. Viscous
and transition flow are non-selective. Generally speaking, when is ten times
[18]
larger than 𝑑𝑝 , Knudsen diffusion becomes dominant . In the Knudsen flow
regime, the mean free path of gases is comparable to the pore diameter.
As equation (1-2) implies, the separation is only effective for those gases with
this regime is usually low, Knudsen flow is not effective in gas separation. As
[21]
equation (1-4) shows , Knudsen permeance is not a function of pressure. It is
proportional to 1⁄√𝑇.
2𝜀𝛾 8
𝐹𝐾 = √ (1-4)
3𝜏𝐿 𝜋𝑅𝑇𝑀
when gas mixtures interact with the inner wall materials. Chemical properties of
size and the permeating molecules which limits the selectivity of the membrane.
18
Chapter 1. Introduction
separation membrane system: (1) the permselectivity or selectivity toward the gases
to be separated. This property affects the recovery of the valuable gas in the feed
and, for the most part, directs the process economics. (2) the permeate flux or
required. (3) the membrane stability or service life which has a strong impact on
zeolite based membranes overcome the limitations associated with organic and
metal membranes. Zeolites have a highly uniform pore structure, good thermal
stability, and strong mechanical properties. Zeolitic materials have been used for
inorganic membranes since 1930s. Zeolites that have been extensively studied for
sodalite [23]. However, gas separation and reactive separation applications of zeolite
rates of mass transfer in membranes. For gas separation, the transport occurs by
permeation rates of the components through the membrane. Gas diffusion obeys
Fick’s Law. The driving force of gas diffusion in a porous membrane is the
The catalytic function can be introduced in one of two ways: inert zeolite
catalytic membrane, the zeolite membrane is not catalytically active and does not
participate in the reaction, and simply acts as a selective separation unit for the
desired products. A separate packed bed of catalyst particles promotes the chemical
reaction on the feed side and the membrane selectively removes one of the products
functionalized with a precious metal, and the membrane effectively performs the
catalytic reaction and separation simultaneously because the catalysis takes place
Based on the configurations of the zeolite membranes and the catalysts, there
[25]
are three types of zeolite catalytic membrane reactor modules , as shown in
20
Chapter 1. Introduction
equilibrium (cases (a) and (b) in Figure 1-10). In the examples shown in case (a),
was enhanced by the removal of product hydrogen from the feed side. This reactor
As shown in case (c) in Figure 1-10, the reactant distributor type membrane
reactor works like an extractor type membrane reactor inversely. A typical example
alkanes like ethane [28-33] and propane [34, 35] to make olefins were carried out in this
the reaction system. There are many benefits to using a distributor membrane for
oxidation reactions, such as achieving higher selectivity, more even reaction heat
catalyst and separator at the same time. In this configuration the reactants are
located at different sides of the membrane and must diffuse through the zeolite layer
to react.
21
Chapter 1. Introduction
22
Chapter 1. Introduction
The objective of this work was to demonstrate the concept of using natural
work also sought to demonstrate the benefits of using ETS-2 as a catalyst substrate
23
Chapter 2. Characterizations and Product Analysis
Testing conditions were: 38 kV, 38 mA, scan speed 2 degree/min, step width
0.02 degree, and scan range 5 to 90 degree. XRD samples data were compared to
the database from the International Center for Diffraction Data-Powder Diffraction
For natural zeolites, both powder and thin surface samples were tested. For
catalysts and supports, all samples were fine powder. All XRD data were collected
at room temperature.
Dispersion
distribution (PSD) of the natural zeolite modernite were calculated from the N2
The specific surface area of samples was calculated using the Brunauer-Emmett-
Teller (BET) method. For the natural zeolite mordenite, the sample dried at 120 °C
was degassed at 150 °C for two hours. Other samples were degased at 200 °C for
in hydrogen at 400 °C for 30 minutes, cooled down under vacuum to 35 °C, and
then evacuated for an additional 10 minutes at this temperature before the hydrogen
isotherm was collected at the same temperature. A total of 14 points were collected
for the hydrogen isotherm between 0.25 and 0.60 kPa. Dispersions were estimated
The XPS measurements were carried out on a Kratos Axis 165 X-ray
at 15 mA and 14 kV). The survey spectra were collected with analyzer pass energy
of 160 eV and a step of 0.4 eV; the high resolution spectra were scanned with a pass
energy of 20 eV and a step of 0.1 eV. During the acquisition of a spectrum, charge
neutralization was applied to compensate the insulating problem of the sample. XPS
signals were fitted by mixed Lorentzian–Gaussian curves using Casa XPS software.
All binding energies are reported after calibration for C1s peak to match 284.8 eV.
25
Chapter 2. Characterizations and Product Analysis
2.1.4 TPR
conductivity detector (TCD) under atmospheric pressure. About 0.5 grams of the
catalyst calcined in air, at 200 °C, for one hour was used for the analysis. The
catalysts were reduced in a flow of 10% H2/Ar (50 ml/min) from room temperature
The SEM used was a Hitachi S-2700 Scanning Electron Microscope (SEM)
equipped with a PGT (Princeton Gamma-Tech) IMIX digital imaging system and
quadrant solid state Backscattered Electron Detector. Samples for EDX were coated
with carbon.
JEOL 6301F field emission scanning electron microscope. Samples for high-
2.1.6 TEM
26
Chapter 2. Characterizations and Product Analysis
microscope operating at 200 kV. Fine powder samples were loaded on copper mesh
Instruments TGA Q500. N2 was used as the inert sweeping gas at a flow rate of 40
ml/min while between 15 to 25 mg of fine powder sample was used for analysis.
The analysis of coke deposits on spent catalysts was conducted using TG-MS.
The TG-MS plots were collected using a TA Instruments Q500 TGA coupled to a
Pfeiffer Omnistar QMA 200 residual gas analyzer. TG scans were run using a
balance purge rate of 5 ml/min Ar and a sample (air) purge rate of 180 ml/min.
Samples were loaded on platinum pans and heated at a rate of 10 °C/min from
The mass spectrometer was configured with a stainless steel capillary heated
to 200 °C and connected to the TGA by means of a stainless steel adapter and a ¼”
branch T-fitting. The capillary was positioned at the center of the tee close to the
furnace exhaust to sample the evolved gases. The exhaust end of the T-fitting was
atmospheric components and led to a fume hood exhaust. Experiments were run in
MID mode with fragments at m/z = 18 and 44 tracked over time. These fragments,
associated with water and carbon dioxide respectively, were found to be the most
diagnostic signals associated with the combustion of the coke. To start data
27
Chapter 2. Characterizations and Product Analysis
few tenths of a degree for the starting temperature can be expected between the TG
and MS plots.
powder sample was pressed into cylindrical rods using a special die and an isostatic
press (Top Industrie) under a pressure of 300 MPa. The rods were about 10 mm in
length and 5 mm in diameter. The pressed rods were heated from room temperature
1000 °C for one hour, the sample was cooled down at a rate of 10 °C/min. The
thermodilatometric data were processed under the form of curves, which report the
analyze gas products. The packed column was Hysep D (80-100 mesh) stainless
28
Part I: The Use of Natural Zeolite Mordenite in a
3.1 Introduction
alkaline earth metals. The name zeolite was created by the Swedish mineralogist
[36]
Axel Fredrick Cronstedt in 1756 and it means “boiling stone” in Greek . More
than 40 naturally occurring zeolites have been identified in the past two centuries.
Chabazite, clinoptilolite, erionite, ferrierite, mordenite, and phillipsite are the six
[37]
most industrially important natural zeolites . Mordenite was first discovered in
[38]
1864 in Morden, King’s County, Nova Scotia, Canada by How and, as per
tradition, was named after the locality. Mordenite mineral deposits occur in more
than 348 locations in 39 countries including Antarctic [39], as shown in Figure 3-1.
New Zealand is one of the countries with the most abundant mordenite
deposits. About 250,000 years ago, massive volcanic ash formed thick sediment
beds in lakes. Hot water transformed those clay beds into soft rocks with ordered
internal structures. Over time the lakes drained due to severe block faulting and the
deeply buried structured clay rocks became zeolites. Mordenite and clinoptilolite
are the predominant natural zeolites in New Zealand. Due to their diagenesis, they
are highly matured and mechanically strong and cannot be broken down into
constituent clays.
All samples of natural zeolite mordenite used in the thesis work are from New
Zealand.
30
Chapter 3. Characterization of Natural Mordenite Zeolite
physically and chemically stable at the required working temperatures and it has
addition, it is very economical due to its abundance in nature. Its pore structure,
and smaller making it ideal for hydrogen separation from larger hydrocarbons.
Zeolite mordenite has a big elliptical 12-ring channel and an 8-ring (2.6 ×
5.7Å) channel along [001] with two interconnected 8-ring channels (3.4 × 4.8Å)
along [010]. Figure 3-2 shows an ideal framework of mordenite. The 8-ring holes
of successive sheets do not align to make channels along [010]. Since the 8-ring
31
Chapter 3. Characterization of Natural Mordenite Zeolite
channel along [001] is smaller than most gases’ kinetic diameter [40] (He: 2.6 Å, H2:
2.8 Å, CO2: 3.3 Å, O2: 3.4Å, N2: 3.6Å), the 8-ring opening acts as a connecting
more visual demonstration is shown in Figure 3-3 [41]. The figure clearly shows that
the 8-ring channels are highly compressed and not accessible to most molecules or
ions, and there is no direct connections between 12-ring channels. Although the
framework structure appears like it would admit larger molecules, the position and
types of the cations inside the framework prevent molecules larger than about 3 Å
32
Chapter 3. Characterization of Natural Mordenite Zeolite
33
Chapter 3. Characterization of Natural Mordenite Zeolite
3.3 XRD
Both the powder and the thin membrane surface samples of natural zeolite
Figure 3-4 shows that the sample of natural zeolite mordenite contained a
quartz impurity. From Figure 3-5, it can be seen that the natural mordenite samples
were thermally stable at elevated temperatures. The XRD patterns show that the
34
Chapter 3. Characterization of Natural Mordenite Zeolite
In Figure 3-6, cross-section images clearly show that there were plenty of
mordenite has a needle-like crystal structure while the layers stacked on each other
indicate the presence of multiple crystals. This multiple-layer structure reveals the
35
Chapter 3. Characterization of Natural Mordenite Zeolite
Figure 3-7 shows the TGA curve of the natural zeolite mordenite. The total
weight loss was about 12.1 wt%. The weight losses at ≤ 150 °C, 150-200 °C, 200-
300 °C, 300-400 °C, 400-670 °C were 6.2 wt% (about 51% of the total weight loss),
1.1 wt% (9% of the total weight loss), 1.3 wt% (10.7% of the total weight loss), 1.1
wt% (9% of the total weight loss), and 2.2 wt% (18.2% of the total weight loss)
respectively. After 660 °C, there was nearly no weight loss (~ 0.2 wt%).
36
Chapter 3. Characterization of Natural Mordenite Zeolite
The weight loss at < 200 °C is caused by dehydration of the zeolite. For the
zeolites with high kinetic diameter, like mordenite, the main dehydration
temperature is lower than 200 °C. The so-called zeolitic water is free water
absorbed in capillary channels and external surface as well. The water dipole is
attracted to the oxygen electron density in the SiO4 and AlO4 tetrahedra. Free water
in large channels is easy to remove because it lacks strong associations, and so the
dehydration temperature is in the low range (~ 100 °C). Water associated with the
remove and is desorbed at higher temperatures (100-200 °C). Once all of the water
associated with the cations is desorbed the remaining weight loss is due to
“chemical” and “structural” water. Hydroxyl groups present in defects and at the
amount of water content and dehydration profile of natural mordenite were in good
agreement with literature reports [42]. The loss of water can cause zeolite membrane
37
Chapter 3. Characterization of Natural Mordenite Zeolite
100
98
96
Weight (%)
94
92
90
88
86
0 100 200 300 400 500 600 700 800 900 1000
Temperature (C)
38
Chapter 3. Characterization of Natural Mordenite Zeolite
20
-20
∆L/L0, ×10-3
-40
-60
-80
-100
-120
0 100 200 300 400 500 600 700 800 900 1000
Temperature (°C)
1000 °C under atmospheric conditions is shown in Figure 3-8. In the region between
room temperature and around 800 °C, the negative slope curve shows shrinkage of
sintering behavior of the sample. The TGA curve in Figure 3-7 shows that from
room temperature up to 700 °C the mordenite sample was continuously losing water,
and the weight loss at 700 °C was about 98% of total weight loss. From the XRD
profile in Figure 3-5, it can be seen the mordenite retained its crystal frameworks
without obvious structural collapse and phase transformation. So it appears that the
continuous dehydration with increased temperature was the main reason for the
shrinkage. In the region between 800 °C and 1000 °C, shrinkage increased
39
Chapter 3. Characterization of Natural Mordenite Zeolite
[43]
significantly and thermal collapse of the structure occurred. Cruciani reported
For the negative thermal expansion behaviors, Sleight proposed four possible
[44]
mechanisms : (a) decrease of M-O bond length leading to a net cell volume
contraction; (b) anisotropic thermal behavior of M-O bonds; (c) extrinsic effect of
For zeolites, the M-O bonds in the frameworks are hinged by two-coordinated
oxygen and are relatively rigid. Their thermal motions have limited freedom in
mode (RUM) model to interpret and quantify the flexibility possessed by zeolites
[45, 46]
, but it was a debated issue [43].
the release of water molecules from the channels. The cell parameters b and c
decreased regularly as the temperature rose. The removal of water molecules also
relocated the initial Ca sites into many positions bonded to the framework oxygen.
The increased interaction with the framework oxygen of Ca sites was intimately
related to the distortion of the 12-ring which is in turn related to the lengthening of
the a cell parameter. Figure 3-9 demonstrates the initial cations relocating and water
40
Chapter 3. Characterization of Natural Mordenite Zeolite
Figure 3-9 Refined natural mordenite structure viewed along [001] at different
temperatures: (a) 25 °C, (b) 200 °C, (c) 450 °C, (d) 830 °C
1 1 𝐶−1 𝑃
𝑃 = + ( ) (3-1)
𝑊( 0 −1) 𝑊𝑚 𝐶 𝑊𝑚 𝐶 𝑃0
𝑃
41
Chapter 3. Characterization of Natural Mordenite Zeolite
𝐸1 −𝐸𝐿
𝐶 = 𝑒𝑥𝑝 ( ) (3-2)
𝑅𝑇
Figure 3-10 shows the BET surface area changes of the mordenite treated at
200 °C for four hours has the largest surface area. As water was removed out from
[47]
the inner channels, the surface area increased. Martucci and coworkers found
that at 200 °C the diameter of the 8-ring along [001] was the largest, decreasing
with further temperature raise. A big ring aperture exposes more internal surfaces
40
35
30
Surface Are, m2/g
25
20
15
10
5
0
0 200 400 600 800 1000
Temperature, °C
temperatures
42
Chapter 3. Characterization of Natural Mordenite Zeolite
shown in Figure 3-11. Many pore size distribution methods are derived from the
method was derived independently from the Kelvin equation, which more
𝑃 𝑁𝑠 𝐴𝑠 +𝑁𝐴 𝐴𝐴 𝜎4 𝜎 10 𝜎4 𝜎 10
𝑅𝑇 𝑙𝑛 ( ) = 𝐾 [ 𝑑 3
− 𝑑 9
− 𝑑 3
+ 𝑑 9
] (3-3)
𝑃0 𝜎 4 (𝑙−𝑑) 3(𝑙− ) 9(𝑙− ) 3( ) 9( )
2 2 2 2
6𝑚𝑐 2 𝛼𝑠 𝛼𝐴
𝐴𝑠 = 𝛼𝑠 𝛼𝐴
(3-4)
( + )
𝜒𝑠 𝜒𝐴
3𝑚𝑐 2 𝛼𝐴 𝜒𝐴
𝐴𝐴 = (3-5)
2
0.858𝑑
𝜎= (3-6)
2
𝑑 = 𝑑𝑠 + 𝑑𝐴 (3-7)
43
Chapter 3. Characterization of Natural Mordenite Zeolite
4 6 8 10 12 14 16 18 20
1.6E-3
1.1E-3 120°C
5.4E-4
0.0
5.1E-2
200°C
3.4E-2
1.7E-2
Dv(w) (cc/Å/g)
0.0
1.8E-3 400°C
1.2E-3
6.0E-4
0.0
6.0E-4 800°C
4.0E-4
2.0E-4
0.0
12.3 1000°C
8.2
4.1
0.0
4 6 8 10 12 14 16 18 20
Pore Width (Å)
method
Figure 3-11 clearly shows the micropore size distribution of the natural
mean pore size increased, and the size distribution became wider. Both pore size
and distribution were well predicted by the removal of water adsorbed in channels
and redistribution of initial cations through water removal, which made N2 more
accessible to the micropores. At 800 °C, a higher number of large pores with a wide
range of sizes are present. These big pores may be caused by the collapse of some
of the micro frameworks. When temperature increased to 1000 °C, micropore size
44
Chapter 3. Characterization of Natural Mordenite Zeolite
surface area and pore size distribution information primarily for mesoporous and
microporous materials. The equations for pore size distribution calculations are
2
𝑟𝑝𝑛
𝑉𝑝𝑛 = ( 𝛥𝑡 ) (𝛥𝑉𝑛 − 𝛥𝑡𝑛 ∑𝑛−1
𝑗=1 𝐴𝑐𝑗 ) (3-8)
𝑟𝐾𝑛 + 𝑛
2
𝑟̅ ̅̅̅−𝑡
𝑟𝑝 𝑟̅
𝑐 = ̅̅̅𝑐 = ̅̅̅
(3-9)
𝑟𝑝 𝑟𝑝
treated at different temperatures. The figure clearly indicates that low temperature
treated samples lack mesopores. However, when treated at 800 °C and 1000 °C, 20
45
Chapter 3. Characterization of Natural Mordenite Zeolite
0 10 20 30 40 50
2E-2
120°C
1E-2
8E-3
4E-3
0
5.4E-2 200°C
3.6E-2
1.8E-2
0.0
Dv(d) (cc/Å/g)
400°C
5.7E-3
3.8E-3
1.9E-3
0.0
1.2E-3 800°C
8.2E-4
4.1E-4
0.0
5.4E-4 1000°C
3.6E-4
1.8E-4
0.0
0 10 20 30 40 50
Pore Diameter (Å)
method
46
Chapter 3. Characterization of Natural Mordenite Zeolite
[50]
falling in the typical range of Si/Al of mordenite . The sample had low
concentration of alkali metals like sodium and potassium, but very high calcium
47
Chapter 3. Characterization of Natural Mordenite Zeolite
distribution measurements show that mordenite has uniform pore size. All these
48
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
Dehydrogenation
4.1 Introduction
packaging and more. The global ethylene capacity reached 150 million tons in 2012,
[4]
and is predicted to have a growth rate of 3.4% per year . The global sales of
Currently, tube furnace steam cracking process is the dominant technology for
ethylene production, which accounts for about 99% of global ethylene production
process. Among the hydrocarbon feedstock for ethylene production, ethane has the
lowest cost and the highest ethylene yield. Heavier feedstock produces more low
value by-products. Steam cracking process for ethylene production occurs at high
(Equation (4-1).
Published: A. M. Avila, Z. Yu, S. Fazli, J. A. Sawada, and S. M. Kuznicki: "Hydrogen-
Selective Natural Mordenite in a Membrane Reactor for Ethane Dehydrogenation", Microporous
and Mesoporous Materials, 2014, 190 (0), 301-308
49
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
(4-1)
The ethane diluted with steam flows through high temperature tubes at high
the high reaction temperature, furnace skin temperature is usually between 1100 °C
to 1200 °C. Such high temperature operation causes serious side reactions in
dehydrogenation
50
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
ethane was diluted with argon and hydrogen. The driving force for hydrogen
separation was Knudsen diffusion. Gobina and coworkers [58-61] studied the packed-
bed membrane reactor. The membrane was a thin layer of Pd-Ag alloy supported
on porous Vycor glass tube. Ethane was diluted with nitrogen. Szegner and
The catalyst was Pt-Sn/Al2O3 pellets. Membrane used in the reactor was inert and
only functioned as a separator. Feed gas was diluted with hydrogen and argon. The
results showed that the conversion of ethane dehydrogenation was higher than that
the reaction system with inert gas could also shift the conversion above the
equilibrium.
removing hydrogen from the reaction system shifting the reaction to the favorable
direction. Hydrogen removal by the selective membrane is based on the fact that
51
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
temperatures the perm-selectivity of hydrogen is just slightly higher than that in the
Knudsen flow, because the Knudsen separation factor is proportional to the square
The unique internal channel system and pore sizes of some zeolitic
has uniform a one-dimensional channel system which has pores that can be small
mordenite as membrane to extract hydrogen from the reaction system. The Pt-
Sn/Al2O3 pellets were used as the catalyst. The PBMR operated in both the
switching the sweeping gas on and off to evaluate the relative effect of the
52
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
4.2 Experimental
The catalyst used for the PBMR was 1wt%Pt-0.3wt%Sn/Al2O3 (Alfa Aesar
product, USA, reduced). The spherical alumina pellet size ranged from 1-2 mm.
The catalyst was dried at 110 °C for four hours before use.
Mordenite membrane disks were prepared from the raw natural monolithic
mordenite fragments from Paradise Quarry Limited, New Zealand. The rock pieces
Trim Saw/Grinder equipped with a MK-303 diamond lapidary saw blade. The
cubes were sliced into ∅19 × 1.5 𝑚𝑚 disks by water jet cutting or by hand. The
disks were polished with a rotating diamond polishing lapidary disk (180 mesh,
Fac-Ette Manufacturing Inc.), washed in ultrasonic bath for 30 minutes, and then
dried at 110 °C for four hours. Finished membrane disks were activated at 750 °C
for four hours before use. Figure 4-2 show the steps of membrane disks preparation
53
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
outer tube was made of quartz (ID 22 mm) and used as the feed side of the
membrane reactor. The permeate side consisted of two concentric ceramic tubes.
The sweep gas flowed through the inner tube and the permeate flux flowed out
through the outer tube (OD 12.6 mm). The catalyst pellets were packed in the feed
side in contact with the membrane and held by the quartz wool. The membrane was
sealed on the end of the ceramic outer tube with high temperature ceramic adhesive
described in the next section. All tube connections were sealed with rubber O-rings
and Swagelok fittings. Outlet flows of both feed side and permeate side were
54
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
Ethane dehydrogenation reactions were carried out at 500 °C and 550 °C. The
450-GC) with automatic sampling valves and TCD detector was used to analyze
both the permeate and the feed side streams. Inlet flow rates (feed and sweep gas)
were controlled by mass flow controllers (Sierra Instruments Inc.). All gas flow
rates (both inlets and outlets) were measured under lab conditions (21 °C,
atmosphere pressure). Inlet flow rates were controlled by mass flow meters. Outlet
flow rates were measured with bubble meters. All gases used were from Praxair
Canada.
Ethane (99.5%) flow rate was 12.0 ml/min. Argon (99.99%) was used as the
sweep gas and its flow rate was 54.0 ml/min. A higher sweep gas flow rate did not
improve the hydrogen removal from the feed side. Feed side pressure was
maintained at 2.7 kPa(g) with a back pressure regulator valve while the permeate
side pressure was atmospheric. Argon was not detected in the retentate stream when
pure helium was used as a feed stream in the membrane reactor at the same
operating conditions. Feed dilution effects were negligible in the membrane reactor
experiments.
55
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
reactor mode by closing off the inlet and outlet of the sweep gas stream. A
schematic demonstration is shown in Figure 4-4. When both inlet and outlet of the
permeate side were shut off, the concentration of permeated species eventually
equalized in both sides. The membrane reactor set-up essentially performed like a
each other.
The reaction rate enhancement achieved in the membrane reactor mode could
be quantified considering the mass balance of species in the reaction zone. The
expressed as:
56
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
where the term on the left-hand side represents the net molar flow rate of
species i through the boundaries of the reaction zone, 𝑟𝑖∗ is the local reaction rate in
𝑚𝑜𝑙/(𝑐𝑚3 ∙ 𝑔𝑐𝑎𝑡 ∙ 𝑠) and is the catalyst weight in the reactor. Integrated both
where 𝐹𝑖 is the molar flow rate of each reacting species and 〈𝑟𝑖∗ 〉𝑉𝑅 is the
The integral mass balance was applied for each product species of ethane
dehydrogenation (H2 and C2H4) in both the conventional reactor and the membrane
57
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
Where 𝜏𝑖,𝑀𝑅/𝑅 is the reaction rate ratio of product species between the
membrane reactor (MR) and the conventional reactor (R) mode, 𝛾𝑖,𝑀𝑅/𝑅 is the
outlet flow rate ratio for product species between the MR and the R mode, and
𝛽𝑖,𝑃/𝑅 is the ratio between product permeation flow rate over its formation rate in
the R mode. The dimensionless number 𝜏𝑖,𝑀𝑅/𝑅 represents the reaction rate
estimated in terms of the flow rate leaving the reaction zone, 𝑄𝑜𝑢𝑡 ; the outlet
permeate flow rate, 𝑄𝑝 and the outlet compositions from the reaction and permeate
All the dimensionless numbers defined for the analysis of the experimental
data are summarized in Table 4-1. They are also related to the Damköhler and
membrane Peclet numbers [64] estimated under reaction conditions. The Damköhler
number relates the rate of reaction to the rate of reactant feed. The membrane Peclet
number is defined in terms of the reactant feed rate and permeation rate of each
product.
58
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
Table 4-1 Dimensionless numbers defined for the comparison between reactor
and membrane reactor mode
Dimensionless number Symbol Definition
〈𝑟𝑖∗ 〉𝑉𝑅,𝑀𝑅
Reaction rate ratio 𝜏𝑖,𝑀𝑅/𝑅 𝜏𝑖,𝑀𝑅/𝑅 =
〈𝑟𝑖∗ 〉𝑉𝑅,𝑅
𝑥H2,𝑜𝑢𝑡
H2/ethane outlet molar ratio 𝜑H2⁄C2 𝜑 H2 ⁄C 2 =
𝑥C2,𝑜𝑢𝑡
ethylene yield and ethylene selectivity. Ethane conversion was calculated in terms
of the reaction products. For the reactor mode the ethane conversion is expressed
as
𝑥𝐶2 ,𝑜𝑢𝑡
𝛼𝐶2 = (1 − ) × 100% (4-6)
𝑥𝐶2 ,𝑜𝑢𝑡 +𝑥𝐶= ,𝑜𝑢𝑡 +0.5𝑥𝐶𝐻4 ,𝑜𝑢𝑡
2
where C2 and C2= represent the ethane and ethylene species respectively. The
59
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
𝑥𝐶= ,𝑜𝑢𝑡
2
𝑌𝐶2= = ( ) × 100% (4-7)
𝑥𝐶2 ,𝑜𝑢𝑡 +𝑥𝐶= ,𝑜𝑢𝑡 +0.5𝑥𝐶𝐻4 ,𝑜𝑢𝑡
2
gas. When a steady state was reached, it was switched to the reactor mode (R) by
shutting off the sweeping gas. Experimental results are shown in Figure 4-5. When
the sweep gas was shut off, hydrogen concentration on the feed side increased, and
ethylene concentration decreased accordingly. When the system was switched back
resumed back to the levels at the beginning of the MR steady state. The
60
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
10.0
Membrane reactor Reactor Membrane reactor
9.0
C2H6 x 1/10
Outlet molar fraction (%) 8.0
7.0
6.0
H2
5.0
4.0
C2H4
3.0
2.0
1.0
0.0
0 20 40 60 80 100 120 140 160
Time Lapse ∆t (min)
Figure 4-5 Dynamics of molar fractions of H2, C2H4, and C2H6 in the reactor outlet as the
system switched reversibly between a membrane reactor and a reactor operating mode at 500
Figure 4-6 shows the steady state concentrations of ethane, hydrogen, and
ethylene for both the membrane reactor (MR) and conventional reactor mode (R).
Figure 4-7 reveals the concentration changes of hydrogen and ethylene. Hydrogen
molar fraction increased from 4.5% to 4.8% when the system was switched from
membrane reactor mode to conventional reactor mode. Ethylene molar fraction was
61
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
higher in the membrane reactor mode than in conventional reactor mode. The
measured data were: [𝑄𝑜𝑢𝑡 ]𝑀𝑅 = 12.6 𝑚𝑙/𝑚𝑖𝑛 , [𝑄𝑜𝑢𝑡 ]𝑅 = 12.9 𝑚𝑙/𝑚𝑖𝑛 ,
9.0 Equilibrium
C2H6 x 1/10
8.0
Membrane reactor Reactor
Outlet molar fraction (%)
7.0
6.0
H2
5.0
4.0 Equilibrium
C2H4
3.0
2.0
1.0
0.0
0 20 40 60 80 100 120 140
Time Lapse ∆t (min)
Figure 4-6 Dynamics of outlet molar fractions of C2H6, H2, and C2H4 between membrane
62
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
6.0
Membrane reactor Reactor
5.5
4.5
Equilibrium
4.0
3.5
3.0
0 20 40 60 80 100 120
Time Lapse ∆t (min)
Figure 4-7 Dynamics of outlet molar fractions of H2 and C2H4 between membrane reactor
As shown in Table 4-2, the reaction rate ratio, 𝜏C=2 ,𝑀𝑅/𝑅 , calculated with
Equation (4-5) in terms of ethylene was 1.05, which indicates that the ethylene
production rate in the MR mode was higher than that in the R mode. The ethylene
yields (𝑌C=2 ) calculated with Equation (4-7) in the R and MR mode were 4.6% and
4.9% respectively. The results indicate that an ethylene yield enhancement of 6.5%
was reached in the MR mode. Ethane conversion increased from 4.9% in the R
rate, 𝛽H2 ,𝑝/𝑅 , was 16%, which meant that about 16% of the hydrogen generated in
the reaction was extracted through the membrane. The value 𝛽H2 ,𝑝/𝑅 =
63
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
The number approaching one means all the hydrogen produced in the
Table 4-2 Experimental values of the dimensionless numbers defined in Table 4-1 for R
and MR modes at 500 °C and 550 °C
Mode R MR R MR R MR
Membrane Reactor
reactor can extract more hydrogen from the reaction system while the hydrogen
generation rate in the conventional reactor mode remains the same under the same
64
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
reaction temperature. Therefore, the ratio of permeation rate/reaction rate, 𝛽H2 ,𝑝/𝑅 ,
increased. The measured data is: [𝑄𝑜𝑢𝑡 ]𝑀𝑅 = 12.1 𝑚𝑙/𝑚𝑖𝑛, [𝑄𝑜𝑢𝑡 ]𝑅 = 12.5 𝑚𝑙/
𝑚𝑖𝑛, permeate rate: 55.8, permeation compositions: 𝑥H2 ,𝑜𝑢𝑡 = 0.43 𝑣%, 𝑥C2=,𝑜𝑢𝑡 =
Figure 4-8 shows the feed and product concentration changes in the two
membrane area. With the large-area membrane, the hydrogen molar fraction
dropped considerably when the system was switched from conventional packed-
bed reactor mode to membrane reactor mode. The ethylene molar fraction values
65
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
10.0
Reactor Membrane reactor
9.0 Equilibrium
C2H6 x 1/10
8.0
Outlet molar fraction (%)
7.0
6.0
H2
5.0
4.0 Equilibrium
C2H4
3.0
2.0
1.0
0.0
0 10 20 30 40 50 60 70 80
Time Lapse ∆t (min)
Figure 4-8 Dynamics of outlet molar fractions of C2H6, H2, and C2H4 between
reactor and membrane reactor mode at 500 °C, WHSV = 0.4 h-1 with bigger
For the larger ratio of effective membrane area to reactor volume, 𝐴/𝑉𝑅 =
0.16 , the dimensionless number is 𝛽H2 ,𝑝/𝑅 = 36% , while for the smaller one
(𝐴/𝑉𝑅 = 0.04), the dimensionless number is 𝛽H2 ,𝑝/𝑅 = 16%, as shown in Table
4-2. Similarly, the H2/C2H6 molar ratio at the reactor outlet 𝜑H2⁄C2 decreased 80%
of its original value when the system was switched from the conventional packed-
bed reactor mode to the membrane reactor mode, while the 𝜑H2⁄C2 value only
decreased 3.8% (from 5.3% in the conventional reactor mode to 5.1% in the
membrane reactor mode) for the lower ratio of 𝐴/𝑉𝑅 = 0.04. Larger permeation
area favored more hydrogen pass through the membrane, and thus more hydrogen
66
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
was extracted from the reaction zone. As a result, the reaction rate, 〈𝑟𝑖∗ 〉𝑉𝑅,𝑀𝑅 , in the
membrane reactor mode increased and the ethane conversion, 𝛼C2 , improved in
corresponding value for the dimensionless number is 𝜏C=2 ,𝑀𝑅/𝑅 = 1.15, indicating a
15% increase in the reaction rate for a membrane reactor mode. Ethylene yield
improved about 17%, from 4.4% in the conventional reactor mode to 5.2% in the
dehydrogenation in the conventional reactor mode and the membrane reactor mode
at 550 °C. The ethylene selectivity at 550 °C decreased compared to that at 500 °C
(see Table 4-2) because more side reactions occurred at the higher temperature. The
reaction rate ratio, 𝜏C=2 ,𝑀𝑅/𝑅 = 1.09, was larger than the corresponding value at
500 °C. The ethane conversion at 550 °C increased from 9.7% in reactor mode to
10.5% in the membrane reactor mode. Other measured data were: [𝑄𝑜𝑢𝑡 ]𝑅 =
12.4 𝑚𝑙/𝑚𝑖𝑛 , [𝑄𝑜𝑢𝑡 ]𝑀𝑅 = 12.2 𝑚𝑙/𝑚𝑖𝑛 , Permeate rate: 54.2, permeate
composition [%v/v]: 𝑥H2 ,𝑜𝑢𝑡 = 0.50 𝑣% , 𝑥C=2 ,𝑜𝑢𝑡 = 0.16 𝑣% , and 𝑥C2 ,𝑜𝑢𝑡 =
0.60 𝑣%.
67
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
11.0 9.0
9.0 H2 7.0
8.0 6.0
C2H4
7.0 5.0
Reactor Membrane reactor
6.0 4.0
5.0 3.0
0 10 20 30 40 50 60 70 80
Time Lapse ∆t (min)
Figure 4-9 Dynamics of outlet molar fractions of C2H6, H2, and C2H4 between
reactor and membrane reactor mode at 550 °C, WHSV = 0.4 h-1
Higher reaction temperature led to higher product concentraion, and thus the
driving force for product permeation increased. Hydrogen and ethylene permeate
larger than the corresponding values at 500 °C with the same 𝐴/𝑉𝑅 ratio (see Table
4-2). At 500 °C, 𝜑H2⁄C2 ratio dropped about 3.8% of the original value when the
system was switched from the conventional ractor mode to the membrane reactor
mode, while it dropped about 11.6% of the original value at 550 °C. Product
original value by switching from reactor mode to membrane reactor mode at 500 °C,
68
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
while the yield increased 8.9% of the original value at 550 °C. The ethylene
As shown in Table 4-3, when the temperature increased from 500 °C to 550
°C, the Damköhler number (𝐷𝑎𝑅 ) doubled as the reaction rate increased twice with
the temperature. The increase of 𝐷𝑎𝑅 is also reflected in the conversion values risen
from ~5% at 500 °C to ~10% at 550 °C, as shown in Table 4-2. However, the larger
formation rate of hydrogen was compensated with a higher hydrogen removal rate.
The hydrogen permeation rate was three times larger at 550 °C than at 500 °C. The
Table 4-3. Consequently, the Damköhler-Peclet number (𝐷𝑎𝑅 𝑃𝑒) decreased from
6.1 to 4.2. Thus the effectiveness of the membrane reactor increased at higher
temperature. The ethylene yield enhancement at 550 °C was 9.4%, which was
higher than that at 500 °C. When the membrane area-to-volume ratio 𝐴/𝑉𝑅
increased from 0.04 to 0.16, the permeation/reaction rate ratio 𝛽H2 ,𝑝/𝑅 increased up
to 36% at 500 °C. This is equivalent to the 𝐷𝑎𝑅 𝑃𝑒 number decreasing to 2.7 . The
membrane reactor can work more effectively as 𝐷𝑎𝑅 𝑃𝑒 number approaches 1. The
ethylene yield enhanced about 17% at 500 °C in the membrane reactor mode.
Higher hydrogen removal rate verse its formation rate resulted in a larger ethylene
yield enhancement. The membrane reactor mode with the smallest 𝐷𝑎𝑅 𝑃𝑒 number
69
Chapter 4. Packed Bed Membrane Reactor for Ethane Dehydrogenation
Table 4-3 Ethylene yield enhancements achieved with the membrane reactor at two different
reaction temperatures and two 𝑨⁄𝑽𝑹 ratios. The corresponding Damköhler-Peclet numbers
were calculated with reaction conditions. WHSV = 0.4 h-1
1
,× 102 0.8 2.4 2.0
𝑃𝑒
4.4 Conclusions
Pt/Al2O3 catalyst packed bed was able to selectively extract hydrogen and to shift
the area/volume ratio of the membrane reactor module, a 15.6% ethylene yield
reactor at 500 °C. The effectiveness of the membrane reactor improved since more
hydrogen than the amount formed was removed from the system. The membrane
reactor showed a higher ethylene yield enhancement at 550 °C than at 500 °C, due
favored the higher ethylene yield improvement. The larger membrane disk at
70
Part II: Ethane Dehydrogenation Catalyst Development
Chapter 5. Ethane Dehydrogenation Catalyst Development
Development
5.1 Introduction
ranging from natural gas to heavy gas oils. The steam cracking technology is a
[15] [66]
highly energy intensive process with low selectivity (~75% per pass) .
Ethylene production from ethane (main component in natural gas or shell gas) is
the most economical and efficient process due to the abundant reserves and easy
exploitation of natural gas, and the high selectivity as shown in Table 5-1 . However,
temperature. For example, a typical furnace outlet temperature for cracking ethane
is approximately 800 °C, while the temperature for cracking naphtha or gas oils is
use of the membrane reactor can improve conversion by removing hydrogen from
the reaction system, and thus shifting the equilibrium to the direction favoring the
production of ethylene.
72
Chapter 5. Ethane Dehydrogenation Catalyst Development
Yield, wt%
H2 4.05 1.55 1.09 0.92 0.71
CH4 3.52 23.27 20.29 14.91 10.58
C2H2 0.47 0.51 0.42 0.47 0.38
C2H4 52.31 37.51 35.81 28.45 25.93
C2H6 35.03 2.8 4.16 4.23 2.82
C3H4 0.02 0.57 0.93 0.76 0.63
C3H6 1.13 14.82 17.24 15.64 14.07
C3H8 0.12 9.96 0.35 0.51 0.37
C4H4 0.05 0.08 0.1 0.13 0.12
C4H6 1.8 2.9 4.08 4.79 5.73
C4H8 0.19 1 2.83 4.52 3.61
C4H10 0.21 0.04 6.07 0.44 0.05
Benzene 0.47 2.12 2.69 7.12 5.44
Toluene 0.07 0.4 0.77 3.1 3.49
Xylenes 0 0.05 0.11 1.16 0.92
Ethylbenzene 0 0.01 0.01 0.62 0.37
Styrene 0.02 0.2 0.24 1.14 1.18
Pyrolysis gasoline 0.32 1.26 1.81 7.96 7.28
Pyrolysis fuel oil 0.16 0.89 0.94 3.09 16.3
E+P+B olefin 53.63 53.33 55.88 48.61 43.61
73
Chapter 5. Ethane Dehydrogenation Catalyst Development
alkanes like butanes, ethylbenzene because methyl C-H bonds are stronger than
methylene C-H bonds. It needs much higher reaction temperature to achieve the
practical conversions. The use of catalyst can lower activation energy because the
energy of C-H bonds. Moreover, C-H bonds in ethane (423 kJ/mol) are much
stronger than C-C bond (347~356 kJ/mol). C-H bonds in ethane are among the
[67]
strongest single bonds in nature . In order to achieve a high yield of ethylene
production, C-H bonds need to be selectively activated while keeping the C-C bond
stable. In this way, the side reactions such as coking and methane production are
temperatures traces back to early 1900s when transition metals such as Co, Ni, Cu,
[68]
Fe, Ag, Pt, and Pd were used as catalysts. Frey and Huppke carried out the
on chromic oxide gel catalyst in 1933, and found the catalysts lost activity at higher
[69]
temperature. Fabian and Robertson examined ethane dehydrogenation in a
A mass spectrometry was used to detect free radicals and atoms evaporating from
the filament. At very low pressure, the dehydrogenation reaction was a first order
reaction. Pt filament lost reactivity quickly and residual oil vapor was observed.
They found water and heavy water (D2O) vapor could reactivate the poisoned
74
Chapter 5. Ethane Dehydrogenation Catalyst Development
[71-73] [70]
dispersion of Pt on alumina support. , , , ,
[74]
and were studied as the supports for the dehydrogenation catalysts.
[75]
, carbon [76], hydrotalcite ( ) [77-81], monoliths [82, 83], [84]
, and
[85]
carbon nanotubes were also exploited as the supports for the dehydrogenation
catalysts.
Sn is a widely used promoter for the supported Pt catalyst. Bell’s group [80, 81, 86, 87]
Besides Sn, Fe [89], Rh [90], and Au [91] were also investigated as a second component
of Pt dehydrogenation catalysts.
been extensively investigated. Loaiza et al. [92] proposed that ethane was adsorbed
75
Chapter 5. Ethane Dehydrogenation Catalyst Development
[66]
Vincent et al. proposed a mechanism of generating methane by-product and
[93]
coke deposit. Adlhart and Uggerud used a Fourier transform ion cyclotron
[94]
elimination mechanism. Ju et al. proposed a and
concluded that the two bonds from the two carbon atoms synchronously broke
steps proposed by Loaiza, Vincent, Adlhart and Uggerud, and Ju et al. are consistent
Yagasaki and Masel [95] found that coke and methane were mainly formed on
at elevated temperatures. The key to supress coke deposits and methane by-product
was to prevent the re-adsorption of ethylene onto the Pt surface [80]. Rodriguez and
caused the density of Pt 5d electrons to shift to 6s and 6p. Zn has the similar
76
Chapter 5. Ethane Dehydrogenation Catalyst Development
larger than late transition metals (Group 8-11 metals, by IUPAC definition). The
Pt.
dissociated hydrogen atom and ethyl group (Reactions (5-1), (5-2), and (5-4)).
Hydrogen atom reacted with absorbed ethane to produce hydrogen gas (Reaction
(5-3)).
(5-1)
(5-2)
(5-3)
77
Chapter 5. Ethane Dehydrogenation Catalyst Development
(5-4)
(5-5)
(5-6)
(5-7)
(5-8)
(5-9)
(5-10)
(5-11)
(5-12)
78
Chapter 5. Ethane Dehydrogenation Catalyst Development
Heterogeneous catalysts are a complex system. Not only the supported active
metal components but also the support itself plays a key role in determining the
catalyst performances. During the reactions, both the active components and the
enhancing the synergetic effects and adjusting the interactions among active
components and the supports [97, 98]. This work reports the development and a study
of a novel catalyst with high reactivity and selectivity for ethane dehydrogenation:
5.2 Experimental
surface titania atoms to charged titanate species whose anionic charge is offset by
sodium or hydrogen ions. A commercial sample marked “filter cake” was used to
prepare the catalysts. The filter cake represents the material directly discharged
from the reactor, filtered, and washed without any other pre-treatment steps. The
filter cake was calcined at 600 °C for four hours and crushed into 30 to 50 mesh
pellets as catalyst support. The ETS-2 support pellets were modified with ZnO and
PdCl2 (Palladium (II) chloride, Pd 59.5%, Alpha Aesar, USA) (in 28wt%
All ETS-2 supports were calcined at 600 °C for four hours before
contents and then impregnated with 1 wt% of Pt. All catalysts were calcined at
500 °C for four hours except for the specified temperatures (600 and 650 °C). In
the same procedure, for the other metal catalysts, the ETS-2 support was
temperature overnight (> 8 hours), and then calcined at 500 °C for four hours. All
reactor as shown in Figure 5-1 was set up. The reactor was a quartz tube (ID 22
mm). The catalyst bed was supported by quartz wool on both sides. Both reduced
and non-reduced catalysts were dried at 110 °C for four hours before use. The
catalysts were reduced in situ with argon diluted hydrogen (40 v% H2). The
range from room temperature to 300 °C with the heating rate 2 °C/min; at 300 °C
80
Chapter 5. Ethane Dehydrogenation Catalyst Development
The amount of the catalyst loaded on each run was 3.0 g. An on-line GC
(Bruker 450-GC) was used. The ethane feed (99.5%, Praxair) flow rate was
controlled by a mass flow meter (Sierra Instruments Inc.). The outlet flow rate was
[𝐶2 𝐻4 ]
𝑆𝐶2 𝐻4 = [𝐶 × 100% (5-13)
2 𝐻4 ]+0.5[𝐶𝐻4 ]
Table 5-2 summarizes the BET specific surface areas of ETS-2, Al2O3
supported Pt catalysts, and Pt dispersion on the catalysts. From the table, it can be
seen that the commercial ETS-2 filter cake dried at 150 °C has a high surface area.
81
Chapter 5. Ethane Dehydrogenation Catalyst Development
When the sample was calcined at 600 °C for four hours, the surface area dropped
When 1 wt% of Pt was introduced in the ETS-2 support, the specific surface
area of 1%Zn/ETS-2 catalyst slightly increased. As the Zn content added into the
support increased, the specific surface areas of the catalysts gradually decreased,
and leveled off after the Zn content reached 5 wt%. The surface area decrease could
82
Chapter 5. Ethane Dehydrogenation Catalyst Development
sizes is illustrated in Figure 5-2. From the figure, it can be seen that the Pt dispersion
on the catalysts dramatically increased with Zn content, and reached the maximum
dispersion. When Zn content reached 5 wt% any additional Zn could not cause any
change in the dispersion, but the dispersion of Pt on the catalyst was still higher
than that on the catalyst without Zn. The Pt crystal size change followed the
shown in Figure 5-3. The introduction of Zn changed the ETS-2 surface properties,
and therefore enhanced the Pt dispersion on the surface and Pt particle size
reduction. The higher dispersion of Pt on the surface, the more active sites they can
60 12
Pt Dispersion on Catalysts, %
50 10
Pt dispersion
Pt Mean Particle Size, nm
40 8
30 6
20 4
10 2
0 0
0 1 2 3 4 5 6 7 8
Zn Content in the Catalysts 1%Pt/ETS-2, wt%
Figure 5-2 Pt mean particle size and dispersion change with Zn content
83
Chapter 5. Ethane Dehydrogenation Catalyst Development
60
50
Pt Particle Size, nm
40
30
20
10
0
0 2 4 6 8 10 12
Pt Dispersion, %
5.3.1.2 XRD
Figure 5-4 shows the XRD patterns of ETS-2 dried at 150 °C and calcined at
600 °C for four hours. For the dried sample, the TiO2 was in the form of anatase.
After calcined at 600 °C for four hours, some anatase phase was converted into
rutile phase. Figure 5-5 indicates the XRD patterns of the catalyst 5%Zn,
1%Pt/ETS-2 treated at different temperatures. From the figure, it can be seen that
new characteristic peaks are present in all catalyst’s patterns, which imply loaded
84
Chapter 5. Ethane Dehydrogenation Catalyst Development
whether those reactions will affect catalyst performance until further experiments
are conducted.
● ▲: Rutile
●: Anatase
●
Intensity (Counts)
ETS-2, 600°C
▲ ▲
● ●● ●
▲
▲▲ ●●
●
●
ETS-2, 150°C
●● ● ●●
0 10 20 30 40 50 60 70 80 90
2-Theta (degree)
Figure 5-4 XRD patterns of ETS-2 dried at 150 °C and calcined at 600 °C for
four hours
85
Chapter 5. Ethane Dehydrogenation Catalyst Development
(f)
(e)
Intensity (Counts)
(d)
(c)
(b)
(a)
0 10 20 30 40 50 60 70 80 90
2-Theta (degree)
Figure 5-5 XRD patterns of: (a) ETS-2 dried at 150 °C , (b) ETS-2 treated at
600 °C, (c) 5%Zn, 1%Pt/ETS-2 (500 °C calcined), (d) 5%Zn, 1%Pt/ETS-2
Figure 5-6 shows the SEM images of ETS-2 before calcination (a) and after
calcination at 600 °C (b). From the two images, it can be seen that the texture of
ETS-2 with and without calcination is made of 1-2 µm of irregular fine particles.
Low temperature dried samples are present as more flaky grains. After calcination
at the elevated temperature the particle of the sample became granular. The change
of morphology may count for the change of the specific surface area.
Figure 5-7 shows the images of the catalyst 5%Zn, 1%Pt/ETS-2 (a) calcined
and (b) tested at 500 °C. There are no visible aggregated Pt particles observed in
86
Chapter 5. Ethane Dehydrogenation Catalyst Development
either image. Figure 5-8 shows the images of the catalyst with the same
compositions but treated at 600 °C. After testing, aggregated Pt particles were
Figure 5-6 Industrial ETS-2 (a) dried at 150°for 4 hours; (b) calcined at 600 °C
for 4 hours
Figure 5-7 Catalyst 5%Zn, 1%Pt/ETS-2 (a) calcined at 500 °C for 4 hours; (b)
tested at 500 °C
87
Chapter 5. Ethane Dehydrogenation Catalyst Development
Figure 5-8 Catalyst 5%Zn, 1%Pt/ETS-2 (a) calcined at 600 °C for 4 hours; (b)
tested at 650 °C
Figure 5-9 shows the image of catalyst 1%Pt/ETS-2 calcined at 500 °C. In the
component will cause activity loss and changes in both stability and selectivity.
different conditions. From Figure 5-10 (a), it can be seen that Pt particles have
aggregated on the catalyst surface. After the catalyst was evaluated for ethane
dehydrogenation at 500 °C for four hours, Pt particles grew larger (b). When the
catalyst calcined at 600 °C, the aggregated Pt particle size were larger than at 500 °C.
Figure 5-10 TEM images of the catalyst 1%Pt, 5%Zn/ETS-2: (a) calcined at
500 °C, (b) spent (run at 500 °C), (c) calcined at 600 °C, (d) spent (run at 600 °C)
5.3.1.4 XPS
The catalyst 5%Zn, 1%Pt/ETS-2 was calcined at 500 °C and 600 °C for four
v% H2, 40ml/min, and 300 °C. Figure 5-11 shows XPS spectrum of the catalyst
89
Chapter 5. Ethane Dehydrogenation Catalyst Development
calcined at 600 °C before reduction. The deconvolutions of the spectrum for core
level Zn 2p3/2 revealed that the binding energies of Zn 2p3/2 could be deconvoluted
to 1022.5 eV and 1021.1 eV. The Zn precursor was Zn(NO3)2. When calcined at
associated to Zn species in ZnO, which showed good agreement with literature data
[99, 100]
. The binding energy 1021.1 eV corresponds to metallic Zn [99-101] generated
from the reduced ZnO. Large negative shift of the binding energy implied the
formation of Pt-Zn alloy [96]. By integrating the two peak areas, it can be calculated
XPS spectrum of the catalyst calcined at 500 °C before reduction was different
from that of the catalyst treated at 600 °C, as shown in Figure 5-12. Only one single
peak at 1021.3 eV was observed, which was ascribed to metallic Zn. Based on the
negative binding energy shift, it was predicted that probably Pt-Zn alloy was
formed.
hydrogen due to the positive free Gibbs energy under the present conditions.
However, the free Gibbs energy becomes negative if hydrogen atoms take part in
reduced first by hydrogen during the reduction process, and then, the reduced
atoms spilled over from Pt [102-106] to the adjacent zinc cations and reduced them to
their metallic state. Furthermore, when Pt and Zn formed an alloy, the enthalpy of
alloy formation was negative. The reduction of zinc cation was thermodynamically
90
Chapter 5. Ethane Dehydrogenation Catalyst Development
favorable. The similar binding energy of metallic Zn and Zn2+ made it difficult to
assess the zinc state by XPS. The Zn LMM Auger transition analysis is more
distinguishable [100].
temperatures indicate that ZnO reacted with support at higher temperature (600 °C)
600 °C
91
Chapter 5. Ethane Dehydrogenation Catalyst Development
Intensity, counts
500 °C
5.3.1.5 TPR
Figure 5-13 illustrates the H2-TPR profiles of the three catalysts: (1) no Zn
92
Chapter 5. Ethane Dehydrogenation Catalyst Development
H2 Consumption, a.u.
TPR curves reflected the difference in the reduction ability of the catalysts
modified with different metals. The Pt metal on the catalyst 1%Pt/ETS-2 was much
hydrogen uptake gradually decreased. Starting from 130 °C for 5%Zn/ETS-2, the
Zn was not reduced until 330 °C. A gradual increased curve after 330 °C revealed
that increasing amounts of Zn were reduced at relatively high temperatures. For the
their reduction behaviors. Two reduction peaks (H2 consumption) appeared at low
temperatures (110 °C and 160 °C), and one peak at 360 °C.
93
Chapter 5. Ethane Dehydrogenation Catalyst Development
Figure 5-14. In the temperature range from room temperature to 600 °C, the total
weight loss was only 1.4 wt%. The MS analysis proved that all the weight loss was
associated with the water that was due to the desorption of the physically adsorbed
reactions was identified. However, a carbon dioxide emission peak around 400 °C
in the TG-MS analysis of the spent alumina supported catalyst was identified, as
shown in Figure 5-15. Carbon dioxide emission indicated there was coke deposit
on the spent catalyst. The amount of water loss of the alumina supported catalyst
1.5E-13
99.5%
1.0E-13
Weight Loss
Ion Current
5.0E-14 99.0%
0.0E+00
98.5%
-5.0E-14
-1.0E-13 98.0%
0 100 200 300 400 500 600
Temperature, °C
94
Chapter 5. Ethane Dehydrogenation Catalyst Development
3.0E-13 100.0%
2.0E-13 98.0%
96.0%
1.0E-13
Weight Loss
Ion Current
94.0%
0.0E+00
92.0%
-1.0E-13
90.0%
-2.0E-13 88.0%
-3.0E-13 86.0%
0 100 200 300 400 500 600
Temperature, °C
Dehydrogenation
The noble metals, Pt and Pd, are the main catalytic components in ethane
the same amount of platinum and palladium were loaded on ETS-2 support
respectively as the catalysts for ethane dehydrogenation. Figure 5-16 shows the
ethane dehydrogenation product distributions on the two catalysts. From the figure,
it can be seen that the Pt/ETS-2 catalyst exhibited a higher ethylene and lower
hydrogen and methane yields at steady state. About 0.5 wt% of by-product methane
95
Chapter 5. Ethane Dehydrogenation Catalyst Development
was produced on both catalysts. Pt was a better active component for ethylene
CH₄ C₂H₄ H₂
7.0
6.0
5.0
Yields, %
4.0
3.0
2.0
1.0
0.0
1%Pd/ETS-2 1%Pt/ETS-2
Catalyats
Figure 5-16 Product spectrum obtained with the 1%Pd/ETS-2 and 1%Pt/ETS-
Figure 5-17 shows the yield changes using 1%Pt/ETS-2 as the catalyst at
h-1 is also shown in the figure. The product yields on 1%Pt/ETS-2 decreased with
the space velocity increasing. Even at higher space velocity (WHSV=0.80 h-1), the
ethylene yield of 1%Pt/ETS-2 was still higher than that of 1%Pd/ETS-2 at the lower
space velocity (WHSV=0.27 h-1). The result further proved that the Pt catalyst had
96
Chapter 5. Ethane Dehydrogenation Catalyst Development
5
1% Pt/ETS-2
4
Yields, %
3
H2
2 C2H4
1 CH4
0
6
5
1% Pd/ETS-2
Yields, %
4
3
2
1
0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
-1
WHSV, h
Figure 5-17 Product yield change with space velocity at 509 °C over catalysts
Dehydrogenation
To study the effect of the supports on the reaction performances of the ethane
supports, ETS-2, Al2O3, and natural zeolite clinoptilolite (impregnated with 1 wt%
reaction under the same conditions are illustrated in Figure 5-18. Catalyst
1%Pt/ETS-2 gave the lowest yield of by-product methane, but the highest yield of
1%Pt/clinoptilolite had the highest hydrogen yield and the lowest ethylene yield.
97
Chapter 5. Ethane Dehydrogenation Catalyst Development
Obviously, ETS-2 was the best support used on the catalyst for ethane
dehydrogenation.
CH₄ C₂H₄ H₂
8.0
7.0
6.0
Yields, %
5.0
4.0
3.0
2.0
1.0
0.0
1%Pt/ETS-2 1%Pt/Al₂O₃ 1%Pt/clinoptilolite
Catalysts
and 1%Pt/clinoptilolite catalysts at 509 °C, and ethane WHSV = 0.28 h-1
Dehydrogenation
and simultaneously suppressed the production of methane generated from the side
Hydrogen yield decreased slightly, while ethylene yield increased slightly. The
98
Chapter 5. Ethane Dehydrogenation Catalyst Development
dehydrogenation reaction.
7
H2
C2H4
Yields, %
6
CH4
5
4
1% Pt/Al2O3
7
6
5
Yields, %
4
3
1% Pt, 1% Zn/Al2O3
2
1
0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00 2.25 2.50
WHSV, h-1
Figure 5-19 Product concentration change with WHSV at 509 °C over catalysts
The product yield changes with time on 1%Pt/ETS-2 and 0.5%Zn, 1%Pt/ETS-
2 are shown in Figure 5-20 and Figure 5-21 respectively. The conversion on
1%Pt/ETS-2 catalyst without Zn was lower than the equilibrium value. The side
99
Chapter 5. Ethane Dehydrogenation Catalyst Development
6.0
5.0
4.0
Yields, %
3.0
2.0
1.0
0.0
20 30 40 50 60 70 80 90 100
Time on Stream, min
The addition of 0.5wt% Zn into 1%Pt/ETS-2 did not promote the conversion
since the ethylene yield was still below equilibrium, and side reactions were still
present (CH4 still presenting) even at relatively high feed flow rate. However, the
Increasing the Zn content to 1wt%, the activity increased and stability was
the middle of the catalyst bed was lower than that near the reactor wall because of
the temperature gradient. So, the overall product yields were between the
equilibrium and the highest point near the wall. However, side reactions (CH4 by-
100
Chapter 5. Ethane Dehydrogenation Catalyst Development
6.0
5.0
4.0
Yields, %
3.0
2.0
1.0
0.0
75 85 95 105
Time on Stream, min
When the WHSV increased from 0.88h-1 to 1.14 h-1, the by-product methane
disappeared, as shown in Figure 5-23. The ethylene yield still stayed around
equilibrium. Meanwhile, hydrogen yield was gradually reduced with reaction time
101
Chapter 5. Ethane Dehydrogenation Catalyst Development
H₂ CH₄ C₂H₄
Equilibrium@509°C Equilibrium@514°C
6.0
5.0
4.0
Yields, %
3.0
2.0
1.0
0.0
10 20 30 40 50
Time on Stream, min
Figure 5-22 Product concentration changes with time over 1%Zn, 1%Pt/ETS-
H₂ CH₄ C₂H₄
Equilibrium@509°C Equilibrium@514°C
6.0
5.0
4.0
Yields, %
3.0
2.0
1.0
0.0
50 60 70 80 90
Time on Stream, min
Figure 5-23 Product concentration changes with time over 1%Zn, 1%Pt/ETS-
wt%, the hydrogen and ethylene yields were similar at low feed flow rate, even
though side reactions still existed. However, when feed flow rate increased,
methane disappeared, and the ethylene and hydrogen yield became identical.
H₂ CH₄ C₂H₄
Equilibrium@509°C Equilibrium@514°C
6.0
5.0
4.0
Yields, %
3.0
2.0
1.0
0.0
15 25 35 45 55 65
Time on Stream, min
Figure 5-24 Product yield changes with time over 2%Zn, 1%Pt/ETS-2 at
103
Chapter 5. Ethane Dehydrogenation Catalyst Development
H₂ CH₄ C₂H₄
Equilibrium@509°C Equilibrium@514°C
6.0
5.0
4.0
Yields, %
3.0
2.0
1.0
0.0
70 80 90 100
Time on Stream, min
Figure 5-25 Product yield changes with time over 2%Zn, 1%Pt/ETS-2 at
space velocity (0.89 h-1). The conversion reached equilibrium state, as shown in
Figure 5-26.
As shown in Figure 5-27, when Zn content increased to 7.5 wt%, the side
reactions disappeared at low space velocity, the product yields were maintained
around the equilibrium, and the activity increased further. The hydrogen yield was
slightly higher than that of the ethylene. Zn modification enhanced the catalyst
stability.
104
Chapter 5. Ethane Dehydrogenation Catalyst Development
H₂ CH₄ C₂H₄
Equilibrium@509°C Equilibrium@514°C
6.0
5.0
4.0
Yields, %
3.0
2.0
1.0
0.0
15 25 35 45 55
Time on Stream, min
Figure 5-26 Product yield changes with time over 5%Zn, 1%Pt/ETS-2 at
H₂ CH₄ C₂H₄
Equilibrium@509°C Equilibrium@514°C
6.0
5.0
4.0
Yields, %
3.0
2.0
1.0
0.0
10 20 30 40
Time on Stream, min
Figure 5-27 Product yield changes with time over 7.5%Zn, 1%Pt/ETS-2 at
The production yield changes with the Zn content were summarized in Figure
5-28. It can be seen that, to reach the reaction equilibrium, more than 1 wt% of Zn
H₂ CH₄ C₂H₄
Equilibrium@509°C Equilibrium@514°C
6.0
5.0
Yields, %
4.0
3.0
2.0
1.0
0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0
Zn, wt%
For Zn modified 1%Pt/ETS-2 catalysts, the yield changes of the main product
ethylene, hydrogen, and by-product methane with different WHSV are shown in
Figure 5-29. As expected, the product yields decreased with WHSV for all catalysts.
However, the extent was quite different. From Figure 5-29, it can be seen that the
106
Chapter 5. Ethane Dehydrogenation Catalyst Development
ethylene yields for all Zn modified catalysts were higher than that for the catalyst
studied. Among the Zn modified catalysts, 1%Pt/ETS-2 with 1 wt%, 2 wt%, and 5
wt% of Zn had similar high ethylene yields at all space velocities, but 1 wt% and 2
wt% of Zn modified catalysts had relatively higher methane yield at low space
velocity. The catalyst 1%Pt, 5%Zn/ETS-2 had relatively stable ethylene yields at
low space velocities, and its methane yields were suppressed to zero at all space
velocities. The catalyst 1%Pt, 7.5%Zn/ETS-2 had the similar properties as 1%Pt,
6 6
5 5
4 4
C2H4, %
H2, %
3 3
2 2
0.5 0.4 0.6 0.8 1.0 1.2 1.4 1.6 0.4 0.6 0.8 1.0 1.2 1.4 1.6
-1
0.4 WHSV, h
0.3
1%Pt/ETS-2
CH4, %
WHSV, h-1
Figure 5-29 C2H4, H2, and CH4 yield change with space velocity on different
107
Chapter 5. Ethane Dehydrogenation Catalyst Development
100
99
Selectivity, %
98
97
WHSV=0.87h-1
WHSV=1.13h-1
96
0 1 2 3 4 5 6 7 8
Zn, wt%
the space velocities 0.87 and 1.13 h-1. From the figure, it can be seen that selectivity
was able to reach 100% even at lower Zn content for high space velocity (1.13 h-1).
To study the promotion effect with the other metals, 1%Pt/ETS-2 was
modified with Zn and Ce. The ethylene, hydrogen and methane yields at various
108
Chapter 5. Ethane Dehydrogenation Catalyst Development
WHSVs are shown in Figure 5-31. From the figure, it can be seen that Zn enhanced
ethylene yield, while Ce it. However, Ce could not eliminate side reactions even at
1%Zn/ETS-2) was higher than that on only Zn and Ce modified catalysts (1%Pt,
1%Zn, 2.5%Ce/ETS-2). Zn and Zn-Ce modified catalysts could suppress the side
also stabilize ethylene yields at higher space velocities. Hydrogen yield change was
6 6
5 5
C2H4, %
4 4
H2, %
3 3
2 2
1 1
0.6 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
0.3 1%Pt/ETS-2
0.2
1%Pt, 2.5%Ce/ETS-2
1%Pt, 1%Zn/ETS-2
0.1
1%Pt, 1%Zn, 2.5%Ce/ETS-2
0.0
Figure 5-31 C2H4, H2, and CH4 yields change with WHSV at 509 °C over
109
Chapter 5. Ethane Dehydrogenation Catalyst Development
To investigate the Ce modification effect for the support Al2O3, the catalysts
modified with 1%Zn and 1%Zn+2.5%Ce were prepared. The ethylene, hydrogen,
and methane yields at the various WHSVs are shown in Figure 5-32. Ethylene yield
increased with space velocity for both catalysts, but both hydrogen and by-product
methane yields decreased with space velocity. For the catalyst 1%Pt, 1%Zn/Al2O3,
Ce inhibited the catalyst activities and thus decreased the ethylene yield. However,
Ce modification did not help suppress by-product methane production for the
5.2 7.2
5.1 7.0
5.0 6.8
4.9 6.6
C2H4, %
H2, %
4.8 6.4
4.7
6.2
4.6
6.0
4.5
5.8
4.4
4.0 1.0 1.5 2.0 2.5 3.0 1.0 1.5 2.0 2.5 3.0
-1
3.5 WHSV, h
3.0
CH4, %
2.5
0.5
1.0 1.5 2.0 2.5 3.0
-1
WHSV, h
Figure 5-32 C2H4, H2, and CH4 yields change with space velocities at 509 °C
110
Chapter 5. Ethane Dehydrogenation Catalyst Development
Sn/Al2O3 catalyst are shown in Figure 5-33. From Figure 5-33, it can be seen that
CH₄ C₂H₄ H₂
7.0
6.0
5.0
Yield, %
4.0
3.0
2.0
1.0
0.0
1% Pd, 1% Sn/Al₂O₃ 1% Pd, 1% Sn, 1%Ce, 1% Zn/Al₂O₃
Catalysts
Figure 5-33 Product spectrum obtained with the 1%Pd, 1%Sn/Al2O3 and
1%Pd, 1%Ce, 1%Zn, 1%Sn/Al2O3 catalysts at 509 °C, and ethane WHSV =
0.28 h-1
111
Chapter 5. Ethane Dehydrogenation Catalyst Development
5.4 Conclusions
The Zn modification helped the Pt disperse on ETS-2 and Al2O3 supports, and
performances. Pt was aggregated at high temperature. ZnO could react with the
support to form new species. In the Pt catalysts, ZnO can be easily reduced by
hydrogen. The new phase formation and strong interaction with the support of ZnO
Pd; ETS-2 support was superior in reaction specificity to Al2O3 and clinoptilolite.
and stability significantly. The side reactions like coking and methane formation
content increase, the activity of the catalysts increased, and the methane production
was greatly suppressed. Pt-Zn alloy was formed and might act as the active species
for ethane dehydrogenation that helped to reduce side reactions and shorten the time
to reach the equilibrium. Residence time was a main factor that affected selectivity
and activity.
112
Chapter 6. Conclusions and Suggestions for Future Work
fragments procured from New Zealand have good machinability and thermal
stability. The crystal structure is stable up to 800 °C. The zeolite mordenite has
uniform pore size distribution and has internal channels which can selectively
separate hydrogen. All these features make the mordenite a good candidate as a
Pt/Al2O3 catalyst packed bed was able to selectively extract hydrogen and shifted
area/volume ratio of the membrane reactor module from 0.04 m-1 to 0.16 m-1, the
ethylene yield increased from 6.5% to 15.6%. The membrane reactor showed a
higher ethylene yield enhancement at 550 °C than at 500 °C, due to the increase in
hydrogen permeation rate. Ethylene yield increased from 6.5% at 500 °C to 8.9%
at 550 °C compared to the conventional packed bed reactor with the area/volume
ratio 0.04 m-1. Lower Damköhler-Peclet number favored the higher ethylene yield
improvement.
113
Chapter 6. Conclusions and Suggestions for Future Work
Zn alloy phase generated on catalyst surface could be active species for ethane
Pt/ETS-2 catalyst improved the ethylene selectivity and stability significantly. The
side reactions like coking and methane formation were effectively suppressed. As
the Zn content increased, the activity of the catalysts increased as well, while the
methane production was greatly suppressed. Residence time was one of the main
Selectivity and permeance are key factors for membrane performance and
selectivity depends on √𝑀𝐵 ⁄𝑀𝐴 , because the square root of the molecular weight
ratio of the two gas molecules is a small number. However, Knudsen flow occurs
in pores 2 < 𝑑 < 100 𝑛𝑚 , which is in meso-pores (2 < 𝑑 < 50 𝑛𝑚) and small
also occurs in the range (𝑑 > 20 𝑛𝑚) that Knudsen flow does, but it is not selective.
114
Chapter 6. Conclusions and Suggestions for Future Work
than Knudsen diffusion, high permeance, and good thermal stability at working
temperatures.
coking and methane formation. However, its activity could be improved (higher
space velocity). Compared to alumina support, its specific surface area is very low,
which affects the dispersion of active metal species like Pt, reduces the active sites
on the surface, and thus limits the activity. Therefore, further modification of the
ETS-2 is not stable at high temperature. It can easily change phase above
600 °C, and it is also reactive with supported metal components at high temperature.
These will cause the catalyst to lose activity and will reduce the run cycles. This is
catalyst and the support, such as optimal metal compositions and chemistry of
115
Chapter 6. Conclusions and Suggestions for Future Work
catalyst preparation, will help understand reaction mechanisms and control catalyst
quality.
116
References
[2] W. R. True: "Global Ethylene Capacity Poised for Major Expansion", Oil &
[3] W. R. True: "Global Ethylene Capacity Continues Advance in 2011", Oil &
[4] L. Bewley: "Ihs Wpc 2012: Shale Reshapes Ethylene Markets", Chemical
[6] C. Carr: "Global Ethylene Outlook: One Product....... Many Strategies", 2013,
http://ihsglobalevents.com/wpc2013/files/2013/03/Carr_20131.pdf.
[8] K. Bullis: "Shale Gas Will Fuel a U.S. Manufacturing Boom", MIT
[10] The Dow Chemical Company: "Dow to Build New Ethylene Production Plant
117
[11] Nexant: "North American Shale Gas: Opportunity or Threat for Global
http://www.chemsystems.com/reports/search/docs/prospectus/STMC12_Sha
le%20Gas_pros.pdf.
[12] K. Holmquist: "Global Ethylene Surplus to Last through 2011", Oil & Gas
Industrial Chemistry", 6th edn, 465-529; 2009, Wiley-VCH Verlag GmbH &
Co. KGaA.
[15] T. Ren, M. Patel, and K. Blok: "Olefins from Conventional and Heavy
Area and Porosity", Pure and Applied Chemistry, 1985, 57 (4), 603-619.
118
[18] H. P. Hsieh: 'Chapter 4 Physical, Chemical and Surface Properties of
1995, Elsevier.
Arbor, 2005.
(11), 2101-2136.
119
[25] C. Téllez and M. Menéndez: '8 Zeolite Membrane Reactors', in "Membranes
Basile and F. Gallucci), 243-273; 2011, John Wiley & Sons, Ltd.
120
Reactor", Angewandte Chemie-International Edition, 2010, 49 (33), 5656-
5660.
Reactors for the Oxidative Activation of Ethane", Catalysis Today, 2006, 118
(1-2), 98-103.
121
[37] M. W. Ackley, S. U. Rege, and H. Saxena: "Application of Natural Zeolites
[38] D. C. L. How: "Xi.-on Mordenite, a New Mineral from the Trap of Nova
122
Frameworks", The Journal of Physical Chemistry B, 1998, 102 (10), 1759-
1767.
[49] G. Horvath and K. Kawazoe: "Method for the Calculation of Effective Pore
234.
123
[52] D. L. Bish and J. W. Carey: "Thermal Behavior of Natural Zeolites", Reviews
124
[59] E. Gobina, K. Hou, and R. Hughes: "Ethane Dehydrogenation in a Catalytic
2059-2072.
414-419.
[64] S. T. Oyama and H. Lim: "An Operability Level Coefficient (Olc) as a Useful
C. Diniz da Costa: "An Analysis of the Peclet and Damkohler Numbers for
125
[66] R. S. Vincent, R. P. Lindstedt, N. A. Malik, I. A. B. Reid, and B. E. Messenger:
Propane, and the Butanes", Industrial and Engineering Chemistry, 1933, 25,
54-59.
16.
471-475.
126
[73] C. De la Cruz and N. Sheppard: "Adsorption and Dehydrogenation of Ethane
1 (2), 329-332.
and Catalytic Properties of Nickel and Platinum Impregnated Eta and Gamma
127
[79] A. Virnovskaia, E. Rytter, and U. Olsbye: "Kinetic and Isotopic Study of
(2), 209-219.
(1-4), 61-67.
Effects of Particle Size and Sn/Pt Ratio on Catalyst Activity, Selectivity, and
128
Stability", Abstracts of Papers of the American Chemical Society, 2012, 244,
1.
187-194.
[89] J. Jia, J. Shen, L. Lin, Z. Xu, T. Zhang, and D. Liang: "A Study on Reduction
Very Short Contact Times', in "Studies in Surface Science and Catalysis", (eds.
[92] A. Loaiza, M. Xu, and F. Zaera: "On the Mechanism of the H–D
129
Cations, Pt-N(+) N=1-21", Chemistry-a European Journal, 2007, 13 (24),
6883-6890.
1125.
[99] S. W. Gaarenstroom and N. Winograd: "Initial and Final State Effects in the
130
[100] J. Silvestre-Albero, A. Sepúlveda-Escribano, F. Rodrı́guez-Reinoso, and J. A.
[102] R. Prins: "Hydrogen Spillover. Facts and Fiction", Chemical Reviews, 2012,
131