169
DOCTORAL THESIS Zeolite Membrane Separation at Low Temperature Pengcheng Ye Chemical technology

Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

DOCTORA L T H E S I S

Department of Civil, Environmental and Natural Resources EngineeringDivision of Chemical Engineering Zeolite Membrane Separation

at Low Temperature

Pengcheng Ye

ISSN 1402-1544ISBN 978-91-7583-557-0 (print)ISBN 978-91-7583-558-7 (pdf)

Luleå University of Technology 2016

Pengcheng Ye Zeolite M

embrane Separation at Low

Temperature

Chemical technology

Page 2: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Thesis for the Degree of Doctor of Philosophy

Zeolite Membrane Separation at Low Temperature

Pengcheng Ye

Chemical Technology

Division of Chemical Engineering Department of Civil, Environmental and Natural Resources Engineering

Luleå University of Technology

SE-971 87 Luleå

April 2016

Page 3: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

ii

Cover image: Outstanding performance of zeolite membrane for N2/He separation at low temperature with a membrane selectivity of 76 at 124 K.

Page 4: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 5: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

iv

Page 6: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

v

ACKNOWLEDGEMENTS

Life is like a large river.

Being a lucky man, I am grateful for meeting several excellent individuals during the journey.

Thanks for all the support that makes this work possible. First of all, I would like to thank my supervisor Prof. Jonas Hedlund for offering me the opportunity to enter this fantastic research group with smart and kind-hearted people. Thanks for helping me and guiding me when the research came to bottlenecks. Your great ambition has always inspired me in my research. I would also like to acknowledge my assistant supervisor Prof. Mattias Grahn, an amiable guy. It is always pleasant to have discussions with you.

Special thanks to my assistant supervisor Dr. Danil Korelskiy for reinforcement in various aspects and frequent discussions of research. In addition, I really appreciate your warm-hearted support in private life as a close friend, especially during the initial tough months.

Moreover, I am extremely grateful to Göran Wallin for all the time spent on my experimental rig construction and persistent modifications. You did a fantastic job! I feel so lucky that I can work with such a competent and enthusiastic man.

I would like express my gratitude to Dr. Linda Sandström, Assoc. Prof. Allan Holmgren, Dr. Liang Yu, Dr. Han Zhou, Dr. Erik Sjöberg and Dr. Ming Zhou in the zeolite membrane group for fruitful discussions and help related to my doctoral thesis topic. It has been great pleasure to work with such erudite and experienced researchers.

I would like to thank all my colleagues at Chemical Technology. You make my life cheerful and amazing in Sweden. In addition, I am thankful to friends from the division and the department for fikas, parties and other delightful moments. There is an army of names on the list here, which will be kept in my heart. It is a beautiful memory for me.

The Swedish Research Council and the Swedish Energy Agency are gratefully acknowledged for the financial support of my work.

Jag vill verkligen säga ett stort tack till mina vänner Adrian Rodriguez, Göran Wallin, Lubomir Novotny, Tommy Karlkvist och Lindsay Ohlin för ert stöd med min svenska språkinlärning och alltid ett vackert sällskap.

Mom and Dad, you are the best parents in the world! I am giving you a huge hug for your unconditional support since the beginning of my life journey. I love you! 我爱你们, 我的家

人们!谢谢过去和将来陪伴在我生命旅途中的人。

Page 7: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

vi

Page 8: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

vii

LIST OF PAPERS

The thesis is based on the following papers:

I. Air Separation at Cryogenic Temperature Using MFI Membranes

Pengcheng Ye, Erik Sjöberg and Jonas Hedlund

Microporous and Mesoporous Materials, 192 (2014) 14-17.

II. Cryogenic Air Separation at Low Pressure Using MFI Membranes

Pengcheng Ye, Danil Korelskiy, Mattias Grahn and Jonas Hedlund

Journal of Membrane Science, 487 (2015) 135-140.

III. Efficient Separation of N2 and He at Low Temperature Using MFI Membranes

Pengcheng Ye, Mattias Grahn, Danil Korelskiy, and Jonas Hedlund

Accepted for publication in AIChE Journal.

IV. A Uniformly Oriented MFI Membrane for Improved CO2 Separation

Ming Zhou, Danil Korelskiy, Pengcheng Ye, Mattias Grahn, and Jonas Hedlund

Angewandte Chemie (International Edition), 53 (2014) 3492-3495.

V. Efficient Ceramic Zeolite Membranes for CO2/H2 Separation

Danil Korelskiy, Pengcheng Ye, Shahpar Fouladvand, Somayeh Karimi, Erik Sjöberg, and Jonas Hedlund

Journal of Materials Chemistry A, 3 (2015) 12500-12506.

VI. Ultra-Thin MFI Membrane with High Performance for Light Olefins/Nitrogen Separations

Liang Yu, Mattias Grahn, Pengcheng Ye, and Jonas Hedlund

To be submitted to Journal of Membrane Science.

VII. A Study of CO2/CO Separation by Sub-micron b-Oriented MFI Membranes

Danil Korelskiy, Mattias Grahn, Pengcheng Ye, Ming Zhou, and Jonas Hedlund

Submitted to RSC Advances.

Page 9: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

viii

My contribution to the appended papers:

I. Most of the planning and establishing the final method. All experimental work, evaluation. Nearly all writing.

II. A large part of planning and construction of the new experimental set-up for cryogenic gas separation. Nearly all experimental work and writing. Data evaluation work apart from the modeling.

III. Nearly all the experimental work and data evaluation apart from the modeling. All the writing apart from the modeling section.

IV. Participation in all the experimental work on membrane evaluation for gas separation. Participation in discussion of the separation results and data evaluation.

V. Participation in the experimental work on membrane evaluation for gas separation. Participation in manuscript preparation and revision, apart from the modeling and the economic analysis.

VI. Participation in the experimental work, e.g., establishing the analysis methods by GC, gas separation experiment. Participation in manuscript preparation.

VII. Participation in the experimental work on membrane evaluation for gas separation. Participation in manuscript preparation.

Page 10: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

ix

Other papers by the author related to but not included in the thesis:

I. An Experimental Study of Micropore Defects in MFI Membranes

Danil Korelskiy, Pengcheng Ye, Han Zhou, Johanne Mouzon and Jonas Hedlund

Microporous and Mesoporous Materials, 192 (2014) 69-75.

Conference contributions (not included in the thesis):

I. Air Separation at Cryogenic Temperature Using MFI Membranes

Pengcheng Ye, Erik Sjöberg and Jonas Hedlund

Poster presentation at the 6th International Zeolite Membrane Meeting in June 2013 in Jeju Island, Korea.

II. An Experimental Study of Micropore Defects in MFI Membranes

Danil Korelskiy, Pengcheng Ye, Han Zhou, Johanne Mouzon and Jonas Hedlund

Poster presentation at the 6th International Zeolite Membrane Meeting in June 2013 in Jeju Island, Korea.

II. Cryogenic Air Separation at Low Pressure Using MFI Membranes

Pengcheng Ye, Danil Korelskiy, Mattias Grahn and Jonas Hedlund

Poster presentation at the International Symposium on Zeolite and Microporous Crystals in June-July 2015 in Sapporo, Japan.

III. High Performance MFI Membranes for Efficient CO2 Removal

Danil Korelskiy, Pengcheng Ye, Shahpar Fouladvand, Somayeh Karimi, Erik Sjöberg, and Jonas Hedlund

Oral presentation at the International Symposium on Zeolite and Microporous Crystals 2015 in June-July 2015 in Sapporo, Japan.

IV. Efficient Separation of N2 and He Using MFI Membranes at Low Temperature

Pengcheng Ye, Danil Korelskiy, Mattias Grahn and Jonas Hedlund

Poster presentation at Euromembrane 2015 in September 2015 in Aachen, Germany.

Page 11: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

x

Page 12: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

xi

ABSTRACT

The energy consumption of separation processes accounts for a large part of the total energy consumption in chemical industry. Membrane separation processes require much less energy than the currently used thermally driven separation processes and could therefore reduce energy consumption in industry considerably. Today, most commercially available membranes are organic polymeric membranes. Inorganic zeolite membranes have several superiorities over polymeric membranes, e.g., higher flux and selectivity, higher chemical and thermal stability, and thus have great potential for a variety of gas and liquid separations. Whereas there have been extensive studies on zeolite membrane separation at high temperature during the past decades, scientific reports on the low temperature applications of zeolite membranes is extremely scarce and there are no reports at cryogenic temperature. This work is pioneering research on investigation of the performance of zeolite membranes for separation of various gas mixtures at unprecedentedly low temperature, down to cryogenic temperature.

In the present work, zeolite membranes were, for the first time, evaluated for gas separation at cryogenic temperature. Air separation by ultra-thin MFI membranes was carried out at a feed pressure ranging from 100 mbar to 5.0 bar over the temperature range of 62–110 K. The membranes were found to be oxygen selective at all the conditions investigated. The observed results were well above the upper bound in the 2008 Robeson selectivity-permeability plot for polymeric membranes when the feed pressure was less than or equal to 1.0 bar. The O2/N2 separation factor reached 5.0 at 67 K and 100 mbar, with a high O2 permeance of 8.6 × 10-7 mol m-2 s-1 Pa-1. The performance of our membranes (in terms of selectivity) was comparable to that recently reported for the most promising polymeric membranes under development by research groups, but 100 times higher in terms of permeance and flux. The membrane selectivity was found to increase with decreasing temperature and feed pressure. The present work has therefore indicated the optimum conditions for air separation using MFI membranes, namely low feed pressures and cryogenic temperatures. A mathematical model showed that the membrane selectivity emanated from O2/N2 adsorption selectivity.

N2/He separation is essential for helium recovery from natural gas and helium reclamation for airships and submarines. MFI zeolite membranes were evaluated for this separation over the temperature range of 85–260 K, possessing high N2-selectivity at all the conditions investigated. When the feed pressure was 5.0 bar and the permeate pressure was 0.5 bar, a highest N2/He separation factor of 62 was observed at 124 K. The N2 permeance was rather high, up to 39 ×10−7 mol m−2 s−1 Pa−1. The separation was attributed to adsorption selectivity to N2, effectively suppressing the transport of He in the zeolite pores and this effect was more significant at low temperature. A mathematical model showed that the largest difference of adsorbed loading over the film at ca. 120 K was probably the main reason for the observed maximum selectivity at this temperature. The model also indicated that the selectivity could

Page 13: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

xii

even be increased by 2–3 times if the membrane was totally defect-free. This work demonstrates that a zeolite membrane process could be rather competitive for N2/He separation.

Synthesis gas generated from biomass is a valuable, renewable resource that can be used for production of clean energy and various chemicals. It is mainly a mixture of CO, CO2, and H2. CO2 is usually an undesired component in the syngas and should, therefore, be removed. In this work, CO2 separation from H2 and CO using MFI zeolite membranes was studied at low temperatures, down to 235 K and at a feed pressure of 9.0 bar. The membrane performance in terms of both selectivity and flux was superior to that reported for the state-of-the-art polymeric and inorganic membranes. The highest separation factor was 202 for CO2/H2 separation at 235 K and 21 for CO2/CO separation at 258 K, significantly higher than that at room temperature. The observed CO2 flux was very high, i.e., 300–420 kg m-2 h-1, in the entire studied temperature range of 235–310 K.

Efficient light olefins/N2 separation technologies are of great interest to recover monomers from N2 purge gas in polymer plants. C3H6/N2 and C2H4/N2 separations were investigated using MFI zeolite membranes in a temperature range of 258–356 K. The membranes were rather selective towards the hydrocarbons. For C3H6/N2 separation, a maximum separation factor of 43 was observed at room temperature with a C3H6 permeance of 22×10-7 mol m-2 s-1 Pa-1. For C2H4/N2 separation, the maximum separation factor was 6 at 277 K with a C2H4 permeance of 57×10-7 mol m-2 s-1 Pa-1. The permeance and the selectivity for C3H6 were much higher than commercial polymeric membranes. For C2H4, the permeance was much higher with similar selectivity compared to commercial polymeric membranes. The findings reveal that zeolite membranes are promising candidates for light olefins/N2 separation in petrochemical processes.

The adsorption properties determine separation performance for systems studied in the present work. The high selectivity emanates from competitive adsorption, e.g., the strongly adsorbing components hinder the permeances of the weakly adsorbing ones and the effect was stronger at low temperature. In addition, gas permeances through zeolite membranes tend to decrease at low temperature most likely due to decreasing diffusivity, especially at cryogenic temperature. However, the permeances of our MFI membranes even at low temperature were still one to two orders of magnitude higher than those reported for inorganic and polymeric membranes. Thus, the high-flux MFI membranes have great superiority in this case. The fairly high permeance was ascribed to the ultra-thin (< 1µm) MFI film and highly permeable support used. The author provides here a promising candidate, ultra-thin zeolite MFI membranes, with high permeance and excellent selectivity for gas separation application at low temperature.

Page 14: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

xiii

CONTENTS

ACKNOWLEDGEMENTS ....................................................................................................... v

LIST OF PAPERS .................................................................................................................... vii

ABSTRACT .............................................................................................................................. xi

CONTENTS ............................................................................................................................ xiii

1. INTRODUCTION .............................................................................................................. 1

1.1. Zeolites ........................................................................................................................... 1

1.2. Membranes ..................................................................................................................... 2

1.2.1. Membrane Terminology ......................................................................................... 2

1.2.2. The Robeson Upper Bound ..................................................................................... 3

1.2.3. Membrane Transport Theory .................................................................................. 4

1.2.4. Separation Mechanisms in Porous Membranes ...................................................... 5

1.3. Zeolite Membranes ......................................................................................................... 5

1.3.1. A Good Zeolite Membrane ..................................................................................... 7

1.3.2. Gas Separation by Zeolite Membranes at Low Temperature ................................. 8

1.4. Separations of Interest for the Present Work .................................................................. 9

1.4.1. Air Separation ......................................................................................................... 9

1.4.2. N2/He Separation .................................................................................................... 9

1.4.3. CO2 Separation from Synthesis Gas ..................................................................... 10

1.4.4. Light Olefins/N2 Separation .................................................................................. 11

1.5. Scope of the Present Work ........................................................................................... 12

2. EXPERIMENTAL ............................................................................................................ 13

2.1. Membrane Preparation ................................................................................................. 13

2.1.1. Weakly Oriented MFI Membranes (Papers I, II, III, V and VI) ........................... 13

2.1.2. b-Oriented MFI Membranes (Papers IV and VII) ................................................ 13

2.2. Membrane Characterization ......................................................................................... 13

2.3. Single Gas Permeation .................................................................................................. 14

2.3.1. O2 and N2 (Papers I and II) ................................................................................... 14

2.3.2. N2 and He (Paper III) ............................................................................................ 14

Page 15: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

xiv

2.4. Mixed-gas Separation Measurements ........................................................................... 15

2.4.1. Air Separation (Papers I and II) ............................................................................ 15

2.4.2. N2/He Separation (Paper III) ................................................................................. 17

2.4.3. CO2 Separation from Synthesis Gas (Papers IV, V and VII) ............................... 17

2.4.4. Light Olefins/N2 Separation (Paper VI) ................................................................ 17

3. MODELING ..................................................................................................................... 19

3.1. Air Separation (Papers I and II) .................................................................................... 19

3.2. The Effects of Flow-through Defects, Concentration Polarization and Pressure Drop over the Support (Papers III, VII and VI) ............................................................................. 20

3.2.1. Transport through Zeolite Pores and Defects ....................................................... 20

3.2.2. Concentration Polarization .................................................................................... 20

3.2.3. Pressure Drop over the Membrane and Support ................................................... 21

4. RESULTS AND DISCUSSION ....................................................................................... 23

4.1. Membrane Characterization ......................................................................................... 23

4.1.1. Permporometry ..................................................................................................... 23

4.1.2. SEM Analysis ....................................................................................................... 24

4.1.3. XRD Analysis ....................................................................................................... 25

4.2. Single Gas Permeation .................................................................................................. 25

4.2.1. O2 and N2 Single Gas Permeation (Papers I and II) .............................................. 25

4.2.2. N2 and He Single Gas Permeation (Paper III) ...................................................... 27

4.3. Separation of Binary Mixtures at Low Temperature .................................................... 28

4.3.1. Air Separation at Cryogenic Temperature (Papers I and II) ................................. 28

4.3.2. N2/He Separation at Low Temperature (Paper III) ............................................... 31

4.3.3. CO2 Separation at Low Temperature (Papers IV, V and VII) .............................. 36

4.3.4. Light Olefins/N2 Separation at Low Temperature (Paper VI) .............................. 40

4.3.5. Concluding Remarks ............................................................................................. 44

5. CONCLUSIONS ............................................................................................................... 45

6. FUTURE WORK .............................................................................................................. 47

REFERENCES ......................................................................................................................... 49

Page 16: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

1

1. INTRODUCTION

1.1. Zeolites

Zeolites are crystalline materials with a uniform and well-defined pore network, which can be found in nature or produced artificially [1]. These materials are composed of silicon, aluminum and oxygen as well as non-framework cations. Different zeolites have different pore sizes ranging from 0.3 to 1.3 nm, depending on the framework type [2]. There are more than 200 different zeolite frameworks defined so far. The International Zeolite Association uses three capital letters to represent different zeolite frameworks, e.g., MFI, LTA and FAU [3]. The membranes used in this work comprise MFI zeolite. The MFI structure, with an average pore diameter of 0.55 nm has two types of intersecting pores. One is zigzag along the a-axis and the other one is straight along the b-axis, see Figure 1.1. As already mentioned, other zeolite structures may have different pore diameters, e.g., 0.38, 0.4 and 0.7 nm for CHA, DDR and FAU zeolites, respectively.

For synthetic zeolites, the Si/Al ratio is determined by the synthesis conditions and it affects the material polarity and adsorption properties. MFI zeolites with Si/Al ratios ranging from 10 to infinity [4] have two sub-groups, namely ZSM-5 with a Si/Al ratio of 10–200 and silicalite-1 with a Si/Al ratio of more than 200. In the present work, MFI zeolite membranes were prepared: ZSM-5 membranes with a Si/Al ratio of ca. 140 were evaluated in Papers I, II, III, V and VI whilst silicalite-1 membranes were evaluated in Papers IV and VII.

Figure 1.1. Schematic representation of the MFI zeolite [5].

Zeolites play rather an important part in industry, mainly for catalysis [1], gas separation [6] and ion exchange [7].

Page 17: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

2

1.2. Membranes

A membrane is “a structure, having lateral dimensions much greater than its thickness, through which mass transfer may occur under a variety of driving forces” [8]. Membranes have gained a tremendous interest for different applications, like separation, membrane reactors and controlled drug delivery. In the separation application, the goal is to allow one component in a mixture to permeate the membrane freely, while hindering the other components from permeating. The driving forces include pressure gradient, concentration gradient, temperature gradient or electrical potential gradient between the two membrane sides. Membranes can be porous or dense, and supported or non-supported. Membranes can be made of polymer, ceramic, metal and liquid. Thin membranes are desirable due to high transport rates. The mechanical strength of thin membranes is normally provided by rigid supports [9]. Typical materials for the substrates are alumina, amorphous silica, stainless steel, glass, porous ceramics, etc. [10].

Membrane technology is an attractive alternative to conventional separation processes due to low capital costs, simple and continuous operation, small space requirement and easy scalability. Moreover, membrane separation processes consume 6–10 times less energy than thermally driven separation processes [11]. For some special applications, for instance purification of thermally labile compounds, membranes possess unique superiority at low operating temperatures. The current industrial membrane separation processes include ultrafiltration, reverse osmosis, microfiltration, gas separation, pervaporation and ion exchange membrane process. A number of industrial plants for membrane gas separation have been installed during last 20 years [9] and the market is expanding rapidly [12], especially in food/beverage industry, the (petro) chemical industry and waste water treatment.

In applications, membranes are always assembled in membrane modules to fulfill industrial requirements, e.g., large membrane area, compactness and easy maintenance. There are several types of modules: plate-and-frame modules, tubular modules, spiral-wound modules and hollow fiber modules, etc. [9]. With designs of multi-stage membrane units or combination of membrane and other separation units, e.g., distillation [13] and crystallizer [14], it is often possible to reach very high purity for products at low cost. However, in many cases single stage membrane processes would be sufficient.

1.2.1. Membrane Terminology

Figure 1.2 illustrates a membrane separation process: a feed stream, comprised of several components, is applied to one side of the membrane. Part of the stream passes through the membrane to the other side, and it is referred to as permeate. The remaining part is named retentate, see Figure 1.2 (a). A stream of inert gas, called sweep gas, is often applied, especially in academic studies, on the permeate side to increase the gradient [9], see Figure 1.2 (b).

Page 18: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

3

Figure 1.2. A schematic diagram of the membrane separation process without (a) and with sweep gas (b).

The membrane flux Ji is defined as the amount of component i permeating the membrane per unit time and per membrane area (kg m-2 h-1 or mol m-2 s-1). The flux divided by the driving force is referred to as permeance Πi (mol m-2 s-1 Pa-1):

Πi=Ji/ΔPi=Fi/(AfilmΔPi), (1.1)

where Fi is the molar flow rate of component i through the membrane, Afilm is the membrane area, and ΔPi is the partial pressure difference of component i across the membrane.

Permeability (Barrer) is a commonly used term for polymeric membranes and equals permeance multiplied by the membrane thickness.

The membrane selectivity (also referred to as the permeance-based selectivity) is determined as

αi/j=Πi/Πj, (1.2)

where Πi and Πj are the permeances of species i and j, respectively.

The separation factor is calculated as follows:

βi/j=(yi/yj)/(xi/xj), (1.3)

where yi and xi are the molar fractions of species i in the permeate and feed streams, respectively.

High selectivity is desired to achieve high purity of the product, whilst high permeability is favored to minimize membrane area and thus capital cost.

1.2.2. The Robeson Upper Bound

Robeson plotted reported permeances and selectivities for a large number (ca. 300) of polymeric membranes of various types and illustrated that there was a trade-off between selectivity and permeance. The best polymeric membranes with high selectivity had relatively low permeability, and vice versa. Polymers with the best combinations of selectivity and

Page 19: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

4

Figure 1.3. Robeson selectivity-permeability upper bound for a number of commercially important gas separations [15].

permeability constitute what is known as the Robeson selectivity-permeability upper bound [16] and it was updated in 2008 [17], see Figure 1.3. The “upper bound line” is empirical and the Robeson upper bound is widely cited and referred to when researchers compare membrane performances for gas separation. It is worth pointing out that the most permeable materials possessing sufficient selectivity, are of major interest for industrial applications [9]. In other words, a highly selective membrane with too low permeability is not industrially applicable.

1.2.3. Membrane Transport Theory

Molecular transport through membranes can be described by the solution-diffusion model and the pore-flow model [9]. The solution-diffusion model describes the transport as: the permeate is first dissolved (absorbed) in the membrane and then transported through the membrane, which is caused by a concentration gradient. The flux can be determined by Fick´s Law:

Ji=-Di (dci/dx), (1.4)

where Ji is the flux of component i; dci/dx is the concentration gradient of the component i; Di is the diffusion coefficient of component i.

The solution-diffusion model can be used for reverse osmosis, pervaporation and gas permeation in dense polymeric membranes.

The pore-flow model is normally applied to describe molecular transport through porous membranes. There are four principal types of diffusion through pores: molecular diffusion, Knudsen diffusion, surface diffusion and configurational diffusion, depending on the sizes of the pores and molecules as well as the operating conditions:

Page 20: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

5

1. Molecular diffusion prevails in large pores, where the diffusant molecules collide with each other rather than with the pore walls. It is governed by viscous flow if the transport is driven by a pressure gradient in a macroporous medium.

2. Knudsen diffusion takes place when the mean free path of the diffusant molecules is larger than the pore size and the molecules collide with the pore walls.

3. Surface diffusion normally prevails in mircropores. The molecules move along the pore walls by hopping from one adsorption site to another.

4. Configurational diffusion dominates when the diffusant molecules have similar size as the pore size.

1.2.4. Separation Mechanisms in Porous Membranes

An intrinsic property of membranes for separation is to have different permeation rates for different species. In a gas separation process using porous membranes, e.g., zeolite membranes, the mechanism can be classified as [18]: preferential adsorption, preferential diffusion and molecular sieving (/size exclusion). These mechanisms could coexist in a separation process; however, one of the mechanisms will most likely dominate. The separation mechanism can also be strongly affected by the operating conditions, e.g., temperature and partial pressure. Generally speaking, the difference in adsorption is more significant at low temperatures whilst diffusion selectivity plays a more important role at high temperatures. Figure 1.4 illustrates temperature effects on separation for gases smaller than membrane pores. At low temperatures, the strongly adsorbing molecules in the mixture feed inhibite the permeance of weakly adsorbing gas. Thus adsorption properties dominate separation performance. At high temperatures, adsorption diminishes whilst diffusion selectivity becomes more important, and thus preferential diffusion turns to be the dominant separation mechanism.

Moreover, the membrane materials normally contain a range of pore sizes along with a certain number of defects, which makes it more complex to address a precise separation mechanism.

Figure 1.4. A schematic clarification of temperature effects on membrane separation mechanisms [19].

1.3. Zeolite Membranes

Zeolite membranes are one common group of inorganic, porous membranes comprised of inter-grown zeolite crystals. These materials can be used for gas [20] or liquid separations [21], as catalysts in membrane reactors [22,23], and as sensors. Zeolite membranes are

Page 21: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

6

regarded as an attractive alternative for gas separations due to the potentially high fluxes as well as high thermal, chemical and mechanical stability. These membranes could potentially be applied for separations at severe conditions where polymeric membranes are not suitable. With the well-defined pore structure, these membranes can also offer high selectivity. One industrial-scale application of zeolite membranes is dehydration of ethanol and i-propanol [24-26]. Although more than 200 different zeolite frameworks have been found, only a few of them have been fabricated as membranes, e.g., MFI [27-29], LTA [24,25], FAU [30,31], CHA [32,33], DDR [34,35] and SOD [36,37]. Even so, most of the membranes possess low permeability due to the high film thickness or/and moderate selectivity due to the membrane defects.

MFI membranes have the possibility to efficiently separate small gas molecules based on adsorption selectivity instead of molecular sieving [38] since the average pore size of MFI zeolite (5.5 Å) is larger than the kinetic diameters of those light gas molecules, see Table 1.1. In contrast, DDR membranes with a window opening size of 3.6 × 4.4 Å can separate small gas molecules based on molecular sieving. A CO2/CH4 selectivity of 100–3000 by DDR membranes was reported [39]. Similar high selectivity can also achieved by CHA (3.8 Å) membranes [33]. However, the main drawback for small pore membranes (e.g., DDR and CHA membranes) is the low membrane permeance by current manufacture methods.

The membrane selectivity is affected by the adsorption properties. It can be enhanced by altering Si/Al ratio of the zeolite or by introducing foreign chemical groups [40]. In addition, flow-through defects in the membrane, e.g., in the form of open grain boundaries, can affect the membrane performance. For high flux membranes, the membrane performance may also be affected by both external mass transfer limitations (concentration polarization) and pressure drop over the support. The former may lead to a depletion of the preferentially permeating species at the membrane/fluid interface. The pressure drop over the support leads to an increased partial pressure of the permeating species on the permeate side of the selective film. Thus driving force decreased resulting in a reduction in flux and sometimes also in selectivity.

Table 1.1. Kinetic diameter [38] and heat of adsorption on silicalite* [6] of some gas and vapor molecules.

Molecule H2O He H2 CO2 O2 N2 CO CH4 C2H4 C3H6 SF6

Kinetic

Diameter (Å) 2.65 2.55 2.89 3.30 3.47 3.64 3.69 3.76 4.16 4.68 5.50

–ΔHads

(kJ mol-1) - - 5.9 27.2 16.3 17.5 16.7 20.9 26.6 39.2 34.3

* The values of heat of adsorption can slightly vary in difference references.

Page 22: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

7

Synthesis techniques to prepare zeolite films on supports can be classified into two main groups: in-situ synthesis and secondary growth. The main advantage for the former method is that the preparation steps can be minimized since supports are just placed in synthesis solutions where crystals nucleate and grow into a film. The secondary growth method involves seeds preparation, seeds deposition on supports and film growth by hydrothermal treatment. This approach prevails in recent publications (including the present work), because it owns significant advantages such as high reproducibility and good control over membrane thickness and orientation [38].

1.3.1. A Good Zeolite Membrane

In order to be an industrially competitive separation process, membranes should be cost-effective, and possess high flux with sufficient selectivity, preferably with high thermal and chemical stability. Polymeric membranes are used in a number of large-scale industrial processes. In contrast, zeolite membranes are potentially more stable, but have not been widely applied in commercial-scale to the present.

To provide sufficient selectivity, a zeolite membrane should essentially be free of defects, i.e., any pores larger than the zeolite pores. The defect-free zeolite membranes contain only zeolite pores inside which the separations take place due to the interplay of the mixture adsorption equilibrium and diffusion [41]. It is however still challenging to make defect-free zeolite membranes [42].

High flux/permeability is essential for a large scale application as it results in lower membrane area to suit the same separation task, which will in turn partly trim the investment cost. In fact, only 10–20% of the total costs attribute to the membrane itself whilst most to modules, for both polymeric and inorganic membrane processes [26]. In addition, expensive supports also contribute to the cost of zeolite membranes to a great extent. For high flux zeolite membranes, the required supports and modules are reduced so as to be cost-competitive for industrial application. The membrane thickness should be as low as possible with the aim of a high permeability. Thin zeolite crystal layer (ca. 1µm) supported on substrates is considered promising for future applications [30]. However, most of the prepared zeolite membranes reported in the literature have a film thickness exceeding 1 μm, resulting in very low flux, and therefore large membrane area needed [30].

Rigid substrates are often utilized to provide mechanical strength for zeolite films. An ideal substrate should provide mechanical support without creating additional mass transport resistance.

Our research group has developed membranes comprised of thin (<1 μm) zeolite films on open graded alumina supports, with both high flux and high selectivity [43]. These membranes have been tested for various gas separations [28,43-47] and liquid separations [48,49], displaying 10–100 times higher permeances than most of membranes previously reported by other groups, with still similar separation factors. For instance, the ultra-thin membranes exhibited a p/o-xylene separation factor of 16 with a high p-xylene permeance of 3 × 10-7 mol m-2 s-1 Pa-1 at 663 K [43]. The outstanding performance of our membranes owes

Page 23: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

8

Figure 1.5. SEM images of membranes prepared with (a, b) and without (c, d) masking technique using the same synthesis conditions [50].

to the ultra-thin film, low defect density and low support resistance [28].The masking technique is essential to produce high quality membranes in our group, by improving both permeance and selectivity. All the support pores are filled with wax before the film grows on the top surface of the support, to reduce harmful invasion of synthesis mixture and leaching from the support [50]. Calcination is implemented to remove the wax and polymer at the end. As it is shown in Figure 1.5, the masked membrane presented a thinner film and less cracks than unmasked membrane. Moreover, the masking technique resulted less deposits in the supports, and thus could considerably improve permeance.

1.3.2. Gas Separation by Zeolite Membranes at Low Temperature

As mentioned above, zeolite membranes possess high thermal stability. They can potentially be used at both very high and very low temperatures. High temperature applications of zeolite membranes have been studied extensively [51]. On the contrary, scientific reports on gas separation using zeolite membranes at low temperatures are rather scarce. Hong et al. [32] studied CO2/H2 separation using SAPO-34 membranes in the temperature range of 253–308 K. At 253 K, the reported CO2/H2 separation factor was as high as 110, however the CO2 flux was only 3 kg m-2 h-1 even though a high trans-membrane pressure difference of 12 bar was applied. Piera et al. [52] separated CO/air at an extremely low CO concentration of 160 ppbv using an MFI-type membrane. The separation factor was 3.1 with an extraordinarily low CO

Page 24: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

9

flux of 1.3 × 10-7 kg m-2 h-1 at 245 K. Van den Bergh et al. [39] investigated permeation of various gases (CO2, CH4, N2, O2) and the mixtures separation, using DDR membranes in the temperature range of 220–373 K. The mixture selectivities increased as temperature was decreased, especially for CO2/CH4, CO2/air and N2O/air mixtures. The O2/N2 selectivity was low amounting to only ca. 2 and the CO2/CH4 separation factor was higher than 1000 at a temperature below 250 K. However, the flux was reported as “very low”, i.e., lower than 7.8 × 10-2 kg m-2 h-1 in all cases. In comparison, the MFI membranes made in our group displayed excellent performance for CO2 separation at high pressure and fairly low temperature [28]. The CO2/H2 separation factor was as high as 32 with a CO2 flux of 332 kg m-2 h-1 at 275 K, at a feed pressure of 10.0 bar∗1 and a permeate pressure of 2.0 bar.

1.4. Separations of Interest for the Present Work

1.4.1. Air Separation

The current air separation technologies include cryogenic distillation, pressure swing adsorption (PSA), chemical absorption and membrane separation [53]. Cryogenic distillation is a highly developed and well-established technique for large-scale production (typically larger than 10 MMscfd∗2 of gas) of pure oxygen. However, the technique is energy-intensive with high investment costs. PSA processes are usually applied in smaller scales (typically 1–10MMscfd), when it is sufficient with lower purity, typically around 95 vol.% oxygen. Chemical absorption process could be rather a competitive alternative if the facility corrosion could be reduced. Polymeric membranes have been developed for oxygen separation from air. Materials that have been used for air separation include poly(1-trimethylsilyl-1-propyne) (PTMSP), Teflon AF 2400, silicone rubber, poly(4-methyl-1-pentene) (TPX), poly(phenylene oxide) (PPO), ethyl cellulose, 6FDA-DAF (polyimde), polysulfone, polyaramide, etc [9]. Most of these membranes exhibit high selectivity but low permeability, i.e., low flux. In addition, most of these membranes have poor durability. Carta et al. [54] have recently reported in Science microporous polymeric membranes with a rather high O2 permeability of 900–1100 Barrer and a good O2/N2 selectivity of 4.5–6.1.

Oxygen-enriched air is one of the greatest markets where membrane technology is competitive. There is a large demand from chemical industry, e.g., oxidation process, steel industry, and medical use.

1.4.2. N2/He Separation

There is a large and increasing demand of helium in the world, especially from aerospace and cryogenic technologies. However, the terrestrial abundance of helium is scarce and the

∗1 The pressures presented in the thesis are all absolute pressures. ∗2 MMscfd is an abbreviation of million standard cubic feet per day, predominantly used in the United States, as a measure of natural gas, liquefied petroleum gas, and other gases that extracted, processed or transported in high quantities. 1 MMscfd = 1180 m3 h-1.

Page 25: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

10

shortage hampers the development of both research and industry [55,56]. Natural gas, as the current major resource of helium, typically contains 0–4% of helium, and only 1/10 of the natural gas fields possess economically feasible concentrations of helium, higher than 0.4%. It is of tremendous importance to develop an efficient method for helium separation as this would open up an opportunity for economically viable recovery of helium from natural gas with lower concentration of helium. Besides methane and other hydrocarbons, natural gas contains small amounts of nitrogen, water vapor, carbon dioxide and helium. The composition of natural gas varies greatly depending on the source [57]. Nitrogen is the most difficult component to separate from helium when recovering helium from natural gas due to the need of deep cooling prior to distillation [58]. Another application calling for N2/He separation is the reclamation of spent helium gas from airships and deep-sea diving systems [59,60].

The current well-established technique for large-scale production of commercial helium (> 99.99%) is based on energy-intensive cryogenic distillation and other multi-stage unit operations that require both high investment and operating costs. Only a few exploratory reports on helium purification via pressure swing adsorption using adsorbents, e.g., activated carbon, are available [61,62]. Membrane processes with compact equipment, low capital investment and low energy consumption appear to be an attractive alternative for helium recovery [59,63]. Various polymeric membranes have been investigated for He/N2 separation, but nearly all membranes showing high helium selectivity display low permeability [16,17,64,65], requiring large membrane areas for the separation task. Many types of inorganic membranes, e.g., porous alumina [66], silica [66], zeolite [67] and MOF [68] membranes, have also been evaluated for He/N2 separation, mainly at ambient temperature. The reported ideal separation selectivities were low and close to the Knudsen selectivity of 2.6 [59]. Only Koresh et al. [69] reported separation selectivity much higher than Knudsen selectivity, i.e., an ideal separation selectivity of 20–40 for molecular sieve carbon membranes with a pore diameter of 0.3–0.5 nm. However, no mixture separation selectivities have been reported, which could deviate strongly from ideal separation selectivities based on single gas permeation.

1.4.3. CO2 Separation from Synthesis Gas

CO2 is not only a greenhouse gas, but also an undesirable impurity in natural gas, biogas, synthesis gas, etc. [70]. It is often necessary to remove CO2 or reduce the CO2 concentration in natural gas because it lowers the calorific value, causes corrosion problems, and increases the transport cost. Synthesis gas contains mainly CO, CO2, and H2. Proportions of these components vary, depending on the raw materials and production process. The ideal synthesis gas stoichiometry to produce methanol is a molar ratio (H2–CO2)/(CO+CO2) of 2 [71]. Synthesis gas generated by gasification of coal or biomass contains too much CO2 from this point of view, so it is necessary to remove some of it prior to the synthesis step, to achieve a desirable composition of the gas mixture. In addition, efficient CO2 separation from syngas is also needed for production of pure H2, which can be used for generation of electricity and chemicals, etc. [72]. Therefore, CO2/H2 and CO2/CO separation were carried out in the present work.

Page 26: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

11

There are several technologies for CO2 separation and capture currently including absorption, adsorption, cryogenic distillation, gas hydrate formation, chemical looping and membranes [70]. Absorption, e.g., amine scrubbing is the most developed technology and widely adapted in chemical and petroleum industries. However, it is energy-intensive with high capital cost and corrosion problems. Membrane technology has shown great potential in pre-combustion capture (CO2/H2 separation) and in post-combustion capture (CO2/N2 separation) [72]. Polymeric membranes, e.g., cellulose acetate membranes have been successfully applied to upgrade natural gas [73], but exhibit several problems, e.g., plasticization at high pressure, swelling and low contaminant resistance. The poly(ethylene oxide) copolymers with CO2/H2 selectivities of 10–12 and a CO2 permeance of ca. 2×10−7 mol m−2 s−1 Pa−1 are the best polymeric membranes developed to date [15,74].

There are only a few research papers available in the literature on CO2 separation from H2 by zeolite membranes. Most of the reported membranes possess quite low CO2 permeance and would be too costly to apply in industry. The specific data on the performance of different zeolite membranes are summarized in the Discussion Section. Our high flux MFI membranes have also been evaluated for CO2/H2 separation at elevated pressures [28]. The separation factor was found to increase with decreasing temperature, within the investigated temperature range of 275 K–396 K. Resting on these findings, it is interesting to explore the membrane behavior for CO2/H2 separation at even lower temperatures.

CO2/CO separation by membrane technology has scarcely been reported. However, Lin et al. [75] evaluated CO2 removal from synthesis gas using polymeric PolarisTM membranes, from laboratory-scale and pilot-scale to commercial-scale. The pilot-scale membrane showed a CO2/CO selectivity of 10–20 and a CO2 permeance of (0.34–1) ×10−7 mol m−2 s−1 Pa−1 at temperatures ranging from 268 to 298 K. Liu et al. [76] reported a CO2/CO selectivity of 5.0 with a CO2 permeance of 1×10−7 mol m−2 s−1 Pa−1 by ZIF-69 membranes at 298 K. In particular, there are hardly any reports in the open literature on the use of zeolite membranes for this separation. Sandström et al. [28] from our group tested MFI membranes for the separation of an equimolar mixture of CO2/H2/CO. The membrane was rather CO2 selective (CO2/CO selectivity ≈10) with a high CO2 permeance of 75×10−7 mol m−2 s−1 Pa−1 at 277 K and a feed pressure of 9.0 bar.

1.4.4. Light Olefins/N2 Separation

Recovery of light hydrocarbons from petrochemical process streams, for instance the purge gas from a polypropylene or polyethylene plant resin degasser, is an interesting application. A typical degas stream from a polypropylene plant contains 20–30 vol.% C3+ hydrocarbons in nitrogen. It would be of great value for the industry to recycle the hydrocarbons instead of burning them [15,77,78].

Membrane separations are particularly appealing for vapor/gas separation. From organic to inorganic, various membrane materials have been evaluated. Silicone rubber is by far the most widely used membrane material for recovering hydrocarbons from the vent streams, as summarized by Baker et al. [15]. Those membranes are well-suited for separation of organic vapor/gas mixtures and the commercialization of this type of membrane continue to stimulate

Page 27: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

12

membrane invention. However, silicone rubber membranes are quite thick (typically ca. 50–100 µm) and exhibit rather low permeances. Besides, the membranes possess low stability. Freshly made thin composite polymeric membranes often lose 50% of the permeance within two weeks for this application. Paul et al. [79] have shown that the performance deterioration is caused by the reordering of the polymer chains in the membrane, thus reducing the number and size of free volume elements that contribute to gas permeation.

1.5. Scope of the Present Work

The aim of the present work was to investigate gas separations using zeolite membrane at low temperature and even cryogenic temperatures because reports at low temperature are scarce and there are no reports on cryogenic temperatures. Ultra-thin MFI zeolite membranes were for the first time evaluated for the following separations under such low temperatures:

I. O2 separation from air. This system was studied in the cryogenic temperature range of 62–110K. The feed and permeate pressures were varied to identify the optimum operating conditions. The adsorption of N2 and O2 was modeled to gain insight into the experimental observation.

II. N2/He separation. The separation was investigated over a temperature range of 85–260 K at different feed and permeate pressures. Modeling was performed to evaluate the influences of membrane defects, concentration polarization and pressure drop over the support.

III. CO2 separation from synthesis gas. CO2/H2 and CO2/CO separation were studied at much lower temperatures than reported before, i.e., down to 235 K, which is not too far from the sublimation temperature of CO2. Weakly oriented membranes were evaluated for the separation of CO2/H2. Novel b-oriented MFI zeolite membranes were developed and evaluated for the separation of CO2 from H2 and CO. Modeling was performed to elucidate the experimental findings and to evaluate the effect of concentration polarization and pressure drop over the support.

IV. Light olefins/N2 separation. C3H6/N2 and C2H4/N2 separation were investigated using zeolite membranes in range of temperatures down to 258 K. The effect of the concentration polarization and pressure drop over the support were investigated by modeling.

Page 28: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

13

2. EXPERIMENTAL

2.1. Membrane Preparation

2.1.1. Weakly Oriented MFI Membranes (Papers I, II, III, V and VI)

Weakly oriented MFI membranes were prepared on a graded porous α-alumina disc (Fraunhofer IKTS, Germany), using a previously reported method [43]. The support has a diameter of 25 mm and is comprised of two layers: a 30 μm thick top layer with a pore size of 100 nm and a 3 mm thick base layer with a pore size of 3 μm. The alumina disc was masked and then seeded with a monolayer of MFI crystals of 50 nm in size as described previously [50]. After that, a hydrothermal treatment was carried out for the MFI film growth. The seeded support was immersed in a synthesis solution with a molar composition of 3TPAOH:25SiO2:1450H2O:100C2H5OH in a polypropylene tube and was kept in the solution at 361 K for 72 h with reflux of evaporated solution. After synthesis, the membrane was rinsed in a 0.1 M NH3 solution overnight. Calcination was performed at 773 K for 6 h at a heating rate of 0.2 K min-1 and a cooling rate of 0.3 K min-1.

2.1.2. b-Oriented MFI Membranes (Papers IV and VII)

A brief description of the preparation procedure of b-oriented MFI membranes will be given here. More details can be found elsewhere [80]. The seeds were synthesized from a mixture of 5TEOS:1TPAOH:500H2O. The solution was shaken at room temperature for 7 days, and then stirred at 403 K for 9 h. The crystals obtained were purified by repetitive centrifugation and re-dispersion in distilled water and were freeze-dried before use. The MFI crystals with a size of 500×450×200 nm3 could be organized as a b-oriented monolayer on the graded porous α-alumina support by dynamic interfacial assembly [81]. The support was immersed in a synthesis solution with a molar composition of 3TPAF:25SiO2:1500H2O for hydrothermal treatment at 373 K  under reflux for 48 h. After that, the membrane was rinsed with a 0.2 M NH3 solution and then dried at 323 K overnight. Calcination was performed at 773 K for 6 h, with a heating rate of 0.2  K min−1 and a cooling rate of 0.3  K min−1.

2.2. Membrane Characterization

Adsorption-branch n-hexane/helium permporometry [47] was used to characterize the amount of flow-through defects in the membranes. This technique has been demonstrated to be an effective and reliable tool for evaluating the amount of micropore defects, down to 0.7 nm, [82,83] and mesopore defects [47] in porous membranes. The results were consistent with Scanning Electron Microscopy (SEM) analysis and mixed-gas separation data. The procedure for the permporometry characterization is described in detail in a previous work [82] and only a brief description is given below. The membrane was mounted in a stainless steel cell and

Page 29: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

14

sealed with graphite gaskets (Eriks, the Netherlands). It was dried at 573 K for 6 h prior to the test and the temperature was set to 323 K during the measurement. The feed pressure was kept at 2.0 bar and the permeate pressure was kept at 1.0 bar. The permeance of the non-adsorbing component helium was measured at a series of relative pressures (Pi/Pi0) of a strongly adsorbing component n-hexane. The relative pressure of n-hexane was raised in a step-wise manner from 0 to ca. 0.4. At each relative pressure, the system was allowed to achieve steady-state. In the absence of n-hexane (Pi/Pi0=0), helium can permeate through all pores, i.e., zeolite pores and defects, resulting in high initial helium permeance. Once introduced, n-hexane first blocked small pores, i.e., zeolite pores, and then larger and larger pores, i.e., defects. Hence, as the concentration of n-hexane in the feed stream increased, the helium permeance decreased. The desired Pi/Pi0 values were achieved by altering the flow rate of helium and the temperature of the saturators filled with n-hexane. The defect width was calculated from n-hexane relative pressure using either the Horvàth–Kavazoe equation (micropore defects) or the Kelvin equation (mesopore defects).

The morphology of the membranes was characterized by SEM using an FEI Magellan 400 field emission XHR-SEM without any coating of the samples.

X-ray diffraction (XRD) characterization of the membranes was performed using a PANalytical Empyrean diffractometer equipped with a Cu LFF HR X-ray tube and a PIXcel3D detector. The data was evaluated using the HighScore Plus 3.0.4 software.

2.3. Single Gas Permeation

To better understand the separation mechanism, single gas permeation experiments were carried out. The MFI membrane was kept in the same steel cell as used for permporometry analysis. Prior to the permeation experiments, the cell with the membrane inside was mounted in a cryostat and the membrane was dried in a helium flow at 573 K for 6 h at a heating rate of ca. 2 K min-1. The temperature was monitored by a type T thermocouple on the feed side.

2.3.1. O2 and N2 (Papers I and II)

A stream of 500 ml/min of O2 (> 99.999%, AGA) or N2 (> 99.999%, AGA) was fed to the membrane at 1.0 bar while the permeate side was kept at 100 mbar (Paper II). In the work described in Paper I, similar conditions were used but the permeate side was kept at 1.0 bar with a stream of 450 ml/min helium (> 99.999%, AGA) as sweep gas.

2.3.2. N2 and He (Paper III)

N2 and He single gas permeation experiments (experiments E4 and E5) were carried out over a temperature range of 58–121 K, see Table 2.1. The feed pressure was kept at 5.0 bar while the permeate side was kept at 0.5 bar, as in E3.

Page 30: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

15

Table 2.1. Feed and permeate pressures used in the binary N2/He mixture separation and single gas permeation experiments.

Experimental Feed Permeate Pressure ratio N2 in the feed He in the feed Feed flowrate

code (bar) (bar) φ (%) (%) (ml min-1)

E1 3.0 1.0 3 52.6 47.4 420

E2 5.0 1.0 5 52.6 47.4 1300

E3 5.0 0.5 10 52.6 47.4 1300

E4 5.0 0.5 10 100 0 1236

E5 5.0 0.5 10 0 100 1630

2.4. Mixed-gas Separation Measurements

The separation experiments were carried out using the experimental set-up illustrated in Figure 2.1. The membrane (kept in the same steel cell) was dried in a helium flow at 573 K for 6 h at a heating rate of 1 K min-1, prior to the separation test. A column packed with dried 5 A zeolite pellets was used to adsorb any traces of moisture in the feed gas. The gas flow rate was controlled by mass flow controllers (Bronkhorst, F-201CV). The feed or permeate pressure could be set at a certain value, either above or below 1 atm by choosing the back pressure regulators (Swagelok, KCB) or vacuum pump system, respectively. Vacuum pumps (Pfeiffer Vacuum, MVP 020 and MVP 040) installed in the retentate and permeate lines reduced the pressures on both sides of the membrane. Pressures below 1 atm were controlled by dosing valves (Pfeiffer Vacuum, EVN 116). The permeate composition was analyzed by a mass spectrometer (Inprocess Instruments, GAM 400) every 5 seconds. The custom-built cryostat (ICEoxford UK, DRYICE25K) could set separation temperature at any value ranging from 22 K to 573 K with a deviation of less than 0.05 K.

2.4.1. Air Separation (Papers I and II)

A stream of synthetic air (purity > 99.999%, AGA), consisting of 79.2% N2 and 20.8% O2, was fed to the membrane. The membrane cell was kept in thermostat at low temperature: a Dewar (KGW, Germany) filled with liquid nitrogen (Paper I) or a cryostat (ICEoxford, DRYICE25K) (Paper II). The permeate flow was measured by a drum type gas meter (Ritter Apparatehau GmbH) (Paper I) or by a digital mass flow meter (Bronkhorst, F-201CV) (Paper II).

Page 31: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

16

(a)

(b)

Figure 2.1. A schematic diagram (a) and a photograph (b) of the set-up for gas permeation and separation experiments at cryogenic temperature.

Page 32: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

17

Paper I: The feed flow rate of synthetic air was 1500 ml/min and the sweep gas flow rate was 450 ml/min. The feed pressures ranged from 1.0 bar to 5.0 bar (no vacuum pumps had been installed when this work was carried out) and the permeate was kept at a constant pressure of 1.0 bar. The membrane cell was kept in a Dewar and liquid nitrogen was slowly added to the Dewar until the membrane temperature reached 77 K. After that, the temperature in the cell increased very slowly (less than 0.2 K/min) when the Dewar was heated by the ambient air. No cryostat was installed when this work was carried out.

Paper II: The feed flow rate of synthetic air was 500 ml/min with a feed pressure of 100 mbar, 400 mbar, 660 mbar and 1.0 bar (at a feed pressure of 100 mbar, the feed flow rate was reduced to 100 ml/min due to the limited capacity of the vacuum pump). A pressure of 10 mbar, 40 mbar, 66 mbar, or 100 mbar, was applied on the permeate side of the membrane. Since the pressure ratio, defined as total feed pressure divided by permeate pressure φ= PFeed/PPerm, may affect the separation performance, a constant pressure ratio of 10 was set in all experiments. No sweep gas was applied. A pressure transmitter (Pfeiffer Vacuum, CMR 362) was used to measure the permeate pressure lower than 100 mbar.

2.4.2. N2/He Separation (Paper III)

During the separation experiments, the gas flow rates of nitrogen and helium were controlled by mass flow controllers (Bronkhorst, F-201CV) and mixed dynamically after going through the adsorption columns. The feed mixture was then introduced to the membrane cell inside the cryostat. The feed pressure was controlled by a back pressure regulator. The permeate pressure was controlled by a dosing valve connected to a vacuum pump. The permeate flow was measured with a mass flow meter (Bronkhorst, F-201CV) and composition was analyzed online by a mass spectrometer.

The feed and permeate pressures used in all mixture separation experiments (experiments denoted E1-E3) are shown in Table 2.1. Mixtures consisting of 52.6 % N2 (>99.999%, AGA) and 47.4 % He (>99.999%, AGA), were fed to the membrane at a total feed pressure of 3.0–5.0 bar. The permeate pressure was varied from 0.5 to 1.0 bar. The pressure ratios, were 3, 5 and 10 for E 1, E2 and E3, respectively.

2.4.3. CO2 Separation from Synthesis Gas (Papers IV, V and VII)

A similar procedure was used for CO2 separation from synthesis gas as for air separation and N2/He separation. CO2/H2 (1:1) or CO2/CO (1:1) was mixed dynamically by two mass flow controllers with a total feed flow rate of 10 l min-1 at a feed pressure of 9.0 bar. Both CO2 and H2 had a purity of 99.999% and CO had a purity of 99.97%. The permeate side was kept at 1.0 bar and no sweep gas was used. The membrane cell was set in a thermostat (VWR, 89203). The permeate volumetric flow was measured by a drum type gas meter (Ritter Apparatehau GmbH) and the composition was analyzed by a mass spectrometer.

2.4.4. Light Olefins/N2 Separation (Paper VI)

The hydrocarbon/N2 mixtures (10 l min-1) were fed to the membrane at a feed pressure of 10.0 bar and a permeate pressure of 1.0 bar. The gases were purchased from AGA gas company

Page 33: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

18

with the purity of N2 >99.999% and hydrocarbons >99.95%. No sweep gas was used. Sub-ambient membrane temperatures were achieved by setting the membrane cell in a thermostat. Gas flow in the permeate stream was measured using a drum type gas meter, and the composition of feed and permeate streams was analyzed by GC (490 Micro GC, Agilent).

Page 34: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

19

3. MODELING∗

3.1. Air Separation (Papers I and II)

In order to gain some insight into the experimental observations, the adsorption of N2 and O2 was modeled. It may be expected that the membrane would be N2 adsorption selective due to the larger heat of adsorption and Langmuir adsorption parameters than those for O2, see Table 3.1. However, it is energetically favorable to adsorb smaller molecules at high loadings since smaller molecules pack more efficiently in the pores at high loadings [84]. To illustrate this, adsorption isotherms for O2 and N2 at 67 K and 90 K were simulated using the multisite Langmuir model [85]. The pure component adsorption isotherm is given by [86]:

( )nnbp

θθ−

=1

, (3.1)

where p is the pressure of the adsorbate in gas phase, θ is the fractional coverage, b is the Langmuir adsorption parameter, n is the number of adsorption sites one adsorbate molecule occupies. The multicomponent version of the multisite Langmuir model is given by [86]

in

j

N

jiiii pybn

−= Σ

=

θθ1

1 , (3.2)

where yi is the gas mole fraction of species i.

Table 3.1. Heat of adsorption and Langmuir adsorption parameters for N2 and O2. [87]

–ΔHads (kJ mol-1) b (bar-1) at 303 K

N2 17.0 0.277

O2 16.3 0.147

∗ The modeling work was not carried out by the author. However, a brief summary is presented in the thesis to help readers understanding the experimental results. More details could be found in the appended papers.

Page 35: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

20

3.2. The Effects of Flow-through Defects, Concentration Polarization and Pressure Drop over the Support (Papers III, VII and VI)

For high flux membranes, the membrane performance may be affected by flow-through defects, external concentration polarization and pressure drop over the support. To elucidate to which extent the experimental results were affected, modeling of mass transfer through the membrane was carried out. The model was not fitted to experimental data.

3.2.1. Transport through Zeolite Pores and Defects

Transport through the zeolite layer was assumed to occur both via transport through zeolite pores and via flow-through defects [82,88,89]. The transport through the zeolite pores was described using the Maxwell-Stefan formalism [90-92]. For a binary mixture, the permeating flux for species 1 through the zeolite pores is given by

( )

++

∆+∆+=

12121221

,21221,112211,1 //1

//1DDxDDx

qDDxqDDxlD

AAJ adsads

z

z

tot

zz

ρ (3.3)

where Az/Atot is the relative area of the membrane surface that is composed of zeolite (close to 1). The small remaining area is composed of flow-through defects. The defect distribution was obtained from permporometry data, see Table 4.1. Further, zρ and zl are the zeolite density and film thickness, respectively. Di is the Maxwell-Stefan diffusivity which in general depends on the adsorbed loading. The adsorbed loadings were determined by the ideal adsorbed solution theory (IAST) using single component Langmuir parameters. xi is the adsorbed phase mole fraction of species i and Δqi, ads is the difference in adsorbed loading between the feed and permeate side of the zeolite film.

Transport via flow-through defects, was calculated using Fick’s law, following the procedure reported previously [82,89]. The total fluxes through defects were then calculated by summing up the fluxes through each defect size interval:

=

nndefi

tot

ntotdefi J

AAJ ,,,, , (3.4)

3.2.2. Concentration Polarization

The concentration polarization index was determined as [93]:

[ ])/exp(1)/exp( cvb

pcv

b

m kJnn

kJnn

−+= , (3.5)

where nm is the mole fraction at the zeolite film - gas interface, nb is the mole fraction in the gas bulk, kc is the mass transfer coefficient, Jv is the volumetric flux and np is the mole fraction in the permeate.

Page 36: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

21

3.2.3. Pressure Drop over the Membrane and Support

Figure 3.1 shows a schematic illustration of a pressure drop profile over a composite membrane and gas film. The flux through the small pore top layer (L1) of the support was modeled as occurring via combination of viscous flow and Knudsen diffusion:

( )1

13,2,1

1

1,1,,1, 2 L

LiieffL

L

Lieff

LiKnLi l

PRT

PPBlP

RTD

J ∆

++

∆=

η, (3.6)

The flux through the coarse pores of the base layer L2 was modeled as occurring only by viscous flow:

( )2

24,3,22, 2 L

LiieffL

Li lP

RTPPB

J ∆

+=

η. (3.7)

In these equations, effjB is the effective permeability for layer j, η is the viscosity of the gas

and ΔPj is the partial pressure difference over layer j.

Figure 3.1. Schematic representation of the composite zeolite membrane and pressure profile

over the gas film (G), zeolite film (Z) and support layers 1 (L1) and 2 (L2).

Page 37: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

22

Page 38: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

23

4. RESULTS AND DISCUSSION

4.1. Membrane Characterization

4.1.1. Permporometry

In order to estimate the amount of defects, the synthesized membranes were characterized by permporometry as described in Section 2.2. Table 4.1 shows the permporometry data and the estimated defect distribution for the membrane used in the work described in Papers II and III.

The initial helium permeance at p/p0 = 0, i.e., the helium permeance in the absence of n-hexane, was about 43×10−7 mol m−2 s−1 Pa−1, which denotes the permeance through zeolite pores and defects. This permeance is about half of that for our best membranes with a total film thickness of about 500 nm [43] and the lower permeance observed here should be a result

Table 4.1. Helium permeance at steady state and relative areas of defects for the membrane, estimated from permporometry data.

P/P0 Helium permeance at steady state Defect width Defect interval Relative area

(10-7 mol m-2 s-1 Pa-1) (nm) (nm) of defects * (%)

0.00 43.15 - - -

2.72×10-4 2.86 0.71 - -

4.55×10-4 2.07 0.73 0.71 – 0.73 0.07

1.39×10-3 0.95 0.80 0.73 – 0.80 0.08

1.13×10-2 0.34 1.04 0.80 – 1.04 0.04

1.39×10-1 0.19 1.78 1.04 – 1.78 0.005

4.31×10-1 0.14 5.14 1.78 – 5.14 0.001

> 5.14 0.0008

Total: 0.20

* Area of defects divided by the membrane area.

Page 39: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

24

of a thicker film. However, the defect distribution for this membrane is consistent with that previously reported for high quality membranes prepared by our group [82]. With increasing relative pressure of n-hexane, first zeolite pores and then defects were blocked by the adsorbed hexane molecules, causing the helium permeance to decrease. At the relative pressure of about 1.13×10-2, the helium permeance reduced by more than two orders of magnitude. This relative pressure corresponds to a defect width of 1.04 nm and thus the recorded amount of helium permeance was the permeance through the defects larger than 1.04 nm. The total amount of defects was only 0.20 % of the total membrane surface area indicating a high quality of the membrane. The majority of defects, i.e., ca. 97% of the total amount of defects, were micropore defects smaller than 1 nm in size. As discussed in our previous work [82], these defects are most likely narrow open grain boundaries. In addition, nearly no defects larger than 5 nm in size could be detected.

The permporometry data for the other membranes used in the present work (Papers I and IV–VII) were comparable to those reported in Table 4.1, indicating high quality, and can be found in the corresponding papers.

4.1.2. SEM Analysis

Figure 4.1 shows top view (a) and cross-sectional (b) SEM micrographs of an as-synthesized membrane prepared in the same batch as the membrane used for work described in Papers I, II and III. The micrographs illustrate that the total film thickness, including support invasion, varies in the range 0.5–1 µm, i.e., the average film thickness exceeds 500 nm, in line with permporometry results. Furthermore, the film is comprised of well-intergrown crystals and no defects can be observed, which also is consistent with permporometry observations. The morphology for this membrane was comparable to that previously reported for high quality membranes prepared by our group [83].

Figure 4.1. Top-view (a) and cross-sectional (b) SEM images of an as-synthesised MFI membrane.

Page 40: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

25

4.1.3. XRD Analysis

Figure 4.2 shows XRD patterns of a weakly a-oriented membrane (black) and a b-oriented membrane (red). The reflection marked with an asterisk (*) emanated from the alumina support. The detected reflections emanated from MFI zeolite and the alumina support, indicating that no amorphous material was present in the membrane. The black curve in Figure 4.2 presents the pattern for a weakly a-oriented membrane, as concluded from the observation of a relatively strong (501) reflection [50]. For the red pattern, the five distinct peaks at 8.82°, 17.74°, 26.78°, 36.00°, and 45.48° were attributed to (020), (040), (060), (080), and (0100) reflections, confirming that the membrane was b-oriented.

Figure 4.2. XRD patterns of a weakly oriented membrane (black) and a b-oriented membrane (red).

4.2. Single Gas Permeation

4.2.1. O2 and N2 Single Gas Permeation (Papers I and II)

Figure 4.3 shows the simulated single component adsorption isotherms at 67 K and 90 K. It is clear that N2 exhibits larger affinity than O2 for the zeolite at low loadings at both temperatures. However, the isotherms of N2 and O2 cross at pressures of ca. 8 × 10-9 and 9× 10-6 bar at 67 and 90 K, respectively. The partial pressures in the experiments were several orders of magnitude higher, and consequently a reversal of the adsorption selectivity, favoring the adsorption of O2 may be expected. This is mainly caused by an entropy effect, i.e., O2 with a smaller molecular size can fill the confined space more easily at high loadings [94]. The reverse preferred adsorption from N2 to O2 at low temperature/high loadings were also observed for other materials, e.g., carbon nanotube [94-96].

The multicomponent version of the multisite Langmuir model was used to estimate the adsorption selectivity on the feed side of the membrane, for the same composition as that in

Page 41: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

26

mixture experiment and a total pressure of 100 mbar. The O2/N2 adsorption selectivities were determined to be 4.2 and 1.8 at 67 and 90 K, respectively. Note that the adsorption selectivity increased when the temperature was decreased.

Figure 4.3. Simulated single component adsorption isotherms of O2 (red: 67 K, black: 90 K) and N2 (blue: 67 K, green: 90 K).

Figure 4.4. Single gas permeances for pure O2 and N2 as a function of temperature for a feed pressure of 1.0 bar and a permeate pressure of 100 mbar.

Single gas permeation experiments with pure O2 and N2 at cryogenic temperature were performed in order to better understand the air separation mechanism. The results obtained in the work described in Papers I (with sweep gas) and II (without sweep gas) were similar.

Figure 4.4 shows the permeance of N2 was higher than that of O2 over the entire temperature range despite the fact that the adsorption should favor O2 slightly. This suggests that

Page 42: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

27

diffusivity of N2 should be greater than that of O2 at all temperatures studied and that the membrane selectivity to O2 for gas mixture separation (see results in Section 4.3.1) should mostly arise from adsorption selectivity.

4.2.2. N2 and He Single Gas Permeation (Paper III)

Figure 4.5 shows the results of pure N2 and He gas permeation experiments at cryogenic temperature. The single gas permeance of He was higher than that of N2 at all temperatures.

Figure 4.5. Single gas permeances for pure N2 and He as a function of temperature at a feed pressure of 5.0 bar and a permeate pressure of 0.5 bar.

Figure 4.6. Single gas permeances for pure N2 at a feed pressure of 1.0 bar (Paper II) and 5.0 bar (Paper III).

Page 43: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

28

In addition, single gas permeances for pure N2 at different feed pressures were compared, see Figure 4.6. With the same feed/permeate pressure ratio (=10), lower N2 permeance was observed at a higher feed pressure of 5.0 bar compared to that at 1.0 bar, in concert with the previous findings [28]. It is probably because the zeolite is close to saturation at both the feed and permeate sides when the feed pressure and permeate pressure are high, resulting a low concentration gradient, i.e., a low driving force. The N2 fluxes do not increase proportionally as the feed pressure increases at high fractional surface coverage, resulting in lower permeances at higher pressures.

4.3. Separation of Binary Mixtures at Low Temperature

4.3.1. Air Separation at Cryogenic Temperature (Papers I and II)

O2/N2 separation experiments were performed over the temperature range of 62–106 K using different feed pressures. The results of feed pressure in range of 1.0 bar–100 mbar (Paper II) are presented in Figure 4.7. The membrane is oxygen selective as expected from the estimated O2/N2 adsorption selectivity, see Section 4.2.1. At each feed pressure, a maximum separation factor was observed at an optimum temperature and the separation factor was approaching one at the highest and lowest temperatures studied at all feed pressures. On the other hand, the O2/N2 separation factor was 0.95 at 296 K and a feed pressure of 1.0 bar (not shown). The membrane was slightly N2 selective, likely because N2 adsorbs somewhat stronger than O2 in MFI pores at room temperature (at low loadings). This is consistent with the heat of adsorption values in Table 3.1 and the adsorption simulation results in Section 4.2.1.

At the lowest feed pressure of 100 mbar, a maximum O2/N2 separation factor of 5.0 (corresponding to a selectivity of 6.3) with an O2 permeance of 8.6 × 10-7 mol m-2 s-1 Pa-1 (corresponding to 1025 Barrer) was observed at the optimum temperature of 67 K. This membrane performance is far above the 2008 upper bound in the Robeson selectivity-permeability plot for polymeric membranes at room temperature [17], see Figure 4.8. The separation performance at the other investigated feed pressures (400 mbar, 660 mbar and 1.0 bar) also lies well above the upper bound. The performance of our membrane (in terms of selectivity and permeability) is comparable to that recently reported by Carta et al. [54] for novel polymeric membranes. However, it is worth pointing out that the measured O2 permeance for our thin MFI membrane is more than 100 times higher than that for the polymeric membranes (5.5 × 10-9 mol m-2 s-1 Pa-1) with a thickness ranging from 95 to 181 µm [54]. In addition, Burdyny and Struchtrup [97] showed a hybrid membrane/cryogenic process could reduce the overall energy consumption and distillation column size for air separation. Therefore, our high performance membrane could be a promising candidate for industrial air separation processes.

Page 44: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

29

Figure 4.7. O2/N2 separation factors as a function of temperature for different feed pressures.

Figure 4.8. O2/N2 selectivity versus O2 permeability observed in present work (solid squares) compared to Robeson´s 1991 and 2008 upper bounds: (1) 1.0 bar, (2) 660 mbar, (3) 400 mbar, (4) 100 mbar for feed pressure. Hollow stars represent data for a polymeric membrane with “exceptional performance” recently reported in Science [54]. The original figure is obtained from J. Membr. Sci., 320 (2008) 390, with permission from Elsevier.

Page 45: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

30

Figure 4.9(a) shows the maximum O2/N2 separation factors observed at the optimum separation temperature at each feed pressure. The feed pressures for the separation experiments described in Paper I and Paper II varied from 1.0 bar to 5.0 bar and from 1.0 bar to 100 mbar, respectively. The maximum O2/N2 separation factors increased significantly as the feed pressure was decreased from 5.0 bar to 100 mbar. In addition, the optimum separation temperatures decreased as the feed pressure was increased. For instance, the maximum separation factor was only 2.7 at the optimum temperature of 95 K when the feed pressure was 5.0 bar, whereas a maximum separation factor of 5.0 was observed at 67 K and at a feed pressure of 100 mbar.

Figure 4.9(b) shows permeances of O2 and N2 measured in the gas mixture separation experiments at the optimum separation temperatures as a function of pressure. The figure also shows O2/N2 membrane selectivities at the optimum separation temperatures. The permeance of O2 was higher than that of N2 in the entire feed pressure range. Neither O2 nor N2 permeances observed in the present work exhibited strict correlation with the feed pressures. An entirely opposite situation was observed in the single gas permeation experiments (see Figure 4.4), i.e., the permeance of O2 was lower than that of N2 at all temperatures. Furthermore, the N2 permeance measured in the gas mixture separation experiments was much lower, by ca. a factor of 7, than that measured in the single gas permeation experiments whilst the O2 permeances were comparable. It shows that selective adsorption of O2 reduces N2 permeance in mixture separation experiments compared to single gas N2 permeance. The observations in the work presented in Papers I and II are consistent. The slightly higher selectivity for a feed pressure of 1.0 bar in the work described in Paper I compared to that in Paper II might emanate from the experimental differences. Helium was used as a sweep gas in the former (Paper I), while no sweep gas was used in the latter (Paper II). Counter diffusion of helium can certainly influence the separation performance to some extent. In addition, different membranes were applied for these two experiments.

Page 46: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

31

Figure 4.9. (a) Optimum separation temperatures and maximum O2/N2 separation factors as a function of feed pressure. (b) Permeances of O2 and N2 measured in the gas mixture separation experiments at optimum separation temperatures as a function of feed pressure.

4.3.2. N2/He Separation at Low Temperature (Paper III)

N2/He separation was carried out at various feed and permeate pressures in the temperature range of 82 to 283 K. Figure 4.10 illustrates separation factors as a function of temperature. The membrane was quite nitrogen selective at all conditions studied. The N2/He separation factors exceeded 10 over a wide temperature range, from 90 to 180 K. For each experimental run (E1, E2 and E3), a maximum separation factor was detected at an optimum temperature. The separation is certainly an adsorption-controlled process. N2 adsorbs more extensively in the zeolite compared to He, suppressing the permeation of He. Bakker et al. [98] measured adsorption of various gases in silicalite-1 crystals over a temperature range of 200–600 K and

Page 47: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

32

Figure 4.10. Experimental N2/He separation factors as a function of temperature for different feed and permeate pressures.

at a pressure of 101 kPa. The calculated values for heat of adsorption –ΔHads of N2 and He were 13.8 kJ mol-1 and 0 kJ mol-1, respectively. Less than 0.1 mmol g-1 He (the lower limit for the method used) could be adsorbed at 200 K. In other words, He was “essentially non-adsorbing” and the adsorption of N2 was enormously stronger compared to that of He at those conditions. At higher pressures and lower temperatures, as in the present work, He is expected to adsorb in silicalite-1, albeit significantly less extensively than N2.

Comparing different separation conditions for feed and permeate pressures, the separation factors increased in order of E1 < E2 < E3, with a pressure ratio of 3, 5 and 10, respectively, see Table 2.1. The maximum separation factor observed in E1 was 30.5 at an optimum temperature of 118 K, 50.7 at 130 K in E2 and 61.7 at 124 K in E3.

The data in experiment E3 are described in more detail in Figure 4.11. In this experiment (and similarly in E1 and E2), the permeance of N2 decreased as the temperature decreased from 260 K to 85 K, and the effect was even more significant for temperatures below 118 K. The (mixture gas) permeances of N2 were substantially higher than those of He over the entire temperature range. In contrast, the single gas permeance of He was higher than that of N2 at all temperatures, see Fig 4.4. In addition, the permeances of N2 measured in the separation experiments and single gas permeation experiments were rather close at similar conditions. In contrast, He permeances in the binary separation experiments were lower than those in the single gas experiments by a factor of as much as 10–200. Thus, the transport of N2 is nearly unaffected by the presence of He in the mixture, whereas the transport of He is strongly dependent on the presence of N2 in the gas. These findings demonstrated that the separation mechanism at the conditions studied should be based on selective adsorption and effective

Page 48: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

33

Figure 4.11. Permeances and selectivity for N2/He as a function of temperature in the mixture separation experiments at a feed pressure of 5.0 bar and a permeate pressure of 0.5 bar (E3) in the temperature range of 85–260 K (a) and 73–85 K (b).

transport of N2, impeding the permeation of He. It is worth noting that the N2 permeances measured at cryogenic temperatures were only a few times lower than those measured at the highest investigated temperature of 260 K, but still more than 100 times larger than those reported for polymeric membranes at room temperature [17]. A maximum selectivity of 75.7 was obtained at an optimum temperature of 124 K with a high N2 permeance of 39 ×10−7 mol m−2 s−1 Pa−1. At these conditions, the N2 concentration in the permeate was as high as 98.4%. A much lower membrane area would be needed for our zeolite membranes with higher permeance and high selectivity as compared to polymeric membranes to carry out a given separation task. The selectivities were approaching one when the temperature was close to

Page 49: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

34

room temperature. The higher N2/He selectivities at cryogenic temperatures should primarily emanate from more extensive adsorption of N2 at cryogenic temperature, resulting in more effective blocking of the He transport. As temperature increase, less N2 adsorbs, liberating the permeation pathways for He. Besides, the same trend was observed for the lowest temperatures studied. The drastic drop in selectivities at a temperature ranging from 118 to 85 K was related to the drastic drop in N2 permeance, most likely due to the reduced diffusion of N2 at those temperatures. In contrast, at temperatures below 85 K, the selectivity was increasing, reaching a very high value of ca. 400 at 73 K, see Figure 4.11 (b). Since the boiling point of N2 at 2.6 bar (= 5bar × 52.6%) is 90 K, N2 gas is most likely condensing in the cell and on the membrane surface at temperatures below the boiling point, completely blocking the transport of He. The condensation of N2 should also result in a much lower mass transfer rate through the membrane. Indeed, the permeances of N2 and He were extremely low at temperatures below 85 K, i.e., 2.7 ×10−8 mol m−2 s−1 Pa−1 and 6.0 ×10−10 mol m−2 s−1 Pa−1, respectively.”

To further understand the behavior of the membranes at conditions E3 (5.0 bar feed pressure and 0.5 bar permeate pressure), the membrane processes was simulated at three temperatures, viz. 106, 124 and 171 K, see Table 4.2. Note that the model was not fitted to the experimental data. The permselectivities obtained from the simulations are in reasonably good agreement with the experimental observations and the modeling results capture the main trend with maximum in selectivity at ca. 120 K. The driving force, expressed as the difference in adsorbed loading of N2 over the zeolite film Δqads, is largest at 124 K. This is probably the main reason for the observed maximum in selectivity at this temperature.

As the lighter molecule, He is preferentially transported through the defects, which decreases the selectivity of the membrane. According to the model, the fractions of He transported via flow-through defects were 93%, 72% and 23% at 106 K, 124 K and 171 K, respectively. For comparison, the relative flux through defects was always below 1% for N2. If the transport via defects could be completely eliminated, the selectivity would reach impressive values of 111, 192 and 38, respectively.

The concentration polarization indices Cmem,N2/Cbulk,N2, were 0.75, 0.54 and 0.50 at 106, 124 and 171 K, respectively, indicating that external mass transfer limitations have affected the membrane process and that the limitations were more severe at higher temperatures where the permeance is higher. The limitations may be diminished by increasing the velocity of the feed. If the feed flow rate could be increased to 50 l min-1, yielding a concentration polarization index of 0.95, the selectivity would nearly double, from 53 to 97 at 124 K. The increase in selectivity primarily originates from a smaller flux of He through the flow-through defects due to the reduced partial pressure of He at the membrane feed interface as the polarization index is increased. However, in our case, the feed flow rate was also limited by the experimental setup, e.g., the capacities of the mass flow controllers and cooling capacity of the cryostat.

The relative pressure drop over the support (compared to the total pressure drop over the membrane) after compensating for concentration polarization was estimated to be 5, 16 and

Page 50: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

35

27% at 106, 124 and 171 K, respectively, thus increasing with increasing permeance through the membrane. In concert with previous findings, the main part of the pressure drop was over the thin top layer of the support [48,89]. Reducing the thickness of this layer from 30 µm to 10 µm would be an effective but still realistic option. Such a reduction in the thickness could diminish the relative pressure drop from 27% to 10% at the highest gas permeance at 171 K and from 16% to 7.5% at 124 K. If both the feed flow rate could be increased to 50 l min-1 and the thickness of the top layer could be reduced to 10 μm, the selectivity at 124 K would nearly double, from 53 to 99.

Table 4.2. Summary of results from simulations and selected experimental data for comparison.

T (K) 106 124 171

Δqads for N2 (mol m-3) 960 2500 1200

Δqads for He (mol m-3) 113 149 196

Selectivity, exp 64 76 18

Selectivity, sim 11 53 29

Relative transport through defects, He (%)

93 72 23

Selectivity, no defects 111 192 38

Concentration polarization index Cmem/Cbulk

0.75 0.54 0.50

Relative pressure drop over support (%)

5 16 27

Page 51: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

36

4.3.3. CO2 Separation at Low Temperature (Papers IV, V and VII)

Separation of binary gas mixtures related to synthesis gas, i.e., CO2/H2 (Papers IV and V) and CO2/CO (Papers IV and VII), using MFI membranes has been studied at temperatures down to 235 K and a feed pressure of 9.0 bar. Both CO2/H2 and CO2/CO separation factors increased significantly with decreasing temperature, as anticipated from a previous work [28].

Several membranes (b-oriented membranes in Paper IV and weakly oriented membranes in Paper V) were evaluated for CO2/H2 separation at low temperature. Taking the work in Paper V as an example (see Figure 4.12 and Figure 4.13 for membrane M1), the observed CO2/H2 separation factor was 20 at 298 K, which is a bit higher than the previously reported separation factor of 16 at 296 K [28]. The separation factor reached as high as 165 at 235 K, the lowest investigated temperature in this work, close to the CO2 sublimation point at the

Figure 4.12. Separation factors for CO2/H2 separation at low temperature (Paper V).

Figure 4.13. Permeances of CO2 and H2 measured for CO2/H2 mixture separation at low temperature (Paper V).

Page 52: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

37

studied experimental conditions. Under this condition, the CO2 concentration in the permeate stream was as high as 99.4%, see Table 4.3. The permeance of CO2 was much higher than that of H2 for the entire temperature range, see Figure 4.13. The Henry coefficient and –ΔHads (see Table 1.1) are much larger for CO2 than those for H2, hence CO2 can adsorb more extensively than H2, thereby blocking H2 transport. This blocking effect is higher at a lower temperature due to the increased CO2 adsorption [28]. The CO2 permeance decreased as the temperature was decreased from 310 to 235 K, however it was still very high in the range of 62–78×10−7 mol m−2 s−1 Pa−1, corresponding to a CO2 flux of 350–420 kg m−2 h−1. The permeance was one to two orders of magnitude higher than those reported for inorganic and polymeric membranes in the literature. At 253 K, the CO2 flux was ca. 390 kg h−1 m−2, 130 times higher than that (3 kg h−1 m−2) reported for SAPO-34 zeolite membranes, which displayed a (lower) separation factor of ca. 110 at similar experimental conditions [32]. Another membrane (denoted M2) with similar defect distribution displayed similar performance for CO2/H2 separation under the same conditions, illustrating good reproducibility of the separation results. The separation factor was even a bit higher, i.e., 84–202 at 235–270 K, with a CO2 flux of 303–448 kg m−2 h−1.

Table 4.3. CO2 flux, permeate concentration and membrane selectivity at different temperature (Paper V).

Permeate concentration (mol%)

T (K) CO2 flux (kg h-1 m-2) CO2 H2 CO2/H2 selectivity

310 423 93.22 6.78 17

300 429 95.36 4.61 26

290 428 96.71 3.28 37

270 420 98.47 1.53 82

260 406 98.91 1.08 117

250 383 99.20 0.80 159

240 364 99.33 0.67 189

235 356 99.40 0.60 210

Page 53: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

38

Figure 4.14. CO2/H2 separation factor versus CO2 permeance at low temperature in this work compared to the previous reported results by zeolite membranes (labeled 1–6, more information contained in Table 4.4).

Table 4.4. Summary of experimental data used to compile Figure 4.14.

No. Temperature Zeolite Type Membrane thickness Reference

(K)

(µm)

1 253 to 296 SAPO-34 5 [32]

2 298 DDR 10 [99]

3 295 MFI (Silicalite-1) 50-60 [100]

4 308 FAU (NaY) 3 [31]

5 298 Ba-ZSM-5 0.55 [46]

6 298 MFI 0.7 [28]

This work 235 to 298 MFI 0.5 Papers IV and V

Figure 4.14 shows the CO2/H2 separation performance of the MFI membrane prepared in this work (Papers IV and V) and that of other zeolite membranes previously reported in the literature. An “upper bound” was also drawn to guide the eyes. Our membrane performance at low temperature is above this upper bound as previously reported for CHA (SAPO-34), DDR,

Page 54: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

39

FAU, Ba-ZSM-5 and MFI membranes. The membrane performances in the work reported in Paper IV and Paper V are similarly excellent and are comparable to that indicated by point 6 [28], also reported by our group. The membranes are more selective at low temperature but the permeance is somewhat lower.

On the other hand, to compare with H2 selective membrane for H2/CO2 mixture separation, the highest reported H2 permeance was 20×10−7 mol m−2 s−1 Pa−1 for a silica membrane with and selectivity of 70 [101]. Li et al. reported a graphene oxide membrane with a H2 permeance of ca. 2×10−7mol m−2 s−1 Pa−1 and selectivity of 3400 [102].

b-Oriented membranes were evaluated for CO2/CO separation at low temperature (Papers IV and VII). As it is shown in Figure 4.15, the separation factor also increased to quite a high level, e.g., 21 as the temperature was decreased to 258 K with a CO2 permeance of 57×10−7 mol m−2 s−1 Pa−1. In comparison with the best commercial polymeric membranes, the zeolite membrane in the present work had ca. 70 times higher permeance with similar membrane selectivity under the same conditions [75]. The separation factor was lower compared to that for CO2/H2, because CO adsorbs more strongly than H2 with a 17 times higher Henry´s constant [103]. The difference of adsorption abilities between CO2 and CO is smaller than that between CO2 and H2.

To assess whether the results obtained in this work were affected by the pressure drop over the support and concentration polarization, modeling of these effects was conducted at 303, 278 and 258 K. As it is shown in Table 4.5, the relative pressure drop over the support was estimated to be 33%, 27% and 21%, respectively, indicating that the support adversely affected the performance of the membrane. Most of the resistance was found to be in the thin top layer of the support. The concentration polarization index was determined to be 0.83, 0.87 and 0.90 at 303, 278 and 258 K, indicating that concentration polarization reduced the performance of the membrane, especially at the higher temperatures.

Figure 4.15. Separation factors and permeances for CO2/CO separation at low temperature.

Page 55: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

40

Table 4.5. Relative pressure drop over support, concentration polarization index, permselectivity for the zeolite film and adsorption selectivity for CO2/CO separation.

T (K) 258 278 303

Relative pressure drop over support (%)

21 27 33

Concentration polarization index, Cmem/Cbulk

0.90 0.87 0.83

Selectivity (exp), overall 26 14 6

Permselectivity, film 41 24 15

Adsorption selectivity on feed side 46 30 18

The intrinsic permselectivities for the zeolite film after correcting for concentration polarization and pressure drop over the support, were 15, 24 and 41 at 303, 278 and 258 K, respectively. These values are significantly higher than the overall selectivities observed in the experiment being 6, 14 and 26, indicating the adverse effect of polarization and the support. The adsorption selectivities on the feed side of the membrane estimated by IAST after removing the effect of concentration polarization were very similar to the experimental selectivities, which confirmed that membrane selectivity can be ascribed to the adsorption selectivtity.

4.3.4. Light Olefins/N2 Separation at Low Temperature (Paper VI)

Separation of C3H6/N2. The C3H6/N2 mixture (20/80) separation was performed at different separation temperatures and a feed pressure of 10.0 bar, see Figure 4.16. The observed C3H6/N2 selectivity was high throughout the whole temperature range studied. A maximum separation selectivity of 80 was observed at 296 K, with a corresponding separation factor of 43. Similar trends, with an optimum separation temperature have been reported previously for similar gas mixtures, for instance for the separation of n-C4H10 or C2H6 from binary mixtures with CH4 [104]. Starting from the highest temperature (326K), the selectivity increased with decreasing temperature, due to increased C3H6/N2 adsorption selectivity. According to IAST, the C3H6/N2 adsorption selectivities on the feed side, after compensating for concentration polarization, were 141, 234 and 238 at 326, 296 and 265 K, respectively, see Table 4.7. The adsorption of C3H6 is favoured as compared to N2 as the temperature decreased because of the higher heat of adsorption of C3H6 (–39.2 kJ mol-1) compared to that for N2 (–17.0 kJ mol-1), resulting in higher C3H6/N2 adsorption selectivity at lower temperature in this range. At lower temperatures saturation loading is approached and the adsorption selectivity does not increase anymore. Figure 4.16 shows that the N2 permeance was very low and decreased even further as the temperature decreased in the presence of C3H6. The highest observed C3H6 permeance

Page 56: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

41

was 39×10-7 mol m-2 s-1 Pa-1 at 326 K and it decreased almost linearly with decreasing temperature. In this range, as the temperature decreased, the adsorbed loading of C3H6 did not increase much. Using the IAST, the loading of C3H6 was estimated (after correcting for concentration polarization) to 11.7, 12.2 and 12.4 kmol/(m3 accessible pore volume) at 326, 296 and 265 K, respectively. This is comparable to the saturation loading of ca. 12.5 kmol m-3. It shows that the zeolite was close to saturation with C3H6 at the conditions of the experiments. The observed decrease in C3H6 permeance with decreasing temperature should mainly be an effect of decreasing diffusivity with decreasing temperature. C3H6/N2 separation selectivities of 8–10 were usually reported for commercial polymeric membranes [15,105], while much higher selectivities were observed for MFI membranes in the present work. Normally, the permeances of light hydrocarbons for polymer membranes or carbon molecular sieve membranes were lower than 3.35×10-7 mol m-2 s-1 Pa-1 [105-109]. At the optimum separation temperature, a C3H6 permeance of 24×10-7 mol m-2 s-1 Pa-1 was observed, i.e. more than 7 times higher than for commercial silicon rubber polymeric membranes [110]. All of this indicates that ultra-thin MFI membranes are very promising for separation of C3H6/N2 mixtures, in terms of permeance and selectivity.

Figure 4.16. The separation factors, separation selectivities and permeances as a function of temperature for C3H6/N2 separation.

Page 57: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

42

Table 4.6. Concentration polarization index, relative pressure drop over support and adsorption selectivity on feed side (estimated by IAST) for olefins/ N2 separation.

C3H6/N2 C2H4/N2

Temperature (K) 265 296 326 258 284 315

Adsorption selectivity on feed side

238 234 141 69 47 34

Concentration polarization index, Cmem/Cbulk

0.97 0.93 0.88 0.96 0.87 0.91

Relative pressure drop over support (%)

3.2 8.4 17.0 8.9 22.0 25.7

The mathematical model showed that the pressure drop over the support was 17 % at the highest flux (326 K), 8.4% at the highest selectivity (296 K) and only 3.2 % at the lowest flux (265 K), see Table 4.6. In concert with our previous findings, the major part of the pressure drop over the support is over the thin, 30 µm, top layer [48]. Modeling also indicated that concentration polarization on the feed side only had a relatively small effect on the outcome of the experiments. The concentration polarization indexes, Cmem/Cbulk, ranged from 0.88 to 0.97, see Table 4.6.

Separation of C2H4/N2. Figure 4.17 shows the results for separation of a 20/80 mixture of C2H4/N2 as a function of temperature. The membrane was also selective towards the olefin in this case. Similar trends were observed for this system as for the C3H6/N2 system when temperature changed, albeit the magnitudes differed. However, the selectivity was significantly lower than observed for C3H6/N2 as C2H4 adsorbs weaker than C3H6 and is therefore less effective in the competitive adsorption with N2 [109,111]. The C2H4/N2 adsorption selectivities at the feed side estimated by IAST, after compensating for concentration polarization, were 34, 47 and 69 at 315, 284 and 258 K, respectively, see Table 4.6. Maximum separation factor and selectivity for C2H4/N2 mixtures were observed at 277 K and were 6 and 8.4, respectively, see Figure 4.17. Polymeric membranes usually have C2H4/N2 seletivities of 7–12 at 233 K [110]. The permeance decreased with decreasing temperature for both components. In addition, the permeance of both components were higher for the C2H4/N2 mixture than for the C3H6/N2 mixture, reflecting the higher diffusion rate of C2H4 coupled with a weaker adsorption and lower ability to block transport of N2 [112,113]. The highest C2H4 permeance was around 80×10-7 mol m-2 s-1 Pa-1 at 315 K, and was 57×10-7 mol m-2 s-1 Pa-1at the optimum temperature of 277 K. This permeance is more than 3 orders of magnitude larger than that of reported for PDMS membranes (ca. 3×10-9 mol m-2 s-1 Pa-1) [109] and more than 2 orders of magnitude larger than that claimed for a polytrimethysilylpropyne membrane in a patent (3×10-8 mol m-2 s-1 Pa-1) [110]. As the

Page 58: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

43

permeance of C2H4 was higher than that of C3H6, the pressure drop over the support and concentration polarization may potentially be a greater issue. The outcomes of the modeling are presented in Table 4.6. Concentration polarization was similar as for C3H6/N2 separation despite the higher flux of hydrocarbon observed in the C2H4/N2 experiments. The higher gas phase diffusivity of C2H4 compared to C3H6 (by ca. 25%) is the reason for the similarities in concentration polarization index observed for the two systems [111]. On the other hand, pressure drop over the support was more severe for the C2H4/N2 mixtures compared to that for C3H6/N2 mixtures because of the higher flux. Similarly as for the C3H6/N2 separation, the performance of the membrane may be further improved by increasing the permeability of the support or by improving mixing on the feed side, e.g. by increasing the velocity of the gas or by introducing turbulence generating elements in the cell, thus minimizing concentration polarization.

Figure 4.17. Separation factors and selectivities and permeances as a function of temperature for a C2H4/N2 separation.

Page 59: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

44

4.3.5. Concluding Remarks

Figure 4.18 summarizes the separation results for all the gas mixture separations studied in the present work. All the separations are adsorption-control processes. The strongly adsorbing molecules preferentially permeated, hindering the transport of the weakly adsorbing molecules. The maximum separation factors for different separations increase in order of: O2/N2 < C2H4/N2 < CO2/CO < C3H6/N2 < N2/He < CO2/H2. This could be explained by the difference of adsorption properties between the gas pairs, see Table 1.1. For example, the difference of properties between CO2 and H2 is huge, and far bigger compared to CO2/CO, e.g., the heat of adsorption of CO2, CO and H2 are -27.2, -16.7 and -5.9 kJ mol-1, respectively. On the other hand, O2 and N2 molecules have rather similar properties, e.g., polarizability, resulting lower adsorption selectivity and thus lower membrane selectivity compared to other gas mixture separations.

In addition, the adsorption selectivities might be favored at low temperature, however, diffusivities decrease as temperature decreases. The adsorption loading becomes saturated at both of membrane sides and membrane selectivities start to decrease when the temperature is too low. The optimum separation temperatures increase in order of: O2/N2 < N2/He < CO2 separations < C2H4/N2 < C3H6/N2. This is related with adsorption properties for the stronger adsorbing molecule in each pair. The absolute values of the heat of adsorption are in a similar order: O2 < N2 < CO2 ≈ C2H4 < C3H6 (see Table 1.1), resulting that the adsorption of C3H6 reaches saturation at a relatively higher temperature compared to that for other gases like O2 and N2. However, the author does not want to draw too far reaching conclusions based on the observation since the separations were carried out under different feed and permeate pressures.

Figure 4.18. Separation factors as a function of temperature for O2/N2, N2/He, CO2/H2, CO2/CO, C3H6/N2 and C2H4/N2 separations.

Page 60: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

45

5. CONCLUSIONS

In the present work, zeolite membranes were for the first time evaluated for gas separation at cryogenic temperature. For air separation, the membranes were found to be O2-selective at all the conditions studied with a feed pressure ranging from 5.0 bar to 100 mbar. The membrane performance was well above the upper bound in the 2008 Robeson selectivity-permeability plot when the feed pressure was lower than 1.0 bar. As the feed pressure decreased, the maximum separation factors for O2/N2 increased while the optimum separation temperatures decreased. The highest O2/N2 selectivity was 6.3, with O2 permeance of 8.6 × 10-7 mol m-2 s-1 Pa-1 when the feed pressure was 100 mbar and the separation temperature was 67 K. The permeance was nearly 100 times higher than that recently reported for promising polymeric membranes. The selectivity to O2 should primarily emanate from O2/N2 adsorption selectivity.

The zeolite membranes were evaluated for N2/He separation at low temperature down to 83 K with different feed/permeate pressure ratios. The membranes were found to be N2-selective at all the conditions studied. A highest separation selectivity of 75.7 was observed at a temperature of 124 K, a feed pressure of 5.0 bar and a permeate pressure of 0.5 bar with a N2 permeance of as high as 39×10−7 mol m−2 s−1 Pa−1 (corresponding to a flux of 83 kg m−2 h−1). The selectivity to N2 should mostly emanate from stronger adsorption of N2 in the zeolite pores, thus blocking the transport of He, especially at cryogenic temperatures, as observed experimentally and by a mathematical model. Furthermore, the N2/He selectivity was found to increase with increasing feed/permeate pressure ratios. The main trends of membrane selectivity with a maximum value at ca. 120 K could be described by the simulations. Modeling results indicated that defects, concentration polarization and pressure drop over the support limited the separation performance significantly. The selectivity at ca. 120 K could increase to 99 by reducing the concentration polarization and increasing the permeability of the support. A zeolite membrane process could be rather competitive for N2/He separation if integrated into the existing cryogenic distillation process.

CO2 separation from binary mixtures related to synthesis gas by zeolite membranes were also investigated at low temperature. The CO2/H2 selectivity was 17 at 298 K and 9.0 bar feed pressure, and was significantly higher at lower temperature. It reached as high as 303 at 235 K with a fairly high CO2 permeance of 53×10−7 mol m−2 s−1 Pa−1. The highest CO2/CO selectivity was 26 at 258 K with a CO2 permeance of 57×10−7 mol m−2 s−1 Pa−1. These membrane performances are among the best results ever reported for CO2 separation. CO2 adsorption inhibited H2 and CO permeances, resulting in membrane selectivity to CO2. A higher selectivity at a lower temperature results from stronger CO2 adsorption. The membrane performance for CO2/CO separation was affected by the pressure drop over the support.

Zeolite membrane was used to separate C3H6 or C2H4 from binary mixtures with N2 at temperatures down to 258 K. The membrane showed a high performance for both separations. The highest C3H6/N2 selectivity was 80, coupled with a permeance of 22×10-7 mol m-2 s-1 Pa-1

Page 61: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

46

at room temperature. The maximum separation selectivity for C2H4/N2 gas mixture was 8.4 at 277 K with a C2H4 permeance of 57×10-7 mol m-2 s-1 Pa-1. The high selectivity was ascribed to strong adsorption of olefin effectively blocking the transport of N2 through the zeolite. The weaker adsorption of C2H4 could not block the pores as effectively as C3H6, but the lighter C2H4 exhibited a larger permeance than C3H6. Pressure drop over the support and concentration polarization had an adverse effect on the membrane performance, in particular for C2H4/N2 separation when the permeance was high. The separation results demonstrate that the ultra-thin MFI zeolite membrane is promising for recovery of light olefins from degassing/vent streams in the petrochemical industries.

To summarize, selective adsorption is the key factor for the membrane selectivity in the present work. The permeance of a weakly adsorbing component was significantly reduced in the presence of a stronger adsorbing component. The strongly adsorbing molecules preferentially permeated and the selectivities were higher at low temperatures. On the other hand, the permeances tended to decrease as temperature decreased under the conditions investigated, but the membrane used in present work still exhibited fairly high permeances, i.e., high flux, mainly due to the ultra-thin porous zeolite film grown on a fully open graded and highly permeable support. Lower membrane area would be needed for our zeolite membranes with higher permeance as compared to polymeric membranes for a given separation task. This work suggests high-flux zeolite membranes have potential applications for selective gas separations at low temperature.

Page 62: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

47

6. FUTURE WORK

It is of great interest to further explore the performance of zeolite membranes at low temperature. The newly built and well-controllable cryostat system makes it possible to investigate various gas mixture separations at a temperature down to 22 K, with either high pressure or low pressure. Our ultra-thin membranes will be evaluated for separation of CH4/He and C2H4/C2H6, etc. at low temperature.

An important aspect of a membrane process is stability. Several zeolite membranes were investigated at low temperature, however, no systematic study has been carried out due to lack of time. It would be of great use to survey membranes´ stability after many cycles of cooling and heating in a long run.

Other types of zeolite membranes may give better performance for some separations. For example, DDR with smaller zeolite pores may work rather than MFI membranes in some cases.

The membranes tested in lab-scale are prepared on disc-shape supports in the present work. Tubular membranes modules are desirable for assemble and scale-up from industrial application point of view. Therefore, it is another key step to manufacture a good membrane on tubular supports.

It would be necessary to estimate the cost in detail before the process is applied in industry, including the operation cost. It could be designed as either membrane processes or combined with other separation units, depending on the economically feasibility in specific cases.

Page 63: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

48

Page 64: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

49

REFERENCES

[1] C.N. Satterfield. Heterogeneous catalysis in industrial practice. (1991).

[2] D.W. Breck, Zeolite molecular sieves: structure, chemistry, and use, Wiley New York, 1973.

[3] International Zeolite Association (IZA). http://www.iza-structure.org/databases/.

[4] R. Szostak, Handbook of molecular sieves: Structures, Springer, 1992.

[5] X.G. Lei, S. Jockusch, M. Francesca Ottaviani, and N.J. Turro. In situ EPR investigation of the addition of persistent benzyl radicals to acrylates on ZSM-5 zeolites. Direct spectroscopic detection of the initial steps in a supramolecular photopolymerization. Photochem.Photobiol.Sci., 2 (2003) 1095.

[6] S. Sircar, A.L. Myers. Gas separation by zeolites. Handbook of Zeolite Science and Technology, (2003) 1063.

[7] A. Hedström. Ion exchange of ammonium in zeolites: a literature review. J. Environ. Eng., 127 (2001) 673.

[8] W. Koros, Y. Ma, and T. Shimidzu. Terminology for membranes and membrane processes. J. Membr. Sci., 120 (1996) 149.

[9] R. Baker. Membrane technology and applications. England: John Wiley and Sons, (2004).

[10] A. Tavolaro, E. Drioli. Zeolite membranes. Adv. Mater., 11 (1999) 975.

[11] M.G. Buonomenna, G. Golemme, Advanced materials for membrane preparation, Advanced Materials for Membrane Preparation, Bentham Science Publishers Ltd., 2012.

[12] V.N. Burganos. Membranes and membrane processes. MRS Bull, 24 (1999) 19.

[13] Z. Olujic, P. Perez, F. De Bruijn, J. de Graauw, and P. Jansen. Augmenting distillation by membranes: Developments and prospects. Chem. Biochem. Eng. Q., 20 (2006) 301.

[14] E. Curcio, A. Criscuoli, and E. Drioli. Membrane crystallizers. Ind. Eng. Chem. Res., 40 (2001) 2679.

[15] R.W. Baker, B.T. Low. Gas separation membrane materials: A perspective. Macromolecules, 47 (2014) 6999.

[16] L.M. Robeson. Correlation of separation factor versus permeability for polymeric membranes. J. Membr. Sci., 62 (1991) 165.

Page 65: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

50

[17] L.M. Robeson. The upper bound revisited. J. Membr. Sci., 320 (2008) 390.

[18] M. Yu, R.D. Noble, and J.L. Falconer. Zeolite membranes: Microstructure characterization and permeation mechanisms. Acc.Chem.Res., 44 (2011) 1196.

[19] Y. Lin, M.C. Duke. Recent progress in polycrystalline zeolite membrane research. Curr. Opin. Chem. Eng., 2 (2013) 209.

[20] E. Sjöberg, L. Sandström, O.G. Öhrman, and J. Hedlund. Separation of CO2 from black liquor derived syngas using an MFI membrane. J. Membr. Sci., 443 (2013) 131.

[21] H. Zhou, D. Korelskiy, T. Leppäjärvi, M. Grahn, J. Tanskanen, and J. Hedlund. Ultrathin zeolite X membranes for pervaporation dehydration of ethanol. J. Membr. Sci., 399 (2012) 106.

[22] P. Dai, D. Sherwood Jr, and B. Martin. Influence of zeolite secondary porosity on performance of resid hydrocraking catalysts. Stud. Surf. Sci. Catal., 83 (1994) 489.

[23] A. Pantazidis, C. Mirodatos. Mechanistic approach of the oxidative dehydrogenation of propane over VMgO catalysts by in situ spectroscopic and kinetic techniques. Stud. Surf. Sci. Catal., 101 (1996) 1029.

[24] K. Sato, K. Sugimoto, and T. Nakane. Mass-production of tubular NaY zeolite membranes for industrial purpose and their application to ethanol dehydration by vapor permeation. J. Membr. Sci., 319 (2008) 244.

[25] K. Sato, K. Sugimoto, and T. Nakane. Synthesis of industrial scale NaY zeolite membranes and ethanol permeating performance in pervaporation and vapor permeation up to 130 C and 570kPa. J. Membr. Sci., 310 (2008) 161.

[26] J. Gascon, F. Kapteijn, B. Zornoza, V. Sebastián, C. Casado, and J. Coronas. Practical approach to zeolitic membranes and coatings: state of the art, opportunities, barriers, and future perspectives. Chem. Mater., 24 (2012) 2829.

[27] H. Guo, G. Zhu, H. Li, X. Zou, X. Yin, W. Yang, et al. Hierarchical growth of large-scale ordered zeolite silicalite-1 membranes with high permeability and selectivity for recycling CO2. Angew. Chem., 118 (2006) 7211.

[28] L. Sandström, E. Sjöberg, and J. Hedlund. Very high flux MFI membrane for CO2 separation. J. Membr. Sci., 380 (2011) 232.

[29] M.P. Bernal, J. Coronas, M. Menéndez, and J. Santamaría. Separation of CO2/N2 mixtures using MFI-type zeolite membranes. AIChE J., 50 (2004) 127.

[30] J. Caro, M. Noack. Zeolite membranes – Recent developments and progress. Microporous Mesoporous Mater., 115 (2008) 215.

[31] K. Kusakabe, T. Kuroda, K. Uchino, Y. Hasegawa, and S. Morooka. Gas permeation properties of ion-exchanged faujasite-type zeolite membranes. AIChE J., 45 (1999) 1220.

Page 66: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

51

[32] M. Hong, S. Li, J.L. Falconer, and R.D. Noble. Hydrogen purification using a SAPO-34 membrane. J. Membr. Sci., 307 (2008) 277.

[33] S. Li, J.  Falconer, and R.  Noble. Improved SAPO-34 Membranes for CO2/CH4 Separations. Adv. Mater., 18 (2006) 2601.

[34] S. Himeno, T. Tomita, K. Suzuki, K. Nakayama, K. Yajima, and S. Yoshida. Synthesis and permeation properties of a DDR-type zeolite membrane for separation of CO2/CH4 gaseous mixtures. Ind. Eng. Chem. Res., 46 (2007) 6989.

[35] T. Tomita, K. Nakayama, and H. Sakai. Gas separation characteristics of DDR type zeolite membrane. Microporous Mesoporous Mater., 68 (2004) 71.

[36] A. Julbe, J. Motuzas, F. Cazevielle, G. Volle, and C. Guizard. Synthesis of sodalite/αAl2O3 composite membranes by microwave heating. Sep.Purif.Technol., 32 (2003) 139.

[37] S. Khajavi, J.C. Jansen, and F. Kapteijn. Application of hydroxy sodalite films as novel water selective membranes. J.Membr.Sci., 326 (2009) 153.

[38] N. Kosinov, J. Gascon, F. Kapteijn, and E.J.M. Hensen. Recent developments in zeolite membranes for gas separation. J.Membr.Sci., 499 (2016) 65.

[39] J. van den Bergh, W. Zhu, J. Gascon, J.A. Moulijn, and F. Kapteijn. Separation and permeation characteristics of a DD3R zeolite membrane. J. Membr. Sci., 316 (2008) 35.

[40] J. Lindmark, J. Hedlund. Modification of MFI membranes with amine groups for enhanced CO2 selectivity. J. Mater. Chem., 20 (2010) 2219.

[41] J. Caro, M. Noack, P. Kölsch, and R. Schäfer. Zeolite membranes – state of their development and perspective. Microporous Mesoporous Mater., 38 (2000) 3.

[42] M.A. Snyder, M. Tsapatsis. Hierarchical nanomanufacturing: from shaped zeolite nanoparticles to high–performance separation membranes. Angew. Chem. Int. Ed., 46 (2007) 7560.

[43] J. Hedlund, J. Sterte, M. Anthonis, A. Bons, B. Carstensen, N. Corcoran, et al. High-flux MFI membranes. Microporous Mesoporous Mater., 52 (2002) 179.

[44] H. Zhou, D. Korelskiy, E. Sjöberg, and J. Hedlund. Ultrathin hydrophobic MFI membranes. Microporous Mesoporous Mater., 192 (2014) 76.

[45] L. Sandström, J. Lindmark, and J. Hedlund. Separation of methanol and ethanol from synthesis gas using MFI membranes. J. Membr. Sci., 360 (2010) 265.

[46] J. Lindmark, J. Hedlund. Carbon dioxide removal from synthesis gas using MFI membranes. J. Membr. Sci., 360 (2010) 284.

Page 67: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

52

[47] J. Hedlund, D. Korelskiy, L. Sandström, and J. Lindmark. Permporometry analysis of zeolite membranes. J. Membr. Sci., 345 (2009) 276.

[48] D. Korelskiy, T. Leppäjärvi, H. Zhou, M. Grahn, J. Tanskanen, and J. Hedlund. High flux MFI membranes for pervaporation. J.Membr.Sci., 427 (2013) 381.

[49] L. Yu, D. Korelskiy, M. Grahn, and J. Hedlund. Very high flux MFI membranes for alcohol recovery via pervaporation at high temperature and pressure. Sep.Purif.Technol., 153 (2015) 138.

[50] J. Hedlund, F. Jareman, A. Bons, and M. Anthonis. A masking technique for high quality MFI membranes. J. Membr. Sci., 222 (2003) 163.

[51] E.R. Geus, H. van Bekkum, W.J. Bakker, and J.A. Moulijn. High-temperature stainless steel supported zeolite (MFI) membranes: preparation, module construction, and permeation experiments. Microporous Mater., 1 (1993) 131.

[52] E. Piera, C.A.M. Brenninkmeijer, J. Santamaría, and J. Coronas. Separation of traces of CO from air using MFI-type zeolite membranes. J. Membr. Sci., 201 (2002) 229.

[53] A. Smith, J. Klosek. A review of air separation technologies and their integration with energy conversion processes. Fuel Process. Technol., 70 (2001) 115.

[54] M. Carta, R. Malpass-Evans, M. Croad, Y. Rogan, J.C. Jansen, P. Bernardo, et al. An efficient polymer molecular sieve for membrane gas separations. Science, 339 (2013) 303.

[55] K.H. Kaplan. Helium shortage hampers research and industry. Phys. Today, 60 (2007) 31.

[56] A. Cho. Helium-3 shortage could put freeze on low-temperature research. Science, 326 (2009) 778.

[57] J. Agrawal, S. Sourirajan. Helium separation by cellulose acetate membranes. J. Appl. Polym. Sci., 13 (1969) 1065.

[58] R.M. Ruemann, A. Fedoritenko, Physical basis of separating helium and nitrogen, Redstone scientific information center redstone arsenal al, 1964.

[59] H. Hsieh, Inorganic membranes for separation and reaction, Alcoa Technical Center, USA, Elsevier, 1996.

[60] V. Abetz, T. Brinkmann, M. Dijkstra, K. Ebert, D. Fritsch, K. Ohlrogge, et al. Developments in membrane research: From material via process design to industrial application. Adv. Eng. Mater., 8 (2006) 328.

[61] N.K. Das, H. Chaudhuri, R.K. Bhandari, D. Ghose, P. Sen, and B. Sinha. Purification of helium from natural gas by pressure swing adsorption. Curr.Sci., 95 (2008) 1684.

Page 68: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

53

[62] G. Bird, W. Granville, Separation of Nitrogen from Helium Using Pressure-Swing Adsorption, in K.D. Timmerhaus (Ed.), Advances in Cryogenic Engineering, Springer, 1995, pp. 463-473.

[63] J. Schrier. Helium separation using porous graphene membranes. Chem. Lett., 1 (2010) 2284.

[64] S. Stern, T. Sinclair, P. Gareis, N. Vahldieck, and P. Mohr. Helium recovery by permeation. Ind. Eng. Chem., 57 (1965) 49.

[65] M.E. Rezac, N. Shanan Moore, and A.B. Georgia. Effect of Temperature on the Transport Properties and Morphology of Polymeric Asymmetric Membranes. Sep.Sci.Technol., 32 (1997) 505.

[66] V. Havredaki, J.H. Petropoulos. Experimental studies in relation to a new theoretical description of the permeation of dilute adsorbable gases through porous membranes. J.Membr.Sci., 12 (1983) 303.

[67] M. Jia, K. Peinemann, and R. Behling. Ceramic zeolite composite membranes. Preparation, characterization and gas permeation. J.Membr.Sci., 82 (1993) 15.

[68] F. Cao, C. Zhang, Y. Xiao, H. Huang, W. Zhang, D. Liu, et al. Helium recovery by a Cu-BTC metal-organic-framework membrane. Ind. Eng. Chem. Res., 51 (2012) 11274.

[69] J.E. Koresh, A. Sofer. Molecular sieve carbon permselective membrane. Part I. Presentation of a new device for gas mixture separation. Sep.Sci.Technol., 18 (1983) 723.

[70] D.M. D'Alessandro, B. Smit, and J.R. Long. Carbon dioxide capture: prospects for new materials. Angew. Chem. Int. Ed., 49 (2010) 6058.

[71] J. Lange. Methanol synthesis: a short review of technology improvements. Catal. Today, 64 (2001) 3.

[72] S.D. Kenarsari, D. Yang, G. Jiang, S. Zhang, J. Wang, A.G. Russell, et al. Review of recent advances in carbon dioxide separation and capture. RSC Adv., 3 (2013) 22739.

[73] S. Basu, A.L. Khan, A. Cano-Odena, C. Liu, and I.F. Vankelecom. Membrane-based technologies for biogas separations. Chem.Soc.Rev., 39 (2010) 750.

[74] H. Lin, E. Van Wagner, B.D. Freeman, L.G. Toy, and R.P. Gupta. Plasticization-enhanced hydrogen purification using polymeric membranes. Science, 311 (2006) 639.

[75] H. Lin, Z. He, Z. Sun, J. Vu, A. Ng, M. Mohammed, et al. CO2-selective membranes for hydrogen production and CO2 capture – Part I: Membrane development. J.Membr.Sci., 457 (2014) 149.

[76] Y. Liu, G. Zeng, Y. Pan, and Z. Lai. Synthesis of highly c-oriented ZIF-69 membranes by secondary growth and their gas permeation properties. J.Membr.Sci., 379 (2011) 46.

Page 69: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

54

[77] R.W. Baker, K.A. Lokhandwala, M.L. Jacobs, and D.E. Gottschlich. Recover feedstock and product from reactor vent streams. Chem.Eng.Prog., 96 (2000) 51.

[78] R.W. Baker. Future directions of membrane gas separation technology. Ind. Eng. Chem. Res., 41 (2002) 1393.

[79] Y. Huang, D.R. Paul. Physical aging of thin glassy polymer films monitored by gas permeability. Polymer, 45 (2004) 8377.

[80] M. Zhou, D. Korelskiy, P. Ye, M. Grahn, and J. Hedlund. A uniformly oriented MFI membrane for improved CO2 separation. Angew. Chem. Int. Ed., 53 (2014) 3492.

[81] M. Zhou, J. Hedlund. Oriented monolayers of submicron crystals by dynamic interfacial assembly. J. Mater. Chem., 22 (2012) 3307.

[82] D. Korelskiy, M. Grahn, J. Mouzon, and J. Hedlund. Characterization of flow-through micropores in MFI membranes by permporometry. J. Membr. Sci., 417-418 (2012) 183.

[83] D. Korelskiy, P. Ye, H. Zhou, J. Mouzon, and J. Hedlund. An experimental study of micropore defects in MFI membranes. Microporous Mesoporous Mater., 186 (2014) 194.

[84] F.R. Siperstein. Determination of azeotropic behavior in adsorbed mixtures. Adsorption, 11 (2005) 55.

[85] T. Nitta, T. Shigetomi, M. Kuro-oka, and T. Katayama. Adsorption isotherm of multi-site occupancy model for homogeneous surface. J.Chem.Eng.Japan, 17 (1984) 39.

[86] D.D. Do, Adsorption analysis: equilibria and kinetics, Singapore, World Scientific, 1998.

[87] G. Sethia, R.S. Pillai, G.P. Dangi, R.S. Somani, H.C. Bajaj, and R.V. Jasra. Sorption of methane, nitrogen, oxygen, and argon in ZSM-5 with different SiO2/Al2O3 ratios: Grand canonical monte carlo simulation and volumetric measurements. Ind. Eng. Chem. Res., 49 (2010) 2353.

[88] J. Hedlund, M. Grahn, D. Korelskiy, M. Rayson, S. Öberg, and P.R. Briddon. Mass transport in porous media from first principles: An experimental and theoretical study. J.Membr.Sci., 415-416 (2012) 271.

[89] M. Grahn, J. Hedlund. Maxwell–Stefan modeling of high flux tubular silicalite-1 membranes for CO2 removal from CO2/H2 gas mixtures. J.Membr.Sci., 471 (2014) 328.

[90] R. Krishna, J.M. van Baten. Investigating the influence of diffusional coupling on mixture permeation across porous membranes. J.Membr.Sci., 430 (2013) 113.

[91] R. Krishna. The Maxwell-Stefan description of mixture diffusion in nanoporous crystalline materials. Microporous Mesoporous Mater., 185 (2014) 30.

[92] R. Krishna, J.M. van Baten. In silico screening of zeolite membranes for CO2 capture. J.Membr.Sci., 360 (2010) 323.

Page 70: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

55

[93] A.M. Avila, H.H. Funke, Y. Zhang, J.L. Falconer, and R.D. Noble. Concentration polarization in SAPO-34 membranes at high pressures. J.Membr.Sci., 335 (2009) 32.

[94] J. Jiang, S.I. Sandier. Nitrogen and oxygen mixture adsorption on carbon nanotube bundles from molecular simulation. Langmuir, 20 (2004) 10910.

[95] J.R. Arnold. Adsorption of gas mixtures. Nitrogen-oxygen on anatase. J.Am.Chem.Soc., 71 (1949) 104.

[96] G. Arora, S.I. Sandler. Air separation by single wall carbon nanotubes: Thermodynamics and adsorptive selectivity. J.Chem.Phys., 123 (2005) 044705.

[97] T. Burdyny, H. Struchtrup. Hybrid membrane/cryogenic separation of oxygen from air for use in the oxy-fuel process. Energy, 35 (2010) 1884.

[98] W.J.W. Bakker, L.J.P. Van Den Broeke, F. Kapteijn, and J.A. Moulijn. Temperature Dependence of One-Component Permeation through a Silicalite-1 Membrane. AIChE J., 43 (1997) 2203.

[99] M. Kanezashi, J. O'Brien‐Abraham, Y. Lin, and K. Suzuki. Gas permeation through DDR‐type zeolite membranes at high temperatures. AIChE J., 54 (2008) 1478.

[100] W.J.W. Bakker, F. Kapteijn, J. Poppe, and J.A. Moulijn. Permeation characteristics of a metal-supported silicalite-1 zeolite membrane. J.Membr.Sci., 117 (1996) 57.

[101] R.M. de Vos, H. Verweij. High-selectivity, high-flux silica membranes for gas separation. Science, 279 (1998) 1710.

[102] H. Li, Z. Song, X. Zhang, Y. Huang, S. Li, Y. Mao, et al. Ultrathin, molecular-sieving graphene oxide membranes for selective hydrogen separation. Science, 342 (2013) 95.

[103] T. Golden, S. Sircar. Gas adsorption on silicalite. J. Colloid. Interface Sci., 162 (1994) 182.

[104] L.J.P. Van Den Broeke, W.J.W. Bakker, F. Kapteijn, and J.A. Moulijn. Binary permeation through a silicalite-1 membrane. AIChE J., 45 (1999) 976.

[105] R.W. Baker, J.G. Wijmans, and J.H. Kaschemekat. The design of membrane vapor-gas separation systems. J.Membr.Sci., 151 (1998) 55.

[106] I. Pinnau, Z. He. Pure- and mixed-gas permeation properties of polydimethylsiloxane for hydrocarbon/methane and hydrocarbon/hydrogen separation. J.Membr.Sci., 244 (2004) 227.

[107] H.H. Park, J. Won, S.G. Pyo, and H. Park. Effect of palladium ion on facilitated olefin transport through silver-polymer complex membranes. Macromol.Res., 19 (2011) 1077.

Page 71: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

56

[108] X. Ma, B.K. Lin, X. Wei, J. Kniep, and Y.S. Lin. Gamma-alumina supported carbon molecular sieve membrane for propylene/propane separation. Ind. Eng. Chem. Res., 52 (2013) 4297.

[109] S.H. Choi, J.H. Kim, and S.B. Lee. Sorption and permeation behaviors of a series of olefins and nitrogen through PDMS membranes. J.Membr.Sci., 299 (2007) 54.

[110] R.W. Baker, K.A. Lokhandwala, and I. Pinnau, Ethylene/nitrogen separation process, U.S. patent, 5,879,431, 1999.

[111] J. Peng, H. Ban, X. Zhang, L. Song, and Z. Sun. Binary adsorption equilibrium of propylene and ethylene on silicalite-1: Prediction and experiment. Chem.Phys.Lett., 401 (2005) 94.

[112] R.W. Elliott, H. Watts. Diffusion of some Hydrocarbons in Air: a Regularity in the Diffusion Coefficients of a Homologous Series. Can. J. Chem., 50 (1972) 31.

[113] F. Kapteijn, W.J.W. Bakker, G. Zheng, J. Poppe, and J.A. Moulijn. Permeation and separation of light hydrocarbons through a silicalite-1 membrane. Application of the generalized Maxwell-Stefan equations. Chem.Eng.J.Biochem.Eng.J., 57 (1995) 145.

Page 73: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 74: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Microporous and Mesoporous Materials 192 (2014) 14–17

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials

journal homepage: www.elsevier .com/locate /micromeso

Air separation at cryogenic temperature using MFI membranes

1387-1811 � 2013 The Authors. Published by Elsevier Inc.http://dx.doi.org/10.1016/j.micromeso.2013.09.016

⇑ Corresponding author. Tel.: +46 920493161.E-mail address: [email protected] (P. Ye).

Open access under CC BY-NC-ND license.

Pengcheng Ye ⇑, Erik Sjöberg, Jonas HedlundChemical Technology, Luleå University of Technology, SE-971 87 Luleå, Sweden

a r t i c l e i n f o a b s t r a c t

Article history:Available online 18 September 2013

Keywords:Zeolite membraneCryogenic temperatureAir separation

In the present work, zeolite membranes were for the first time evaluated for separations at cryogenictemperatures. MFI membranes were evaluated with a feed of synthetic air, at varying feed pressuresbetween 1 and 5 bar and a sweep stream of helium at a constant pressure of 1 bar for temperatures inthe range 77–110 K. When the feed pressure was 1 bar, the highest O2/N2 separation factor was 3.9, cor-responding to a separation selectivity of 4.1, with an oxygen permeance of 6.7 � 10�7 mol m�2 s�1 Pa�1 atthe optimum temperature 79 K. This membrane performance is just above the upper bound in the 2008Robeson selectivity–permeability plot for polymeric membranes. As the feed pressure increased, themaximum separation factors for O2/N2 decreased while the optimum separation temperatures increased.It is inferred that the separation is governed by condensation and effective transport of oxygen in the zeo-lite pores under these conditions. However, the adsorption of nitrogen, and thereby the transport of nitro-gen, likely increases as the pressure is increased, which reduce the selectivity. To test this hypothesis, thefeed was diluted with 33% helium and the total feed pressure was maintained at 1 bar, corresponding to apartial pressure of air of 0.66 bar. In this case, the maximum separation factor increased to 4.3, whichsupports the proposed separation mechanism.

� 2013 The Authors. Published by Elsevier Inc. Open access under CC BY-NC-ND license.

1. Introduction

The current air separation technologies include cryogenic distil-lation, pressure swing adsorption (PSA), membrane separation andchemical absorption [1]. Cryogenic distillation, with high invest-ment costs, is a well-developed and established technique forlarge-scale production (typically larger than 10 MMscfd of gas) ofpure oxygen. PSA processes are usually applied in smaller scale(typically 1–10 MMscfd), when it is sufficient with lower purity,typically around 95 vol.% oxygen. Chemical absorption processmay be a rather competitive alternative if the facility corrosioncould be reduced. Polymeric membranes have been developedfor separation of oxygen from air. Materials that have been usedfor air separation include poly(1-trimethylsilyl-1-1propyne)(PTMSP), Teflon AF 2400, silicone rubber, poly(4-methyl-1-pen-tene) (TPX), poly(phenylene oxide) (PPO), ethyl cellulose, 6FDA-DAF (polyimde), polysulfone, polyaramide, etc. [2]. Robesonplotted reported permeances and selectivities for a large numberof polymeric membranes of various types and illustrated that thereis a trade-off between selectivity and permeance. Polymers withhigh selectivity have relatively low permeance, and vice versa.

Polymers with the best combinations of selectivity and perme-ances constitute what is known as the Robeson selectivity–perme-ability upper bound [3]. However, the most permeable polymerswith sufficient selectivity are preferred in gas separationapplications.

Unlike polymeric membranes, zeolite membranes are inorganicand porous and have gained great interest for gas separations dueto their potential for high permeances and fluxes as well as highthermal, chemical and mechanical stability compared to polymericmembranes. Our research group has developed zeolite membranescomprised of thin (<500 nm) MFI films on open graded aluminasupports, with high flux and high selectivity [4]. These membraneshad 10–100 times higher permeances than previous reportedmembranes, and still the separation factors were similar.

There are only a few reports in the scientific literature describ-ing gas separations at low temperatures using zeolite membranes.Hong et al. [5] investigated CO2/H2 separations by SAPO-34 mem-branes at temperatures in the range of 253–308 K. At 253 K, theCO2/H2 separation selectivity was as high as 110 but the CO2 fluxwas only 3 kg/(m2 h) even though the trans-membrane pressuredifference was as high as 12 bar. van den Bergh et al. [6] studiedsingle component and mixture permeation of various gases,including CO2, CH4, N2 and O2, through DDR membranes at a tem-perature of 220–373 K and a feed pressure of 101–400 kPa. Themixture selectivity at low temperatures was much higher thanthe ideal selectivity for all mixtures containing CO2. For instance,for temperatures below 250 K, the CO2/CH4 separation factor was

Page 75: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Fig. 1. Top view (a) and cross-sectional (b) SEM images an as-synthesised MFImembrane.

P. Ye et al. / Microporous and Mesoporous Materials 192 (2014) 14–17 15

higher than 1000. However, only modest oxygen selectivity (ca 2)for mixtures of O2 and N2 was observed and it was reported thatthe flux was ‘‘very low’’ in all cases. Piera et al. [7] used an MFImembrane to remove traces (160 ppbv) of CO from air. At 245 K,a CO/air separation factor up to 3 was reported. To our best knowl-edge, there are no publications on separations using zeolite mem-branes at cryogenic temperatures, i.e. below 123 K, in the openliterature.

In the present work, zeolite membrane separation was for thefirst time evaluated at cryogenic temperatures, for air separation.

2. Experimental section

2.1. Membrane preparation and characterization

MFI membranes with a film thickness of about 0.4 lm wereprepared using a previously reported method [4] and only a briefdescription of the preparation procedure is given here. A gradedporous a-alumina disc (Fraunhofer IKTS, Germany), with a diame-ter of 25 mm diameter is used as support. It is comprised of a30 lm thick top layer with a pore size of 100 nm and a 3 mm thickbase layer with a pore size of 3 lm. The alumina disc was maskedand seeded with the procedure described in detail elsewhere [8].The seeded support was immersed in a synthesis solution with amolar composition of 3TPAOH:25SiO2:1450H2O:100C2H5OH andheated in an oil bath kept at 88 �C for 72 h, with reflux of evapo-rated solution. Afterwards, the membrane was rinsed in a 0.1NH3 solution overnight and calcined at 500 �C for 6 h, with a heat-ing rate of 0.2 �C/min and a cooling rate of 0.3 �C/min.

The morphology of the membranes was investigated by Scan-ning Electron Microscopy (SEM) using an FEI Magellan 400 fieldemission XHR-SEM.

2.2. Permeation measurements

The membrane was mounted in a stainless steel cell, sealedwith graphite gaskets. A type T thermocouple inside the cell wasused to measure the temperature. The membrane was dried in he-lium flow at 300 �C for 6 h before the permeation measurements.

First, adsorption-branch n-hexane/helium permporometry[9,10] was used to characterize the amount of flow-through defectsin the membranes. After the permporometry experiment, themembrane was dried again using the same procedure as that de-scribed above.

In the next step, O2 and N2 single gas permeation experimentswere carried out. The feed pressure of O2 (>99.999%, AGA) or N2

(>99.999%, AGA) was 1 bar with a helium (>99.999%, AGA) sweepgas at 1 bar pressure. The feed flow rate was 1500 ml (STP)/minand the sweep flow rate was 450 ml (STP)/min. The membrane cellwas kept in a Dewar for liquid nitrogen (KGW, Germany) and liquidnitrogen was slowly added to the Dewar until the membrane tem-perature reached 77 K. After terminating the addition of liquidnitrogen to the Dewar, the temperature in the cell increased veryslowly, less than 0.2 K/min. The permeate flow was measured bya drum type gas meter (Ritter Apparatebau GmbH) and the compo-sition was analysed by mass spectrometry (GAM 400, InprocessInstruments) every 0.5 min.

In addition, binary separation experiments was carried out byfeeding synthetic air (purity > 99.999%, AGA), consisting of 79.2%N2 and 20.8% O2, to the membrane at varying feed pressures rang-ing from 1 to 5 bar. To the permeate side of the membrane, a sweepstream of helium at a constant pressure of 1 bar, was fed. In onecase, the feed of synthetic air was diluted with 33% helium. Theseparation factor was estimated from the measured composition:

a ¼ ðyi=yjÞ=ðxi=xjÞ ð1Þ

where yi and xi are molar fractions of species i in the permeate andfeed streams, respectively.

The separation selectivity (also referred as permeance-basedselectivity) was determined as:

ap ¼ Pi=Pj ð2Þ

where Pi and Pj are permeance of species i and j in the permeatestream, respectively.

3. Results and discussion

Fig. 1 shows top view (a) and cross-sectional (b) SEM micro-graphs of one membrane prepared in the same batch as the mem-brane used for separation experiments. The micrographs illustratethat the film is comprised of well intergrown crystals and that it isabout 400 nm thick.

Table 1 shows the permporometry results for the membraneused for separations in the present work. The total amount of de-fects is only 0.60% of the total membrane surface area, and only1.6% of these defects are larger than 1.04 nm. Moreover, almostno defects larger than 4.3 nm can be detected. The defect distribu-tion for this membrane is consistent with that of high qualitymembranes reported previously [10].

Fig. 2a a shows single gas permeation data for O2 and N2 atcryogenic temperatures. The boiling points for N2 and O2 at 1 barpressure (as in the experiment) are 77.4 and 90.2 K, respectively.The oxygen feed must be in liquid state at temperatures below90.2 K and the oxygen single ‘‘gas’’ experiment is actually a

Page 76: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Table 1Helium permeance through defects and relative areas of defects for the membrane, estimated from permporometry data.

P/P0 Permeance via defects (mol m�2 s�1 Pa�1) Defect width (nm) Defect interval (nm) Relative area of defectsa (%)

2.26 � 10�4 8.15 � 10�7 0.71 – –3.72 � 10�4 6.08 � 10�7 0.73 0.71–0.73 0.221.12 � 10�3 2.60 � 10�7 0.80 0.73–0.80 0.301.13 � 10�2 1.19 � 10�7 1.04 0.80–1.04 0.071.12 � 10�1 0.74 � 10�7 1.78 1.04–1.78 0.0113.49 � 10�1 0.62 � 10�7 4.23 1.78–4.31 0.0013

>4.31 0.0045Total 0.60

a Area of defects divided by the membrane area.

Fig. 2. (a) Single gas permeances for O2 and N2 at cryogenic temperatures with afeed pressure of O2 and N2 respectively at 1 bar and a helium sweep gas at 1 barpressure. (b) Permeances and separation factors for O2/N2 at cryogenic tempera-tures with a feed of pure air at 1 bar pressure and a helium sweep gas at 1 barpressure.

16 P. Ye et al. / Microporous and Mesoporous Materials 192 (2014) 14–17

pervaporation experiment at the lowest temperatures investigatedhere. We speculate that both O2 and N2 are condensing in the zeo-lite pores (and already in the membrane cell in case of oxygen atthe lowest temperatures) at temperatures lower than 100 and103 K, respectively, since the permeance of both molecules isquite low and nearly constant and only varies between 3–6and 2–4 � 10�7 mol m�2 s�1 Pa�1, respectively. For the sake ofsimplicity, the terms condense/condensation are used to describethe phenomena that occurs when molecules adsorb in the poresin high concentration, with a density close to that of the bulkliquid. However, at temperatures exceeding 100 and 103 K for O2

and N2, respectively, the permeance is increasing rapidly with

temperature, which is indicating a distinctively different diffusionmechanism, perhaps not involving condensed, but only adsorbedspecies.

Fig. 2b illustrates the results of the separation experiments witha feed of synthetic air at 1 bar pressure and a sweep stream of he-lium at a pressure of 1 bar. In this case, the feed flow rate was1500 ml (STP)/min and the sweep flow rate was 450 ml (STP)/min. In a narrow temperature range, the membrane was quite oxy-gen selective. A maximum O2/N2 separation factor of 3.9 with anoxygen permeance of 6.7 � 10�7 mol m�2 s�1 Pa�1 (similar as inthe single gas experiment and corresponding to 740 Barrer) wasobserved at the optimum temperature 79 K. The membrane perfor-mance is just above the upper bound in the 2008 Robeson selectiv-ity–permeability plot for polymeric membranes [3]. On the otherhand, the O2/N2 separation factor at 296 K is 0.95, with an oxygenpermeance of 24.7 � 10�7 mol m�2 s�1 Pa�1. The membrane isslightly N2 selective at room temperature, likely because nitrogenadsorbs somewhat more than oxygen, and adsorption (rather thancondensation) controls the separation in this case. The O2/N2 sep-aration factors at cryogenic temperatures are higher, likely dueto condensation effects, as compared to those at room tempera-ture. At the same time, the permeances at cryogenic temperatureare only a few times lower at cryogenic temperatures as comparedto those at room temperature. The diffusivities are probably muchlower, but this seems partially counteracted by much higher con-centrations of (condensed) species in the pores at cryogenictemperatures.

Fig. 3 shows that, the maximum O2/N2 separation factors, ob-served at the optimum separation temperature at each pressure,decreased significantly as the feed pressure was increased from 1to 5 bar. At the same time, the optimum separation temperatures,at which the maximum separation factor was observed, increasedas the feed pressure was increased. For instance, when the feedpressure was 5 bar, the maximum separation factor was 2.7 atthe optimum temperature 95 K. These observations are consistentwith a separation mechanism governed by selective condensationand transport of oxygen in the zeolite pores, and at the same time,the adsorption of nitrogen should be limited. However, nitrogenadsorbs a little more than oxygen in MFI pores. Isosteric heats ofadsorption for nitrogen and oxygen on ZSM-5 (Si/Al ratio = 80)are 22.2 kJ/mol and 14.1 kJ/mol, respectively. At 303.15 K, Henry’sconstant for nitrogen adsorption is 2.8 times higher than that ofoxygen on ZSM-5 with an Si/Al ratio of 80, and 1.1 times onZSM-5 with an Si/Al ratio of = 200 [11]. The membrane preparedin the present work have a Si/Al ratio of about 139 [12], i.e. in-be-tween these two values. Therefore, when the feed pressure is in-creased, the adsorption of nitrogen should increase more thanthat of oxygen and, consequently, the transport of nitrogen shouldincrease, and the separation factor should decrease, as observedexperimentally. To test the suggested separation mechanism, astream of synthetic air diluted with 33.3% helium was fed to themembrane at feed pressure of 1 bar, to arrive at a partial air

Page 77: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Fig. 3. (a) Optimum separation temperatures and maximum O2/N2 separationfactors as a function of feed pressure of synthetic air. (b) Permeances of O2 and N2 asa function of feed pressure at optimum separation temperatures.

P. Ye et al. / Microporous and Mesoporous Materials 192 (2014) 14–17 17

pressure of 0.66 bar. The maximum separation factor increased to4.3 at 77.4 K, as compared to 3.9 at 79 K at an air pressure of1 bar. The increased separation factor upon helium dilution sup-ports the suggested separation mechanism.

Fig. 3b shows permeances of oxygen and nitrogen as a functionof pressure at the optimum separation temperature. When the feedpressure increases from 1 to 5 bar, the permeance at the (increas-ing) optimum separation temperature increase, likely due to high-er diffusivity at higher temperature. Meanwhile, the permeancesare higher at a (partial) feed pressure of 0.66 bar compared to1 bar. The observed oxygen permeance is 61% higher at a partialfeed pressure of 0.66 bar, than at 1 bar. This is consistent withthe proposed separation mechanism, i.e. condensation of oxygenin the pores, resulting in the same concentration of liquid oxygen,and thereby transport, of oxygen trough the pores at both pres-sures, which should result in 51% higher oxygen permeance(0.66�1 = 1.51), which is close to the observed increase. The ob-served nitrogen permeance is 43% higher at a partial feed pressureof 0.66 bar, than at 1 bar. The higher permeance could be caused bya similar effect as that proposed for oxygen, but instead caused by

that the adsorption of nitrogen is nearly constant, due to the con-stant condensation of oxygen, at both a partial feed pressure of0.66 bar and a pressure of 1 bar. However, in the whole (partial)pressure range from 0.66–5 bar, the separation selectivity, whichis the ratio of oxygen permeance over nitrogen permeance, de-creases when the feed pressure increase, i.e. the nitrogen perme-ance is increasing more than the oxygen permeance when thepressure is increased, which supports the proposed separationmechanism.

In the present work, the membrane was cooled down fromroom temperature to cryogenic temperatures, and warmed upagain several times. The initial separation performance remainedafter many cycles during a period of 3 months, which indicatesgood mechanical resistance upon cooling.

4. Conclusion

Zeolite membranes were for the first time evaluated for separa-tion at cryogenic temperatures, for air separation. At each feedpressure, the membranes were quite oxygen selective in a narrowtemperature range. The membrane separation performance justexceeded the upper bound in the Robeson plot for polymeric mem-branes. As the feed pressure increased, the maximum separationfactors for O2/N2 decreased, while the optimum separation temper-atures increased. It is inferred that oxygen condenses in, and istransported effectively through, the zeolite pores. At the sametime, the adsorption and transport of nitrogen is limited, likelysuppressed by the condensed oxygen. However, as the pressureis increased, the adsorption of nitrogen increases more than thatof oxygen, which is accompanied by higher nitrogen permeanceand reduced selectivity.

Acknowledgment

The authors are grateful to the Swedish Research Council andthe Swedish Foundation for Strategic Research for financialsupport.

References

[1] A.R. Smith, J. Klosek, Fuel Process. Technol. 70 (2001) 115–134.[2] R.W. Baker, Membrane Technology and Applications, second ed., Wiley,

Chichester, 2004.[3] L.M. Robeson, J. Membr. Sci. 320 (2008) 390–400.[4] J. Hedlund, J. Sterte, M. Anthonis, A. Bonsb, B. Carstensenc, N. Corcoranc, D.

Coxc, H. Deckmanc, W.D. Gijnstb, P. de Moorb, F. Laic, J. McHenryc, W.Mortierb, J. Reinosod, J. Peterse, Microporous Mesoporous Mater. 52 (2002)179–189.

[5] M. Hong, S.G. Li, J.L. Falconer, R.D. Noble, J. Membr. Sci. 307 (2008) 277–283.[6] J. van den Bergh, W. Zhu, J. Gascon, J.A. Moulijn, F. Kapteijn, J. Membr. Sci. 316

(2008) 35–45.[7] E. Piera, C.A.M. Brenninkmeijer, J. Santamaria, J. Coronas, J. Membr. Sci. 201

(2002) 229–232.[8] J. Hedlund, F. Jareman, A.J. Bons, J. Membr. Sci. 222 (2003) 163–179.[9] J. Hedlund, D. Korelskiy, L. Sandström, J. Membr. Sci. 345 (2009) 276–287.

[10] D. Korelskiy, M. Grahn, J. Mouzon, J. Hedlund, J. Membr. Sci. 417–418 (2012)183–192.

[11] G. Sethia, R.S. Pillai, G.P. Dangi, R.S. Somani, H.C. Bajaj, R.V. Jasra, Ind. Eng.Chem. Res. 49 (2010) 2353–2362.

[12] L. Sandström, J. Lindmark, J. Hedlund, J. Membr. Sci. 360 (2010) 265–275.

Page 78: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

PAPER II

Cryogenic Air Separation at Low Pressure Using MFI Membranes

Pengcheng Ye, Danil Korelskiy, Mattias Grahn and Jonas Hedlund

Journal of Membrane Science, 487 (2015) 135-140.

Page 79: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 80: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Cryogenic air separation at low pressure using MFI membranes

Pengcheng Ye n, Danil Korelskiy, Mattias Grahn, Jonas HedlundChemical Technology, Luleå University of Technology, SE-971 87 Luleå, Sweden

a r t i c l e i n f o

Article history:Received 27 January 2015Accepted 15 March 2015Available online 4 April 2015

Keywords:Zeolite membraneCryogenic temperatureAir separationLow pressure

a b s t r a c t

Ultra-thin MFI membranes were for the first time evaluated for air separation at low feed pressuresranging from 100 to 1000 mbar at cryogenic temperature. The membrane separation performance atoptimum temperature at all investigated feed pressures was well above the Robeson upper bound forpolymeric membranes at near room temperature. The O2/N2 separation factor at optimum temperatureincreased as the feed pressure was decreased and reached 5.0 at 100 mbar feed pressure and amembrane temperature of 67 K. The corresponding membrane selectivity was 6.3, and the O2

permeance was as high as 8.6�10�7 mol m�2 s�1 Pa�1. This permeance was about 100 times higherthan that reported for promising polymeric membranes. The membrane selectivity and high O2

permeance was most likely a result of O2/N2 adsorption selectivity. The increase in O2/N2 separationfactor with decreasing pressure and temperature was probably due to increased adsorption selectivity atreduced temperature. This work has demonstrated the potential of MFI zeolite membranes for O2/N2

separations at cryogenic temperature.& 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND

license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

Development of more efficient methods for separation ofvarious gas mixtures are currently of tremendous interest. Mem-brane gas separation technology is one such method that isalready applied in industrial scale for H2 recovery, CO2 separation,dehydration of air and natural gas, and air separation [1]. For thelatter application, polymeric membranes have been developed forseparation of O2 from air. Most of these membranes show highselectivity but low permeability, i.e., low flux [2]. Robeson hasinvestigated the performance of different polymeric membranesand found a trade-off between the desired selectivity and perme-ability. The best polymeric membranes with high selectivity hadrelatively low permeability, and vice versa. This trade-off is knownas the Robeson selectivity–permeability upper bound [3] and itwas updated in 2008 [4]. It is here worth pointing out that forindustrial applications, materials with as high permeability aspossible, possessing sufficient selectivity, are of major interest[2]. In other words, a very selective membrane with very lowpermeability is not very industrially useful. Recently, Carta et al.[5] have reported microporous polymeric membranes with ratherhigh oxygen permeability of 900–1100 Barrer and good O2/N2

selectivities of 4.5–6.1.Zeolite membranes being inorganic porous membranes are

known to have high permeabilities in gas separations. Due to the

well-defined pore structure, zeolite membranes can also offer highselectivities. In addition, these membranes have high chemical,mechanical and thermal stability, i.e., can potentially be used atboth very high and very low temperatures offering a greatadvantage over polymeric membranes. High temperature applica-tions of zeolite membranes have been studied extensively. On thecontrary, scientific reports on gas separation using zeolite mem-branes at low temperatures are rather scarce. Hong et al. [6]studied CO2/H2 separation using SAPO-34 membranes in thetemperature range of 253–308 K. At 253 K, the reported CO2/H2

separation factor was as high as 110, however the CO2 flux wasonly 3 kg m�2 h�1 even though the trans-membrane pressuredifference was as high as 12 bar. Piera et al. [7] separated CO/airat an extremely low CO concentration of 160 ppbv using an MFI-type membrane in the temperature range of 243–303 K. Theseparation factor was 3.1 with an extremely low CO flux of1.3�10�7 kg m�2 h�1 at 245 K. Van den Bergh et al. [8] investi-gated permeation and separation properties of various gases (CO2,CH4, N2, and O2) and their mixtures, using DDR membranes in thetemperature range of 220–373 K. The mixture selectivitiesincreased as temperature was decreased, especially for CO2/CH4,CO2/air and N2O/air mixtures. The CO2/CH4 separation factor washigher than 1000 at temperatures below 250 K. At the same time,the O2/N2 selectivity was low amounting to only ca. 2 and the fluxwas reported as “very low”, i.e. lower than 7.8�10�2 kg m�2 h�1

in all cases. Our research group has developed ultra-thin(E0.5 μm) MFI membranes on open graded alumina supports[9]. Due to the low thickness and well-defined pore structure, themembranes can display both high flux and high selectivity. These

Contents lists available at ScienceDirect

journal homepage: www.elsevier.com/locate/memsci

Journal of Membrane Science

http://dx.doi.org/10.1016/j.memsci.2015.03.0630376-7388/& 2015 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

n Corresponding author. Tel.: þ46 920493161.E-mail address: [email protected] (P. Ye).

Journal of Membrane Science 487 (2015) 135–140

Page 81: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

membranes display excellent performance for CO2 separation fromsynthesis gas at high pressure and low temperature [10]. Theseparation factor for CO2/H2 was as high as 32 with a CO2 flux of332 kg m�2 h�1 at 275 K. We have also shown [11] that separationfactors exceeding 100 can be reached below 243 K. The greaterseparation performance of the membranes at lower temperaturewas attributed to stronger adsorption of CO2 in the zeolite pores,effectively blocking the transport of H2. Recently, we have eval-uated these membranes for air separation at cryogenic tempera-tures, i.e., below 123 K and the report [12] is the first on zeolitemembrane separation at cryogenic temperature. The experimentswere carried out in the temperature range of 77–110 K and thefeed pressure was varied between 1–5 bar. Liquid nitrogen wasused for cooling, which limited the lowest achievable experimen-tal temperature to 77 K, and no vacuum system was used, whichlimited the lowest feed and permeate pressures to 1 bar. Heliumwas used as a sweep gas on the permeate side at a pressureof 1 bar. The membranes were oxygen selective, and when thefeed pressure was decreased, the maximum O2/N2 separationfactor increased whereas the optimum separation temperaturesdecreased. At the lowest investigated feed pressure of 1 bar andthe optimum temperature of 79 K, the highest observed O2/N2

separation factor was 3.9, with an O2 permeance of 6.7�10�7 mol m�2 s�1 Pa�1. We have now installed a vacuum systemto enable experiments at reduced pressure, and a cryostat, whichenables experiments at temperatures down to 22 K. In the presentwork, the membranes are evaluated for air separation at signifi-cantly lower feed pressures and membrane temperatures than inthe previous work [12] using the new experimental set-up. Amodel for the adsorption in the zeolite pores was also developedto better understand the separation performance of the mem-branes.

2. Experimental

2.1. Membrane preparation

MFI membranes were prepared on graded porous α-aluminadiscs (Fraunhofer IKTS, Germany), using a previously reportedmethod [9]. The support had a diameter of 25 mm and wascomprised of a 30 μm thick top layer with a pore size of 100 nmand a 3 mm thick base layer with a pore size of 3 μm. The aluminadisc was masked and then seeded with a monolayer of MFIcrystals of 50 nm in size as described previously [13]. After that,a hydrothermal treatment was carried out, where the seededsupport was immersed in a synthesis solution with a molarcomposition of 3TPAOH:25SiO2:1450H2O:100C2H5OH and kept inthe solution at 361 K for 72 h. Thereafter, the membrane wasrinsed in a 0.1 NH3 solution overnight and calcined at 773 K for 6 hat a heating rate of 0.2 K min�1 and a cooling rate of 0.3 K min�1.

2.2. Membrane characterization

Adsorption-branch n-hexane/helium permporometry [14] wasused to characterize the amount of flow-through defects in themembranes. This technique has recently been demonstrated to bean effective and reliable way to evaluate micropore defects downto 0.7 nm [15,16] and mesopore defects [14] in MFI membranes.The procedure for the permporometry characterization isdescribed in our previous work [15].

The morphology of the membranes was characterized by SEMusing an FEI Magellan 400 field emission XHR-SEM without anycoating of the samples.

2.3. Gas separation measurements

The membrane was mounted in a stainless steel cell and sealedwith graphite gaskets (Eriks, the Netherlands). A type T thermo-couple inserted inside the cell was used to monitor the tempera-ture. Prior to the gas separation, the membrane was dried in ahelium flow at 573 K for 6 h at a heating rate of 1 K min�1.

The separation experiments were carried out using the experi-mental set-up illustrated in Fig. 1. A flow of 500 ml min�1

(purity499.999%, AGA) of synthetic air, consisting of 79.2% N2

and 20.8% O2, was fed to the membrane at feed pressures of100 mbar, 400 mbar, 660 mbar and 1000 mbar (at a feed pressureof 100 mbar, the feed flow rate was reduced to 100 ml min�1 dueto the limited capacity of the vacuum pump). A pressure of10 mbar, 40 mbar, 66 mbar, or 100 mbar, was applied on thepermeate side of the membrane, in order to keep the ratio of feedpressure to permeate pressure at 10. Since the pressure ratio mayaffect the separation performance, the pressure ratio was keptconstant in all experiments. The gas flow rate was controlled bymass flow controllers (Bronkhorst, F-201CV). Vacuum pumps(Pfeiffer Vacuum, MVP 020 and MVP 040) installed after thepressure regulators in the retentate and permeate lines reducedthe pressures on both sides of the membrane. The pressures werecontrolled by dosing valves (Pfeiffer Vacuum, EVN 116). Themembrane cell was kept in a custom-built cryostat (ICEoxfordUK, DRYICE25K) and the separation temperature could be set atany value ranging from 22 K to 573 K with a deviation of less than0.05 K. The permeate flow was measured with a mass flow meter(Bronkhorst, F-201CV). The permeate composition was analyzedby a mass spectrometer (GAM 400, InProcess Instruments) every5 s. A pressure transmitter (Pfeiffer Vacuum, CMR 362) was used tomeasure the permeate pressure. A column packed with 5A zeolitepellets was dried before the experiment and was used (at roomtemperature) to adsorb any traces of moisture in the feed gas.

The permeance Πi (mol m�2 s�1 Pa�1) was defined as

Π i ¼ Fi=ðAf ilmΔPiÞ; ð1Þwhere Fi is the molar flow rate of component i through themembrane, Afilm is the membrane area, and ΔPi is the partialpressure difference of component i.

The separation factor was determined as follows:

βi=j ¼ ðyi=yjÞ=ðxi=xjÞ; ð2Þ

where yi and xi are the molar fractions of species i in the permeateand feed streams, respectively.

The membrane selectivity (also referred to as the permeance-based selectivity) was calculated as

αi=j ¼Π i=Π j; ð3Þ

where Πi and Πj are the permeances of species i and j,respectively.

O2 and N2 single gas permeation experiments were also carriedout. A stream of 500 ml min�1 of O2 (499.999%, AGA) or N2

(499.999%, AGA) was fed to the membrane at 1000 mbar whilethe permeate side was kept at 100 mbar.

3. Modelling

In order to gain some insight into the experimental observa-tions, the adsorption of N2 and O2 was modelled. Heats ofadsorption and Langmuir adsorption parameters for N2 and O2 inhigh silica MFI [17] are given in Table 1. It may be expected thatthe membrane would be N2 adsorption selective because of thelarger heat of adsorption and Langmuir adsorption parametersthan those for O2. However, for mixture adsorption where the

P. Ye et al. / Journal of Membrane Science 487 (2015) 135–140136

Page 82: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

main difference between the molecules is the size, the isothermswill cross at a certain pressure. It is energetically favorable toadsorb smaller molecules at high loadings since smaller moleculespack more efficiently in the pore system at high loadings [18].Therefore, at pressures significantly higher than the pressurewhere the isotherms cross, a reversal in adsorption selectivitymay be observed, i.e. an adsorption azeotrope. To illustrate this,adsorption isotherms for O2 and N2 at 67 K and 90 K weresimulated using the multisite Langmuir model [19]. This modelis capable of capturing adsorption azeotropes [18]. The purecomponent adsorption isotherm is given by [20]

p¼ θnb 1�θ

� �n; ð4Þ

where p is the pressure of the adsorbate in gas phase, θ is thefractional coverage (adsorbed loading divided by the saturationloading), b is the Langmuir adsorption parameter, n is the numberof adsorption sites one adsorbate molecule occupies. The satura-tion capacity for N2 was taken from [21] and that of O2 was takenas 1.16 times that of N2 because the volume of an O2 molecule is ca.1.16 times smaller than that of a N2 molecule, as judged from thekinetic diameters of the molecules [3]. b-Values and heat ofadsorption reported by Sethia et al. [17] were extrapolated to thetemperatures of interest in this work. n was taken as 3 for O2 and

3�1.16¼3.48 for N2 because a N2 molecule is ca. 1.16 times largerthan a O2 molecule. The multicomponent version of the multisiteLangmuir model is given by [20]

θi ¼ nibipyi 1� ΣN

j ¼ 1θj

� �ni

; ð5Þ

where index i refers to species i in the mixture and yi is the gasmole fraction of species i.

4. Results and discussion

Table 2 shows permporometry data for the membrane used inthis work. The initial helium permeance at p/p0¼0, i.e., the heliumpermeance measured in the absence of n-hexane, was about43�10�7 mol m�2 s�1 Pa�1. This permeance is about half of thatfor our best membranes comprised of a film with a total thicknessof about 500 nm [9] and the lower permeance indicates that the

Table 1Heat of adsorption and Langmuir adsorption parameters for N2 and O2 [17].

ΔHads (kJ/mol) b (bar�1) at 303 K

N2 �17.0 0.277O2 �16.3 0.147

Table 2Helium permeance at equilibrium and relative areas of defects for the membrane,estimated from permporometry data.

p/p0 Helium permeance atequilibrium(10�7 mol m�2 s�1 Pa�1)

Defectwidth(nm)

Defectinterval(nm)

Relative areaof defects a

(%)

0.00 43.15 – – –

2.72�10�4 2.86 0.71 – –

4.55�10�4 2.07 0.73 0.71–0.73 0.071.39�10�3 0.95 0.80 0.73–0.80 0.081.13�10�2 0.34 1.04 0.80–1.04 0.041.39�10�1 0.19 1.78 1.04–1.78 0.0054.31�10�1 0.14 5.14 1.78–5.14 0.001

45.14 0.0008Total 0.20

a Area of defects divided by the membrane area.

Fig. 1. Gas permeation and separation set-up.

P. Ye et al. / Journal of Membrane Science 487 (2015) 135–140 137

Page 83: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

total thickness of the film in the membrane used in the presentwork exceeds 500 nm. However, the defect distribution for thismembrane is consistent with that previously reported for highquality membranes prepared by our group [15]. The total amountof defects was only 0.20% of the total membrane surface areaindicating a high quality of the membrane. The majority of defects,i.e., 97% of the total amount, were micropore defects smaller than1 nm in size. As discussed previously [16], these defects are mostlikely narrow open grain boundaries. In addition, nearly no defectslarger than 5 nm in size could be detected.

Fig. 2 shows top view (a) and cross-sectional (b) SEM micro-graphs of an as-synthesized membrane prepared in the samebatch as the membrane used for the separation experiments. Themicrographs illustrate that the total film thickness, includingsupport invasion, varies in the range 0.5–1 mm, i.e. the averagefilm thickness exceeds 500 nm, in line with permporometryresults. Furthermore, the film is comprised of well-intergrowncrystals and no defects can be observed, which also is in-line withpermporometry observations.

Fig. 3 shows the simulated single component adsorption isothermsat 67 K and 90 K and it is clear that N2 shows larger affinity than O2 forthe zeolite at low loadings at both temperatures. However, theisotherms of N2 and O2 cross at pressures of ca. 8�10�9 and9�10�6 bar at 67 and 90 K, respectively. The partial pressures inthe experiments were orders of magnitude higher, and consequently areversal of the adsorption selectivity, i.e. an adsorption azeotrope,

favoring the adsorption of O2 may be expected in the membraneseparation experiments. The multicomponent version of the multisiteLangmuir model was used to estimate the adsorption selectivity onthe feed side of the membrane, for the same composition as the feedand a total pressure of 100mbar. The adsorbed loadings werecalculated from Eq. (5), and thereafter the O2/N2 adsorption selectiv-ities were determined to be 4.2 and 1.8 at 67 and 90 K, respectively.Note that the adoption selectivity increased when the temperaturewas decreased under the same feed pressure.

Fig. 4 shows the separation results using different feed pressuresover the temperature range of 62–106 K. The membrane is oxygenselective as expected from the estimated O2/N2 adsorption selectivity.At each feed pressure, a maximum separation factor was observed atan optimum temperature and the separation factor was approachingone at the highest and lowest temperatures studied at all feedpressures. As shown in Table 3, synthetic air condensed at 82 K and66 K at 1000 and 100 mbar feed pressure, respectively. Condensationof the feed mixture was probably the reason for the drastic drop ofseparation factor at the lowest investigated temperatures. In order toelucidate these observations, single gas permeation experimentswith pure O2 and N2 were performed, see Fig. 5. The permeance ofN2 was higher than that of O2 over the entire temperature rangedespite the fact that the adsorption should favor O2. This suggeststhat diffusivity of N2 should be greater than that of O2 at alltemperatures studied and that the membrane selectivity shouldmostly emanate from adsorption selectivity. In addition, at the lowest

Fig. 2. Top view (a) and cross-sectional (b) SEM images of an as-synthesized MFI membrane.

Fig. 3. Simulated single component adsorption isotherms of O2 (red: 67 K, black: 90 K) and N2 (blue: 67 K, green: 90 K). (For interpretation of the references to color in thisfigure legend, the reader is referred to the web version of this article.)

P. Ye et al. / Journal of Membrane Science 487 (2015) 135–140138

Page 84: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

temperatures, the single component permeances of both O2 and N2

were very low (i.e., below 8.0�10�7 mol m�2 s�1 Pa�1). This isprobably a result of condensation of the entire feed at these lowtemperatures. As shown in Table 3, O2 condense at 90 K and N2 at77 K at 1 bar feed pressure in single component experiments.

At the lowest feed pressure of 100 mbar, a maximum O2/N2

separation factor of 5.0 (corresponding to a selectivity of 6.3)with an O2 permeance of 8.6�10�7 mol m�2 s�1 Pa�1 (corre-sponding to 1025 Barrer) was observed at the optimum

temperature of 67 K. This membrane performance is well abovethe 2008 upper bound in the Robeson selectivity–permeabilityplot for polymeric membranes at room temperature [4], seeFig. 6. The separation performance at the other investigated feedpressures (400 mbar, 660 mbar and 1000 mbar) also lies wellabove the upper bound. The performance of our membrane (interms of selectivity and permeability) is comparable to thatrecently reported by Carta et al. [5] for novel polymeric mem-branes. However, it is worth pointing out that the measured O2

permeance for our thin MFI membrane is more than 100 timeslarger than that for the polymeric membranes (lower than5.5�10�9 mol m�2 s�1 Pa�1) with a thickness ranging from95 to 181 mm [5].

Fig. 7a shows the maximum O2/N2 separation factors observedat the optimum separation temperature at each feed pressure. Themaximum O2/N2 separation factors increased significantly as thefeed pressure was decreased from 1000 mbar to 100 mbar. Inaddition, the optimum separation temperatures decreased as thefeed pressure was increased. For instance, the maximum separa-tion factor was only 3.7 at the optimum temperature of 89 K whenthe feed pressure was 1000 mbar, whereas the maximum separa-tion factor observed at 67 K and at a feed pressure of 100 mbarwas 5.0. As discussed above, the separation is most probablygoverned by O2/N2 adsorption selectivity, and this selectivity ishigher at low temperatures.

Fig. 7b shows permeances of O2 and N2 measured in the gasmixture separation experiments at the optimum separationtemperatures as a function of pressure. The figure also showsO2/N2 membrane selectivities at the optimum separation tem-peratures. The permeance of O2 was higher than that of N2 inthe entire feed pressure range. Accordingly, the O2/N2 mem-brane selectivities, defined as the ratio of O2 permeance over N2

permeance, were above unity at all feed pressures, which showsthat the membrane was oxygen selective. Both O2 and N2

permeances observed in the present work showed no strictcorrelation with the feed pressures. However, the membraneselectivity increased as the feed pressure and temperature wasdecreased. This is consistent with the results from the previous

Fig. 4. O2/N2 separation factor as a function of temperature for different feedpressures.

Table 3Dew points of N2, O2 and synthetic air at the different feed pressures.

Feed pressure(mbar)

Single gas N2 dewpoint (K)

Single gas O2 dewpoint (K)

Mixture gas dewpoint (K)

1000 77.4 90.2 –

1000 – – 81.6660 – – 78.1400 – – 74.4100 – – 65.9

Fig. 5. Single gas permeances for pure O2 and N2 as a function of temperature for afeed pressure of 1000 mbar and a permeate pressure of 100 mbar.

Fig. 6. O2/N2 selectivity versus O2 permeability observed in present work (solidblack square) compared to Robeson's 1991 upper bound and 2008 upper bound:(1) 1000 mbar, (2) 660 mbar, (3) 400 mbar, and (4) 100 mbar for feed pressure.Black stars represent data for a polymer membrane with “exceptional perfor-mance” recently reported in Science [5]. The original figure is obtained from [4],with permission from Elsevier.

P. Ye et al. / Journal of Membrane Science 487 (2015) 135–140 139

Page 85: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

work [12]. An entirely opposite situation was observed in thesingle gas permeation experiments (see Fig. 5) i.e., the per-meance of O2 was lower than that of N2 at all temperatures.Furthermore, the N2 permeance measured in the gas mixtureseparation experiments was much lower, by ca. a factor of 7,than that measured in the single gas permeation experiments.In contrast, the O2 permeances were comparable. The observa-tions in the present work are consistent with those reported inour previous work [12] using feed pressures ranging from 1 barto 5 bar. The slightly higher selectivity (at a higher optimumtemperature) for a feed pressure of 1 bar in present workcompared to the previous work might emanate from theexperimental differences. Helium was used as a sweep gas inthe previous work, while no sweep gas was used in the presentwork. Counter diffusion of helium may certainly influence theseparation performance to some extent.

5. Conclusions

For the first time, ultra-thin MFI membranes were evaluated atcryogenic temperature for air separation at low feed pressure downto 100 mbar. The membranes were found to be oxygen selective at allthe conditions studied. The observed O2/N2 selectivity and oxygenpermeance were well above the upper bound in the 2008 Robesonselectivity–permeability plot. The permeance was nearly 100 timeshigher than that recently reported for promising polymeric mem-branes. The selectivity to O2 should primarily emanate from O2/N2

adsorption selectivity. In addition, with decreasing feed pressure andtemperature, the O2/N2 membrane selectivity was found to increase,which can most likely be attributed to greater O2/N2 adsorptionselectivity at lower temperatures. The present work has thereforeindicated the optimum conditions for air separation using MFImembranes, namely low feed pressures and low temperatures.

Acknowledgments

The Swedish Research Council (Grant no. RMA08-0018), theSwedish Foundation for Strategic Research (Grant no. 2010-5317)and the Swedish Energy Agency (Grant no. 2013-006587) aregratefully acknowledged for financially supporting this work.Furthermore, the authors are grateful to Göran Wallin for technicalassistance.

References

[1] P. Bernardo, E. Drioli, G. Golemme, Membrane gas separation: a review/state ofthe art, Ind. Eng. Chem. Res. 48 (2009) 4638.

[2] R. Baker, Membrane Technology and Applications, John Wiley and Sons,England, 2004.

[3] L.M. Robeson, Correlation of separation factor versus permeability for poly-meric membranes, J. Membr. Sci. 62 (1991) 165.

[4] L.M. Robeson., The upper bound revisited, J. Membr. Sci. 320 (2008) 390.[5] M. Carta, R. Malpass-Evans, M. Croad, Y. Rogan, J.C. Jansen, P. Bernardo, et al.,

An efficient polymer molecular sieve for membrane gas separations, Science339 (2013) 303.

[6] M. Hong, S. Li, J.L. Falconer, R.D. Noble, Hydrogen purification using a SAPO-34membrane, J. Membr. Sci. 307 (2008) 277.

[7] E. Piera, C.A.M. Brenninkmeijer, J. Santamarıa, J. Coronas, Separation of tracesof CO from air using MFI-type zeolite membranes, J. Membr. Sci. 201 (2002)229.

[8] J. van den Bergh, W. Zhu, J. Gascon, J.A. Moulijn, F. Kapteijn, Separation andpermeation characteristics of a DD3R zeolite membrane, J. Membr. Sci. 316(2008) 35.

[9] J. Hedlund, J. Sterte, M. Anthonis, A. Bons, B. Carstensen, N. Corcoran, et al.,High-flux MFI membranes, Microporous Mesoporous Mater. 52 (2002) 179.

[10] L. Sandström, E. Sjöberg, J. Hedlund, Very high flux MFI membrane for CO2

separation, J. Membr. Sci. 380 (2011) 232.[11] M. Zhou, D. Korelskiy, P. Ye, M. Grahn, J. Hedlund., A uniformly oriented MFI

membrane for improved CO2 separation, Angew. Chem. Int. Ed. 53 (2014)3492.

[12] P. Ye, E. Sjöberg, J. Hedlund, Air separation at cryogenic temperature using MFImembranes, Microporous Mesoporous Mater. 192 (2014) 14.

[13] J. Hedlund, F. Jareman, A. Bons, M. Anthonis., A masking technique for highquality MFI membranes, J. Membr. Sci. 222 (2003) 163.

[14] J. Hedlund, D. Korelskiy, L. Sandström, J. Lindmark, Permporometry analysis ofzeolite membranes, J. Membr. Sci. 345 (2009) 276.

[15] D. Korelskiy, M. Grahn, J. Mouzon, J. Hedlund, Characterization of flow-through micropores in MFI membranes by permporometry, J. Membr. Sci.417–418 (2012) 183.

[16] D. Korelskiy, P. Ye, H. Zhou, J. Mouzon, J. Hedlund., An experimental study ofmicropore defects in MFI membranes, Microporous Mesoporous Mater. 186(2014) 194.

[17] G. Sethia, R.S. Pillai, G.P. Dangi, R.S. Somani, H.C. Bajaj, R.V. Jasra, Sorption ofmethane, nitrogen, oxygen, and argon in ZSM-5 with different SiO2/Al2O3

ratios: grand canonical monte carlo simulation and volumetric measurements,Ind. Eng. Chem. Res. 49 (2010) 2353.

[18] F.R. Siperstein, Determination of azeotropic behavior in adsorbed mixtures,Adsorption 11 (2005) 55.

[19] T. Nitta, T. Shigetomi, M. Kuro-oka, T. Katayama, Adsorption isotherm of multi-site occupancy model for homogeneous surface, J. Chem. Eng. Jpn. 17 (1984)39.

[20] D.D. Do, Adsorption Analysis: Equilibria and Kinetics, World Scientific,Singapore, 1998.

[21] R. Krishna, J.M. van Baten., Maxwell–Stefan modeling of slowing-down effectsin mixed gas permeation across porous membranes, J. Membr. Sci. 383 (2011)289.

Fig. 7. (a) Optimum separation temperatures and maximum O2/N2 separation factors as a function of feed pressure. (b) Permeances of O2 and N2 measured in the gasmixture separation experiments at optimum separation temperatures as a function of feed pressure.

P. Ye et al. / Journal of Membrane Science 487 (2015) 135–140140

Page 86: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

PAPER III

Efficient Separation of N2 and He at Low Temperature Using MFI Membranes

Pengcheng Ye, Mattias Grahn, Danil Korelskiy, and Jonas Hedlund

Accepted for publication in AIChE Journal.

Page 87: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 88: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Efficient Separation of N2 and He at Low Temperature Using MFI Membranes

Pengcheng Ye*, Mattias Grahn, Danil Korelskiy and Jonas Hedlund

Chemical Technology, Luleå University of Technology, SE-971 87 Luleå, Sweden

*Corresponding author. Tel: +46920493161; E-mail: [email protected].

Abstract

Ultra-thin MFI membranes were evaluated for N2/He separation over the temperature range of

85–260 K for the first time. The membranes were rather nitrogen selective at all the

conditions investigated. A highest N2/He selectivity of 75.7 with a high N2 flux of 83 kg m−2

h−1 was observed at 124 K. The separation was attributed to adsorption selectivity to N2,

effectively hindering the transport of He in the zeolite pores. The exceedingly high permeance

even at low temperatures was ascribed to the ultra-thin (< 1µm) membrane used. As the

pressure ratios increased, a better separation performance was obtained. A mathematical

model showed the largest difference of adsorbed loading over the film at ca. 120 K was the

main reason for the observed maximum selectivity. Further, the modelling indicated the

selectivity would increase 2-3 times by reducing the influence of defects, concentration

polarisation and pressure drop over the support.

Keywords: Zeolite membrane; Cryogenic temperature; Helium recovery; Gas separation;

Mass transport.

Page 89: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Introduction

There is a large and increasing demand of helium in the world, especially from aerospace and

cryogenic technologies. However, the terrestrial abundance of helium is scarce and the

shortage hampers the development of both research and industry 1,2. Natural gas, as the

current major resource of helium, typically contains 0-4% of helium, and only 1/10 of the

natural gas fields possess economically feasible concentrations of helium, higher than 0.4%. It

is of tremendous importance to develop an efficient method for helium separation as this

would open up an opportunity for economically viable recovery of helium from natural gas

with lower concentration of helium. Besides methane and other hydrocarbons, natural gas

contains a small amount of nitrogen, water vapour, carbon dioxide and helium. The

composition of natural gas varies greatly depending on the source 3. Nitrogen is the most

difficult component to remove from helium when recovering helium from natural gas due to

the need of deep cooling 4. Another application calling for N2/He separation is the reclamation

of spent helium gas from airships and deep-sea diving systems 5,6.

The current well-established technique for large-scale production of commercial helium (>

99.99%) is based on energy-intensive cryogenic distillation and other multi-stage unit

operations that require both high investment and operating costs. A few exploratory reports on

helium purification via pressure swing adsorption using adsorbents, e.g., activated carbon, are

available 7,8. Membrane processes with compact equipment, low capital investment and low

energy consumption compared to thermally driven processes appear to be an attractive

alternative for helium recovery 5,9. Various polymeric membranes have been investigated for

He/N2 separation, but nearly all membranes showing high helium selectivity display low

permeability, i.e., low flux 10-13, requiring large membrane areas for the separation task. It is

worth pointing out that membranes with as high permeability as possible and sufficient

Page 90: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

selectivity are of major interest for industrial applications 14. Many types of inorganic

membranes, e.g., porous alumina 15, silica 15, zeolite 16 and MOF 17 membranes, have also

been evaluated for He/N2 separation, mainly at ambient temperature. The reported ideal

separation selectivities were low and close to the Knudsen selectivity of 2.6 5. Only Koresh et

al. 18 reported separation selectivity much higher than Knudsen selectivity; i.e., an ideal

separation selectivity of 20-40 for molecular sieve carbon membranes with a pore diameter of

0.3-0.5 nm. However, no mixture separation selectivities have been reported, which could

deviate strongly from ideal separation selectivities based on single gas permeation.

As one type of inorganic membranes, zeolite membranes have several potential advantages

over polymeric membranes, e.g., higher flux, higher chemical and thermal stability. Due to the

well-defined pore structure, zeolite membranes can also potentially offer high selectivities.

Our research group has developed zeolite membranes comprised of thin (~ 500 nm) MFI

films deposited on graded alumina supports. The membranes normally exhibit 10-100 times

higher fluxes than most of the previously reported membranes with similar separation factors

19. However, thinner membranes with high performance have been recently prepared by other

groups20-23.With such a high flux, a much smaller membrane area would be needed for the

same separation task, as was demonstrated in our recent works 24,25. Our ultra-thin membranes

have been investigated for CO2 removal from synthesis gas 26,27, showing outstanding

performance. In addition, the high thermal stability of zeolite membranes provides an

opportunity to operate at both high and low temperatures, offering a great advantage over

polymeric membranes. Numerous studies on zeolite membrane separation at high temperature

have been published, whereas scientific reports on low temperature applications of zeolite

membranes are quite scarce in the open literature, especially cryogenic temperatures, i.e.,

below 123K. Recently, our ultra-thin MFI membranes were for the first time evaluated for air

Page 91: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

separation at cryogenic temperature. The observed O2/N2 selectivity and the oxygen

permeance were well above the upper bound in the 2008 Robeson selectivity-permeability

plot 28,29. In the present work, the zeolite membranes were for the first time investigated for

N2/He separation at low temperatures and at different feed and permeate pressures.

Experimental

Membrane preparation

MFI membranes were prepared on a graded porous α-alumina disc (Fraunhofer IKTS,

Germany), using a previously reported method 19. Only a brief description of the procedure

will be given here. More details are found in the Supplementary Material. The support has a

diameter of 25 mm and is comprised of two layers: a 30 μm thick top layer with a pore size of

100 nm and a 3 mm thick base layer with a pore size of 3 μm. The alumina disc was masked

and then seeded with a monolayer of MFI crystals of 50 nm in size as described previously 30.

After that, a hydrothermal treatment was carried out for the MFI film growth. The seeded

support was immersed in a synthesis solution with a molar composition of

3TPAOH:25SiO2:1450H2O:100C2H5OH in a polypropylene tube and was kept in the solution

at 361 K for 72 h with reflux of evaporated solution. After synthesis, the membrane was

rinsed in a 0.1 NH3 solution overnight. Calcination was performed at 773 K for 6 h at a

heating rate of 0.2 K min-1 and a cooling rate of 0.3 K min-1.

The total film thickness of the synthesised MFI membrane, including support invasion, varied

in the range of 0.5–1 µm 29. The film was comprised of well-intergrown crystals and no

defects could be observed by Scanning Electron Microscopy (SEM). Adsorption-branch n-

hexane/helium permporometry 31 was used to characterize the amount of flow-through defects

Page 92: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

in the membranes. The procedure for the permporometry characterization is described in

detail in a previous work 32 and only a brief description is given below. The membrane was

mounted in a steel cell and dried at 573 K for 6 h before the test and the temperature was set

to 323 K during the measurement. The feed pressure was kept at 2 bar and the permeate

pressure was kept at 1 bar. The permeance of the non-adsorbing component helium was

measured at a series of relative pressures (Pi/Pi0) of a strongly adsorbing component; n-

hexane. The relative pressure of n-hexane was raised in a step-wise manner from 0 to ca. 0.4.

At each relative pressure, the system was allowed to achieve steady-state. In the absence of n-

hexane (Pi/Pi0=0), helium can permeate through all pores, i.e., zeolite pores and defects,

resulting in high initial helium permeance. Once introduced, n-hexane will first block small

pores, i.e., zeolite pores, and then larger and larger pores, i.e., defects. Hence, as the

concentration of n-hexane in the feed stream increases, the helium permeance is decreasing.

The desired Pi/Pi0 values were achieved by altering the flow rate of helium and the

temperature of the saturators filled with n-hexane. The defect width was calculated from n-

hexane relative pressure using either the Horvàth–Kavazoe equation (micropore defects) or

the Kelvin equation (mesopore defects). The same membrane as used in the present work has

earlier been evaluated for air separation at cryogenic temperature, and more details regarding

the membrane properties of this particular membrane have been reported earlier 29.

Single gas permeation and binary gas separation measurements

Fig. 1. shows a principle scheme of the experimental set-up used for single gas permeation

and N2/He mixture separation experiments in the present work. Two on-line columns packed

with 5A zeolite pellets were dried at 583 K in a flow of helium (purity > 99.999%, AGA)

before the experiment and were then used to adsorb any traces of moisture in the feed gas

during the experiment. The MFI membrane was placed in a stainless steel cell and sealed with

Page 93: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

graphite gaskets (Eriks, the Netherlands). Prior to the permeation experiments, the cell with

the membrane inside was mounted in a cryostat and the membrane was dried in a helium flow

at 573 K for 6 h at a heating rate of ca. 2 K min-1. The temperature was monitored by a type T

thermocouple on the feed side. During the separation experiments, the gas flow rates of

nitrogen and helium were controlled by mass flow controllers (Bronkhorst, F-201CV) and

mixed dynamically after going through the adsorption columns. The feed mixture was then

introduced to the membrane cell inside the cryostat. The feed pressure was controlled by a

back pressure regulator (Swagelok, KCB). The permeate pressure was controlled by a dosing

valve (Pfeiffer Vacuum, EVN 116) connected to a vacuum pump (Pfeiffer Vacuum, MVP 040).

The pressures on the feed and permeate sides were monitored using pressure gauges

(Swagelok).The custom-built cryostat (ICEoxford UK, DRYICE25K) could set separation

temperature at any value ranging from 22 K to 573 K with a deviation of less than 0.05 K.

The permeate flow was measured with a mass flow meter (Bronkhorst, F-201CV) and

composition was analysed online by a mass spectrometer (GAM 400, InProcess Instruments)

every 5 seconds.

The feed and permeate pressures used in all mixture separation experiments (experiments

denoted E1-E3) are shown in Table 1. Mixtures consisting of 52.6 % N2 (>99.999%, AGA)

and 47.4 % He (>99.999%, AGA), were fed to the membrane at a total feed pressure of 3-5

bar. The permeate pressure was varied from 0.5 to 1 bar. The pressure ratios, defined as total

feed pressure divided by permeate pressure φ= PFeed/PPerm, were 3, 5 and 10 for E 1, E2 and

E3, respectively.

The permeance Πi (mol m−2 s−1 Pa−1) was defined as

Πi=Fi/(AfilmΔPi), (1)

Page 94: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

where Fi is the molar flow rate of component i through the membrane, Afilm is the membrane

area, and ΔPi is the partial pressure difference of component i across the membrane.

The membrane selectivity (also referred to as the permeance-based selectivity) was

determined as

αi/j=Πi/Πj, (2)

where Πi and Πj are the permeances of species i and j, respectively.

The separation factor was calculated as follows:

βi/j=(yi/yj)/(xi/xj), (3)

where yi and xi are the molar fractions of species i in the permeate and feed streams,

respectively.

To better understand the mechanism of N2/He separation, single gas permeation experiments

(experiments E4 and E5) were carried out over a temperature range of 58-121 K, see Table 1.

The feed pressure was kept at 5 bar while the permeate side was kept at 0.5 bar, as in E3.

Modelling

Flow-through defects in the membrane, e.g., in the form of open grain boundaries, can affect

the performance of the membrane. For high flux membranes, the membrane performance may

also be affected by both external mass transfer limitations (concentration polarisation) and

pressure drop over the support. To elucidate to which extent the experimental results were

affected by these factors, modelling of mass transfer through the membrane was carried out.

The model was not fitted to experimental data. The parameter values used as inputs for the

modelling were summarized in Table 2.

Transport through the zeolite layer was assumed to occur both via transport through zeolite

Page 95: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

pores and via flow-through defects 25,32,33, where the mixed gas transport through the zeolite

pores was described using the Maxwell-Stefan formalism 34-36 in combination 37,38 with the

ideal adsorbed solution theory (IAST). For a binary mixture, the transport expressed via the

permeating flux for species 1 and 2 through the zeolite pores is given by

( )

++

∆+∆+=

12121221

,21221,112211,1 //1

//1DDxDDx

qDDxqDDxlD

AAJ adsads

z

z

tot

zz

ρ (4)

( )

++

∆+∆+=

12121221

,11212,212122,2 //1

//1DDxDDx

qDDxqDDxlD

AAJ adsads

z

z

tot

zz

ρ (5)

where Az/Atot is the relative area of the membrane surface that is composed of zeolite, and this

ratio will be close to unity. The small remaining area will be composed of flow-through

defects. The defect distribution and total relative defect areas of defects were obtained from

permporometry data, see Table 3. Further, zρ and zl are the zeolite density and film thickness,

respectively. Di is the Maxwell-Stefan diffusivities for species i, determined by:

−=

RTE

DD diffAii

,0 exp

(6)

however, the diffusion in zeolites is typically also dependent on the adsorbed concentration. It

decreases with increasing loading, in particular at high loadings as is the case for N2. This

effect is commonly expressed using the Reed-Ehrlich approach39:

( )( )z

ii

zidiffA

ii RTE

DDfε

ε/1

1exp1

,0 +

+

−=

(7)

where z is the coordination number, and with:

( )( )i

iii θ

fθβε

+−=

1221

(8)

and

Page 96: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

( )( )( )iiii fθθβ /11141 −−−= (9)

( )ii aa θf 21 exp= (10)

where a1 and a2 are two constants. For He, we used the constants reported for H2 previously

of 1.0 and 0.35, respectively, see Krishna et al. for details 40,41. The activation energies of

diffusion (EA, diff) were 3.1 kJ mol-1 and 1.0 kJ mol-1 for N2 and He, respectively 42,43. The

diffusivities and parameters for describing the dependence of diffusivities with loadings have

been reported for N2 previously 44. For He, we used a diffusivity at zero loading (D0i) of

10.8×10-8 m2 s-1 (at 300 K), i.e., 1.3 times that of H2 42,44 and the dependence of diffusivity on

adsorbed loading was assumed to be identical to that of H2 42. In general, correlation effects

(diffusional coupling) are strong in the MFI framework. This manifests itself in that a stronger

adsorbing, but slow diffusing specie, slows down the transport of a less adsorbing, faster

diffusing specie in mixture permeations as shown in numerous examples previously 37,45,46. In

the Maxwell-Stefan formalism, the correlation effects are given from the exchange coefficient,

D12. This is in turn usually obtained from the self-exchange diffusivities D11 and D22 for N2

and He, respectively, via the Vignes interpolation 44, as:

( ) ( ) 21221112

xx DDD = (11)

where x1 and x2 are the adsorbed mole fractions of N2 and He, respectively. Further, Δqi, ads, is

the difference in adsorbed loading of component i between the feed and permeate side of the

zeolite film. The partial pressures on the feed side of the zeolite film were corrected for

concentration polarisation and the partial pressures on the permeate side were corrected for

the pressure drop from the support, as described below by Eqs. 17-23.The adsorbed loadings

were determined by IAST using single component Langmuir parameters (saturation loading

Page 97: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

and Langmuir adsorption coefficients) reported by Krishna as input 35. The Langmuir

adsorption coefficients were extrapolated to the temperatures of interest using heat of

adsorptions. The heat of adsorption for He was assumed to be half that of H2, i.e., -2.9 kJ mol-

1 42. The adsorbed phase mole fraction of species i, xi, was determined as x1=q1/(q1+q2).

Kapteijn et al. have previously analysed single-component permeation data as a function of

temperature showing a maximum in the flux and used this data to determine the ratio between

activation energy of diffusion and heat of adsorption 47 :

)11

ln(

)1()1(

,

,

,,,

feedi

permi

feedipermi

ads

diffA

QE

θθ

θθ

−−

−−−=

(12)

where, Qads is heat of adsorption; feedi,θ and permi,θ are the fractional adsorbed loadings on feed

side and permeate side, respectively.

Transport via flow-through defects, such as open grain boundaries, was calculated following

the procedure reported previously 25,32. In short, the transport was estimated using Fick’s law:

z

filmiidefi l

PRTDJ ,

,

∆= , (13)

where, according to gas translational theory, the diffusivity of a gas molecule in a porous

medium can be expressed as

∆−=

RTE

MRTr

D idefi exp82

pτσ

, (14)

where σ is the geometrical probability that the gas molecule will move in the desired direction.

This parameter was assumed to be 1/3; referring to the case of no spatial obstructions in the

pathway of the diffusing molecule moving in three-dimensional space, rdef is the average

Page 98: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

diameter of the defect and is obtained from permporometry data and τ is the tortuosity of the

defects.

The activation energy of diffusion in the defects was assumed to decrease with increasing

defect width according to the van der Waal atom-slab model:

3−=∆ krEi , (15)

where r is half width of a channel (defect) and k is the proportionality coefficient given from

the ratio 3, )2//( −∆ zpdiffA dE in this case, where dzp is the average diameter of the zeolite pores.

The total fluxes through defects were then calculated by summing up the fluxes through each

defect size interval, as obtained from permporometry data:

=

nndefi

tot

ntotdefi J

AA

J ,,,, , (16)

where totn AA / is the relative area of defects in a particular defect size interval and n is the

index of the defect size interval.

The effect of concentration polarisation was assessed by determining the concentration

polarisation index 48:

[ ])/exp(1)/exp( cvb

pcv

b

m kJnn

kJnn

−+= , (17)

where nm is the mole fraction at the zeolite film - gas interface, nb is the mole fraction in the

gas bulk, kc is the mass transfer coefficient, Jv is the volumetric flux and np is the mole

fraction in the permeate. The mass transfer coefficient is usually determined from correlations

of the Sherwood (Sh) number as a function of the Reynolds (Re) and Schmidt (Sc) numbers.

These numbers are dependent on, e.g., flow conditions (laminar or turbulent flow) and system

geometry. Perdana et al. 49 have determined correlations for external mass transfer in a Wicke-

Kallenbach cell:

Page 99: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

30.133.058.0Re24.283.2

+=

CDwScSh , (18)

where w is the compartment height, i.e., the thickness of the gas volume on the feed side of

the membrane, and DC is the diameter of the membrane inside the gaskets, for details see

literature 49. The mass transfer coefficient (kc) was finally obtained from the Sherwood

number:

Ddk

Sh hc= , (19)

where dh is the hydraulic diameter of the cell and D is the diffusivity of N2 in He.

Fig. 2. shows a schematic illustration of a pressure drop profile for the preferred permeating

species over a composite membrane and gas film. Pi,b is the partial pressure of species i in gas

bulk on the feed side. Pi,1 is the partial pressure at the zeolite film surface on the feed side. Pi,4

is the partial pressure in the permeate bulk. Further, Pi,1 –Pi,2 is the pressure drop over the

zeolite film layer, Pi,2 –Pi,3 is the pressure drop over the 30 µm thick top layer of the support,

whereas Pi,3 – Pi,4 is the pressure drop over the 3 mm thick base layer of the support. The

pressure drop over the support was assessed using the same procedure as reported by several

groups before 32,50-52. The flux through the small pore top layer (L1) of the support was

modelled as occurring via combination of viscous flow and Knudsen diffusion:

( )1

13,2,1

1

1,1,,1, 2 L

LiieffL

L

Lieff

LiKnLi l

PRT

PPBlP

RTD

J ∆

++

∆=

η, (20)

where the effective Knudsen diffusivity effKnD is given by:

i

effiKn M

TKD 97, = , (21)

where K is the Knudsen structural parameter for layer L1, and Mi is the molar mass of

component i. The Knudsen structural parameter was estimated as:

Page 100: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

11

1L

L

L rKτε

= , (22)

where εj is the porosity of layer j, τj is the tortuosity of the pores in layer j and rj is the average

pore radius in layer j.

The flux through the coarse pores of the bottom layer L2 was modelled as occurring only by

viscous flow:

( )2

24,3,22, 2 L

LiieffL

Li lP

RTPPB

J ∆

+=

η. (23)

In these equations, effjB is the effective permeability for layer j, η is the viscosity of the gas

and ΔPj is the partial pressure difference over layer j. The Knudsen structural parameter and

effective permeabilities used have been determined previously 53, see Table 2.

In the simulations, ideal gas behaviour was assumed on both feed and permeate sides of the

membrane. Gas phase diffusivities and viscosities were adjusted to the desired temperature

using the Di /23T∝ relationship and Sutherland’s equation 54, respectively. The viscosities

were calculated as molar fraction weighted averages, with pure component viscosities (at 300

K) of 1.78×10-5 and 1.99×10-5 kg m-1 s-1 for N2 and He, respectively..

Results and Discussion

n-Hexane/helium permporometry has been demonstrated to be an effective and reliable tool

for evaluating the amount of micropore defects down to 0.7 nm 32,55 and mesopore defects 31

in zeolite membranes, correlating well with Scanning Electron Microscopy (SEM) analysis

and gas mixture separation data. Permporometry data for the particular membrane used in the

present work have been published elsewhere 29. Helium permeance and estimated defect

distribution calculated from the permporometry data are shown in Table 3. The initial helium

Page 101: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

permeance at p/p0 = 0, i.e., the helium permeance in the absence of n-hexane, was about

43×10−7 mol m−2 s−1 Pa−1. This permeance is about half of that for our best membranes with a

total film thickness of about 500 nm 19 and the lower permeance observed here was a result of

a thicker film. However, the defect distribution for this membrane is consistent with that

previously reported for high quality membranes prepared by our group 32. At the relative

pressure of about 1.13×10-2, the helium permeance reduced by more than two orders of

magnitude due to adsorbed n-hexane molecules blocking transport via the zeolite pores. This

relative pressure corresponds to a defect width of 1.04 nm. According to the estimated defect

distribution (Table 3), the majority of defects, i.e., 96.5% of the total amount, were micropore

defects smaller than 1 nm in size. In addition, the total amount of defects was only 0.20% of

the total membrane surface area and nearly no defects larger than 5 nm in size could be

detected, indicating high quality of the membrane.

Fig. 3. illustrates the N2/He separation results at various feed and permeate pressures in the

temperature range of 82 to 283 K. The membrane was quite nitrogen selective at all

conditions studied. The N2/He separation factors exceeded 10 over a wide temperature range,

from 90 to 180 K. For each experimental run (E1, E2 and E3), a maximum separation factor

was detected at an optimum temperature. This should be related to much stronger adsorption

of N2 in the zeolite compared to that of He, which hindered the transport of He through the

membrane. Bakker et al. 42 measured adsorption of various gases in silicalite-1 crystals over a

temperature range of 200-600 K and at a pressure of 101 kPa. The calculated values for heat

of adsorption of N2 and He were -13.8 kJ mol-1 and 0 kJ mol-1, respectively. Less than 0.1

mmol g-1 He (the lower limit for the method used) could be adsorbed at 200 K. In other words,

He was “essentially non-adsorbing” and the adsorption of N2 was exceedingly stronger

compared to that of He at those conditions. At higher pressures and lower temperatures, as in

Page 102: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

the present work, He is expected to adsorb in silicalite-1, albeit significantly weaker than N2.

Comparing different separation conditions for feed and permeate pressures, the separation

factors increased in order of E1 < E2 < E3, with a pressure ratio of 3, 5 and 10, respectively,

see Table 1. The maximum separation factor observed in E1 was 30.5 at an optimum

temperature of 118 K, 50.7 at 130 K in E2 and 61.7 at 124 K in E3. The separation factors

increased as the pressure ratio increased since the process was in the pressure-ratio-limited

region (α >> φ).

The data in experiment E3 are described in more detail in Fig.4. In this experiment (and

similarly in E1 and E2), the permeance of N2 decreased as the temperature decreased from

260 K to 85 K, and the effect was even more significant for temperatures below 118 K. The

permeances of N2 were substantially higher than those of He over the entire temperature range.

In contrast, the single gas permeance of He was higher than that of N2 at all temperatures, see

Fig. 5. At the same time, the permeances of N2 measured in the separation experiments and

single gas permeation experiments at similar conditions were very similar. In contrast, He

permeances in the binary separation experiments were lower than those in the single gas

experiments by a factor of 10-200. Thus, the transport of N2 is nearly unaffected by the

presence of He in the mixture, whereas the transport of He is strongly dependent on the

presence of N2 in the gas. These findings demonstrated that the separation mechanism at the

conditions studied should be based on selective adsorption and effective transport of N2,

impeding the transport of He under these conditions. It is worth noting that the N2 permeances

measured at cryogenic temperatures were only a few times lower than those measured at the

highest investigated temperature of 260 K, but still more than 100 times larger than those

reported for polymeric membranes at room temperature 11. N2 permeances at 260 K still were

3-150 times higher than those reported for zeolite membranes at room temperature 56. A

Page 103: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

maximum selectivity of 75.7 was obtained at an optimum temperature of 124 K with a high

N2 permeance of 39 ×10−7 mol m−2 s−1 Pa−1. A much lower membrane area would be needed

for our zeolite membranes with higher permeance and high selectivity as compared to

polymeric membranes to carry out a given separation task. The selectivities were approaching

one when the temperature was close to 260 K. No study of the separation performance of our

membranes at temperatures above 260 K was carried out due to experimental limitations; the

increasing permeance at high temperature requires much higher feed flow rate, which is,

however, beyond the capacity of the mass flow controllers in our current experimental setup.

The higher N2/He selectivities at cryogenic temperatures, as compared to those at

temperatures close to room temperature, should primarily emanate from more extensive

adsorption of N2 at cryogenic temperature, resulting in more effective blocking of the He

transport. The same trend was observed for the lowest temperatures studied. The drastic drop

in selectivities below 118 K was related to the drastic drop in N2 permeance at that

temperature, when N2 gas perhaps started to condense in the membrane pores. The drastic

drop in selectivities at a temperature ranging from 118 to 85 K (see Figure 4a) was related to

the drastic drop in N2 permeance, most likely due to the reduced diffusion of N2 at those

temperatures. In contrast, at temperatures below 85 K, the selectivity was increasing, reaching

a very high value of ca. 400 at 73 K, see Figure 4b. Since the boiling point of N2 at 2.6 bar (=

5bar × 52.6%) is 90 K, N2 gas is most likely condensing in the cell and on the membrane

surface at temperatures below the boiling point. This liquid may block defects so that nearly

no He molecules could pass through the membrane. The condensation of N2 should also result

in a much lower mass transfer rate through zeolite pores. Indeed, the permeances of N2 and

He were extremely low at temperatures below 85 K, i.e., 2.7 ×10−8 mol m−2 s−1 Pa−1 and 6.0

×10−10 mol m−2 s−1 Pa−1, respectively.

Page 104: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

As reports on heat of adsorption and activation energy of diffusion for He in zeolite are scarce,

we used the approach expressed by Eq. 12 to analyse our single-component permeation data.

The ratios for He were determined as 0.95, 0.97, 0.98, 0.99 at 85, 106, 124 and 170 K,

respectively, meaning the flux becomes independent of temperature at high temperature. It is

consistent with the experimental observations, see Fig. 5. For our N2 single gas permeance

data, the adsorbed loadings are indeed high and the effect of adsorbed loading on the

diffusivity of N2 expressed via the Reed Ehrlich model (see Eq. 7) is greater than that of the

decreased diffusivity with temperature expressed via the activation energy of diffusion. For

our He single gas permeation data, the adsorbed loadings are low and the effect of adsorbed

loading on the single gas diffusion of He is negligible. However, for our N2 single gas data,

the adsorbed loading is very high and the approach (see Eq. 12) reported by Kapteijn et al is

probably not suitable for analysing the data as it would likely result in severe errors in the

estimation of the EA, diff/Qads – ratio.

To further understand the behaviour of the membranes at conditions E3 (5 bar feed pressure

and 0.5 bar permeate pressure), the membrane processes was simulated at three temperatures,

viz. 106, 124 and 171 K to capture the maximum in selectivity observed at ca. 124 K. Note

that the model was not fitted to the experimental data. The results from the simulations are

summarised in Table 4. Simulated permeances and selectivity (red) are shown together with

experimental data (black) in Figure 4a. The permselectivities obtained from the simulations

are in reasonably good agreement with the experimental observations and the modelling

results capture the main trend with maximum in selectivity at ca. 120 K. From the simulations

it may be inferred that the driving force, expressed as the difference in adsorbed loading of N2

over the zeolite film Δqads, is largest at 124 K, see Table 4. This is probably one of the main

reasons for the observed maximum in selectivity at this temperature. Moreover, N2 adsorbs

Page 105: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

much stronger in zeolite ZSM-5 than He. According to the IAST, the adsorbed concentration

of N2 on the feed side of the membrane is 100 to 300 times larger than that of He, indicating

that the N2 is effectively blocking the transport of He through the zeolite pores. Based on Eq.

11, as the adsorbed concentration of N2 is much higher than that of He, the self-diffusivity of

N2 will control D12, i.e. the transport of He will be slowed down by the slower N2. The degree

of correlation for He, is given by the ratio D2/D12, and correlations are significant if this ratio

is >>0. In this work this ratio was very high viz. 6900, 990 and 148 at 106, 124 and 170 K,

respectively, further highlighting the strong slowing down effect of N2 on He diffusion in

these experiments.

Another contributing factor is the transport via flow-through defects. As the lighter molecule,

He is preferentially transported through the defects, which decreases the selectivity of the

membrane. The fractions of He being transported via flow-through defects were found to be

93%, 72% and 23% at 106 K, 124 K and 171 K, respectively. For comparison, the relative

flux through defects was always below 1% for N2. In the model, it is assumed that the

transport of N2 and He through the defects are independent of each other. This may be an

oversimplification at the lowest temperature where there is likely a substantial adsorption of

N2 in the defects. Such an adsorbed layer would certainly lower the flux of He through the

defects to some extent, and we propose that this may be a contributing factor for the observed

large difference in simulated and experimentally measured permselectivity at the lowest

temperature. That is, in the experiments, the defects would be partially blocked by adsorbed

N2, decreasing the flux of He through the defects, and increasing the selectivity as compared

to the estimated selectivity. Nevertheless, the model suggests that the selectivity of the

membrane could be improved if the transport through flow-through defects could be

eliminated. If the transport via defects could be completely eliminated, the selectivity would

Page 106: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

reach impressive values of 111, 192 and 38, respectively, at the three temperatures considered.

The discrepancy between simulated and experimental results may also be ascribed to that the

inputs to the simulations are for defect free silicalite-1, whereas the real membrane is a high

silica (Si/Al ≈ 139) ZSM-5 membrane. The membrane prepared at alkaline conditions is

known to possess intracrystalline defects in the zeolite, which in turn may affect both the

adsorption and diffusion properties. Furthermore, the Langmuir adsorption coefficients and

diffusivities have been extrapolated from ca. 300 K to the range 171–106 K. However, as

stated previously, there is still reasonably good agreement between the experiment and

simulations with the main trends captured by the simulations despite the fact that the model

was not fitted to experimental data.

Concentration polarisation may be an issue for membrane separation processes, especially for

high flux membranes. The concentration polarisation indices, i.e., Cmem,N2/Cbulk,N2, determined

by Eq. 17 were 0.75, 0.54 and 0.5 at 106, 124 and 171 K, respectively, indicating that external

mass transfer limitations have affected the results and that the limitations were more severe at

higher temperatures where the permeance is higher. As the correlation for determining the

external mass transfer coefficient was used at higher Reynolds numbers (Re in the range 600-

1200) than the range for which it was derived (Re < 50), we do not want to draw too far

reaching conclusions for these results. However, it should be pointed out that these Reynolds

numbers (Re < 1200) still correspond to laminar flow as the transition from laminar to mixed

flow occurs at ca Re=2300 for flow in pipes. However, external mass transfer limitations may

be diminished by increasing the velocity of the feed. If the feed flow rate could be increased

to 50 l min-1, yielding a concentration polarisation index of 0.95, the model shows that the

selectivity would nearly double, from 53 to 97 at 124 K. The increase in selectivity primarily

originates from a smaller flux of He through the flow-through defects due to the reduced

Page 107: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

partial pressure of He at the membrane feed interface as the polarisation index is increased.

For N2, the increase in the partial pressure at the membrane surface leads to an increase in the

adsorbed concentration on the feed side and thus an increased driving force over the zeolite

film. However, at the same time the adsorbed concentrations are getting close to saturation at

these conditions and consequently the increased driving force is offset by a decrease in the

diffusivity because of the high adsorbed concentration. The net effect of reducing the

concentration polarisation on the transport of N2 through the membrane is therefore small. A

feed flow rate of 50 l min-1 is reasonable for the separation as the resulting average superficial

velocity in the cell would be ca. 5 m s-1 at a hydraulic diameter of ca. 6.5 mm. The main

limiting factor for high superficial velocities is the pressure drop and we have previously

shown that the pressure drop in a tubular membrane with 7 mm inner diameter and a length of

2.4 m is very small, even at superficial velocities of 15 m s-1 25. In our case, the feed flow rate

was limited by the experimental setup, e.g., the capacities of the mass flow controllers and

cooling capacity of the cryostat.

For high flux membranes, the performance of the membrane may also be affected by the

pressure drop over the support. In the present work, the relative pressure drop over the

support (compared to the total pressure drop over the membrane) was determined by Eqs. 4, 5,

20-23 after adjusting the gas composition for concentration polarisation. The relative pressure

drop over the support was estimated to be 5, 16 and 27% at 106, 124 and 171 K, respectively,

thus increasing with increasing permeance through the membrane as expected. In concert with

previous findings, the main part of the pressure drop was over the thin top layer of the support

25,53. Reducing the thickness of this layer from 30 µm to 10 µm would be an effective but still

realistic option for reducing the influence of the support. Such a reduction in the thickness

could diminish the relative pressure drop from 27% to 10% at the highest gas permeance at

Page 108: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

171 K and from 16% to 7.5% at 124 K. However, if both the feed flow rate could be increased

to 50 l min-1 and the thickness of the top layer could be 10 μm, the selectivity at 124 K would

nearly double, increasing from 53 to 99. Again, the largest effect originates from a decreased

transport of He via flow-through defects because the partial pressure of He at the zeolite/fluid

interface is decreased as the concentration polarisation index approaches one.

Conclusion

Ultra-thin MFI membranes were for the first time evaluated for N2/He separation at low

temperature down to 73 K with different feed-to-permeate pressure ratios. The membranes

were found to be nitrogen selective at all the conditions studied. The highest separation factor

of 61.7 (corresponding separation selectivity 75.7) was observed at a temperature of 124 K, a

feed pressure of 5 bar and a permeate pressure of 0.5 bar with a N2 permeance of as high as

39×10−7 mol m−2 s−1 Pa−1. The selectivity to N2 should mostly emanate from stronger

adsorption of N2 in the zeolite pores, thus blocking the transport of He, especially at

cryogenic temperatures, as observed experimentally and by a mathematical model.

Furthermore, the N2/He selectivity was found to increase with increasing feed/permeate

pressure ratios due to reduced pressure ratio limitations. The present work has therefore

identified the optimum conditions for N2/He separation using MFI membranes, namely high

feed pressure ratio and cryogenic temperatures. The permeances decreased as temperature

decreased, but the membrane used in present work still showed fairly high permeances, i.e.,

high flux, mainly due to the ultra-thin porous zeolite film grown on a fully open graded and

highly permeable support. Lower membrane area would be needed for our zeolite membranes

with higher permeance as compared to polymeric membranes for a given separation task.

Modelling results showed that defects, concentration polarisation and pressure drop over the

support limited the separation performance significantly. The main trends of membrane

Page 109: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

selectivity with a maximum value at ca. 120 K could be described by the simulations. Further,

simulations show that the selectivity at ca. 120 K could be nearly doubled to 99 by reducing

the concentration polarisation and increasing the permability of the support.

Acknowledgment

The Swedish Energy Agency (Project no. 38336-1) is gratefully acknowledged for the

financial support. Furthermore, the authors are grateful to Göran Wallin for technical

assistance in the lab.

Page 110: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Literature Cited

1. Kaplan KH. Helium shortage hampers research and industry. Phys Today. 2007;60:31-32.

2. Cho A. Helium-3 shortage could put freeze on low-temperature research. Science.

2009;326:778-779.

3. Agrawal J, Sourirajan S. Helium separation by cellulose acetate membranes. J Appl Polym

Sci. 1969;13:1065-1068.

4. Ruemann R, Fedoritenko A. Physical Basis of Separationg Helium and Nitrogen. Redstone

scientific information center redstone arsenal al: DTIC Document;1964. RSIC-226.

5. Hsieh H. Inorganic membranes for separation and reaction. Alcoa Technical Center, USA:

Elsevier, 1996.

6. Abetz V, Brinkmann T, Dijkstra M, Ebert K, Fritsch D, Ohlrogge K, Paul D, Peinemann K-,

Nunes SP, Scharnagl N, Schossig M. Developments in membrane research: From material via

process design to industrial application. Adv Eng Mater. 2006;8:328-358.

7. Das NK, Chaudhuri H, Bhandari RK, Ghose D, Sen P, Sinha B. Purification of helium

from natural gas by pressure swing adsorption. Curr Sci. 2008;95:1684-1687.

8. Bird G, Granville W. Separation of Nitrogen from Helium Using Pressure-Swing

Adsorption. In: Timmerhaus KD . Advances in Cryogenic EngineeringSpringer, 1995:463-

473.

9. Schrier J. Helium separation using porous graphene membranes. Chem Lett. 2010;1:2284-

2287.

Page 111: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

10. Robeson LM. Correlation of separation factor versus permeability for polymeric

membranes. J Membr Sci. 1991;62:165-185.

11. Robeson LM. The upper bound revisited. J Membr Sci. 2008;320:390-400.

12. Stern S, Sinclair T, Gareis P, Vahldieck N, Mohr P. Helium recovery by permeation. Ind

Eng Chem. 1965;57:49-60.

13. Rezac ME, Shanan Moore N, Georgia AB. Effect of Temperature on the Transport

Properties and Morphology of Polymeric Asymmetric Membranes. Sep Sci Technol.

1997;32:505-525.

14. Baker R. Membrane technology and applications. England: John Wiley and Sons. 2004.

15. Havredaki V, Petropoulos JH. Experimental studies in relation to a new theoretical

description of the permeation of dilute adsorbable gases through porous membranes. J Membr

Sci. 1983;12:303-312.

16. Jia M, Peinemann K, Behling R. Ceramic zeolite composite membranes. Preparation,

characterization and gas permeation. J Membr Sci. 1993;82:15-26.

17. Cao F, Zhang C, Xiao Y, Huang H, Zhang W, Liu D, Zhong C, Yang Q, Yang Z, Lu X.

Helium recovery by a Cu-BTC metal-organic-framework membrane. Ind Eng Chem Res.

2012;51:11274-11278.

18. Koresh JE, Sofer A. Molecular sieve carbon permselective membrane. Part I. Presentation

of a new device for gas mixture separation. Sep Sci Technol. 1983;18:723-734.

Page 112: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

19. Hedlund J, Sterte J, Anthonis M, Bons A, Carstensen B, Corcoran N, Cox D, Deckman H,

De Gijnst W, de Moor P. High-flux MFI membranes. Microporous Mesoporous Mater.

2002;52:179-189.

20. Agrawal KV, Topuz B, Pham TCT, Nguyen TH, Sauer N, Rangnekar N, Zhang H,

Narasimharao K, Basahel SN, Francis LF, Macosko CW, Al-Thabaiti S, Tsapatsis M, Yoon

KB. Oriented MFI membranes by gel-less secondary growth of sub-100 nm MFI-nanosheet

seed layers. Adv Mater. 2015;27:3243-3249.

21. Agrawal KV, Topuz B, Jiang Z, Nguenkam K, Elyassi B, Francis LF, Tsapatsis M,

Navarro M. Solution-processable exfoliated zeolite nanosheets purified by density gradient

centrifugation. AIChE J. 2013;59:3458-3467.

22. Pham TC, Kim HS, Yoon KB. Growth of uniformly oriented silica MFI and BEA zeolite

films on substrates. Science. 2011;334:1533-1538.

23. Pham TCT, Nguyen TH, Yoon KB. Gel-free secondary growth of uniformly oriented

silica MFI zeolite films and application for xylene separation. Angew Chem Int Ed.

2013;52:8693-8698.

24. Korelskiy D, Ye P, Fouladvand S, Karimi S, Sjöberg E, Hedlund J. Efficient ceramic

zeolite membranes for CO2/H2 separation. J Mater Chem A. 2015;3:12500-12506.

25. Grahn M, Hedlund J. Maxwell–Stefan modeling of high flux tubular silicalite-1

membranes for CO2 removal from CO2/H2 gas mixtures. J Membr Sci. 2014;471:328-337.

26. Sandström L, Sjöberg E, Hedlund J. Very high flux MFI membrane for CO2 separation. J

Membr Sci. 2011;380:232-240.

Page 113: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

27. Zhou M, Korelskiy D, Ye P, Grahn M, Hedlund J. A uniformly oriented MFI membrane

for improved CO2 separation. Angewandte Chemie - International Edition. 2014;53:3492-

3495.

28. Ye P, Sjöberg E, Hedlund J. Air separation at cryogenic temperature using MFI

membranes. Microporous Mesoporous Mater. 2014;192:14-17.

29. Ye P, Korelskiy D, Grahn M, Hedlund J. Cryogenic air separation at low pressure using

MFI membranes. J Membr Sci. 2015;487:135-140.

30. Hedlund J, Jareman F, Bons A, Anthonis M. A masking technique for high quality MFI

membranes. J Membr Sci. 2003;222:163-179.

31. Hedlund J, Korelskiy D, Sandström L, Lindmark J. Permporometry analysis of zeolite

membranes. J Membr Sci. 2009;345:276-287.

32. Korelskiy D, Grahn M, Mouzon J, Hedlund J. Characterization of flow-through

micropores in MFI membranes by permporometry. J Membr Sci. 2012;417-418:183-192.

33. Hedlund J, Grahn M, Korelskiy D, Rayson M, Öberg S, Briddon PR. Mass transport in

porous media from first principles: An experimental and theoretical study. J Membr Sci.

2012;415-416:271-277.

34. Krishna R, van Baten JM. Investigating the influence of diffusional coupling on mixture

permeation across porous membranes. J Membr Sci. 2013;430:113-128.

35. Krishna R. The Maxwell-Stefan description of mixture diffusion in nanoporous crystalline

materials. Microporous Mesoporous Mater. 2014;185:30-50.

Page 114: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

36. Krishna R, van Baten JM. In silico screening of zeolite membranes for CO2 capture. J

Membr Sci. 2010;360:323-333.

37. Van De Graaf JM, Kapteijn F, Moulijn JA. Modeling permeation of binary mixtures

through zeolite membranes. AIChE J. 1999;45:497-511.

38. Kapteijn F, Moulijn JA, Krishna R. The generalized Maxwell-Stefan model for diffusion

in zeolites: Sorbate molecules with different saturation loadings. Chem Eng Sci.

2000;55:2923-2930.

39. Reed DA, Ehrlich G. Surface diffusion, atomic jump rates and thermodynamics. Surf Sci.

1981;102:588-609.

40. Krishna R, Van Baten J. Diffusion of alkane mixtures in zeolites: Validating the Maxwell-

Stefan formulation using MD simulations. The Journal of Physical Chemistry B.

2005;109:6386-6396.

41. Krishna R, Paschek D, Baur R. Modeling the occupancy dependence of diffusivities in

zeolites. Microporous and mesoporous materials. 2004;76:233-246.

42. Bakker WJW, Van Den Broeke LJP, Kapteijn F, Moulijn JA. Temperature Dependence of

One-Component Permeation through a Silicalite-1 Membrane. AIChE J. 1997;43:2203-2214.

43. Selassie D, Davis D, Dahlin J, Feise E, Haman G, Sholl DS, Kohen D. Atomistic

simulations of CO2 and N2 diffusion in silica zeolites: The impact of pore size and shape. J

Phys Chem C. 2008;112:16521-16531.

44. Krishna R, van Baten JM. Maxwell-Stefan modeling of slowing-down effects in mixed

gas permeation across porous membranes. J Membr Sci. 2011;383:289-300.

Page 115: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

45. Bakker WJW, Kapteijn F, Poppe J, Moulijn JA. Permeation characteristics of a metal-

supported silicalite-1 zeolite membrane. J Membr Sci. 1996;117:57-78.

46. Hong M, Li S, Falconer JL, Noble RD. Hydrogen purification using a SAPO-34

membrane. J Membr Sci. 2008;307:277-283.

47. Kapteijn F, van de Graaf, Jolinde M, Moulijn JA. One‐component permeation maximum:

Diagnostic tool for silicalite‐1 membranes? AIChE J. 2000;46:1096-1100.

48. Avila AM, Funke HH, Zhang Y, Falconer JL, Noble RD. Concentration polarization in

SAPO-34 membranes at high pressures. J Membr Sci. 2009;335:32-36.

49. Perdana I, Tyoso BW, Bendiyasa IM, Rochmadi, Wirawan SK, Creaser D. Effect of

external mass transport on permeation in a Wicke-Kallenbach cell. Chem Eng Res Des.

2009;87:1438-1447.

50. Thomas S, Schäfer R, Caro J, Seidel-Morgenstern A. Investigation of mass transfer

through inorganic membranes with several layers. Catal Today. 2001;67:205-216.

51. Jareman F, Hedlund J, Creaser D, Sterte J. Modelling of single gas permeation in real MFI

membranes. J Membr Sci. 2004;236:81-89.

52. De Bruijn FT, Sun L, Olujic Ž, Jansens PJ, Kapteijn F. Influence of the support layer on

the flux limitation in pervaporation. J Membr Sci. 2003;223:141-156.

53. Korelskiy D, Leppäjärvi T, Zhou H, Grahn M, Tanskanen J, Hedlund J. High flux MFI

membranes for pervaporation. J Membr Sci. 2013;427:381-389.

54. Johnson RW. Handbook of fluid dynamics. Boca Raton, FL: CRC Press, 1998.

Page 116: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

55. Korelskiy D, Ye P, Zhou H, Mouzon J, Hedlund J. An experimental study of micropore

defects in MFI membranes. Microporous Mesoporous Mater. 2014;186:194-200.

56. Hernández MG, Salinas-Rodríguez E, Gómez SA, Roa-Neri JAE, Alfaro S, Valdés-

Parada FJ. Helium permeation through a silicalite-1 tubular membrane. Heat Mass Transfer.

2015;51:847-857.

Page 117: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

List of Figure Captions

Fig. 1. Gas permeation and separation set-up.

Fig. 2. Schematic representation of the composite zeolite membrane and pressure profile over

the gas film (G), zeolite film (Z) and support layers 1 (L1) and 2 (L2).

Page 118: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Fig. 3. Experimental N2/He separation factors as a function of temperature for different feed

and permeate pressures.

Page 119: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Fig. 4. Permeances and selectivity for N2/He as a function of temperature in the mixture

separation experiments at a feed pressure of 5.0 bar and a permeate pressure of 0.5 bar (E3) in

the temperature range of a) 85-260 K and b) 73-85 K. The experimental results are presented

with black symbols and the simulated data with red symbols.

Fig. 5. Single gas permeances for pure N2 and He as a function of temperature at a feed

pressure of 5.0 bar and a permeate pressure of 0.5 bar.

Page 120: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

List of tables

Table 1. Feed and permeate pressures used in the binary mixture separation and single gas

permeation experiments.

Experimental Feed Permeate Pressure ratio N2 in the feed

He in the

feed

Feed

flowrate

code (bar) (bar) φ (%) (%) (ml min-1)

E1 3.0 1.0 3 52.6 47.4 420

E2 5.0 1.0 5 52.6 47.4 1300

E3 5.0 0.5 10 52.6 47.4 1300

E4 5.0 0.5 10 100 0 1236

E5 5.0 0.5 10 0 100 1630

Page 121: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Table 2. Summary of parameter values used for modeling.

Parameters values Parameters values for N2 values for He

zρ (kg m-3) 1796 ΔEA,diff (kJ mol-1) 3.1 1.0

zl ( nm) 750 Qads (kJ mol-1) -15.0 -2.9

σ 1/3 D0i at 300K (m2 s-1) 1.35×10-8 10.8×10-8

τ 1.3 η at 300K (kg m-1 s-1) 1.78×10-5 1.99×10-5

dzp (nm) 0.55 satθ , site A (molecules uc-1) 15 77

w (mm) 2.4 satθ , site B (molecules uc-1) 15 120

DC (mm) 19 b, site A (Pa-1) 7×10-7 5×10-9

z 2.5 b, site B (Pa-1) 7.64×10-9 7.95×10-11

KL1 (m) 29.4 × 10-10 x* 2.163 4.4804

effLB 1 (m2) 14.5 × 10-17 y* 7.6895 -2.917

effLB 2 (m2) 6.5 × 10-13 a1 1.9 1.0

a2 -0.3 0.35

*The parameters are used to determine self-exchange diffusivity Dii : Di/Dii = x+y iθ

Page 122: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Table 3. Helium permeance at steady state and relative areas of defects for the membrane,

estimated from permporometry data.

P/P0 Helium permeance at steady state Defect width Defect interval Relative area of defects a

(10-7 mol m-2 s-1 Pa-1) (nm) (nm) (%)

0.00 43.15 - - -

2.72×10-4 2.86 0.71 - -

4.55×10-4 2.07 0.73 0.71 – 0.73 0.07

1.39×10-3 0.95 0.80 0.73 – 0.80 0.08

1.13×10-2 0.34 1.04 0.80 – 1.04 0.04

1.39×10-1 0.19 1.78 1.04 – 1.78 0.005

4.31×10-1 0.14 5.14 1.78 – 5.14 0.001

> 5.14 0.0008

Total: 0.20

a Area of defects divided by the membrane area.

Page 123: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Table 4. Summary of results from simulations and selected experimental data for comparison.

T=106 K T=124 K T=171 K

Δqads for N2 (mol m-3) 960 2500 1200

Δqads for He (mol m-3) 113 149 196

Selectivity, exp 64 76 18

Selectivity, sim 14 53 27

Relative transport through

defects, He (%)

93 72 23

Selectivity, no defects 111 192 38

Concentration polarisation

index Cmem/Cbulk

0.75 0.54 0.5

Relative pressure drop over

support (%)

5 16 27

Page 124: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

PAPER IV

A Uniformly Oriented MFI Membrane for Improved CO2 Separation

Ming Zhou, Danil Korelskiy, Pengcheng Ye, Mattias Grahn, and Jonas Hedlund

Angewandte Chemie (International Edition), 53 (2014) 3492-3495.

Page 125: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 126: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Zeolite Materials Hot PaperDOI: 10.1002/anie.201311324

A Uniformly Oriented MFI Membrane for Improved CO2 Separation**Ming Zhou,* Danil Korelskiy, Pengcheng Ye, Mattias Grahn, and Jonas Hedlund

Abstract: Membrane separation of CO2 from natural gas,biogas, synthesis gas, and flu gas is a simple and energy-efficient alternative to other separation techniques. But resultsfor CO2-selective permeance have always been achieved byrandomly oriented and thick zeolite membranes. Thin, ori-ented membranes have great potential to realize high-flux andhigh-selectivity separation of mixtures at low energy cost. Wenow report a facile method for preparing silica MFI mem-branes in fluoride media on a graded alumina support. In theresulting membrane straight channels are uniformly verticallyaligned and the membrane has a thickness of 0.5 mm. Themembrane showed a separation selectivity of 109 for CO2/H2

mixtures and a CO2 permeance of 51 � 10�7 mol m�2 s�1 Pa�1 at�35 8C, making it promising for practical CO2 separation frommixtures.

The challenge of developing effective separation and purifi-cation technologies that have much smaller energy footprintsis greater for carbon dioxide (CO2) than for other gases.[1] Inaddition to its involvement in climate change, CO2 is animpurity in precombustion natural gas (CO2/CH4), biogas(natural gas produced from biomass), synthesis gas (CO2/H2,the main source of hydrogen in refineries), and postcombus-tion flue gas (CO2/N2).[2] The separation of CO2 from thesemixtures is of great interest from an environmental andenergy perspective: Effectively capturing CO2 from powerplants can have a positive impact on reducing greenhouse gasemissions. CO2 separation from synthesis gas generated frombiomass is necessary to achieve a suitable composition of thegas prior to the synthesis of fuels such as methanol. Removalof CO2 from biogas and natural gas is necessary since CO2

reduces the heating value, causes pipe corrosion, and occupiesvolume in the pipeline. Membrane technology combines highenergy efficiency with simple process equipment, which playsa key role in making the separation economically feasible.Polymeric membranes are already commercially available fornatural gas upgrading;[3] however, polymeric membranesgenerally suffer from low resistance to contaminants, physicalaging, and membrane plasticization at high pressures ofCO2.

[4]

Zeolites are amongst the most widely reported physicaladsorbents for CO2 capture in the patent and journalliterature.[5] They can bring their intrinsic properties, such as

well-defined molecular-sized pores, high thermal, chemical,and mechanical stabilities, and hydrophobicity/philicity, intothe membrane. Ideally, zeolite membranes should be large-scale ordered and defect-free, exhibiting high thermal stabil-ity, selectivity, and permeability.[6] Zeolite SAPO-34,[7, 8]

FAU,[9, 10] DDR,[11] T-type,[12] Ba-ZSM-5,[13] and MFI[14–17]

membranes have been reported for CO2/CH4, CO2/N2, andCO2/H2 separations. These membranes either have demon-strated high selectivity for CO2 over other gas molecules withlow permeability owing to the thickness of the zeolite filmsand the random arrangement of zeolite crystals in themembrane, or have demonstrated high flux of CO2 withmoderate selectivity due to defects present in the membranes.High silica MFI or pure silica MFI zeolite membranes with5.6 � 5.3 � straight channels vertical (b-oriented) to theporous support have demonstrated very promising separa-tions of xylene isomers.[18–20] The separation occurs bya molecular sieving mechanism, because p-xylene and o-xylene molecules are roughly 0.58 nm and 0.68 nm in size,respectively. These b-oriented MFI membranes were sup-ported by homemade porous a-Al2O3 or silica disks, usingtrimer-tetrapropylammonium hydroxide (TPAOH) or tetrae-thylammonium hydroxide (TEAOH)/(NH4)2SiF6 as struc-ture-directing agents (SDAs), and have fewer defects, such aspinholes or cracks, than previously reported MFI membraneswith random, c-, or [h0h]-orientations. Here in this study, forthe first time a uniformly b-oriented MFI membrane isproduced on a commercially available a-Al2O3 disk witha graded pore structure; TPA + F is used as an SDA. The as-prepared MFI membrane shows a significantly low level ofdefects and an improved separation performance for CO2/H2

mixtures.It is well known that zeolite crystals prepared at neutral

pH using F� as the mineralizing agent contain fewer defects[21]

and are more hydrophobic.[22] There are only a few reports onthe preparation of zeolite films using F� as the mineralizingagent. Examples are the preparations of zeolite Beta[23] andMFI[24] films on silicon wafers and porous alumina supports,respectively. However, these films were synthesized fromrandomly oriented seed layers or grown in unfavorable gels,which resulted in films with no or only weak orientation of thecrystals.

The close-packed and uniformly b-oriented MFI seed(1.2 � 1.0 � 0.45 mm3) monolayer was produced on a glass plateby previously reported methods[25] (Figure 1a,d). The seedlayer was immersed into TPA + F gel with a tetraethylorthosilicate (TEOS)/TPAOH/H2O/HF molar compositionof 1:0.2:200:0.2, the reaction was carried out at 150 8C for24 h. The SEM images of the MFI film are shown inFigure 1b,e. The surface morphology was smooth and con-tinuous and showed considerable in-plane growth along the c-axis; the surface crystals typically have dimensions of 10 mm

[*] Dr. M. Zhou, D. Korelskiy, P. Ye, M. Grahn, Prof. J. HedlundChemical Technology, Lule� University of TechnologyPorsçn Campus, 97187 Lule� (Sweden)E-mail: [email protected]

[**] We acknowledge Bio4Energy and the Swedish Foundation forStrategic Research for financially supporting this work. The Knutand Alice Wallenberg Foundation is acknowledged for financiallysupporting the Magellan SEM instrument.

.AngewandteCommunications

3492 � 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2014, 53, 3492 –3495

Page 127: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

along c-axis and 2–3 mm along a-axis, which is in agreementwith their typical long coffin shape. The thickness of this filmwas 1.0 mm. In the same synthesis solution but without addedHF, the seed layer underwent out-of-plane fast growth toform a film with a thickness of 2.5 mm after 24 h of growth at150 8C (Figure 1c,f). Corresponding XRD patterns of seedlayer and MFI films are shown in Figure 1g. The MFI filmgrown in TPA + F gel was b-oriented; five distinct peaks at8.828, 17.748, 26.788, 36.008, and 45.488 were attributed to(020), (040), (060), (080), and (0100) reflections. The MFI filmgrown in TPA solution is a- and b-oriented as evidenced byfive new peaks observed at 8.758, 17.618, 26.588, 35.728, and45.108, which were attributed to (200), (400), (600), (800) and(1000) reflections, consistent with the SEM observations. TheMFI films grown in TPA + F gel appeared quite transparent,which might be attributed to their thin and smooth structureand there was no additional crystal form on the back side ofthe glass plate. In contrast, the MFI films prepared in TPAsolution seemed semitransparent; their film surface wasirregular and the back side of the glass plate was also coveredwith an MFI film by in situ growth, as illustrated bytransmittance spectra and the insert photograph in Figure 1 h.

MFI crystals with a smaller size (0.5 � 0.45 � 0.2 mm3) wereorganized in a b-oriented monolayer on the surface ofa porous alumina disk with a diameter of 25 mm by previouslyreported methods[26–28] (Figure 2a). The seeded aluminasupport was immersed in a TPA + F gel with a SiO2/TPAOH/H2O/HF composition of 1:0.12:60:0.12, using colloi-dal SiO2 30 % as the silica source; the reaction was carried outat 100 8C in oil bath under reflux for 48 h. The resulting MFI

membrane showed primarily in-plane growth with surfacecrystals typically possessing a length of 1.8 mm along the c-axisand 0.6 mm along the a-axis (Figure 2 b). The thickness of thismembrane was 0.5 mm (Figure 2c). In the correspondingXRD pattern (Figure 2d) the five distinct peaks at 8.828,17.748, 26.788, 36.008, and 45.488 were attributed to (020),(040), (060), (080), and (0100) reflections, which confirmedthat the membrane was b-oriented. The four peaks marked byasterisks (*) on the XRD trace indicate peaks of the a-alumina support. The side view of the supported film at lowermagnification revealed its hierarchical structure (Figure 2e),which displayed three sets of pores with sizes in the sequenceof 5.6 � 5.3 �, 100 nm, and 3 mm, respectively (Figure 2 f).

Methods have been developed previously for evaluatingthe defects in membranes; they include: measuring the ratioof single-gas permeances,[6] permporometry,[29] and fluores-cence confocal optical microscopy.[30] Figure 3a shows the n-hexane/helium adsorption-branch permporometry data forthe b-oriented MFI membrane produced in this study andcompares it with the data from a randomly oriented MFImembrane with same thickness prepared in an earlierstudy.[31] At p/p0 = 0, the helium permeance is 67 �10�7 molm�2 s�1 Pa�1 for the randomly oriented membraneand 64 � 10�7 molm�2 s�1 Pa�1 for the b-oriented membrane.When n-hexane is added to the feed (p/p0 = 0.01), the zeolite

Figure 1. Top (a–c) and side (d–f) view SEM images of glass-supported MFI seed layers with crystal sizes of 1.2� 1.0 � 0.45 mm3

(a,d), after growth in TPA +F gel (b, e) and in clear TPA solution (c, f)at 150 8C for 24 h. g) Corresponding XRD patterns. h) Transmittancespectra of glass plates coated with MFI films, which were prepared bygrowth of seed layer in TPA +F gels and in TPA solution (asindicated), and a photograph of the glass plates (insert).

Figure 2. a) SEM image of an MFI seed layer with a crystal size of0.5 � 0.45 � 0.2 mm3 on a porous alumina support. b–e) SEM imagesand XRD pattern of b-oriented MFI membrane after growth of the seedlayer in TPA + F gel at 100 8C for 48 h. f) Representation of orientedMFI channel systems on a graded alumina support.

AngewandteChemie

3493Angew. Chem. Int. Ed. 2014, 53, 3492 –3495 � 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org

Page 128: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

pores are blocked and the helium permeance is dramaticallyreduced. Any remaining helium flow at this point is permeat-ing through defects larger than about 1.6 nm. For randomlyand b-oriented membranes, the helium permeance decreasedby 99.2% and 99.8%, respectively, when the relative pressureof n-hexane increased from 0 to 0.01. Table 1 summarizes thetotal amount of defects, which account for 0.49% and 0.13%of the total surface areas of the randomly and b-oriented MFImembranes, respectively. This indicates the b-oriented mem-brane produced in this study has fewer defects than therandomly oriented membrane with 0.5 mm thickness reportedearlier by our group.[32] The addition of F� is not the onlyreason that the defect area is reduced: the MFI membranegrown in fluoride gel with random orientation has a defect

area of 0.24 %.[24] The uniform b-orientation con-tributes to the decrease of the defect area further to0.13 %.

Separation selectivity and permeance are twoimportant parameters used to evaluate membraneseparation performance. The MFI membrane usedin this study separate CO2/H2 mixture by a compet-itive adsorption[33] and diffusion[15] mechanism. Fig-ure 3b,d shows separation results for 50:50 CO2/H2

and 50:50 CO2/CO mixtures for 0.5 mm thick b-oriented MFI membrane. The membrane showedhighest CO2/H2 selectivity at �35 8C, with a value of109. This is because zeolites are selective to CO2 overH2 at lower temperatures due to strong CO2

adsorption and blocking of H2 permeation. Theselectivity drops with increasing temperature: at�10 8C the observed separation selectivity was 66and at 25 8C the selectivity was about 31. The CO2

permeances were very high in the range of 51–66 �10�7 molm�2 s�1 Pa�1. The membrane showed thehighest CO2/CO selectivity of 21 at �15 8C witha CO2 permeance of 57 � 10�7 mol m�2 s�1 Pa�1. It isimportant that the flux of the membrane is high, sothat the required membrane area can be reducedand industrial applications economically moreaffordable. In the molecular sieving separation ofH2/CO2 mixture the (kinetic diameters of H2 andCO2 are 0.289 nm and 0.33 nm, respectively), thehighest reported H2 permeance was 20 �

10�7 molm�2 s�1 Pa�1 for a silica membrane[34] and ca. 2 �10�7 molm�2 s�1 Pa�1 for a graphene oxide membrane.[35]

An upper bound can be used to compare the separationperformance of the new membrane with that of previouslyreported membranes. Figure 3c shows a comparison of our b-oriented MFI membrane with previously reported zeolitemembranes with random orientations for the separation ofCO2/H2 mixtures. The selectivity of the b-oriented MFImembrane is above the upper bound and its separationperformance is superior to that of the randomly orientedSAPO-34, FAU, Ba-ZSM-5, and MFI membranes.

In summary, a b-oriented zeolite MFI membrane witha thickness of 0.5 mm was fabricated on a commerciallyavailable a-alumina support by a facile TPA-fluoride route.This membrane showed CO2/H2 and CO2/CO mixture sepa-ration performances that are higher than those of the state-of-the-art random-orientation zeolite membranes; this makes itattractive for practical CO2 separation from mixtures.

Experimental SectionSynthesis of MFI seeds: Plate-shaped MFI crystals with crystal sizesof 1.2 � 1.0 � 0.45 mm3 and 0.5 � 0.45 � 0.2 mm3 were synthesized frommixtures with a TEOS/TPAOH/H2O molar composition of 1:0.2:100but with different hydrolysis time. Two identical solutions wereshaken at room temperature for 1 day and 7 days, respectively, beforethey were subjected to hydrothermal treatment at 130 8C for 9 hunder constant stirring. Crystals synthesized from both solutions werepurified by repetitive centrifugation and re-dispersion in distilleddeionized water and freeze-dried before use.

Figure 3. a) Permporometry patterns for randomly and b-oriented MFI mem-branes; insert summarizes the defect areas of the two membranes. b,d) Resultsfor the separation of 50:50 CO2/H2 (b) and CO2/CO (d) gas mixtures with the b-oriented MFI membrane. c) Comparison of selectivity versus CO2 permeance forthe b-oriented MFI membrane, the randomly oriented MFI membrane, and othertypes of zeolite membranes for CO2/H2 gas mixture separation; the straight linedenotes the upper bound of zeolite membranes for CO2/H2 separation, data inthe dotted oval labeled 10 are the results for the b-oriented MFI membrane fromthis study. See Table 1 for an explanation of points 1–10.

Table 1: Explanation of data for points 1–10 in Figure 3.

No. T [8C] Membrane Thickness[mm]

Orientation Ref.

1 �20 to 23 SAPO-34 5 random [8]2 60 FAU 15–18 random [9]3 35 FAU (NaY) 3 random [10]4 25 Ba-ZSM-5 0.55 random [13]5 60 MFI

(modified)0.55 random [36]

6 25 MFI 7 random [15]7 20 MFI 40–50 c, (h0l) [14]8 25 MFI 0.5 random [24]9 25 MFI 0.7 random [17]10 �35 to 25 MFI 0.5 b this

work

.AngewandteCommunications

3494 www.angewandte.org � 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2014, 53, 3492 –3495

Page 129: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Assembly of monolayer: Monolayers of b-oriented MFI crystals(1.2 � 1.0 � 0.45 mm3) were assembled on glass plates by rubbing thepowders of freeze-dried crystals onto the surfaces using a Nitrile-gloved finger.[25] The porous-alumina-supported MFI seed layers(0.5 � 0.45 � 0.2 mm3) were produced by dynamic interfacial assem-bly[26] or by polymer-mediated assembly.[27, 28]

Growth of membranes: The glass-supported seed layer wasimmersed in a gel with a TEOS/TPAOH/H2O/HF molar composition1:0.2:200:0.2 and the reaction was carried out in autoclave at 150 8Cfor 24 h. The alumina-supported seed layer was subjected to hydro-thermal treatment in a gel with a SiO2/TPAOH/H2O/HF molarcomposition of 1:0.12:60:0.12, using colloidal SiO2 30% as the silicasource; the reaction was carried out in an oil bath at 100 8C underreflux for 48 h. After synthesis, the membranes were rinsed witha 0.2m ammonia solution and dried at 50 8C over night. Calcinationwas then performed at 500 8C for 6 h, with a heating rate of0.2 8Cmin�1 and a cooling rate of 0.3 8C min�1.

Permporometry and separations: After calcination, the mem-brane was mounted in a stainless steel cell sealed with graphitegaskets and copper gaskets. Then, the measurement of n-hexane/helium adsorption-branch permporometry was performed. Prior tothe separation experiments, the membrane was dried for 6 h at 300 8Cin a flow of nitrogen. CO2/H2 (1:1) and CO2/CO (1:1) mixtures werefed to the membrane at total feed pressure of 9 bar. The totalpermeate pressure was kept at 1 bar. The composition in both feedand permeate streams is known from GC analysis. No sweep gas wasused.

Received: December 31, 2013Published online: March 3, 2014

.Keywords: carbon dioxide · gas separation · membranes ·thin films · zeolites

[1] P. Nugent, Y. Belmabkhout, S. D. Burd, A. J. Cairns, R. Luebke,K. Forrest, T. Pham, S. Q. Ma, B. Space, L. Wojtas, M. Eddaoudi,M. J. Zaworotko, Nature 2013, 495, 80 – 84.

[2] D. M. D’Alessandro, B. Smit, J. R. Long, Angew. Chem. 2010,122, 6194 – 6219; Angew. Chem. Int. Ed. 2010, 49, 6058 – 6082.

[3] S. Basu, A. L. Khan, A. Cano-Odena, C. Liu, I. F. J. Vankelecom,Chem. Soc. Rev. 2010, 39, 750 – 768.

[4] Y. Xiao, B. T. Low, S. S. Hosseini, T. S. Chung, D. R. Paul, Prog.Polym. Sci. 2009, 34, 561 – 580.

[5] S. Choi, J. H. Drese, C. W. Jones, ChemSusChem 2009, 2, 796 –854.

[6] M. Yu, R. D. Noble, J. L. Falconer, Acc. Chem. Res. 2011, 44,1196 – 1206.

[7] S. G. Li, J. L. Falconer, R. D. Noble, Adv. Mater. 2006, 18, 2601 –2603.

[8] M. Hong, S. Li, J. L. Falconer, R. D. Noble, J. Membr. Sci. 2008,307, 277 – 283.

[9] J. Caro, M. Noack, Microporous Mesoporous Mater. 2008, 115,215 – 233.

[10] K. Kusakabe, T. Kuroda, K. Uchino, Y. Hasegawa, S. Morooka,AIChE J. 1999, 45, 1220 – 1226.

[11] S. Himeno, T. Tomita, K. Suzuki, K. Nakayama, K. Yajima, S.Yoshida, Ind. Eng. Chem. Res. 2007, 46, 6989 – 6997.

[12] X. L. Zhang, L. F. Qiu, M. Z. Ding, N. Hu, F. Zhang, R. F. Zhou,X. S. Chen, H. Kita, Ind. Eng. Chem. Res. 2013, 52, 16364 –16374.

[13] J. Lindmark, J. Hedlund, J. Membr. Sci. 2010, 360, 284 – 291.[14] H. L. Guo, G. S. Zhu, H. Li, X. Q. Zou, X. J. Yin, W. S. Yang,

S. L. Qiu, R. Xu, Angew. Chem. 2006, 118, 7211 – 7214; Angew.Chem. Int. Ed. 2006, 45, 7053 – 7056.

[15] M. Kanezashi, Y. S. Lin, J. Phys. Chem. C 2009, 113, 3767 – 3774.[16] M. P. Bernal, J. Coronas, M. Men�ndez, J. Santamar�a, AIChE J.

2004, 50, 127 – 135.[17] L. Sandstrçm, E. Sjçberg, J. Hedlund, J. Membr. Sci. 2011, 380,

232 – 240.[18] Z. P. Lai, G. Bonilla, I. Diaz, J. G. Nery, K. Sujaoti, M. A. Amat,

E. Kokkoli, O. Terasaki, R. W. Thompson, M. Tsapatsis, D. G.Vlachos, Science 2003, 300, 456 – 460.

[19] T. C. T. Pham, H. S. Kim, K. B. Yoon, Science 2011, 334, 1533 –1538.

[20] T. C. T. Pham, T. H. Nguyen, K. B. Yoon, Angew. Chem. 2013,125, 8855 – 8860; Angew. Chem. Int. Ed. 2013, 52, 8693 – 8698.

[21] J. L. Guth, H. Kessler, R. Wey, Stud. Surf. Sci. Catal. 1986, 28,121 – 128.

[22] K. Zhang, R. P. Lively, J. D. Noel, M. E. Dose, B. A. McCool,R. R. Chance, W. J. Koros, Langmuir 2012, 28, 8664 – 8673.

[23] L. Tosheva, M. Hçlzl, T. H. Metzger, V. Valtchev, S. Mintova, T.Bein, Mater. Sci. Eng. C 2005, 25, 570 – 576.

[24] H. Zhou, D. Korelskiy, E. Sjçberg, J. Hedlund, MicroporousMesoporous Mater. 2014, in press.

[25] J. S. Lee, J. H. Kim, Y. J. Lee, N. C. Jeong, K. B. Yoon, Angew.Chem. 2007, 119, 3147 – 3150; Angew. Chem. Int. Ed. 2007, 46,3087 – 3090.

[26] M. Zhou, J. Hedlund, J. Mater. Chem. 2012, 22, 3307 – 3310.[27] M. Zhou, J. Hedlund, J. Mater. Chem. 2012, 22, 24877 – 24881.[28] B. Q. Zhang, M. Zhou, X. F. Liu, Adv. Mater. 2008, 20, 2183 –

2189.[29] J. Hedlund, D. Korelskiy, L. Sandstrom, J. Lindmark, J. Membr.

Sci. 2009, 345, 276 – 287.[30] G. Bonilla, M. Tsapatsis, D. G. Vlachos, G. Xomeritakis, J.

Membr. Sci. 2001, 182, 103 – 109.[31] D. Korelskiy, T. Leppajarvi, H. Zhou, M. Grahn, J. Tanskanen, J.

Hedlund, J. Membr. Sci. 2013, 427, 381 – 389.[32] J. Hedlund, J. Sterte, M. Anthonis, A. J. Bons, B. Carstensen, N.

Corcoran, D. Cox, H. Deckman, W. De Gijnst, P. P. de Moor, F.Lai, J. McHenry, W. Mortier, J. Reinoso, Microporous Meso-porous Mater. 2002, 52, 179 – 189.

[33] T. C. Golden, S. Sircar, J. Colloid Interface Sci. 1994, 162, 182 –188.

[34] R. M. de Vos, H. Verweij, Science 1998, 279, 1710 – 1711.[35] H. Li, Z. N. Song, X. J. Zhang, Y. Huang, S. G. Li, Y. T. Mao,

H. J. Ploehn, Y. Bao, M. Yu, Science 2013, 342, 95 – 98.[36] J. Lindmark, J. Hedlund, J. Mater. Chem. 2010, 20, 2219 – 2225.

AngewandteChemie

3495Angew. Chem. Int. Ed. 2014, 53, 3492 –3495 � 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org

Page 130: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

PAPER V

Efficient Ceramic Zeolite Membranes for CO2/H2 Separation

Danil Korelskiy, Pengcheng Ye, Shahpar Fouladvand, Somayeh Karimi, Erik Sjöberg, and Jonas Hedlund

Journal of Materials Chemistry A, 3 (2015) 12500-12506.

Page 131: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 132: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Journal ofMaterials Chemistry A

PAPER

Publ

ishe

d on

14

May

201

5. D

ownl

oade

d on

29/

12/2

015

15:1

8:28

.

View Article OnlineView Journal | View Issue

Efficient ceramic

Chemical Technology, Lulea University of

E-mail: [email protected]

Cite this: J. Mater. Chem. A, 2015, 3,12500

Received 24th March 2015Accepted 14th May 2015

DOI: 10.1039/c5ta02152a

www.rsc.org/MaterialsA

12500 | J. Mater. Chem. A, 2015, 3, 125

zeolite membranes for CO2/H2

separation

D. Korelskiy,* P. Ye, S. Fouladvand, S. Karimi, E. Sjoberg and J. Hedlund

Membranes are considered one of the most promising technologies for CO2 separation from industrially

important gas mixtures like synthesis gas or natural gas. In order for the membrane separation process

to be efficient, membranes, in addition to being cost-effective, should be durable and possess high flux

and sufficient selectivity. Current CO2-selective membranes are low flux polymeric membranes with

limited chemical and thermal stability. In the present work, robust and high flux ceramic MFI zeolite

membranes were prepared and evaluated for separation of CO2 from H2, a process of great importance

to synthesis gas processing, in a broad temperature range of 235–310 K and at an industrially relevant

feed pressure of 9 bar. The observed membrane separation performance in terms of both selectivity and

flux was superior to that previously reported for the state-of-the-art CO2-selective zeolite and polymeric

membranes. Our initial cost estimate of the membrane modules showed that the present membranes

were economically viable. We also showed that the ceramic zeolite membrane separation system would

be much more compact than a system relying on polymeric membranes. Our findings therefore suggest

that the developed high flux ceramic zeolite membranes have great potential for selective, cost-effective

and sustainable removal of CO2 from synthesis gas.

Introduction

Efficient and sustainable CO2 separation and capture technol-ogies are currently of tremendous interest for several reasons.Firstly, CO2 is a greenhouse gas, and combustion of fossil fuelsis one of the major sources of CO2 emissions. Secondly, CO2 isan undesired component in many industrial gas streams, suchas natural gas, biogas (methane produced from biomass), andsynthesis gas, including bio-syngas produced by biomass gasi-cation.1 Removal of CO2 from syngas is a requirement forfurther processing, such as production of liquid fuels, e.g.,methanol,2 and hydrogen at reneries, petrochemical plants,and Integrated Gasication Combined Cycle (IGCC) powerplants.3 Today, CO2 is removed primarily by absorption, e.g.,amine scrubbing, which is rather an energy-intensive methodwith high capital costs.4 In addition, the used absorbents arecorrosive and environmentally unfriendly, and the absorptionunit is quite large and complex.

Over the past decades, membrane separation technologieshave gained an increasing interest for the reasons of high effi-ciency, sustainability and low energy consumption. Currently,membranes are considered to be one of the most promisingCO2 separation and capture technologies with great marketpotential.4,5 For instance, the amount of energy required for a90% recovery of CO2 using an efficient membrane has been

Technology, SE-97187 Lulea, Sweden.

00–12506

estimated to be ca. 16% of the power produced by the powerplant,6 whereas the energy required by an amine absorption/desorption process is ca. 50% of the power.7 From thecommercial point of view, polymeric membranes have been themost successful membrane type thus far.4 For instance, theMTR Polaris™ membranes8 have been the rst commercialpolymeric membranes able to separate CO2 from synthesis gas.Today's best polymeric membranes can achieve CO2/H2 selec-tivities of 10–12 with a CO2 permeance of ca. 2 � 10�7 mol s�1

m�2 Pa�1 at room temperature.9 Such a low permeance coupledwith the fairly poor selectivity necessitates the use of quite largemembrane areas for a given separation task. In addition, poly-meric membranes suffer from plasticisation induced by CO2,which signicantly reduces the membrane selectivity andstability over time.4

Among ceramic membranes, zeolite membranes are espe-cially attractive and promising.5 These membranes are micro-porous aluminosilicate membranes with a well-dened poresystem.10 Due to the porous structure, zeolite membranes candisplay much higher uxes than polymeric membranes,11 i.e., amuch smaller membrane area would be needed for a givenseparation task. Additionally, ceramic zeolite membranes offeran advantage over polymeric membranes in terms of highchemical and thermal stability.12

Despite the great interest in synthesis gas upgrading usingmembranes, the number of studies devoted to evaluation ofzeolite membranes for this application is small.5 Whereashighly CO2-selective zeolite membranes have been developed,

This journal is © The Royal Society of Chemistry 2015

Page 133: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Paper Journal of Materials Chemistry A

Publ

ishe

d on

14

May

201

5. D

ownl

oade

d on

29/

12/2

015

15:1

8:28

. View Article Online

e.g., SAPO-34 membranes13 with a CO2/H2 separation factor of110 at 253 K and a feed pressure of 12 bar, there are only a fewreports on high ux zeolite membranes. Our research group hasextensive experience in preparing ultra-thin (ca. 0.5–1 mm) highux MFI zeolite membranes,14 and these membranes have beenevaluated for various gas2,14–19 and liquid20 separations. In thepresent work, these membranes were evaluated for separationof CO2 from H2 (CO2/H2 mixtures are typically considered as amodel system for synthesis gas21) in a wide temperature range of235–310 K and at a feed pressure of 9 bar.

ExperimentalMembrane synthesis

Supported zeolite membranes comprised of an H-ZSM-5 lmwith a thickness of ca. 0.5 mm and a Si/Al ratio of 139 (ref. 17)were prepared as described in detail in our earlier work.14 Aporous graded a-alumina disc (Fraunhofer IKTS, Germany) wasused as the support. Prior to the lm synthesis, the supportswere masked as described elsewhere22 and then seeded withcolloidal MFI crystals of 50 nm in diameter. The lm synthesiswas carried out for 36 h at 100 �C in a solution with a molarcomposition of 3TPAOH : 25SiO2 : 1450H2O : 100C2H5OH.Aer the synthesis, the membranes were rinsed with a0.1 M Ammonia solution for 24 h and then calcined for 6 h at500 �C at a heating rate of 0.2 �C min�1 and a cooling rate of0.3 �C min�1.

Membrane characterisation

Scanning electron microscopy (SEM) characterisation of themembranes was carried out using a Magellan 400 (the FEICompany, Eindhoven, the Netherlands) instrument with nocoating. Cross-sections of the membranes were obtained byfracture with a pair of cutting pliers.

X-ray diffraction (XRD) characterisation of the membraneswas performed using a PANalytical Empyrean diffractometerequipped with a Cu LFF HR X-ray tube and a PIXcel3D detector.The data evaluation was performed using HighScore Plus 3.0.4.

The prepared membranes were also characterised by n-hexane/helium permporometry15 as described in detail in ourearlier work23 and in brief below. Themembranes were sealed ina stainless steel cell using graphite gaskets (Eriks, the Nether-lands). In order to remove any adsorbed compounds, themembranes were heated to 300 �C at a heating rate of 1 �Cmin�1 and kept at this temperature for 6 h in a ow of purehelium. Permporometry characterisation was carried out at 50�C and a total pressure difference across the membrane of 1 barwith the permeate stream kept at atmospheric pressure. Therelative pressure of n-hexane was raised in a step-wise mannerfrom 0 to ca. 0.4. At each relative pressure, the system wasallowed to achieve steady-state. For removing n-hexane from thepermeate stream, a condenser kept at �40 �C followed by acolumn packed with activated carbon was used. The permeatevolumetric ow rate was measured with a soap bubble owmeter. A detailed procedure for estimation of the relative areasof defects from permporometry data is given in our earlier

This journal is © The Royal Society of Chemistry 2015

work.23 In brief, the defect width was calculated from n-hexanerelative pressure using either the Horvath–Kavazoe equation(micropore range defects) or the Kelvin equation (mesoporerange defects). For each defect interval, the average defect widthwas then calculated. Based on the average defect width, theaverage helium diffusivity in each defect interval was estimatedusing the gas-translational model. Knowing the diffusivity, thehelium molar ux was further calculated from Fick's law.Finally, the defect area was estimated as the ratio betweenhelium molar ow and ux through the defects in that partic-ular interval.

Separation experiments

Separation experiments were carried out using an equimolarmixture of CO2 and H2. The membrane was in the same cell asused for the permporometry experiment. The total feed pres-sure was kept at 9 bar, whereas the total permeate pressurewas atmospheric. All experiments were performed withoutsweep gas. Prior to the experiments, the membrane wasushed with pure helium for 6 h at 300 �C in order to removeany adsorbed species. The permeate volumetric ow rate wasmeasured with a drum-type gasmeter (TG Series, RitterApparatebau GmbH) and the permeate composition was ana-lysed on-line with a mass spectrometer (GAM 400, InProcessInstruments).

The ux of component i, Ji (mol s�1 m�2), was estimatedfrom the measured molar ow of the corresponding componentthrough the membrane, Fi (mol s�1) as

Ji ¼ Fi/A,

where A is the membrane area (m2).The permeance of component i, Pi (mol s�1 m�2 Pa�1), was

calculated from the ux of the corresponding componentthrough the membrane as

Pi ¼ Ji/DPi,

where DPi (Pa) is the partial pressure difference of component iacross the membrane.

The separation factor bi/j was estimated as

bi/j ¼ (yi/yj)/(xi/xj),

where x and y are the molar fractions in the feed and permeate,respectively.

The membrane selectivity ai/j was estimated as

ai/j ¼ Pi/Pj.

Results and discussionMembrane characterisation

The fabricated membranes were H-ZSM-5 zeolite lms with a Si/Al ratio of about 139 (ref. 17) supported on commerciala-alumina discs (Fraunhofer IKTS, Germany). The synthesisprocedure is described in the Experimental. Cross-sectional and

J. Mater. Chem. A, 2015, 3, 12500–12506 | 12501

Page 134: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Fig. 2 An XRD pattern of membrane M2. The reflection marked withan asterisk emanates from the alumina support.

Journal of Materials Chemistry A Paper

Publ

ishe

d on

14

May

201

5. D

ownl

oade

d on

29/

12/2

015

15:1

8:28

. View Article Online

top-view SEM images of an as-synthesised membrane are shownin Fig. 1. The zeolite lm appears to be even with a thickness ofca. 0.5 mm. The crystals composing the lm are well-intergrownwith a size of ca. 200 nm. No large defects (>5 nm) could bedetected by SEM, indicating high quality of the membrane.Fig. 2 shows an XRD pattern of membrane M2. The detectedreections were solely the expected reections emanating fromMFI zeolite and alumina (the support) indicating that no otherphase was present in the membrane.

In order to estimate the amount of defects, the membraneswere characterised by n-hexane/helium permporometry15,23 asdescribed in the Experimental. In this technique, helium per-meance through the membrane is measured as a function of n-hexane relative pressure. Table 1 reports permporometry datafor membrane M1. The helium permeance at a relative pressureof n-hexane of 0, i.e., the permeance through zeolite pores and

Fig. 1 Cross-sectional (a) and top-view (b) SEM images of an as-synthesised membrane.

12502 | J. Mater. Chem. A, 2015, 3, 12500–12506

defects, was as high as 53 � 10�7 mol s�1 m�2 Pa�1, whichshows that the zeolite pores are open and rather permeable. Asthe relative pressure of n-hexane was increased, rst zeolitepores and then increasingly larger defects were blocked by n-hexane, and, therefore, the helium permeance decreased. Theamount of defects in terms of relative areas was estimated fromthe permporometry data as described in the Experimental. Thetotal amount of defects in the membrane was very low, consti-tuting less than 0.1% of the total membrane area, indicating avery high quality of the membrane. The main type of defects (ca.99.4% of all defects) was micropore defects, i.e., defects < 2 nmin size. Such defects are most likely narrow open grain bound-aries, as discussed in detail in our previous work.24 Essentiallyno large defects (>5 nm) were detected by permporometry,which is consistent with the SEM observations.

Separation experiments

Fig. 3 shows permeances of CO2 and H2 measured formembrane M1 as a function of temperature when a 50/50 (v/v)mixture of CO2/H2 was fed to the membrane. The permeance ofCO2 was high and much greater than that of H2 in the entiretemperature range. This is a result of the fact that CO2 isadsorbing much stronger in the membrane than H2,2 therebyblocking the transport of H2 and rendering the membrane CO2-selective. The highest CO2 permeance of ca. 78 � 10�7 mol s�1

m�2 Pa�1 was observed at the higher temperatures, i.e., 290–310K. In general, the measured CO2 permeances were consistentwith those previously reported by our group,2 and one to twoorders of magnitude higher than those reported for zeolite andpolymeric membranes in the literature. With decreasingtemperature, the permeances of both CO2 and H2 decreased,most likely due to decreasing diffusivity, as discussed in ourearlier work.2 However, the permeance of CO2 was reduced to asignicantly lower extent than that of H2 resulting in increasingselectivity of the membrane to CO2 with decreasing tempera-ture, which can be ascribed to increasing adsorption of CO2

with decreasing temperature. At the lowest investigatedtemperature of 235 K, the permeance of H2 was as low as 0.3 �

This journal is © The Royal Society of Chemistry 2015

Page 135: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Table 1 Permporometry data for membrane M1

P/P0He permeance (10�7

mol s�1 m�2 Pa�1)Defect interval(nm)

Relative areaof defectsa (%)

0 53 —3.8 � 10�4 1.25 0.71–0.73 0.066.7 � 10�4 0.70 0.73–0.80 0.032.1 � 10�3 0.36 0.80–1.04 0.011.1 � 10�2 0.23 1.04–1.78 0.0031.5 � 10�1 0.11 1.78–5.43 04.5 � 10�1 0.11 >5.43 0.0006

Total: 0.10

a Area of defects per total membrane area.

Fig. 3 Permeances of CO2 and H2 measured for membrane M1 as afunction of temperature.

Paper Journal of Materials Chemistry A

Publ

ishe

d on

14

May

201

5. D

ownl

oade

d on

29/

12/2

015

15:1

8:28

. View Article Online

10�7 mol s�1 m�2 Pa�1, whereas the permeance of CO2 was stillas high as 62 � 10�7 mol s�1 m�2 Pa�1.

Fig. 4 illustrates CO2/H2 separation factors recorded formembrane M1 as a function of temperature. With decreasingtemperature, the separation factor was increasing to as high as165 at the lowest investigated temperature of 235 K. At theseconditions, the CO2 concentration in the permeate was as high

Fig. 4 CO2/H2 separation factor recorded for membrane M1 as afunction of temperature.

This journal is © The Royal Society of Chemistry 2015

as 99.4%. Table 2 shows the CO2 uxes, the concentration ofCO2 and H2 in the permeate stream and the CO2/H2 membraneselectivities. The latter term denotes the ratio of CO2 and H2

permeances (not to be confused with the separation factor). Theobserved CO2 ux was very high, i.e., 350–420 kg m

�2 h�1, in theentire temperature range. As discussed in our earlier work,2 thehigh CO2 ux is a result of the very low zeolite lm thickness,strong CO2 adsorption and high CO2 diffusivity in the zeolitepores, and a relatively high CO2 partial pressure difference of3.5 bar across the membrane. The CO2 ux was decreasing withdecreasing temperature, i.e., similar to the CO2 permeance.Since the membrane was highly CO2-selective in the entiretemperature range, the CO2 concentration in the permeate wasclose to 100%, see Table 2. Consequently, the partial pressure ofCO2 in the permeate was nearly constant at 1 bar, resulting inalmost the same partial pressure difference of CO2 across themembrane (ca. 3.5 bar) at all temperatures. As a result, the CO2

ux was varying with temperature in an almost identicalmanner as the CO2 permeance. At 253 K, the separation factorwas almost as high as 120 with a CO2 ux of ca. 400 kg h�1 m�2,which is 133 times higher than that (3 kg h�1 m�2) reported forthe highly CO2-selective SAPO-34 zeolite membranes at similarexperimental conditions.13 It is also worth noting that the totalduration of the separation experiments was ca. 6 h. During thistime, the membrane was constantly exposed to a high ow ofgas at elevated pressure. Despite this, no indication of deterio-rating membrane quality was observed indicating goodmembrane stability at these experimental conditions. Evalua-tion of the long-term stability of the membranes would,however, require an industrial gas supply due to the largeconsumption of gas and the associated high costs, which wasbeyond the scope of the present work.

In order to study reproducibility of the separation results,another membrane (denoted M2) with defect distributionsimilar to that for membrane M1 was evaluated for CO2/H2

separation in the temperature range of 235–270 K using a feedpressure of 9 bar. The separation data for membrane M2 sum-marised in Table 3 were well comparable to those for membraneM1, illustrating good reproducibility of the separation results.

Table 2 CO2 flux, permeate concentration and CO2/H2 membraneselectivity observed for membrane M1

T (K)CO2 ux(kg h�1 m�2)

Permeateconcentration(mol%)

CO2/H2 membraneselectivityCO2 H2

310 423 93.22 6.78 17300 429 95.36 4.61 26290 428 96.71 3.28 37270 420 98.47 1.53 82260 406 98.91 1.08 117250 383 99.20 0.80 159240 364 99.33 0.67 189235 356 99.40 0.60 210

J. Mater. Chem. A, 2015, 3, 12500–12506 | 12503

Page 136: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Table 3 CO2/H2 separation data recorded for membrane M2

T (K)CO2 ux(kg h�1 m�2)

CO2/H2 separationfactor

CO2/H2 membraneselectivity

270 448 84 107260 407 114 145250 404 129 165240 341 189 242235 303 202 258

Fig. 5 Summary of the best CO2/H2 separation data reported forzeolite membranes in the literature2,13,16,31–33 as well as the dataobtained in our previous work25 and in the present work.

Journal of Materials Chemistry A Paper

Publ

ishe

d on

14

May

201

5. D

ownl

oade

d on

29/

12/2

015

15:1

8:28

. View Article Online

A summary of the best CO2/H2 separation data reported forzeolite membranes in the literature is depicted in Fig. 5. Fig. 5also shows the data obtained in the present work for randomlyoriented MFI membranes, and in our previous work25 for b-oriented MFI membranes. The separation performance of themembranes prepared in the present work is well above theupper bound for the best zeolite membranes reported previ-ously. The observed separation performance was also greaterthan that of high quality b-oriented MFI membranes recentlyprepared by our group.25 Since the amount of defects in bothtypes of membranes was nearly identical, the difference in theseparation performance between the randomly oriented and b-oriented MFI membranes should most likely emanate from the

Table 4 A comparison between the cost of commercial-scale MTR Polmembranes prepared in the present work for separation of 300 ton CO

Parameter Polar

CO2 permeance (10�8 mol s�1 m�2 Pa�1) 20 (rModule type SpiraMembrane area in one module (m2) 20 (rMembrane area needed (m2) 395No. of modules needed 20Cost of membranes and module (USD per m2) 10 (rTotal cost of modules with membranes (USD) 39 50

a The cost of the module was estimated by Fraunhofer IKTS (Dr Ing. H. R

12504 | J. Mater. Chem. A, 2015, 3, 12500–12506

difference in the adsorption affinity of the membranes for CO2.The b-oriented MFI membranes reported in our previous work25

were prepared in a uoride medium at near-neutral pH,whereas the membranes in the present work were synthesisedin an alkaline medium. MFI zeolites prepared in a uoridemedium have been shown26,27 to be less hydrophilic thansimilar MFI zeolites prepared in a hydroxide medium due to thelower amount of Si–OH groups. In addition, the b-oriented MFI-F membranes prepared in our previous work25 should mostprobably contain less aluminium in the structure than thepresent MFI-OHmembranes as the leaching of aluminium fromthe support during the lm synthesis should be reduced at near-neutral pH. The lower aluminium content should also result ina less hydrophilic nature of the b-orientedMFI-Fmembranes. Atthe same time, the adsorption affinity of MFI zeolites for CO2

has been demonstrated28–30 to increase with increasing hydro-philicity. Thus, the present randomly oriented MFI-OHmembranes, being somewhat more hydrophilic, should havegreater adsorption affinity for CO2 than the b-oriented MFI-Fmembranes, and, hence, should be more selective to CO2, asobserved in the present work. It is also worth noting that in aprevious work19 we compared randomly oriented MFImembranes prepared in uoride and alkaline media. A similartrend was observed for CO2/H2 separation, i.e., the MFI-OHmembranes were more selective to CO2 than the MFI-Fmembranes. In contrast, the latter membranes were moreselective to n-butanol, as should be expected for a less hydro-philic membrane. It should also be noted that the preparationprocedure for the randomly oriented MFI membranes is ratherwell-established and it is much simpler than that for the b-oriented MFI membranes. Hence, at this moment, therandomly oriented high ux MFI membranes should be easierto scale-up.

Cost estimation

In order to evaluate the economic viability of our membranes,the estimated cost of the membrane modules was comparedwith that of the commercially available spiral-wound modulesused in a natural gas processing plant.34 The latter moduleswere assumed to contain MTR Polaris™ membranes recentlyevaluated for CO2/H2 separation in commercial scale.3 Thezeolite membrane modules were assumed to contain zeolite

aris™ membrane modules and the cost of modules with high flux MFI

2 per day

is membranes MFI membranes

ef. 3) 775l-wound Multichannel tubes (19 channels)ef. 3) 10 (ref. 35)

101

ef. 34) 2600a

0 26 500

ichter, personal communication, 6 March 2015).

This journal is © The Royal Society of Chemistry 2015

Page 137: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Fig. 6 A comparison between the size of an amine scrubber system,polymeric membrane system and high flux MFI membrane systemperforming the same separation task. The background picture wasadapted from Dortmundt and Doshi.36 The ceramic membranemodule image was provided by Inopor®.35

Paper Journal of Materials Chemistry A

Publ

ishe

d on

14

May

201

5. D

ownl

oade

d on

29/

12/2

015

15:1

8:28

. View Article Online

membranes supported on 19-channel a-alumina tubes with thesame CO2 permeance as measured experimentally for the disc-shaped membranes in the present work. The results of the costcomparison are summarised in Table 4. The costs were esti-mated for amembrane process with a separation capacity of 300ton CO2 per day at a CO2 partial pressure difference across themembrane of 10 bar and room temperature. For this purpose, apolymeric membrane process would need as many as 20membrane modules, whereas a ceramic MFI zeolite membraneprocess would only require one module. The estimationdemonstrates that the total cost of modules with membranes inthe case of high ux MFI membranes was approx. 30% lowerthan that of high performance commercial polymericmembranes. This is due to the much greater permeance of theMFI membranes resulting in a very low membrane area neededfor the separation process. Furthermore, the MFI membranesdisplay much higher CO2/H2 selectivity (26 and 210 at 300 and235 K, respectively, see Table 2) than polymeric membranes (10–12 at room temperature). It is also worth noting that theequipment needed for the high ux MFI membrane process willbe very compact, as schematically illustrated in Fig. 6. Hence,the prepared MFI membranes have great market potential forseparation of CO2 from synthesis gas.

Conclusions

Ultra-thin randomly oriented high ux MFI zeolite membraneswere prepared and evaluated for CO2/H2 separation at atemperature ranging from 235 to 310 K and a feed pressure of 9bar. The observed membrane separation performance in termsboth selectivity and ux was superior to that previously reportedfor CO2-selective zeolite and polymeric membranes. An initialestimate of the cost of membrane modules revealed that thepresent membranes were more economically attractive than

This journal is © The Royal Society of Chemistry 2015

commercial-scale polymeric membranes. In addition, theceramic zeolite membrane separation system would be muchmore space efficient than a system relying on polymericmembranes. The ndings of the present work therefore suggestthat the developed high ux MFI zeolite membranes have greatpotential for selective and cost-effective removal of CO2 fromsynthesis gas.

Acknowledgements

The Swedish Foundation for Strategic Research (SSF, grant no.RMA08-0018), the Swedish Research Council (VR, grant no.2010-5317), the Swedish Research Council Formas (Grant no.213-2013-1684), the Swedish Energy Agency (Grant no. 2013-006587) and Bio4Energy are gratefully acknowledged fornancially supporting this work. Amirfarrokh Farzaneh isgratefully acknowledged for fruitful discussions.

Notes and references

1 X. Yin, D. Y. C. Leung, J. Chang, J. Wang, Y. Fu and C. Wu,Energy Fuels, 2004, 19, 305–310.

2 L. Sandstrom, E. Sjoberg and J. Hedlund, J. Membr. Sci., 2011,380, 232–240.

3 H. Lin, Z. He, Z. Sun, J. Vu, A. Ng, M. Mohammed, J. Kniep,T. C. Merkel, T. Wu and R. C. Lambrecht, J. Membr. Sci., 2014,457, 149–161.

4 S. D. Kenarsari, D. Yang, G. Jiang, S. Zhang, J. Wang,A. G. Russell, Q. Wei and M. Fan, RSC Adv., 2013, 3, 22739–22773.

5 M. Pera-Titus, Chem. Rev., 2013, 114, 1413–1492.6 T. C. Merkel, H. Lin, X. Wei and R. Baker, J. Membr. Sci., 2010,359, 126–139.

7 P. Luis, T. Van Gerven and B. Van der Bruggen, Prog. EnergyCombust. Sci., 2012, 38, 419–448.

8 Membrane Technology & Research (MTR), http://www.mtrinc.com.

9 R. W. Baker and B. T. Low, Macromolecules, 2014, 47, 6999–7013.

10 J. Caro and M. Noack, Microporous Mesoporous Mater., 2008,115, 215–233.

11 E. Sjoberg, L. Sandstrom, O. G. W. Ohrman and J. Hedlund,J. Membr. Sci., 2013, 443, 131–137.

12 J. Caro, M. Noack and E. Stefan, in Advances in NanoporousMaterials, Elsevier, 2010, vol. 1, pp. 1–96.

13 M. Hong, S. Li, J. L. Falconer and R. D. Noble, J. Membr. Sci.,2008, 307, 277–283.

14 J. Hedlund, J. Sterte, M. Anthonis, A. J. Bons, B. Carstensen,N. Corcoran, D. Cox, H. Deckman, W. de Gijnst, P. P. deMoor, F. Lai, J. McHenry, W. Mortier and J. Reinoso,Microporous Mesoporous Mater., 2002, 52, 179–189.

15 J. Hedlund, D. Korelskiy, L. Sandstrom and J. Lindmark, J.Membr. Sci., 2009, 345, 276–287.

16 J. Lindmark and J. Hedlund, J. Membr. Sci., 2010, 360, 284–291.

17 L. Sandstrom, J. Lindmark and J. Hedlund, J. Membr. Sci.,2010, 360, 265–275.

J. Mater. Chem. A, 2015, 3, 12500–12506 | 12505

Page 138: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Journal of Materials Chemistry A Paper

Publ

ishe

d on

14

May

201

5. D

ownl

oade

d on

29/

12/2

015

15:1

8:28

. View Article Online

18 P. Ye, E. Sjoberg and J. Hedlund, Microporous MesoporousMater., 2014, 192, 14–17.

19 H. Zhou, D. Korelskiy, E. Sjoberg and J. Hedlund,Microporous Mesoporous Mater., 2014, 192, 76–81.

20 D. Korelskiy, T. Leppajarvi, H. Zhou, M. Grahn, J. Tanskanenand J. Hedlund, J. Membr. Sci., 2013, 427, 381–389.

21 M. Grahn and J. Hedlund, J. Membr. Sci., 2014, 471, 328–337.22 J. Hedlund, F. Jareman, A. J. Bons and M. Anthonis, J.

Membr. Sci., 2003, 222, 163–179.23 D. Korelskiy, M. Grahn, J. Mouzon and J. Hedlund, J. Membr.

Sci., 2012, 417–418, 183–192.24 D. Korelskiy, P. Ye, H. Zhou, J. Mouzon and J. Hedlund,

Microporous Mesoporous Mater., 2014, 186, 194–200.25 M. Zhou, D. Korelskiy, P. Ye, M. Grahn and J. Hedlund,

Angew. Chem., Int. Ed., 2014, 53, 3492–3495.26 Z. Qin, L. Lakiss, L. Tosheva, J.-P. Gilson, A. Vicente,

C. Fernandez and V. Valtchev, Adv. Funct. Mater., 2014, 24,257–264.

12506 | J. Mater. Chem. A, 2015, 3, 12500–12506

27 K. Zhang, R. P. Lively, J. D. Noel, M. E. Dose, B. A. McCool,R. R. Chance andW. J. Koros, Langmuir, 2012, 28, 8664–8673.

28 J. A. Dunne, R. Mariwala, M. Rao, S. Sircar, R. J. Gorte andA. L. Myers, Langmuir, 1996, 12, 5888–5895.

29 J. A. Dunne, M. Rao, S. Sircar, R. J. Gorte and A. L. Myers,Langmuir, 1996, 12, 5896–5904.

30 R. Krishna and J. M. van Baten, J. Membr. Sci., 2010, 360,323–333.

31 M. Kanezashi, AIChE J., 2008, 54, 1478–1486.32 K. Kusakabe, T. Kuroda, K. Uchino, Y. Hasegawa and

S. Morooka, AIChE J., 1999, 45, 1220–1226.33 W. J. W. Bakker, F. Kapteijn, J. Poppe and J. A. Moulijn, J.

Membr. Sci., 1996, 117, 57–78.34 R. W. Baker and K. Lokhandwala, Ind. Eng. Chem. Res., 2008,

47, 2109–2121.35 Inopor®, http://www.inopor.com.36 D. Dortmundt and K. Doshi, Chem. Eng. World, 2003, 38, 55–

66.

This journal is © The Royal Society of Chemistry 2015

Page 139: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 140: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

PAPER VI

Ultra-Thin MFI Membranes for Light Olefins/Nitrogen Separation

Liang Yu, Mattias Grahn, Pengcheng Ye, and Jonas Hedlund

To be submitted to Journal of Membrane Science.

Page 141: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 142: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Ultra-thin MFI membranes for olefin/nitrogen separation

Liang Yu*, Mattias Grahn, Pengcheng Ye and Jonas Hedlund.

Chemical Technology, Luleå University of Technology, SE-971 87 Luleå, Sweden

*Corresponding author: Email address: [email protected]; Tel.: +46 920493002; fax: +46 920491199;

Abstract

The recovery of light hydrocarbons such as propylene and ethylene from vent streams in polymer

plants is desirable since it opens up for more efficient conversion of the monomers with high economic

value. Consequently, polymer membrane vapour-gas separation systems have been used for this

purpose for decades [1,2]. However, an alternative is zeolite membranes. In this work, ultra-thin MFI

zeolite membranes (0.5 µm) were used to separate propylene or ethylene from binary 20/80

olefin/nitrogen mixtures at different temperatures. The membranes were olefin selective with high

permeance at all investigated temperatures. At room temperature, the permeance of propylene was

22×10-7 mol m-2 s-1 Pa-1 and the separation factor was 43, which corresponds to a separation selectivity

of around 80. For a mixture of 20 mol.% ethylene in nitrogen, the maximum separation factor was 6

(corresponds to a separation selectivity of 8.4) at 277 K with an ethylene permeance of 57×10-7 mol m-2

s-1 Pa-1. The membrane selectivity was governed by more extensive adsorption of olefin, especially

propylene, as compared to nitrogen. Comparing with ethylene, propylene has higher heat of adsorption,

which probably caused the higher propylene/nitrogen selectivity compared to ethylene/nitrogen

selectivity. The permeance and the selectivity for propylene were much higher than for commercial

polymeric membranes. For ethylene, the permeance was much higher, and the selectivity was

comparable to commercial polymeric membranes. Modelling showed that the pressure drop over the

support limited the flux through the membranes especially at higher temperatures and in particular for

the ethylene/nitrogen system with high flux. Further, modelling indicated that the result obtained at

Page 143: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

high temperatures, where the flux was high, was also affected by concentration polarization. However,

for the propylene/nitrogen system at the optimum separation temperature, the pressure drop over the

support and the concentration polarisation were small. The results show that ultra-thin MFI zeolite

membranes are promising candidates for light olefins/nitrogen separation in polymer plants.

Keywords: MFI membrane; Zeolite membrane; Olefin/nitrogen separation; Ethylene; Propylene.

1. Introduction

Vapour-gas separation systems emerged in petrochemical and refinery industries due to the high

economic value of the recovered hydrocarbons. Recovery of light hydrocarbons from petrochemical

process streams, for example the purge gas from a polypropylene or polyethylene plant resin degasser,

is an interesting application. Typically, the purge stream from a polypropylene plant contains 20-30 %

C3+ hydrocarbons in nitrogen, and it would be of great value for the industry to recycle this part and use

the monomers for polymer production instead of burning it [1,3,4].

Membrane separations are particularly appealing for gas purification due to their low energy

consumption, good selectivity and low costs. From organic to inorganic, various materials of

membranes have been evaluated for separation of organic vapour/gas mixtures. Silicone rubber is by

far the most widely used membrane material for recovering hydrocarbons from vent streams, as

summarized by Baker et al. [1]. Silicon rubber membranes are well-suited for separation of organic

vapour/gas mixtures and the commercialization of this type of membrane continue to stimulate

membrane invention and industrial demand. However, these membranes are quite thick (typically ca.

50–100 µm), which partially explains why these membranes show rather low permeances. Furthermore,

the selectivity is only 8-12 for propylene/nitrogen and ethylene/nitrogen mixtures [1, 5]. Consequently,

large membrane area and many modules are needed, which increases the capital cost of the facility.

Besides, the membranes show rather low stability, freshly made thin composite polymer membranes

Page 144: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

will often lose 50% of the permeance within two weeks in this application. Paul et al. have shown that

the performance deterioration results from the reordering of the polymer chains in the membrane, thus

reducing the number and size of free volume elements that contribute to gas permeation [6]. Therefore,

a key challenge for polymer membranes is to increase the long term stability. Enhanced performance

may be obtained by combinations of polymers with inorganic particles in mixed-matrix membranes,

but difficulties, for example achieving good dispersion [7], are frequently encountered for this type of

membranes.

Over the past decades, the development of inorganic membranes, especially zeolite membranes, has

gained increased interest due to the potentially high stability, permeability and selectivity. MFI zeolite

is one of the most explored zeolites for membrane applications. MFI zeolite has a well-defined system

of pores with a size of ca. 0.55 nm in diameter. Furthermore, the Si/Al ratio of this zeolite can be varied

considerably making it possible to tailor the properties to a large extent, for instance hydrophobicity

and ion-exchange capacity, which could open up for using of the membranes in a broad range of liquid

and gas mixture separation applications [8,9]. Thus far, however, most reports on zeolite membranes

described relatively thick zeolite films; consequently permeances are typically quite low, comparable to

polymeric membranes. However, to be competitive, zeolite membranes must show much higher

permeance than polymeric membranes [10].

Our group has developed a technique for preparing ultra-thin (ca. 500 nm) MFI zeolite membranes on

open graded supports showing very high permeances e.g. a CO2 permeance of 93×10-7 mol m-2 s-1 Pa-1

for 50/50 CO2/H2 binary mixture with a separation factor of 16.2 [11]. In addition, very high fluxes

have been reported for p-xylene [12] and also for ethanol and n-butanol [13,14].

In the present work, we evaluated our ultra-thin MFI membranes for separation of propylene/nitrogen

and ethylene/nitrogen mixtures (20 mol% olefins in nitrogen) for the first time. The separations were

Page 145: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

performed at different temperatures to identify the optimum separation conditions. The effect of

concentration polarization and pressure drop over the support were estimated by modelling.

2. Experimental

2.1. Membrane preparation and characterisation

The membranes were prepared using a seeding method on masked supports as described in detail

previously [15]. Porous graded α-alumina discs (Fraunhofer IKTS, Germany) with a diameter of 25

mm comprised of a 30 µm thick top layer with a pore size of 100 nm and a 3 mm thick base layer with

a pore size of 3 µm were used as supports. The supports were masked with polymethylmethacrylate

(PMMA, Mw = 100 000 g mol-1, q = 1.8, CM 205, Polykemi AB, Sweden) and a Fisher Tropsh wax

(Sasolwax C105, Carbona AB, Sweden) to prevent invasion of zeolite into the support as described in

previous work [15] and were thereafter seeded with 50 nm silicalite-1 crystals as described in detail

elsewhere [12]. MFI zeolite films were grown in a synthesis mixture with a molar composition of

3TPAOH:25SiO2:1450H2O:100EtOH kept in an oil bath at 100 °C for 36 h under reflux. After

synthesis, the membranes were rinsed in a 0.1 M NH3 solution overnight and then calcined at 500 °C

for 6 h at a heating rate of 0.2 °C min-1 and a cooling rate of 0.3 °C min-1.

The membranes were characterized using n-hexane/helium adsorption-branch permporometry [16] in

order to estimate the amount of flow-through defects. The membranes were mounted in a stainless steel

cell sealed with graphite gaskets (Eriks, the Netherlands). A detailed description of the experimental

and data evaluation procedures are given in our previous work [17]. Scanning electron microscopy

(SEM, FEI Magellan 400 field emission XHR-SEM) was used to investigate the morphology of the

membranes.

2.2. Separation experiments

Page 146: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

The membranes were dried for 6 h at 300 °C under a flow of helium with a heating rate of 1 °C min-1

and cooling naturally before the separation experiments. A Wicke-Kallenbach cell was used and the

temperature inside the cell was monitored by a type K thermocouple and sub-ambient membrane

temperatures were achieved by placing the membrane cell in thermostated silicone oil. The desired feed

flow rate was achieved by using digital mass flow controllers (Bronkhorst HI-TEC). The feed pressure

was controlled by a backpressure regulator on the retentate line. The pressure on both sides of the

membrane was monitored by a digital pressure gauge. Permeation of gas mixtures was performed in a

continuous flow mode. No sweep gas was used and the permeate was kept at atmospheric pressure. The

permeate flow rate was measured using a drumtype gas meter (Ritter Apparatebau GmbH), and the

composition of feed and permeate streams was analyzed using an online GC (490 Micro GC, Agilent).

The olefin/nitrogen mixtures were fed to the membrane at a pressure of 10 bar and a flow rate of 10 l

min-1. Consequently, the pressure ratio, ϕ = PFeed/PPerm, was 10. Nitrogen single gas permeation

experiments were also performed in continuous flow mode without sweep gas. The nitrogen flow rate

was 8 l min-1 with a differential pressure of 7 bar over the membrane. The gases were purchased from

AGA gas company with the purity of nitrogen >99.999% and hydrocarbons >99.95%.

2.3. Modelling

The performance of high flux composite membranes may sometimes be affected by concentration

polarization on the feed side and pressure drop over the support. Concentration polarization leads to a

depletion of the preferentially permeating species at the membrane/fluid interface and pressure drop

over the support leads to an increased partial pressure of the permeating species on the permeate side of

the selective film. Consequently, the driving force may decrease, resulting in decreased flux and

sometimes also decreased selectivity. To elucidate if concentration polarization or pressure drop over

the support may have affected the outcome of the separation experiments, these effects were estimated

by modelling.

Page 147: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Concentration polarization can be described by the concentration polarization index nm/nb and it may be

estimated from the equation [18]:

[ ])/exp(1)/exp( cvb

pcv

b

m kJnn

kJnn

−+= (1)

where nm is the mole fraction at the zeolite film – gas interface and nb is the mole fraction in the gas

bulk, kc is the mass transfer coefficient, further Jv is the volumetric flux and np is the mole fraction in

the permeate. The mass transfer coefficient is obtainable from the Sherwood number, which in turn

may be obtained from correlations of Reynolds and Schmidt numbers. These correlations are dependent

on the geometry of system, flow conditions etc. To the best of our knowledge, there is only one

correlation available, capturing the complex geometry and special flow pattern inside a Wicke-

Kallenbach cell [19]:

30.133.058.0Re24.283.2

+=

CDwScSh (2)

where w is the compartment height (2.4 mm in our case) i.e. the thickness of the gas volume on the

feed side of the membrane and DC is the diameter (19 mm) of the membrane inside the gaskets, for

details see Perdana et al. [19]. Finally, the mass transfer coefficient was obtained from the Sherwood

number:

DdkSh hc= (3)

where dh is the hydraulic diameter of the cell and D is the diffusivity of propylene or ethylene in

nitrogen.

The pressure drop over the support was determined using the same procedure as has been reported by

several groups, including our group, before [13]. In short, transport through the 3 mm base layer with 3

Page 148: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

µm pores was assumed to occur via viscous flow, whereas transport through the thin (30 µm) top layer

with 100 nm pores was assumed to occur via a combination of viscous flow and Knudsen diffusion, see

e.g. Korelskiy et al. for details [13].

The Soave-Redlich-Kwong (SRK) equation of state was used for calculating fugacities on the high

pressure feed side whereas ideal gas behaviour was assumed on the permeate side. Gas phase

diffusivities and viscosities were adjusted to the desired temperature using the Di /23T∝ relationship

and Sutherland’s equation [20], respectively. The viscosities were calculated as molar fraction

weighted averages. The ideal adsorbed solution theory (IAST) was used for estimating adsorbed

amounts in the membrane during separation [21]. Saturation loadings, Langmuir affinity coefficients

and adsorption enthalpies were combined from literature data [22-24].

3. Results and discussion

3.1. Membrane characterization

Top view and cross-sectional SEM images of an as-synthesised MFI membrane are shown in Fig. 1.

The top view shows that the film is continuous film and is comprised of well intergrown zeolite

crystals with a maximum size of about 400 nm at the top surface of the membrane. No defects could be

observed by SEM in the membranes. The cross-sectional image shows that the film appears to be rather

even with a total thickness including invasion of around 420 nm. Moreover, the support is open and the

invasion of zeolite is limited to about 80 nm. The Si/Al ratio of the zeolite has been determined to 139

previously for another membrane prepared by the same method [25]. The aluminium originated from

the support being slightly leached during synthesis.

Table 1 shows n-hexane/helium adsorption-branch permporometry data for the membrane used in this

work. The initial helium permeance, i.e., at zero relative pressure of n-hexane, was high, amounting to

62×10-7 mol s-1 m-2 Pa-1. This high permeance indicates that the zeolite film is thin and that the pores

Page 149: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

are fully open and permeable. The helium permeance drops dramatically as hexane is introduced and

thereby blocking helium transport through the zeolite pores. A large drop in helium permeance as n-

hexane introducing is characteristic for high membrane quality. Consequently, the estimated defect

distribution showed that the total amount of defects was quite low, ca. 0.17% of the total membrane

area, and that more than 98 % (based on area) of the defects were smaller than 1 nm. Furthermore,

nearly no mesopore defects were detected by permporometry for this membrane. In summary, these

results show that the quality of this membrane is high, similar to that reported earlier by our group [17].

Fig. 1. Top view (left) and cross section (right) SEM images of an as-synthesised MFI membrane.

Table 1. Helium permeance measured by permporometry and relative areas of flow-through defects

estimated from permporometry data.

p/p0 Helium permeance

(10-7 mol s-1 m-2 Pa-1) Defect width

(nm) Defect interval

(nm) Relative

area of defects (%)

0 62.00 2.17×10-4 2.24 0.71 – – 3.67×10-4 1.51 0.73 0.71 - 0.73 0.08 1.11×10-3 0.64 0.80 0.73 - 0.80 0.07 1.12×10-2 0.27 1.04 0.80 - 1.04 0.02 2.60×10-2 0.24 1.19 1.04 - 1.19 0.001 1.35×10-1 0.21 1.78 1.19 - 1.78 0.0006 4.21×10-1 0.17 5.01 1.78 - 5.01 0.000

> 5.01 0.0011

Total 0.17

Page 150: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

3.2. Single gas permeation experiments

The single gas permeance of nitrogen as a function of the temperature is presented in Fig. 2. The

nitrogen permeance increased almost linearly with decreasing temperature. This is expected at least

within some temperature range as observed here. In this range, as the temperature decreases, the

adsorbed loading of nitrogen probably increases faster than the diffusivity of nitrogen decreases with

temperature since the heat of adsorption for nitrogen –20 kJ/mol is more negative (larger absolute value)

than the activation energy for diffusion –15 kJ/mol [26].

Fig. 2. Single gas permeance of N2 as a function of temperature with a differential pressure of 7 bar

over the membrane.

3.3. Separation of propylene/nitrogen mixtures

Propylene/nitrogen mixture separation experiments were performed at varying temperatures and a fixed

pressure ratio of 10 over the membrane. Fig. 3 illustrates the results for a 20/80 mixture of propylene in

nitrogen.

Page 151: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Fig. 3. Separation factors, separation selectivities, permeances and fluxes observed for a feed

comprised of a propylene/nitrogen (20/80) mixture as a function of temperature.

The observed propylene/nitrogen selectivity was high throughout the whole temperature range studied.

A maximum separation selectivity of 80 was observed at 296 K, with a corresponding separation factor

of 43. Similar trends, with an optimum separation temperature have been reported previously for

similar gas mixtures, for instance for the separation of n-C4H10 or C2H6 from binary mixtures with CH4

[27]. Starting from the highest temperature (326K), the selectivity increased with decreasing

temperature approaching the optimum, due to increased propylene/nitrogen adsorption selectivity.

According to IAST, the propylene/nitrogen adsorption selectivities on the feed side, after compensating

for concentration polarization, are 141, 234 and 238 at 326, 296 and 265 K, respectively, see Table 2.

The adsorption of propylene is favoured as compared to nitrogen as the temperature decreased towards

Page 152: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

the optimum because of the higher heat of adsorption of propylene (–39 kJ/mol) compared to that for

nitrogen (–20 kJ/mol) [24,26], resulting in higher propylene/nitrogen adsorption selectivity at lower

temperature in this range. At lower temperatures than optimum, saturation loading is reached and the

adsorption selectivity does not increase anymore. Fig. 3 shows that the nitrogen permeance was very

low and decreased even further as the temperature decreased in the presence of propylene. In strong

contrast, the single gas nitrogen permeance was very high and increased with decreasing temperature.

This clearly shows that propylene reduced nitrogen permeance, due to competitive adsorption. Fig. 3

shows that the highest observed propylene permeance was 39×10-7 mol m-2 s-1 Pa-1 at 326 K and that

the permeance decreased almost linearly with decreasing temperature in contrast to nitrogen single gas

permeance. In this range, as the temperature decreases, the adsorbed loading of propylene is not

increasing much. Using the IAST, the loading of propylene was estimated (after correcting for

concentration polarization) to 11.7, 12.2 and 12.4 kmol/(m3 accessible pore volume) at 326, 296 and

265 K, respectively. This is comparable to the saturation loading of ca. 12.5 kmol m-3. It shows that the

zeolite is close to saturation with propylene at the conditions of the experiments. The observed decrease

in propylene permeance with decreasing temperature should mainly be an effect of decreasing

diffusivity with decreasing temperature. For commercial polymeric membranes, propylene/nitrogen

separation selectivities of 8-10 at 263 K or even lower temperatures are usually reported [1,2], while

we observe much higher selectivities for MFI membrane in the present work, and besides, the optimum

propylene/nitrogen separation temperature is around room temperature. Consequently, an advantage

with the zeolite membrane for propylene/nitrogen separation is that much higher selectivity is reached

at much higher temperature (296 K), enabling operation of the process at room temperature and save

the cost of cooling. Normally, the permeance of light hydrocarbons for polymer membranes or carbon

molecular sieve membranes are lower than 1000 GPU (1 GPU=3.35×10-10 mol m-2 s-1 Pa-1) [2, 28-31],

i.e. lower than 3.35×10-7 mol m-2 s-1 Pa-1. At the optimum separation temperature, we observed a

propylene permeance of about 24×10-7 mol m-2 s-1 Pa-1, i.e. more than 7 times higher than for

Page 153: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

commercial silicon rubber polymeric membranes [2]. The very high selectivity and permeance of MFI

membranes observed here show that zeolite membranes are highly promising for this separation. As

usual, the flux follows the same trend as the permeance. For propylene, a maximum flux of 70 kg m-2 h-

1 was observed at 326 K, and it decreased to 15 kg m-2 h-1 at 265 K (Fig. 3). At the optimum separation

temperature (296 K), the nitrogen flux was only 2.2 kg m-2 h-1. All of this indicates that ultra-thin MFI

membranes are very promising for separation of propylene/nitrogen mixtures; in terms of permeance,

flux and selectivity.

Table 2. Concentration polarization index, relative pressure drop over support and adsorption selectivity on

feed side (estimated by IAST) for olefins/nitrogen separation.

Propylene/nitrogen Ethylene/nitrogen

Temperature (K) 265 296 326 258 284 315

Concentration polarization

index (Cm/Cb) 0.97 0.93 0.88 0.96 0.87 0.91

Relative pressure drop over

support (%) 3.2 8.4 17 8.9 22 25.7

Adsorption selectivity on feed

side (estimated by IAST) 238 234 141 69 47 34

The mathematical model showed that the pressure drop over the support was 17 % at the highest flux

(326 K), 8.4% at the highest selectivity (296 K) and only 3.2 % at the lowest flux (265 K), see Table 2.

In concert with our previous findings, the major part of the pressure drop over the support is over the

thin, 30 µm, top layer [13]. Reducing the thickness of the top layer is therefore a tempting option for

reducing the influence of the support. It should be noted that pressure drop over the support affects the

membrane performance negatively both in terms of flux and selectivity [13,14].

Modelling also shows that concentration polarization on the feed side only had a relatively small effect

on the outcome of the experiments. The concentration polarization indexes, Cm/Cb, ranged from 0.88 at

Page 154: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

the highest flux to 0.97 at the lowest flux, see Table 2. Increasing the feed flow rate, and thereby

increasing the superficial velocity, could improve the performance of the membrane at the higher

temperatures where concentration polarization was significant. The Reynolds numbers obtained in this

work was higher than the numbers obtained by Perdana et al [19] when deriving the correlation

presented in equation 2, thus the absolute values of the concentration polarization index should be

taken with some caution.

3.4. Separation of ethylene/nitrogen mixtures

Fig. 4 shows the results for separation of a 20/80 mixture of ethylene/nitrogen as a function of

temperature.

Page 155: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Fig. 4. Separation factors and selectivities, permeances and fluxes observed for feed comprised of an

ethylene/nitrogen (20/80) mixture as a function of temperature.

The membrane was also selective towards the olefin in this case. Similar trends were observed for this

system as for the propylene/nitrogen system when temperature changed, albeit the magnitudes differed.

However, the selectivity was significantly lower than observed for propylene/nitrogen as ethylene

adsorbs weaker than propylene and is therefore less effective in the competitive adsorption with

nitrogen [31,32]. The ethylene/nitrogen adsorption selectivities at the feed side estimated by IAST,

after compensating for concentration polarization, are 34, 47 and 69 at 315, 284 and 258 K,

respectively, see Table 2. As for propylene/nitrogen, a maximum in both separation factor and

selectivity was observed also for ethylene/nitrogen mixtures. Maximum separation factor and

selectivity was observed at 277 K and were 6 and 8.4, respectively, see Fig. 4. Polymeric membranes

usually have ethylene/nitrogen seletivities of 7-12 at 233 K [5]. Consequently, an advantage with the

zeolite membrane ethylene/nitrogen separation is that comparable selectivity is reached at much higher

temperature (277 K), enabling operation of the process at higher temperature and less need of costly

cooling. The permeance decreased with decreasing temperature for both components. In addition, the

permeance of both components were higher for the ethylene/nitrogen mixture than for the

propylene/nitrogen mixture, reflecting the higher diffusion rate of ethylene compared to propylene

coupled with a weaker adsorption and lower ability to block transport of nitrogen [33,23]. The highest

ethylene permeance was around 80×10-7 mol m-2 s-1 Pa-1 at 315 K, and it was decreased to 28×10-7 mol

m-2 s-1 Pa-1 at 258 K as shown in Fig. 4. At the temperature where the largest separation factor was

observed (277 K), the ethylene permeance was 57×10-7 mol m-2 s-1 Pa-1. This is more than 3 orders of

magnitude larger than the permeance around 10 GPU, i.e. 3×10-9 mol m-2 s-1 Pa-1, reported for PDMS

membranes [31] and more than 2 orders of magnitude larger the permeance 100 GPU, i.e. 3×10-8 mol

m-2 s-1 Pa-1, claimed for a polytrimethysilylpropyne membrane in a patent [5]. The flux of ethylene was

also very high around 79 kg m-2 h-1 at 277 K, see Fig. 4. As usual, the flux of ethylene and nitrogen

Page 156: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

followed the same trend as the permeance. As the flux of ethylene was higher than that of propylene,

the pressure drop over the support and concentration polarization may potentially be a greater issue.

The outcomes of the modelling are presented in Table 2. Concentration polarization were similar as for

propylene despite the higher flux of hydrocarbon observed in the ethylene/nitrogen experiments

compared to the propylene/nitrogen experiments. The higher gas phase diffusivity of ethylene

compared to propylene (by ca. 25%) is the reason for the similarities in concentration polarization

index observed for the two systems [32]. On the other hand, pressure drop over the support was more

severe for the ethylene/nitrogen mixtures compared to that for propylene/nitrogen mixtures because of

the higher flux observed for the former system. Similarly as for the propylene/nitrogen separation, the

performance of the membrane may be further improved by improving the permeability of the support

or by improving mixing on the feed side, e.g. by increasing the velocity of the gas or by introducing

turbulence generating elements in the cell, thus minimizing concentration polarization.

5. Conclusions

Ultra-thin (0.5 µm) MFI zeolite membranes were used to separate propylene or ethylene from binary

mixtures with nitrogen at different temperatures. The membrane showed a high performance for both

separations. The highest propylene/nitrogen separation factor and separation selectivity were 43 and 80,

respectively, coupled with a permeance of 22×10-7 mol m-2 s-1 Pa-1 at room temperature. The maximum

separation factor and separation selectivity for ethylene/nitrogen gas mixture were 6 and 8.4,

respectively at 277 K with an ethylene permeance of 57×10-7 mol m-2 s-1 Pa-1. The high permeance was

a result of the ultra-thin zeolite films and open supports. The olefin selectivity was ascribed to

olefin/nitrogen adsorption selectivity. The weaker adsorption of ethylene compared to propylene

resulted in lower ethylene/nitrogen selectivity compared to propylene/nitrogen selectivity, but on the

other hand the lighter ethylene showed a larger permeance than propylene. Pressure drop over the

support and concentration polarization had an adverse effect on the membrane performance, in

Page 157: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

particular when the flux was high. The results show that ultra-thin MFI zeolite membranes are very

promising candidates for recovery of light olefins from vent streams in the petrochemical industries.

Acknowledgements

The Swedish Energy Agency is gratefully acknowledged for financially supporting this work.

References

[1] R.W. Baker, B.T. Low, Gas Separation Membrane Materials: A Perspective, Macromolecules 47

(2014) 6999−7013.

[2] R.W. Baker, J.G. Wijmans, J.H. Kaschemekat, The design of membrane vapor-gas separation

systems, J. Membr. Sci. 151 (1998) 55-62.

[3] R.W. Baker, Future Directions of Membrane Gas Separation Technology, Ind. Eng. Chem. Res. 41

(2002) 1393-1411.

[4] R.W. Baker, K.A. Lokhandwala, M.L. Jacobs, D.E. Gottschlich, Recover feedstock and product

from reactor vent streams, Chem. Eng. Prog. 2000, 96, 51−57.

[5] R.W. Baker, K.A. Lokhandwala, I. Pinnau, Ethylene/nitrogen separation process, U.S. patent

5879431, 1999.

[6] Y. Huang, D.R. Paul, Physical aging of thin glassy polymer films monitored by gas permeability,

Polymer 45 (2004) 8377–8393.

[7] M.A. Aroona, A.F. Ismail, T. Matsuura, M.M. Montazer-Rahmati, Performance studies of mixed

matrix membranes for gas separation: A review, Sep. Purif. Technol. 75 (2010) 229–242.

[8] J. Caro, M. Noack, Zeolite membranes – Recent developments and progress, Micropor. Mesopor.

Mater. 115 (2008) 215–233.

Page 158: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

[9] N. Kosinov, J. Gascon, F. Kapteijn, E.J.M. Hensen, Recent developments in zeolite membranes for

gas separation, J. Membr. Sci. 499 (2016) 65–79.

[10] D. Korelskiy, P. Ye, S. Fouladvand, S. Karimi, E. Sjöberg, J. Hedlund, Efficient ceramic zeolite

membranes for CO2/H2 separation, J. Mater. Chem. A, 3 (2015) 12500–12506.

[11] L. Sandström, E. Sjöberg, J. Hedlund, Very high flux MFI membrane for CO2 separation, J.

Membr. Sci. 380 (2011) 232– 240.

[12] J. Hedlund, J. Sterte, M. Anthonis, A.J. Bons, B. Carstensen, N. Corcoran, D. Cox, H. Deckman,

W. deGijnst, P.P. deMoor, F. Lai, J. McHenry, W. Mortier, J. Reinoso, High-flux MFI membranes,

Micropor. Mesopor. Mater. 52 (2002) 179–189.

[13] D. Korelskiy, T. Leppäjärvi, H. Zhou, M. Grahn, J. Tanskanen, J. Hedlund, High flux MFI

membranes for pervaporation, J. Membr. Sci. 427 (2013) 381–389.

[14] L. Yu, D. Korelskiy, M. Grahn, J. Hedlund, Very high flux MFI membranes for alcohol recovery

via pervaporation at high temperature and pressure, Separat. Purif. Technol. 153 (2015) 138–145.

[15] J. Hedlund, F. Jareman, A.J. Bons, M. Anthonis, A masking technique for high quality MFI

membranes, J. Membr. Sci. 222 (2003) 163–179.

[16] J. Hedlund, D. Korelskiy, L. Sandström, J. Lindmark, Permporometry analysis of zeolite

membranes, J. Membr. Sci. 345 (2009) 276–287.

[17] D. Korelskiy, M. Grahn, J. Mouzon, J. Hedlund, Characterization of flow-through micropores in

MFI membranes by permporometry, J. Membr. Sci. 417–418 (2012) 183–192.

[18] R.W. Baker, J.G. Wijmans, A.L. Athayde, R. Daniels, J.H. Ly, M. Le, The effect of concentration

polarization on the separation of volatile organic compounds from water by pervaporation, J. Membr.

Sci. 137 (1997) 159-172.

Page 159: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

[19] I. Perdana, B.W. Tyoso, I.M. Bendiyasa, Rochmadi, S.K. Wirawan, D. Creaser, Effect of external

mass transport on permeation in a Wicke-Kallenbach cell, Chem. Eng. Res. Des. 87 (2009) 1438–1447.

[20] R.W. Johnson, The handbook of fluid dynamics, Boca Raton, FL, CRC Press, 1998.

[21] A.L. Myers, J.M. Prausnitz, Thermodynamics of mixed-gas adsorption. AIChE J. 11 (1965)

121−127.

[22] R. Krishna, The Maxwell–Stefan description of mixture diffusion in nanoporous crystalline

materials, Micropor. Mesopor. Mater. 185 (2014) 30–50

[23] F. Kapteijn, W.J.W. Bakker, G. Zheng, J. Poppe, J.A. Moulijn, Permeation and separation of light

hydrocarbons through a silicalite-1 membrane: Application of the generalized Maxwell-Stefan

equations, Chem. Eng. J. 57 (1995) 145–153.

[24] P. Pascual, P. Ungerer, B. Tavitian, A. Boutin, Development of a transferable guest-host force

field for adsorption of hydrocarbons in zeolites. II. Prediction of alkenes adsorption and alkane/alkene

selectivity in silicalite, J. Phys. Chem. B 108 (2004) 393–398.

[25] L. Sandström, J. Lindmark, J. Hedlund, Separation of methanol and ethanol from synthesis gas

using MFI membranes, J. Membr. Sci. 360 (2010) 265–275.

[26] F.V.S. Lopes, C.A. Grande, A.M. Ribeiro, J.M. Loureiro, O. Evaggelos, V. Nikolakis, A.E.

Rodrigues, Adsorption of H2, CO2, CH4, CO, N2 and H2O in activated carbon and zeolite for hydrogen

production, Sep. Sci. Technol. 44 (2009) 1045–1073.

[27] L.J.P. van den Broeke, W.J.W. Bakker, F. Kapteijn, J.A. Moulijn, Binary permeation through a

silicalite-1 membrane, AIChE J. 45 (1999) 976-985.

[28] I. Pinnau, Z. He, Pure- and mixed-gas permeation properties of polydimethylsiloxane for

hydrocarbon/methane and hydrocarbon/hydrogen separation, J. Membr. Sci. 244 (2004) 227–233.

Page 160: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

[29] H.H. Park, J. Won, S.G. Pyo, H. Park, Effect of palladium ion on facilitated olefin transport

through silver-polymer complex membranes, Macromol. Res. 19 (2011) 1077–1081.

[30] X. Ma, B.K. Lin, X. Wei, J. Kniep, Y.S. Lin, Gamma-alumina supported carbon molecular sieve

membrane for propylene/propane separation, Ind. Eng. Chem. Res. 52 (2013) 4297−4305.

[31] S.-H. Choi, J.-H. Kim, S.-B. Lee, Sorption and permeation behaviors of a series of olefins and

nitrogen through PDMS membranes, J. Membr. Sci. 299 (2007) 54–62.

[32] J. Peng, H.Y. Ban, X.T. Zhang, L.J. Song, Z.L. Sun, Binary adsorption equilibrium of propylene

and ethylene on silicalite-1: prediction and experiment, Chem. Phys. Lett. 401 (2005) 94–98.

[33] R.W. Elliott, H. Watts, Diffusion of some hydrocarbon in air: a regularity in the diffusion

coefficients of a homologous series, Canadian J. Chem. 50 (1972) 31-34.

Page 161: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 162: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

PAPER VII

A Study of CO2/CO Separation by Sub-micron b-Oriented MFI Membranes

Danil Korelskiy, Mattias Grahn, Pengcheng Ye, Ming Zhou, and Jonas Hedlund

Submitted to RSC Advances.

Page 163: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering
Page 164: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

RSC Advances

ARTICLE

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 1

Please do not adjust margins

Please do not adjust margins

Chemical Technology, Luleå University of Technology, SE-97187 Luleå, Sweden. E-mail: [email protected]

Received 00th January 20xx, Accepted 00th January 20xx

DOI: 10.1039/x0xx00000x

www.rsc.org/

A study of CO2/CO separation by sub-micron b-oriented MFI membranes D. Korelskiy,* M. Grahn, P. Ye, M. Zhou, and J. Hedlund

Separation of CO2 and CO is of great importance for many industrial applications. Today, CO2 is removed from CO mainly by adsorption or physical or chemical absorption systems that are energy-intensive and expensive. Membranes are listed among the most promising sustainable and energy-efficient alternatives for CO2 separation. Here, we study CO2/CO separation by novel sub-micron b-oriented MFI zeolite membranes in a temperature range of 258–303 K and at a feed pressure of 9 bar. At all experimental conditions, the membranes were CO2-selective and displayed high CO2 permeance ranging from 17,000 to 22,000 gpu. With decreasing temperature, the CO2/CO selectivity was increasing, reaching a maximum of 26 at 258 K. We also developed a mathematical model to describe the membrane process, and it indicated that the membrane separation performance was limited by the mass transfer resistance in the support and concentration polarisation on the feed side of the membrane.

Introduction

CO is used for manufacture of a wide range of valuable chemicals, such as Oxo-alcohols, phosgene (intermediate for polyurethane), methanol, acetic acid or various liquid hydrocarbon fuels via the Fischer-Tropsch process. CO itself is often produced by gasification (partial oxidation) of carbon-containing materials, e.g. coal, biomass, etc., or steam reforming of natural gas. In addition to CO, the produced gas mixture, referred to as synthesis gas, contains H2, CO2 and minor impurities, e.g. N2, H2O, H2S, etc. To utilise for the synthesis of chemicals, CO usually has to be purified, mainly from CO2. For instance, the gas used for Oxo-synthesis should not contain more than 0.5% of CO2

1. Separation of CO and CO2 is, however, challenging since the properties of the two gases are rather similar. Most of currently available CO2 separation methods rely on adsorption or absorption, e.g. pressure swing adsorption or amine scrubbing 2. Despite certain advantages, the sorption-based separation systems are cumbersome, rather energy-intensive and require large capital investments 3. Moreover, the absorbents used today are often corrosive and may result in releasing hazardous substances into the environment. Membranes are an appealing alternative to many common separation technologies, including adsorption, absorption and cryogenic distillation. High efficiency, sustainability and low energy demand are defining characteristics of membrane-based separation methods. Moreover, membrane separation processes are simple and continuous, requiring a minimum of

process equipment. For instance, we have recently demonstrated 4 that only one module housing 10 m2 of membrane area could replace an entire amine scrubbing system for separation of 300 ton CO2 per day from synthesis gas. It is worth pointing out that membranes have been listed among the most promising CO2 separation and capture technologies 3. The CO2-selective membranes currently available on the market are polymeric membranes. These membranes can be manufactured in large quantities at moderate costs. At the same time, the large quantities, i.e. large membrane areas, are a necessity for a given separation task because the membranes have low permeability, often coupled with a fairly poor selectivity. For example, commercial-scale PolarisTM membranes have a CO2 permeance of 100–300 gpu and a CO2/CO selectivity of 10–20 at a temperature ranging from 268 to 298 K 5. Even state-of-the-art CO2-selective polymeric membranes display a CO2 permeance of at most 1000 gpu 6. In addition, CO2 has a detrimental effect on the membrane stability, especially at high concentrations, reducing the membrane performance and lifespan 3. Among inorganic CO2-selective membranes, zeolite membranes are particularly attractive 7. These membranes have a well-defined pore system with pores ranging from 0.3 to 1.3 nm in size 8. Being porous, zeolite membranes can exhibit much higher fluxes than dense polymeric membranes 9. As a consequence, much lower membrane areas would suffice for a given separation task. In addition, the chemical and thermal stability of zeolite membranes is vastly superior to that of polymeric membranes 10. In spite of the great interest in separation of the main components of synthesis gas, i.e. CO, H2 and CO2, the number of research reports on zeolite membranes for this separation is small 7. In particular, CO2/CO membrane separation has

Page 165: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

ARTICLE Journal Name

2 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Please do not adjust margins

scarcely been studied. Our research group 11 has previously evaluated a sub-micron randomly oriented MFI membrane for the separation of an equimolar mixture of CO2, H2 and CO. The membrane was CO2-selective with a high CO2 permeance up to ca. 24,000 gpu. At a temperature of 277 K and a feed pressure of 9 bar, a CO2/CO membrane selectivity of ca. 10 was reported. Recently, we 12 have developed a sub-micron b-oriented MFI membrane employing a fluoride route (neutral pH) for the synthesis. The developed membranes displayed high CO2/H2 and CO2/CO selectivities up to ca. 110 and 20, respectively. Here, we extend our previous work on CO2/CO separation using sub-micron b-oriented MFI membranes by developing a model to elucidate the experimental findings and describe the effect of the support and concentration polarisation on the membrane separation performance. In addition, the morphology of the membranes, especially the grain boundaries and grain size, was carefully studied by HR-SEM. Permporometry data were used to estimate the width of the open grain boundaries and the grain size, and the findings were compared with the HR-SEM data.

Experimental Membrane synthesis

Sub-micron (ca. 0.5 μm) b-oriented MFI zeolite membranes were prepared as described in detail in our previous work 12. The membranes were supported on a commercial graded α-alumina disc (Fraunhofer IKTS, Germany) with a diameter of 25 mm. A brief summary of the membrane preparation procedure is given below. In the first step, the supports were seeded with plate-shaped MFI crystals with a size of 500×450×200 nm. The seed crystals were synthesised from a mixture containing tetraethyl orthosilicate (TEOS), tetrapropylammonium hydroxide (TPAOH) and water with a respective molar ratio of 1:0.2:100. The mixture was kept at 403 K for 9 h. In the second step, zeolite films were grown on the seeded supports by hydrothermal synthesis at 373 K for 48 h using a synthesis mixture with a molar composition of SiO2 : 0.12TPAOH : 60H2O : 0.12HF. After synthesis, the membranes were rinsed with a 0.2 M ammonia solution and dried at 323 K overnight. The following day, the membranes were calcined at 773 K for 6 h at a heating rate of 0.2 K min-1 and a cooling rate of 0.3 K min-

1. Membrane characterisation

The morphology of the membranes was characterised by High Resolution Scanning Electron Microscopy (HR-SEM) using a Magellan 400 SEM (the FEI Company, the Netherlands). Cross-sections of the membranes were obtained by fracturing the membranes with a pair of cutting pliers. No conductive coating was applied to the samples. The quality of the membranes in terms of defects was evaluated by n-hexane/helium permporometry 13 as described in detail in our earlier work 14 and in brief here. The membranes were mounted in a stainless steel cell using graphite gaskets (Eriks, the Netherlands) for sealing. In order

to remove any adsorbed compounds, the membranes were heated to 573 K at a heating rate of 1 K min-1 and kept at this temperature for 6 h in a flow of pure helium (99.999%, AGA). The permporometry experiment was performed at a temperature of 323 K, a total feed pressure of 2 bar and a total permeate pressure of 1 bar. The relative pressure of n-hexane (99%, Alfa Aesar) in the feed was increased in a step-wise manner from 0 to ca. 0.4. At each relative pressure, the system was allowed to equilibrate. In order to remove n-hexane from the permeate stream, a condenser kept at 233 K followed by a column packed with activated carbon were employed. A soap bubble flow meter was used to measure the n-hexane-free permeate volumetric flow rate. For each n-hexane relative pressure, a defect width was calculated by either the Horvàth–Kavazoe equation (micropore-range defects) or the Kelvin equation (mesopore-range defects). For each defect interval, the average defect width was then calculated. Based on the average defect width, the average helium diffusivity in each defect interval was estimated using the gas-translational model. Knowing the diffusivity, the helium molar flux was further calculated from Fick’s law. The area of defects was estimated as the ratio between helium molar flow and flux through the defects in that particular interval. In the final step, the relative area of defects was calculated by dividing the area of defects with the total membrane area. More details about the evaluation procedure of the permporometry data can be found in our previous work 14. CO2/CO separation experiments

Mixed-gas separation experiments were carried out using a 50/50 (v/v) mixture of CO2 (99.999%, AGA) and CO (99.97%) at a total feed pressure of 9 bar and a total permeate pressure of 1 bar. The membrane was kept in the cell used for the permporometry test. The experiments were carried out at a temperature ranging from 258 to 303 K. Prior to the experiments, the membrane was flushed at 573 K for 6 h using a stream of pure helium (99.999%, AGA) to remove any adsorbed species. The volumetric flow rate of the permeate was measured with a drum-type gasmeter (TG Series, Ritter Apparatebau GmbH), and the composition of the permeate was analysed with a mass spectrometer (GAM 400, InProcess Instruments). The flux of component i, Ji (mol s-1 m-2), was estimated from the measured molar flow rate of the corresponding component through the membrane, Fi (mol s-1) as

AFJ ii = , (1)

where A is the membrane area (m2). The permeance of component i, iΠ (mol s-1 m-2 Pa-1), was calculated from the flux of the corresponding component through the membrane as

iii PJ Δ=Π , (2)

Page 166: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Journal Name ARTICLE

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 3

Please do not adjust margins

Please do not adjust margins

where ΔPi (Pa) is the partial pressure difference of component i across the membrane. It should be noted that for gas permeation and separation, permeance is normally reported in gpu (gas permeation units), where 1 gpu is equal to 3.35×10-10 mol s-1 m-2 Pa-1. The CO2/CO membrane selectivity, CO/CO2

α , was estimated as

COCOCO/CO /22

ΠΠ=α . (3)

Modelling

For composite membranes, as in this work, high flux may cause both concentration polarisation on the feed side and a pressure drop over the support. These effects may result in a decrease in the performance of the membrane, in terms of both flux and selectivity 15, 16. To assess to what extent these effects may have influenced the separation performance of the membrane in the present work, a mathematical model was used. The model has been described in detail elsewhere15, 17, 18 and will only be described briefly below. The effect of concentration polarisation may be evaluated by determining the concentration polarisation index (CPI) as

[ ])/exp(1)/exp(CPI cvb

pcv kJ

nn

kJ −+= , (4)

where nm is the mole fraction at the zeolite film–gas interface and nb is the mole fraction in the gas bulk, kc is the mass transfer coefficient, Jv is the volumetric flux and np is the mole fraction in the permeate. The mass transfer coefficient may be obtained from mass transfer correlations relating the Sherwood (Sh) number to the Reynolds (Re) and Schmidt (Sc) numbers. Perdana et al. have reported a correlation for mass transfer in a Wicke-Kallenbach cell, which is used in the present work 19.

30.133.058.0 ScRe24.283.2Sh

+=

CDw

, (5)

where w is the compartment height, i.e. the thickness of the gas film on the feed side of the membrane and DC is the diameter of the membrane inside the gaskets, for more details see Perdana et al. 19. The mass transfer coefficient was finally retrieved from the Sherwood number:

Ddk

Sh hc=, (6)

where dh is the hydraulic diameter of the cell and D is the gas phase diffusivity of CO2. The pressure drop across each layer of the support was determined using procedures described in detail elsewhere15,

20-22. The transport through the 3 mm thick base layer with 3 µm pores was assumed to occur via viscous flow, whereas transport through the 30 µm thick top layer was assumed to occur via a combination of viscous flow and Knudsen diffusion. Effective permeabilities and Knudsen transport coefficients for

each of the support layer have been determined in our previous work 15. Gas phase diffusivities and viscosities were adjusted to the desired temperature using the Di /23T∝ relationship and Sutherland’s equation 23, respectively. The viscosities of the mixtures were calculated as molar fraction weighted averages. The ideal adsorbed solution theory (IAST) 24 was used to estimate adsorbed loadings on the feed side of the membrane during the separation experiment. Saturation loadings, Langmuir affinity coefficients and adsorption enthalpies were taken from the literature25-28.

Results and discussion Membrane characterisation

Figure 1 shows cross-sectional and top-view SEM images of an as-synthesised membrane. The MFI film was even and continuous, with a very low thickness of ca. 0.5 µm. The crystals composing the film were shown to be well-intergrown with a characteristic coffin shape. The length of the surface crystals along the a-axis was ca. 0.5 μm, whereas the length of the crystals along the c-axis was ca. 2 μm. Defects in the form of cracks and pinholes, which are normally larger than 5 nm, were not detected by SEM. The absence of such defects indicates high quality of the membrane. The b-orientation of as-synthesised membranes was corroborated by X-Ray Diffraction (XRD), and the data were reported in our previous work12.

Page 167: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

ARTICLE Journal Name

4 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Please do not adjust margins

Fig. 1. Cross-sectional (a) and top-view (b) SEM images of an as-synthesised membrane. Table 1 shows the helium permeance through the synthesised membrane measured at each relative pressure of n-hexane in the permporometry experiment. The table also shows the relative areas of defects estimated for each defect interval. In the permporometry experiment, the initial measuring point is recorded at a relative pressure of n-hexane of 0. The helium permeance at this point denotes the combined permeance through zeolite pores and defects. For this membrane, the initial helium permeance was high amounting to ca. 20,000 gpu, indicating that the zeolite pores are open and permeable. With increasing relative pressure of n-hexane, first MFI pores and then defects were blocked by n-hexane causing the helium permeance to decrease. The total relative area of defects in the membrane was very small, scarcely above 0.1% of the total membrane area, indicating a very high quality of the membrane. The few detected defects were essentially micropore defects, accounting for ca. 99.6% of all defects. Virtually no mesopore defects were detected by permporometry, which is consistent with the SEM observations. In our previous work, we showed that the micropore defects detected by permporometry in our randomly oriented high-flux MFI membranes were open grain boundaries. In order to identify whether the detected micropore defects in the b-oriented MFI membranes prepared in this work were as well open grain boundaries, we used a rough geometrical model of the membrane microstructure. In the model, the grains were assumed to have a cuboid habit with a quadratic cross-section with a width of x, see Figure 2. The grain width was assumed to be constant throughout the film thickness. The micropore defects with an average width of d (where d = (0.7 + 2)/2 = 1.35 nm), were assumed to be evenly distributed between the grains, as shown in Figure 3. Therefore, the relative area of micropore defects a can calculated as

( )( )

.2dxxdxd

a+

++= (7)

Since d should be much smaller than x, Equation 7 can be simplified to

.2xd

a = (8)

The relative area of micropore defects a can be estimated from permporometry data obtained experimentally. Knowing the relative area, the grain size x can easily be estimated using Equation 8. In other words, the grain size x can also be estimated from permporometry data.

Fig. 2. A geometrical model of the membrane microstructure. For this particular membrane, the relative area of micropore defects a was ca. 0.13% yielding a grain size x of ca. 2 µm. In spite of the great simplicity of the model, the grain size estimated from the permporometry data was in excellent agreement with the grain sizes observed by SEM (see Figure 1b). The latter indicates that the micropore defects detected by permporometry in the prepared membrane should mainly be open grain boundaries. Table 1. Permporometry data for synthesised membrane. P/P0 He permeance

(1000 gpu) Defect interval (nm)

Relative area of defectsa (%)

0 20 – 2.3×10-4 0.52 0.71 – 0.74 0.059 3.7×10-4 0.35 0.74 – 0.81 0.043 1.1×10-3 0.20 0.81 – 1.05 0.029 1.1×10-2 0.035 1.05 – 1.80 0.0011 1.2×10-1 0.022 1.80 – 4.41 0.000092

Page 168: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

Journal Name ARTICLE

This journal is © The Royal Society of Chemistry 20xx J. Name., 2013, 00, 1-3 | 5

Please do not adjust margins

Please do not adjust margins

3.7×10-1 0.019 > 4.41 0.00046 Total: 0.13 a Area of defects per total membrane area CO2/CO separation experiments

Figure 3 shows mixed-gas permeances of CO2 and CO as a function of temperature. High mixed-gas permeances up to 22,000 gpu were observed. In the entire temperature range, the permeance of CO2 through the membrane was higher than that of CO. With decreasing temperature, the permeances of both CO2 and CO were decreasing, most likely due to reduced diffusivities at lower temperatures. It should be noted, however, that even at the lowest studied temperature of 258 K the permeance of CO2 was as high as ca. 17,000 gpu.

Fig. 3. CO2 and CO permeances measured in the temperature range of 258–303 K at a feed pressure of 9 bar and a permeate pressure of 1 bar.

Fig. 4. CO2/CO membrane selectivity determined in the temperature range of 258–303 K at a feed pressure of 9 bar and a permeate pressure of 1 bar. Figure 4 shows CO2/CO mixed-gas membrane selectivity as a function of temperature. At the highest studied temperature of 303 K, the CO2/CO selectivity was ca. 6, and it was increasing with decreasing temperature, reaching a maximum of 26 at the lowest studied temperature of 258 K. The high

selectivity to CO2 most likely emanates from the greater adsorption affinity of the membrane for CO2 compared to CO. Table 2 reports the CO2 and CO fluxes, and the concentration of CO2 and CO in the permeate stream. The observed CO2 flux was very high, i.e. 320–440 kg m-2 h-1, in the entire temperature range. The high CO2 flux is most likely a result of the very low thickness of zeolite film, strong CO2 adsorption and high CO2 diffusivity in the zeolite pores, and a relatively high CO2 partial pressure difference of 3.5 bar across the membrane. Table 2. CO2 and CO fluxes and permeate concentration observed in the present work.

T (K) Flux (kg h-1 m-2) Permeate concentration (mol. %)

CO2 CO CO2 CO 303 438 51 84.75 15.63 298 420 41 86.63 13.33 293 434 38 87.90 12.05 288 436 32 89.12 10.40 283 370 26 90.15 9.80 278 367 21 91.70 8.30 273 366 18 92.80 7.15 268 347 15 93.75 6.22 263 321 12 94.60 5.40 258 319 10 95.35 4.60

We have previously shown15, 16 that the pressure drop over the support and concentration polarisation on the feed side may adversely affect the separation performance of high flux membranes. The effect of these phenomena was evaluated for three different temperatures, viz. 303, 278 and 258 K. The selected points represent high, medium and low flux, respectively. Table 3 shows the results of the evaluation. The relative pressure drop over the support, ∆P, (i.e. pressure drop over support/pressure drop over whole composite membrane) was estimated to be 33%, 27% and 21% at 303, 278 and 258 K, respectively. These values are in agreement with our previous findings for high-flux MFI zeolite membranes 15, 16, indicating that the support adversely affected the performance of the membrane. Most of the resistance was found to be in the thin top layer of the support having 100 nm pores. Reducing the thickness of this top layer from 30 to, e.g. 5–10 µm would be a feasible way of reducing the negative influence of the support17. The concentration polarisation index was determined to be 0.83, 0.87 and 0.9 at 303, 278 and 258 K, indicating that concentration polarisation reduced the performance of the membrane, especially at the higher temperatures, where the high flux was higher. Table 3. Relative pressure drop over the support, concentration polarisation index (CPI), permselectivity for the zeolite film and adsorption selectivity according to IAST estimated for three different temperatures.

T (K) ∆P (%) CPI αperm, film αads

303 33 0.83 15 18 278 27 0.87 24 30

Page 169: Zeolite Membrane Separation - DiVA portalltu.diva-portal.org/smash/get/diva2:990452/FULLTEXT02.pdfDO CTO RL TH ESIS Department of Civil, Environmental and Natural Resources Engineering

ARTICLE Journal Name

6 | J. Name., 2012, 00, 1-3 This journal is © The Royal Society of Chemistry 20xx

Please do not adjust margins

Please do not adjust margins

258 21 0.9 41 46 The permselectivities for the zeolite film after correcting for concentration polarisation and pressure drop over the support, αperm, film, were ca. 15, 24 and 41 at 303, 278 and 258 K, respectively. These values differ from the overall selectivities, ca. 6, 14 and 26, measured at the same temperatures (see Figure 3). It is therefore evident that the membrane selectivity was significantly reduced by concentration polarisation and pressure drop over the support. To better understand the experimental findings, we estimated the adsorption selectivity on the feed side of the membrane using the IAST, after removing the effect of concentration polarisation, see Table 3. The CO2/CO adsorption selectivity was increasing with decreasing temperature, as expected because of the more negative heat of adsorption of CO2 compared to that of CO 26. Moreover, the estimated adsorption selectivities were very similar to the permselectivities. The later fact indicates that the membrane selectivity to CO2 should mainly be ascribed to the stronger adsorption of CO2 in the zeolite film.

Conclusions Ultra-thin b-oriented oriented high-flux MFI zeolite membranes were prepared and evaluated for CO2/CO separation at a temperature ranging from 258 to 303 K, a feed pressure of 9 bar and a permeate pressure of 1 bar. The membranes were CO2-selective in the entire temperature range. The highest CO2/CO selectivity of 26 was observed at the lowest studied temperature of 258 K. The selectivity of the membrane to CO2 was attributed to the stronger CO2 adsorption, which was supported by a mathematical model based on IAST. The model also indicated that the separation performance of the membrane was reduced by the pressure drop across the support and concentration polarisation on the feed side. Without these effects, a CO2/CO selectivity of 41 would be achieved at the lowest studied temperature of 258 K.

Acknowledgements The Swedish Research Council Formas (Grant no. 213-2013-1684), the Swedish Energy Agency (Project no. 38336-1) and Bio4Energy are gratefully acknowledged for financially supporting this work.

Notes and references 1. C. Higman and M. v. d. Burgt, Gasification, Gulf

Professional Pub./Elsevier Science, Amsterdam ; Boston, 2nd edn., 2008.

2. X. Y. Chen, H. Vinh-Thang, A. A. Ramirez, D. Rodrigue and S. Kaliaguine, RSC Advances, 2015, 5, 24399-24448.

3. S. D. Kenarsari, D. Yang, G. Jiang, S. Zhang, J. Wang, A. G. Russell, Q. Wei and M. Fan, RSC Advances, 2013, 3, 22739-22773.

4. D. Korelskiy, P. Ye, S. Fouladvand, S. Karimi, E. Sjöberg and J. Hedlund, J. Mater. Chem. A, 2015, 3, 12500-12506.

5. H. Lin, Z. He, Z. Sun, J. Vu, A. Ng, M. Mohammed, J. Kniep, T. C. Merkel, T. Wu and R. C. Lambrecht, J. Membr. Sci., 2014, 457, 149-161.

6. R. W. Baker and B. T. Low, Macromolecules, 2014, 47, 6999-7013.

7. M. Pera-Titus, Chem. Rev., 2013, 114, 1413-1492. 8. J. Caro and M. Noack, Micropor. Mesopor. Mater., 2008,

115, 215-233. 9. E. Sjöberg, L. Sandström, O. G. W. Öhrman and J. Hedlund,

J. Membr. Sci., 2013, 443, 131-137. 10. J. Caro, M. Noack and E. Stefan, in Advances in Nanoporous

Materials, Elsevier, 2010, vol. 1, pp. 1-96. 11. L. Sandström, E. Sjöberg and J. Hedlund, J. Membr. Sci.,

2011, 380, 232-240. 12. M. Zhou, D. Korelskiy, P. Ye, M. Grahn and J. Hedlund,

Angew. Chem., Int. Ed., 2014, 53, 3492-3495. 13. J. Hedlund, D. Korelskiy, L. Sandström and J. Lindmark, J.

Membr. Sci., 2009, 345, 276-287. 14. D. Korelskiy, M. Grahn, J. Mouzon and J. Hedlund, J.

Membr. Sci., 2012, 417–418, 183-192. 15. D. Korelskiy, T. Leppäjärvi, H. Zhou, M. Grahn, J. Tanskanen

and J. Hedlund, J. Membr. Sci., 2013, 427, 381-389. 16. L. Yu, D. Korelskiy, M. Grahn and J. Hedlund, Sep. Purif.

Technol., 2015, 153, 138-145. 17. M. Grahn and J. Hedlund, J. Membr. Sci., 2014, 471, 328-

337. 18. P. Ye, D. Korelskiy, M. Grahn and J. Hedlund, Submitted to

AIChE Journal. 19. I. Perdana, B. W. Tyoso, I. M. Bendiyasa, Rochmadi, S. K.

Wirawan and D. Creaser, Chem. Eng. Res. Des., 2009, 87, 1438-1447.

20. F. T. de Bruijn, L. Sun, Z. Olujic, P. J. Jansens and F. Kapteijn, J. Membr. Sci., 2003, 223, 141-156.

21. F. Jareman, J. Hedlund, D. Creaser and J. Sterte, J. Membr. Sci., 2004, 236, 81-89.

22. S. Thomas, R. Schäfer, J. Caro and A. Seidel-Morgenstern, Catal. Today, 2001, 67, 205-216.

23. R. W. Johnson, The handbook of fluid dynamics, CRC Press, Boca Raton, Fla., 1998.

24. A. L. Myers and J. M. Prausnitz, AIChE J., 1965, 11, 121-127. 25. L. Ohlin, P. Bazin, F. Thibault-Starzyk, J. Hedlund and M.

Grahn, The Journal of Physical Chemistry C, 2013, 117, 16972-16982.

26. S. K. Wirawan, M. Petersson and D. Creaser, ASEAN Journal of Chemical Engineering, 2008, 8.

27. R. Krishna and J. M. van Baten, J. Membr. Sci., 2011, 383, 289-300.

28. R. Krishna, Micropor. Mesopor. Mater., 2014, 185, 30-50. 29. D. Korelskiy, P. Ye, H. Zhou, J. Mouzon and J. Hedlund,

Micropor. Mesopor. Mater., 2014, 186, 194-200.