You are on page 1of 268

ABSTRACT

ZHAO, JUNJIE. Metal-Organic Framework Thin Films on Fibers Enabled by Atomic Layer
Deposition. (Under the direction of Dr. Gregory N. Parsons).

Metal-organic frameworks (MOFs) are crystalline porous materials that consist of metal

clusters and organic bridging linkers. With ultra-high surface area and large porosity, MOFs

are promising for many applications such as gas adsorption, separations and catalysis.

However, the powder form of MOFs obtained from common solvothermal synthesis requires

further handling and manipulation, and is difficult to be integrated into functional materials

and devices for applications such as gas filtration, membrane separations and chemical

sensing. Surface-immobilized MOF thin films, especially on natural and synthetic fiber

scaffolds, could simplify the deployment, enable new device configurations, and expand the

applications of MOFs.

This research is motivated to develop new methods to synthesize MOF thin films on

polymer fibers and other functional substrate materials, and to understand the growth

mechanism of MOF thin films on functionalized surfaces. Metal oxide thin films including

Al2O3, ZnO and TiO2 deposited via atomic layer deposition (ALD) were used as nucleation

layers or templating materials for MOF thin films. Both macroscopic uniformity and

microscopic conformality of the MOF coatings are improved compared to previous methods.

The functional properties of MOFs are also fully maintained, as the MOF-coated fibers show

excellent performance for capturing toxic industrial chemicals and catalytically degrading

chemical warfare agents. The methods developed in this work can also be generally applied

to a wide range of MOFs and fibrous substrates.


Through an extensive investigation for the reactions of MOF precursors on ALD metal

oxide surfaces, we also discovered an ultra-fast reaction path for MOF formation at room

temperature. Hydroxy double salts (HDSs) converted from ZnO were found as key

intermediate species to enable the rapid formation of MOFs via anion exchange. The

space-time-yield of this process is over 3×104 kg∙m-3∙d-1, a significant enhancement for

advancing the industial applications and commertialization of MOFs. New MOF-based

composites can also be synthesized using this facile convertion from HDSs to MOFs.

The synthesis routes we have developed represent breakthroughs for MOF thin films and

scalable MOF production. Several applications of MOF-functionalized fibrous materials in

gas adsorption and heterogeneous catalysis will be presented in this dissertation. The

synthesis method and growth mechanism of milimeter-scale HKUST-1 [Cu3(BTC)2] single

crystals and the ZnO 3D hierarchical nanostructures for photocatalysis will also be discussed.
© Copyright 2016 by Junjie Zhao

All Rights Reserved


Metal-Organic Framework Thin Films on Fibers Enabled by Atomic Layer Deposition

by
Junjie Zhao

A dissertation submitted to the Graduate Faculty of


North Carolina State University
in partial fulfillment of the
requirements for the Degree of
Doctor of Philosophy

Chemical Engineering

Raleigh, North Carolina

2016

APPROVED BY:

_____________________________ ___________________________
Dr. Gregory N. Parsons Dr. Saad A. Khan
Chair of Advisory Committee

_____________________________ ___________________________
Dr. Michael D. Dickey Dr. Chih-Hao Chang
DEDICATION

To my parents, Jiannong Zhao and Xueying Wu,

and my fiancée Yizhou Chen

ii
BIOGRAPHY

Junjie Zhao was born in Suzhou, Jiangsu Province, China in 1988. He obtained a

Bachelor of Engineering degree in Chemical Engineering & Techonology as well as a

Bachelor of Arts degree in English from Zhejiang Univeristy, Hangzhou, China in June 2011.

In August 2011, Junjie was enrolled in the Ph.D. program in the Department of Chemical &

Biomolecular Engineering at North Carolina State Univeristy. His research focus is on

metal-organic framework (MOF) thin films, MOF-functionalized polymer fibers for

adsorption and catalysis, scalable rapid room-temperature synthesis for MOFs and the

application of atomic layer deposition for nanostructures. During his Ph.D. studies, Junjie has

achieved the American Vacuum Society Thin Film Division Graduate Student Award, the AIF

Annual Best Paper Award at NCSU and the 2nd Place Oral Presentation in 2016 Schoenborn

Graduate Research Symposium. He also visited the National Institute of Standards and

Technology as a guest researcher for research collabration in 2016. He will continue his

research career as a postdoctoral scholar at Massachusetts Institute of Technology in

Semptember 2016.

iii
ACKNOWLEDGEMENT

I am very grateful to my advisor, Prof. Gregory N. Parsons, for his tremendous support

and guidance for my research and helpful advice for my career development. I also

appreciate the help and inspiring comments from the members of my committee, Prof. Saad

A. Khan, Prof. Michael D. Dickey and Prof. Chih-Hao Chang.

I would like to thank former members in Parsons research group, Dr. Mark Losego, Dr.

Bo Gong, Dr. Do Han Kim, Dr. Kyoungmi Lee, Dr. Christina McClure, Dr. Will Sweet, Dr.

Berc Kalanyan, Dr. Sarah Atanasov, Dr. Philip Williams, Dr. Moataz Mousa for their

mentoring and support. Dr. Mark Losego helped me a lot in both research and technical

writing. I am thankful to current group members, Dr. Chris Oldham, Dr. Alex Brozena, Mike

Maniti, Paul Lemaire, Zack Mundy, Erinn Needham, James Daubert, Eric Stevents, Jennifer

Ovental, Wenyi Xie, Mariah Ritz, Dennis Lee and Heather Barton, for their helpful

discussion and research collaboration. I also want to thank Trent Blevins, Fahim Sidi,

William Nunn, Anqi Qu and Ian Woodward, for their help in the past projects.

I also would like to acknowledge the staff at the analytical instrumentation facility (AIF)

at NCSU, Roberto Garcia, Ching-Chang Chung, Chuck Mooney, Yang Liu, Xiahan Sang and

Chuanzhen Zhou for their training and help with some characterization. I appreciate the

research collaboration with Gregory Peterson (ECBC), Dr. Howard Walls (RTI), Dr. Sarah

Shepherd (RTI), Robert Yaga (RTI), Dr. Pamela Chu (NIST), Dr. Brent Sperling (NIST), Dr.

M. Douglas LeVan (Vanderbilt), Trent Tovar (Vanderbilt), Dr. Michael Dickey, Yiliang Lin,

Dr. Chih-hao Chang, Abhijeet Bagal and Xu Zhang.

Last but not least, I am very grateful to my family. Without your support this would not

iv
be possible.

v
TABLE OF CONTENTS

LIST OF TABLES ................................................................................................................................ ix


LIST OF FIGURES ................................................................................................................................ x
CHAPTER 1. Introduction – Synthesis, Characterization and Applications of Metal-Organic
Framework Functionalized Polymer Fibers ........................................................................................... 1
1.1 Introduction .................................................................................................................................. 2
1.2 Synthetic Methods ........................................................................................................................ 5
1.2.1 Solvothermal Synthesis ......................................................................................................... 5
1.2.1.1 Direct Solvothermal Growth .............................................................................................. 5
1.2.1.2 Solvothermal Synthesis on Chemically Functionalized Surface ........................................ 6
1.2.1.3 Solvothermal Synthesis on MOF-Seeded Surface.............................................................. 8
1.2.1.4 Solvothermal Synthesis on Atomic Layer Deposited Nucleation Layers .......................... 9
1.2.2 Layer-by-Layer Synthesis ................................................................................................... 14
1.2.2.1 Direct Layer-by-Layer Synthesis ..................................................................................... 14
1.2.2.2 Layer-by-Layer Synthesis on Chemically Functionalized Surface .................................. 16
1.2.2.3 Layer-by-Layer Synthesis on Atomic Layer Deposited Nucleation Layers ..................... 19
1.2.3 MOF Conversion from Thin Film Templates ..................................................................... 21
1.2.3.1 Conversion from ALD Metal Oxide Thin Films .............................................................. 21
1.2.3.2 Room-Temperature Conversion from Hydroxy Double Salts .......................................... 23
1.2.4 Electrospinning .................................................................................................................... 25
1.2.5 Co-Extrusion Spinning ........................................................................................................ 33
1.2.6 Inkjet Printing ...................................................................................................................... 35
1.2.7 Spay Drying Method ........................................................................................................... 36
1.3 Characterization Methods ........................................................................................................... 39
1.3.1 Attenuated Total Reflection (ATR) Infrared spectroscopy ................................................. 39
1.3.2 UV-Vis Spectroscopy .......................................................................................................... 41
1.3.3 Analysis of MOF Mass Fraction ......................................................................................... 42
1.3.4 Permeation Test ................................................................................................................... 44
1.3.5 Surface Roughness Analysis ............................................................................................... 46

vi
1.3.6 Adhesion Test ...................................................................................................................... 46
1.4 Applications................................................................................................................................ 49
1.4.1 Capture of Hazardous Gases ............................................................................................... 49
1.4.2 Membrane Separation .......................................................................................................... 52
1.4.3 Destruction of Chemical Warfare Agents ........................................................................... 54
1.4.4 Antibacterial Protection ....................................................................................................... 57
1.4.5 Catalytic NO Release .......................................................................................................... 58
1.5 Outlook ....................................................................................................................................... 59
References ........................................................................................................................................ 60
CHAPTER 2. Highly Adsorptive, MOF-Functionalized Nonwoven Fiber Mats for Hazardous Gas
Capture Enabled by Atomic Layer Deposition .................................................................................... 68
Abstract ............................................................................................................................................ 69
2.1 Introduction ................................................................................................................................ 69
2.2 Experimental Methods ............................................................................................................... 71
2.3 Results and Discussion ............................................................................................................... 74
2.4 Conclusion .................................................................................................................................. 84
References ........................................................................................................................................ 85
Supporting Information .................................................................................................................... 87
CHAPTER 3. Conformal and highly adsorptive metal–organic framework thin films via
layer-by-layer growth on ALD-coated fiber mats ................................................................................ 96
Abstract ............................................................................................................................................ 97
3.1 Introduction ................................................................................................................................ 98
3.2 Experimental Methods ............................................................................................................. 100
3.3 Results and Discussion ............................................................................................................. 103
3.4 Conclusions .............................................................................................................................. 114
References ...................................................................................................................................... 116
Supporting Information .................................................................................................................. 120
CHAPTER 4. Facile Conversion of Hydroxy Double Salts to Metal-Organic Frameworks Using
Metal Oxide Particles and ALD Thin Film Templates....................................................................... 121
Abstract .......................................................................................................................................... 122
4.1 Introduction .............................................................................................................................. 122

vii
4.2 Results and Discussion ............................................................................................................. 124
4.3 Conclusion ................................................................................................................................ 134
References ...................................................................................................................................... 135
Supporting Information .................................................................................................................. 137
CHAPTER 5. Conformal Metal-Organic Framework Thin Films on Nanofibers for Ultra-Fast
Degradation of Chemical Warfare Agents ......................................................................................... 160
Abstract .......................................................................................................................................... 161
5.1 Introducion ............................................................................................................................... 161
5.2 Results and Discussion ............................................................................................................. 164
5.3 Conclusion ................................................................................................................................ 175
References ...................................................................................................................................... 176
Supporting Information .................................................................................................................. 179
CHAPTER 6. Diffusion of CO2 in large single crystals of Cu-BTC MOF ........................................ 199
Abstract .......................................................................................................................................... 200
6.1 Introduction .............................................................................................................................. 200
6.2 Experimental Methods ............................................................................................................. 202
6.3 Results and Discussion ............................................................................................................. 205
6.4 Conclusion ................................................................................................................................ 216
References ...................................................................................................................................... 217
Supporting Information .................................................................................................................. 220
CHAPTER 7. ALD-Assisted Synthesis of ZnO 3D Hierarchical Nanostructures for Enhanced
Photocatalytic Dye Degradation ......................................................................................................... 228
Abstract .......................................................................................................................................... 229
7.1 Introduction .............................................................................................................................. 229
7.2 Experimental Methods ............................................................................................................. 231
7.3 Results and Discussion ............................................................................................................. 235
7.4 Conclusion ................................................................................................................................ 243
References ...................................................................................................................................... 244

viii
LIST OF TABLES

Table 1.1 Summary of synthesis methods and properties of MOF-functionalized fibers ................. 38
Table 1.2 CO2 permeance and permselectivity for CO2/N2 for pure Ultem and Ultem/ZIF-8 hollow
fiber membranes ................................................................................................................................... 54
Table 2.1 Characterization of MOF-functionalized fiber mats: fiber mass gain (mg/cm2) after MOF
integration, BET surface area (m2/g) of MOF-fiber materials and NH3 dynamic loading (mol/kg)
calculated from breakthrough curves. .................................................................................................. 82
Table 4.1 Routes for HDS-Driven Room-Temperature Synthesis of Various MOF Materials. ..... 131
Table S4.1 ICP-OES for the concentration of Cu and Zn in the HKUST-1 crystals prepared through
the rapid room-temperature MOF synthesis. ...................................................................................... 149
Table S4.2 ICP-OES for the concentration of Cu and Zn in the filtrate after collecting the MOF
powder by filtration. ........................................................................................................................... 149
Table 5.1 Material properties and catalytic performance towards CWA simulant degradation. ..... 172
Table S5.1 MOF mass fraction in MOF-nanofiber composites calculated from BET results and
ICP-OES analysis ............................................................................................................................... 188
Table S5.2 Kinetic constants and half-lives of DMNP hydrolysis with UiO MOF powders and UiO
MOF thin films on PA-6@TiO2 nanofibers. ...................................................................................... 193
Table 6.1 Micropore diffusion fitting parameters for various Cu-BTC crystal sizes at 0.5% CO2. 213
Table 6.2 Micropore diffusion fitting parameters for 0.7 mm Cu-BTC single crystals and BPL
activated carbon at different CO2 concentrations. .............................................................................. 215
Table 7.1 Feature sizes and photocatalytic performance of ZnO 3D hierarchical nanostructrues. . 242

ix
TABLE OF FIGURES

Figure 1.1 Schematic illustration of the constituents for metal-organic frameworks. ......................... 3
Figure 1.2 Some common and representative MOF structures. .......................................................... 4
Figure 1.3 SEM image of HKUST-1 grown on chemithermomechanical pulp fibers using direct
solvothermal synthesis ........................................................................................................................... 6
Figure 1.4 Carboxymethylation on cellulose for anchoring HKUST-1 MOF coatings ...................... 7
Figure 1.5 SEM of HKUST-1 grown on cotton fibers pretreated via carboxymethylation. Insert
image is an EDX spectrum showing the presence of Cu ........................................................................ 7
Figure 1.6 SEM images of (a) ZIF-90 seeds deposited onto Torlon hollow fiber surface by dip
coating, (b-e) polycrystalline ZIF-90 thin film on Torlon hollow fiber. (f) XRD patterns for
Torlon/ZIF-90 membrane and the bare Torlon substrate ....................................................................... 9
Figure 1.7 Schematic illustration of the process sequence for ALD Al2O3 ....................................... 10
Figure 1.8 (a-c) schematic of the synthesis approach using ALD nucleation layers for solvothermal
synthesis. (a) polymer fiber, (b) polymer fiber coated with ALD Al2O3 with rich surface hydroxyl
groups, (c) conformal MOF coatings formed on ALD-coated fiber. (d) SEM image of conformal
HKUST-1 crystals grown on ALD-coated PP fiber ............................................................................. 11
Figure 1.9 Illustration of the synthesis route for solvothermal growth of Zr-based UiO MOFs on
nylon nanofibers with ALD TiO2 nucleation layers ............................................................................. 13
Figure 1.10 SEM images and cross-sectional TEM images of (a,d) UiO-66 grown on PA@TiO2
nanofibers, (b,e) UiO-66-NH2 grown on PA@TiO2 nanofibers, and (c,f) UiO-67 grown on PA@TiO2
nanofibers ............................................................................................................................................. 13
Figure 1.11 SEM image of HKUST-1 particles grown silk fibers using ultrasound-enhanced
layer-by-layer method .......................................................................................................................... 14
Figure 1.12 SEM images of HMTI-1 grown on silk fibers using (a-b) LbL method without
irradiation enhancement and (c-d) sonication-assisted LbL synthesis ................................................. 15
Figure 1.13 SEM images of silk fibers coated with MOF-5 using direct LbL synthesis (a) without
ultrasound irradiation and (b) with sonication ...................................................................................... 16
Figure 1.14 Surface functionalization of polyester fibers using polyvinylamine (PVA) and
bromoacetic acid (BAA) for layer-by-layer deposition of HKUST-1 thin films ................................. 17
Figure 1.15 SEM images of (a-b) untreated polyester fibers and (c-d) 40 cycles of HKUST-1
layer-by-layer growth on chemically functionalized polyester fibers .................................................. 18
Figure 1.16 SEM images of (a) untreated cotton fibers and (b-f) HKUST-1 coated cotton fibers ... 19

x
Figure 1.17 Illustration of the synthesis approach using ALD Al2O3 as nucleation layer for LbL
synthesis of HKUST-1 conformal thin films........................................................................................ 20
Figure 1.18 (a-b) SEM images of ALD-coated PP fibers with 40 cycles of LbL HKUST-1. (c-d)
Cross-sectional TEM images showing the conformal HKUST-1 thin films deposited on ALD-coated
PP fibers ............................................................................................................................................... 20
Figure 1.19 Schematic for MOF conversion from ALD metal oxide thin films ............................... 22
Figure 1.20 SEM images of conformal ZIF-8 and MIL-53-NH2(Al) thin films grown on PAN
nanofibers using ALD-to-MOF conversion ......................................................................................... 23
Figure 1.21 Schematic for MOF conversion from hydroxy ALD Metal Oxide Thin Films ............. 24
Figure 1.22 SEM Images and cross-sectional TEM image of HKUST-1 thin films grown on (a-b)
PP microfibers and (c-d) PAN nanofibers using HDS-to-MOF conversion ........................................ 25
Figure 1.23 Schematic of a horizontal electrospinning apparatus with rotational counter electrode 26
Figure 1.24 (a-b) Optical micrograph and SEM image of electrospun HKUST-1/PS fibers. (c)
Photo of MIL-100(Fe)/PVP fibers on a polypropylene nonwoven substrate. (d) SEM image of
electrospun MIL-100(Fe)/PVP fibers. (e-f) SEM images of HKUST-1/PAN fibers on a PAN
nonwoven substrate .............................................................................................................................. 27
Figure 1.25 (a) Optical photograph, (b) SEM image and (c) TEM image of ZIF-8–PVP nanofibers28
Figure 1.26 Schematic of electrospinning process and secondary growth for synthesizing MOF
membranes ........................................................................................................................................... 29
Figure 1.27 SEM image of (a) as-spun PS nanofibers with HKUST-1, (b) HKUST-1 membrane
after 4 cycles of secondary solvothermal synthesis on the PS/HKUST-1 nanofiber scaffold, (c)
as-spun PS nanofibers with ZIF-8, (d) ZIF-8 membrane after 5 cycles of secondary growth on
PS/ZIF-8 electrospun fibers.................................................................................................................. 30
Figure 1.28 SEM images of electrospun (a) UiO-66/PAN and (b) MIL-101(Cr)/PAN nanofiber
composites ............................................................................................................................................ 31
Figure 1.29 (a) N2 isotherms for UiO-66 powder and PVCi/UiO-66 nanofiber membrane
synthesized by electrospinning plus 3 cycles of secondary solvothermal growth at 100℃ ................. 33
Figure 1.30 Schematic of the setup for co-extrusion spinning to synthesize hollow fibers .............. 34
Figure 1.31 SEM images of (a) pure Ultem (b-c) Ultem/ZIF-8 hollow fiber membrane .................. 35
Figure 1.32 Photographs and SEM images of HKUST-1 printed on paper fibers for (a) 3 cycles and
(b-d) 8 cycles, respectively ................................................................................................................... 36
Figure 1.33 SEM images of UiO-66 sprayed on silk microfibers ..................................................... 37
Figure 1.34 Schematic of IR beam path in an ATR-IR ..................................................................... 40
Figure 1.35 ATR-IR spectra for cotton fabric coated with different number of cycles of HKUST-1 . .
.............................................................................................................................................................. 41

xi
Figure 1.36 (a) UV-Vis absorption spectra for different numbers of HKUST-1 layer-by-layer
cycles on polyester fibers pretreated with polyvinylamine and bromoacetic acid. (b) absorbance at
704 nm plotted as a function of cycle number ..................................................................................... 42
Figure 1.37 (a) Permeance for UiO-66/PVCi nanofiber membrane for He, N2, Ar and SF6. (b) a
schematic of the apparatus for permeation test .................................................................................... 45
Figure 1.38 AFM images and roughness profiles of (a) untreated polyester fiber surface and (b)
chemically modified polyester fibers deposited with 40 cycles of HKUST-1 layer-by-layer growth . 46
Figure 1.39 Mass change of ALD-coated PBT fiber mats with HKUST-1 thin films during the
compress air blowing test ..................................................................................................................... 47
Figure 1.40 (a) Photographs of cotton fiber mats printed with HKUST-1 dot pattern (5
printing-drying cycles) before and after exposure to NH3, HCl and H2S. (b) Relative mass change for
the HKUST-1 thin film printed on QCM during NH3 adsorption/desorption tests. (c) Adsorption
capacities of the HKUST-1 films printed on QCM for different gases. ............................................... 49
Figure 1.41 NH3 breakthrough curves for bare PP fibers (PP, green blank circle), PP fibers with
ALD Al2O3 coatings (PP/ALD, orange blank diamond), HKUST-1 grown on untreated PP fibers
(MOF-PP, blue circle), HKUST-1 grown on PP/ALD fiber mats (MOF-PP/ALD, red diamond) ...... 50
Figure 1.42 (a) Ammonia breakthrough curves for LbL HKUST-1 thin films grown on
ALD-coated PP fibers and control sample without MOF coating. (b) Ammonia dynamic loading for
PP/ALD fiber mats with 0 ~ 40 cycles of LbL HKUST-1. Black square points were calculated from
the breakthrough data before saturation, and red circle points were calculated based on the
breakthrough curves with desorption parts. Error bar shows standard deviation ................................. 51
Figrue 1.43 Ammonia dynamic loading on HKUST-1 MOF coated PP microfibers and PAN
nanofibers ............................................................................................................................................. 52
Figure 1.44 (a-c) Single gas permeances for Torlon/ZIF-90 hollow fiber membranes measured at (a)
35 oC and 50 psia, (b) 70 oC with varying pressure, (c) 50 psia with varying temperature. (d)
Permeances of hydrocarbon molecules measured at 22 oC by pervaporation ...................................... 53
Figure 1.45 (a) SEM images of silk fibers coated with UiO-66@LiOtBu. (b-d) Catalytic hydrolysis
profiles of CWA simulants using silk@[UiO-66@LiOtBu] composite catalysts ................................ 55
Figure 1.46 (a) Reaction schematic of DMNP catalytic hydrolysis using MOF-nanofiber catalyst.
(b) UV/Visible absorption spectra to monitor DMNP degradation. (c-e) DMNP conversion during
hydrolysis with different MOF powders and MOF-functionalized nanofiber catalysts ....................... 56
Figure 1.47 Photos of the petri dishes for antibacterial tests using HKUST-1 coated silk fibers. The
results were compared with untreated silk, gentamycin (GM), amoxicillin (AMX), cefepime (FEP)
and oxacillin (OX). Samples in Group I were prepared via LbL synthesis without sonication, while
Group IV were synthesized with ultrasound irradiation ....................................................................... 57
Figure 1.48 (Top) Catalytic reaction to form NO from S-nitrosocysteamine. (Bottom) NO release
as function of time with different catalytic materials ........................................................................... 58

xii
Figure 2.1 (a-c) Schematic illustration of the synthesis route. (a) Polymer fiber substrate. (b)
Al2O3-coated polymer fiber via atomic layer deposition (ALD). The cross section in the dashed
square illustrates the conformal coating of ALD Al2O3 with hydroxyl surface termination. (c) MOFs
integated on Al2O3-coated polymer fiber using solvothermal MOF synthesis. (d) SEM image of
HKUST-1 MOF crystals grown on MOF-PP/ALD.............................................................................. 75
Figure 2.2 Comparison of HKUST-1 MOF grown on untreated and ALD-coated polypropylene
fiber mats. (a) SEM image of HKUST-1 MOF grown on untreated polypropylene fiber mats
(MOF-PP). Insert image is a photograph of MOF-PP (scale bar represents 1cm). b) SEM image of
HKUST-1 MOF grown on ALD-Al2O3-coated polypropylene fiber mats (MOF-PP/ALD). Insert
image is a photograph of MOF-PP/ALD (scale bar represents 1cm). c) X-ray diffraction of MOF-PP
and MOF-PP/ALD. d) Mass increase percentage based on substrate dry weight. Interquartile range
and average value were calculated based on 16 MOF-PP samples and 12 MOF-PP/ALD samples. e)
Brunauer–Emmett–Teller (BET) surface area of MOF-fiber materials and calculated BET surface
area for MOF part (error bars represent standard deviation). BET surface area for PP fiber substrates
is 1.3~1.5 m2/g. The values were measured for uncoated and ALD coated fibers, and the values were
indistinguishable. .................................................................................................................................. 77
Figure 2.3 NH3 breakthrough curves for untreated polypropylene fiber mats (PP, ○ ),
ALD-Al2O3-coated polypropylene fiber mats (PP/ALD, ◇), HKUST-1 MOF grown on untreated
polypropylene fiber mats (MOF-PP, ● ), HKUST-1 MOF grown on ALD-Al2O3-coated
polypropylene fiber mats (MOF-PP/ALD, ◆). .................................................................................. 79
Figure 2.4 MOF-fiber mats with different ALD coating thicknesses. a) MOF crystal size
distribution on the top surface. b) MOF crystal size distribution on the cross-section. Crystal size
distributions were analyzed based on 50 measured crystal sizes on each corresponding SEM image. c)
Mass increase of the fiber mats after MOF integration. d) BET surface area of the
MOF-functionalized fiber mats and the surface area of MOF component in the fiber mats. e) NH 3
dynamic loading on MOF-functionalized fiber mats. f) NH3 breakthrough time on
MOF-functionalized fiber mats. ........................................................................................................... 81
Figure 2.5 HKUST-1 MOF grown on different ALD coatings and different polymer fibers. a) SEM
image of MOF on ALD-ZnO-coated polypropylene fiber mats (MOF-PP/ALD(ZnO)). b) SEM
image of MOF on ALD-TiO2-coated PP fiber mats (MOF-PP/ALD(TiO2)). c) SEM image of MOF
on ALD-Al2O3- coated PP fiber mats (MOF-PP/ALD(Al2O3)). d) SEM image of MOF on
ALD-Al2O3-coated polybutylene terephthalate (PBT) fiber mats (MOF-PBT/ALD(Al2O3)). e) SEM
image of MOF on ALD-Al2O3-coated cotton fiber mats (MOF-Cotton/ALD(Al2O3)). ....................... 83
Figure S2.1 Surface wetting ability of nonwoven polypropylene fiber mats with different number
of ALD Al2O3 cycles (0~500 cycles). 50 ALD cycles generates a very thin (~6nm) Al2O3 coating,
but this film was “thick” enough to change the wettability of PP substrates from originally
hydrophobic to hydrophilic. Contact angles of PP in ethanol are 0 degree regardless of ALD cycles.87

xiii
Figure S2.2 (a) Mass increase of PP/ALD fiber mats due to MOF growth at different temperatures
(100~140 ℃ ). (b) total BET surface area and calculated MOF surface area of HKUST-1
functionalized PP/ALD fiber mats prepared at different temperatures (100~140℃). ......................... 88
Figure S2.3 (a) Optical image of MOF-PP/ALD fiber mats prepared at different temperatures
(100~140℃). (b) SEM image of the side products found in MOF-PP/ALD fiber mats synthesized at
100℃. (c) SEM image of the flower-like CuO poly-crystals found in MOF-PP/ALD fiber mats
synthesized at 140℃. ........................................................................................................................... 89
Figure S2.4 X-ray diffraction data of MOF-PP/ALD fiber mats prepared at different temperatures
(100~140ºC) and PP substrate. ............................................................................................................ 90
Figure S2.5 Mass change of MOF funbctionalized fiber mats during the adhesion tests (data shown
are based on MOF-PBT/ALD fiber mats). The mass loss stabilizes after <4 min of air flow, so
empirically we take m/mo(t=4min) to evaluate the attachment of MOF crystals to the fiber substrate.
Mass change of MOF-PP/ALD during the adhesion tests is similar to MOF-PBT/ALD. ................... 91
Figure S2.6 Rapid micro-breakthrough analysis equipment. ............................................................ 92
Figure S2.7 Representative SEM images showing the top surface and the cross-section of
MOF-functionalized fibrous materials based on polypropylene fibers with different ALD coating
thickness. .............................................................................................................................................. 93
Figure S2.8 MOF crystal size distribution in PP fiber mats with different ALD thickness. ............. 94
Figure S2.9 SEM image of MOF crystals grown inside ALD-Al2O3 coated cotton fibers. .............. 95
Figure S2.10 (a) SEM image of UiO-66 MOF grown on PP/ALD fiber mats. (b) SEM image of
Zn-MOF-74 grown on PP/ALD fiber mats. (c) SEM image of Mg-MOF-74 grown on PP/ALD fiber
mats. ..................................................................................................................................................... 95
Scheme 3.1 Schematic of the synthesis route. Polymer fiber substrates were coated with 50 cycles
of ALD-Al2O3, forming a core@shell structure of “Fiber@Al2O3”. HKUST-1 MOF thin film was
grown onto ALD-coated polymer fibers via layer-by-layer synthesis method. ................................... 99
Scheme 3.2 Schematic of the rapid micro-breakthrough analysis system for NH3 and H2S
breakthrough tests............................................................................................................................... 103
Figure 3.1 (a,b,e) SEM images for ALD-Al2O3-coated PP fibers with 40 cycles of LbL HKUST-1
MOF (PP@ALD@LbL40). (c,d) Cross-sectional TEM images for PP@ALD@LbL40 showing the
core@shell structure. (f-i) Energy Dispersive X-ray analysis for PP@ALD@LbL40 showing the
presence of carbon (f) from the polypropylene and the HKUST-1 MOF, oxygen (g) from the ALD
Al2O3 and the HKUST-1 MOF, aluminum (h) from the ALD coating and copper (i) from the MOF. ....
............................................................................................................................................................ 104

xiv
Figure 3.2 SEM images of (a) 20 cycles of LbL HKUST-1 MOF on untreated PP fibers and (b) 20
cycles of LbL HKUST-1 MOF on ALD-Al2O3-coated PP fibers (PP@ALD). MOF thin film grown
on untreated PP substrate exhibits poor uniformity, while MOF coating on PP@ALD shows good
conformality with complete coverage on fibers. ................................................................................ 106
Figure 3.3 (a) Optical images of ALD-Al2O3-coated PP fibers with 0~40 cycles of LbL HKUST-1
MOF. (b) X-ray diffraction data of ALD-Al2O3-coated PP fibers with 0, 20, 40 cycles of LbL
HKUST-1 MOF. Green triangles (▼) represent the peaks for HKUST-1 MOF. (c) Thickness of the
MOF coating (0~40 LbL cycles) on ALD-Al2O3-coated PP fibers measured from cross-section TEM
images. Error bars represent a 95% confidence interval based on 25 data points measured for each
fiber mat. Solid line is a linear-fitted line to the data points. (d) Percent mass increase of LbL
HKUST-1 MOF (0~40 cycles) based on the dry weight of the fiber substrates (△m/mo). Dashed line
is a linear-fitted line to the data points. .............................................................................................. 107
Figure 3.4 (a) Fourier transform infrared (FTIR) spectra for HKUST-1 thin films deposited on
ALD-Al2O3 coated silicon wafers using layer-by-layer (LbL) method. Al2O3 coated silicon wafers
were dipped sequentially in LbL precursor solutions for 1 hour with 5-minute ethanol wash steps in
between. (b) Plot of IR absorbance at ~1370 cm-1 (carboxylate symmetric stretching vibrations) vs.
number of LbL cycles. A linear increase was observed for the absorbance at ~1370 cm -1 after 2 LbL
cycles. ................................................................................................................................................. 110
Figure 3.5 Brunauer-Emmett-Teller (BET) surface area (in units of m2/gMOF+Fiber) of
ALD-Al2O3-coated PP fibers with 0~40 cycles of LbL HKUST-1 MOF. ......................................... 111
Figure 3.6 (a) NH3 breakthrough curves for ALD-coated PP fiber mat with no LbL MOF (■), 20
cycles of LbL MOF (▲) and 40 cycles of LbL MOF (●). (b) NH3 dynamic loading on ALD-coated
PP fiber mats with 0, 20, 40 cycles of LbL MOF. Square points (■) were calculated based on the
corresponding breakthrough curve before saturation, and circle points (●) were calculated with the
desorption part. Error bar represents standard deviation. ................................................................... 111
Figure 3.7 (a-c) SEM images of 20 cycles of LbL HKUST-1 MOF grown on ALD-Al2O3-coated
polypropylene (a), polyethylene terephthalate (b) and cotton (c). (d) Percent mass gain for 20 LbL
cycles on PP@ALD, PET@ALD and Cotton@ALD. (e) BET surface area of PP@ALD@LbL20,
PET@ALD@LbL20 and Cotton@ALD@LbL20. ............................................................................ 113
Figure S3.1 Schematic of the homemade hot-wall viscous-flow ALD reactor used for ALD Al2O3
coatings on fibers. In an ALD cycle, trimethylaluminum (TMA) and water are dosed sequentially
into the chamber, with a purge step of inert gas (N2) in between. Deposition temperature is controlled
by the furnace, and all the gas lines and valves are wrapped with heating tapes to prevent precursor
condensation. ...................................................................................................................................... 120

xv
Figure 4.1 (a) Schematic of the rapid room-temperature synthesis route for Cu3(BTC)2. ZnO reacts
with Cu(NO3)2 to form (Zn,Cu) hydroxy double salt. The (Zn,Cu) HDS converts to Cu3(BTC)2 via
fast anion exchange. (b) Powder XRD patterns for Cu3(BTC)2 synthesized in rapid room-temperature
method (black) and solvothermal method (red), and simulated Cu3(BTC)2 pattern (blue). (c) Percent
yield in 1 minute of reaction (black circle) and space-time-yield for the rapid room temperature
synthesis (red diamond). Insert image in (c) is an SEM image of a Cu3(BTC)2 crystal showing the
octahedral shape. ................................................................................................................................ 125
Figure 4.2 (a) XRD patterns for Cu3(BTC)2 powder (black), ALD ZnO surface after exposure to
Cu(NO3)2 for 1 min (blue) and subsequently to H3BTC for 30 s (red). Blue dot lines represent the
(Zn,Cu) hydroxy double salt. (b) FTIR difference spectra for ALD ZnO in the as-deposited form
(black, Si as background), after exposure to Cu(NO3)2 for 1 min (magenta, previous spectrum as
background), and after exposure to H3BTC for 30 s (green, previous spectrum as background) and
the final Cu3(BTC)2 spectrum (orange, Si as background). (c-d) HAADF STEM images for the cross
section of the Cu3(BTC)2 grown on ALD ZnO coated silicon wafer. The green box in (c) indicates
the location of (d-i). (e-i) High resolution EDX mapping images of the cross section. Scale bars in
(b-g) represent 50nm. ......................................................................................................................... 128
Figure 4.3 (a) Schematic of the fabrication procedure for HKUST-1 patterns. (b-e) SEM image and
EDX mapping images for a star-shape HKUST-1 pattern. ................................................................ 130
Figure 4.4 (a) Schematic of the rapid room-temperature synthesis route for MOF coatings onto
various form factors. (b-d) SEM images of HKUST-1 deposited onto PS spheres, silicon wafer and
PAN nanofibers, respectively. ............................................................................................................ 132
Figure 4.5 (a) BET surface area and (b) NH3 dynamic loading for untreated PP microfibers and
PAN nanofibers, and MOF-coated PP and PAN fibers (MOF-PP, MOF-PAN, respectively)........... 133
Figure S4.1 Size distribution of ZnO nanoslurries dispersed in deionized water measured by
dynamic light scattering (DLS). The average particle size is 252±23 nm, and the polydispersity index
is 0.732. .............................................................................................................................................. 145
Figure S4.2 (a-b) SEM images of HKUST-1 crystals obtained from the rapid room-temperature
synthesis. (c) Crystal size distribution analyzed from SEM images (100 measured data points). The
average crystal size is 1.17±0.40 μm. ................................................................................................. 146
Figure S4.3 (a) SEM image of HKUST-1 crystals prepared via rapid synthesis (dispersed on a
silicon wafer). (b-d) Energy dispersive X-ray (EDX) mapping images for C, O and Cu in the
HKUST-1 crystals. (e) EDX spectrum showing that no Zn can be detected by EDX (blue dash lines
indicate where Zn can be expected). .................................................................................................. 147
Figure S4.4 (a-c) ToF-SIMS surface mapping (negative ion mode) images for a layer of densely
packed HKUST-1 crystals. These results confirm the presence of (a) oxygen, (b) carbon and (c)
copper. (d) ToF-SIMS depth profile (positive ion mode) for C+, Cu+ and 65Cu+ in a layer of densely
packed HKUST-1 crystals. Zn-containing residue was not found by either positive or negative
detection mode. .................................................................................................................................. 148

xvi
Figure S4.5 N2 adsorption and desorption isotherm for the HKUST-1 powder prepared via rapid
synthesis. ............................................................................................................................................ 150
Figure S4.6 SEM images for HKUST-1 grown on top of (Zn,Cu) hydroxy double salt. ................ 151
Figure S4.7 Powder XRD pattern for the Cu-BDC MOF converted from (Zn,Cu) hydroxy nitrate
HDS at room temperature (red). The simulated Cu-BDC pattern is shown in black. The XRD pattern
of the Cu-BDC obtained from HDS agrees well with the simulated pattern and reported powder XRD
patterns for this MOF.6,7 The results from the synthesis of HKUST-1 and Cu-BDC from (Zn,Cu)
HDS indicate that (Zn,Cu) hydroxy nitrate is an intermediate that preferentially converts to Cu-based
MOFs. ................................................................................................................................................. 152
Figure S4.8 Powder XRD patterns for the (Zn,Zn) HDS (blue) and IRMOF-3 converted from
(Zn,Zn) HDS at room temperature (red). The simulated IRMOF-3 pattern is shown in black. ......... 153
Figure S4.9 Powder XRD patterns for the (Zn,Zn) hydroxyl acetate HDS synthesized with DMF
(blue) and ZIF-8 converted from (Zn,Zn) HDS (red). The simulated ZIF-8 pattern is shown in black.
............................................................................................................................................................ 154
Figure S4.10. (a) Photo of circular and star-shape HKUST-1 patterns. The red box in (a) represents
the location of Image (b). (b-d) Optical micrograph and SEM images of star patterns made from
HKUST-1. The red box in (c) shows the location of (d). ................................................................... 155
Figure S4.11 SEM images for (a) untreated PS microspheres, (b) PS microspheres with ALD ZnO
coating, and (c-d) HKUST-1 grown on ZnO-coated PS microspheres. ............................................. 156
Figure S4.12 (a-c) SEM images for ZnO-coated PP fibers (a) before and (b-c) after HKUST-1
rapid synthesis. Insert photo in (b) shows the macroscopic uniformity of MOF growth on the PP fiber
mat. (d) Cross-sectional TEM image shows uniform HKUST-1 coating on ZnO-coated PP fibers.
(e-f), SEM images for ZnO-coated PAN nanofibers (e) before and (f) after HKUST-1 rapid synthesis.
Insert optical image in (f) shows the uniform MOF growth on the PAN nanofiber mat. .................. 157
Figure S4.13 (a) XRD patterns for ALD ZnO coated polypropylene fiber mat (PP/ZnO, red) and
HKUST-1 grown on PP/ZnO (MOF-PP, blue). (b) XRD patterns for ALD ZnO coated PAN
nanofiber mat (PAN/ZnO, red) and HKUST-1 grown on PAN/ZnO (MOF-PAN, blue) .................. 158
Figure S4.14 (a) NH3 breakthrough curves for untreated PP and MOF-PP fiber mats. (b) H2S
breakthrough curves for untreated PP and MOF-PP fiber mats. (c) NH3 breakthrough curves for
untreated PAN and MOF-PAN fiber mats. ........................................................................................ 159
Figure 5.1 Schematic of the synthetic procedure for Zr-based MOF functional coatings on
polyamide-6 nanofibers. The MOF crystal structures are illustrated in the dashed box .................... 165
Figure 5.2 (a) Photo of a free-standing PA-6@TiO2@UiO-66-NH2 nanofiber mat. (b-d) SEM
images, (e-i) energy dispersive X-ray mapping images. .................................................................... 165

xvii
Figure 5.3 (a-c) SEM images of (a) PA-6@TiO2@UiO-66, (b) PA-6@TiO2@ UiO-66-NH2 and (c)
PA-6@TiO2@UiO-67. (d-f) Cross-sectional TEM images of (d) PA-6@TiO2@UiO-66, (e)
PA-6@TiO2@UiO-66-NH2 and (f) PA-6@TiO2@UiO-67. (g-i) XRD patterns of PA-6 nanofibers
before and after ALD, MOF-coated nanofibers and MOF powders. (j-l) N2 adsorption and desorption
isotherms for PA-6@TiO2 nanofibers with and without MOF coatings and Zr-based MOF powders. ...
............................................................................................................................................................ 169
Figure 5.4 (a) Catalytic reaction of DMNP hydrolysis using Zr-based MOF powder and MOF
functionalized nanofiber catalyst. (b) UV/Visible absorption spectra for monitoring degradation of
DMNP. (c-e) Conversion of DMNP to p-nitrophenoxide versus reaction time using MOF powder
and functionalized nanofiber catalyst. ................................................................................................ 171
Figure 5.5 (a) Catalytic reaction of GD hydrolysis using MOF-nanofiber catalyst. (b) Conversion
of GD versus reaction time during catalysis. Dashed lines are fitted results assuming first order
reaction kinetics. ................................................................................................................................. 173
Figure S5.1 SEM images of UiO-66-NH2 grown on (a-b) untreated PA-6 nanofibers and (c-d) ALD
TiO2 coated PA-6 nanofibers. Depositing a thin ALD TiO2 nucleation layer significantly improves
the growth uniformity and crystal coverage on the fiber surface. ...................................................... 186
Figure S5.2. SEM images of (a-b) electrospun PA-6 nanofibers and (c-d) ALD TiO2 coated PA-6
nanofibers. The average diameter of untreated PA-6 nanofibers measured from SEM images is
37±16 nm. ........................................................................................................................................... 187
Figure S5.3 DMNP percent conversion as a function of time during the hydrolysis with untreated
PA-6 and ALD TiO2 coated PA-6 (PA-6@TiO2) nanofibers. Estimated t1/2 values are 3950 min and
1170 min for PA-6 and PA-6@TiO2, respectively. ............................................................................ 190
Figure S5.4 Kinetic analysis of DMNP degradation with (a-b) UiO-66, (c-d) UiO-66-NH2, (e-f)
UiO-67. The red curves in (a), (c), and (e) are plotted using the rate constants derived from the linear
fitting in (b), (d), and (f), respectively, based on the assumption of first order reaction kinetics. We
found that the reaction kinetics with UiO-66-NH2 does not fit well to first-order rate equation. ...... 191
Figure S5.5 Kinetic analysis of DMNP degradation with (a-b) PA-6@TiO2@UiO-66, (c-d)
PA-6@TiO2@UiO-66-NH2, (e-f) PA-6@TiO2@UiO-67. The red curves in (a), (c), and (e) are
plotted using the rate constants derived from the linear fitting in (b), (d), and (f), respectively, based
on the assumption of first order reaction kinetics. The reaction kinetics with
PA-6@TiO2@UiO-66-NH2 does not fit well to first-order rate equation. ......................................... 192
Figure S5.6 SEM images and EDX spectra of (a-b) PA-6@TiO2@UiO-66 nanofibers, (c-d)
PA-6@TiO2@UiO-66 nanofibers and (e-f) PA-6@TiO2@UiO-67 nanofibers after DMNP
degradation experiment. SEM images and EDX results confirm that significant amounts of MOF
coatings remain in the MOF-nanofiber composites even after strong agitation during the DMNP
hydrolysis tests. .................................................................................................................................. 194
31
Figure S5.7 P NMR spectra of GD during hydrolysis with PA-6@TiO2@UiO-66 nanofibers.... 195

xviii
Figure S5.8 31P NMR spectra of GD during hydrolysis with PA-6@TiO2@UiO-66-NH2 nanofibers.
............................................................................................................................................................ 196
31
Figure S5.9 P NMR spectra of GD during hydrolysis with PA-6@TiO2@UiO-67 nanofibers.... 197
Figure 6.1 (a-b) Molecular representation of the Cu-BTC MOF along the [100] direction and [111]
direction. Color code: Cu (yellow); O (red); C (black); H (not shown). (c-d) Optical microscopic and
(e-f) SEM images of Cu-BTC single crystals..................................................................................... 206
Figure 6.2 XRD patterns of a Cu-BTC single crystal for the planes parallel to (100) (black) and the
planes parallel to (111) (red), and simulated powder diffraction pattern for Cu-BTC. ...................... 207
Figure 6.3 Increase of Cu-BTC crystal size as a function of reaction time ..................................... 209
Figure 6.4 Gravimetric CO2 isotherm on large Cu-BTC single crystals at 298 K fit by a Toth
isotherm model. Also shown are data points from literature isotherms of CO2 on powder Cu-BTC at
298 K. ................................................................................................................................................. 210
Figure 6.5 Amplitude ratio curves for CSFR experiments on large Cu-BTC single crystals versus a
powder sample. ................................................................................................................................... 211
Figure 6.6 CSFR curves for different Cu-BTC crystal sizes at 0.5% CO2 concentrations. ............. 212
Figure 6.7 CSFR curves for various CO2 concentrations on Cu-BTC crystals of approximately 1.5
mm. ..................................................................................................................................................... 214
Figure S6.1 Schematic of the concentration-swing frequency response apparatus ......................... 220
Figure S6.2 Optical microscopic images of Cu-BTC crystals obtained from a three-day
solvothermal synthesis. The crystal size ranges from 500 μm to 1.3 mm. (Crystal size measurements
are labelled on the images.) ................................................................................................................ 221
Figure S6.3 (a) N2 isotherms for Cu-BTC single crystals and fine powder. (b) BET surface area
and (c) pore volume calculated from N2 isotherms. ........................................................................... 222
Figure S6.4 Thermogravimetric analysis (TGA) for Cu-BTC single crystals (red) and fine powder
(black)................................................................................................................................................. 223
Figure S6.5 The correlation between the mass per crystal and the crystal dimension. The error bar
indicates the standard deviation of the crystal sizes in a mass measurement group. .......................... 224
Figure S6.6 Crystal mass increase rate (dm/dt) as a function of reaction time based on Eqnation S3.
............................................................................................................................................................ 226
Figure S6.7 CO2 CSFR curves on BPL activated carbon at varying concentrations....................... 227
Figure 7.1 Schematic of the fabrication procedure for the ZnO 3D hierarchical nanostructures. (a)
Silicon wafer spun-coated with antireflection layer and positive photoresist. (b) Periodic
nano-pillars fabricated using laser interference lithography. (c) Conformal ZnO thin films deposited
onto the nano-pillars via ALD. (d) ZnO 3D hierarchical nanostructure obtained from hydrothermal
synthesis of ZnO nano-wires on ALD ZnO coated nano-pillars. ....................................................... 231

xix
Figure 7.2 Water droplet contact angles of the periodic nano-pillars (500 nm high) (a) before ZnO
ALD and (b) after ZnO ALD (200 cycles). Insert SEM images represent the nano-pillars before and
after ZnO ALD, respectively. Scale bars in SEM images represent 1μm. ......................................... 236
Figure 7.3 (a-d) SEM images of ALD ZnO coated nano-pillars (a) before and after (b) 60 min, (c)
120 min, and (d) 150 min of hydrothermal synthesis. (e) ZnO nano-wire length plotted as a function
of hydrothermal reaction time. ........................................................................................................... 237
Figure 7.4 (a) Schematic of ALD ZnO coated flat Si wafer (control) and nano-pillars with different
heights. (b) Normalized concentration as a function of time during the photocatalytic dye degradation.
(c) Apparent first order rate constant (black square) and half-life of MO (red circle) as a function of
the normalized illuminated surface area. ............................................................................................ 238
Figure 7.5 Comparison of photocatalytic performance between nanostructure with and without
ZnO nano-wires. (a) MO degradation curves for flat ZnO surface on Si wafer and ZnO nano-wires
on ZnO coated Si wafer. (b-d) MO degradation curves for ALD ZnO coated nanopillars with and
without ZnO nano-wires. The heights of nanopillars in (b-d) are 250 nm, 500 nm and 800 nm,
respectively......................................................................................................................................... 240

xx
CHAPTER 1 is a reprint of a manuscript in preparation.

CHAPTER 1. Introduction – Synthesis, Characterization and

Applications of Metal-Organic Framework Functionalized

Polymer Fibers

Junjie Zhao, Dennis T. Lee, Gregory N. Parsons*

Department of Chemical & Biomolecular Engineering, North Carolina State University,

Raleigh, NC 27695, U.S.A

*Corresponding author: gnp@ncsu.edu

1
1.1 Introduction

Metal organic frameworks (MOFs) are crystalline porous networks composed of

metal-containing clusters and organic bridging ligands.1 Common building blocks for MOFs

are shown in Figure 1.1, and representative MOF structures are shown in Figure 1.2. Due to

the wide variety of constituents, over 20,000 different MOF structures have been reported in

the past decade.2 These porous materials have garnered tremendous interest because MOFs

generally exhibit ultrahigh surface area (typically 1000 ~ 10,000 m2/g) and large pore

volume.2–6 The pore size, geometry and internal functionality can also be designed rationally

owing to the advantage of significant synthetic versatility.2,7 Post-synthetic processes have

also been established for modifying MOF structures to further enhance the properties.8–10

With these advantages, MOFs have been applied to many applications. For example,

adsorption of harmful gases,11–13 storage of hydrogen and methane,14,15 and sequestration of

carbon dioxide.16 Since new features (luminosity, structural flexibility, pH-sensitivity, and

electronic properties) of MOFs were discovered and taken advantage of, applications have

been extended to separations, drug delivery, chemical sensing and catalysis.7

MOFs obtained from common solvothermal synthesis are in powder forms that are not

soluble in most organic solvents. Therefore further handling and manipulation are required

for practical use. In addition, it is difficult to be integrated MOF powders into functional

materials and application-orientated devices.17,18 Surface-immobilized MOF thin films can

potentially overcome the inherent limitation of powders, and simplify the deployment. MOF

thin films also enable new composite structure and novel device configurations, and expand

the applications of MOFs as well.19,20

2
There has been growing interest in integrating MOFs with natural and synthetic fibers

since 2009,21 and this field is rapidly developing. Fibrous substrates are flexible and porous,

and often exhibit substantial external surface area and good gas permeability. These features

make fibers promising candidates as scaffolds for MOF thin films. A wide variety of

synthetic approaches have been developed in the past few year, and MOF-functionalized

fibers have also been demonstrated for the applications in hazardous gas capture,22–25

membrane separations,26–28 catalytic degradation of toxic compounds,29,30 and functional

biomedical materials.31,32

Figure 1.1 Schematic illustration of the constituents for metal-organic frameworks.

While recent reviews on topics of MOF composites and MOF thin films have partly

discussed about MOF integration on fibers,17,18,20 there is not a comprehensive review to date

3
to cover all the thoughtful strategies of synthesis, the characterization techniques and the

growing number of applications for MOF-coated fibers. In this review, our focus is on the

discussion of different synthesis approaches as well as the associated characterization

methods and application for MOF thin films on textiles and polymer hollow fibers.

Properties of the MOF-fiber composites are compared and listed in Table 1.1, including MOF

mass fraction, overall BET surface area (per unit mass of MOF plus fiber), microscopic

heterogeneity and crystal density on fiber surfaces observed in electronic microscopic

images.

Figure 1.2 Some common and representative MOF structures.

4
1.2 Synthetic Methods

1.2.1 Solvothermal Synthesis

1.2.1.1 Direct Solvothermal Growth

Küsgens et al. were the first few researchers who investigated MOF integration on fibers.

They investigated the growth of HKUST-1 MOF on chemithermomechanical pulp (CTMP),

bleached southern pine kraft pulp and unbleached kraft pulp fibers using direct solvothermal

synthesis.21 8 mmol of Cu(NO3)2 was dissolved in 50 mL of water, and 8 mmol of H3BTC

was dissolved in 100 mL DMF/ethanol (1:1 v/v). 1 g of pulp fibers were added to the mixed

precursor solution. After 15 min of stirring, the suspension was heated to 85oC for 24 h. The

mass fraction of HKUST-1 in the composite was determined by TGA, and the MOF weight

percent was 20.0 wt% and 10.7% for CTMP pulp fibers and unbleached pulp fibers,

respectively. Küsgens et al. found that higher lignin (containing COOH groups) content in

the pulp fibers results in higher mass loading for HKUST-1, consistent with the previous

study of HKUST-1 growth on COOH-terminated SAMs.33 However, the carboxylic acid

terminal groups in the pulp fibers were still not enough for a dense MOF coating (Figure 1.3).

Consequently, the MOF mass fraction and the surface coverage of HKUST-1 on the pulp

fibers were both low.

5
Figure 1.3 SEM image of HKUST-1 grown on chemithermomechanical pulp fibers using

direct solvothermal synthesis.21 (Scale bar adapted from Ref. 17)

Ullah et al. reported carbon nanofibers (CNFs) functionalized with MIL-53(Al) using

direct solvothermal synthesis.34 2.0 g of Al(NO3)2 and 0.949 g of H2BDC were mixed in 40

mL of DMF. 0.4 g of carbon nanofibers was then added to the mixed solution. The reaction

proceeded at 150 oC for 5 h. The resulting CNF@MIL-53 composite exhibits a BET surface

area of 140 m2/g. However, the BET surface area of their bulk MIL-53 powder (138 m2/g) is

quite low compared to reported values for this MOF (1140 – 1235 m2/g),35,36 probably due to

insufficient activation. The CNF@MIL-53 was characterized for CO2 adsorption, and shows

0.10 mmol/g CO2 uptake at 100 kPa.

1.2.1.2 Solvothermal Synthesis on Chemically Functionalized Surface

Carboxymethyltation has been used to add COOH terminal groups to cellulose fiber

surfaces, providing anchoring sites for Cu2+ ions as seeds for Cu-based MOFs.37 This surface

functionalization was based on the reaction between sodium chloroacetate and the hydroxyl

6
groups on cellulose with NaOH catalyst (Figure 1.4). Subsequently, Cu2+ ions from

Cu(OAc)2 precursor can be bound to the carboxylate groups on the treated cellulose surface

via ligand exchange. These Cu-carboxylate seeds enabled the nucleation and growth of

HKUST-1 crystals on cotton fibers when H3BTC linkers were further added to the growth

solution. XRD, XPS and ATR-FTIR results confirm the formation of HKUST-1 on chemical

modified cotton. 37 However, SEM images (Figure 1.5) show low density of MOF crystals on

the fiber surface. There was still space for improving the microscopy growth uniformity.

Figure 1.4 Carboxymethylation on cellulose for anchoring HKUST-1 MOF coatings.37

Figure 1.5 SEM of HKUST-1 grown on cotton fibers pretreated via carboxymethylation.

Insert image is an EDX spectrum showing the presence of Cu. 37

7
1.2.1.3 Solvothermal Synthesis on MOF-Seeded Surface

Using MOF crystals as seeds for secondary solvothermal synthesis has been

demonstrated for achieving continuous MOF films or membranes with high MOF mass

loading.27,38,39 MOF seeds can be deposited onto fiber surface by dip coating or incorporated

in polymer via electrospinning. The latter method will be discussed in the part on

electrospinning processes. Brown et al. synthesized ZIF-90 crystals with narrow size

distribution and deposited these seeds onto a Torlon (a polyamide-imide) hollow fiber by dip

coating the fiber in a ZIF-90 ethanolic suspension (4g/L).27 The seeded fiber was then

transferred to the mixed methanol solution (100 mL) containing imidazole carboxyaldehyde

(ImCA, 0.10 M) and Zn(NO3)2∙6H2O (0.025 M). The solvothermal synthesis proceeded at 65
o
C for 4 h, and a continuous polycrystalline ZIF-90 thin film was obtained on the Torlon

hollow fiber (Figure 1.6). In comparison, the attempts to grow ZIF-90 direct on Torlon fibers

failed to form uniform intergrown ZIF-90 crystals. Therefore seeding plays an important role

to achieve high-quality defect-free MOF thin films. SEM images show the thickness of the

ZIF-90 thin film is about 5 μm.27

8
Figure 1.6 SEM images of (a) ZIF-90 seeds deposited onto Torlon hollow fiber surface by

dip coating, (b-e) polycrystalline ZIF-90 thin film on Torlon hollow fiber. (f) XRD patterns

for Torlon/ZIF-90 membrane and the bare Torlon substrate.27

1.2.1.4 Solvothermal Synthesis on Atomic Layer Deposited Nucleation Layers

Inspired by the early work on MOF thin films grown on inorganic substrates,33,40 our

group developed a general approach to form metal oxide thin films on polymer fibers and use

the deposited inorganic surface for MOF nucleation and growth.23,24,30,41 Atomic layer

deposition (ALD) was used to deposit conformal metal oxide coatings onto fibrous substrates

in a vapor-phase process. Figure 1.7 shows a typical cycle of an example ALD reaction

9
sequence. In the first half-reaction, the precursor reacts with the substrates and leads to a

change in surface terminal groups which are ready to react with the co-reactant in the second

half reaction. Between each half-reaction, vapor byproduct and precursor residue are

removed from chamber using an inert gas purge step. The cycle repeats to build up a film

with thickness controlled by the number of ALD cycles used.42 Previous work in our group

has demonstrated that ALD is powerful in modifying polymers as well as enhancing the

wetting ability and conductivity of fibers.43–46 Recent work in our group has shown that ALD

nucleation layers significantly improve the conformality of MOF coatings on polymer fiber.

Figure 1.7 Schematic illustration of the process sequence for ALD Al2O3.

Zhao et al. showed that ALD Al2O3 (200 cycles deposited at 60 oC, approximately 24 nm)

thin films deposited on polypropylene (PP) fibers enable conformal growth of HKUST-1

micron-scale crystal on the fiber surface using solvothermal synthesis (Figure 1.8).23

10
Compared to untreated PP fibers, the MOF mass loading is increased from 9.61 mg/cm2 to

14.78 mg/cm2 with the ALD Al2O3 nucleation layer. The MOF mass fraction (75.3%) is also

among the highest values ever reported. The thickness of ALD Al2O3 films was found to

affect the MOF crystal size distribution on the fibers. The MOF mass loadings also vary on

different ALD surfaces (Al2O3, ZnO and TiO2) on PP fibers. This method was also applied to

polybutylene terephthalate (PBT) and cotton fibers, and further extended to MOF-74 and

UiO-66 coatings.23

Figure 1.8 (a-c) schematic of the synthesis approach using ALD nucleation layers for

solvothermal synthesis. (a) polymer fiber, (b) polymer fiber coated with ALD Al2O3 with

rich surface hydroxyl groups, (c) conformal MOF coatings formed on ALD-coated fiber. (d)

SEM image of conformal HKUST-1 crystals grown on ALD-coated PP fiber. 23

11
Our group have also investigated the growth mechanism of HKUST-1 on different metal

oxide nucleation surfaces on fibers during solvothermal synthesis.25,41 HKUST-1 growth rate

was found fastest on ALD ZnO coated PP fibers, because (Zn,Cu) hydroxy double salt

intermediate was quickly formed on the ZnO coatings after exposing to Cu(NO3)2. Further

anion exchange leads to the fast kinetics for HKUST-1 formation on ALD ZnO. A rapid

room-temperature synthesis route has also been developed based on this mechanism, which

will be discussed in the later part of this review.

Zhao et al. also applied ALD nucleation layers to nanofiber systems for solvothermal

sysnthesis of Zr-based MOFs including UiO-66, UiO-66-NH2, and UiO-67 (Figure 1.9).30

ALD TiO2 (100 cycles at 100 oC, approximately 5 nm) was deposited onto electrospun PA-6

nanofibers. Solvothermal synthesis (85 oC) with HCl modulator was performed on the

TiO2-coated PA-6 nanofibers (PA-6@TiO2). SEM and cross-sectional TEM images show that

conformal MOF coatings were formed onto the nanofiber surface for all the three types of

Zr-based UiO MOFs (Figure 1.10). The fact that the MOF crystals were grown directly on

and around the nanofibers indicates the strong attachment to the substrates. These

UiO-coated nanofibers were also applied to catalytic degradation of chemical warfare agents

and showed excellent activity comparable to the UiO MOF powders.30

12
Figure 1.9 Illustration of the synthesis route for solvothermal growth of Zr-based UiO

MOFs on nylon nanofibers with ALD TiO2 nucleation layers.30

Figure 1.10 SEM images and cross-sectional TEM images of (a,d) UiO-66 grown on

PA@TiO2 nanofibers, (b,e) UiO-66-NH2 grown on PA@TiO2 nanofibers, and (c,f) UiO-67

grown on PA@TiO2 nanofibers.30

13
1.2.2 Layer-by-Layer Synthesis

1.2.2.1 Direct Layer-by-Layer Synthesis

Similar to direct solvothermal synthesis, substrate materials with rich COOH groups are

desired for direct layer-by-layer (LbL) method without pretreatment on fiber surfaces. Abbasi

et al. investigated ultrasound-enhanced LbL process to grow HKUST-1 on silk.31 The silk

substrates were sequentially dipped in Cu(OAc)2 aqueous solution and H3BTC in

DMF/ethanol mixed solution, with washing steps in water between the exposure to reactants.

Sonication was used to accelerate the LbL reactions and to improve the homogeneity of MOF

coatings on fibers. Although the MOF growth uniformity was improved compared to the

early reports using direct solvothermal synthesis, this method was not able to form

continuous MOF films (Figure 1.11). While the authors claimed that the COOH groups are

the anchoring sites for the MOFs, the density of the carboxylic acid groups present in the

chain terminals of polypeptides in natural silk is actually very low.47 As a result, silk may not

be an ideal substrate for direct MOF synthesis.

Figure 1.11 SEM image of HKUST-1 particles grown silk fibers using

ultrasound-enhanced layer-by-layer method.31

14
Direct LbL synthesis was also applied for growing [Pb(NO3)2(4-bpdh)(H2O)]n (HMTI-1)

on silk fibers.48 0.033 M of Pb(NO3)2 aqueous solution and 0.033 M of

2,5-bis(4-pyridyl)-3,4-diaza-2,4-hexadiene (4-bpdh) methanol solution were used as the

precursor solutions for LbL. In each LbL cycle, the silk substrate was dipped sequentially in

the two reactant solutions for 5 min, with rinsing steps (2 min) in water between the

immersion steps in precursors. 20 LbL cycles were deposited onto silk fibers. SEM images

(Figure 1.12) show high crystal density in the MOF film, but the film growth is not

conformal on the fiber surface. Using ultrasound irradiation to enhance the LbL synthesis,

the uniformity of HMTI-1 coating was improved. Decreased MOF crystal size was also

observed. 48

Figure 1.12 SEM images of HMTI-1 grown on silk fibers using (a-b) LbL method without

irradiation enhancement and (c-d) sonication-assisted LbL synthesis .48

Khanjani et al. also investigated the synthesis of MOF-5 coatings on silk fibers using

direct LbL method.49 7.74 mmol Zn(OAc)2 DMF solution and 3.05 mmol H2BDC DMF

15
solution were used as precursor solutions. SEM images indicate that the morphology of the

MOF-5 coatings synthesized with sonication is more uniform than that obtained without

ultrasound irradiation (Figure 1.13). The MOF-5 coated silk fibers were tested for dye

adsorption, and the large uptake for cargo red dye confirms the porosity of the MOF-5

coatings. 49

Figure 1.13 SEM images of silk fibers coated with MOF-5 using direct LbL synthesis (a)

without ultrasound irradiation and (b) with sonication. 49

1.2.2.2 Layer-by-Layer Synthesis on Chemically Functionalized Surface

Meilikhov et al. showed a chemical functionalization approach to modify polyester

microfiber surface for layer-by-layer deposition of HKUST-1 thin films.50 The low density of

16
COOH groups in the polymer chain terminals was improved by the sequential reactions with

polyvinylamine (PVA) and bromoacetic acid (BAA) (Figure 1.14). The enriched carboxylic

acid terminal groups provide anchoring sites for the secondary building unites of HKUST-1.

In a layer-by-layer growth cycle, the modified polyester fibers were first immersed in 2 mM

Cu(OAc)2 ethanol solution for 10 min, and then washed with pure ethanol for 10 min.

Subsequently, the fibers were exposed to 0.2 mM H3BTC ethanol solution for 20 min, and

washed with pure ethanol again for 10 min. 40 cycles of LbL HKUST-1 thin films were

deposited onto the polyester fibers.

Figure 1.14 Surface functionalization of polyester fibers using polyvinylamine (PVA) and

bromoacetic acid (BAA) for layer-by-layer deposition of HKUST-1 thin films.50

XRD and IR spectroscopy confirm the formation of HKUST-1 on the modified polyester

fibers, and UV-Vis absorbance spectra show linear absorbance at 704 nm increases as a

function of the number of LbL cycles. MOF mass loading was determined by dissolving the

MOF thin films and measuring the Cu concentration via ICP-OES. Meilikhov et al. found

that for the polyester fibers treated with PVA and BAA the Cu uptake is 69 ± 9

17
μg∙gFiber-1∙cycle-1. Taking the Cu concentration in bulk HKUST-1 into account, this value

corresponds to MOF mass uptake of ~0.22 mg∙gFiber-1∙cycle-1. With 40 LbL cycles, the MOF

weight percent in the composite is approximately 0.88%. The authors claimed that

continuous HKUST-1 thin films have been formed on the modified polyester fibers. However,

large MOF agglomerates were also observed in the SEM images (Figure 1.15). The film

roughness is also not as good as those grown on SAM-functionalized Au substrates.50

Figure 1.15 SEM images of (a-b) untreated polyester fibers and (c-d) 40 cycles of

HKUST-1 layer-by-layer growth on chemically functionalized polyester fibers.50

Neufeld et al. describes a synthesis approach similar to the previously reported

procedures for functionalizing cotton surfaces with carboxylate groups37 and growing

Cu-BTC using layer-by-layer method.24,31,50 In this work, high concentration of the reactant

solutions were used, and the solvents for each precursors were different from typical

layer-by-layer synthesis. Specifically, 0.40 M of Cu(OAc)2 aqueous solution and 0.20 M of

H3BTC DMF/ethanol/water (equal parts) solution were used for the sequential dipping

18
process. In contrast to the continuous smooth thin films grown with low concentration

precursors,24 the modified recipe forms dense micron-scale crystals on the pretreated cotton

surface (Figure 1.16).

Figure 1.16 SEM images of (a) untreated cotton fibers and (b-f) HKUST-1 coated cotton

fibers.32

1.2.2.3 Layer-by-Layer Synthesis on Atomic Layer Deposited Nucleation Layers

Since ALD metal oxide thin films on polymer fibers have been demonstrated as good

nucleation layers for MOF solvothermal growth, our group have also investigated the growth

of MOF films on ALD-coated fibers using LbL method.24 Zhao et al. deposited ALD Al2O3

thin films on PP fibers (50 cycles at 60 oC, approximately 6 nm) and applied LbL synthesis

with ultrasound irradiation (Figure 1.17). Conformal and smooth HKUST-1 thin films were

19
obtained, and no agglomerates of MOF crystals were observed in SEM images (Figure 1.18).

The sub-micron crystal size of the LbL HKUST-1 is also much smaller than the crystals

synthesized via solvothermal methods. MOF film thickness and mass loading were found to

increase linearly with the number of LbL cycles, indicating an excellent control over the

MOF film growth. Low concentration of precursor solution and sufficient rinsing between

the precursor exposure steps are critical for high-quality MOF thin films.24

Figure 1.17 Illustration of the synthesis approach using ALD Al2O3 as nucleation layer for

LbL synthesis of HKUST-1 conformal thin films.24

Figure 1.18 (a-b) SEM images of ALD-coated PP fibers with 40 cycles of LbL HKUST-1.

(c-d) Cross-sectional TEM images showing the conformal HKUST-1 thin films deposited on

ALD-coated PP fibers.24

20
Zhao et al. also investigated the growth rate of HKUST-1 thin films in the LbL synthesis.

Film thicknesses were analyzed from cross-sectional TEM images, and a growth rate was

found to be 3.0 nm/cycle (or ~288 ng∙cm-2 per cycle). This value is close to the dimension of

one monolayer of HKUST-1 MOF. The growth rate measured from MOF mass uptake on

fibers ~290 ng∙cm-2 per cycle is also similar to the results from film thickness analysis. This

synthesis method was also applied to polyethylene terephthalate and cotton fibers and

generated uniform HKUST-1 films on these substrates as well.24

1.2.3 MOF Conversion from Thin Film Templates

1.2.3.1 Conversion from ALD Metal Oxide Thin Films

Metal oxides have been previously shown as possible precursors than can be directly

converted to MOF structures in the reaction with organic linkers under solvothermal

conditions.51,52 The same strategy can also be applied to MOF thin films on fibers.41,53 Figure

1.19 shows the general approach for MOF conversion from ALD metal oxide thin films.

Bechelany et al. reported the synthesis of ZIF-8 and MIL-53-NH2(Al) on PAN nanofibers

using this method.53 50 nm of ALD ZnO and 42 nm of ALD Al2O3 were deposited onto PAN

nanofibers respectively prior to MOF conversion. 2-methylimidazole methanolic solution

was used for ZIF-8 conversion from ZnO with microwave-assisted heating at 100 oC for 1.5 h,

while 2-aminoterephthalic acid aqueous solution was used for MIL-53-NH2(Al) conversion

from Al2O3 with the same heating condition and reaction time. 53

21
Figure 1.19 Schematic for MOF conversion from ALD metal oxide thin films.

The excellent conformality of ALD thin films can be translated into good microscopic

homogeneity of the MOF coatings. Bechelany et al. showed dense crystalline ZIF-8 and

MIL-53-NH2(Al) thin films formed on PAN nanofibers using this ALD-to-MOF conversion

(Figure 1.20). Since this conversion is localized,52 no agglomerates of MOFs are formed in

the fiber networks. Conventional heating and microwave-assisted heating were also

compared for this type of synthesis. The growth uniformity and crystal coverage on fibers for

samples prepared using conventional heating are not as good as those with microwave

irradiation. Nevertheless, this process provides opportunities for new types of conformal

MOF thin films.53

22
Figure 1.20 SEM images of conformal ZIF-8 and MIL-53-NH2(Al) thin films grown on

PAN nanofibers using ALD-to-MOF conversion. 53

Our group also observed ALD-to-MOF conversions when we were investigating surface

reactions between MOF precursors and ALD metal oxides.41 Lemaire et al. found that the

ALD Al2O3 coating on PP fibers can be converted to MIL-96 after exposure to H3BTC

solution with mixed ethanol/water (1:1 v/v) solvent at 120 oC for 20 h. Similarly, ALD ZnO

films on PP fibers were converted to Zn-BTC. These results further confirmed the general

applicability of ALD-to-MOF conversion for MOF thin films on fibrous substrates.41

1.2.3.2 Room-Temperature Conversion from Hydroxy Double Salts

Hydroxy double salts (HDSs) are layered compounds containing 2-D cationic sheets and

interlamellar anions, and can exhibit fast anion exchange rate.54 Zhao et al. found that HDSs

23
can also be converted to MOFs at room temperature through rapid anion exchange.25 (Zn, Cu)

hydroxy nitrate formed in situ from ZnO powder or films can further react with the organic

linkers and form HKUST-1 or Cu(BDC) (Figure 1.21). This process significantly reduces the

synthesis temperature and exhibits ultra-high space-time-yield (up to 3.6 × 104 kg·m−3·d−1).25

Figure 1.21 Schematic for MOF conversion from hydroxy ALD Metal Oxide Thin Films.

Using this fast HDS-to-MOF conversion, MOF thin films were synthesized on PP

microfibers and PAN nanofibers. SEM and TEM images show that dense HKUST-1 crystal

coatings were formed on the fiber surfaces (Figure 1.22). Similar to ALD-to-MOF

conversion, the conversion from HDS intermediates is also localized. The quality of the

MOF films as well as the microscopic uniformity is comparable to the conformal MOF

coatings synthesized via other method. Since this process can be performed at room

temperature, it is very promising for scale-up production of MOF-functionalize composites.25

24
Figure 1.22 SEM Images and cross-sectional TEM image of HKUST-1 thin films grown on

(a-b) PP microfibers and (c-d) PAN nanofibers using HDS-to-MOF conversion.25

1.2.4 Electrospinning

In a typical electrospinning process, the polarization of polymer solution draws fibers

from a feed droplet at the outlet of the metallic needle to the counter electrode collector

(Figure 1.23). The solvent evaporates during the process, and dry fiber mat can be obtained.55

This technique has been well developed for making polymer micro- and nanofibers. Recently,

electrospinning has also been demonstrated for the synthesis of MOF-incorporated fibers.

25
Figure 1.23 Schematic of a horizontal electrospinning apparatus with rotational counter

electrode.55 Vertical electrospinning setups are similar, since the basic principle is the same.

Rose et al. are the first few researchers to look into the synthesis of MOF-functionlized

textiles using electrospinning.55 Three polymer systems were investigated including

polyvinylpyrrolidone (PVP) in ethanol, polystyrene (PS) in THF and polyacrylonitrile (PAN)

in DMF, and the mass ratios in the mixed solutions for electrospinning are

HKUST-1/PAN/DMF = 1:0.25:4.7, HKUST-1/PS/THF = 1:1.6:4.3 and MIL-100(Fe)/PVP/

EtOH = 1:0.25:5.3. The feed flow rate was 4 mL/h. A high voltage of 10 – 20 kV was applied

and the collector was 10 – 15 cm from the needle. Up to 40 wt% was achieved for

HKUST-1/PS fibers, but the huge size difference between the MOF crystal and the fiber

diameter resulted in non-uniform MOF distribution in the fiber (Figure 1.24b). The

investigators further studied electrospinning PVP fibers with reduced diameters (100 – 500

nm) for physically trapping the MOF crystals in the fiber web (Figure 1.24d). However, the

instability of PVP in polar solvent was a great concern. The attachment of these micron-scale

MOF crystals to the nanofibers may also be weaker in this case than those chemically-bound

26
MOF coatings or incorporated MOF particles. While a high MOF weight percent (80%) was

achieved for HKUST-1/PAN system, large chunks of MOFs were observed in SEM images,

indicating poor uniformity of MOF distribution.

Figure 1.24 (a-b) Optical micrograph and SEM image of electrospun HKUST-1/PS fibers.

(c) Photo of MIL-100(Fe)/PVP fibers on a polypropylene nonwoven substrate. (d) SEM

image of electrospun MIL-100(Fe)/PVP fibers. (e-f) SEM images of HKUST-1/PAN fibers

on a PAN nonwoven substrate.55

Ostermann et al. found that MOF nanocrystals with narrow distribution are critical in

order to achieve a homogeneous MOF distribution in the electrospun MOF-polymer fibers.56

Using previous solvothermal synthesis method, they prepared ZIF-8 nanoparticles with

27
narrow distribution. 400 mg ~ 2000 mg of ZIF-8 was mixed with 12 wt% PVP methanol

solution. After dilution or concentration, the final solution with 3.5 wt% PVP concentration

was fed through a needle at the rate of 0.35 mL/h. The voltage for electrospinning was 5 kV,

and the aluminum foil collector was placed 6-8 cm from the needle. SEM and TEM images

(Figure 1.25) show that ZIF-8 nanoparticles are homogeneously embedded in the PVP

nanofibers. 56% ZIF-8 mass fraction was achieved in the MOF-nanofiber structure. As a

result, the overall BET surface area for the PVP-ZIF-8 exceeds 500 m2/g.56

Figure 1.25 (a) Optical photograph, (b) SEM image and (c) TEM image of ZIF-8–PVP

nanofibers.56

To avoid the disadvantage of PVP dissolution in polar solvents, Ostermann et al. also

applied the same process for polystyrene-ZIF-8 nanofibers. With 25 wt% MOF mass fraction,

the PS-ZIF-8 nanofiber composite exhibits overall BET surface area of 210 m2/g.

28
Polyethylene oxide (PEO) was also attempted for electrospinning MOF-nanofibers. However,

PEO-ZIF-8 shows no accessible MOF surface area because the polymer completely blocks

the MOF pores. The polymer layer covering the MOF particles was found to be an additional

diffusion barrier, so a low thickness is crucial for the polymer layer to maintain the

accessibility of the MOF pores in these electrospun MOF-nanofiber structures.56

Wu et al. also investigated electrospinning and explored membrane synthesis using

electrospun MOF-fibers as scaffold. PS in THF solution was used for HKUST-1 and ZIF-8,

and PVP in ethanol solution was used for MIL-101(Fe) and Zn2(bpdc)2(bpee)

(bpdc=4,4’-biphenyldicarboxylate; bpee=1,2-bipyridylethene). The electrospinning process is

similar to the method developed by Ostermann et al (Figure 1.26). With approximately 20 wt%

MOF mass fraction, the electrospun PS/HKUST-1 fibers exhibit BET surface area of 311

m2/g, and PS/ZIF-8 fibers have BET surface area of 303 m2/g. The homogeneous distribution

of MOF in the fibers is similar to the MOF-nanofiber structures Ostermann et al. have

reported.

Figure 1.26 Schematic of electrospinning process and secondary growth for synthesizing

MOF membranes.38

29
Using the as-spun nanofibers with MOF seeds embedded, further secondary

solvothermal growth can improve the MOF mass fraction up to 50% and enable the

formation of MOF membranes (Figure 1.27).38 Since MOF crystals uniformly cover the

entire surface of the fiber mat, nanofiber skeleton is no longer visible in SEM images after

repeated secondary growth. As a result of increased MOF weight percent, the overall BET

surface area for the MOF membranes is 944 m2/g for PS/HKUST-1 membrane after 4 cycles

of secondary solvothermal synthesis on the PS/HKUST-1 nanofiber scaffold. Similarly, the

PS/ZIF-8 membrane obtained after 5 cycles of secondary growth exhibits surface area of 935

m2/g. Wu et al. also investigated the potential of the PS/ZIF-8 membrane for CO2/N2

separation. However, due to macroscopic pores and defects in the membrane, the separation

factor of N2/CO2 was limited to 2.4. 38

Figure 1.27 SEM image of (a) as-spun PS nanofibers with HKUST-1, (b) HKUST-1

membrane after 4 cycles of secondary solvothermal synthesis on the PS/HKUST-1 nanofiber

scaffold, (c) as-spun PS nanofibers with ZIF-8, (d) ZIF-8 membrane after 5 cycles of

secondary growth on PS/ZIF-8 electrospun fibers.38

30
Figure 1.28 SEM images of electrospun (a) UiO-66/PAN and (b) MIL-101(Cr)/PAN

nanofiber composites.57

Ren et al. reported the fabrication of MOF-nanofiber composites containing Zr- and

Cr-based MOFs using similar electrospinning method. 100 mg of UiO-66 and MIL-101(Cr)

were mixed with 500 mg of 10 wt% polyacrylonitrile (PAN) DMF solution respectively, and

used for electrospinning.57 The feed rate of mixed suspension was controlled at 1.25 – 2.5

mL/h, and the voltage for electrospinning was set to 10 kV. Electrospun nanofibers were

collected on an aluminum foil which was 10 cm away from the solution outlet. The author

pointed out that the MOF mass fraction in the mixed suspension has to be controlled under

40 wt% in order to obtain proper viscosity for electrospinning. Although the authors

31
calculated the MOF weight percent in the MOF-fiber composites based on the MOF mass

fraction in the mixed suspension, they neglected the evaporation of DMF during the

electrospinning and post-synthetic drying processes. The actual MOF mass fraction in the

final composite form could be as high as 67%, assuming the removal of all the solvent in the

electronspun MOF-fibers. The ratio of the overall (MOF+fiber) BET surface area to the MOF

surface area is another good estimation of the actual MOF mass fraction in the composites.

The MOF weight percent calculated based on their BET surface area is 68.7% for UiO-66 on

PAN and 43.3% for MIL-101(Cr) on PAN, respectively. As a result of the high MOF mass

loading, the MOF-fiber composites also exhibited over 50% H2 adsorption capacity

compared with the bulk MOF powders. 57

Armstrong et al. reported a similar approach using electrospinning technique. UiO-66

crystals were mixed with 10 wt% poly(vinyl cinnamate) (PVCi) dichloromethane solution,

and fed through a needle at 4.2 mL/h rate. The voltage for electrospinning was 22 kV, and the

collector was 6 cm away from the needle. In addition to electrospinning, UV irradiation and

secondary growth were also applied to the MOF-nanofiber mats. During the UV irradiation,

PVCi was crosslinked via a photo-induced cycloaddition reaction, which was confirmed by

FTIR. As a result, the stability of PVCi nanofibers in DMF was significantly improved after

photo-crosslinking. Using the UiO-66 embedded in the PVCi nanofibers as the crystal seed,

the author also performed repeated secondary growth to obtain high MOF loading. After 3

cycles of secondary solvthermal growth at 100℃ on the electrospun MOF-nanofiber mat,

overall BET surface area of 430±1 m2/g was obtained. Based the overall surface area and

that for the UiO-66 powder (1188 m2/g), we estimate the MOF mass fraction in this

32
UiO-66/PVCi composite structure to be 36.2%. 39

Figure 1.29 (a) N2 isotherms for UiO-66 powder and PVCi/UiO-66 nanofiber membrane

synthesized by electrospinning plus 3 cycles of secondary solvothermal growth at 100℃.39

1.2.5 Co-Extrusion Spinning

Husain et al. demonstrated a spinning technique by co-extruding two polymer solutions

and applied this method for asymmetric polymer/zeolite hollow fibers.58 Later this process

was used to synthesize polymer/MOF hollow fibers.28 Figure 1.30 illustrates the

experimental setup for co-extrusion spinning. Basically, sheath dope (for the

33
MOF-containing skin) and core dope (for the porous polymer hollow fiber substrate)

solutions are extruded simultaneously with the presence of bore fluid for supporting the

nascent hollow fiber. The active sheath layer is formed when the solvent is evaporated in the

air gap, and the rapid phase separation in the subsequent wet quench process enables the

formation of large porosity in the core polymer layer. 28

Figure 1.30 Schematic of the setup for co-extrusion spinning to synthesize hollow fibers.58

Dai et al. synthesized Ultem (a polyetherimide) hollow fibers with a functional

ZIF-8-containing sheath layer using the co-extrusion spinning process.28 SEM images show

no obvious boundary between the core and sheath layer, indicating good adhesion. Since the

ZIF-8 crystals were embedded in the sheath layer like common flat polymer mixed matrix

membranes, this type of MOF/fiber structure is essentially designed for gas separation

applications.

34
Figure 1.31 SEM images of (a) pure Ultem (b-c) Ultem/ZIF-8 hollow fiber membrane.28

1.2.6 Inkjet Printing

Taking advantage of evaporation/solvent induced crystalization,22,59 Zhuang et al.

developed a precursor solution and used it as the ink for printing MOFs on papers and

textiles.60 Ethylene glycol was found to stablize the precursor solution with

ethanol/dimethylsulfoxide (DMSO) mixted solvents. The precursor ink were printed onto

PET foils, office paper and cotton textiles using a commertial printer, and the substrate

materials were dried at 80℃ for 2 mins. The thickness of the MOF coating and the density

of MOF crystals can be controled by the number of printing-drying cycles. This unique

process can achieve very good growth uniformity on the top surface (Figure 1.32) of the

substrates and can be easily applied for patterning on large substrates. However, due to the

limited penetration depth of the precursor ink into the complex fiber networks, it may be an

issue for this process to obtain good spacial homogeneity for textiles. Large MOF

agglomerates were also observed in the voids of paper substrates. 60

35
Figure 1.32 Photographs and SEM images of HKUST-1 printed on paper fibers for (a) 3

cycles and (b-d) 8 cycles, respectively. 60

1.2.7 Spay Drying Method

López-Maya et al. reported a process to physically disperse UiO-66 particles onto

fibrous substrates using spray drying method.29 Methanolic suspension containing

[Zn4O(dmcapz)3] or UiO-66 was sprayed through an atomizer during the electrospinning

process for silk fibers. The MOF particles spayed onto the substrates in each process cycle

are physically immobilized in the as-spun layers of silk fibers.29 While this method can

achieve high MOF mass fraction (50%), the MOF crystals are not chemically bound to the

fiber surface. The agglomerates of MOF particles (Figure 1.33) in the composites also exhibit

a smaller external surface area that is accessible during heterogeneous catalysis, compared

with the conformal MOF thin films grown onto the fiber scaffolds.

36
Figure 1.33 SEM images of UiO-66 sprayed on silk microfibers.29

37
Table 1.1 Summary of synthesis methods and properties of MOF-functionalized fibers.
Type of MOF Substrate Material Synthesis Method MOF wt% Overall BET Microscopic Crystal References

Surface Area Homogeneity Density on


2
(m /g) Fibers

HKUST-1 Cotton microfiber Layer-by-layer synthesis after carboxymetylation (sodium 18 ±2 N/A Good High 32
chloroacetate + NaOH)
HKUST-1 Chemithermomechanical pulp Direct Solvothermal 20.0 314 Good Low 21
(CTMP) microfiber
HKUST-1 Unbleached kranft pulp Direct Solvothermal 10.7 165 N/A N/A 21
microfibers
HKUST-1 Polyester microfiber Layer-by-layer synthesis after surface functionalization 0.88 <b> N/A Poor High 50

HKUST-1 Polystyrene microfiber Electrospinning 40 N/A Poor Low 55

HKUST-1 Polyacrylonitrile nanofiber Electrospinning 80 1200 Poor Low 55

HKUST-1 Polystyrene nanofiber Electrospinning 20 <c> 311 Good Low 38

HKUST-1 Polystyrene nanofiber Electrospinning + secondary solvothermal growth 50 <a> 944 Good High 38

HKUST-1 Cotton microfiber Solvothermal synthesis after carboxymetylation N/A N/A Poor Low 37

HKUST-1 Silk microfiber Direct layer-by-layer synthesis N/A N/A Good Low 31

HKUST-1 Paper Inkjet printing 10.1 N/A Good High 60

HKUST-1 Cotton microfiber Inkjet printing 2.7 N/A Good High 60

HKUST-1 Polypropylene microfiber Solvothermal on ALD nucleation layers 75.3 695 Good High 23

HKUST-1 Polybutylene terephthalate Solvothermal on ALD nucleation layers 52.5 <a> 485 Good High 23
microfiber
HKUST-1 Cotton microfiber Solvothermal on ALD nucleation layers 13.1 <a> 121 Poor Low 23

HKUST-1 Polypropylene microfiber Layer-by-layer on ALD nucleation layers 17 93 Good High 24

HKUST-1 Polyethylene terephthalate Layer-by-layer on ALD nucleation layers 1.2 12.2 Good High 24
microfiber
HKUST-1 Cotton microfiber Layer-by-layer on ALD nucleation layers 5.3 18.2 Good High 24

HKUST-1 Polypropylene microfiber MOF conversion from hydroxy double salt 10.6 <a> 201 Good High 25

HKUST-1 Polyacrylonitrile nanofiber MOF conversion from hydroxy double salt 27.7 <a> 524 Good High 25

MIL-100(Fe) Polyvinylpyrrolidone nanofiber Electrospinning N/A N/A Poor Low 55

MIL-101(Cr) Polyacrylonitrile nanofiber Electrospinning 43.3 <a> 1134 Good High 57

MIL-101(Fe) Polyvinylpyrrolidone nanofiber Electrospinning 67 <c> N/A Good Low 38

MIL-53(Al) Carbon nanofiber Direct Solvothermal N/A 140 N/A N/A 34

MIL-53-NH2(Al) Polyacrylonitrile nanofiber MOF Conversion from ALD Metal Oxide N/A N/A Good High 53

MOF-5 Silk microfiber Direct layer-by-layer synthesis N/A N/A Good Low 49

MOF-74(Mg) Polypropylene microfiber Solvothermal on ALD nucleation layers N/A N/A Good High 23

MOF-74(Zn) Polypropylene microfiber Solvothermal on ALD nucleation layers N/A N/A Good High 23

Pb(4-bpdh)(NO3)∙ Silk microfiber Direct layer-by-layer synthesis 61.5 N/A Poor High 48
(H2O)
UiO-66 Polyacrylonitrile nanofiber Electrospinning 68.7 <a> 815 Good High 57

UiO-66 Poly(vinyl cinnamate) nanofiber Electrospinning + UV cross-linking, + secondary growth 36.2 <a> 430 Poor Low 39

UiO-66 Polypropylene microfiber Solvothermal on ALD nucleation layers N/A N/A Good High 23

UiO-66 Silk microfiber Spray Drying Method 50 670 Poor Low 29

UiO-66 Polyamide-6 nanofiber Solvothermal on ALD nucleation layers 8.6 140 Good High 30

UiO-66@LiOtBu Silk microfiber Spray Drying Method N/A 130 Poor Low 29

UiO-66-NH2 Polyamide-6 nanofiber Solvothermal on ALD nucleation layers 13 196 Good High 30

UiO-67 Polyamide-6 nanofiber Solvothermal on ALD nucleation layers 19.4 447 Good High 30

ZIF-8 Polyvinylpyrrolidone nanofiber Electrospinning 56 530 Good High 56

ZIF-8 Polystyrene nanofiber Electrospinning 25 210 Good High 56

ZIF-8 Polystyrene nanofiber Electrospinning 20 <c> 303 Good Low 38

ZIF-8 Polystyrene nanofiber Electrospinning + secondary solvothermal growth 50 935 Good High 38

ZIF-8 Ultem (a polyetherimide) hollow Co-extrusion spinning 0.10 <b> N/A Good High 28
fiber
ZIF-8 Polyacrylonitrile nanofiber MOF Conversion from ALD Metal Oxide N/A N/A Good High 53

ZIF-90 Torlon® (a polyamide-imide) Solvothermal on MOF Seeds N/A N/A Good High 27
hollow fiber
Zn2(adc)2(dabco)2 Paper Inkjet printing N/A N/A N/A N/A 60

Zn2(bpdc)2(bpee) Polyvinylpyrrolidone nanofiber Electrospinning 50 <c> N/A Good Low 38

Zn4O(dmcapz)3 Silk microfiber Spray Drying Method 50 690 Poor Low 29

<a> MOF wt% calculated based on BET surface area. <b> MOF wt% estimated from metal
concentration. <c>MOF wt% estimated based on the MOF to polymer mass ratio in the
mixed solution for electrospinning.

38
1.3 Characterization Methods

Common characterization techniques for bulk MOF powders, including XRD, BET,

SEM, TEM and TGA, are generally applicable to MOF-functionalized fibers. In this part, we

will highlight a few analysis methods that are particularly useful for fibrous materials. The

methods presented here can be applied for investigation of surface composition, thin film

growth rate, MOF mass fraction, permeation property, surface roughness and mechanical

stability (MOF attachment).

1.3.1 Attenuated Total Reflection (ATR) Infrared spectroscopy

Developed by Harrick and Fahrenfort more than 50 years ago,61,62 ATR technique relies

on the total reflection of IR beam in an internal reflection element (IRE) and the evanescent

wave formed perpendicular to the reflecting surface in the IRE (Figure 1.34).63,64 To form

total reflection within the IRE, the angle of incidence θ needs to be larger than the critical

angle (θcrit):
𝑛
𝜃𝑐𝑟𝑖𝑡 = 𝑠𝑖𝑛−1 ( 2 ) (1)
𝑛1

where n1 and n2 are the refractive indices of the IRE and the sample, respectively.64 The

evanescence wave can penetrate a small distance (dp) into the sample and be absorbed by the

sample. The penetration depth is given by:

𝜆
𝑑𝑝 = (2)
2𝜋√𝑛12 𝑠𝑖𝑛2 𝜃−𝑛22

where λ is the wavelength of incident IR beam, θ is the angle of incidence, n1 and n2 are the

refractive indices of the IRE and the sample, respectively.64

39
Figure 1.34 Schematic of IR beam path in an ATR-IR.

Compared to the common transmission FTIR spectroscopy, ATR FTIR is advantageous

for probing the surface of the sample (rather than measuring the bulk) and particularly

powerful for in situ analysis of chemisorption and reactions at solid/liquid and solid/gas

interfaces.64 For thin MOF-nanofiber composites, transmission FTIR may still be useful for

characterization.39 However, ATR FTIR is generally suitable for chemical analysis and

qualitative studies of the MOF thin film growth for all MOF-functionalized fibers.

Neufeld et al. reported using ATR-IR to monitor the growth of HKUST-1 on cotton

fabrics by tracking the peak intensity at 728 cm-1 (δ(CH)), 1371 cm-1 (νs(COO)) and 1644

cm-1 (νa(COO)) (Figure 1.35). Increased peak intensity was observed at these wavenumbers,

indicating the additional MOF loadings onto the substrate during each cycle of deposition.

40
Figure 1.35 ATR-IR spectra for cotton fabric coated with different number of cycles of

HKUST-1. 32

1.3.2 UV-Vis Spectroscopy

UV-Vis absorption spectroscopy and diffuse reflectance UV-Vis spectroscopy can be

used to qualitatively and quantitatively (with other complementary analysis) monitor the

growth of MOF thin films on fibers. Especially for MOF thin films grown by layer-by-layer

method which are often too thin for XRD measurements, UV-Vis has been demonstrated as a

useful alternative approach to analyze the film growth.32,50 Meilikhov et al. used UV-Vis to

monitor the growth of HKUST-1 thin films deposited onto chemically functionalized

polyester fibers using layer-by-layer method (Figure 1.36). The absorbance at 704 nm

increases linearly as a function of the number of cycles after the second cycle, indicating the

linear growth of film thickness.50 Similarly, Neufeld et al. used diffuse reflectance UV-Vis

spectroscopy for characterizing HKUST-1 layer-by-layer synthesis on surface-functionalized

cotton fibers. The increase of absorbance at 704 nm was also found as more cycles were

deposited.32

41
Figure 1.36 (a) UV-Vis absorption spectra for different numbers of HKUST-1

layer-by-layer cycles on polyester fibers pretreated with polyvinylamine and bromoacetic

acid. (b) absorbance at 704 nm plotted as a function of cycle number.50

1.3.3 Analysis of MOF Mass Fraction

Analyzing the MOF mass fraction in the MOF-fiber composites is sometimes

challenging, when the actual mass change is very small to measure directly by weighing

methods and/or there is water adsorption on MOFs and hygroscopic fabric materials.

Meilikhov et al. used inductively coupled plasma optical emission spectroscopy

(ICP-OES) to measure the Cu concentration in the MOF-fiber composite after dissolving the

HKUST-1 layers. Using the known Cu concentration in the bulk MOF structure, MOF weight

percent can be calculated from the measured Cu concentration.50 Neufeld et al. described a

similar method using atomic absorption spectroscopy (AAS) to determine the mass fraction

of HKUST-1 in the MOF-fiber composites.32 MOF coated fabrics were sonicated in 3.5×

10-3 M HCl aqueous solution for 30 mins to fully dissolve the HKUST-1 coatings. The

42
copper concentration measured by AAS was used for the calculation of MOF mass fraction.

This method provides an accurate approach to quantify the amount of MOFs assuming all the

MOF coatings can be effectively removed. However, it is still necessary to measure the total

dry mass of MOF-coated fibers for calculating the weight percent of MOFs.

Our group developed a method to use BET surface area of MOF-coated fibers and that of

bulk MOFs to calculate the MOF mass fraction. Previously, we reported the following

equation to correlate the MOF mass fraction with the overall BET surface per unit mass of

MOF plus fibers:

𝑆𝐴𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 × 𝑚𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 −𝑆𝐴𝑓𝑖𝑏𝑒𝑟 × 𝑚𝑓𝑖𝑏𝑒𝑟


𝑆𝐴𝑀𝑂𝐹 ≈ (3)
𝑚𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 − 𝑚𝑓𝑖𝑏𝑒𝑟

where m is the mass and SA is the surface area for each component.23 With known surface

area for the uncoated fiber substrate, the MOF-coated fibers and the bulk MOF material, the

MOF mass fraction (ω) can be calculated using Equation X:

𝑆𝐴𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 = 𝑆𝐴𝑀𝑂𝐹 ∙ 𝜔 + 𝑆𝐴𝑓𝑖𝑏𝑒𝑟 ∙ (1 − 𝜔) (4)

𝑆𝐴𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 −𝑆𝐴𝑓𝑖𝑏𝑒𝑟
or 𝜔= (5)
𝑆𝐴𝑀𝑂𝐹 − 𝑆𝐴𝑓𝑖𝑏𝑒𝑟

We found that the MOF weight percentage calculated in this approach is similar to what

was measured by the ICP-OES method.30 This method provides a simple and straightforward

way to calculate the MOF weight percent when instruments for elemental analysis are not

available. One prerequisite is that both the MOF coatings on the fibers and the bulk MOF

powder are fully activated for BET measurements. It is also recommended to use the MOF

crystals collected from the same batch for MOF synthesis on fibers for BET analysis in order

43
to obtain a more accurate MOF mass fraction.

1.3.4 Permeation Test

The gas permeation property is important for MOF-fiber composites especially in

applications of garments, membranes and gas filters. The permeance (P/l)i (in unites of gas

permeation unit or GPU, 1 GPU = 1×10–6cm3(STP)/cm2∙s∙cmHg) is defined with the equation

below:

𝑃 𝑄𝑖
( )𝑖 = (3)
𝑙 ∆𝑃 ∙ 𝐴

where P is permeability in units of Barrer (1 Barrer = 1×10–10 cm3(STP)∙cm / cm2∙s∙cmHg), l

(cm) is membrane thickness, Qi (cm3/s) is permeation rate for gas species i at STP, ∆𝑃

(cmHg) is pressure drop across the membrane, A (cm2) is area of membrane.65

The ideal separation factor (or selectivity) αA/B can be calculated based on the

permeability or permeance of species A and B:65


𝑃
𝑃𝐴 ( 𝑙 )𝐴
αA/B = = 𝑃 (4)
𝑃𝐵 ( 𝑙 )𝐵

Armstrong et al. tested the gas permeation properties of their UiO-66/PVCi nanofiber

membrane using an apparatus similar to those typically used for polymer mixed matrix

membranes (Figure 1.37b).39 The nanofiber membrane was fitted with rubber o-rings in a

stainless steel permeation cell. The transmembrane pressure drop was controlled at 1 psi, and

the permeation rate was measured by a 100 mL bubble flow meter. The results (Figure 1.37a)

indicate that the permeation in the UiO-66/PVCi nanofiber membrane follows Knudsen flow

mechanism.39 Knudsen flow mechanism is given in Eqn. 5:

44
𝜀 1 1
𝐹𝐾 = ( ) ( ) ( ) 𝐷 (5)
𝜏 𝐿 𝑅𝑇

where FK is permeance, 𝜀 is porosity, 𝜏 is tortuosity, L is membrane thickness, R is ideal

gas constant, T is temperature, D is pore diffusivity.66 If the Knudsen diffusion is the

dominating mechanism for the mass transport, the pore diffusivity of Eqn. 5 can be replaced

with Knudsen diffusivity (DK): 66

2 8𝑅𝑇 1/2
𝐷𝐾 = ( ) 𝑟𝑃 (6)
3 𝜋𝑀𝑤

where Mw is the gas molecular weight, 𝑟𝑃 is the pore radius. Substitution of D in Eqn. 5

with DK gives Eqn. 7: 39

𝜀 𝑟𝑃 1 1
𝐹𝐾 = 1.06 ∙ ∙ ∙ ∙ (7)
𝜏 𝐿 √𝑅𝑇 √𝑀𝑤

Figure 1.37 (a) Permeance for UiO-66/PVCi nanofiber membrane for He, N2, Ar and SF6.

(b) a schematic of the apparatus for permeation test.39

45
1.3.5 Surface Roughness Analysis

Meilikhov et al. has shown using AFM to characterize the surface roughness of the MOF

thin films.50 AFM images were acquired by scanning a surface area of 10 × 10 μm2, and the

roughness profiles were reported for both untreated fiber surface and MOF-coated fiber

surface (Figure 1.38). The author pointed out that bending roughness curve is a result of the

cylindrical fiber shape. Consequently, it is still challenging to interpret the AFM data and

report the roughness values for the MOF thin films alone. Further research is needed in this

area to address these challenges and develop a protocol for analyzing roughness for MOF

thin films on fibers.

Figure 1.38 AFM images and roughness profiles of (a) untreated polyester fiber surface and

(b) chemically modified polyester fibers deposited with 40 cycles of HKUST-1

layer-by-layer growth.50

1.3.6 Adhesion Test

The adhesion of MOF particles or thin films to the fiber substrate is a critical property

46
that can affect the application and may cause particle shedding issues with weak attachment.

Several thoughtful ideas have been reported in literature including blowing test, tensile test

and so on.23,32

Zhao et al. reported a novel blowing test to quantify the attachment of MOF crystals to

the fiber substrate, as traditional bending and rubbing tests for textiles resulted negligible

amount of mass loss due to good MOF adhesion.23 For the blowing test, compressed air (~ 40

psi) was forced through a MOF-coated fiber mats for more than 4 min. The mass change was

recorded as a function of blowing time. The loosely attached MOF crystals are expected to be

knocked off by the compress air. For example, Figure 1.39 shows the mass change of a

HKUST-1 coated PBT fiber mats during the blowing tests. The remaining mass percentage

after 4 min was used to analyze the MOF adhesion, since the mass loss tends to stabilize after

4 min of compress air blowing.23

Figure 1.39 Mass change of ALD-coated PBT fiber mats with HKUST-1 thin films during

the compress air blowing test.23

47
Neufeld et al. used cyclic tensile test to analyze the adhesion property.32 The samples

were stretched with a tensile force of 2 N (maximum strain = ~7.1% (2 mm / 28 mm)) and
32
relaxed afterwards. After 5 cycles of tensile test, the materials were imaged using SEM.

While this measurement demonstrates that the mechanical robustness of the cotton substrate

was maintained, the actual stress on MOF particles is much smaller than that applied

macroscopically on the fabric and may not be enough to cause particle shedding. SEM

Images after the tensile tests cannot provide quantitative information about the MOF mass

loss either.

48
1.4 Applications

1.4.1 Capture of Hazardous Gases

Since MOFs have large surface area, porosity and high adsorption capacity for toxic

industrial chemicals (TICs),11,12,67 integrating MOFs onto fabrics enables new composite

structures for filters, masks and suits for TIC protection.23–25,60 Zhuang et al. printed

HKUST-1 on cotton textiles and reported the color change as an indication of TIC uptake.

After exposure to NH3, HCl and H2S, the original turquoise color of HKUST-1 coatings

changes to dark blue, yellow and brown, respectively (Figure 1.40). Quantitative analysis of

TIC adsorption capacity was performed on the HKUST-1 thin films printed on quartz crystal

microbalance (QCM). The printed HKUST-1 coating shows large uptake for several TICs

and CO2. Zhuang et al. also observed the deformation of HKUST-1 films after the adsorption

tests for NH3, HCl and H2S, as the XRD patterns show no MOF peaks.60

Figure 1.40 (a) Photographs of cotton fiber mats printed with HKUST-1 dot pattern (5

printing-drying cycles) before and after exposure to NH3, HCl and H2S. (b) Relative mass

change for the HKUST-1 thin film printed on QCM during NH3 adsorption/desorption tests.

(c) Adsorption capacities of the HKUST-1 films printed on QCM for different gases. 60

49
Our group have also looked into the applications of MOF-functionalized fiber mats for

TIC protection, and quantified the gas adsorption capacity for the MOF-coated fabrics

directly using micro-breakthrough analysis (Figure 1.41).23–25 Zhao et al. reported that the

dynamic loading of NH3 for HKUST-1 coated PP fibers (pretreated with ALD Al2O3) is 5.28

± 0.40 molNH /kg(MOF+fiber). The adsorption capacity for the MOF film calculated based the
3

MOF mass fraction is 6.9 ± 0.5 mol/kg (MOF), similar to that for bulk HKUST-1.23 This result

demonstrates that the functionality of bulk MOF crystals is fully maintained after integration

on fibers, and is very promising for TIC protection for environmental preservation and

industrial safety.

Figure 1.41 NH3 breakthrough curves for bare PP fibers (PP, green blank circle), PP fibers

with ALD Al2O3 coatings (PP/ALD, orange blank diamond), HKUST-1 grown on untreated

PP fibers (MOF-PP, blue circle), HKUST-1 grown on PP/ALD fiber mats (MOF-PP/ALD,

red diamond). 23

Zhao et al. also evaluated the TIC removal performance for the HKUST-1 thin films

grown on ALD-coated PP fibers via layer-by-layer process.24 Both breakthrough time and

50
NH3 dynamic loadings were found to increase as a function of the number of LbL cycles

(Figure 42). With 40 LbL cycles (~17 wt%), the MOF-coated PP/ALD fibers show an

ammonia dynamic loading of 1.37 molNH3/kgMOF+fiber (equivalent to 7.63 molNH3/kgMOF) and

a H2S dynamic loading of 1.49 molH2S/kgMOF+fiber (equivalent to 9.46 molH2S/kgMOF). The high

adsorption capacities indicate the excellent performance of these MOF-functionalized fibers

for TIC removal.24

Figure 1.42 (a) Ammonia breakthrough curves for LbL HKUST-1 thin films grown on

ALD-coated PP fibers and control sample without MOF coating. (b) Ammonia dynamic

loading for PP/ALD fiber mats with 0 ~ 40 cycles of LbL HKUST-1. Black square points

were calculated from the breakthrough data before saturation, and red circle points were

calculated based on the breakthrough curves with desorption parts. Error bar shows standard

deviation. 24

MOF-functionalized nanofibers have also been evaluated for TIC removal. The

HKUST-1 thin films grown on PAN nanofibers via rapid room-temperature HDS-to-MOF

conversion were tested and compared to the HKUST-1 coated PP microfibers prepared using

51
the same approach. The higher MOF mass loading on PAN nanofibers leads to significant

improvement of NH3 adsorption capacity (Figure 1.43). The high dynamic loadings of NH3

as well as H2S on these MOF-functionalized fiber mats suggest that these composites are

very promising for gas filtration and protective suites. 25

Figrue 1.43 Ammonia dynamic loading on HKUST-1 MOF coated PP microfibers and

PAN nanofibers.25

1.4.2 Membrane Separation

In addition to the MOF-incorporated polymer mixed matrix membranes (MMM), free

standing MOF membranes supported by MOF-fiber scaffolds and MOF thin films on hollow

fibers are another two types of composite structures for MOF-based membranes. Polymer

MMMs with MOF fillers have been reviewed previously,68,69 and are outside the scope of

MOF-fiber composites. So the latter two membrane structures will be mainly discussed here

for separation applications.

Brown et al. synthesized continuous ZIF-90 thin films on Torlon (a polyamide-imide)

hollow fiber and measured the single-gas permeation properties using a time-lag method.27

52
They found that the permeance through the membrane decreases with increased kinetic

diameter of gas molecules (Figure 1.44a), indicating that the mass transport within ZIF-90

pores but not through defects is the permeation mechanism. The selectivity for CO2/N2 (3.5)

and CO2/CH4 (1.5) is higher than the corresponding Knudsen selectivity (0.8 and 0.6

respectively). Brown et al. also found that the permeance for all gases increases with feed

pressure and decreases as a function of temperature (Figure 1.44b). They also investigated

the permeation of hydrocarbon molecules through the Torlon/ZIF-90 membranes by

pervaporation method, and found that the membranes are selective for n-hexane over

benzene and cyclohexane. 27

Figure 1.44 (a-c) Single gas permeances for Torlon/ZIF-90 hollow fiber membranes

measured at (a) 35 oC and 50 psia, (b) 70 oC with varying pressure, (c) 50 psia with varying
o
temperature. (d) Permeances of hydrocarbon molecules measured at 22 C by

pervaporation.27

53
Dai et al. investigated the separation performance of their Ultem/ZIF-8 hollow fibers

synthesized by co-extrusion spinning.28 It is very interesting that the ZIF-8 functionalized

Ultem hollow fiber shows higher CO2 permeance and selectivity towards CO2/N2 at various

test conditions (Table 1.2). The enhancement of selectivity was ascribed to the active ZIF-8

fillers in the sheath layer. Mixed gas permeation was also tested at multiple temperatures and

pressures for this hollow fiber composite. While the selectivity for CO2/N2 decreased slightly,

increased permeance of CO2 in the Ultem/ZIF-8 membrane was observed.28

Table 1.2 CO2 permeance and permselectivity for CO2/N2 for pure Ultem and Ultem/ZIF-8

hollow fiber membranes.28

1.4.3 Destruction of Chemical Warfare Agents

The applications of MOFs in heterogeneous catalysis have been widely investigated. The

catalytic effect of the metal nodes and organic linkers in the bulk MOF structures has been

demonstrated, while post-synthetic modification to introduce addition functionalities or

incorporate nanoparticles can further enhance the catalytic performance of MOFs.70,71 One

specific application of MOF catalyst is destruction of highly toxic compounds like chemical

warfare agents (CWA), and a few MOFs have been reported with high activity towards these

reactions.72–78 Integrating MOFs onto textiles enables new composite materials for protecting

54
first-responders, soldiers and the general public from CWA threats.29,30

López-Maya et al. investigated the degradation of CWA simulants using

silk@[UiO-66@LiOtBu] composite catalysts. MOF particles (50 wt%) sprayed onto the silk

fibers were further modified with LiOtBu. The half-lives of diisopropylfluorophosphate

(DIFP), dimethylmethylphosphonate (DMMP) and 2-chloroethylethylsulfide (CEES) are 20

min, 50 min and 8 min, respectively (Figure 1.45). Experimental results also confirmed the

heterogeneous nature of these composite catalysts.29

Figure 1.45 (a) SEM images of silk fibers coated with UiO-66@LiOtBu. (b-d) Catalytic

hydrolysis profiles of CWA simulants using silk@[UiO-66@LiOtBu] composite catalysts.29

Our group synthesized conformal Zr-based MOF thin films, including UiO-66,

UiO-66-NH2 and UiO-67, on nylon nanofibers, and applied them for CWA degradation.30

Hydrolylsis of DMNP (a CWA simulant) in an aqueous buffer solution containing

N-ethylmorpholine (pH = 10) was evaluated with UiO MOF powders and MOF-coated

55
nanofiber catalysts. Both PA-6@TiO2@UiO-66-NH2 and PA-6@TiO2@UiO-67 enable

ultra-fast reaction rates for DMNP degradation with half-lives less than 8 min (Figure 1.46).

The catalytic activity of the UiO thin films were found comparable to that of the bulk UiO

powders. Catalytic degradation of O-Pinacolyl Methylphosphonofluoridate (also known as

Soman or GD) was further investigated with PA-6@TiO2@UiO-66-NH2. The results in drop

test show GD conversion in 1 h is up to 46%. This is also the first demonstration for

destruction of real nerve agents using MOF-functionalized fibers.30

Figure 1.46 (a) Reaction schematic of DMNP catalytic hydrolysis using MOF-nanofiber

catalyst. (b) UV/Visible absorption spectra to monitor DMNP degradation. (c-e) DMNP

conversion during hydrolysis with different MOF powders and MOF-functionalized

nanofiber catalysts.30

56
1.4.4 Antibacterial Protection

Textiles with immobilized nanoparticles can be potentially applied to hygienic clothing

for protection from bacterial hazard. Abbasi et al. have investigated the antimicrobial activity

of HKUST-1 coated silk fibers.31 The inhibition zones in the Müller-Hinton agar were used to

evaluate the antibacterial performance (Figure 1.47). HKUST-1/Silk fibers synthesized via

sonication-enhanced layer-by-layer method (IV-cycle 4 and IV-cycle 10) exhibit distinct

zones of inhibition against E. coli (Gram-negative bacteria) and S. aureus (Gram-positive

bacteria). The increased MOF mass loading as a result of increased number of LbL cycles

was also found to promote the antibacterial activity. The leaching of Cu ions into the aqueous

medium was considered to be possible mechanism. 31

Figure 1.47 Photos of the petri dishes for antibacterial tests using HKUST-1 coated silk

fibers. The results were compared with untreated silk, gentamycin (GM), amoxicillin (AMX),

cefepime (FEP) and oxacillin (OX). Samples in Group I were prepared via LbL synthesis

without sonication, while Group IV were synthesized with ultrasound irradiation. 31

57
1.4.5 Catalytic NO Release

Nitric oxide (NO) has biomedical application due to its antiplatelet, antibacterial and

anti-inflammatory effects.32 Reynolds and coworkers have demonstrated HKUST-1 as

effective catalyst for producing NO from S-nitrosothiols,79 applied HKUST-1 coated cotton

fabrics for catalytic NO release (Figure 1.48). Cotton swatches with HKUST-1 coatings were

found to generate a large total amount of NO (7.1 ± 1.2 μmolNO / mgMOF) over a long period

of release time. The authors also found that the surface-immobilized HKUST-1 crystals

exhibit a shorter induction period compared to mixed matrix membrane with the same type

of MOF, indicating fast diffusion rate in the MOF-fiber structure. Diffuse reflectance UV-Vis

spectra and ICP-AES analysis for the solvated Cu2+ confirm the structural integrity of the

MOF-coated fabric after the catalytic NO release. This unique application is promising for

localized NO release for biomedical uses such as wound healing.32

Figure 1.48 (Top) Catalytic reaction to form NO from S-nitrosocysteamine. (Bottom) NO

release as function of time with different catalytic materials.32

58
1.5 Outlook

A wide variety of synthesis methods have been discussed and compared in this review.

Surface functionalization processes including chemical modification and ALD for metal

oxide nucleation layers are critical for solvothermal and layer-by-layer synthesis in order to

obtain highly conformal MOF thin films. MOF Conversion from ALD oxides and HDSs

opens doors for developing novel MOF thin films on fibers especially at low processing

temperature. Electospinning, co-extrusion, and spay-drying approaches can take advantage of

commercially available MOF powders. However, challenges are still remaining. Compared to

the thousands of MOF structures reported in literature so far, the number of MOFs that have

been explored for MOF thin films on fibers is still very limited. There is huge opportunity for

the future development in this field. To date, standard metrics have not been fully established.

In addition, many applications are yet to explore for MOF-functionalized fibers.

59
References

(1) James, S. L. Metal-organic frameworks. Chem. Soc. Rev. 2003, 32 (5), 276–288.

(2) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The Chemistry and
Applications of Metal-Organic Frameworks. Science 2013, 341 (6149), 1230444.

(3) Chui, S. S. A Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n.


Science 1999, 283 (5405), 1148–1150.

(4) Park, K. S.; Ni, Z.; Côté, A. P.; Choi, J. Y.; Huang, R.; Uribe-Romo, F. J.; Chae, H. K.;
O’Keeffe, M.; Yaghi, O. M. Exceptional chemical and thermal stability of zeolitic
imidazolate frameworks. Proc. Natl. Acad. Sci. 2006, 103 (27), 10186–10191.

(5) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud,
K. P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks
with Exceptional Stability. J. Am. Chem. Soc. 2008, 130 (42), 13850–13851.

(6) Farha, O. K.; Eryazici, I.; Jeong, N. C.; Hauser, B. G.; Wilmer, C. E.; Sarjeant, A. A.;
Snurr, R. Q.; Nguyen, S. T.; Yazaydın, A. Ö.; Hupp, J. T. Metal–Organic Framework
Materials with Ultrahigh Surface Areas: Is the Sky the Limit? J. Am. Chem. Soc. 2012,
134 (36), 15016–15021.

(7) Meek, S. T.; Greathouse, J. A.; Allendorf, M. D. Metal-Organic Frameworks: A


Rapidly Growing Class of Versatile Nanoporous Materials. Adv. Mater. 2011, 23 (2),
249–267.

(8) Wang, Z.; Cohen, S. M. Postsynthetic Covalent Modification of a Neutral


Metal−Organic Framework. J. Am. Chem. Soc. 2007, 129 (41), 12368–12369.

(9) Kim, M.; Cahill, J. F.; Fei, H.; Prather, K. A.; Cohen, S. M. Postsynthetic Ligand and
Cation Exchange in Robust Metal–Organic Frameworks. J. Am. Chem. Soc. 2012, 134
(43), 18082–18088.

(10) Brozek, C. K.; Dincă, M. Cation exchange at the secondary building units of
metal-organic frameworks. Chem. Soc. Rev. 2014, 43 (16), 5456–5467.

(11) Britt, D.; Tranchemontagne, D.; Yaghi, O. M. Metal-organic frameworks with high
capacity and selectivity for harmful gases. Proc. Natl. Acad. Sci. U. S. A. 2008, 105
(33), 11623–11627.

(12) Glover, T. G.; Peterson, G. W.; Schindler, B. J.; Britt, D.; Yaghi, O. MOF-74 building
unit has a direct impact on toxic gas adsorption. Chem. Eng. Sci. 2011, 66 (2), 163–

60
170.

(13) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective gas adsorption and separation in
metal-organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1477–1504.

(14) Rosi, N. L.; Eckert, J.; Eddaoudi, M.; Vodak, D. T.; Kim, J.; O’Keeffe, M.; Yaghi, O.
M. Hydrogen storage in microporous metal-organic frameworks. Science 2003, 300
(5622), 1127–1129.

(15) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O. M.
Systematic design of pore size and functionality in isoreticular MOFs and their
application in methane storage. Science 2002, 295 (5554), 469–472.

(16) Xie, Y.; Wang, T.-T.; Liu, X.-H.; Zou, K.; Deng, W.-Q. Capture and conversion of CO2
at ambient conditions by a conjugated microporous polymer. Nat. Commun. 2013, 4.

(17) Boehringer, B.; Fischer, R.; Lohe, M. R.; Rose, M.; Kaskel, S.; Kuesgens, P. MOF
Shaping and Immobilization; Farrusseng, D., Ed.; Wiley-V C H Verlag Gmbh:
Weinheim, 2011.

(18) Bradshaw, D.; Garai, A.; Huo, J. Metal-organic framework growth at functional
interfaces: thin films and composites for diverse applications. Chem. Soc. Rev. 2012,
41 (6), 2344–2381.

(19) Allendorf, M. D.; Schwartzberg, A.; Stavila, V.; Talin, A. A. A roadmap to


implementing metal-organic frameworks in electronic devices: challenges and critical
directions. Chem. Weinh. Bergstr. Ger. 2011, 17 (41), 11372–11388.

(20) Furukawa, S.; Reboul, J.; Diring, S.; Sumida, K.; Kitagawa, S. Structuring of metal–
organic frameworks at the mesoscopic/macroscopic scale. Chem. Soc. Rev. 2014, 43
(16), 5700–5734.

(21) Kuesgens, P.; Siegle, S.; Kaskel, S. Crystal Growth of the Metal-Organic Framework
Cu3(BTC)2 on the Surface of Pulp Fibers. Adv. Eng. Mater. 2009, 11 (1-2), 93–95.

(22) Zhuang, J.-L.; Ceglarek, D.; Pethuraj, S.; Terfort, A. Rapid Room-Temperature
Synthesis of Metal–Organic Framework HKUST-1 Crystals in Bulk and as Oriented
and Patterned Thin Films. Adv. Funct. Mater. 2011, 21 (8), 1442–1447.

(23) Zhao, J.; Losego, M. D.; Lemaire, P. C.; Williams, P. S.; Gong, B.; Atanasov, S. E.;
Blevins, T. M.; Oldham, C. J.; Walls, H. J.; Shepherd, S. D.; et al. Highly Adsorptive,
MOF-Functionalized Nonwoven Fiber Mats for Hazardous Gas Capture Enabled by
Atomic Layer Deposition. Adv. Mater. Interfaces 2014, 1 (4), 1400040.

61
(24) Zhao, J.; Gong, B.; Nunn, W. T.; Lemaire, P. C.; Stevens, E. C.; Sidi, F. I.; Williams, P.
S.; Oldham, C. J.; Walls, H. J.; Shepherd, S. D.; et al. Conformal and highly adsorptive
metal–organic framework thin films via layer-by-layer growth on ALD-coated fiber
mats. J. Mater. Chem. A 2015, 3 (4), 1458–1464.

(25) Zhao, J.; Nunn, W. T.; Lemaire, P. C.; Lin, Y.; Dickey, M. D.; Oldham, C. J.; Walls, H.
J.; Peterson, G. W.; Losego, M. D.; Parsons, G. N. Facile Conversion of Hydroxy
Double Salts to Metal-Organic Frameworks Using Metal Oxide Particles and Atomic
Layer Deposition Thin-Film Templates. J. Am. Chem. Soc. 2015, 137 (43), 13756–
13759.

(26) Wu, Y.; Li, F.; Liu, H.; Zhu, W.; Teng, M.; Jiang, Y.; Li, W.; Xu, D.; He, D.; Hannam,
P.; et al. Electrospun fibrous mats as skeletons to produce free-standing MOF
membranes. J. Mater. Chem. 2012, 22 (33), 16971–16978.

(27) Brown, A. J.; Johnson, J. R.; Lydon, M. E.; Koros, W. J.; Jones, C. W.; Nair, S.
Continuous Polycrystalline Zeolitic Imidazolate Framework-90 Membranes on
Polymeric Hollow Fibers. Angew. Chem. Int. Ed. 2012, 51 (42), 10615–10618.

(28) Dai, Y.; Johnson, J. R.; Karvan, O.; Sholl, D. S.; Koros, W. J. Ultem((R))/ZIF-8 mixed
matrix hollow fiber membranes for CO2/N-2 separations. J. Membr. Sci. 2012, 401,
76–82.

(29) López-Maya, E.; Montoro, C.; Rodríguez-Albelo, L. M.; Aznar Cervantes, S. D.;
Lozano-Pérez, A. A.; Cenís, J. L.; Barea, E.; Navarro, J. A. R. Textile/Metal–
Organic-Framework Composites as Self-Detoxifying Filters for Chemical-Warfare
Agents. Angew. Chem. Int. Ed. 2015, 54 (23), 6790–6794.

(30) Zhao, J.; Lee, D. T.; Woodward, I. R.; Oldham, C. J.; Walls, H. J.; Yaga, R. W.; Hall,
M. G.; Peterson, G. W.; Parsons, G. N. Conformal Metal-Organic Framework Thin
Films on Nanofibers for Ultra-Fast Degradation of Chemical Warfare Agents. 2016, in
Preparation..

(31) Abbasi, A. R.; Akhbari, K.; Morsali, A. Dense coating of surface mounted CuBTC
Metal-Organic Framework nanostructures on silk fibers, prepared by layer-by-layer
method under ultrasound irradiation with antibacterial activity. Ultrason. Sonochem.
2012, 19 (4), 846–852
.
(32) Neufeld, M. J.; Harding, J. L.; Reynolds, M. M. Immobilization of Metal–Organic
Framework Copper(II) Benzene-1,3,5-tricarboxylate (CuBTC) onto Cotton Fabric as a
Nitric Oxide Release Catalyst. ACS Appl. Mater. Interfaces 2015, 7 (48), 26742–
26750.

62
(33) Zacher, D.; Baunemann, A.; Hermes, S.; Fischer, R. A. Deposition of microcrystalline
[Cu3(btc)2] and [Zn2(bdc)2(dabco)] at alumina and silica surfaces modified with
patterned self assembled organic monolayers: evidence of surface selective and
oriented growth. J. Mater. Chem. 2007, 17 (27), 2785–2792.

(34) Ullah, S.; Shariff, A. M.; Bustam, M. A.; Elkhalifah, A. E. I.; Murshid, G.; Riaz, N.;
Shimekit, B. Effect of Modified MIL-53 with Multi-Wall Carbon nanotubes and nano
Fibers on CO2 Adsorption. In Process and Advanced Materials Engineering; Ahmed, I.,
Ed.; Trans Tech Publications Ltd: Stafa-Zurich, 2014; Vol. 625, pp 870–873.

(35) Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.; Henry, M.; Bataille, T.;
Férey, G. A Rationale for the Large Breathing of the Porous Aluminum Terephthalate
(MIL-53) Upon Hydration. Chem. Eur. J. 2004, 10 (6), 1373–1382.

(36) Patil, D. V.; Rallapalli, P. B. S.; Dangi, G. P.; Tayade, R. J.; Somani, R. S.; Bajaj, H. C.
MIL-53(Al): An Efficient Adsorbent for the Removal of Nitrobenzene from Aqueous
Solutions. Ind. Eng. Chem. Res. 2011, 50 (18), 10516–10524.

(37) Pinto, M. da S.; Augusto Sierra-Avila, C.; Hinestroza, J. P. In situ synthesis of a


Cu-BTC metal-organic framework (MOF 199) onto cellulosic fibrous substrates:
cotton. Cellulose 2012, 19 (5), 1771–1779.

(38) Wu, Y.; Li, F.; Liu, H.; Zhu, W.; Teng, M.; Jiang, Y.; Li, W.; Xu, D.; He, D.; Hannam,
P.; et al. Electrospun fibrous mats as skeletons to produce free-standing MOF
membranes. J. Mater. Chem. 2012, 22 (33), 16971–16978.

(39) Armstrong, M. R.; Arredondo, K. Y. Y.; Liu, C.-Y.; Stevens, J. E.; Mayhob, A.; Shan,
B.; Senthilnathan, S.; Balzer, C. J.; Mu, B. UiO-66 MOF and Poly(vinyl cinnamate)
Nanofiber Composite Membranes Synthesized by a Facile Three-Stage Process. Ind.
Eng. Chem. Res. 2015, 54 (49), 12386–12392.

(40) Stavila, V.; Volponi, J.; Katzenmeyer, A. M.; Dixon, M. C.; Allendorf, M. D. Kinetics
and mechanism of metal-organic framework thin film growth: systematic investigation
of HKUST-1 deposition on QCM electrodes. Chem. Sci. 2012, 3 (5), 1531–1540.

(41) Lemaire, P. C.; Zhao, J.; Williams, P. S.; Walls, H. J.; Shepherd, S. D.; Losego, M. D.;
Peterson, G. W.; Parsons, G. N. CuBTC MOF Nucleation Mechanisms on Metal Oxide
Powders and Thin Films formed by Atomic Layer Deposition. ACS Appl. Mater.
Interfaces 2016.

(42) Parsons, G. N.; George, S. M.; Knez, M. Progress and future directions for atomic
layer deposition and ALD-based chemistry. Mrs Bull. 2011, 36 (11), 865–871.

63
(43) Peng, Q.; Sun, X.-Y.; Spagnola, J. C.; Hyde, G. K.; Spontak, R. J.; Parsons, G. N.
Atomic layer deposition on electrospun polymer fibers as a direct route to Al2O3
microtubes with precise wall thickness control. Nano Lett. 2007, 7 (3), 719–722.

(44) Hyde, G. K.; Scarel, G.; Spagnola, J. C.; Peng, Q.; Lee, K.; Gong, B.; Roberts, K. G.;
Roth, K. M.; Hanson, C. A.; Devine, C. K.; et al. Atomic Layer Deposition and Abrupt
Wetting Transitions on Nonwoven Polypropylene and Woven Cotton Fabrics.
Langmuir 2010, 26 (4), 2550–2558.

(45) Gong, B.; Spagnola, J. C.; Arvidson, S. A.; Khan, S. A.; Parsons, G. N. Directed
inorganic modification of bi-component polymer fibers by selective vapor reaction and
atomic layer deposition. Polymer 2012, 53 (21), 4631–4636.

(46) Kalanyan, B.; Losego, M. D.; Oldham, C. J.; Parsons, G. N. Low-Temperature Atomic
Layer Deposition of Tungsten using Tungsten Hexafluoride and Highly-diluted Silane
in Argon. Chem. Vap. Depos. 2013, 19 (4-6), 161–166.

(47) Liu, X.; Zhang, K.-Q. Silk Fiber — Molecular Formation Mechanism, Structure-
Property Relationship and Advanced Applications. In Oligomerization of Chemical
and Biological Compounds; Lesieur, C., Ed.; InTech, 2014.

(48) Khanjani, S.; Morsali, A. Layer by layer growth of nano porous lead(II) coordination
polymer on natural silk fibers and its application in removal and recovery of iodide.
Crystengcomm 2012, 14 (23), 8137–8142.

(49) Khanjani, S.; Morsali, A. Ultrasound-promoted coating of MOF-5 on silk fiber and
study of adsorptive removal and recovery of hazardous anionic dye “congo red.”
Ultrason. Sonochem. 2014, 21 (4), 1424–1429.

(50) Meilikhov, M.; Yusenko, K.; Schollmeyer, E.; Mayer, C.; Buschmann, H.-J.; Fischer, R.
A. Stepwise deposition of metal organic frameworks on flexible synthetic polymer
surfaces. Dalton Trans. 2011, 40 (18), 4838–4841.

(51) Zhan, W.; Kuang, Q.; Zhou, J.; Kong, X.; Xie, Z.; Zheng, L. Semiconductor@Metal–
Organic Framework Core–Shell Heterostructures: A Case of ZnO@ZIF-8 Nanorods
with Selective Photoelectrochemical Response. J. Am. Chem. Soc. 2013, 135 (5),
1926–1933.

(52) Khaletskaya, K.; Turner, S.; Tu, M.; Wannapaiboon, S.; Schneemann, A.; Meyer, R.;
Ludwig, A.; Van Tendeloo, G.; Fischer, R. A. Self-Directed Localization of ZIF-8 Thin
Film Formation by Conversion of ZnO Nanolayers. Adv. Funct. Mater. 2014, 24 (30),
4804–4811.

64
(53) Bechelany, M.; Drobek, M.; Vallicari, C.; Chaaya, A. A.; Julbe, A.; Miele, P. Highly
crystalline MOF-based materials grown on electrospun nanofibers. Nanoscale 2015, 7
(13), 5794–5802.

(54) Meyn, M.; Beneke, K.; Lagaly, G. Anion-exchange reactions of hydroxy double salts.
Inorg. Chem. 1993, 32 (7), 1209–1215.

(55) Rose, M.; Boehringer, B.; Jolly, M.; Fischer, R.; Kaskel, S. MOF Processing by
Electrospinning for Functional Textiles. Adv. Eng. Mater. 2011, 13 (4), 356–360.

(56) Ostermann, R.; Cravillon, J.; Weidmann, C.; Wiebcke, M.; Smarsly, B. M.
Metal-organic framework nanofibers via electrospinning. Chem. Commun. 2011, 47
(1), 442–444.

(57) Ren, J.; Musyoka, N. M.; Annamalai, P.; Langmi, H. W.; North, B. C.; Mathe, M.
Electrospun MOF nanofibers as hydrogen storage media. Int. J. Hydrog. Energy 2015,
40 (30), 9382–9387.

(58) Husain, S. Mixed Matrix Dual Layer Hollow Fiber Membranes For Natural Gas
Separation.

(59) Ameloot, R.; Gobechiya, E.; Uji-i, H.; Martens, J. A.; Hofkens, J.; Alaerts, L.; Sels, B.
F.; De Vos, D. E. Direct Patterning of Oriented Metal-Organic Framework Crystals via
Control over Crystallization Kinetics in Clear Precursor Solutions. Adv. Mater. 2010,
22 (24), 2685 – 2688.

(60) Zhuang, J.-L.; Ar, D.; Yu, X.-J.; Liu, J.-X.; Terfort, A. Patterned Deposition of
Metal-Organic Frameworks onto Plastic, Paper, and Textile Substrates by Inkjet
Printing of a Precursor Solution. Adv. Mater. 2013, 25 (33), 4631–4635.

(61) Harrick, N. J. Study of Physics and Chemistry of Surfaces from Frustrated Total
Internal Reflections. Phys. Rev. Lett. 1960, 4 (5), 224–226.

(62) Fahrenfort, J. Attenuated total reflection. A new principle for the production of useful
infra-red reflection spectra of organic compounds. Spectrochim. Acta 1961, 17, 698–
709.

(63) Hind, A. R.; Bhargava, S. K.; McKinnon, A. At the solid/liquid interface: FTIR/ATR
— the tool of choice. Adv. Colloid Interface Sci. 2001, 93 (1–3), 91–114.

(64) Andanson, J.-M.; Baiker, A. Exploring catalytic solid/liquid interfaces by in situ


attenuated total reflection infrared spectroscopy. Chem. Soc. Rev. 2010, 39 (12), 4571–
4584.

65
(65) Ismail, A. F.; Rana, D.; Matsuura, T.; Foley, H. C. Examples of CMSM Preparation,
Characterization and Testing. In Carbon-based Membranes for Separation Processes;
Springer New York, 2011; pp 93–108.

(66) Ismail, A. F.; Rana, D.; Matsuura, T.; Foley, H. C. Other Carbon-Based Membranes. In
Carbon-based Membranes for Separation Processes; Springer New York, 2011; pp
145–246.

(67) DeCoste, J. B.; Peterson, G. W. Metal-Organic Frameworks for Air Purification of


Toxic Chemicals. Chem. Rev. 2014, 114 (11), 5695–5727.

(68) Qiu, S.; Xue, M.; Zhu, G. Metal–organic framework membranes: from synthesis to
separation application. Chem. Soc. Rev. 2014, 43 (16), 6116–6140.

(69) Seoane, B.; Coronas, J.; Gascon, I.; Etxeberria Benavides, M.; Karvan, O.; Caro, J.;
Kapteijn, F.; Gascon, J. Metal-organic framework based mixed matrix membranes: a
solution for highly efficient CO2 capture? Chem. Soc. Rev. 2015, 44 (8), 2421–2454.

(70) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T.
Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38 (5), 1450–
1459.

(71) Metal Organic Frameworks as Heterogeneous Catalysts; Llabrés i Xamena, F.,


Gascon, J., Eds.; RSC Catalysis Series; Royal Society of Chemistry: Cambridge, 2013.

(72) Montoro, C.; Linares, F.; Quartapelle Procopio, E.; Senkovska, I.; Kaskel, S.; Galli, S.;
Masciocchi, N.; Barea, E.; Navarro, J. A. R. Capture of Nerve Agents and Mustard Gas
Analogues by Hydrophobic Robust MOF-5 Type Metal-Organic Frameworks. J. Am.
Chem. Soc. 2011, 133 (31), 11888–11891.

(73) Padial, N. M.; Quartapelle Procopio, E.; Montoro, C.; Lopez, E.; Enrique Oltra, J.;
Colombo, V.; Maspero, A.; Masciocchi, N.; Galli, S.; Senkovska, I.; et al. Highly
Hydrophobic Isoreticular Porous Metal-Organic Frameworks for the Capture of
Harmful Volatile Organic Compounds. Angew. Chem.-Int. Ed. 2013, 52 (32), 8290–
8294.

(74) Katz, M. J.; Mondloch, J. E.; Totten, R. K.; Park, J. K.; Nguyen, S. T.; Farha, O. K.;
Hupp, J. T. Simple and Compelling Biomimetic Metal-Organic Framework Catalyst
for the Degradation of Nerve Agent Simulants. Angew. Chem.-Int. Ed. 2014, 53 (2),
497–501.

(75) Mondloch, J. E.; Katz, M. J.; Isley Iii, W. C.; Ghosh, P.; Liao, P.; Bury, W.; Wagner, G.

66
W.; Hall, M. G.; DeCoste, J. B.; Peterson, G. W.; et al. Destruction of chemical
warfare agents using metal–organic frameworks. Nat. Mater. 2015, 14 (5), 512–516.

(76) Moon, S.-Y.; Liu, Y.; Hupp, J. T.; Farha, O. K. Instantaneous Hydrolysis of
Nerve-Agent Simulants with a Six-Connected Zirconium-Based Metal-Organic
Framework. Angew. Chem.-Int. Ed. 2015, 54 (23), 6795–6799.

(77) Peterson, G. W.; Moon, S.-Y.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Hupp, J. T.;
Farha, O. K. Tailoring the Pore Size and Functionality of UiO-Type Metal-Organic
Frameworks for Optimal Nerve Agent Destruction. Inorg. Chem. 2015, 54 (20), 9684–
9686.

(78) Liu, Y.; Moon, S.-Y.; Hupp, J. T.; Farha, O. K. Dual-Function Metal–Organic
Framework as a Versatile Catalyst for Detoxifying Chemical Warfare Agent Simulants.
ACS Nano 2015, 9 (12), 12358–12364.

(79) Harding, J. L.; Reynolds, M. M. Metal Organic Frameworks as Nitric Oxide Catalysts.
J. Am. Chem. Soc. 2012, 134 (7), 3330–3333.

67
CHAPTER 2 is a reprint of a manuscript published Advance Materials Interfaces.

CHAPTER 2. Highly Adsorptive, MOF-Functionalized Nonwoven

Fiber Mats for Hazardous Gas Capture Enabled by Atomic Layer

Deposition

Junjie Zhao†, Mark D. Losego†, Paul C. Lemaire†, Philip S. Williams†, Bo Gong†,

Sarah E. Atanasov†, Trent M. Blevins†, Christopher J. Oldham†, Howard J. Walls‡,

Sarah D. Shepherd‡, Matthew A. Browe#, Gregory W. Peterson#

and Gregory N. Parsons†*


Department of Chemical and Biomolecular Engineering, North Carolina State University,

911 Partners Way, Raleigh, NC 27695, U.S.A (E-mail: gnp@ncsu.edu)



RTI International, 3040 East Cornwallis Road, Research Triangle Park, NC 27709, U.S.A.
#
Edgewood Chemical Biological Center, 5183 Blackhawk Road, Aberdeen Proving Ground,

MD 21010, U.S.A.

68
Abstract

While metal-organic frameworks (MOFs) show great potential for gas adsorption and

storage, their powder form limits deployment opportunities. Integration of MOFs on

polymeric fibrous scaffolds will enable new applications in gas adsorption, membrane

separation, catalysis, and toxic gas sensing. Here, we demonstrate a new synthesis route for

growing MOFs on fibrous materials that achieves high MOF loadings, large surface areas

and high adsorptive capacities. We find that a nanoscale coating of Al2O3 formed by atomic

layer deposition (ALD) on the surface of nonwoven fiber mats facilitates nucleation of MOFs

on the fibers throughout the mat. Functionality of MOFs is fully maintained after integration,

and MOF crystals are well attached to the fibers. Breakthrough tests for HKUST-1 MOFs

[Cu3(BTC)2] on ALD-coated polypropylene fibers reveal NH3 dynamic loadings up to

5.93±0.20 mol/kg(MOF+fiber). Most importantly, this synthetic approach is generally applicable

to a wide range of polymer fibers (e.g., PP, PET, cotton) and MOFs (e.g., HKUST-1,

MOF-74, and UiO-66).

2.1 Introduction

Hazardous gas adsorption has important implications for human health, industrial safety,

and environmental protection. Traditional adsorbents including zeolites, activated carbons,

and silica gels can abate hazardous gas emission, but there is growing need for materials with

higher capacity and better selectivity, while simultaneously providing multifunctional

performance (i.e. capture and alert capability).1,2 Metal organic frameworks (MOFs),

composed of metal cluster secondary building units and organic bridging ligands,3 exhibit

69
high surface area, good thermal stability, and have significant synthetic versatility, enabling

structures with tunable pore sizes and adjustable internal functionality.4 MOF synthesis

usually follows wet solvothermal batch methods, producing powders that require further

manipulation and handling, and may be difficult to implement for some applications.5

Integration of MOFs on polymeric fibrous scaffolds could simplify handling, regeneration

and deployment. It could also broaden practical use for filtration, separations, catalysis,

sensing and other applications.6

Several methods have been used to integrate MOF crystals on polymer fiber matrices,

including encapsulation in electrospun fibers7,8 and immobilization on fibers via solvothermal

synthesis9,10, layer-by-layer method6,11 or microwave irradiation12. Unfortunately, these

methods generally lead to small MOF loading fractions, poor MOF crystal quality, and low

Brunauer–Emmett–Teller (BET) surface areas—limiting applicability. For example,

Kuesgens et al. grew HKUST-1 crystals on pulp fibers using direct solvothermal synthesis

and found the surface area to be limited to 314 m2/g.9 Wu et al. encapsulated HKUST-1

MOFs in electrospun fibers, and also only achieved a value of 311 m2/g before repeating

secondary growth.8 These values are at least 2 times lower than pure HKUST-1 powders

grown solvothermally (692~1460m2/g).13-17 To date, no prior report has assessed the

hazardous gas adsorption capacity for MOF-functionalized fibers.

In this work, we introduce a novel synthesis route for achieving high-quality MOF

growth on polymer fibers. We use atomic layer deposition (ALD) to form a nanoscale-thick

conformal coating on polymer fibers for nucleation of solvothermally grown MOF crystals.

This approach is first demonstrated with HKUST-1 on ALD-Al2O3-coated polypropylene

70
fibers. MOF nucleation on fibers is compared between untreated and ALD-coated substrates,

while the quality of MOFs is benchmarked to pure powders. Hazardous gas (ammonia)

adsorption capacity was characterized via breakthrough tests. We also show the general

applicability of this technique to a wide range of polymer fibers (e.g., PP, PET, and cotton)

and MOFs (e.g., HKUST-1, MOF-74, and UiO-66).

2.2 Experimental Methods

Polymeric Fibrous Materials

Non-woven polypropylene (PP), polybutylene terephthalate (PBT) and cotton fiber mats

were acquired from Nonwovens Cooperative Research Center (NCRC), North Carolina State

University. Non-woven PP fiber mats are 0.30mm thick, with fiber diameter ranging from

0.6μm to 9.0μm. Non-woven PBT fiber mats are 0.76mm thick, with fiber diameter ranging

from 8.2μm to 18μm. Non-woven cotton fiber mats are 1.25mm thick, with fiber diameter

ranging from 13μm to 16μm.

Atomic layer deposition (ALD) of Al2O3

ALD Al2O3 was deposited onto polymeric fibrous substrates using a homemade hot-wall

viscous-flow vacuum reactor. Deposition pressure was kept at ~1Torr, and the temperature

was 60℃. In a typical ALD Al2O3 cycle, tri-methyl aluminum (TMA) was first dosed to the

reaction chamber for 1s, followed with 30s of N2 purge between doses. After TMA dose and

N2 purge, deionized water was dosed with another 30s of N2 purge. 200 cycles of ALD Al2O3

were selected as the standard coating thickness, while 50, 100 and 500 cycles were also used

71
to study the effect of ALD thickness on MOF growth.

Atomic layer deposition (ALD) of ZnO and TiO2

ALD ZnO and TiO2 were conducted in the same hot-wall viscous-flow vacuum reactor

as that used for ALD Al2O3. The deposition pressures were both ~1Torr, and the temperatures

were kept at 100℃. In an ALD ZnO cycle, the substrate was exposed to 2s of di-ethyl zinc

(DEZ) and 2s of deionized water alternatively, with 60s of N2 purge between dose steps. In

an ALD TiO2 cycle, precursors (TiCl4 and H2O) were dosed alternatively to the reaction

chamber for 1s, with 40s of N2 purge between dose steps. 200 cycles of ALD ZnO or TiO2

were deposited onto the fiber mats as the promoter of MOF growth.

Solvothermal growth of HKUST-1 MOF on fibrous materials

HKUST-1 MOFs were synthesized solvothermally using copper nitrate trihydrate and

1,3,5-benzene-tricarboxylic (BTC) acid as precursors. 0.42g BTC was dissolved in 12mL of

ethanol, and 0.87g copper nitrate trihydrate was dissolved in 12mL of deionized water. Fiber

mats were placed in a Teflon-lined pressure vessel then covered by the liquid reagents which

readily soaked into the substrates. The pressure vessel was sealed and crystal growth

proceeded at 120°C for 20 hours. After MOF growth, the fiber mats were dried in vacuum

over at 120°C for 12 hours. Dry samples were weighed, and mass increases were calculated

based on the dry weights of the starting substrate materials and the products.

72
Materials characterization

Scanning electron microscope (SEM) images of MOF-fiber materials were taken with an

FEI Phenom® bench-top SEM. Samples were sputter-coated with 5~10nm of Au-Pd before

imaging. X-ray diffraction (XRD) was measured with a Rigaku SmartLab X-ray diffraction

tool (Cu Kα X-ray source). BET surface areas of the MOF-fiber materials were measured

using a Quantachrome Autosorb-1C surface area and pore size analyzer. Samples were

vacuum dried at 120°C for 12h before the measurement of nitrogen adsorption isotherm.

BET surface areas were calculated based on isotherm within the P/Po range of 0.05~0.31.

Ammonia breakthrough test

Rapid micro-breakthrough analysis equipment (shown in Fig. S6) was used to NH3

adsorption of our MOF-fiber materials. NH3 was added into a ballast and pressurized, and

subsequently mixed with moisturized air stream to reach the challenge concentration of 1000

mg/m3 with 50% relative humidity (RH). The challenge gas mixture then passed through an

adsorbent column containing ~40mg of MOF-fiber material. The adsorbent column was kept

at 20°C in a themostatted water bath. The effluent stream was analyzed with a continuously

measuring gas chromatograph (HP5890 Series II) equipped with a photoionization detector.18

Ammonia breakthrough data analysis

Ammonia breakthrough time is defined as the time for the effluent signal to reach 5% of the

feed concentration. Dynamic loadings (DL in units of mol/kg) were calculated based on the

following equations.18

73
𝐶𝑓𝑒𝑒𝑑 𝐹𝑓𝑒𝑒𝑑 ∙𝑡𝑡𝑜𝑡𝑎𝑙
𝑁𝑓𝑒𝑒𝑑 = (1)
𝑀𝑤
𝑡𝑡𝑜𝑡𝑎𝑙 𝐶𝑜𝑢𝑡 𝐹𝑓𝑒𝑒𝑑
𝑁𝑜𝑢𝑡 = ∫0 𝑑𝑡 (2)
𝑀𝑤
𝑁𝑓𝑒𝑒𝑑 −𝑁𝑜𝑢𝑡
𝐷𝐿 = (3)
𝑚𝑎𝑑𝑠

Nfeed is defined as total moles of adsorbate in the feed, while Nout is defined as the moles of

adsorbate detected in the effluent stream. Both Nfeed and Nout have units of mol. Cfeed and Cout

are the concentrations of adsorbate in the feed and in the downstream, and have unites of

g/m3. Ffeed is defined as the feed flow rate in units of m3/min. Time (t) has units of min. Mw

is the molecular weight of the adsorbate in g/mol. mads is defined as the mass of adsorbent

and has units of kg.

2.3 Results and Discussion

Figure 2.1a-c illustrates the process scheme for the case of HKUST-1 grown on

ALD-Al2O3-coated polypropylene fibers. Polypropylene (PP) fibers (Figure 2.1a) in a

nonwoven mat were conformally coated with ALD Al2O3 (200 cycles at 60°C), creating a

core/shell “PP/ALD” structure with hydroxyl surface termination (Figure 2.1b).19 This ALD

coating can change wettability of fibers for thorough permeation of solvothermal solvents

(Figure S2.1),17 and improve reactivity of the fiber surfaces for MOF nucleation.20

Depending on the the deposition conditions, ALD coatings can increase surface roughness,21

which may also promote MOF nucleation. HKUST-1 [Cu3(BTC)2] was grown onto

ALD-coated fibers solvothermally using copper nitrate trihydrate and

1,3,5-benzene-tricarboxylic (BTC) acid as precursors in a water/ethanol (50/50 vol%)

74
solution (Figure 2.1c). Fiber mats were placed in a Teflon-lined pressure vessel and

immersed in the liquid reagents which readily wetted the coated fiber substrates. The vessel

was sealed, and MOF growth proceeded under the optimal condition at 120°C for 20 hours

(optimization shown in SI). Figure 2.1d shows HKUST-1 cyrstals grown densely and

conformally on an ALD-coated PP fibers. This resulting MOF-functionalized fiber mat is

referred to as“MOF-PP/ALD”.

Figure 2.1 (a-c) Schematic illustration of the synthesis route. (a) Polymer fiber substrate. (b)

Al2O3-coated polymer fiber via atomic layer deposition (ALD). The cross section in the

dashed square illustrates the conformal coating of ALD Al2O3 with hydroxyl surface

termination. (c) MOFs integated on Al2O3-coated polymer fiber using solvothermal MOF

synthesis. (d) SEM image of HKUST-1 MOF crystals grown on MOF-PP/ALD.

75
In Figure 2.2 we compare the growth of HKUST-1 on untreated and ALD-coated PP

fiber mats. MOFs grown on the untreated PP fibers (referred to as “MOF-PP”) show

macroscopic non-uniformity and appear to have nucleated homogeneously (i.e. in solution)

throughout the fiber matrix (Figure 2.2a). When coated with ALD Al2O3, the fiber mats show

uniform MOF coverage indicative of heterogeneous nucleation on the fiber surface. X-ray

diffraction data in Figure 2.2c confirms HKUST-1 crystal formation for both structures.

Figure 2.2d shows the percent mass increase after MOF growth on untreated and

ALD-coated PP fiber substrates. Mass increase percentage (m/m0) is 336±38% for

PP/ALD, and 228±83% for untreated PP respectively. The results show that the ALD

treatment enhances both the amount and uniformity of MOF growth. Normalizing the MOF

mass loading per unit substrate external surface area, we find m = 14.78 mg/cm2 on

PP/ALD to be 54% larger than on untreated PP (m = 9.61 mg/cm2). Mass gain is also more

reproducible on the ALD-Al2O3-coated fiber mats, as indicated by the boxplot’s interquartile

range (IQR) = 69% for PP/ALD vs 145% for untreated PP. Figure 2.2e shows results from

N2 isotherms and Brunauer–Emmett–Teller (BET) surface area analysis. The MOF-PP/ALD

fiber mats show an average surface area (SAMOF+fiber) of 695 ±76 m2/g(MOF+fiber), which is ~60%

higher than that of MOF-PP fibers (434 ± 198 m2/g(MOF+fiber)). This is consistent with the

higher MOF mass gain on PP/ALD substrates. The surface area of MOF-PP/ALD fiber mat

is also >2× larger than prior reports for MOFs on fibers.8,9

76
Figure 2.2 Comparison of HKUST-1 MOF grown on untreated and ALD-coated

polypropylene fiber mats. (a) SEM image of HKUST-1 MOF grown on untreated

polypropylene fiber mats (MOF-PP). Insert image is a photograph of MOF-PP (scale bar

represents 1cm). b) SEM image of HKUST-1 MOF grown on ALD-Al2O3-coated

polypropylene fiber mats (MOF-PP/ALD). Insert image is a photograph of MOF-PP/ALD

(scale bar represents 1cm). c) X-ray diffraction of MOF-PP and MOF-PP/ALD. d) Mass

increase percentage based on substrate dry weight. Interquartile range and average value

were calculated based on 16 MOF-PP samples and 12 MOF-PP/ALD samples. e) Brunauer–

Emmett–Teller (BET) surface area of MOF-fiber materials and calculated BET surface area

for MOF part (error bars represent standard deviation). BET surface area for PP fiber

substrates is 1.3~1.5 m2/g. The values were measured for uncoated and ALD coated fibers,

and the values were indistinguishable.

77
We compare the quality of the MOFs nucleated on fibers to that of pure powders. The

surface area of the MOF component in the MOF-fiber material (SAMOF) is given by:

𝑆𝐴𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 × 𝑚𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 −𝑆𝐴𝑓𝑖𝑏𝑒𝑟 × 𝑚𝑓𝑖𝑏𝑒𝑟


𝑆𝐴𝑀𝑂𝐹 ≈ (4)
𝑚𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 − 𝑚𝑓𝑖𝑏𝑒𝑟

where SA is the surface area and m is the mass of each component. From this analysis (Figure

2.2e), we find SAMOF = ~716 ± 305 m2/g(MOF) for MOF-PP, while SAMOF = 924 ± 94 m2/g(MOF)

for MOF-PP/ALD. HKUST-1 powders similarly produced in our lab show SAMOF = 1066

m2/g. It is consistent with as-synthesized high quality HKUST-1 crystals,13-17 and essentially

similar to the MOFs we grew on PP/ALD fibers. Although post-synthesis activation maybe

further increases the MOF surface area, vacuum drying at high temperature is not suitable for

the polymer fiber substrate. Lower surface area on untreated PP is likely the result of

nonstoichiometric crystal phases or other unwanted side products.

To assess the quality of interaction between MOFs and the fiber substrates, we

performed compressed air blowing tests. We forced compressed air at ~40 psi for 4 min

through a standard-size MOF-fiber mat (10 cm2) and measured the mass change versus

air-flow time. Results are plotted in Figure S2.5. Generally, the mass loss stabilizes within a

few minutes, and the overall loss of MOF-PP/ALD is limited to ~15% of the starting mass.

Many samples show >85% MOF retention. Subsequent laboratory handling of MOF-fiber

mats after forced air testing resulted in no noticeable MOF detachment. In addition, we tested

our MOF-fiber mats in bending and rubbing tests. The amount of particles coming off during

these tests was too small to quatifiy. Rubbing test may also lead to fiber abrasion. So

compressed air blowing test was chosen to be our standard methods for MOF attachement

78
testing.

PP PP/ALD
1
0.8
0.6

C/Co
0.4
0.2
MOF-PP MOF-PP/ALD
0
0 5000 10000 15000 20000
Weighted Time (min/g)

Figure 2.3 NH3 breakthrough curves for untreated polypropylene fiber mats (PP, ○),

ALD-Al2O3-coated polypropylene fiber mats (PP/ALD, ◇ ), HKUST-1 MOF grown on

untreated polypropylene fiber mats (MOF-PP, ● ), HKUST-1 MOF grown on

ALD-Al2O3-coated polypropylene fiber mats (MOF-PP/ALD, ◆).

To evaluated MOF-functionalized fibers for hazardous gas adsorption, we used ammonia

breakthrough analysis (Figure S2.6).18 Results in Figure 3 show downstream NH3

concentration changes as a function of time after the feed flow starts. Ammonia breakthrough

is defined as the time at which the downstream signal reaches 5% of the feed concentration.

Dynamic loading (calculation shown in SI) indicates the total ammonia sorption capacity for

NH3 at saturation. The decay in NH3 signal after feed flow is terminated shows NH3

desorption rate, indicating the retention and therefore strength of sorption. Figure 2.3 shows

that the breakthrough time for MOF-PP/ALD is up to 58% longer than MOF-PP. Control PP

mats without MOFs exhibit immediate breakthrough. From the breakthrough results, the

ammonia dynamic loadings are 5.28±0.40 molNH3/kg(MOF+fiber) for MOF-PP/ALD and 4.57±

79
0.27 molNH3/kg(MOF+fiber) for MOFs on untreated fibers. Calculation results show the NH3

dynamic loading for the MOF component in MOF-PP/ALD fiber mats is 6.9 ± 0.5

mol/kg(MOF), similar to pure MOF powders synthesized in our lab (7.23 mol/kg) and

consistent with the reported values.[16] The similar values for the NH3 dynamic loadings

indicate the adsorption takes place in the bulk MOF crystals on the fibers.

In addition to ammonia, we also evaluated the performance MOF-functionalized fibers

for H2S removal. Compared with H2S adsorption by MOF powders,22 the H2S dynamic

loading for our MOF-fiber mats show good H2S loading capacity, demonstrating that the

MOF-fiber constructs have functional capacity to adsorb hazardous gases beyond NH3.

We investigated the effect of ALD coating thickness, and find that MOF crystal size

distribution and extent of growth into the fiber mat can be systematically changed with ALD

thickness. Figure S2.7 shows top-view and cross-sectional SEM images of MOFs grown on

PP fiber mats with 0, 50, 100, 200 and 500 cycles of Al2O3 ALD (roughly correspond to 0 nm,

6 nm, 12 nm, 24 nm, and 60 nm of Al2O3 respectively). MOF crystal size distributions

analyzed from both top-view and cross-sectional SEM are plotted in Figure 2.4a,b and Figure

S2.8. On the top layer of fibers (Figure 2.4a), MOF crystal sizes decrease from 24±8μm to 9

±3μm with narrowing distributions, as ALD coating thickness increases from 0 to 500

cycles. Similar trend is also found in the cross-sections of the MOF-fiber mats with different

ALD thickness (Figure 2.4b). The amount of MOFs grown into the fiber mat rises as the

number of ALD cycles increases to ~100 cycles (Figure S2.7), leading to the similar trends of

MOF loading and BET surface area shown in Figure 2.4c,d. The NH3 breakthrough tests,

80
summarized in Figure 2.4e,f show that NH3 adsorption capacity correlates closely with BET

surface area, regardless of the different crystal size distribution.

Figure 2.4 MOF-fiber mats with different ALD coating thicknesses. a) MOF crystal size

distribution on the top surface. b) MOF crystal size distribution on the cross-section. Crystal

size distributions were analyzed based on 50 measured crystal sizes on each corresponding

SEM image. c) Mass increase of the fiber mats after MOF integration. d) BET surface area of

the MOF-functionalized fiber mats and the surface area of MOF component in the fiber mats.

e) NH3 dynamic loading on MOF-functionalized fiber mats. f) NH3 breakthrough time on

MOF-functionalized fiber mats.

In addition to ALD coating thickness, we also tested the effect of ALD chemistry. ALD

81
ZnO and TiO2 coated PP fiber mats were functionalized with HKUST-1 MOFs, and the

results shown in Figure 2.5a,b are compared with MOFs on ALD Al2O3 coated PP. We find

that MOF crystals on ZnO are small, whereas those on TiO2 are generally larger than the

MOFs on Al2O3. This indicates that MOF crystal morphology can be adjusted using different

ALD coatings. We hypothesize that the wettability, surface roughness and isoelectric points

(IEP) of different ALD coatings may all affect the nucleation of MOF crystals. The

mechanism of MOF nucleation on different ALD materials is currently under investigation.

The resulting mass increase, BET surface area, and NH3 dynamic loading of these MOF-fiber

materials are shown in Table 2.1. The NH3 dynamic loading of MOF-PP/ALD(ZnO)

(5.93±0.20 mol/kg(MOF+fiber)) represents the best adsorptive capacity of MOF-fiber materials

so far. MOF-PP/ALD(TiO2) fiber mats exhibit less adsorptive capacity than

MOF-PP/ALD(Al2O3) and MOF-PP/ALD(ZnO). This may be explained with the low surface

area of the MOF component in this material (~670 m2/g(MOF)), which may suggest the

formation of impurities or the necessity of careful post-synthesis activation.

Table 2.1 Characterization of MOF-functionalized fiber mats: fiber mass gain (mg/cm2) after

MOF integration, BET surface area (m2/g) of MOF-fiber materials and NH3 dynamic loading

(mol/kg) calculated from breakthrough curves.

Fiber Mass Gain Total BET Surface NH3 Dynamic


MOF-Fiber Composites
(mg/cm2) Area (m2/g) Loading (mol/kg)
MOF-PP/ALD(ZnO) 25.21 764±11 5.93±0.20
MOF-PP/ALD(TiO2) 12.85 509±151 2.75±0.02
MOF-PP/ALD(Al2O3) 14.78 694±76 5.28±0.40
MOF-PBT/ALD(Al2O3) 17.17 485 2.52±0.23
MOF-Cotton/ALD(Al2O3) 13.59 121 1.70±1.14

82
Figure 2.5 HKUST-1 MOF grown on different ALD coatings and different polymer fibers.

a) SEM image of MOF on ALD-ZnO-coated polypropylene fiber mats

(MOF-PP/ALD(ZnO)). b) SEM image of MOF on ALD-TiO2-coated PP fiber mats

(MOF-PP/ALD(TiO2)). c) SEM image of MOF on ALD-Al2O3- coated PP fiber mats

(MOF-PP/ALD(Al2O3)). d) SEM image of MOF on ALD-Al2O3-coated polybutylene

terephthalate (PBT) fiber mats (MOF-PBT/ALD(Al2O3)). e) SEM image of MOF on

ALD-Al2O3-coated cotton fiber mats (MOF-Cotton/ALD(Al2O3)).

Finally, we demonstrate our approach can readily extend to various fibers and MOFs.

Figure 2.5d,e and Table 2.1 show HKUST-1 growth on ALD Al2O3 coated polybutylene

terephthalate (PBT) and cotton fiber mats. Cotton cellulose has hydroxyl groups that promote

ALD Al2O3 growth,23 but our results show HKUST-1 mass uptake on ALD treated cotton is

less than on other ALD-modified fibers. Initial MOF loading on cotton/ALD fiber mats is

83
comparable with that on PP/ALD and PBT/ALD fiber mats (Table 2.1), but MOFs grown on

cotton/ALD fibers tend to show poor adhesion. We hypothesize that the ALD layer enhances

nucleation, but nucleii attachement is less robust, possibly resulting from the smooth surface

after ALD treatment. This is supported by Figures 2.5(e) and S2.9 showing MOF growth in

the crevaces on the cotton fibers. The details of MOF nucleation on different surfaces are

currently under investigation. Preliminary examples of Zn-MOF-74, Mg-MOF-74 and

UiO-66 on ALD-coated nonwoven polymer fibrous substrates are given in Figure S2.10.

Integration of these MOFs on nonwoven fiber mats could potentially broaden the practical

use for filtration, separations, catalysis, sensing and other applications.

2.4 Conclusion

In conclusion, we demonstrate a new synthesis route using ALD coatings to facilitate

MOF nucleation on fibrous materials. Al2O3 ALD coatings improve the MOF macroscopic

uniformity and its coverage on fiber surfaces. MOF-PP/ALD fiber mats exhibit higher BET

surface areas and higher NH3 dynamic loading capacities than MOFs grown on untreated PP

fiber mats. These results significantly advance capability for functional MOFs on fiber

substrates. Variations in ALD surface modification are found to affect MOF crystallite size

and distribution but ultimately have minimal effect on ammonia removal performance. The

process scheme presented here provides a robust platform technology for integrating various

MOFs onto polymeric fibrous substrates for the use in human protection, industrial safety,

and environmental preservation.

84
References

(1) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Chem. Soc. Rev. 2009, 38, 1477–1504.

(2) Lu, Z.-Z.; Zhang, R.; Li, Y.-Z.; Guo, Z.-J.; Zheng, H.-G. J. Am. Chem. Soc. 2011, 133,
4172–4174.

(3) James, S. L. Chem. Soc. Rev. 2003, 32, 276–288.

(4) Meek, S. T., Greathouse, J. A.; Allendorf, M. D. Adv. Mater. 2011, 23, 249–267.

(5) Bradshaw, D.; Garai, A.; Huo, J. Chem. Soc. Rev. 2012, 41, 2344–2381.

(6) Meilikhov, M.; Yusenko, K.; Schollmeyer, E.; Mayer, C.; Buschmann, H.-J.; Fischer,
R. A. Dalton Trans. 2011, 40, 4838–4841.

(7) Ostermann, R.; Cravillon, J.; Weidmann, C.; Wiebcke, M.; Smarsly, B. M. Chem.
Commun. 2011, 47, 442–444.

(8) Wu, Y.; Li, F.; Liu, H.; Zhu, W.; Teng, M.; Jiang, Y.; Li, W.; Xu, D.; He, D.; Hannam,
P.; Li, G. J. Mater. Chem. 2012, 22, 16971–16978.

(9) Kuesgens, P.; Siegle, S.; Kaskel, S. Adv. Eng. Mater. 2009, 11, 93–95.

(10) Pinto, M. da S.; Augusto Sierra-Avila, C.; Hinestroza, J. P. Cellulose 2012, 19, 1771–
1779.

(11) Abbasi, A. R.; Akhbari, K.; Morsali, A. Ultrason. Sonochem. 2012, 19, 846–852.

(12) Centrone, A.; Yang, Y.; Speakman, S.; Bromberg, L.; Rutledge, G. C.; Hatton, T. A. J.
Am. Chem. Soc. 2010, 132, 15687–15691.

(13) Chui, S. S. Science 1999, 283, 1148–1150.

(14) Basu, S.; Maes, M.; Cano-Odena, A.; Alaerts, L.; De Vos, D. E.; Vankelecom, I. F. J. J.
Membr. Sci. 2009, 344, 190–198.

(15) Britt, D.; Tranchemontagne, D.; Yaghi, O. M. Proc. Natl. Acad. Sci. U. S. A. 2008, 105,
11623–11627.

(16) Peterson, G. W.; Wagner, G. W.; Balboa, A.; Mahle, J.; Sewell, T.; Karwacki, C.J. J.
Phys. Chem. C 2009, 113, 13906–13917.

85
(17) Levasseur, B.; Petit, C.; Bandosz, T. J. ACS Appl. Mater. Interfaces 2010, 2,
3606-3613.

(18) Glover, T. G.; Peterson, G. W.; Schindler, B. J.; Britt, D.; Yaghi, O. Chem. Eng. Sci.
2011, 66, 163-170.

(19) Hyde, G. K.; Scarel, G.; Spagnola, J. C.; Peng, Q.; Lee, K.; Gong, B.; Roberts, K. G.;
Roth, K. M.; Hanson, C. A.; Devine, C. K.; Stewart, S. M.; Hojo, D.; Na, J.-S.; Jur, J.
S.; Parsons, G. N. Langmuir 2010, 26, 2550–2558.

(20) Zacher, D.; Baunemann, A.; Hermes, S.; Fischer, R. A. J. Mater. Chem. 2007, 17,
2785–2792.

(21) Jur, J. S.; Spagnola, J. C.; Lee, K.; Gong, B.; Peng, Q.; Parsons, G. N. Langmuir 2010,
26, 8239–8244.

(22) Petit, C.; Mendoza, B.; Bandosz, T. J. ChemPhysChem 2010, 11, 3678-3684.

(23) Hyde, G. K.; Park, K. J.; Stewart, S. M.; Hinestroza, J. P.; Parsons, G. N. Langmuir
2007, 23, 9844-9849.

86
Supporting Information

160

Water Contact Angle (degree)


140

120

100

80

60

40

20

0
0 100 200 300 400 500
Number of ALD Cycles

Figure S2.1 Surface wetting ability of nonwoven polypropylene fiber mats with different

number of ALD Al2O3 cycles (0~500 cycles). 50 ALD cycles generates a very thin (~6nm)

Al2O3 coating, but this film was “thick” enough to change the wettability of PP substrates

from originally hydrophobic to hydrophilic. Contact angles of PP in ethanol are 0 degree

regardless of ALD cycles.

87
Synthesis Optimization

The temperature of MOF growth on fibers was optimized based on PP fiber mats with 200

cycles of ALD Al2O3. We studied the temperatures ranging from 100℃ to 140℃. MOF

crystals start to form on fibers when T≥110℃, and mass gain increases almost linearly with

temperature from 39% to 336% within the range of 100℃~120℃ (Shown in Fig. S2.2a).

The amount of MOF crystals grown on fabric substrates decreases when T≥120℃. This is

reasonable because at increased temperature the reaction rate is raised and may be higher

than the diffusion rate of precursor molecules into the fiber networks. Crystals quickly

forming on the outer surfaces of the substrate may block the path for molecular diffusion,

giving rise to low internal MOF growth. BET surface area of fibrous substrate plus MOFs

shown in Fig. S2.2b is consistent with the mass gains.

Figure S2.2 (a) Mass increase of PP/ALD fiber mats due to MOF growth at different

temperatures (100~140℃). (b) total BET surface area and calculated MOF surface area of

HKUST-1 functionalized PP/ALD fiber mats prepared at different temperatures

(100~140℃).

88
50μm 50μm

Figure S2.3 (a) Optical image of MOF-PP/ALD fiber mats prepared at different

temperatures (100~140℃). (b) SEM image of the side products found in MOF-PP/ALD fiber

mats synthesized at 100℃. (c) SEM image of the flower-like CuO poly-crystals found in

MOF-PP/ALD fiber mats synthesized at 140℃.

XRD (shown in Fig. S2.4) and SEM (Fig. S2.3) results show that one-dimensional crystals

dominantly formed at 100℃ are not MOF-199. These low dimensional crystals are likely to

be the reported chain polymer [Cu(BTC-H)-(H2O)3]n.1

89
140℃

130℃
Intensity (a.u.)

120℃

110℃

100℃

PP substrate

5 10 15 20 25 30 35 40 45
2θ (deg)

Figure S2.4 X-ray diffraction data of MOF-PP/ALD fiber mats prepared at different

temperatures (100~140ºC) and PP substrate.

Characteristic peaks of PP are not readily observed after MOF growth at 110 ~130 °C,

because the densely packed MOF crystals on fibers could diminish the X-ray diffraction

signal from the substrate. In a control experiment, we characterized the crystallinity of PP

fiber substrates after soaking them in ethanol/water (50:50 vol%) for 20 hours. XRD results

show the crystallinity of these PP fibers does not change in the solvothermal conditions. This

is also consistent with PP peaks being present after MOF growth at 100 ºC and 140 ºC where

MOF loading is relatively small (Figure S2.2a and S2.3a).

90
1.00

m/mo 0.90

0.80

0.70

0.60
0 1 2 3 4 5 6 7
Time (min)

Figure S2.5 Mass change of MOF funbctionalized fiber mats during the adhesion tests

(data shown are based on MOF-PBT/ALD fiber mats). The mass loss stabilizes after <4 min

of air flow, so empirically we take m/mo(t=4min) to evaluate the attachment of MOF crystals to

the fiber substrate. Mass change of MOF-PP/ALD during the adhesion tests is similar to

MOF-PBT/ALD.

91
Figure S2.6 Rapid micro-breakthrough analysis equipment.

92
Figure S2.7 Representative SEM images showing the top surface and the cross-section of

MOF-functionalized fibrous materials based on polypropylene fibers with different ALD

coating thickness.

93
Figure S2.8 MOF crystal size distribution in PP fiber mats with different ALD thickness.

94
Figure S2.9 SEM image of MOF crystals grown inside ALD-Al2O3 coated cotton fibers.

10μm

Figure S2.10 (a) SEM image of UiO-66 MOF grown on PP/ALD fiber mats. (b) SEM

image of Zn-MOF-74 grown on PP/ALD fiber mats. (c) SEM image of Mg-MOF-74 grown

on PP/ALD fiber mats.

(1) Pech, R.; Pickardt, J. Acta Crystallogr. C 1988, 44, 992.

95
CHAPTER 3 is a reprint of a manuscript published Journal of Materials Chemistry A.

CHAPTER 3. Conformal and highly adsorptive metal–organic

framework thin films via layer-by-layer growth on ALD-coated

fiber mats

Junjie Zhao†, Bo Gong†, William T. Nunn†, Paul C. Lemaire†, Eric C. Stevens†,

Fahim I. Sidi†, Philip S. Williams†, Christopher J. Oldham†, Howard J. Walls‡,

Sarah D. Shepherd‡, Matthew A. Browe‡, Gregory W. Peterson‡, Mark D. Losego§,

and Gregory N. Parsons†,*


Department of Chemical and Biomolecular Engineering, North Carolina State University,

911 Partners Way, Raleigh, NC 27695, U.S.A (E-mail: gnp@ncsu.edu)



RTI International, 3040 East Cornwallis Road, Research Triangle Park, NC 27709, U.S.A.
#
Edgewood Chemical Biological Center, 5183 Blackhawk Road, Aberdeen Proving Ground,

MD 21010, U.S.A.
§
School of Materials Science and Engineering, Georgia Institute of Technology, 771 Ferst Dr.

NW, Atlanta, GA 30332, U.S.A.

96
Abstract

Integration of metal-organic frameworks (MOFs) on textiles shows promise for enabling

facile deployment and expanding MOF applications. While MOFs deposited on flat

substrates can show relatively smooth surface texture, most previous reports of MOFs

integrated on fibers show poor conformality with many individual crystal domains. Here we

report a new low-temperature (<70°C) method to deposit uniform and smooth MOF thin

films on fiber surfaces using an energy enhanced layer-by-layer (LbL) method with an ALD

Al2O3 nucleation layer. Cross-sectional TEM images show a well-defined core@shell

structure of the MOF-functionalized fiber, and SEM shows a flat MOF surface texture. We

analyze the thickness and mass increase data of LbL HKUST-1 MOF thin films on

ALD-coated polypropylene fibers and find the growth rate to be 288~290 ng∙cm-2 per LbL

cycle. Unlike planar LbL MOF embodiments where adsorption capacities are difficult to

quantify, the large volume quantity on a typical fiber mat enables accurate surface area

measurement of these unique MOF morphologies. After 40 LbL cycles the MOFs on fibers

exhibit N2 adsorption BET surface areas of up to 93.6 m2/gMOF+fiber (~535 m2/gMOF) and

breakthrough test results reveal high dynamic loadings for NH3 (1.37 molNH3/kgMOF+fiber) and

H2S (1.49 molH2S/kgMOF+fiber). This synthesis route is applicable to many polymer fibers, and

the fiber@ALD@MOF structure is promising for gas filtration, membrane separation,

catalysis, chemical sensing and other applications.

97
3.1 Introduction

Metal-organic frameworks (MOFs) are crystalline organic/inorganic materials consisting

of nanoscale metal ions or cluster building blocks coordinated by organic linkers. MOFs

exhibit large surface area and high porosity,1 and many allow for post-synthetic chemical

modification.2–4 Utilizing these characteristics, MOFs have been applied to gas adsorption

and storage,5–8 separations,9–11 and catalysis.12–16 However, since MOFs are generally

synthesized in the form of insoluble powders, methods for depositing MOF thin films are in

high demand for applications such as adsorptive gas filters, smart membranes, chemical

sensors and catalytic coatings.17,18 Integration of MOFs into application-oriented

configurations could also enable new device fabrication and simplify their deployment,

pushing forward the commercialization of MOFs.

The layer-by-layer (LbL) method (or so-called “liquid phase epitaxy”) generates thin

MOF coatings of defined thickness with good homogeneity.19,20 This synthetic approach is

advantageous for controlling MOF structural interpenetration,21 and has been used to build

chemical sensors22 and photo-switches for molecule release23. While most studies of the LbL

method are based on planar substrate surfaces, including SAM surfaces on Au substrates,24–27

silicon20,22 and alumina substrates20, little is known about the LbL process on fibers.

In most examples, even relatively modest MOF loading requires substrates selected

specifically from carboxylic-group-consisting polymers (polyester28 and silk fibers29),

presumably to help promote MOF nucleation. For example, polyester fibers can be enriched

with -COOH surface groups using polyvinylamine and bromoacetic acid sequentially to

improve MOF growth.28 MOFs formed on silk via LbL method appear as isolated crystals

98
with a rough surface texture.29 Until recently,30 very few reports quantify BET surface areas

or gas adsorption capacities of MOFs on fibers, and no reports are available that analyze LbL

MOFs on fibers.

Our recent work30 introduced atomic layer deposition (ALD) as a MOF nucleation layer

to dramatically improve solvothermal MOF growth on various polymer fibrous materials.

Based on self-limiting reactions, ALD enables thin film coatings on complex 3D surfaces

with a control of thickness at the sub-nanometer scale.31 With abundant hydroxyl groups,

ALD coatings improve the wettability of hydrophobic polymers,32 and also provide

anchoring sites for the metal-containing units in MOFs.20 In this work, we show that ALD

layers on fibers also promote LbL MOF growth, yielding very uniform and smooth MOF thin

films on flexible polymer fibers.

Scheme 3.1 Schematic of the synthesis route. Polymer fiber substrates were coated with 50

cycles of ALD-Al2O3, forming a core@shell structure of “Fiber@Al2O3”. HKUST-1 MOF

thin film was grown onto ALD-coated polymer fibers via layer-by-layer synthesis method.

99
3.2 Experimental Methods

Nonwoven fiber mat materials

Nonwoven polypropylene (PP), polyethylene terephthalate (PET) and cotton fiber mats

were used as received from Nonwovens Cooperative Research Center (NCRC), North

Carolina State University. Fiber diameters of PP, PET, and cotton fiber mats are 0.6μm ~

9.0μm, 33μm ~ 35μm and 13μm ~ 16μm respectively.

Atomic layer deposition (ALD) of Al2O3

Nonwoven fiber mats were coated with Al2O3 by ALD using a homemade hot-wall

viscous-flow vacuum reactor (Figure S3.1). Deposition temperature was kept at 60℃, and

pressure was ~1Torr. A typical ALD Al2O3 cycle started with trimethyl aluminum (TMA, 98%

STREM Chemicals, Inc.) dose for 1s and subsequent N2 purge for 30s. After TMA dose and

N2 purge, deionized water was dosed to the chamber for 1s, followed with 60s of N2 purge.

50 cycles of ALD Al2O3 were selected as the standard coating thickness for layer-by-layer

MOF growth.

Layer-by-layer (LbL) growth of HKUST-1 MOF on fibrous materials

1mM 1,3,5-benzenetricarboxylic (BTC, 98%, Acros Organics) acid and 1mM copper

acetate monohydrate (99%, Sigma Aldrich) were dissolved separately in two vessels with

150 mL ethanol. Precursor solutions and rinsing solvent (ethanol) were placed in a sonicated

water bath during the LbL process. In a typical LbL cycle, non-woven fiber substrates were

dipped in BTC solution for 5 minutes, followed with ethanol rinse for 1 minute. The

100
substrates were then transferred to Cu(OAc)2 solution for 5 minutes and subsequently rinsed

in ethanol for 1 minute. During each transfer between vessels, the samples were dried in air

for ~10 s. Precursor solutions were refreshed every 10 cycles to avoid a temperature effect

when the sonication bath became heated. We observed heated sonication (>70°C) led to poor

quality of MOF thin films. The rinse solution was refreshed every 4 cycles to maintain

cleanliness. A good rinse process is critical for fully removing any unreacted species and/or

unattached nuclei and maintaining controlled LbL growth.

Materials characterization

Scanning electron microscopy (SEM) and energy dispersive X-ray analysis (EDX) were

performed on a JEOL JSM 6010 SEM. The synthesized fiber@ALD@MOF material

structures were sputter-coated with a thin layer of Au-Pd (5~10nm) before SEM imaging.

Microtomed fiber mats were imaged in cross section using a JEOL 2010F transmission

electron microscope (TEM). Rigaku SmartLab X-ray diffraction (XRD) tool (Cu Kα X-ray

source) was used for crystalline phase analysis. Fourier transform infrared spectrometer

(FTIR, Thermo Scientific Nicolet 6700) was used to monitor LbL MOF growth on ALD-

Al2O3 coated silicon wafers. Brunauer–Emmett–Teller (BET) surface area was measured on a

Quantachrome Autosorb-1C surface area and pore size analyzer. Fiber@ALD@MOF

samples were dried in vacuum (~1 × 10−5 𝑇𝑜𝑟𝑟)at room temperature for 12h before BET

measurement. An eleven-point nitrogen adsorption isotherm was measured for MOF-coated

fiber mats at 77 K within P/Po range of 0.05~0.30.

101
Breakthrough test for NH3 and H2S

A custom-built rapid, micro-breakthrough system (shown in Scheme 3.2) was used for

characterizing NH3 and H2S adsorption of our fiber@ALD@MOF materials. Challenge gas

(NH3 or H2S) injected into a ballast was pressurized and subsequently mixed with a

moisturized air stream to achieve the target concentration of 1000 mg/m3 with 50% relative

humidity (RH). The challenge gas mixture then flowed through an adsorbent column loaded

with fiber@ALD@MOF material (~40 mg). The temperature of the adsorbent column was

maintained at 20°C in a water bath. The downstream concentration of the challenge gas was

detected with a continuously measuring gas chromatograph (HP5890 Series II) equipped

with a photoionization detector for NH3 (or a flame photometric detector for H2S).33

Breakthrough data analysis

Dynamic loadings (DL in units of mol/kg) of NH3 or H2S on MOF-coated fibers were

calculated from the breakthrough curves using the following equations.33

𝐶𝑓𝑒𝑒𝑑 𝐹𝑓𝑒𝑒𝑑 ∙ 𝑡𝑡𝑜𝑡𝑎𝑙


𝑁𝑓𝑒𝑒𝑑 = (1)
𝑀𝑤
𝑡𝑡𝑜𝑡𝑎𝑙 𝐶
𝑜𝑢𝑡 𝐹𝑓𝑒𝑒𝑑
𝑁𝑜𝑢𝑡 = ∫ 𝑑𝑡 (2)
0 𝑀𝑤
𝑁𝑓𝑒𝑒𝑑 − 𝑁𝑜𝑢𝑡
𝐷𝐿 = (3)
𝑚𝑎𝑑𝑠

Nfeed is defined as the total moles of challenge gas flowing through the adsorbent, while Nout

is the total moles of target gas detected in the effluent stream. Cfeed and Cout (in g/m3) are the

concentrations of challenge gas in the feed and the downstream respectively. Feed flow rate

(Ffeed) has units of m3/min, and time (t) has units of min. Mw (g/mol) is the molecular weight

102
of the challenge gas, and mads (kg) is the adsorbent mass.

Scheme 3.2 Schematic of the rapid micro-breakthrough analysis system for NH3 and H2S

breakthrough tests.

3.3 Results and Discussion

Conformal LbL MOF Thin Films Enabled by ALD-Al2O3 Nucleation Layer

Figure 3.1 displays SEM and TEM images showing the microscopic morphology of

MOF thin films grown on ALD-Al2O3 coated polypropylene (PP) fibers. 50 cycles of ALD

Al2O3 was deposited on PP fibers at 60℃, forming a Fiber@ALD core@shell structure

(PP@ALD). Subsequently, 40 cycles of LbL HKUST-1 MOF was synthesized on top of the

ALD coating, adding an extra shell layer to PP@ALD (shown in Scheme 3.1). We refer to

these MOF-functionalized materials as PP@ALD@LbL40. Figures 3.1a and 3.1b show good

uniformity and complete fiber coverage of the LbL MOF thin film. During solvothermal

MOF film growth on ALD-coated fibers, MOF crystals may nucleate homogeneously within

the fiber mat in the voids between fibers.30 However, the LbL method shows only MOFs

directly deposited on the fibers. Sufficient ethanol rinsing in sonication bath between dipping

103
steps in LbL process can remove unreacted precursor molecules and/or unattached nuclei

from the fiber mesh, ensuring MOF formation happens only on the reactive surface of the

fibers. A high magnification SEM image shown in Figure 3.1b suggests the average MOF

crystal size is ≤1 μm. In comparison, MOF crystals grown on ALD-coated PP fibers via

solvothermal synthesis are typically >5 μm.30 This indicates the LbL method is more

advantageous in controlling the surface roughness of the MOF coating.

Figure 3.1 (a,b,e) SEM images for ALD-Al2O3-coated PP fibers with 40 cycles of LbL

HKUST-1 MOF (PP@ALD@LbL40). (c,d) Cross-sectional TEM images for

PP@ALD@LbL40 showing the core@shell structure. (f-i) Energy Dispersive X-ray analysis

for PP@ALD@LbL40 showing the presence of carbon (f) from the polypropylene and the

HKUST-1 MOF, oxygen (g) from the ALD Al2O3 and the HKUST-1 MOF, aluminum (h)

from the ALD coating and copper (i) from the MOF.

Figures 3.1c-d show cross-sectional TEM images of the PP@ALD@LbL40 core@shell

structures. The well-defined thin layer (~10 nm thick) sandwiched between the fiber substrate

104
and the thick coating corresponds to the ALD-Al2O3 thin film. The ALD growth rate (~2

Å/cycle) and the smoothness of the coating deposited at 60℃ are consistent with previous

analysis of low temperature ALD on these fiber materials.34 With abundant hydroxyl

termination, this ALD layer is expected to facilitate MOF nucleation.30 On top of the ALD

coating, a layer with less uniform TEM contrast corresponds to the LbL HKUST-1 MOF thin

film. This MOF layer exhibits good conformality, with an average thickness of 117 nm. The

higher resolution TEM image in Figure 1d shows that the LbL MOF thin film is porous and

consists of nanocrystals with dimensions of about 5 nm to 8 nm. Similar nanoparticles were

also observed and reported by Wöll and co-workers for their work on LbL HKUST-1 coating

on magnetic nanoparticles.35 These nanoparticles are likely CuO produced from the

HKUST-1 during electron beam irradiation in the TEM.

The energy dispersive X-ray analysis (EDX) images for PP@ALD@LbL40 are shown in

Figures 3.1(f-i). The results show a uniform aluminum signal from the fibers, indicating the

ALD Al2O3 coating is conformal and is fully maintained after MOF integration. Copper is

also detected uniformly on the fibers, further confirming complete and uniform MOF

coverage. Carbon is expected from the polypropylene and from the MOF organic linker

(1,3,5-benzenetricarboxylic acid). The ALD thin films and the MOF coatings are both likely

to contribute to the oxygen signal detected by EDX.

To examine MOF nucleation mechanisms, we performed 20 MOF LbL cycles on PP

fibers with and without Al2O3 ALD pretreatment. SEM images of these coated fibers are

compared in Figure 2. PP fibers without an Al2O3 coating (Fig. 3.2a) exhibit non-uniform and

patchy MOF growth, whereas MOFs grown PP@ALD fibers (Fig. 3.2b) are uniform and

105
smooth. The improved nucleation and growth on the ALD-treated substrate is consistent with

that reported for solvothermal MOF growth on polymer fibers.30 Virgin PP fibers are devoid

of reactive functional groups and, therefore, are relatively inert to MOF nucleation.

Consequently, MOF nucleation likely occurs on random surface defects. Once the nucleation

seed is attached to the untreated fiber, the MOF will continue to grow preferentially at this

site leading to a patchy coverage.

(a) (b)

2μm 2μm

Figure 3.2 SEM images of (a) 20 cycles of LbL HKUST-1 MOF on untreated PP fibers and

(b) 20 cycles of LbL HKUST-1 MOF on ALD-Al2O3-coated PP fibers (PP@ALD). MOF

thin film grown on untreated PP substrate exhibits poor uniformity, while MOF coating on

PP@ALD shows good conformality with complete coverage on fibers.

Growth Rate of LbL HKUST-1 MOF on ALD Al2O3 Coatings

Figure 3.3a is an optical image showing the color change due to the growth of LbL

MOFs on ALD-coated PP fiber mats. The ALD Al2O3 (50 cycles) does not change the visual

color of the PP fiber substrate. With increased number of LbL cycles, the fiber mats transition

from white to turquoise blue. If dried in vacuum, the MOF functionalized fiber mats would

turn deep purple, consistent with the loss of water ligands on the copper paddle wheels, as

observed for solvothermally prepared HKUST-1 MOFs during vacuum degassing. The X-ray

106
diffraction data shown in Figure 3.3b further confirm the HKUST-1 crystal structure. The

diffraction signal intensity increases with the number of LbL cycles.

Figure 3.3 (a) Optical images of ALD-Al2O3-coated PP fibers with 0~40 cycles of LbL

HKUST-1 MOF. (b) X-ray diffraction data of ALD-Al2O3-coated PP fibers with 0, 20, 40

cycles of LbL HKUST-1 MOF. Green triangles (▼) represent the peaks for HKUST-1 MOF.

(c) Thickness of the MOF coating (0~40 LbL cycles) on ALD-Al2O3-coated PP fibers

measured from cross-section TEM images. Error bars represent a 95% confidence interval

based on 25 data points measured for each fiber mat. Solid line is a linear-fitted line to the

data points. (d) Percent mass increase of LbL HKUST-1 MOF (0~40 cycles) based on the dry

weight of the fiber substrates (△m/mo). Dashed line is a linear-fitted line to the data points.

107
To quantify the growth rate of the LbL process, we analyzed MOF film thickness using

high-resolution cross-sectional TEM (Fig. 3.3c) and characterized the mass change vs.

number of LbL cycles (Fig. 3.3d). Figure 3.3c shows the thickness of 0~40 cycles of LbL

MOF coatings on ALD-Al2O3 coated PP fibers. For the TEM analysis, at least 5

cross-sectional images were collected for each sample type, and 5 data points were measured

on each image. The MOF coating thickness increases linearly with the number of LbL cycles,

and the slope corresponds to a growth rate of 3.0 nm/cycle. Using the reported density for

HKUST-1 (0.96 g∙cm-3),36 the change in thickness corresponds to a mass gain of ~288

ng∙cm-2/cycle. The thickness/cycle value we find is close to 2.634 nm, the periodicity of the

HKUST-1 MOF unit cell in the [100] direction,19,36 suggesting the LbL method produces one

MOF “monolayer” per cycle. Previous results show HKUST-1 LbL growth on alumina to

be ~2.5 nm/cycle.20 Small differences in these values may point to subtle but important

differences in the LbL growth reactions. We believe that more detailed studies of half-cycle

reaction saturation and effects of process conditions on MOF growth rate need to be

performed to define true “layer by layer” deposition reaction conditions.

We also measured the mass gain for different numbers of LbL cycles on ALD-coated PP

fibers, and calculated percent mass increase (∆𝑚/𝑚𝑜 ) based on the dry mass of the

substrates. The percent mass gain (Fig. 3.3d) also scales linearly with the number of LbL

cycles, with a slope of 0.435% per cycle (or 4.35× 10−3 gMOF/gfiber per cycle). Considering

the BET surface area of ALD-coated PP fibers is ~1.5 m2/g, the mass gain corresponds to

~290 ng ∙ cm-2/cycle, showing excellent consistency between the mass gain and TEM

thickness analysis results.

108
In addition to PP fibers, we studied the growth rate of LbL HKUST-1 thin films on

ALD-Al2O3 coated silicon wafers by monitoring the characteristic peaks for carboxylate in

Fourier transform infrared (FTIR) spectra. In a typical LbL cycle, Al2O3 coated silicon wafers

were dipped sequentially in BTC and Cu(OAc)2 ethanolic solutions for 1 hour, with 5-minute

ethanol wash steps in between. FTIR spectra were collected for 1, 2, 4, 6, 8 and 10 cycles of

LbL MOF growth. The asymmetric (1645 and 1590 cm-1) and symmetric (1450 and 1370

cm-1) stretching vibration modes shown in Figure 3.4a represent the carboxylate linkers in

the MOF thin films.37 The absorbance of the carboxylate peak at 1370 cm-1 was plotted vs.

number of LbL cycles in Figure 3.4b. A linear increase in IR absorbance at 1370 cm-1 was

observed after the first 2 LbL cycles, indicating a layer-by-layer growth fashion after the

initial nucleation delay.

Adsorption Capacity of Fiber@ALD@MOF

Figure 3.5 shows the BET surface area of PP@ALD@MOF fibers (determined from

BET using an eleven point N2 isotherm, p/po=0.05~0.30, at 77K) plotted versus number of

MOF LbL cycles. ALD Al2O3 coated PP fiber mats had a surface area of ~1.5 m2/g. The

MOF coating added substantial surface area to the substrate, with values exceeding 93

m2/gMOF+fiber after 40 LbL cycles. Note that this surface area is normalized to the total mass of

MOF+fiber, so the surface area per unit mass of MOF is substantially larger. Following

procedures developed previously,30 we analyzed the MOF mass fraction on

fiber@ALD@MOF samples using careful drying and weighing protocols before and after

MOF LbL growth. With this method, after 40 LbL cycles the MOF mass fraction is ~17%,

109
giving a surface area of ~535 m2/gMOF. While the surface area in the range of ~500 m2/gMOF

shows good porosity, the value is still ~2× smaller than typical surface area for bulk

HKUST-1 crystals prepared via solvothermal synthesis. The trend in surface area vs. LbL

cycle in Figure 5 shows the surface area increases non-linearly, with a larger increase during

later LbL cycles. This indicates that the surface area of the growing MOF crystals improves

as growth proceeds, suggesting that MOF thin film deposited in initial LbL cycles may be

amorphous or have a poorer crystallinity than that coated in subsequent LbL cycles.

Figure 3.4 (a) Fourier transform infrared (FTIR) spectra for HKUST-1 thin films deposited

on ALD-Al2O3 coated silicon wafers using layer-by-layer (LbL) method. Al2O3 coated

silicon wafers were dipped sequentially in LbL precursor solutions for 1 hour with 5-minute

ethanol wash steps in between. (b) Plot of IR absorbance at ~1370 cm-1 (carboxylate

symmetric stretching vibrations) vs. number of LbL cycles. A linear increase was observed

for the absorbance at ~1370 cm-1 after 2 LbL cycles.

110
Figure 3.5 Brunauer-Emmett-Teller (BET) surface area (in units of m2/gMOF+Fiber) of

ALD-Al2O3-coated PP fibers with 0~40 cycles of LbL HKUST-1 MOF.

(a) (b)

Figure 3.6 (a) NH3 breakthrough curves for ALD-coated PP fiber mat with no LbL MOF

(■), 20 cycles of LbL MOF (▲) and 40 cycles of LbL MOF (●). (b) NH3 dynamic loading on

ALD-coated PP fiber mats with 0, 20, 40 cycles of LbL MOF. Square points (■) were

calculated based on the corresponding breakthrough curve before saturation, and circle points

(●) were calculated with the desorption part. Error bar represents standard deviation.

111
We also evaluate the adsorption capacity of the Fiber@ALD@MOF materials for

hazardous gases via breakthrough tests. Figure 3.6a shows the NH3 breakthrough curves for 0,

20, 40 cycles of LbL HKUST-1 MOF on ALD-coated PP fibers. Ammonia concentration

detected downstream changes as a function of time, and breakthrough is defined as the time

when the downstream concentration reaches 5% of the feed concentration (Co). Without

MOF coating, PP@ALD fibers show immediate breakthrough when exposed to ammonia,

consistent with zero NH3 adsorption. As the MOF mass increases, breakthrough time

increases, indicating a larger adsorption capacity. PP@ALD@LbL40 fiber mat exhibits

breakthrough time of ~700 min/gMOF+fiber and ~4200 min/gMOF, which is ~80% of the

previously reported values for MOF-fiber mats prepared via solvothermal method.30

Dynamic loadings are calculated for both saturation adsorption with and without

desorption. Saturation dynamic loadings (without desorption) represents the total sorption

capacity, which includes the physisorption and chemisorption capacity. Dynamic loading

with desorption reveals the amount of NH3 retained in the adsorbent even after the feed gas is

terminated, i.e. the amount of NH3 that is chemisorbed. Results from dynamic loading with

and without desorption are shown in Figure 6b. For both measurements, the NH3 dynamic

loading increase almost linearly with the number of LbL cycles. With 40 LbL cycles,

PP@ALD@LbL40 exhibits a dynamic loading (without desorption) of 1.37

molNH3/kgMOF+fiber, equivalent to 7.63 molNH3/kgMOF. With no MOF present, the PP@ALD

samples show nearly zero (0.12 molNH3/kgfiber) dynamic NH3 loading without desorption.

Dynamic loading of 7.63 molNH3/kgMOF agrees well with reported data for bulk HKUST-1

powder (6.6~8.9 molNH3/kgMOF),38 and similar to values reported for solvothermal MOFs

112
grown on fibers.30 With consideration of desorption, PP@ALD@40 can still retain 0.92 mole

of ammonia per kilogram of MOF+fiber.

In addition to ammonia adsorption, we also tested PP@ALD@LbL40 fiber mats for H2S

absorption. The dynamic loading without desorption is up to 1.49 molH2S/kgMOF+fiber (or 9.46

molH2S/kgMOF), and that dynamic loading with desorption is up to 1.44 molH2S/kgMOF+fiber.

Figure 3.7 (a-c) SEM images of 20 cycles of LbL HKUST-1 MOF grown on

ALD-Al2O3-coated polypropylene (a), polyethylene terephthalate (b) and cotton (c). (d)

Percent mass gain for 20 LbL cycles on PP@ALD, PET@ALD and Cotton@ALD. (e) BET

surface area of PP@ALD@LbL20, PET@ALD@LbL20 and Cotton@ALD@LbL20.

LbL MOF on ALD-coated PET and Cotton Fibers

The Fiber@ALD@MOF core@shell structure is not limited to polypropylene substrates.

113
We also deposited LbL MOF thin films on ALD-coated polyethylene terephthalate (PET) and

cotton non-woven fiber mats. Figures 7a-c compare the microscopic morphology of the LbL

MOF thin films grown on PP@ALD, PET@ALD and Cotton@ALD. MOF coating on these

substrates all exhibits good uniformity, high coverage and smooth surface texture. Figure 7d

shows percent mass increase for 20 LbL cycles on different polymer substrates. We find the

percent mass gain data decreases as the diameter of the substrate fiber increases. Assuming

the MOF thickness increase per cycle (∆ℎ, in units of cm/cycle) is same for different fibers,

the percent mass gain will be inversely proportional to the fiber diameter (Eqn. 4).

∆𝑚 𝜌𝑀𝑂𝐹 ∙𝑉𝑀𝑂𝐹 𝜌𝑀𝑂𝐹 ∙(∆ℎ∙𝑆𝐴𝑓𝑖𝑏𝑒𝑟 ) 4 𝜌𝑀𝑂𝐹 ∆ℎ


= = 𝑑𝑓𝑖𝑏𝑒𝑟 = ∙ (4)
𝑚𝑜 𝜌𝑓𝑖𝑏𝑒𝑟 ∙𝑉𝑓𝑖𝑏𝑒𝑟 𝜌𝑓𝑖𝑏𝑒𝑟 ∙(𝑆𝐴𝑓𝑖𝑏𝑒𝑟 ∙ ) 𝜌𝑓𝑖𝑏𝑒𝑟 𝑑𝑓𝑖𝑏𝑒𝑟
4

where ∆𝑚 is mass increase per cycle due to MOF deposition in units of mg/cycle, 𝑚𝑜 is

the substrate weight of ALD-coated fibers in mg, 𝜌𝑀𝑂𝐹 and 𝜌𝑓𝑖𝑏𝑒𝑟 (g∙cm-3) are the densities

of the MOF thin film and the ALD-coated fibers respectively, 𝑆𝐴𝑓𝑖𝑏𝑒𝑟 is the surface area of

the fibers in cm2, 𝑑𝑓𝑖𝑏𝑒𝑟 is the diameter of the fibers in cm. The trend predicted by

equation (4) is consistent with results in Figure 7d, showing decreasing MOF mass fraction

for cotton and PET substrates with larger average fiber diameter. The total BET surface

area (m2/gMOF+fiber) shown in Figure 7e also scales with MOF mass fraction, indicating that

the MOF surface area on these substrates is similar to the MOFs formed on polypropylene.

3.4 Conclusions

We demonstrate fabrication of fiber@ALD@MOF core@shell structures with

conformal and smooth MOF surfaces using a combination of controlled inorganic atomic

114
layer deposition and MOF layer-by-layer synthesis on natural and synthetic polymer fibers.

We use ALD Al2O3 thin film as a nucleation layer for LbL HKUST-1 MOF crystals, and

SEM and cross-sectional TEM images confirm good uniformity and high fiber coverage of

the MOF thin films. After 40 LbL cycles on Al2O3-coated PP fibers, the BET surface area for

the MOF thin films is ~535 m2/gMOF. Although the value is not as high as crystals prepared

via solvothermal methods, we expect the MOF quality and surface area per gram to increase

as growth proceeds further. Analysis of the MOF coating thickness shows the growth rate is

3.0 nm/cycle. Calculation based on both the thickness and the percent mass gain data reveal a

consistent value of 288~290 ng∙cm-2/cycle for the growth rate of LbL HKUST-1 on

ALD-coated PP fibers. Furthermore, the PP@ALD@MOF fiber mats formed by ALD/LbL

have high adsorption capacity for NH3 and H2S. The dynamic loadings of the MOF thin film

are comparable to those of bulk MOF powder, indicating the good quality of the MOF

coating. We also confirm this synthesis route is applicable to other polymer fibers, such as

PET and cotton. In addition to hazardous gas removal, this fiber@ALD@MOF structure is

also promising for catalysis, chemical sensing and many other applications.

115
References

(1) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The Chemistry and
Applications of Metal-Organic Frameworks. Science 2013, 341 (6149), 1230444.

(2) Kim, M.; Cahill, J. F.; Fei, H.; Prather, K. A.; Cohen, S. M. Postsynthetic Ligand and
Cation Exchange in Robust Metal–Organic Frameworks. J. Am. Chem. Soc. 2012, 134
(43), 18082–18088.

(3) Bloch, W. M.; Burgun, A.; Coghlan, C. J.; Lee, R.; Coote, M. L.; Doonan, C. J.;
Sumby, C. J. Capturing snapshots of post-synthetic metallation chemistry in metal–
organic frameworks. Nat. Chem. 2014, 6 (10), 906–912.

(4) Brozek, C. K.; Dincă, M. Cation exchange at the secondary building units of metal–
organic frameworks. Chem Soc Rev 2014, 43 (16), 5456–5467.

(5) Rosi, N. L.; Eckert, J.; Eddaoudi, M.; Vodak, D. T.; Kim, J.; O’Keeffe, M.; Yaghi, O.
M. Hydrogen storage in microporous metal-organic frameworks. Science 2003, 300
(5622), 1127–1129.

(6) Britt, D.; Tranchemontagne, D.; Yaghi, O. M. Metal-organic frameworks with high
capacity and selectivity for harmful gases. Proc. Natl. Acad. Sci. U. S. A. 2008, 105
(33), 11623–11627.

(7) Murray, L. J.; Dinca, M.; Long, J. R. Hydrogen storage in metal-organic frameworks.
Chem. Soc. Rev. 2009, 38 (5), 1294–1314.

(8) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.;
Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal-Organic Frameworks. Chem.
Rev. 2012, 112 (2), 724–781.

(9) Pan, L.; Olson, D. H.; Ciemnolonski, L. R.; Heddy, R.; Li, J. Separation of
Hydrocarbons with a Microporous Metal–Organic Framework. Angew. Chem. Int. Ed.
2006, 45 (4), 616–619.

(10) Li, J.-R.; Kuppler, R. J.; Zhou, H.-C. Selective gas adsorption and separation in
metal-organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1477–1504.

(11) Couck, S.; Denayer, J. F. M.; Baron, G. V.; Remy, T.; Gascon, J.; Kapteijn, F. An
Amine-Functionalized MIL-53 Metal-Organic Framework with Large Separation
Power for CO2 and CH4. J. Am. Chem. Soc. 2009, 131 (18), 6326–6327.

(12) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T.

116
Metal-organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38 (5), 1450–
1459.

(13) Wang, C.; Xie, Z.; deKrafft, K. E.; Lin, W. Doping Metal-Organic Frameworks for
Water Oxidation, Carbon Dioxide Reduction, and Organic Photocatalysis. J. Am.
Chem. Soc. 2011, 133 (34), 13445–13454.

(14) Xie, Y.; Wang, T.-T.; Liu, X.-H.; Zou, K.; Deng, W.-Q. Capture and conversion of CO2
at ambient conditions by a conjugated microporous polymer. Nat. Commun. 2013, 4.

(15) Sun, D.; Fu, Y.; Liu, W.; Ye, L.; Wang, D.; Yang, L.; Fu, X.; Li, Z. Studies on
Photocatalytic CO2 Reduction over NH2-UiO-66(Zr) and Its Derivatives: Towards a
Better Understanding of Photocatalysis on Metal–Organic Frameworks. Chem. – Eur.
J. 2013, 19 (42), 14279–14285.

(16) Wang, S.; Yao, W.; Lin, J.; Ding, Z.; Wang, X. Cobalt Imidazolate Metal–Organic
Frameworks Photosplit CO2 under Mild Reaction Conditions. Angew. Chem. Int. Ed.
2014, 53 (4), 1034–1038.

(17) Bo, L.; Fischer, R. A. Liquid-phase epitaxy of metal organic framework thin films. Sci.
China-Chem. 2011, 54 (12), 1851–1866.

(18) Bradshaw, D.; Garai, A.; Huo, J. Metal-organic framework growth at functional
interfaces: thin films and composites for diverse applications. Chem. Soc. Rev. 2012,
41 (6), 2344–2381.

(19) Munuera, C.; Shekhah, O.; Wang, H.; Wöll, C.; Ocal, C. The controlled growth of
oriented metal–organic frameworks on functionalized surfaces as followed by
scanning force microscopy. Phys. Chem. Chem. Phys. 2008, 10 (48), 7257–7261.

(20) Stavila, V.; Volponi, J.; Katzenmeyer, A. M.; Dixon, M. C.; Allendorf, M. D. Kinetics
and mechanism of metal-organic framework thin film growth: systematic investigation
of HKUST-1 deposition on QCM electrodes. Chem. Sci. 2012, 3 (5), 1531–1540.

(21) Shekhah, O.; Wang, H.; Paradinas, M.; Ocal, C.; Schuepbach, B.; Terfort, A.; Zacher,
D.; Fischer, R. A.; Woell, C. Controlling interpenetration in metal-organic frameworks
by liquid-phase epitaxy. Nat. Mater. 2009, 8 (6), 481–484.

(22) Lu, G.; Hupp, J. T. Metal-Organic Frameworks as Sensors: A ZIF-8 Based Fabry-Perot
Device as a Selective Sensor for Chemical Vapors and Gases. J. Am. Chem. Soc. 2010,
132 (23), 7832–7833.

(23) Heinke, L.; Cakici, M.; Dommaschk, M.; Grosjean, S.; Herges, R.; Bräse, S.; Wöll, C.

117
Photoswitching in Two-Component Surface-Mounted Metal–Organic Frameworks:
Optically Triggered Release from a Molecular Container. ACS Nano 2014, 8 (2),
1463–1467.

(24) Shekhah, O.; Wang, H.; Kowarik, S.; Schreiber, F.; Paulus, M.; Tolan, M.; Sternemann,
C.; Evers, F.; Zacher, D.; Fischer, R. A.; et al. Step-by-step route for the synthesis of
metal-organic frameworks. J. Am. Chem. Soc. 2007, 129 (49), 15118–15119.

(25) Shekhah, O.; Wang, H.; Zacher, D.; Fischer, R. A.; Wöll, C. Growth Mechanism of
Metal–Organic Frameworks: Insights into the Nucleation by Employing a
Step-by-Step Route. Angew. Chem. Int. Ed. 2009, 48 (27), 5038–5041.

(26) Liu, B.; Ma, M.; Zacher, D.; Betard, A.; Yusenko, K.; Metzler-Nolte, N.; Woell, C.;
Fischer, R. A. Chemistry of SURMOFs: Layer-Selective Installation of Functional
Groups and Post-synthetic Covalent Modification Probed by Fluorescence Microscopy.
J. Am. Chem. Soc. 2011, 133 (6), 1734–1737.

(27) Arslan, H. K.; Shekhah, O.; Wohlgemuth, J.; Franzreb, M.; Fischer, R. A.; Woell, C.
High-Throughput Fabrication of Uniform and Homogenous MOF Coatings. Adv.
Funct. Mater. 2011, 21 (22), 4228–4231.

(28) Meilikhov, M.; Yusenko, K.; Schollmeyer, E.; Mayer, C.; Buschmann, H.-J.; Fischer, R.
A. Stepwise deposition of metal organic frameworks on flexible synthetic polymer
surfaces. Dalton Trans. 2011, 40 (18), 4838–4841.

(29) Abbasi, A. R.; Akhbari, K.; Morsali, A. Dense coating of surface mounted CuBTC
Metal-Organic Framework nanostructures on silk fibers, prepared by layer-by-layer
method under ultrasound irradiation with antibacterial activity. Ultrason. Sonochem.
2012, 19 (4), 846–852.

(30) Zhao, J.; Losego, M. D.; Lemaire, P. C.; Williams, P. S.; Gong, B.; Atanasov, S. E.;
Blevins, T. M.; Oldham, C. J.; Walls, H. J.; Shepherd, S. D.; et al. Highly Adsorptive,
MOF-Functionalized Nonwoven Fiber Mats for Hazardous Gas Capture Enabled by
Atomic Layer Deposition. Adv. Mater. Interfaces 2014, 1, 1400040 DOI:
10.1002/admi.201400040.

(31) Parsons, G. N.; George, S. M.; Knez, M. Progress and future directions for atomic
layer deposition and ALD-based chemistry. MRS Bull. 2011, 36 (11), 865–871.

(32) Hyde, G. K.; Scarel, G.; Spagnola, J. C.; Peng, Q.; Lee, K.; Gong, B.; Roberts, K. G.;
Roth, K. M.; Hanson, C. A.; Devine, C. K.; et al. Atomic Layer Deposition and Abrupt
Wetting Transitions on Nonwoven Polypropylene and Woven Cotton Fabrics.
Langmuir 2010, 26 (4), 2550–2558.

118
(33) Glover, T. G.; Peterson, G. W.; Schindler, B. J.; Britt, D.; Yaghi, O. MOF-74 building
unit has a direct impact on toxic gas adsorption. Chem. Eng. Sci. 2011, 66 (2), 163–
170
.
(34) Jur, J. S.; Spagnola, J. C.; Lee, K.; Gong, B.; Peng, Q.; Parsons, G. N.
Temperature-Dependent Subsurface Growth during Atomic Layer Deposition on
Polypropylene and Cellulose Fibers. Langmuir 2010, 26 (11), 8239–8244.

(35) Silvestre, M. E.; Franzreb, M.; Weidler, P. G.; Shekhah, O.; Woell, C. Magnetic Cores
with Porous Coatings: Growth of Metal-Organic Frameworks on Particles Using
Liquid Phase Epitaxy. Adv. Funct. Mater. 2013, 23 (9), 1210–1213.

(36) Chui, S. S. A Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n.


Science 1999, 283 (5405), 1148–1150.

(37) Petit, C.; Mendoza, B.; Bandosz, T. J. Hydrogen Sulfide Adsorption on MOFs and
MOF/Graphite Oxide Composites. Chemphyschem 2010, 11 (17), 3678–3684.

(38) Peterson, G. W.; Wagner, G. W.; Balboa, A.; Mahle, J.; Sewell, T.; Karwacki, C. J.
Ammonia Vapor Removal by Cu3(BTC)2 and Its Characterization by MAS NMR. J.
Phys. Chem. C 2009, 113 (31), 13906–13917.

119
Supporting Information

Pressure
Gauge

Mass Flow
Controller Samples

N2 Load
Lock

Gate
Valve
Furnace
TMA
H2O

To Pump

Figure S3.1 Schematic of the homemade hot-wall viscous-flow ALD reactor used for ALD

Al2O3 coatings on fibers. In an ALD cycle, trimethylaluminum (TMA) and water are dosed

sequentially into the chamber, with a purge step of inert gas (N2) in between. Deposition

temperature is controlled by the furnace, and all the gas lines and valves are wrapped with

heating tapes to prevent precursor condensation.

120
CHAPTER 4 is a reprint of a manuscript published Journal of American Chemical Society.

CHAPTER 4. Facile Conversion of Hydroxy Double Salts to

Metal-Organic Frameworks Using Metal Oxide Particles and

ALD Thin Film Templates

Junjie Zhao †, William T. Nunn †, Paul C. Lemaire †, Yiliang Lin †, Michael D. Dickey †,

Howard J. Walls ‡, Gregory W. Peterson #, Mark D. Losego §, and Gregory N. Parsons †,*


Department of Chemical and Biomolecular Engineering, North Carolina State University,

911 Partners Way, Raleigh, NC 27695, U.S.A (E-mail: gnp@ncsu.edu)



RTI International, 3040 East Cornwallis Road, Research Triangle Park, NC 27709, U.S.A.
#
Edgewood Chemical Biological Center, 5183 Blackhawk Road, Aberdeen Proving Ground,

MD 21010, U.S.A.
§
School of Materials Science and Engineering, Georgia Institute of Technology, 771 Ferst Dr.

NW, Atlanta, GA 30332, U.S.A.

121
Abstract

Rapid room-temperature synthesis of metal-organic frameworks (MOFs) is highly

desired for industrial implementation and commercialization. Here, we find that a (Zn,Cu)

hydroxy double salt (HDS) intermediate formed in situ from ZnO particles or thin films

enables rapid growth (< 1 min) of HKUST-1 (Cu3(BTC)2) at room temperature. The

space-time-yield reaches over 3×104 kg∙m-3∙d-1, which is at least one order of magnitude

greater than any prior report. The high anion exchange rate of (Zn,Cu) hydroxy nitrate HDS

drives the ultra-fast MOF formation. Similarly, we obtained Cu-BDC, ZIF-8 and IRMOF-3

structures from HDS’s, demonstrating synthetic generality. Using ZnO thin films deposited

via atomic layer deposition (ALD), MOF patterns are obtained on pre-patterned surfaces, and

dense HKUST-1 coatings are grown onto various form factors including polymer spheres,

silicon wafers, and fibers. Breakthrough tests show the MOF-functionalized fibers have high

adsorption capacity for toxic gases. This rapid synthesis route is also promising for new

MOF-based composite materials and applications.

4.1 Introduction

Metal-organic frameworks (MOFs) are a class of crystalline porous materials that exhibit

high surface area and large pore volume.1 Versatile combinations of MOF constituents enable

structural design and pore size control,2 and post synthetic modification can introduce

additional internal functionality.3 While MOFs are promising for gas adsorption and

separation,4 catalysis/photocatalysis,5 chemical sensing6 and other applications,7 the poor

synthesis rates (space-time-yield typically less than 300 kg∙m-3∙d-1) and harsh processing

122
conditions (high temperature and pressure) of traditional solvothermal methods still remain

as major hurdles for industrial implementation of MOFs and MOF-functionalized

composites.8 Therefore, new synthetic routes are highly desired to enable fast MOF

formation at low reaction temperature.

Several strategies have been attempted to tackle this challenge, including

mechanochemical methods,9 irradiation-assisted synthesis,10 and electrochemical

approaches.11 While these methods can reduce the reaction time and synthesis temperature

for MOF production, large amounts of external energy input are required to enable the

chemical reactions. Recently, metal oxides and hydroxides have been reported to act as

nucleating agents or sources of cations for room-temperature MOF synthesis.12-13 However,

the incorporation of residual oxide/hydroxide seeds within the MOF crystals due to

insufficient conversion affect the purity and properties of these MOFs. Although extending

reaction time to over 1 hour12 or increasing reaction temperature to >95℃13 can reduce the

oxide/hydroxide residue in the final product, alternative approaches with rapid conversion at

~25 °C are still needed.

Here we report a novel synthesis method using hydroxy double salts (HDS’s) as

intermediates. HDS’s are layered compounds consisting of cationic sheets connected by

inorganic/organic interlamellar anions.14 These materials are generally synthesized by

reacting one divalent metal oxide (MeO) with another different divalent cation (M2+). Similar

to layered double hydroxides, HDS’s show excellent anion exchangeability, and are

promising for catalysis15 and pharmaceutical applications.16 We find, for the first time, that

HDSs can undergo facile conversion and enable rapid formation of MOFs at room

123
temperature. Our space-time-yield (STY) for HKUST-1 (Cu3(BTC)2) reaches up to 3.6×104

kg∙m-3∙d-1, which is at least one order of magnitude greater than any prior report.8,11-12

4.2 Results and Discussion

Figure 4.1a depicts our rapid room-temperature approach for synthesizing HKUST-1

powder. Cu(NO3)2 aqueous solution and DMF are added to ZnO nanoslurries (dispersed in

H2O, average particle size = 252±23 nm, Figure S4.1) at room temperature to form a light

blue suspension of (Zn,Cu) hydroxy nitrate HDS. H3BTC ethanolic solution is immediately

added to the suspension, turning it turquoise in about 10 seconds (Video S4.1 in the

supporting information). This turquois color is a definitive visual indication of HKUST-1

formation. The powder product was immediately filtered and washed with ethanol after 1

minute of reaction, and subsequently dried in a vacuum oven at 120℃ for 6 hours.

Figure 4.1b compares simulated and experimental X-ray diffraction (XRD) patterns for

Cu3(BTC)2 powders obtained from rapid room-temperature synthesis and 85 °C solvothermal

synthesis. All peak positions match the simulated pattern, indicating the product synthesized

from HDS is crystalline HKUST-1. The MOF powder synthesized at 85℃shows a somewhat

larger (111) peak, possibly a result of dehydration and rehydration. The sharp XRD peaks

confirm good crystallinity of the HKUST-1.

124
Figure 4.1 (a) Schematic of the rapid room-temperature synthesis route for Cu3(BTC)2.

ZnO reacts with Cu(NO3)2 to form (Zn,Cu) hydroxy double salt. The (Zn,Cu) HDS converts

to Cu3(BTC)2 via fast anion exchange. (b) Powder XRD patterns for Cu3(BTC)2 synthesized

in rapid room-temperature method (black) and solvothermal method (red), and simulated

Cu3(BTC)2 pattern (blue). (c) Percent yield in 1 minute of reaction (black circle) and

space-time-yield for the rapid room temperature synthesis (red diamond). Insert image in (c)

is an SEM image of a Cu3(BTC)2 crystal showing the octahedral shape.

125
SEM images (insert image in Figure 4.1c and Figure S4.2a-b) reveal the MOF particles

obtained from the rapid synthesis have an average size of 1.17±0.40 μm (Figure S4.2c). The

octahedron crystal shape is also consistent with previous observations for this fcc-type MOF

crystal.17 After the rapid MOF synthesis, ZnO residue was not observed in SEM images or in

X-ray diffraction patterns, and no Zn was detected on the MOF surface by energy dispersive

X-ray (EDX) analysis (Figure S4.3) or by time-of-flight secondary ion mass spectroscopy

(ToF-SIMS) (Figure S4.4). Inductively coupled plasma optical emission spectroscopy

(ICP-OES) was also used to characterize the concentration of Zn2+ residue in the HKUST-1

powder after degrading the crystals in strong acids and H2O2.18 ICP-OES data (Table S4.1)

reveal Zn2+ concentration is generally less than 0.70 wt%, demonstrating high purity within

the HKUST-1 crystals.

We evaluated the BET surface area of the HKUST-1 powder based on the N2 adsorption

isotherm measured at 77 K. Figure S4.5 shows the N2 adsorption and desorption curves for

the powder prepared via rapid synthesis. The average BET surface area (1895±84 m2/g) is

high compared to previous reports,17,19 also suggesting the high purity of the HKUST-1

powder, as any residue with low surface area will precipitously reduce the overall BET

surface area.

Figure 4.1c shows STY (kg∙m-3∙d-1) for the rapid room-temperature synthesis and the

yield after 1 min of reaction. Both the STY and the yield increase linearly as a function of the

n(ZnO):n(Cu(NO3)2) molar ratio (n(H3BTC):n(Cu(NO3)2) was kept at 1:1.8). Previous

high-throughput synthesis methods for HKUST-1 show STY ranging from 225 to 1842

kg∙m-3∙d-1.11-12 Our synthesis strategy reaches an STY up to 3.6×104 kg∙m-3∙d-1, an

126
improvement of more than one order of magnitude. The yield also reaches up to 98% in just

1 minute of reaction, showing significant promise for scale-up processing. While a further

increase of the n(ZnO):n(Cu(NO3)2) ratio can lead to a higher apparent yield, a corresponding

drop in the BET surface area indicates incomplete HDS conversion due to insufficient

H3BTC reactant. By maintaining the ratio n(ZnO):n(Cu(NO3)2) ≤ 0.8, all HKUST-1 samples

show BET surface area exceeding 1800 m2/g.

We hypothesized that HDS could be the critical intermediate during the rapid

room-temperature MOF synthesis, because (Zn,Cu) hydroxy nitrate HDS can be converted

from ZnO and has been reported with high reaction rate for anion exchange.14 To understand

the reaction mechanism for the rapid MOF growth, we investigated particularly how HDS

forms and how it converts to MOF. Specifically, ZnO thin films (34.4±0.6nm) were deposited

on IR-transparent silicon wafers via atomic layer deposition (ALD) and soaked in Cu(NO3)2

aqueous solution for 1 min, followed by XRD analysis. The resulting XRD pattern (Figure 2a)

matches well with prior reports for (Zn,Cu) hydroxy nitrate.14 Subsequently, the HDS was

exposed to H3BTC solution (DMF : H2O : EtOH = 1 : 1 : 1 volume ratio) for only 30 s, and

analyzed by XRD. The XRD peaks representative for Cu3(BTC)2 in the corresponding

pattern and octahedron crystals found on the surface in SEM images (Figure S6) confirm the

formation of the targeting MOF. The decreased intensity of HDS XRD peaks indicates HDS

is partially consumed after the short exposure to H3BTC.

127
Figure 4.2 (a) XRD patterns for Cu3(BTC)2 powder (black), ALD ZnO surface after

exposure to Cu(NO3)2 for 1 min (blue) and subsequently to H3BTC for 30 s (red). Blue dot

lines represent the (Zn,Cu) hydroxy double salt. (b) FTIR difference spectra for ALD ZnO in

the as-deposited form (black, Si as background), after exposure to Cu(NO3)2 for 1 min

(magenta, previous spectrum as background), and after exposure to H3BTC for 30 s (green,

previous spectrum as background) and the final Cu3(BTC)2 spectrum (orange, Si as

background). (c-d) HAADF STEM images for the cross section of the Cu3(BTC)2 grown on

ALD ZnO coated silicon wafer. The green box in (c) indicates the location of (d-i). (e-i) High

resolution EDX mapping images of the cross section. Scale bars in (b-g) represent 50nm.

128
Figure 4.2b shows FTIR difference spectra collected for ALD ZnO thin film in the

as-deposited form and after sequential exposure to Cu(NO3)2 and H3BTC solutions. After

soaking in Cu(NO3)2 aqueous solution for 1 min, the negative–going mode between 400 cm-1

and 480 cm-1 confirms loss of ZnO, and the appearance of υ(NO− -1 -1
3 ) (1360 cm and 1420 cm )

and the distinct O-H group modes (3300 cm-1 ~ 3620 cm-1) indicate the formation of hydroxy

nitrate.20 After subsequent exposure to the H3BTC solution for 30 s, the peaks for NO−
3 and

O-H diminish while the symmetric (1378 cm-1) and asymmetric (1649 cm-1) stretching

modes for the carboxylate groups appear. This spectrum change clearly reveals the anion

exchange process in the (Zn,Cu) HDS, and further supports our proposed reaction path

(Figure 4.1a); ZnO reacts with Cu(NO3)2 to form (Zn,Cu) hydroxy nitrate HDS, followed by

rapid anion exchange between NO−


3 and OH

in the HDS and BTC3-, yielding ultra-fast

formation of HKUST-1 at room temperature.

We further studied the reactions on ZnO surface when it is exposed simultaneously to

both Cu(NO3)2 and H3BTC. ALD ZnO (58.0±0.6 nm) deposited on a silicon wafer was

soaked in the mixed precursor solution for 5 minutes without stirring. Cross-sections of the

resulting thin films were imaged using high angle annular dark field (HAADF) scanning

transmission electron microscope (STEM) (Figure 4.2c-d), and high-resolution EDX

mapping (Figure 4.2e-i) was used to analyze the elemental distribution in the cross section.

Figure 4.2c shows the abrupt interface between the nucleation layer and the silicon substrate,

and Figure 4.2f indicates Zn is only present in the nucleation layer. The nucleation layer also

reveals a uniformly distributed Cu signal in addition to Zn and O, further evidence for the

formation of (Zn,Cu) HDS. The C, O and Cu signals in the MOF layer are consistent with

129
Cu3(BTC)2. These results therefore indicate that the HDS is an important intermediate in the

rapid MOF synthesis, even in a one–pot batch procedure.

Figure 4.3 (a) Schematic of the fabrication procedure for HKUST-1 patterns. (b-e) SEM

image and EDX mapping images for a star-shape HKUST-1 pattern.

To demonstrate the generality of HDS for MOF synthesis, we synthesized three other

MOFs via HDS intermediates. Table 1 summarizes these reaction routes (XRD in Figure

S4.7-4.9). We formed (Zn,Zn) hydroxy acetate HDS by reacting ZnO with

Zn(CH3CO2)2∙2H2O in deionized water at room temperature for 24h.21 ZIF-8 and IRMOF-3

were obtained from this (Zn,Zn) HDS within 10 min at room temperature. Furthermore,

(Zn,Cu) HDS converts quickly (within 2 min) to Cu-BDC MOF. In addition to ZnO, CuO

and NiO can also react with Zn2+, Cu2+, Ni2+ and Co2+ salts to form HDS.14 However, most

reported methods to synthesize these HDS’s are slow (several days) and may require elevated

temperature. Therefore these HDS are not desirable intermediates for rapid room temperature

MOF synthesis. Important factors for analogous schemes will also include the solubility of

130
organic linkers and the mobility of linkers in HDS lattices.

The rapid room-temperature synthesis method is not limited to forming bulk MOF

powder, but also applicable for MOF patterns and thin film coatings. Figure 3a briefly

describes the fabrication procedure for patterning HKUST-1. We prepared SU-8 (negative

photoresist) patterns on the ALD ZnO surface using photolithography, and then exposed the

patterned ZnO surface to Cu(NO3)2 and H3BTC solutions sequentially. SEM and EDX

mapping images (Figure 4.3b-e) of a star-shape pattern reveal the selective HKUST-1 growth

on ZnO surface. More optical micrographs of the MOF patterns are shown in Figure S4.10.

Table 4.1 Routes for HDS-Driven Room-Temperature Synthesis of Various MOF

Materials.

Metal Oxide Metal Salt HDS Organic Linker MOF HDS-to-MOF Time (min)

ZnO Cu(NO3)2 (Zn,Cu) HDS H3BTC HKUST-1 [Cu3(BTC)2] 1


ZnO Cu(NO3)2 (Zn,Cu) HDS H2BDC Cu-BDC [Cu(BDC)] 2
ZnO Zn(OAc)2 (Zn,Zn) HDS 2-mIM ZIF-8 [Zn(mIM)2] 10
ZnO Zn(OAc)2 (Zn,Zn) HDS H2BDC-NH2 IRMOF-3 [Zn4O(BDC-NH2)3] 10

131
Figure 4.4 (a) Schematic of the rapid room-temperature synthesis route for MOF coatings

onto various form factors. (b-d) SEM images of HKUST-1 deposited onto PS spheres, silicon

wafer and PAN nanofibers, respectively.

Figure 4.4a illustrates the general approach to grow HKUST-1 thin films onto various

form factors. Taking advantage of ALD,22 we can deposit conformal ZnO thin films with

controlled thickness on substrate materials with varied morphologies, such as polystyrene

(PS) spheres, silicon wafers and polyacrylonitrile (PAN) fibers. Within 1 minute of exposure

to the HKUST-1 mother solution, dense coatings of HKUST-1 are also obtained on the

abovementioned ZnO-coated substrates (Figure 4.4b-d). Note that the substrate morphology

is maintained with the conformal ALD ZnO thin films (200 cycles, ~36 nm). The coatings of

densely packed crystals shown in Figure 4.3b-d therefore solely represent the HKUST-1

MOF. More SEM and TEM images for MOF coated PS spheres, polypropylene (PP)

microfibers and PAN nanofibers are shown in Figure S4.11-4.12.

132
Previously, we showed that ALD Al2O3 promotes HKUST-1 solvothermal synthesis and

layer-by-layer growth.23 Here, using HDS formed from ALD ZnO thin films, we reduced the

synthesis time to the scale of minutes. This is also the first example showing fast fabrication

of MOF-functionalized fibrous materials at room temperature. By avoiding long exposure to

heated organic solvents, we expect this synthesis route will enable more MOF-based

composite structures, especially for delicate substrate materials that degrade at high

temperature.

To demonstrate the functionality of these MOF-integrated materials, we characterized

and compared the adsorption capacity of MOF-coated PP microfibers and PAN nanofibers

(referred to as MOF-PP and MOF-PAN, respectively). BET analysis based on N2 isotherm

(Figure 4.5a) reveals the overall surface area is 201 m2/gMOF+Fiber and 524 m2/gMOF+Fiber for

MOF-PP and MOF-PAN, respectively. The larger external surface area on nanofibers enables

higher MOF mass loading, leading to a larger overall BET surface area.

Figure 4.5 (a) BET surface area and (b) NH3 dynamic loading for untreated PP microfibers

and PAN nanofibers, and MOF-coated PP and PAN fibers (MOF-PP, MOF-PAN,

respectively).

133
Breakthrough tests were also performed to quantify the performance for hazardous gas

removal. Figure 4.5b compares the NH3 dynamic loadings on PP and PAN fiber mats with

and without MOF coatings. Untreated PP and PAN fibers can barely retain the NH3 challenge

gas, while MOF-PP and MOF-PAN exhibit 36 and 18 higher dynamic loadings towards

ammonia than the corresponding untreated fibers. In addition to NH3, our

MOF-functionalized fibers also show high adsorption capacity for H2S (Figure S4.14),

another highly toxic industrial chemical. These results all suggest the fibrous materials with

MOF coatings are promising for gas filtration and protective garments.

4.3 Conclusion

In summary, we have discovered an ultra-fast room-temperature MOF synthesis strategy

using hydroxy double salt intermediates. (Zn,Cu) hydroxy nitrate HDS formed in situ from

ZnO shows a high rate of anion exchange in the linker solution and drives the rapid

formation of HKUST-1. The space-time-yield for HKUST-1 reaches 3.6×104 kg∙m-3∙d-1,

more than one order of magnitude higher than previous published reports.11-12 Similar

synthetic strategy has been applied to Cu-BDC, ZIF-8 and IRMOF-3, demonstrating the

synthetic generality. Using ALD ZnO thin films, we have successfully obtained HKUST-1

patterns and dense MOF coatings on PS microspheres, silicon wafers, and PP and PAN fiber

mats in a fast processing rate at room temperature. The HDS-driven MOF synthesis approach

reported here is expected to dramatically improve MOF production rates and widely expand

the material set of MOF-functionalized composites.

134
References

(1) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. Science 2013, 341, 1230444.

(2) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O. M.
Science 2002, 295, 469.

(3) (a) Sadakiyo, M.; Yamada, T.; Kitagawa, H. J. Am. Chem. Soc. 2014, 136, 13166; (b)
Brozek, C. K.; Dincă, M. Chem. Soc. Rev. 2014, 43, 5456.

(4) (a) Britt, D.; Tranchemontagne, D.; Yaghi, O. M. Proc. Natl. Acad. Sci. U. S. A. 2008, 105,
11623; (b) Li, J. -R.; Kuppler, R. J.; Zhou, H. -C. Chem. Soc. Rev. 2009, 38, 1477; (c) Li, K.;
Olson, D. H.; Seidel, J.; Emge, T. J.; Gong, H.; Zeng, H.; Li, J. J. Am. Chem. Soc. 2009, 131,
10368.

(5) (a) Lee, J.; Farha, O. K.; Roberts, J.; Scheidt, K. A.; Nguyen, S. T.; Hupp, J. T. Chem. Soc.
Rev. 2009, 38, 1450; (b) Wang, C.; Xie, Z.; deKrafft, K. E.; Lin, W. J. Am. Chem. Soc. 2011,
133, 13445.

(6) Kreno, L. E.; Leong, K., Farha, O. K.; Allendorf, M.; Van Duyne, R. P.; Hupp, J. T. Chem.
Rev. 2012, 112, 1105.

(7) (a) Meek, S. T.; Greathouse, J. A.; Allendorf, M. D. Adv. Mater. 2011, 23, 249; (b)
Horcajada, P.; Gref, R.; Baati, T.; Allan, P. K.; Maurin, G.; Couvreur, P.; Ferey, G.; Morris,
R.E.; Serre, C. Chem. Rev. 2012, 112, 1232.

(8) Stock, N.; Biswas, S. Chem. Rev. 2011, 112, 933.

(9) (a) Beldon, P. J.; Fábián, L.; Stein, R. S.; Thirumurugan, A.; Cheetham, A. K.; Friščić, T.
Angew. Chem. Int. Ed. 2010, 49, 9640; (b) Klimakow, M.; Klobes, P.; Thünemann, A. F.;
Rademann, K.; and Emmerling, F. Chem. Mater. 2010, 22, 5216.

(10) (a) Khan, N. A.; Jhung, S. H. Bull. Korean Chem. Soc. 2009, 30, 2921; (b) Zou, F.; Yu,
R.; Li, R.; Li, W. ChemPhysChem 2013, 14, 2825.

(11) (a) Müller, U.; Pütter, H.; Hesse, M.; Wessel, H.; Schubert, M.; Huff, J.; Guzmann, M.
International Patent WO/2005/049892, 2005; (b) Müller, U.; Schubert, M.; Teich, F.; Pütter,
H.; Schierle-Arndt, K.; Pastré, J. J. Mater. Chem. 2006, 16, 626.

(12) Majano, G.; Pérez-Ramírez, J. Adv. Mater. 2013, 25, 1052.

(13) Zanchetta, E.; Malfatti, L.; Ricco, R.; Styles, M. J.; Lisi, F.; Coghlan, C. J.; Doonan, C.
J.; Hill, A. J.; Brusatin, G.; Falcaro, P. Chem. Mater. 2015, 27, 690.

135
(14) Meyn, M.; Beneke, K.; Lagaly, G. Inorg. Chem. 1993, 32, 1209.

(15) Hara, T.; Ishikawa, M.; Sawada, J.; Ichikuni, N.; Shimazu, S. Green Chem. 2009, 11,
2034.

(16) (a) Bull, R. M. R.; Markland, C.; Williams, G. R; O’Hare, D. J. Mater. Chem. 2011, 21,
1822. (b) Taj, S. F.; Singer, R.; Nazir, T.; Williams, G. R. Rsc Adv. 2013, 3, 358.

(17) Chui, S. S.-Y.; Lo, S. M.-F.; Charmant, J. P. H.; Orpen, A. G.; Williams, I. D. Science
1999, 283, 1148.

(18) Gotthardt, M. A.; Schoch, R.; Wolf, S.; Bauer, M.; Kleist, W. Dalton Trans. 2015, 44,
2052.

(19) (a) Peterson, G. W.; Wagner, G. W.; Balboa, A.; Mahle, J.; Sewell, T.; Karwacki, C. J. J.
Phys. Chem. C 2009, 113, 13906; (b) Wong-Foy, A. G.; Matzger, A. J.; Yaghi, O. M. J. Am.
Chem. Soc. 2006, 128, 3494.

(20) Secco, E. A.; Worth, G. G. Can. J. Chem. 1987, 65, 2504.

(21) Morioka, H.; Tagaya, H.; Karasu, M.; Kadokawa, J.; Chiba, K. Inorg. Chem. 1999, 38,
4211.

(22) Parsons, G. N.; George, S. M.; Knez, M. MRS Bull. 2011, 36, 865.

(23) (a) Zhao, J.; Losego, M. D.; Lemaire, P. C.; Williams, P. S.; Gong, B.; Atanasov, S. E.;
Blevins, T. M.; Oldham, C. J.; Walls, H. J.; Shepherd, S. D.; Browe, M. A.; Peterson, G. W.;
Parsons, G. N. Adv. Mater. Interfaces 2014, 1, 1400040; (b) Zhao, J.; Gong, B.; Nunn, W. T.;
Lemaire, P. C.; Stevens, E. C.; Sidi, F. I.; Williams, P. S.; Oldham, C. J.; Walls, H. J.;
Shepherd, S. D.; Browe, M. A.; Peterson, G. W.; Losego, M. D.; Parsons, G. N. J. Mater.
Chem. A 2015, 3, 1458.

136
Supporting Information

Experimental Section

Synthesis of HKUST-1 MOF Using (Zn,Cu) Hydroxy Double Salt Intermediate

0.293 g of ZnO powder (≥99.0%, Sigma-Aldrich) was dispersed in 8 mL of deionized

water using sonication for 5~10 min to form the nanoslurry. 1.74 g of Cu(NO3)2∙3H2O

(99-104%, Sigma-Aldrich) was dissolved in 8 mL deionized water

(n(ZnO):n(Cu(NO3)2)=1:2), and 0.840g of 1,3,5-benzenetricarboxylic acid (H3BTC, 95%,

Aldrich) was dissolved in 16 mL ethanol (200 proof, Koptec) separately. After the Cu(NO3)2

aqueous solution and H3BTC ethanolic solution were prepared, ZnO nanoslurries were mixed

with 16 mL of DMF (Fisher Scientific), to which the Cu(NO3)2 aqueous solution was added

and next the H3BTC ethanolic solution under magnetic stirring. After 1 min of reaction, the

powder product was immediately filtered with a polypropylene membrane filter (0.45 μm

pore size, Whatman) and washed with ethanol (50 mL, 3 times). The HKUST-1 product was

dried in a fume hood for 30 min and then dried in vacuum oven (~0.01 inHg) at 120℃ for 6

hours. No HKUST-1 formed in the control experiments without ZnO nanoslurries.

Synthesis of Cu-BDC MOF Using (Zn,Cu) Hydroxy Double Salt Intermediate

The synthesis of Cu-BDC MOF form (Zn,Cu) HDS is similar to HKUST-1. 0.293 g of

ZnO powder was dispersed in 6 mL of deionized water using sonication for 5~10 min. 1.74 g

of Cu(NO3)2∙3H2O was dissolved in 6 mL deionized water (n(ZnO):n(Cu(NO3)2)=1:2), and

0.673g of 1,4-benzenedicarboxylic acid (H2BDC, 98%, Aldrich) was dissolved in 25 mL

DMF separately. After the Cu(NO3)2 aqueous solution and H2BDC DMF solution were

137
prepared, ZnO nanoslurries were mixed with 6 mL of DMF, to which the Cu(NO3)2 aqueous

solution was added and next the H2BDC DMF solution under fast magnetic stirring. After 2

min of reaction, the turquoise powder product was immediately filtered and dried in air.

Powder XRD pattern for the Cu-BDC product was measured when the sample was still

damp.

Synthesis of IRMOF-3 using (Zn,Zn) Hydroxy Double Salt Intermediate

(Zn,Zn) hydroxy double salt was synthesized at room temperature using a previously

reported method.1 Specifically, 0.407 g (5 mmol) of ZnO was dispersed in 5 mL of deionized

water, and 1.10 g (5 mmol) of Zn(CH3CO2)2∙2H2O (≥99%, Sigma-Aldrich) was dissolved in

5 mL of deionized water. Zinc acetate aqueous solution was added to the ZnO suspension

under strong magnetic stirring at room temperature. After 24 hours, the mixture converted

into a gel-like viscous fluid as the (Zn,Zn) hydroxy acetate HDS formed. 1.05 g of (Zn,Zn)

HDS suspension was transferred to a beaker, and mixed with 25 mL of 2-aminoterephthalic

acid (0.272 g, 1.5 mmol, ≥99%, Acros Organics) DMF solution at room temperature for 10

min. The product was then filtered and dried in air. Powder XRD pattern for the product was

collected when the sample was still damp.

Synthesis of ZIF-8 using (Zn,Zn) Hydroxy Double Salt Intermediate

The synthesis of ZIF-8 from (Zn,Zn) HDS is similar to IRMOF-3. (Zn,Zn) hydroxy

double salt was synthesized at room temperature by mixing 5 mL of ZnO aqueous suspension

(1 M) with 5 mL of Zn(CH3CO2)2 aqueous solution (1 M) and 5 mL of DMF for 24 h. 3 mL

138
of (Zn,Zn) HDS suspension was added to 9 mL of 2-methylimidazole (0.493 g, 6 mmol, 99%,

Aldrich) DMF solution under fast magnetic stirring at room temperature for 10 min. The

product was then filtered and dried in air. Powder XRD pattern for the product was collected

when the sample was still damp.

Atomic Layer Deposition (ALD) for ZnO Thin Films

ALD ZnO thin films were deposited onto polystyrene microspheres, silicon wafers,

polypropylene and polyacrylonitrile fiber mats using a homemade hot-wall viscous-flow

ALD reactor that has been described in our previous work.2 The deposition pressure was ~2

Torr, and the temperature was 100 ℃. In a typical ALD ZnO cycle, diethyl zinc (DEZ, 95%,

STREM Chemicals) was first dosed to the reactor chamber for 2 s, followed by 60 s of N2

(dried with an Entegris gatekeeper) purge. After DEZ dose and N2 purge, deionized water

was dosed for 2 s, and another 60s of N2 purge completed the ALD cycle.

HKUST-1 Precursor Solutions

0.870 g of Cu(NO3)2∙3H2O was dissolved in 12 mL deionized water (Solution A), and

0.420g of H3BTC was dissolved in 12 mL ethanol (Solution B).

For reaction mechanism studies and MOF patterning, Solution B was mixed with equal

volumes of DMF and deionized H2O, making a H3BTC solution (referred to as Solution B’)

with mixed solvents (DMF:EtOH:H2O=1:1:1).

In the dip coating processes for growing MOFs onto different substrates, 3 mL of

Solution A was first mixed with 3mL DMF and then 3mL of Solution B in a 20 mL

139
scintillation vial (VWR International) under mild magnetic stirring for 1 min. The mixed

precursor solution (Solution M) was then used for growing HKUST-1 coatings on

ZnO-coated substrates.

Fabrication Procedure for HKUST-1 Patterns

Negative photoresist SU-8 2050 (Microchem) was used as received and spun-coated

(3000 rpm for 30 s) onto the ALD ZnO coated Si wafer. After soft baking at 65℃ for 1 min

and 95℃ for 7 min, the wafer was exposed to UV lamp (INTELLI-RAY 400, 60% intensity)

for 5 s. The wafer was then baked at 65 ℃ for 1 min and 95 ℃ for 6 min, and

subsequently dipped into SU-8 developer. The pre-patterned sample was rinsed in IPA and

ethanol, and dried in compressed air. During MOF patterning, the sample was soaked in

Solution A for 1 min, and washed in ethanol for 1 min. The wafer was then transferred to

Solution B’ for 45~60s, followed with gentle ethanol rinse for 5 min. The patterned sample

was finally dried in compressed air.

Synthesis of HKUST-1 Coatings on Polystyrene Spheres

80 μL Polybead® polystyrene (PS) microspheres in aqueous suspension (3 μm diameter,

2.5% solids in water, Polysciences Inc.) were spun-coat onto silicon wafers pretreated with

O2 plasma. The prepared sample was then coated with 200 cycles of ALD ZnO, and then

dipped in Solution M for 1 min. The sample was then slowly lifted out of the MOF precursor

solution, and carefully transferred and soaked in ethanol for 5 min. After ethanol rinsing, the

140
samples were dried in air.

Synthesis of HKUST-1 Coatings on Silicon Wafers

200 cycles of ALD ZnO were deposited onto silicon wafers using the abovementioned

process. After ALD coating, Si wafers were soaked in Solution M for 1 min, and then

carefully transferred to ethanol for 5 min of soaking and finally dried in air.

Synthesis of HKUST-1 Coatings on Fibers

Nonwoven PP microfiber mats were used as received from the Nonwovens Cooperative

Research Center (NCRC) at North Carolina State University. Electrospun PAN nanofibers

were used as received from RTI International. 200 cycles of ALD ZnO were deposited onto

these fiber mats using the process described above. After ALD coating, the PP/ZnO and

PAN/ZnO mats were soaked in the Solution M for 1 min and 5 min respectively, dried in air

for 1 hour, and rinsed in methanol for solvent exchange for 1 day. The MOF coated fibers

were finally dried at room temperature under vacuum.

Space-Time-Yield (STY) Calculation

The space-time-yield (STY, in units of kg·m-3·d-1) was calculated using the following

equation:

𝑚𝑀𝑂𝐹
𝑆𝑇𝑌 = × 1.44 × 106 (1)
𝑉𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛 ∙ 𝜏

where mMOF is the dry mass (g) for the MOF powder obtained from the rapid synthesis,

141
Vsolution is the total volume (cm3) for the mixed precursor solution, and 𝜏 is the residence

time (min).

Material Characterization

Scanning electron microscopic (SEM) images were taken using a JEOL JSM 6010 SEM

and an FEI Verios 460L field emission SEM. A thin layer of Au-Pd (5~10 nm) was

sputter-coated onto all samples before SEM imaging. Energy dispersive X-ray (EDX)

analysis for HKUST-1 powder was performed using an FEI Verios 460L field emission SEM

equipped with an Oxford energy dispersive X-ray spectrometer. The cross sections of the

MOF-coated silicon wafers were prepared with an FEI Quanta 3D FEG focused ion beam

(FIB), and imaged using an FEI Titan 80-300 probe aberration corrected scanning

transmission electron microscope (STEM). High-resolution EDX was analyzed using the

SuperX Energy Dispersive Spectrometry (SuperX EDS) system installed on the FEI Titan

STEM. Time-of-flight secondary ion mass spectroscopy (TOF-SIMS) analysis for HKUST-1

powder was performed using a TOF-SIMS V Instrument (ION TOF, Inc. Chestnut Ridge,

NY). Perkin Elmer 8000 Inductively-coupled plasma-optical emission spectroscopy

(ICP-OES) was used for analyzing the concentration of Cu and Zn in the MOF powder and

the residue solution collected after filtration. Before ICP-OES analysis, the MOF powder was

digested in a mixture of 5 mL HNO3, 3 mL HCl and 0.5 mL H2O2 for 1 hour using a CEM

Mars 5 microwave digesting system.3 X-ray diffraction (XRD) was conducted with a Rigaku

SmartLab X-ray diffraction tool (Cu Kα X-ray source) for crystalline phase analysis. MOF

powder diffraction patterns were also simulated using Mercury 3.0 software and the

142
crystallographic information files from Cambridge Crystallographic Data Centre (CCDC

112954 for HKUST-1, CCDC 687690 for Cu-BDC, CCDC 175574 for IRMOF-3, and CCDC

602542 for ZIF-8). A Quantachrome Autosorb-1C surface area and pore size analyzer was

used for measuring N2 isotherm at 77K. Samples were dried in vacuum (~1 × 10−5 𝑇𝑜𝑟𝑟) at

room temperature for 12h before N2 adsorption measurement. BET surface area was

calculated based on the N2 adsorption data within a relative pressure range of P/Po = 0.02 ~

0.10.4 A Thermo Scientific Nicolet 6700 Fourier transform infrared spectrometer was used

for analyzing MOF growth on IR silicon wafers. Dynamic light scattering (Zetasizer Nano,

Malvern Instruments Ltd) was used to measure the size distribution of ZnO nanoslurries.

Breakthrough Test for NH3 and H2S

The adsorption performance of our MOF-functionalized fiber mats was characterized

with a custom-built rapid, micro-breakthrough system.1 Challenge gas (NH3 or H2S in

moisturized air, 1000 mg/m3 concentration, 50% relative humidity) was injected into an

adsorbent column loaded with MOF-fiber material (~20 mg). The column temperature was

kept at 20°C in a water bath. The downstream concentration was analyzed with a

continuously measuring gas chromatograph (HP5890 Series II) equipped with a

photoionization detector for NH3 (or a flame photometric detector for H2S).5

Breakthrough Data Analysis

Dynamic loading (DL in units of mol/kg) of challenge gas on MOF-functionalized fibers

was calculated using the following equations.5

143
𝐶𝑓𝑒𝑒𝑑 𝐹𝑓𝑒𝑒𝑑 ∙ 𝑡𝑡𝑜𝑡𝑎𝑙
𝑁𝑓𝑒𝑒𝑑 = (2)
𝑀𝑤
𝑡𝑡𝑜𝑡𝑎𝑙 𝐶
𝑜𝑢𝑡 𝐹𝑓𝑒𝑒𝑑
𝑁𝑜𝑢𝑡 = ∫ 𝑑𝑡 (3)
0 𝑀𝑤
𝑁𝑓𝑒𝑒𝑑 − 𝑁𝑜𝑢𝑡
𝐷𝐿 = (4)
𝑚𝑎𝑑𝑠

Nfeed (mol) is the total moles of challenge gas injected into the adsorbent column, Nout (mol) is

the total moles of challenge gas detected in the downstream. Cfeed and Cout (g/m3) are the

concentrations of challenge gas in the feed and the downstream respectively. Ffeed (m3/min) is

the feed flow rate, t (min) is test time. Mw (g/mol) is the molecular weight of the challenge

gas, and mads (kg) is the adsorbent mass.

144
Figure S4.1 Size distribution of ZnO nanoslurries dispersed in deionized water measured

by dynamic light scattering (DLS). The average particle size is 252±23 nm, and the

polydispersity index is 0.732.

145
Figure S4.2 (a-b) SEM images of HKUST-1 crystals obtained from the rapid

room-temperature synthesis. (c) Crystal size distribution analyzed from SEM images (100

measured data points). The average crystal size is 1.17±0.40 μm.

146
Figure S4.3 (a) SEM image of HKUST-1 crystals prepared via rapid synthesis (dispersed

on a silicon wafer). (b-d) Energy dispersive X-ray (EDX) mapping images for C, O and Cu

in the HKUST-1 crystals. (e) EDX spectrum showing that no Zn can be detected by EDX

(blue dash lines indicate where Zn can be expected).

147
Figure S4.4 (a-c) ToF-SIMS surface mapping (negative ion mode) images for a layer of

densely packed HKUST-1 crystals. These results confirm the presence of (a) oxygen, (b)

carbon and (c) copper. (d) ToF-SIMS depth profile (positive ion mode) for C+, Cu+ and 65Cu+

in a layer of densely packed HKUST-1 crystals. Zn-containing residue was not found by

either positive or negative detection mode.

148
Table S4.1 ICP-OES for the concentration of Cu and Zn in the HKUST-1 crystals prepared

through the rapid room-temperature MOF synthesis.

Reaction Cu Concentration* Zn Concentration n(Zn):n(Cu)


n(ZnO):n(Cu(NO3)2) Time (min) in MOF (wt%) in MOF (wt%) in MOF
1:5 1 20.9 0.034 1:628
1:3 1 28.3 0.43 1:67
2:5 1 24.2 0.40 1:62
1:2 1 29.2 0.67 1:45
*The difference of Cu concentration in the MOF is probably a result of solvent inclusion.

Higher Cu concentration is observed for the samples completely activated after vacuum

drying in a BET tube.

Table S4.2 ICP-OES for the concentration of Cu and Zn in the filtrate after collecting the

MOF powder by filtration.

Reaction Cu Concentration Zn Concentration n(Zn):n(Cu)


n(ZnO):n(Cu(NO3)2)
Time (min) in filtrate* (wt%) in filtrate* (wt%) in filtrate
1:3 1 14.8 6.60 1:2.3
1:2 1 11.3 9.71 1:1.2
*The solvent was further evaporated from the filtrate. While the actual concentration of Cu2+

and Zn2+ both increased during solvent removal, the ratio (n(Zn):n(Cu)) did not change in the

concentrated filtrate assuming Cu2+ and Zn2+ are not volatile.

ICP-OES analysis for the filtrate collected after synthesis shows that Zn2+ concentration

in the filtrate is at least one order of magnitude greater than that in the MOF powder, and that

n(Zn):n(Cu) in the filtrate is larger than the initial molar ratio in the mixed reactants. These

results can be explained by the conversion of ZnO to (Zn,Cu) HDS and the release of

Zn-containing species into the solution during the HDS-to-MOF conversion.

149
Figure S4.5 N2 adsorption and desorption isotherm for the HKUST-1 powder prepared via

rapid synthesis.

150
Figure S4.6 SEM images for HKUST-1 grown on top of (Zn,Cu) hydroxy double salt.

151
Figure S4.7 Powder XRD pattern for the Cu-BDC MOF converted from (Zn,Cu) hydroxy

nitrate HDS at room temperature (red). The simulated Cu-BDC pattern is shown in black.

The XRD pattern of the Cu-BDC obtained from HDS agrees well with the simulated pattern

and reported powder XRD patterns for this MOF.6,7 The results from the synthesis of

HKUST-1 and Cu-BDC from (Zn,Cu) HDS indicate that (Zn,Cu) hydroxy nitrate is an

intermediate that preferentially converts to Cu-based MOFs.

152
Intensity (a.u.)

(Zn,Zn) hydroxy acetate HDS

IRMOF-3 converted from (Zn,Zn) HDS

(731)
(200)

(220)

Simulated IRMOF-3 Pattern


(440)
(420)

(600)(531)

(751)
(533)
(444)
(711)
(642)
(111)

(311)
(222)

(620)

(733)

(753)
(400)

(511)

5 10 15 20 25 30 35 40 45
2degree)

Figure S4.8 Powder XRD patterns for the (Zn,Zn) HDS (blue) and IRMOF-3 converted

from (Zn,Zn) HDS at room temperature (red). The simulated IRMOF-3 pattern is shown in

black.

153
Intensity (a.u.) (Zn,Zn) Hydroxy Acetate HDS

(211) ZIF-8 Converted from (Zn,Zn) HDS


(110)

(310)
(222)

(332)
Simulated ZIF-8 Pattern
(411)
(321)
(220)
(200)

5 10 15 20 25 30 35 40 45
2(degree)
Figure S4.9 Powder XRD patterns for the (Zn,Zn) hydroxyl acetate HDS synthesized with

DMF (blue) and ZIF-8 converted from (Zn,Zn) HDS (red). The simulated ZIF-8 pattern is

shown in black.

154
Figure S4.10. (a) Photo of circular and star-shape HKUST-1 patterns. The red box in (a)

represents the location of Image (b). (b-d) Optical micrograph and SEM images of star

patterns made from HKUST-1. The red box in (c) shows the location of (d).

155
Figure S4.11 SEM images for (a) untreated PS microspheres, (b) PS microspheres with

ALD ZnO coating, and (c-d) HKUST-1 grown on ZnO-coated PS microspheres.

156
Figure S4.12 (a-c) SEM images for ZnO-coated PP fibers (a) before and (b-c) after

HKUST-1 rapid synthesis. Insert photo in (b) shows the macroscopic uniformity of MOF

growth on the PP fiber mat. (d) Cross-sectional TEM image shows uniform HKUST-1

coating on ZnO-coated PP fibers. (e-f), SEM images for ZnO-coated PAN nanofibers (e)

before and (f) after HKUST-1 rapid synthesis. Insert optical image in (f) shows the uniform

MOF growth on the PAN nanofiber mat.

157
Figure S4.13 (a) XRD patterns for ALD ZnO coated polypropylene fiber mat (PP/ZnO, red)

and HKUST-1 grown on PP/ZnO (MOF-PP, blue). (b) XRD patterns for ALD ZnO coated

PAN nanofiber mat (PAN/ZnO, red) and HKUST-1 grown on PAN/ZnO (MOF-PAN, blue)

158
Figure S4.14 (a) NH3 breakthrough curves for untreated PP and MOF-PP fiber mats. (b)

H2S breakthrough curves for untreated PP and MOF-PP fiber mats. (c) NH3 breakthrough

curves for untreated PAN and MOF-PAN fiber mats.

(1) Morioka, H.; Tagaya, H.; Karasu, M.; Kadokawa, J.; Chiba, K. Inorg. Chem.
1999, 38, 4211.
(2) Zhao, J.; Gong, B.; Nunn, W. T.; Lemaire, P. C.; Stevens, E. C.; Sidi, F. I.; Williams, P.
S.; Oldham, C. J.; Walls, H. J.; Shepherd, S. D.; Browe, M. A.; Peterson, G. W.; Losego,
M. D.; Parsons, G. N. J. Mater. Chem. A 2015, 3, 1458.
(3) Gotthardt, M. A.; Schoch, R.; Wolf, S.; Bauer, M.; Kleist, W. Dalton Trans. 2015, 44,
2052.
(4) Wong-Foy, A. G.; Matzger, A. J.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 3494.
(5) Glover, T. G.; Peterson, G. W.; Schindler, B. J.; Britt, D.; Yaghi, O. M. Chem. Eng. Sci.
2011, 66, 163.
(6) Carson, C. G.; Hardcastle, K.; Schwartz, J.; Liu, X.; Hoffmann, C.; Gerhardt, R. A.;
Tannenbaum, R. Eur. J. Inorg. Chem. 2009, 2009, 2338.
(7) Rodenas, T.; Luz, I.; Prieto, G.; Seoane, B.; Miro, H.; Corma, A.; Kapteijn, F.; Llabrés i
Xamena, F. X.; Gascon, J. Nat. Mater. 2015, 14, 48.

159
CHAPTER 5 is a reprint of a manuscript in preparation.

CHAPTER 5. Conformal Metal-Organic Framework Thin Films

on Nanofibers for Ultra-Fast Degradation of Chemical Warfare

Agents

Junjie Zhao†, Dennis T. Lee†, Robert W. Yaga‡, Morgan G. Hall#,

Ian R. Woodward†, Christopher J. Oldham†, Howard J. Walls‡,

Gregory W. Peterson#,* and Gregory N. Parsons†,*


Department of Chemical and Biomolecular Engineering, North Carolina State University,

911 Partners Way, Raleigh, NC 27695, U.S.A (E-mail: gnp@ncsu.edu)



RTI International, 3040 East Cornwallis Road, Research Triangle Park, NC 27709, U.S.A.
#
Edgewood Chemical Biological Center, 5183 Blackhawk Road, Aberdeen Proving Ground,

MD 21010, U.S.A. (E-mail: gregory.w.peterson.civ@mail.mil)

160
Abstract

The threat associated with chemical warfare agents (CWAs) motivates the development

of novel materials to provide enhanced protection in a reduced burden. Metal-organic

frameworks (MOFs) have recently been shown as highly effective catalysts for detoxifying

CWAs, but challenges still remain for integrating MOFs into functional filter media and/or

protective garments. Here, we report a series of MOF-based nanofiber composites for fast

degradation of CWAs. We found TiO2 coatings deposited via atomic layer deposition (ALD)

onto polyamide-6 nanofibers enable the formation of conformal Zr-based MOF thin films

including UiO-66, UiO-66-NH2 and UiO-67. Cross-sectional TEM images show that these

MOF crystals nucleate and grow directly on and around the nanofibers, with strong

attachment to the substrates. These MOF-functionalized nanofibers exhibit excellent

reactivity for detoxifying CWAs. The half-lives of a CWA simulant compound and nerve

agent soman (GD) are as short as 7.3 min and 2.3 min, respectively. These results therefore

provide the earliest report of MOF/nanofiber textile composites capable of ultra-fast

degradation of CWAs.

5.1 Introducion

Chemical warfare agents (CWAs) are highly toxic compounds that can injure,

incapacitate or even kill human beings. Detoxification of chemical warfare agents (CWAs)

has been of great social significance due to the past accidental or deliberate emissions and

remaining threat posed to civilian and military personnel.1,2 Materials that can efficiently

degrade these lethal chemicals are therefore highly desired to protect soldiers,

161
first-responders and the general public. Activated carbon loaded with various impregnants

including metal salts, amines and acids are currently used for broad spectrum protection.3

However, the complementary functionalities of impregnants can react with each other within

the carbon pores, leading to a decay of performance over time.1 In this regard, new sorbent

materials are of great interest to provide prolonged period of protection in a reduced volume

and weight.

Metal-organic frameworks (MOFs) are a class of crystalline networks consisting

metal-containing building blocks and organic linking units. Their ultra-high surface area,

large porosity and amenability to design and tailor internal chemical functionalities make

MOFs promising candidates for sorption and detoxification of CWAs.4 The high

concentration of metal ion/clusters in the MOF structures also provides numerous

Lewis-acidic sites that mimic the active moieties in enzymes to promote CWA destruction.5

Early studies were focused on the adsorption and catalytic reaction toward CWA and

simulants using Cu-BTC [Cu3(BTC)2] (also referred to as HKUST-1).6–8 While Cu-BTC

shows reasonable removal rate toward sulfur mustard (HD), sarin (GB) and their simulants,

the high water affinity and instability in ambient moisture affect its long-term use.6–8 Later, a

few hydrophobic or water-stable MOFs have been developed and assessed for their catalytic

activity for CWA destruction. Navarro et al. developed several hydrophobic MOFs including

[Zn4O(3,5-dimethyl-4-carboxypyrazolato)3] and [Ni8(OH)4(H2O)2(L)6]n (L =

pyrazolate-based ligands), and evaluated their adsorption of volatile organic compounds.6,9

Farha, Hupp and their co-workers have reported a series of stable MOFs containing

[Zr6O4(OH)4] clusters for fast degradation of soman (GD) and VX and simulants including

162
dimethyl 4-nitrophenyl phosphate (DMNP) and 2-chloroethyl ethyl sulfide (CEES) with

half-lives as short as 0.5 min. These MOFs include UiO-66(67) and their derivatives,

NU-1000, MOF-808 and PCN-222/MOF-545.4,5,10–13

While bulk MOF powders exhibit exceptional catalytic properties for CWA destruction,

many practical issues still need to be addressed before these materials can be widely applied.

For example, the catalytic reactions to degrade CWAs were observed to occur mainly on the

external surface of MOF crystals,11 indicating that the diffusion of CWAs is limited into the

MOF pores. Meanwhile, particles tend to aggregate and consequently lose many accessible

catalytic sites. Thus, it is a challenge to minimize the amount of MOF powder while

maintaining substantial long-term protection. Simplifying the handling and deployment is

also an important issue for these MOF catalysts.

Polymeric nanofibers, often acquired from electrospinning, can exhibit very high

external surface area, excellent water vapor transport property and good mechanical

strength.14,15 Therefore nanofibers are promising scaffolds for MOF-functionalized

composites designed for CWA protection. As far as we know, only one report about

MOF-textile composites for catalytic destruction of CWA simulants has appeared in

literature, where UiO-66 particles are physically attached to silk microfibers.16 Conformal,

high-quality MOF thin films covalently bound to nanofibers is yet to be explored for CWA

destruction application. In this paper, we report a series of novel MOF-nanofiber structures

for this application. Half-lives of nerve agent soman (O-Pinacolyl

Methylphosphonofluoridate, also known as GD) are as short as 2.3 min using our

MOF-functionalized nanofiber catalysts. This is also the first demonstration to detoxify real

163
CWA compound with MOF-fabric composites, as previous work has only investigated

simulants.16

5.2 Results and Discussion

Figure 5.1 describes the procedure to synthesize our MOF-nanofiber composites.

Free-standing polyamide-6 (PA-6) nanofiber mat obtained from electrospinning was coated

with a conformal thin layer (~ 5 nm thick) of TiO2 using atomic layer deposition (ALD). This

ALD TiO2 layer is expected to promote MOF heterogeneous nucleation on fibers,17 and

provide some contribution to catalytic CWA detoxification.18,19 Three Zr-based MOFs

including UiO-66, UiO-66-NH2 and UiO-67 (crystal structures shown in the dashed box in

Figure 5.1) were chosen because previous reports has demonstrated the excellent stability

and good catalytic properties of these MOFs.5,10,11,20 We used concentrated HCl as the

modulator for the solvothermal synthesis which is similar to reported recipes for these

MOFs,21 but developed our own conditions to achieve optimized growth of these Zr-based

MOF thin films onto the nanofibers. Specifically, 0.343 mmol of ZrCl4 and 0.343 mmol of

dicarboxylic acid linkers (1:1 molar ratio) were dissolved in 20 mL of DMF. 25 μL of

deionized water and concentrated HCl (1.33 mL for UiO-66 and UiO-66-NH2, 0.67 mL for

UiO-67 respectively) were added to the solution. Subsequently, TiO2-coated PA-6 nanofiber

(PA-6@TiO2) mat was transferred into the mixed solution, which was then heated to 85 °C

for 24 h. MOF-coated nanofiber mat was collected, washed and activated after the

solvothermal synthesis.

164
Figure 5.1 Schematic of the synthetic procedure for Zr-based MOF functional coatings on

polyamide-6 nanofibers. The MOF crystal structures are illustrated in the dashed box.

Figure 5.2a is a photo of a free-standing PA-6@TiO2 nanofiber mat coated with

UiO-66-NH2 MOF thin films (referred to as PA-6@TiO2@UiO-66-NH2). The structural

integrity and flexibility were fully maintained after the solvothermal synthesis (Video S5.1).

SEM images (Figures 5.2b-d) shows that UiO-66-NH2 nanocrystals (average size = 126±25

nm) were grown conformally on the PA-6@TiO2 nanofibers. While we and other groups

have attempted various methods to deposit UiO-66 onto polymeric fibers in the past,16,17,22-24

our results in Figure 5.2 uniquely demonstrate good growth uniformity and high crystal

coverage on the fibers. The challenge of integrating Zr-based MOFs onto nanofibers is

solved using ALD TiO2 nucleation layers and modified HCl-modulated solvothermal

synthesis. In contrast, the growth of UiO-66-NH2 on PA-6 nanofibers without TiO2

165
nucleation layers is patchy and does not result in a conformal thin film on the nanofiber

surface as that on TiO2 coated PA-6 nanofibers using the same synthesis conditions (Figure

S5.1).

Figure 5.2 (a) Photo of a free-standing PA-6@TiO2@UiO-66-NH2 nanofiber mat. (b-d)

SEM images, (e-i) energy dispersive X-ray mapping images.

Energy dispersive X-ray analysis (Figures 5.2e-i) also confirms uniform MOF growth on

the nanofibers. C, O and N signals originate from the polyamide substrate and the

2-aminoterephthalate linker in the MOF structures. The ALD TiO2 thin film deposited onto

the PA-6 nanofibers also contributes to the EDX intensity of O. In addition, Zr and Ti

mapping images (Figures 5.2g-h) agree well with the secondary electron image (Figure 2d),

showing excellent conformality of the UiO-66-NH2 MOF coatings and the ALD TiO2 thin

166
films, respectively.

The uniform growth of UiO-66 and UiO-67 was also achieved on PA-6@TiO2

nanofibers using our solvothermal synthesis (these two MOF-nanofiber composites are

referred as PA-6@TiO2@UiO-66 and PA-6@TiO2@UiO-67, respectively). Figures 5.3a-c are

SEM images of PA@TiO2 nanofiber mat coated with three different Zr-based MOF thin films.

Although the crystal shapes shown in our MOF-nanofiber structure are distinct from the

octahedron morphology often reported for these fcu-type Zr-based MOFs, X-ray diffraction

patterns of the MOF thin films on nanofibers match well with those of MOF powders

(Figures 5.3g-i), confirming the formation of targeted MOF structures. We speculated that the

spherical shapes of these crystals reflect the rapid kinetics for growing the MOF coatings

because slow reactions could more readily allow formation of crystal facets.21

Figures 5.3d-h are cross-sectional TEM images of the MOF-coated PA@TiO2 nanofibers.

Tubular features observed in these images represent the core@shell structures of PA@TiO2

nanofibers sliced along the axial direction. The diameters of PA-6 nanofibers measured from

TEM images range from 15 nm to 55 nm, consistent with the average fiber diameter (37±16

nm) measured from SEM images (Figure S5.2). The ALD TiO2 thin films deposited onto the

PA-6 nanofibers is conformal and uniform as shown in Figures 5.3d-h. The average thickness

of the TiO2 coatings is 5.7±1.3 nm, corresponding to an ALD growth rate of ~ 0.6 Å per

cycle. The spherical MOF crystals are found to nucleate and grow directly on or around the

PA-6@TiO2 nanofibers, indicating formation of strong covalent binding to the substrates. No

noticeable particle shedding during the handling after synthesis also confirms good adhesion

of our MOF thin films to the nanofibers.

167
The quality of the Zr-based MOF thin films grown on nanofibers were characterized

using XRD and BET. The sharp XRD peaks associated with the targeted MOFs (Figures

5.3g-i) demonstrate the good crystallinity of our MOF coatings. BET surface area analysis

was performed on the MOF powders collected from the solvothermal synthesis as well as on

the MOF-nanofiber composites. All samples were washed and activated, and characterized

with N2 adsorption/desorption measurement (Figure 5.3j-l). BET surface area of the

UiO-type MOF powders (Table 5.1) is consistent reported values for all the three MOFs. The

large surface area is a result of proper solvent exchange and activation processes to fully

remove solvent and other guest molecules within the pores of the MOFs.

The BET surface area of PA-6@TiO2 nanofibers coated with UiO-66, UiO-66-NH2 and

UiO-67 is also given in Table 5.1. The surface area for the MOF-nanofiber composites is in

excess of 10 times larger than the nanofiber substrates alone, demonstrating the high porosity

of the MOF coatings on the nanofibers. We find that analyzing the MOF mass fraction in the

composites is challenging with weighing methods. The net mass increase due to MOF

loading is relatively small and often comparable to the expected mass change due to water

uptake by the hygroscopic nylon nanofibers. However, assuming surface area of the MOF

component in the MOF-nanofiber system is equivalent to that of the simultaneously

produced MOF powder, we estimate the MOF mass loading using the calculation methods

described previously.17 The calculated MOF mass fraction is 8.8%, 14.7%, and 15.4% for

PA-6@TiO2@UiO-66, PA-6@TiO2@UiO-66-NH2 and PA-6@TiO2@UiO-67, respectively.

These results were further confirmed by elemental analysis using inductively coupled plasma

optical emission spectroscopy (ICP-OES) (Table S5.1).

168
Figure 5.3 (a-c) SEM images of (a) PA-6@TiO2@UiO-66, (b) PA-6@TiO2@ UiO-66-NH2

and (c) PA-6@TiO2@UiO-67. (d-f) Cross-sectional TEM images of (d)

PA-6@TiO2@UiO-66, (e) PA-6@TiO2@UiO-66-NH2 and (f) PA-6@TiO2@UiO-67. (g-i)

XRD patterns of PA-6 nanofibers before and after ALD, MOF-coated nanofibers and MOF

powders. (j-l) N2 adsorption and desorption isotherms for PA-6@TiO2 nanofibers with and

without MOF coatings and Zr-based MOF powders.

169
To evaluate the catalytic property of our MOF-nanofiber composites for CWA

destruction, we first analyzed the hydrolysis of simulant 4-nitrophenyl phosphate (DMNP).

Since nerve agents are extremely toxic and requires special caution to handle, investigation

of simulants with phosphate ester bonds is a more accessible approach for almost all

laboratories to test the efficacy of the catalysts and compare the kinetics.4 Here, we

investigated the catalytic degradation of DMNP (Figure 5.4a) in an aqueous buffer solution

of N-ethylmorpholine (0.45M, pH = 10). 2.5 mg of MOF powders or 14 mg of MOF-coated

nanofibers was used the catalysts, and the reaction kinetics was characterized using a

procedure similar to the methods described in previous reports.4,5

Figure 5.4b shows a typical serious of spectra measured during the catalytic hydrolysis

of DMNP. The decreased absorption at 270 nm represents the reduced concentration of

DMNP, and the increase of absorbance at 407 nm corresponds to the formation of

p-nitrophenoxide.5 We monitored the reaction progress by tracking the absorbance at 407 nm,

and calculated the concentration of p-nitrophenoxide based on Lambert-Beer Law (Equation

1). The percent conversion of DMNP during catalytic degradation using MOF powders or

MOF-nanofiber composites is plotted as a function of time in Figures 5.4c-e. For Zr-based

MOF powders, we observed 95%, 98% and 96% DMNP conversion in 60 min of reaction

when UiO-66, UiO-66-NH2 and UiO-67 were used respectively. Half-life (t1/2) of DMNP

with MOF powder catalysts (Table 5.1) shows similar trend to reported data. UiO-66-NH2

exhibits the fastest degradation rate (t1/2 = 2.8 min) among the three MOF powders, while

UiO-67 also significantly reduces the half-life of DMNP compared to UiO-66. We speculate

that the amine moieties in UiO-66-NH2 synergistically enhance the catalytic activity, while

170
the large pore size of UiO-67 may allow further diffusion of DMNP molecules into the

catalyst.

Figure 5.4 (a) Catalytic reaction of DMNP hydrolysis using Zr-based MOF powder and

MOF functionalized nanofiber catalyst. (b) UV/Visible absorption spectra for monitoring

degradation of DMNP. (c-e) Conversion of DMNP to p-nitrophenoxide versus reaction time

using MOF powder and functionalized nanofiber catalyst.

171
Table 5.1 Material properties and catalytic performance towards CWA simulant

degradation.

BET Surface MOF wt% DMNP t1/2 Ref.


2
Area (m /g) (min)
1512 100 22 This work
UiO-66 a
1580 100 24, 45 5, 11
1334 100 2.8 This work
UiO-66-NH2
1200a 100 0.7 11
2250 100 7.7 This work
UiO-67
2500a 100 3.8 11
PA-6@TiO2@UiO-66 143.9 8.8 135 This work
PA-6@TiO2@UiO-66-NH2 205.9 14.7 7.3 This work
PA-6@TiO2@UiO-67 356.2 15.4 7.4 This work
PA-6@TiO2 11.5 0 1170 This work
PA-6 11.5 0 3950 This work
*BET surface area values are adapted from Ref. 21.

We also characterized the catalytic performance of the MOF-functionalized nanofibers

for degrading CWA simulant DMNP. The t1/2 values of DMNP using MOF-nanofiber

composite catalysts and control materials are summarized in Table 5.1. For untreated PA-6

nanofibers, DMNP shows negligible rate of hydrolysis with an estimated t1/2 value over 65 h

(Figure S5.3). With ALD TiO2 coatings, PA-6@TiO2 reduces the half-life to ~20 h, consistent

with the reported catalytic effect of TiO2 for degrading CWAs.18,19 Compared with PA-6 and

PA-6@TiO2, MOF-coated PA-6@TiO2 nanofibers exhibit significantly enhanced catalytic

performance. Both PA-6@TiO2@UiO-66-NH2 and PA-6@TiO2@UiO-67 enable short

half-life of DMNP (< 8 min) and high conversion (> 90%) in 60 min, demonstrating the

172
excellent catalytic properties of these MOF-nanofiber composites. PA-6@TiO2@UiO-66

shows a slower DMNP hydrolysis rate than the other two composite structures, because of

the smaller MOF mass loading and the lower catalytic activity of UiO-66 compared to

UiO-66-NH2 and UiO-67. Notably the DMNP t1/2 values for the PA-6@TiO2@UiO-66-NH2

and PA-6@TiO2@UiO-67 are all within the same magnitude as those for the corresponding

MOF powder catalysts, indicating that the functionality of MOFs are well maintained after

growth on nanofibers. Detailed analyses of the reaction kinetics are shown in Figure

S5.4-S5.5 and Table S5.2. SEM images and EDX spectra taken for the MOF-nanofiber

composites after DMNP degradation (Figure S5.6) show that significant amounts of Zr-based

MOF coatings remain in the composite structures even after strong agitation during the

DMNP hydrolysis tests. These results also demonstrate the good adhesion of our MOF thin

films to the nanofiber substrates.

Figure 5.5 (a) Catalytic reaction of GD hydrolysis using MOF-nanofiber catalyst. (b)

Conversion of GD versus reaction time during catalysis. Dashed lines are fitted results

assuming first order reaction kinetics.

173
In addition to simulant DMNP, we further tested our MOF-nanofiber composites for the

destruction of the highly toxic nerve agent O-pinacolyl methylphosphonofluoridate (GD)

(Figure 5.5a). For this test, 2.6 µL of GD was added to approximately 14 mg of

MOF-nanofiber catalyst in an NMR tube, followed with vigorous shaking in order to mix GD

well with MOF-nanofibers in the solution. The tube was then analyzed for GD degradation
31
and product formation using P NMR. Figure 5b is a plot of the conversion of GD during

catalytic degradation using MOF-functionalized nanofibers. The doublet peaks at

approximately 40 and 33 ppm associated with GD immediately decrease upon exposure to

the MOF-nanofibers (Figures S5.7-S5.9). The pinacolyl methylphosphonic acid (PMPA)

peak (approximately 27 ppm) begins growing at the same time, indicating detoxification of

the CWA to a non-toxic product. The half-lives of GD are 3.0 min (with

PA-6@TiO2@UiO-66), 3.7 min (with PA-6@TiO2@UiO-66-NH2) and 2.3 min (with

PA-6@TiO2@UiO-67), respectively. All three MOF-nanofiber composites show fast GD

destruction (t1/2 ≤ 4 min) and high conversion (≥80%) within 10 min. The fastest reaction

rate with PA-6@TiO2@UiO-67 is possibly because the large pore size of UiO-67 allows

diffusion of reactants into the pores while the catalytic reactions occur mainly on the external

surface of UiO-66 and UiO-66-NH2.11 As far as we know, this is the first demonstration of

MOF-fiber composites for detoxifying real CWA compounds. Our results are also very

promising since as the catalytic performance of these MOF thin films on nanofibers is

comparable or even exceeds the reported bulk MOF powders.4,11 These results clearly reveal

that conformal MOF thin films grown onto nanofiber substrates can achieve excellent

catalytic activity even with small MOF loadings (< 20%). This advantage will eventually

174
benefit the end users by providing substantial catalytic efficacy at a reduced burden.

5.3 Conclusion

In conclusion, we have demonstrated that nanofibers functionalized with Zr-based MOF

thin films are active catalytic materials for decomposing the chemical warfare agent simulant

DMNP and nerve agent GD. ALD TiO2 coatings deposited onto PA-6 nanofibers enabled

direct nucleation and crystal growth of MOFs onto the fiber substrates, leading to conformal

MOF thin films. Cross-sectional TEM images show that these MOF crystals nucleate and

grow directly on or around the nanofibers, indicating strong attachment to the substrates.

XRD and BET confirm that the MOF coatings are highly crystalline with large adsorptive

capacity. We have shown that our MOF-nanofiber composite textiles enable ultra-fast

detoxification of simulant DMNP and nerve agent GD. The half-lives of DMNP and GD are

as short as 7.3 min and 2.3 min, respectively, indicating great promise of our

MOF-functionalized nanofibers for CWA protection. The synthesis method and the

MOF-nanofiber composite structures we have presented here will also offer new

opportunities to advance the development of gas filters, chemical sensors, and potentially

smart textile materials to protect against harmful air pollutants.

175
References

(1) DeCoste, J. B.; Peterson, G. W. Metal–Organic Frameworks for Air Purification of


Toxic Chemicals. Chem. Rev. 2014, 114 (11), 5695–5727.

(2) Enserink, M. U.N. Taps Special Labs to Investigate Syrian Attack. Science 2013, 341
(6150), 1050–1051.

(3) Romero, J. V.; Smith, J. W. H.; Sullivan, B. M.; Mallay, M. G.; Croll, L. M.; Reynolds,
J. A.; Andress, C.; Simon, M.; Dahn, J. R. Gas Adsorption Properties of the Ternary
ZnO/CuO/CuCl2 Impregnated Activated Carbon System for Multigas Respirator
Applications Assessed through Combinatorial Methods and Dynamic Adsorption
Studies. ACS Comb. Sci. 2011, 13 (6), 639–645.

(4) Mondloch, J. E.; Katz, M. J.; Isley Iii, W. C.; Ghosh, P.; Liao, P.; Bury, W.; Wagner, G.
W.; Hall, M. G.; DeCoste, J. B.; Peterson, G. W.; et al. Destruction of chemical
warfare agents using metal–organic frameworks. Nat. Mater. 2015, 14 (5), 512–516.

(5) Katz, M. J.; Mondloch, J. E.; Totten, R. K.; Park, J. K.; Nguyen, S. T.; Farha, O. K.;
Hupp, J. T. Simple and Compelling Biomimetic Metal-Organic Framework Catalyst
for the Degradation of Nerve Agent Simulants. Angew. Chem.-Int. Ed. 2014, 53 (2),
497–501.

(6) Montoro, C.; Linares, F.; Quartapelle Procopio, E.; Senkovska, I.; Kaskel, S.; Galli, S.;
Masciocchi, N.; Barea, E.; Navarro, J. A. R. Capture of Nerve Agents and Mustard Gas
Analogues by Hydrophobic Robust MOF-5 Type Metal-Organic Frameworks. J. Am.
Chem. Soc. 2011, 133 (31), 11888–11891.

(7) Roy, A.; Srivastava, A. K.; Singh, B.; Shah, D.; Mahato, T. H.; Gutch, P. K.; Halve, A.
K. Degradation of sarin, DEClP and DECNP over Cu-BTC metal organic framework.
J. Porous Mater. 2013, 20 (5), 1103–1109.

(8) Peterson, G. W.; Wagner, G. W. Detoxification of chemical warfare agents by CuBTC.


J. Porous Mater. 2014, 21 (2), 121–126.

(9) Padial, N. M.; Quartapelle Procopio, E.; Montoro, C.; Lopez, E.; Enrique Oltra, J.;
Colombo, V.; Maspero, A.; Masciocchi, N.; Galli, S.; Senkovska, I.; et al. Highly
Hydrophobic Isoreticular Porous Metal-Organic Frameworks for the Capture of
Harmful Volatile Organic Compounds. Angew. Chem.-Int. Ed. 2013, 52 (32), 8290–
8294.

(10) Moon, S.-Y.; Liu, Y.; Hupp, J. T.; Farha, O. K. Instantaneous Hydrolysis of
Nerve-Agent Simulants with a Six-Connected Zirconium-Based Metal-Organic

176
Framework. Angew. Chem.-Int. Ed. 2015, 54 (23), 6795–6799.

(11) Peterson, G. W.; Moon, S.-Y.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Hupp, J. T.;
Farha, O. K. Tailoring the Pore Size and Functionality of UiO-Type Metal-Organic
Frameworks for Optimal Nerve Agent Destruction. Inorg. Chem. 2015, 54 (20), 9684–
9686.

(12) Moon, S.-Y.; Wagner, G. W.; Mondloch, J. E.; Peterson, G. W.; DeCoste, J. B.; Hupp, J.
T.; Farha, O. K. Effective, Facile, and Selective Hydrolysis of the Chemical Warfare
Agent VX Using Zr-6-Based Metal-Organic Frameworks. Inorg. Chem. 2015, 54 (22),
10829–10833.

(13) Liu, Y.; Moon, S.-Y.; Hupp, J. T.; Farha, O. K. Dual-Function Metal–Organic
Framework as a Versatile Catalyst for Detoxifying Chemical Warfare Agent Simulants.
ACS Nano 2015, 9 (12), 12358–12364.

(14) Gibson, P.; Schreuder-Gibson, H.; Rivin, D. Transport properties of porous membranes
based on electrospun nanofibers. Colloids Surf. Physicochem. Eng. Asp. 2001, 187–
188, 469–481.

(15) Huang, Z. M.; Zhang, Y. Z.; Kotaki, M.; Ramakrishna, S. A review on polymer
nanofibers by electrospinning and their applications in nanocomposites. Compos. Sci.
Technol. 2003, 63 (15), 2223–2253.

(16) López-Maya, E.; Montoro, C.; Rodríguez-Albelo, L. M.; Aznar Cervantes, S. D.;
Lozano-Pérez, A. A.; Cenís, J. L.; Barea, E.; Navarro, J. A. R. Textile/Metal–
Organic-Framework Composites as Self-Detoxifying Filters for Chemical-Warfare
Agents. Angew. Chem. Int. Ed. 2015, 54 (23), 6790–6794

(17) Zhao, J.; Losego, M. D.; Lemaire, P. C.; Williams, P. S.; Gong, B.; Atanasov, S. E.;
Blevins, T. M.; Oldham, C. J.; Walls, H. J.; Shepherd, S. D.; et al. Highly Adsorptive,
MOF-Functionalized Nonwoven Fiber Mats for Hazardous Gas Capture Enabled by
Atomic Layer Deposition. Adv. Mater. Interfaces 2014, 1 (4), 1400040 DOI:
10.1002/admi.201400040.

(18) Wagner, G. W.; Chen, Q.; Wu, Y. Reactions of VX, GD, and HD with Nanotubular
Titania. J. Phys. Chem. C 2008, 112 (31), 11901–11906.

(19) Wagner, G. W.; Peterson, G. W.; Mahle, J. J. Effect of adsorbed water and surface
hydroxyls on the hydrolysis of VX, GD, and HD on titania materials: the development
of self-decontaminating paints. Ind. Eng. Chem. Res. 2012, 51 (9), 3598–3603.

(20) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud,

177
K. P. A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks
with Exceptional Stability. J. Am. Chem. Soc. 2008, 130 (42), 13850–13851.

(21) Katz, M. J.; Brown, Z. J.; Colon, Y. J.; Siu, P. W.; Scheidt, K. A.; Snurr, R. Q.; Hupp, J.
T.; Farha, O. K. A facile synthesis of UiO-66, UiO-67 and their derivatives. Chem.
Commun. 2013, 49 (82), 9449–9451.

(22) Armstrong, M. R.; Arredondo, K. Y. Y.; Liu, C.-Y.; Stevens, J. E.; Mayhob, A.; Shan,
B.; Senthilnathan, S.; Balzer, C. J.; Mu, B. UiO-66 MOF and Poly(vinyl cinnamate)
Nanofiber Composite Membranes Synthesized by a Facile Three-Stage Process. Ind.
Eng. Chem. Res. 2015, 54 (49), 12386–12392.

(23) Ren, J.; Musyoka, N. M.; Annamalai, P.; Langmi, H. W.; North, B. C.; Mathe, M. Int.
J. Hydrog. Energy 2015, 40 (30), 9382–9387.

(24) Zhang, Y.; Yuan, S.; Feng, X.; Li, H.; Zhou, J.; Wang, B. J. Am. Chem. Soc. 2016, 138
(18), 5785–5788.

178
Supporting Information

Experimental Section

Synthesis of Zr-based MOFs

UiO-66, UiO-66-NH2 and UiO-67 were synthesized using slightly modified recipes

based on previous reports.1,2

For UiO-66, 0.080 g (0.343 mmol) of ZrCl4 (≥99.5%, Alfa Aesar) was added to 20 mL of

DMF (Fisher) in a glass scintillation vial, followed with 1 min of sonication and 5 min of

magnetic stirring to dissolve ZrCl4 well. 0.057 g (0.0343 mmol) of benzene-1,4-dicarboxylic

acid (H2BDC, 98%, Sigma) was then added to the ZrCl4 solution under stirring. Subsequently,

25 μL of deionized water and 1.33 mL of concentrated hydrochloric acid (NF/FCC grade,

Fisher) were added to the mixed solution. The vial was then placed in a box furnace (Thermo

Scientific) and heated to 85 °C for 24 h. After the solvothermal synthesis, UiO-66 powder

product was filtered using polypropylene membrane (0.45 μm pore size, Whatman) and

washed with DMF twice. Anhydrous ethanol (200 proof, VWR) was used for solvent

exchange, and the solvent was replaced every 24 h for a total of two times. After solvent

exchange, the UiO-66 powder was collected via filtration and dried in a dessicator at room

temperature at reduced pressure for 12 h. To further activate the MOF, ~ 0.3 g of powder

product dried in BET at 80 °C for 12 hour and then at 110 °C for 12~18 h (heating

temperature was slowly ramped from 80 °C to 110 °C).

For UiO-66-NH2, the recipe and synthesis procedure are similar to UiO-66. 0.062 g

(0.343 mmol) of 2-aminoterephthalic acid (H2BDC-NH2, 99%, Sigma) was used as the

organic linker instead of H2BDC. The amounts of ZrCl4, DMF, H2O and HCl were kept same

179
as UiO-66. Reaction temperature was controlled at 85 °C for 24 h. After synthesis, the

washing step, solvent exchange and drying process for UiO-66-NH2 were similar to UiO-66.

For UiO-67, 0.083 g (0.343 mmol) of biphenyl-4,4’-dicarboxylic acid (H2BPDC, 97%,

Sigma) was used as ligand. Same amounts of ZrCl4, DMF, H2O were used as UiO-66. The

addition of HCl modulator was reduced to 0.67 mL for 20 mL of mixed solution. Reaction

proceeded at 85 °C for 24 h. After synthesis, the washing step, solvent exchange and drying

process for UiO-66-NH2 were similar to UiO-66.

Electrospinning of Polyamide-6 Nanofiber Mat

Polyamide-6 (referred to as PA-6, Scientific Polymer Products) was dissolved overnight

in a solvent mixture of glacial acetic acid (Sigma Aldrich) and formic acid (spectroscopy

grade, Fluka) (glacial acetic acid : formic acid = 2 : 1 wt/wt) to reach a final PA-6

concentration of 12 wt%. The polymer solution was then loaded into a syringe with a

27 gauge stainless steel needle attached and placed into a custom-built electrospinning

system described previously.3 An atmosphere of bone dry grade carbon dioxide gas (Airgas)

was mixed with water vapor to form ~50% RH at an ambient temperature of ~22°C. This RH

controlled, CO2 atmosphere provides a stable environment for electrospinning the fibers. The

spinning conditions were a voltage gradient of 1.57 kV/cm, a 25 cm spinning distance

(needle to collection substrate distance), and 90 minutes elapsed spinning time. The polymer

solution was continuously fed through the needle at a rate to maintain a stable

electrospinning jet. PA-6 nanofibers were deposited onto a paper ring (10 cm outside

diameter and 6.5 cm inside diameter) sitting on a 3.5-inch (~8.9 cm) petri dish filled with

180
deionized water. The water provided a collection surface for the forming fibers during

electrospinning. After electrospinning the paper ring with deposited nanofiber mat was

carefully removed and dried overnight at ambient conditions.

Atomic Layer Deposition of TiO2 on PA-6 Nanofibers

A thin film of TiO2 was deposited onto free-standing PA-6 nanofiber mat using atomic

layer deposition (ALD) before growing Zr-based MOF coatings. A custom-built hot-wall

viscous-flow reactor that has been described previously was used for ALD TiO2 processes.4

In a typical ALD TiO2 cycle, titanium (IV) chloride (TiCl4, 99%, STREM Chemicals) was

first dosed to the reactor chamber for 1 s, followed with 40 s of N2 (99.999%, Airgas, further

purified with an Entegris gatekeeper). After TiCl4 dose and N2 purge, deionized H2O was

dosed to the chamber for 1 s, and the chamber was subsequently purged with N2 for 60 s. The

process temperature was controlled at 100°C, and the pressure was about 1.3 Torr. All PA-6

nanofibers were coated with 100 cycles of ALD TiO2 before solvothermal MOF growth.

Synthesis of Zr-based MOF Coatings on Nanofibers

Precursor solutions for Zr-based MOF coatings (i.e. UiO-66, UiO-66-NH2 and UiO-67)

were prepared using the recipe and procedure described above for Zr-based MOF powders.

~10 mg of free standing PA-6 nanofiber mat coated with ALD TiO2 was transferred into a 20

mL scintillation vial and soaked in the MOF precursor solution. The vial was then placed in a

box furnace and heated to 85 °C for 24 h. After the solvothermal synthesis, the MOF-coated

PA-6 nanofiber mat was transferred into a fine aluminum mesh and washed with 100 mL of

181
DMF twice. After DMF wash, the MOF-coated nanofiber mat was further exchanged in

anhydrous ethanol, and the solvent was replaced every 12 h for a total of three times. To fully

activate the MOF coating, the MOF-coated nanofiber mat was first dried in a dessicator at

room temperature at reduced pressure for 12 h, and then in a BET at 80 °C for 12 hour and

subsequently at 110 °C for 12~18 h (heating temperature was slowly increased from 80 °C to

110 °C).

Material Characterization

Scanning electron microscopic (SEM) images were obtained using an FEI Verios

460L field emission SEM. All the samples were sputter-coated with Au-Pd (5~10 nm) before

imaging. Energy dispersive X-ray (EDX) mapping images was taken using an Oxford energy

dispersive X-ray spectrometer attached to the FEI Verios 460L FESEM. Cross sections of the

MOF-coated nanofibers were obtained via microtoming, and imaged using a JEOL 2010F

field emission transmission electron microscope (TEM). X-ray diffraction (XRD) was

performed using a Rigaku SmartLab X-ray diffraction tool (Cu Kα X-ray source) for

crystalline phase analysis. Glass sample holder was used for powder samples, and aluminum

holder was used for nanofiber samples. Powder diffraction patterns for Zr-based MOFs were

also simulated using Mercury 3.0 software based on the crystallographic information files

from Cambridge Crystallographic Data Centre (CCDC 733458 for UiO-66, CCDC 733458

for UiO-67, no CIF available for UiO-66-NH2). BET surface area of MOF powders was

measured with a Micromeritics ASAP 2020 surface area and porosimetry analyzer, and

MOF-coated nanofiber samples were analyzed using a Quantachrome Autosorb-1C surface

182
area and pore size analyzer. Samples were dried in vacuum at 80 °C for 12 hour and

subsequently at 110 °C for 12~18 h (heating temperature was slowly increased from 80 °C to

110 °C) before N2 adsorption measurement. BET surface area was calculated based on the

N2 adsorption data within a relative pressure range of P/Po = 0.02 ~ 0.07 (this range meets

the criteria for determining BET surface area of microporous materials described in

literature).5,6

Degradation of Dimethyl 4-Nitrophenyl Phosphate (DMNP)

The kinetics of catalytic degradation of dimethyl 4-nitrophenyl phosphate (DMNP) was

characterized using a procedure similar to the methods described in previous reports.7,8 2.5

mg of MOF powder sample or 14 mg of MOF-coated nanofiber sample was first dispersed in

1 mL of N-ethylmorpholine aqueous solution (0.45 M) in a 2 mL Eppendorf tube under fast

magnetic stirring (1100 rpm stir rate set on a Thermo Scientific stir plate) for 20 min. 4 μL

(~5.6 mg, 0.023 mmol) of DMNP was then added to the MOF suspension. The Eppendorf

tube was kept on the stir plate (1100 rpm stir rate) during the reaction. A 20 μL aliquot was

taken from the reaction mixture each time and diluted in 10 mL of N-ethylmorpholine

aqueous solution (0.15 M) for UV-visible absorbance spectroscopy. The reaction progress

was evaluated by monitoring the p-nitrophenoxide (degraded product from DMNP)

absorbance at 407 nm. The concentration of p-nitrophenoxide was calculated based on

Lambert-Beer Law (Equation S1).


𝐼
A = −lg = 𝜀𝐶𝑙 (S1)
𝐼0

183
Where A is the absorbance in units of absorbance unit (a.u.), I0 is the incident light intensity,

I is the transmitted light intensity, 𝜀 is the molar absorptivity coefficient

(𝜀 = 18330 𝑀−1 𝑐𝑚−1 ), C is the analyte concentration in unites of M, l is the length of light

path in units of cm.

The percent conversion of DMNP was obtained from the ratio of the p-nitrophenoxide

concentration before dilution to the initial DMNP concentration in the reaction mixture.

Degradation of O-Pinacolyl Methylphosphonofluoridate (GD)

Approximately 14 mg of MOF-coated PA-6@TiO2 nanofibers was weighed on a Mettler

Toledo AB104 calibrated digital scale. Once weighed, the samples were placed into a

humidity chamber (Thunder Scientific Corporation Series 2500) at 25oC to condition them at

50% relative humidity (RH) for at least 16 h. Upon humidification completion, the samples

were transferred into a 4-mm glass NMR tube. 700 µL of deionized water and 47 µL of

N-ethylmorpholine (buffer) were then added to the NMR tube. After 2.6 µL of O-pinacolyl

methylphosphonofluoridate (GD) was added onto the inner wall of the NMR tube, the tube

was capped and vigorously shaken in order to ensure mixture of GD with MOF-nanofibers in

the solution. The time of solution agitation was recorded in order to account for the Δ𝑡 of mix

31
time vs initial scan time. P NMR spectra were obtained over time at ambient temperature

(24-25°C) using a Varian INOVA 400 NB NMR spectrometer equipped with a Doty

184
Scientific 4-mm Liquid NMR probe to monitor the reaction and to identify the products.

The samples were not spun in order to limit slurry formation and obtain good reproducibility.

Samples scan times employed using a 90o pulse widths of 4μs and an overall acquisition time

of 8 min per acquisition. We noticed that the half-lives of GD in the reactions without

N-ethylmorpholine buffer exceed 300 min. Under same reaction conditions, the half-lives of

GD are less than 4 min with the presence of N-ethylmorpholine. We speculate that the HF

formed during the reaction is likely to poison the active sites on the MOF coatings, while the

N-ethylmorpholine buffer can neutralize HF and maintain the catalytic properties of the

MOFs.

185
Figure S5.1 SEM images of UiO-66-NH2 grown on (a-b) untreated PA-6 nanofibers and

(c-d) ALD TiO2 coated PA-6 nanofibers. Depositing a thin ALD TiO2 nucleation layer

significantly improves the growth uniformity and crystal coverage on the fiber surface.

186
Figure S5.2. SEM images of (a-b) electrospun PA-6 nanofibers and (c-d) ALD TiO2 coated

PA-6 nanofibers. The average diameter of untreated PA-6 nanofibers measured from SEM

images is 37±16 nm.

187
Table S5.1 MOF mass fraction in MOF-nanofiber composites calculated from BET results

and ICP-OES analysis

Overall MOF Substrat MOF Measure Measure MOF


BET Surfac e wt% d Zr wt% d Zr wt% wt%
Surfac e Area Surface Calculate in in MOF Calculate
e Area (m2/g) Area d from composit (%) d from
(m2/g) (m2/g) BET e (%) ICP
Results Results
(%) (%)
PA-6@TiO2
143.9 1512 11.5 8.82 2.81 33.0 8.51
@UiO-66
PA-6@TiO2
@UiO-66-N 205.9 1334 11.5 14.7 4.80 29.7 16.2
H2
PA-6@TiO2
356.2 2250 11.5 15.4 4.51 27.2 16.6
@UiO-67

Previously, we reported the following equation to correlate the MOF mass fraction with

the overall BET surface per unit mass of MOF plus fibers:
𝑆𝐴𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 × 𝑚𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 −𝑆𝐴𝑓𝑖𝑏𝑒𝑟 × 𝑚𝑓𝑖𝑏𝑒𝑟
𝑆𝐴𝑀𝑂𝐹 ≈ (𝑆2)
𝑚𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 − 𝑚𝑓𝑖𝑏𝑒𝑟

where m is the mass and SA is the surface area for each component.10 With known surface

area for the uncoated fiber substrate, the MOF-coated fibers and the bulk MOF material, the

MOF mass fraction (ω) can be calculated using Equation X:

𝑆𝐴𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 = 𝑆𝐴𝑀𝑂𝐹 ∙ 𝜔 + 𝑆𝐴𝑓𝑖𝑏𝑒𝑟 ∙ (1 − 𝜔) (𝑆3)

𝑆𝐴𝑀𝑂𝐹+𝑓𝑖𝑏𝑒𝑟 −𝑆𝐴𝑓𝑖𝑏𝑒𝑟
or 𝜔= (𝑆4)
𝑆𝐴𝑀𝑂𝐹 − 𝑆𝐴𝑓𝑖𝑏𝑒𝑟

We found that the MOF weight percentage calculated in this approach is similar to what

was measured by the ICP-OES method. This method provides a simple and straightforward

way to calculate the MOF weight percent when instruments for elemental analysis are not

188
available. One prerequisite is that both the MOF coatings on the fibers and the bulk MOF

powder are fully activated for BET measurements. It is also recommended to use the MOF

crystals collected from the same batch for MOF synthesis on fibers for BET analysis in order

to obtain a more accurate MOF mass fraction.

189
Figure S5.3 DMNP percent conversion as a function of time during the hydrolysis with

untreated PA-6 and ALD TiO2 coated PA-6 (PA-6@TiO2) nanofibers. Estimated t1/2 values

are 3950 min and 1170 min for PA-6 and PA-6@TiO2, respectively.

190
Figure S5.4 Kinetic analysis of DMNP degradation with (a-b) UiO-66, (c-d) UiO-66-NH2,

(e-f) UiO-67. The red curves in (a), (c), and (e) are plotted using the rate constants derived

from the linear fitting in (b), (d), and (f), respectively, based on the assumption of first order

reaction kinetics. We found that the reaction kinetics with UiO-66-NH2 does not fit well to

first-order rate equation.

191
Figure S5.5 Kinetic analysis of DMNP degradation with (a-b) PA-6@TiO2@UiO-66, (c-d)

PA-6@TiO2@UiO-66-NH2, (e-f) PA-6@TiO2@UiO-67. The red curves in (a), (c), and (e)

are plotted using the rate constants derived from the linear fitting in (b), (d), and (f),

respectively, based on the assumption of first order reaction kinetics. The reaction kinetics

with PA-6@TiO2@UiO-66-NH2 does not fit well to first-order rate equation.

192
Table S5.2 Kinetic constants and half-lives of DMNP hydrolysis with UiO MOF powders

and UiO MOF thin films on PA-6@TiO2 nanofibers.

k Calculated Observed MOF wt% TOF

(min-1) t1/2 (min) t1/2 (min) (s-1)

UiO-66 0.046 15 22 100 0.0054

UiO-66-NH2 0.23 3.0 2.8 100 0.045

UiO-67 0.076 9.1 7.7 100 0.020

PA-6@TiO2@UiO-66 0.0061 113 135 8.8 0.0018

PA-6@TiO2@UiO-66-NH2 0.078 8.9 7.3 14.7 0.021

PA-6@TiO2@UiO-67 0.097 7.1 7.4 15.4 0.024

*The amounts of catalysts used in DMNP degradation are 2.5 mg for UiO MOF powders and

14 mg for MOF-nanofiber composites. Turnover frequency (TOF) was calculated at t1/2,

although initial reaction rates give slightly higher TOF values.

193
Figure S5.6 SEM images and EDX spectra of (a-b) PA-6@TiO2@UiO-66 nanofibers, (c-d)

PA-6@TiO2@UiO-66 nanofibers and (e-f) PA-6@TiO2@UiO-67 nanofibers after DMNP

degradation experiment. SEM images and EDX results confirm that significant amounts of

MOF coatings remain in the MOF-nanofiber composites even after strong agitation during

the DMNP hydrolysis tests.

194
31
Figure S5.7 P NMR spectra of GD during hydrolysis with PA-6@TiO2@UiO-66

nanofibers.

195
31
Figure S5.8 P NMR spectra of GD during hydrolysis with PA-6@TiO2@UiO-66-NH2

nanofibers.

196
31
Figure S5.9 P NMR spectra of GD during hydrolysis with PA-6@TiO2@UiO-67

nanofibers.

197
(1) Schaate, A.; Roy, P.; Godt, A.; Lippke, J.; Waltz, F.; Wiebcke, M.; Behrens, P. Chem. –
Eur. J. 2011, 17 (24), 6643–6651.

(2) Katz, M. J.; Brown, Z. J.; Colon, Y. J.; Siu, P. W.; Scheidt, K. A.; Snurr, R. Q.; Hupp, J.
T.; Farha, O. K. Chem. Commun. 2013, 49 (82), 9449–9451.

(3) Andrady, A. L.; Ensor, D. S. Electrospinning in a controlled gaseous environment. US


Patent 7,297,305.

(4) Zhao, J.; Gong, B.; Nunn, W. T.; Lemaire, P. C.; Stevens, E. C.; Sidi, F. I.; Williams, P.
S.; Oldham, C. J.; Walls, H. J.; Shepherd, S. D.; et al. J. Mater. Chem. A 2015, 3 (4),
1458–1464.

(5) Walton, K. S.; Snurr, R. Q. J. Am. Chem. Soc. 2007, 129 (27), 8552–8556.

(6) Thommes, M.; Kaneko, K.; Neimark, A. V.; Olivier, J. P.; Rodriguez-Reinoso, F.;
Rouquerol, J.; Sing, K. S. W. Pure Appl. Chem. 2015, 87 (9–10), 1051–1069.

(7) Katz, M. J.; Mondloch, J. E.; Totten, R. K.; Park, J. K.; Nguyen, S. T.; Farha, O. K.;
Hupp, J. T. Angew. Chem. Int. Ed Engl. 2014, 53 (2), 497–501.

(8) Mondloch, J. E.; Katz, M. J.; Isley Iii, W. C.; Ghosh, P.; Liao, P.; Bury, W.; Wagner, G.
W.; Hall, M. G.; DeCoste, J. B.; Peterson, G. W.; et al. Nat. Mater. 2015, 14 (5), 512–
516.

(9) Peterson, G. W.; Moon, S.-Y.; Wagner, G. W.; Hall, M. G.; DeCoste, J. B.; Hupp, J. T.;
Farha, O. K. Inorg. Chem. 2015, 54 (20), 9684–9686.

(10) Zhao, J.; Losego, M. D.; Lemaire, P. C.; Williams, P. S.; Gong, B.; Atanasov, S. E.;
Blevins, T. M.; Oldham, C. J.; Walls, H. J.; Shepherd, S. D.; et al. Adv. Mater.
Interfaces 2014, 1400040.

198
CHAPTER 6 is a reprint of a manuscript in preparation.

CHAPTER 6. Diffusion of CO2 in large single crystals of Cu-BTC

MOF

Trenton M. Tovar, †,§ Junjie Zhao‡,§William T. Nunn‡, Heather F. Barton‡,

Gregory W. Peterson#, Gregory N. Parsons‡,*, and M. Douglas LeVan†,*


Department of Chemical and Biomolecular Engineering, Vanderbilt University, Nashville,

Tennessee 37212, U.S.A. (m.douglas.levan@vanderbilt.edu)



Department of Chemical and Biomolecular Engineering, North Carolina State University,

911 Partners Way, Raleigh, NC 27695, U.S.A (E-mail: gnp@ncsu.edu)


#
Edgewood Chemical Biological Center, 5183 Blackhawk Road, Aberdeen Proving Ground,

MD 21010, U.S.A.

§
T. M. Tovar and J. Zhao contributed equally to this work.

199
Abstract

Carbon dioxide adsorption in metal-organic frameworks has been widely studied for

applications in carbon capture and sequestration. A critical component that has been largely

overlooked is the measurement of diffusion rates. This paper describes a new reproducible

procedure to synthesize millimeter-scale Cu-BTC single crystals using concentrated reactants

and an acetic acid modulator. Microscopic images, XRD patterns, BET surface area, and

TGA results all confirm the high quality of these Cu-BTC single crystals. The large crystal

size aids in the accurate measurement of micropore diffusion coefficients.

Concentration-swing frequency response performed at varying gas-phase concentrations

gives diffusion coefficients that show very little dependence on loading up to pressures of 0.1

bar. The measured micropore diffusion coefficient for CO2 in Cu-BTC is 1.7 x 10-9 m2/s.

6.1 Introduction

The continuously increasing amount of CO2 emissions due to anthropogenic activities

has caused a sharp rise of CO2 levels in the atmosphere.1 While the search for alternative

clean energy sources continues, developing new technologies for CO2 capture and

sequestration is still necessary and has significant impact for controlling global warming.

Metal-organic frameworks (MOFs) are highly porous crystalline networks containing metal

clusters interconnected by organic linkers.2 MOFs typically have high surface area that

exceeds traditional adsorbent materials.3-7 In addition, the numerous options of metal centers

and linkers enable rational design and synthesis of MOF structures to achieve enhanced

adsorption capacities and catalytic properties. For example, MOFs have shown promising

200
performance for CO2/N2 separation for post-combustion CO2 removal from flue gas, CO2/H2

separation for pre-combustion capture, CO2/CH4 separation for natural gas upgrading, and

direct capture of CO2 from air.1,8,9 MOFs also show excellent catalytic activities for

converting CO2 to valuable compounds.10-13 As a result, MOFs are promising candidates for

CO2 capture and sequestration.

Despite the amount of research on CO2 adsorption and separations and catalysis using

MOFs, a major deficiency in the literature is the measurement of diffusion rates, a critical

property that can significantly affect the overall performance for the abovementioned

applications. CO2 diffusion constants have been reported for a few MOFs. Salles et al.14

measured diffusion rates ranging from 10-8 to 10-9 m2/s over a range of loadings on

MIL-47(V). Sabouni et al.15 reported 7x10-12 m2/s at 298 K for CPM-5, and Saha et al.16

reported diffusion constants on the order of 10-9 m2/s for MOF-177. Even for the same MOF

(MOF-5), the reported diffusion coefficients show a wide range from 10-9 m2/s to 10-8

m2/s.16,17

A big challenge for obtaining accurate diffusion coefficients is to decouple the presence

of different mass transfer mechanisms related to adsorbent morphology. For example, the

impact of surface barriers on MOF thin films was explored by Heinke et al.18 Fletcher et al.19

found a linear driving force behavior for CO2 in Ni2(4,4’-bipy)3(NO3)4, but macropore

diffusion was determined to be the controlling resistance for Co/DOBDC and Ni/DOBDC

pellets by Hu et al.20 using zero-length column techniques and for Cu-BTC pellets by Liu et

al.21 using concentration-swing frequency response (CSFR).

In this paper, we report a novel method to measure CO2 micropore diffusion coefficients

201
using CSFR to analyze millimeter-scale MOF single crystals. We use Cu-BTC (HKUST-1) as

a model to demonstrate this approach, as it shows one of the highest CO2 capacities (at

hydrated conditions) among all MOFs.9 High selectivity for CO2/H2 and CO2/N2 separations

have also been reported for this MOF.22-25

We developed a new synthesis technique to achieve mm-scale Cu-BTC single crystals,

and the crystal sizes can be well controlled. Microscopic images, XRD patterns, BET surface

area and TGA results were used to characterize the quality of these Cu-BTC single crystals.

Diffusion of CO2 is measured by the CSFR technique to determine rate controlling

mechanisms and diffusion constants, and rate behavior is compared with that in an activated

carbon.

6.2 Experimental Methods

Materials

Commercial adsorbents used in this study were Cu-BTC powder purchased from Sigma

Aldrich and BPL activated carbon (lot no 4814-5) in 6x16 mesh form obtained from Calgon

Carbon Corp. Research grade helium, pure CO2, and 1% CO2 in He were purchased from Air

Liquide.

Cu-BTC Single Crystal Synthesis

A low temperature solvothermal method was used to synthesize large Cu-BTC single

crystals. 0.49 g copper nitrate trihydrate (Cu(NO3)2∙3H2O) was dissolved in 3 mL of

deionized water. 0.24 g benzene-1,3,5-tricarboxylic acid (H3BTC) was dissolved in 3 mL of

202
ethanol (slight heating was needed to fully dissolve this organic linker). The Cu(NO3)2

solution was first mixed with 3 mL of N,N-dimethylformamide (DMF) in a 20 mL

scintillation vial. The H3BTC solution and 12 mL of glacial acetic acid (modulator) were

subsequently added to the mixed solution as well. The scintillation vial was then placed in a

furnace where the reaction was allowed to proceed at 55 °C for 3 days. After the

solvothermal synthesis, the Cu-BTC crystals were removed by hand from the glass surface of

the vial. All the crystals were soaked in ethanol for at least 3 days for solvent exchange.

Crystal Characterization

Cu-BTC single crystals were imaged using an Olympus BX60 optical microscope

equipped with a ProgRes C5 camera. Optical microscopic images were captured and

analyzed with ProgRes® CapturePro 2.8.8 software. SEM images of Cu-BTC crystal were

with an FEI Phenom® bench-top SEM. Crystals were mounted on conductive carbon taps

and sputter-coated with 5-10nm of Au-Pd before imaging. X-ray diffraction patterns of the

Cu-BTC single crystals were collected using a PANalytical Empyrean X-ray diffractometer.

BET surface area of Cu-BTC single crystals and powders was measured using a

Quantachrome Autosorb-1C surface area and pore size analyzer. 10 mg of Cu-BTC was

vacuum dried at room temperature for 12-18 h before the measurement of N2 adsorption

isotherm. BET surface area was calculated based on the isotherm within the P/Po range of

0.02-0.10. Thermogravimetric analysis (TGA) was measured with a Discovery TGA from TA

Instruments.

Pure CO2 isotherms were measured using a Rubotherm gravimetric analyzer. A sample

203
was degassed at 25 °C for approximately 12 hours using a turbo molecular pump until no

weight change was detected. The sample was then dosed with CO2 at a specified pressure and

allowed to equilibrate until no more weight change was detected. After equilibration, the CO2

pressure was increased to the next specified value. This process was repeated for the pressure

range of the isotherm, from 0.5 to 800 mmHg. The CO2 loading was calculated by the weight

gain of the sample at each pressure step.

Diffusion Measurement

CO2 diffusion was measured using a concentration-swing frequency response apparatus

(Figure S1). Approximately 15-20 mg of adsorbent was regenerated for 8 hours under

vacuum (at 25 °C for Cu-BTC cyrstals, at 250 °C for BPL activated carbon) for accurate

weight measurement. The sample was then loaded into a shallow bed and placed under

vacuum at room temperature with a 1 sccm He flow for 16 hours. Then a CO2 stream was

mixed with the He stream upstream of the adsorbent bed. Both streams were controlled with

MKS mass flow controllers, and the ratio of the two flow rates determined the gas-phase

concentration. The pressure in the adsorbent bed was controlled at 1 bar by a MKS Baratron

pressure controller, and effluent gas from the adsorbent bed was sampled by an Agilent 5975

mass spectrometer.

To run an experiment, the system was allowed to reach steady-state as determined by a

constant signal from the mass spectrometer. Then the mass flow controllers were used to

introduce sinusoidal perturbations to the flowrates of each stream but 180 degrees out of

phase. This resulted in a feed stream to the adsorbent bed that has a constant flow rate with a

204
sinusoidal concentration swing around the desired steady-state gas-phase concentration.

The perturbations were performed at different frequencies in the range of 0.001 to 0.1 Hz.

The collected data are presented as plots of the amplitude ratio (AR) as a function of the

perturbation frequency, where the AR is calculated from the amplitude of the gas exiting the

adsorbent bed divided by the amplitude of the gas entering the adsorbent bed. The diffusion

mechanism and rate parameters can be extracted by fitting the data to a mathematical model

derived from transfer functions. Detailed mathematical models for a CSFR apparatus have

been described elsewhere.26,27

6.3 Results and Discussion

Crystal Characterization.

Figures 6.1c-f show the optical micrographs and SEM images of the Cu-BTC single

crystals. Truncated cube and truncated octahedron shapes were observed for this fcu-type

MOF crystal. The octagon (or square) facets represent the planes viewed down the <100>

directions, while the hexagon-shape facets correspond to the planes along the <111>

directions. The size of Cu-BTC crystals obtained from a 3-day synthesis ranges from 500 μm

to 1.3 mm (Figure S6.2). These crystals were sorted by size for later CO2 diffusion studies.

The quality of the Cu-BTC single crystals was characterized using XRD, BET and TGA.

Figure 6.2 shows the XRD patterns for the crystal planes parallel to (100) and (111)

respectively. The peak positions match well with the corresponding peaks present in the

simulated powder diffraction pattern and reported powder patterns in the literature.37 The fact

that only peaks associated with targeted parallel crystal planes appear in the diffraction

205
patterns and that optical micrographs show negligible defects confirm the formation of

high-quality single crystals.

Figure 6.1 (a-b) Molecular representation of the Cu-BTC MOF along the [100] direction

and [111] direction. Color code: Cu (yellow); O (red); C (black); H (not shown). (c-d)

Optical microscopic and (e-f) SEM images of Cu-BTC single crystals.

206
BET surface area and pore volume for the Cu-BTC single crystals were measured and

compared with the Cu-BTC fine powder prepared from 85 °C solvothermal synthesis (Figure

S6.3). The surface area is 1980 and 1825 m2/g for the single crystals and the fine powder

respectively. Pore volumes are 0.85 and 0.79 cm3/g for the large crystals and powder samples

respectively. Both the surface area and pore volume compare well with values from the

literature.22

Figure 6.2 XRD patterns of a Cu-BTC single crystal for the planes parallel to (100) (black)

and the planes parallel to (111) (red), and simulated powder diffraction pattern for Cu-BTC.

TGA was used to evaluate the thermal stability of the Cu-BTC single crystals (Figure

S6.4). Cu-BTC powder shows a steeper initial mass drop compared to the large crystals due

to the shorter diffusion length in small particles for solvent evaporation. We also noticed that

the mm-scale crystal exhibits a sharper mass decrease than the fine powder in the

207
decomposition regime (300-385 °C), possibly because of slower heat transfer and less

impurities in the large crystals. Our measured decomposition temperature regime also agrees

well with reported values,29,30 indicating the good quality of these MOF crystals.

Crystal Growth Mechanism

The crystal size was measured after specific growth times (1-5 d), and Figure 6.3 shows

that the size of the largest crystal increases as a function of synthesis time. No crystals form

during the nucleation period (1 d), and the crystal size quickly ramps up to ~1.3 mm at 3 days.

After 3 days, the change in crystal size is very small, indicating the termination of crystal

growth. We used a Gompertz function (Equation S1), a common mathematical model to

describe such growth behavior, and the data fit well to the model. Taking the mass-size

correlation into account (Figure S6.5 and Equation S2), this growth function can be

converted into a plot similar to the classic La Mer diagram (Figure S6.6).31

At the beginning of the synthesis, secondary building units (SBUs) of Cu-BTC MOF

start to form in the solution. As the reaction proceeds and the SBU concentration reaches

supersaturation (after ~1 d), heterogeneous nucleation occurs on the wall of the scintillation

vial as it is more thermodynamically favorable. The crystal growth rate depends on the SBU

concentration and stops once the concentration drops below the supersaturation point (after

~4 d).

While previous work reported the use of nitric acid to inhibit the deprotonation of

H3BTC linker and thus slow down the crystal growth rate in order to achieve large single

crystals,32 our method is more effective and reproducible by using a supersaturated

208
concentration of MOF reactants and an acetic acid modulator. The formation of an acetate

complex competes with the generation of SBU in the solution, allowing the slow and steady

growth of our single crystals. By avoiding high acidity of the precursor solution, we also

obtained high crystal quality as shown above.

Figure 6.3 Increase of Cu-BTC crystal size as a function of reaction time.

Diffusion Measurements

To ensure that CO2 adsorption in the single crystals is similar to Cu-BTC powder, a CO2

isotherm was measured and compared with literature values. Figure 6.4 shows the CO2

isotherm measured at 25 °C on Cu-BTC crystals of various sizes. The isotherm fits well to a

Toth model.33 Figure 6.4 also shows some of the representative CO2 isotherms reported in the

literature. We plotted these curves based on parameters for a Sips fit from Aprea et al.34 and a

Langmuir fit from Hamon et al.35 Select data points at 1 bar were also adapted from Yazaydin

et al.,22 Millward and Yaghi,36 and Wang et al.37 While most isotherms in the literature were

209
measured at high pressure, due to the large capacity of Cu-BTC, CO2 isotherms at low

pressures are important for comparison to diffusion measurements at low CO2 concentrations.

It is clear that the isotherm for the large Cu-BTC single crystals compares favorably with the

MOF powder at low CO2 pressure.

Figure 6.4 Gravimetric CO2 isotherm on large Cu-BTC single crystals at 298 K fit by a

Toth isotherm model. Also shown are data points from literature isotherms of CO2 on powder

Cu-BTC at 298 K.

Figure 6.5 compares the CSFR curves for the Cu-BTC single crystals and the powder at

a gas-phase concentration of 0.5% CO2 and total pressure of 1 bar. The CSFR curve for the

Cu-BTC powder has a steeper slope than the single crystals, indicating faster CO2 uptake in

the powder. This is reasonable because the diffusion length in the powder (crystal size < 50

μm) is much shorter than for the mm-scale single crystals. We used the micropore diffusion

model reported previously to fit the curves,26 and found that this model describes the single

crystals well but shows noticable deviation from the powder data. The fact that the micropore

210
diffusion model fails to fit the powder data well is possibly because other diffusive

mechanisms, such as surface barriers or external mass transfer, may also be involved in

addition to micropore diffusion. For the mm-scale Cu-BTC single crystals, micropore

diffusion is the dominant factor that limits mass transfer rates, ensuring that the measured

diffusion coefficients truly represent the diffusion in the MOF micropores. Therefore, single

crystals are excellent candidates for micropore diffusion measurements.

1.0

0.8
Amplitude Ratio

0.6

0.4
Cu-BTC crystals
Cu-BTC powder
0.2

0.0 2 4 6 8 2 4 6 8 2 4 6 8
0.001 0.01 0.1 1
Frequency (Hz)

Figure 6.5 Amplitude ratio curves for CSFR experiments on large Cu-BTC single crystals

versus a powder sample.

The micropore diffusion model uses two parameters to fit the CSFR curves; the

regressed values are the local isotherm slope, K, and the micropore diffusion coefficient

parameter Ds/R2, where R is the diffusion length scale. As the regressed diffusion coefficient

211
is dependent on particle size, accurate diffusion measurements require precise knowledge of

the crystal size, which is well controlled by the synthesis technique. Three sets of Cu-BTC

crystals with particle sizes of 0.7±0.1 mm, 1.0±0.1 mm, and 1.3±0.1 mm were prepared,

where particle size corresponds to the longest crystal dimension. These were used for CSFR

measurements using 0.5% CO2 concentration. Figure 6 shows the CSFR curves for each set

of crystals, and the fitted parameters for the micropore diffusion model are given in Table 1.

The regressed isotherm slopes show good agreement between the different crystal sizes. The

Ds/R2 values decrease as the crystal size increases, as expected.

Figure 6.6 CSFR curves for different Cu-BTC crystal sizes at 0.5% CO2 concentrations.

The micropore diffusion coefficients are calculated by multiplying the regressed

diffusion parameter by the square of the diffusion length scale. However, the model assumes

a spherical particle shape that does not accurately represent the crystals. We found that the

height of crystals is approximately half of the longest dimension (L), indicating that the

212
effective diffusion length should be smaller than L. We measured the average crystal mass

per crystal size and correlated it as a function of L (Figure S6.5). A power law fit the data

well with an exponent of 2.14 instead of the expected cubic function for a perfect sphere. To

account for the shape factor, an effective diffusion length (Reff) was defined as

4
mass   Reff3  L2meas
.14
(1)
3

where Lmeas (in units of mm) is the length of the longest crystal dimension, ρ (in units of

g/cm3) is the crystal density, α = 0.52 mg∙mm-2.14. For single crystals, the measured density is

1.29 g/cm3, similar to reported values for Cu-BTC.1,10 When normalized by Reff, the

micropore diffusion coefficients agree fairly well and give an average value of 0.0017 mm2/s

(or 1.7 x 10-9 m2/s). This value is on the same order of magnitude as several other CO2

diffusion rates reported in other MOFs.

Table 6.1 Micropore diffusion fitting parameters for various Cu-BTC crystal sizes at 0.5%

CO2.

Crystal Size K Ds/R2 Ds


(mm) (mol/kg∙bar) (1/s) (mm2/s)
0.7 7.1 0.011 0.0014
1 7.1 0.0082 0.0017
1.3 6.6 0.0066 0.0020

An important aspect of diffusion in nanoporous materials is the impact of surface

concentration on diffusivity. Figure 6.7 shows the CSFR curves measured at various

gas-phase concentrations using 18 mg of Cu-BTC single crystals with diameters of ~0.7 mm.

The steady-state CO2 concentrations used in this series of experiments were 0.1%, 0.5%, and

213
10%. We found that the micropore diffusion model accurately fits the data at each tested CO2

concentration. The curves are nearly identical for the 0.1 and 0.5% experiments while the 10%

curve is shifted to higher amplitude ratios indicating a decrease in the local isotherm slope.

Figure 6.7 CSFR curves for various CO2 concentrations on Cu-BTC crystals of

approximately 1.5 mm.

Table 2 summarizes the fitted parameters for the CSFR curves at each concentration as

well as for a similar set of experiments on BPL activated carbon (Figure S7), a

predominantly microporous material that follows the micropore diffusion model well for

various adsorbates.38 For comparison, we also calculated the K values from the Toth fit of the

CO2 isotherm for the Cu-BTC crystals (Figure 4) and list them in Table 2. The K values

measured from CSFR experiments are slightly higher than those from the CO2 isotherm, but

the trends are very similar. For the 0.1% and 0.5% experiment, the isotherm slopes are

similar, which agrees with the near linear shape of the CO2 isotherm at such low

214
concentrations. For the 10% experiment, the K value significantly decreases for both the

CSFR and adsorption isotherm measurements.

Table 6.2 Micropore diffusion fitting parameters for 0.7 mm Cu-BTC single crystals and

BPL activated carbon at different CO2 concentrations.

Cu-BTC BPL Cu-BTC


CSFR CSFR isotherm
Conc K Ds/R2 K Ds/R2 K
% (mol/kg∙bar) (1/s) (mol/kg∙bar) (1/s) (mol/kg∙bar)
0.1 6.9 0.0090 4.9 0.0024 5.9
0.5 6.7 0.0086 4.2 0.0072 5.8
10 5.8 0.0089 2.8 0.0142 5.1

It is very interesting that the diffusion coefficients for Cu-BTC single crystals and BPL

activated carbon particles show completely different trends. For the single crystal Cu-BTC

samples there is no significant difference from tests using CO2 concentrations from 0.1% to

10%. However, for BPL activated carbon, the diffusion coefficient increases dramatically as

a function of CO2 concentration. These different trends are reasonable since the Cu-BTC

single crystals have narrow distribution of pore sizes, ordered lattice structure, and therefore,

more homogenous adsorption sites. Thus, the adsorbate-adsorbent interaction at low partial

pressure remains similar at increased CO2 pressures and loadings. This agrees well with

literature, as Wang et al.37 reported the heat of adsorption of CO2 remains virtually constant

as loading increases after an initial decrease. In contrast, the BPL activated carbon is very

heterogeneous, so the first CO2 molecules will associate with stronger adsorption sites and

will diffuse more slowly than at higher pressures.

215
In comparison with the MOF, the BPL activated carbon particles were larger,

approximately 2 mm in diameter. After normalizing by the diffusion length scale, the

diffusion coefficient for the single MOF crystals is significantly slower than for the activated

carbon. This could be due to strong intermolecular interaction between CO2 and the open

metal sites of Cu-BTC. It could also be due to the small uniform 9 Å pores of Cu-BTC,

whereas the activated carbon has a wide distribution of pore sizes.

6.4 Conclusion

A new synthesis procedure to produce large millimeter size single crystals of Cu-BTC

has been reported. Crystal size could be accurately controlled by length of the synthesis time.

Properties measured by XRD, BET, TGA, and CO2 isotherms on the Cu-BTC crystals agree

well with Cu-BTC powders produced by traditional solvothermal synthesis. The increased

length scale of the microporous domain makes diffusion measurements easier as it limits the

impact of multiple mass transfer mechanisms. CSFR experiments were performed which

verified micropore diffusion as the rate controlling diffusive mechanism. Diffusion

coefficients measured at varying concentrations show very little dependence on loading up to

about 0.1 bar, the highest pressure considered. At 0.5% gas-phase concentration of CO2 in He

at 1 bar, the measured diffusion coefficient is approximately 1.7 x 10-9 m2/s.

216
References

(1) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm, Z. R.;
Bae, T.-H.; Long, J. R. Chem. Rev. 2012, 112 (2), 724-781.

(2) Kim, J.; Chen, B.; Reinike, T. M.; Li, H.; Eddaoudi, M.; Moler, D. B.; O’Keeffe, M.;
Yaghi, O. M. J. Am. Chem. Soc. 2001, 123 (34), 8239-8247.

(3) Furukawa, H.; Ko, N.; Go, Y. B.; Aratani, N.; Choi, S. B.; Choi, E.; Yazaydin, A. O.;
Snurr, R. Q.; O’Keeffe, M.; Kim, J.; Yaghi, O. M. Science 2010, 329 (5990), 424-428.

(4) Farha, O. K.; Yazaydin, A. O.; Eryazici, I.; Malliakas, C. D.; Hauser, B. G.; Kanatzidis,
M. G.; Nguyen, S. T.; Snurr, R. Q.; Hupp, J. T. Nat. Chem. 2010, 2, 944-948.

(5) Farha, O. K.; Eryazici, I.; Jeong, N. C.; Hausert, B. G.; Wilmer, C. E.; Sarjeant, A. A.;
Snurr, R. Q.; Nguyent, S. T.; Yazaydin, A. O.; Hupp, J. T. J. Am. Chem. Soc. 2012, 134
(36), 15016-15021.

(6) Grunker, R.; Bon, V.; Muller, P.; Stoeck, U.; Krause, S.; Mueller, U.; Senkovska, I.;
Kaskel, S. Chem. Commun. 2014, 50, 3450-3452.

(7) Wang, T. C.; Bury, W.; Gomez-Gualdron, D. A.; Vermeulen, N. A.; Mondloch, J. E.;
Deria, P.; Zhang, K.; Moghadam, P. Z.; Sarjeant, A. A.; Snurr, R. Q.; Stoddart, J. F.;
Hupp, J. T.; Farha, O. K. J. Am. Chem. Soc. 2015, 137 (10), 3585-3591.

(8) Li, J.-R.; Ma, Y.; McCarthy, M. C.; Sculley, J.; Yu, J.; Jeong, H.-K.; Balbuena, P. B.;
Zhou, H.-C. Coordin. Chem. Rev. 2011, 255, 1791-1823.

(9) Keskin, S.; van Heest, T. M.; Sholl, D. D. ChemSusChem, 2010, 3, 879-891.

(10) Wang, C.; Xie, Z; deKrafft, K. E.; Lin, W. J. Am. Chem. Soc. 2011, 133 (34),
13445-13454.

(11) Fu, Y.; Sun, D.; Chen, Y.; Huang, R.; Ding, Z.; Fu, X.; Li, Z. Angew. Chem.-Int. Ed.
2012, 51 (14), 3364-3367.

(12) Sun, D.; Fu, Y.; Liu, W.; Ye, L.; Wang, D.; Yang, L.; Fu, X.; Li, Z. Chem.-Eur. J. 2013,
19 (42), 14279-14285.

(13) Wang, S.; Yao, W.; Lin, J.; Ding, Z.; Wang, X. Angew. Chem.-Int. Ed. 2014, 53 (4),
1034-1038.

217
(14) Salles, F.; Jobic, H.; Devic, T.; Llewellyn, P. L.; Serre, C.; Ferey, G.; Maurin, G. ACS
Nano 2010, 4 (1), 143-152.

(15) Sabouni, R.; Kazemian, H.; Rohani, S. Micropor Mesopor Mater. 2013, 175, 85-91.

(16) Saha, D.; Bao, Z.; Jia, F.; Deng, S. Environ. Sci. Technol. 2010, 44, 1820-1826.

(17) Zhao, Z.; Li, Z.; Lin, Y. S. Ind. Eng. Chem. Res. 2009, 48, 10015-10020.

(18) Heinke, L.; Gu, Z.; Woll, C. Nature Communications 2014, 5, 4562, 1-6.

(19) Fletcher, A. J.; Cussen, E. J.; Prior, T. J.; Rosseinsky, M. J.; Kepert C. J.; Thomas, K. M.
J. Am. Chem. Soc. 2001, 123 (41), 10001-10011.

(20) Hu, X.; Brandani, S.; Benin A. I.; Willis, R. R. Ind. Eng. Chem. Res. 2015, 54,
5777-5783.

(21) Liu, J.; Wang, Y.; Benin, A. I.; Jakubczak, P.; Willis, R. R.; LeVan, M. D. Langmuir
2010, 26 (17), 14301-14307.

(22) Yazaydin, A. O.; Benin, A. I.; Faheem, S. A.; Jakubczak, P.; Low, J. J.; Willis R. R.;
Snurr, R. Q. Chem. Mater. 2009, 21, 1425-1430.

(23) Yang, Q.; Zhong, C. J. Phys. Chem. B 2006, 110, 17776-17783.

(24) Liu, B.; Smit, B. Langmuir 2009, 25 (10), 5918-5926.

(25) Yang, Q.; Xue, C.; Zhong C.; Chen, J.-F. AIChE J. 2007, 53 (11), 2832-2840.

(26) Wang, Y.; LeVan, M. D. Ind. Eng. Chem. Res. 2008, 47, 3121-3128.

(27) Wang, Y.; LeVan, M. D. AIChE J. 2011, 57 (8), 2054-2069.

(28) Zhao, J.; Nunn, W. T.; Lemaire, P. C.; Lin, Y.; Dickey, M. D.; Oldham, C. J.; Walls, H.
J.; Peterson, G. W.; Losego, M. D.; Parsons, G. N. J. Am. Chem. Soc. 2015, 137 (43),
13756–13759.

(29) Chowdhury, P.; Bikkina, C.; Meister, D.; Dreisbach, F.; Gumma, S. Micropor. Mesopor.
Mat. 2009, 117, 406-413.

(30) Liang, Z.; Marshall, M.; Chaffee, A. L. Energ. Fuel. 2009, 23, 2785-2789.

218
(31) Stock, N.; Reinsch, H.; Schilling, L.-H. CHAPTER 2. Synthesis of MOFs. In RSC
Catalysis Series; Llabrés i Xamena, F., Gascon, J., Eds.; Royal Society of Chemistry:
Cambridge, 2013; pp 9–30.

(32) Li, L.; Sun, F.; Jia, J.; Borjigin, T.; Zhu, G. Growth of large single MOF crystals and
effective separation of organic dyes. Crystengcomm 2013, 15 (20), 4094–4098.

(33) Do, D. D. Adsorption Analysis: Equilibrium and Kinetics, Imperial College Press:
London, 1998.

(34) Hamon, L.; Jolimaitre, E.; Pirngruber, G. D. Ind. Eng. Chem. Res. 2010, 49,
7497-7503.

(35) Aprea, P.; Caputo, D.; Gargiulo, N.; Iucolano, F.; Pepe, F. J. Chem. Eng. Data 2010, 55,
3655-3661.

(36) Millward, A. R.; Yaghi, O. M. J. Am. Chem. Soc. 2005, 127, 17998-17999.

(37) Wang, Q. M.; Shen, D.; Bulow, M.; Lau, M. L.; Deng, S.; Fitch, F. R.; Lemcoff, N. O.;
Semanscin, J. Micropor. Mesopor. Mat. 2002, 55, 217-230.

(38) Wang, Y.; Mahle, J. J.; Furtado, A. M. B.; Glover, T. G.; Buchanan, J. H.; Peterson G.
W.; LeVan, M. D. Langmuir 2013, 29, 2935-2945.

219
Supporting Information

Figure S6.1 Schematic of the concentration-swing frequency response apparatus.

220
Figure S6.2 Optical microscopic images of Cu-BTC crystals obtained from a three-day

solvothermal synthesis. The crystal size ranges from 500 μm to 1.3 mm. (Crystal size

measurements are labelled on the images.)

221
Figure S6.3 (a) N2 isotherms for Cu-BTC single crystals and fine powder. (b) BET surface

area and (c) pore volume calculated from N2 isotherms.

222
Figure S6.4 Thermogravimetric analysis (TGA) for Cu-BTC single crystals (red) and fine

powder (black).

223
Figure S6.5 The correlation between the mass per crystal and the crystal dimension. The

error bar indicates the standard deviation of the crystal sizes in a mass measurement group.

224
Gompertz function for fitting the growth curve:

𝐿 = 𝑎 ∗ exp(− exp(−𝑘(𝑡 − 𝑡𝑐 ))) Equation S1

Where L (mm) is the size of the large crystal obtained from particular reaction time, 𝑎 (mm)

is the pre-exponential factor, k (d-1) is the kinetic factor, t and tc (in units of d) are the growth

time and transition time (for fitting purpose), respectively. After fitting the data to the

Gompertz function, we get 𝑎 = 1.34 mm, tc = 1.78 d, k = 2.18 d-1.

Figure S4 shows the correlation between the mass and crystal dimension. We used a power

law to fit the data:

𝑚 = 𝜌 ∗ 𝐿𝑛 Equation S2

Where m (mg) is the mass per crystal, L (mm) is the size of the large crystal (crystal

dimension), 𝑛 is power, 𝜌 (𝑚𝑔 ∙ 𝑚𝑚−𝑛 ) is the density factor. After fitting, we get n = 2.14,

𝜌 = 0.52 𝑚𝑔 ∙ 𝑚𝑚−2.14 .

The reaction kinetics can be derived by substituting L in Eqn S2 with Eqn 1 and taking a

derivative of Eqn S2:

𝑑𝑚 𝑑𝐿
= 𝜌𝑛𝐿 𝑛−1 ∗
𝑑𝑡 𝑑𝑡

𝑑𝑚
= 𝜌𝑛𝐿 𝑛−1 ∗ 𝐿 ∗ {− exp[−𝑘(𝑡 − 𝑡𝑐 )]}′
𝑑𝑡

𝑑𝑚
= 𝜌𝑛𝐿 𝑛 ∗ {− exp[−𝑘(𝑡 − 𝑡𝑐 )]} ∗ (−𝑘)
𝑑𝑡
𝑑𝑚
= 𝜌𝑛𝐿 𝑛 ∗ exp[−𝑘(𝑡 − 𝑡𝑐 )] Equation S3
𝑑𝑡

225
Figure S6.6 Crystal mass increase rate (dm/dt) as a function of reaction time based on

Eqnation S3.

226
Figure S6.7 CO2 CSFR curves on BPL activated carbon at varying concentrations.

227
CHAPTER 7 is a reprint of a manuscript in preparation.

CHAPTER 7. ALD-Assisted Synthesis of ZnO 3D Hierarchical

Nanostructures for Enhanced Photocatalytic Dye Degradation

Junjie Zhao†, Abhijeet Bagal‡, Qiaoyin Yang‡, Xu Zhang‡ and Chih-Hao Chang‡,

Mark D. Losego† and Gregory N. Parsons†,*


Department of Chemical and Biomolecular Engineering, North Carolina State University,

911 Partners Way, Raleigh, NC 27695, U.S.A (E-mail: gnp@ncsu.edu)



Department of Mechanical and Aerospace Engineering, North Carolina State University,

911 Oval Drive, Raleigh, NC 27695, U.S.A.

228
Abstract

ZnO 3D hierarchical nanostructures can overcome limitations of free-standing

nano-slurries, and show promising photocatalytic performance. However, very few reports

have systematically investigated how surface–attached nanostructure dimensions affect

photocatalyic performance. In this work, interference lithography was used to fabricate

nano-pillars with controllable heights (250 ~ 800 nm) as templates for photocatalytic ZnO

coatings via atomic layer deposition (ALD). Conformal ALD ZnO thin films promote

uniform nucleation of ZnO nano-wires synthesized using hydrothermal method. The length

of nano-wires increase systematically with growth time (60 ~ 150 min). We find that taller

nano-pillars enhance the rate for photocatalytic degradation of methyl orange, and the

addition of hierarchical nanowires increases the first order degradation rate constant by up to

61%.

7.1 Introduction

Semiconductor-based photocatalysts have attracted tremendous interest over the past

few decades for organic synthesis, water splitting, and decontamination and disinfection

applications.1–4 ZnO is a II-VI semiconductor (large direct bad gap of 3.37eV at room

temperature)5 with similar photocatalytic activity compared to TiO2, and is often considered

as a possible alternative photocatalyst to TiO2 because of its low cost and non-toxic

property.4,6 Specifically in applications for waste water treatment, ZnO has shown promising

photocatalytic performance for degrading azo dyes,6,7 and other organic contaminants.8-10

While using ZnO nano-slurries as photocatalyst can achieve high reaction rate, particle

229
aggregation in solution and difficulty in separating catalyst after use for recycling are major

practical challenges.6,11 Immobilizing ZnO nanostructure photocatalysts on substrate

materials could potentially overcome these limitations.

Many thoughtful strategies have been reported for growing ZnO nanostructures onto

functional substrate materials.11-17 Among them, hierarchical nanostructures are of particular

interest as these structures can exhibit large surface area and light trapping properties.12 For

example, Gong et al., Lee et al. and Colmenares et al. reported ZnO nano-wires/nano-leaves

grown on ZnO seeded fiber surfaces using hydrothermal method.11,16,17 Hung et al.

fabricated ZnO pore-array films using nanosphere mask and electrochemical deposition, and

further decorated the structure with ZnO nano-wires and Ag nanoparticles.12 Xu et al.

synthesized ZnO nanoplate-nanowire hierarchical structures using eletrodeposition and

hydrothermal synthesis.13 Compared with flat ZnO surfaces, nano-wires, nano-plates or

nanosheets, higher photocatalytic activity has been reported for ZnO 3D hierarchical

nanostructures.11-13 However, because of challenges in systematic tuning of periodic

hierarchical nanostructures, very few reports have investigated the role of nanostructure

dimension on photocatalyic activity and overall reaction rates.

Here, we report a new interference lithography–based procedure to synthesize ZnO 3D

hierarchical nanostructures with controllable periodicity and feature size. A series of nano-pillar

structures with varied heights were coated with conformal ALD ZnO thin films. The pillars were

further decorated with hydrothermally grown ZnO nano-wires. The rate of photocatalytic

degradation of methyl orange (MO) was investigated vs feature height and nanowire length.

230
Figure 7.1 Schematic of the fabrication procedure for the ZnO 3D hierarchical

nanostructures. (a) Silicon wafer spun-coated with antireflection layer and positive

photoresist. (b) Periodic nano-pillars fabricated using laser interference lithography. (c)

Conformal ZnO thin films deposited onto the nano-pillars via ALD. (d) ZnO 3D hierarchical

nanostructure obtained from hydrothermal synthesis of ZnO nano-wires on ALD ZnO coated

nano-pillars.

7.2 Experimental Methods

Fabrication of Periodic Nano-Pillars

Periodic nano-pillar structures (500 nm periodicity) were patterned using laser

interference lithography (shown in Figure 7.1(a-b)).18 An antireflection coating (ARC

i-CON-16, Brewer Science, Inc.) and a layer of positive photoresist (Sumitomo PFi88A7)

were spun-coated on silicon wafers. The thickness of the antireflection coating was

231
controlled at 97 nm, while the thickness of the photoresist layer was tuned to control the

height of the nano-pillars (250 nm, 500 nm and 800 nm, respectively). A HeCd laser source

(wavelength = 325 nm) and a Lloyd’s mirror interferometer were used for interference

lithography. The periodic nano-pillar patterns were obtained after two orthogonal exposures

and post-exposure development, and the periodicity was controlled constant at 500 nm.

Atomic Layer Deposition of ZnO

ALD ZnO was deposited onto the nano-pillars (Figure 7.1c) using a customized hot-wall

viscous-flow vacuum reactor. The deposition pressure was controlled at ~2 Torr, and the

temperatures were kept at 100℃. In an ALD ZnO cycle, diethyl zinc (DEZ) was first dosed

to the chamber for 2 s, followed with 60 s of purge with dry N2 (99.999%, further filtered

through an Entegris gatekeeper). Deionized water was subsequently dosed to the chamber for

2 s, followed with another 60 s of N2 purge. 200 cycles of ALD ZnO were deposited onto the

nano-pillars as nucleation layers for hydrothermal synthesis of ZnO nano-wires.

Hydrothermal Growth of ZnO Nano-Wires

To prepare the precursor solution for the hydrothermal synthesis, 0.02 M zinc nitrate

hexahydrate (Zn(NO3)2·6H2O, 99%, Aldrich) and 0.02 M hexamethylene tetramine (C6H12N4,

99%, Aldrich) were mixed in 100 mL of deionize water in a 250 mL PTFE beaker.

232
Substrates with ZnO coated nano-pillars were transferred into the aqueous precursor solution.

The surface with nanostructures was kept facing down in the solution using a PTFE sample

holder during the synthesis. The PTFE beaker was covered with a piece of aluminum foil

(with several holes made on the foil), and placed in a furnace with a set reaction temperature

of 80 ◦C. The reaction time was controlled (60 min, 90 min, 120 min and 150 min) to tune the

lengths of nanowires. After hydrothermal growth, the samples were rinsed with deionized

water for 2 min and dried in N2 flow at room temperature.

Materials characterization

The thicknesses of ALD thin films on flat silicon wafers were measured with a

fixed-angle spectroscopic ellipsometer (J.A. Woollam Co., Inc). Contact angle of static water

on sample surfaces were measured using a Ramé-Hart Model 200 contact angle goniometer

with a CCD camera. At least three different spots were measured on each sample and the

average value of the contact angles was calculated. Scanning electron microscopic (SEM)

images of the ZnO 3D hierarchical nanostructures were taken using a field emission SEM

(JEOL JSM-6400F). Samples were sputter-coated with a thin layer of Au-Pd (5~10 nm)

before imaging. The dimension of the nano-pillars and nano-wires were measured using

ImageJ (1.48v) software.

233
Photocatalytic Characterization

ZnO nanostructure photocatalyst sample on silicon substrate (1  1 cm2) was placed into

a 50 mL Pyrex glass vial containing 30 mL of methyl orange (MO) in deionized water (15

mg/L). The solution was continuously stirred for 30 minutes in the dark to promote

equilibrium adsorption of the dye onto the catalytic surface. Photocatalysis was tested under

UV exposure (Intelli-Ray 400, Uvitron International, 245 μW/cm2) with continuous stirring

at constant temperature ~ 30 ◦C using a water bath. Every 30 min, 0.5 mL of solution aliquot

was collected to characterize MO concentration via UV–Vis absorption at 466 nm. After

each UV-Vis measurement, the sampled solution was added back to the Pyrex vial to

maintain constant reactant volume.

Analysis of Photocatalytic Reaction Kinetics

The photocatalytic degradation data was modeled using pseudo-first-order reaction

kinetics:

−𝑟𝑀𝑂 = 𝑘𝑎𝑝𝑝 𝐶𝑀𝑂 (1)

𝑑𝐶𝑀𝑂
𝑟𝑀𝑂 = (2)
𝑑𝑡

where 𝑟𝑀𝑂 (in units of 𝑚𝑜𝑙 ∙ 𝐿−1 ∙ ℎ−1 ) is the rate of the photocatalytic reaction,

𝑘𝑎𝑝𝑝 (in units of ℎ−1 ) is the apparent first order rate constant. 𝐶𝑀𝑂 (in units of 𝑚𝑜𝑙 ∙ 𝐿−1 )

is the concentration of MO during the reaction. t (in units of h) is the reaction time.

Replacing the 𝑟𝑀𝑂 in Equation 1 with 𝑑𝐶𝑀𝑂 /𝑑𝑡 and integrating the both sides of the

equation, we derived Equation 3:

234
𝐶𝑀𝑂
− 𝑙𝑛 ( ) = 𝑘𝑎𝑝𝑝 𝑡 (3)
𝐶0

where 𝐶0 (in units of 𝑚𝑜𝑙 ∙ 𝐿−1 ) is the initial concentration of MO. By plotting −𝑙𝑛(𝐶𝑀𝑂 /

𝐶0 ) as a function of t and linearly fitting the data, we obtained 𝑘𝑎𝑝𝑝 from the slope of the

fitted line. The half-life of MO (𝑡1/2 , in units of h) during the photocatalytic degradation is

given by:

𝑙𝑛2
𝑡1/2 = (4)
𝑘𝑎𝑝𝑝

7.3 Results and Discussion

Surface Wettability Tuned by ZnO ALD

Periodic nano-pillars were fabricated on silicon wafers using laser interference

lithography. Insert SEM image in Figure 7.2 shows the cylindrical morphology of these

nano-pillars. The nano-pillars used for wettability comparison have a periodicity of ~ 700 nm,

and the height of the pillars was controlled at 500 nm. Figure 7.2(a-b) compare the water

droplet contact angles of the nano-pillars before and after ZnO ALD. Because of the

hydrophobicity of the photoresist, the surface before ALD ZnO exhibits a contact angle of

103◦. In comparison, 200 cycles of ALD ZnO (~36 nm thick) improve the surface wettability

significantly, reducing the contact angle to less than 15◦. This wettability change is a result of

–OH terminal groups on the ALD ZnO surface, consistent with previous reports.19,20 The

improved surface wettability is important for the hydrothermal synthesis to achieve

conformal ZnO nano-wire growth on the nanostructure substrate.11

235
Figure 7.2 Water droplet contact angles of the periodic nano-pillars (500 nm high) (a)

before ZnO ALD and (b) after ZnO ALD (200 cycles). Insert SEM images represent the

nano-pillars before and after ZnO ALD, respectively. Scale bars in SEM images represent

1μm.

ZnO Nano-wire Growth on ALD Coated Nano-pillars

Figure 7.3(a-d) show SEM images of ALD ZnO coated nano-pillars before and after

hydrothermal synthesis at 80 °C with reaction time from 60 to 150 min. With the ALD ZnO

nucleation layer, conformal and uniform ZnO nano-wires were successfully grown onto the

nano-pillars, producing a hierarchical 3D structure. Using the SEM images in Figures 3(b-d),

the average nanowire length vs growth time is plotted in Figure 3e. The results show the

nano–wire length increases linearly with growth time, possibly with some delay during

growth initiation. This linear growth trend is promising for controlling the feature size of

these nano-wires and for tuning the void space between the nano-pillars within the

hierarchical nanostructure. It is important to note that the ZnO nano-wires maintained a

constant diameter (32.0±1.5 nm) for reaction time between 60 min and 150 nm. The

preferential growth of ZnO nano-wires in the axial direction also agrees with previous

reports.11

236
Figure 7.3 (a-d) SEM images of ALD ZnO coated nano-pillars (a) before and after (b) 60

min, (c) 120 min, and (d) 150 min of hydrothermal synthesis. (e) ZnO nano-wire length

plotted as a function of hydrothermal reaction time.

Effect of Nano-pillar Height on Photocatalytic Property

Taking advantage of controlled lithography, conformal ALD and well-tuned ZnO

nano-wire length, we fabricated a series of 3D hierarchical nanostructures with various

feature sizes and characterized their photocatalytic performance. More specifically, we

investigated the effect of nano-pillar height for ALD ZnO coated nano-pillars without

nano-wires, and the effect of nano-wires for the hierarchical nanostructures. In the following

discussion, we will refer to ALD coated nano-pillars as “NPxxx/ALD” (xxx is the height of

237
the nano-pillars in nm), while nano-pilar/nano-wire hierarchical structures will be referred to

as “NPxxx/ALD-NWyyy” (yyy is the hydrothermal growth time in min for ZnO nano-wires).

Similarly for the control group, ALD coated Si wafer will be referred to as “Si/ALD”, and

nano-wire decorated Si/ALD will be “Si/ALD-NWyyy”.

Figure 7.4 (a) Schematic of ALD ZnO coated flat Si wafer (control) and nano-pillars with

different heights. (b) Normalized concentration as a function of time during the

photocatalytic dye degradation. (c) Apparent first order rate constant (black square) and

half-life of MO (red circle) as a function of the normalized illuminated surface area.

238
Figure 7.4(a) schematically illustrates the repeating units of the nano-pillars with

different heights. By simply changing the thickness of the positive photoresist layer (Figure

1(a)) and using the same interference lithography process, nano-pillar height can be tuned

between 250 nm and 800 nm. After 200 cycles of ALD ZnO, these nanostructures were

tested for photocatalytic MO degradation.

Figure 7.4(b) shows the normalized MO concentration vs UV exposure time with and

without photocatalyst present. With only MO solution present, the MO concentration

decreases slightly, likely due to UV assisted MO hydrolysis. Adding a planar ZnO layer on Si

produced a small rate enhancement, whereas the nanostructured catalysts (NP250/ALD,

NP500/ALD or NP800/ALD) all showed significant enhancement, with faster rates for the

taller nano-pillars.

To quantify the kinetic data, we used a first-order reaction model for MO degradation to

derive kapp and t1/2. The external geometric surface area of a repeating unit (S, in units of nm2)

under illumination was estimated from the measured feature geometry then normalized it to a

unit flat Si wafer (S0, in units of nm2). Figure 7.4(c) plots kapp from the data fits vs S/S0 and

shows that kapp increases from 0.25 to 0.45 h–1 as S/S0 increases from 1.0 to 4.0. This can be

explained as the illuminated surface area of the nanostructures (or available reaction sites) is

a contributing factor to kapp.21 As t1/2 is inversely proportional to kapp (Equation 4), the

half-life of MO decreases when the surface area of the nanostructure photocatalysts

increases.

239
Effect of Hierarchical Structure on Photocatalytic Reaction Rate

We further investigated how photocatalytic performance can be improved when ZnO

nano-wires are decorated onto the nano-pillars. Figure 5 compares the MO degradation rate

with a group of nano-pillar structures with and without nano-wires. The growth time for the

ZnO nano-wires were controlled at 120 min (~ 105 nm long) for comparison within this

group of samples. kapp and t1/2 are summarized in Table 7.1.

Figure 7.5 Comparison of photocatalytic performance between nanostructure with and

without ZnO nano-wires. (a) MO degradation curves for flat ZnO surface on Si wafer and

ZnO nano-wires on ZnO coated Si wafer. (b-d) MO degradation curves for ALD ZnO coated

nanopillars with and without ZnO nano-wires. The heights of nanopillars in (b-d) are 250 nm,

500 nm and 800 nm, respectively.

240
Figure 7.5(a) shows that ZnO nano-wires improve the photocatalytic reaction kinetics

significantly for ALD ZnO coated flat Si wafer substrates. The apparent first order rate

constant is increased from 0.25 h-1 to 0.37 h-1, as the nano-wires provide more reaction sites

compared to the flat Si/ALD. The kapp of Si/ALD-NW120 lies within the range between

NP500/ALD and NP800/ALD (Table 7.1), indicating the effective surface area of

Si/ALD-NW120 should be in the range of 2.76 ~ 3.81S0. While the external geometric

surface area of the nano-wires is at least one order of magnitude higher than that of flat

Si/ALD surface (S0), the limited diffusion of dye molecules into the nm-scale pores between

the nano-wires hinders the overall MO degradation rate. In addition, the UV intensity decays

exponentially along the axial direction of the nano-wires (parallel to the direction of UV

source) according to the Lambert-Beer Law.21 The effective surface area of the nano-wires is

therefore less than the external geometric surface area.

Figure 7.5(b) compares short ALD-coated nano-pillars (NP250/ALD) with nano-wire

decorated nano-pillar hierarchical nanostructures (NP250/ALD-NW120). Similar to the

comparison between Si/ALD and Si/ALD-NW120, NP250/ALD-NW120 also enables a

faster MO degradation than NP250/ALD, enhancing the kapp from 0.26 h-1 to 0.42 h-1 (Table

7.1). This 61% increase in reaction rate demonstrates that the 3D hierarchical nanostructures

promote the kinetics of photocatalyic MO degradation. Improved photocatalytic activity was

also observed for NP500/ALD-NW120 and NP800/ALD-NW120 (Figure 7.5(c-d)) compared

to the corresponding nano-pillar structures without nano-wires. However, the extent of

increase in kapp decreases to 25% and 8% for NP500/ALD-NW120 and NP800/ALD-NW120,

respectively (Table 7.1). This indicates that effective reaction sites are not significantly

241
improved when the height of these 3D hierarchical nanostructures is increased, possibly a

combined result of diffusion limitation and UV intensity decay from the top of the structure

to the bottom.

We also studied the effect of nano-wire length on photocatalytic property using

nano-pillar substrates with same feature sizes (NP500/ALD). The synthesis time of

nano-wires on NP500/ALD ranged from 60 min to 120 min. The kapp and t1/2 values are

summarized in Table 1. An improved photocatalytic performance is observed when the

nano-wire length is increased. We noticed that the kapp values of NP500/ALD-NW60 and

NP500/ALD-NW90 are slightly less than NP500/ALD. This could be because of the light

scattering and reflection in the hierarchical nanostructure. An optical model will be needed to

understand the UV intensity profile in this type of complex structures.

Table 7.1 Feature sizes and photocatalytic performance of ZnO 3D hierarchical

nanostructrues.

Nano-pillar Nano-wire
Photocatalyst kapp (h-1) t1/2 (h)
Height (nm) Length (nm)
No Catalyst N/A N/A 0.0084 83
Si/ALD 0 0 0.25 2.8
Si/ALD-NW120 0 120 0.37 1.9
NP250/ALD 250 0 0.26 2.6
NP250/ALD-NW120 250 120 0.42 1.6
NP500/ALD 500 0 0.34 2.1
NP500/ALD-NW60 500 60 0.27 2.5
NP500/ALD-NW90 500 90 0.33 2.1
NP500/ALD-NW120 500 120 0.42 1.6
NP800/ALD 800 0 0.45 1.6
NP800/ALD-NW120 800 120 0.48 1.4

242
7.4 Conclusion

In summary, we have demonstrated a new fabrication procedure for synthesizing ZnO

3D hierarchical nanostructures. Using laser interference lithography, we have prepared a

series of nano-pillars with desired heights. Taking advantage of ALD, conformal ZnO thin

films have been deposited onto the nano-pillars in order to both make the surface hydrophilic

and serve as nucleation layers for ZnO nano-wires. We have shown that within the linear

growth window we can easily tune the nano-wire length by adjusting the hydrothermal

synthesis time. We have fabricated a series of hierarchical nanostructures and studied the

effect of nano-pillar height and the effect of hierarchical structure on the photocatalytic

kinetics of MO degradation. Without nano-wires, taller nano-pillars show higher

photocatalytic activity as a result of increased reaction sites. When ZnO nano-wires are

introduced to the system, the apparent first order rate constant can be further increased up to

61% compared to nano-pillars without nano-wires. Limitation of diffusion of dye molecules

into the nm-scale pores and UV light intensity decay in the hierarchical nanostructures affects

the overall photocatalyic reaction rate. Mathematic modeling of mass transport coupled with

reaction kinetics and light intensity profile in the complex structure will be needed in the

future to further understand photocatalysis in hierarchical nanostructures. This work provides

insights for designing and controlling nano-features for enhancing reactivity, which can also

be applied to similar structures for dye sensitized solar cells, nanostructure batteries and

other photocatalytic systems.

243
References

(1) Hoffmann, M. R.; Martin, S. T.; Choi, W.; Bahnemann, D. W. Environmental


Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95 (1), 69–96

(2) Khan, S. U. M.; Al-Shahry, M.; Ingler, W. B. Efficient photochemical water splitting
by a chemically modified n-TiO2 2. Science 2002, 297 (5590), 2243–2245.

(3) Maeda, K.; Takata, T.; Hara, M.; Saito, N.; Inoue, Y.; Kobayashi, H.; Domen, K.
GaN:ZnO Solid Solution as a Photocatalyst for Visible-Light-Driven Overall Water
Splitting. J. Am. Chem. Soc. 2005, 127 (23), 8286–8287.

(4) Hernández-Alonso, M. D.; Fresno, F.; Suárez, S.; Coronado, J. M. Development of


alternative photocatalysts to TiO2: Challenges and opportunities. Energy Environ. Sci.
2009, 2 (12), 1231–1257.

(5) Janotti, A.; Van de Walle, C. G. Fundamentals of zinc oxide as a semiconductor. Rep.
Prog. Phys. 2009, 72 (12), 126501.

(6) Akyol, A.; Bayramoglu, M. Preparation and characterization of supported ZnO


photocatalyst by zincate method. J. Hazard. Mater. 2010, 175 (1-3), 484–491.

(7) Daneshvar, N.; Salari, D.; Khataee, A. . Photocatalytic degradation of azo dye acid red
14 in water on ZnO as an alternative catalyst to TiO2. J. Photochem. Photobiol. Chem.
2004, 162 (2–3), 317–322.

(8) Evgenidou, E.; Fytianos, K.; Poulios, I. Photocatalytic oxidation of dimethoate in


aqueous solutions. J. Photochem. Photobiol. Chem. 2005, 175 (1), 29–38.

(9) Hariharan, C. Photocatalytic degradation of organic contaminants in water by ZnO


nanoparticles: Revisited. Appl. Catal. Gen. 2006, 304, 55–61.

(10) Daneshvar, N.; Aber, S.; Seyed Dorraji, M. S.; Khataee, A. R.; Rasoulifard, M. H.
Photocatalytic degradation of the insecticide diazinon in the presence of prepared
nanocrystalline ZnO powders under irradiation of UV-C light. Sep. Purif. Technol.
2007, 58 (1), 91–98.

(11) Gong, B.; Peng, Q.; Na, J.-S.; Parsons, G. N. Highly active photocatalytic ZnO
nanocrystalline rods supported on polymer fiber mats: Synthesis using atomic layer
deposition and hydrothermal crystal growth. Appl. Catal. Gen. 2011, 407 (1–2), 211–
216.

(12) Hung, S.-T.; Chang, C.-J.; Hsu, M.-H. Improved photocatalytic performance of ZnO

244
nanograss decorated pore-array films by surface texture modification and silver
nanoparticle deposition. J. Hazard. Mater. 2011, 198, 307–316.

(13) Xu, F.; Shen, Y.; Sun, L.; Zeng, H.; Lu, Y. Enhanced photocatalytic activity of
hierarchical ZnO nanoplate-nanowire architecture as environmentally safe and facilely
recyclable photocatalyst. Nanoscale 2011, 3 (12), 5020–5025.

(14) Yang, J.; Fei, G. T.; Li, H.; Ouyang, H. Facile Synthesis of 3D Porous Flower-like
ZnO Micro/nanostructure Films and Their Photocatalytic Performance. Chin. J. Chem.
Phys. 2012, 25 (3), 339–344.

(15) Iqbal, D.; Kostka, A.; Bashir, A.; Sarfraz, A.; Chen, Y.; Wieck, A. D.; Erbe, A.
Sequential Growth of Zinc Oxide Nanorod Arrays at Room Temperature via a
Corrosion Process: Application in Visible Light Photocatalysis. Acs Appl. Mater.
Interfaces 2014, 6 (21), 18728–18734.

(16) Lee, H. U.; Park, S. Y.; Lee, S. C.; Seo, J. H.; Son, B.; Kim, H.; Yun, H. J.; Lee, G. W.;
Lee, S. M.; Nam, B.; et al. Highly photocatalytic performance of flexible 3
dimensional (3D) ZnO nanocomposite. Appl. Catal. B Environ. 2014, 144, 83–89.

(17) Colmenares, J. C.; Kuna, E.; Jakubiak, S.; Michalski, J.; Kurzydłowski, K.
Polypropylene nonwoven filter with nanosized ZnO rods: Promising hybrid
photocatalyst for water purification. Appl. Catal. B Environ. 2015, 170–171, 273–282.

(18) Smith, H. I. Low cost nanolithography with nanoaccuracy. Phys. E Low-Dimens. Syst.
Nanostructures 2001, 11, 104–109.

(19) Hyde, G. K.; Scarel, G.; Spagnola, J. C.; Peng, Q.; Lee, K.; Gong, B.; Roberts, K. G.;
Roth, K. M.; Hanson, C. A.; Devine, C. K.; et al. Atomic Layer Deposition and Abrupt
Wetting Transitions on Nonwoven Polypropylene and Woven Cotton Fabrics.
Langmuir 2010, 26 (4), 2550–2558.

(20) Lee, K.; Jur, J. S.; Kim, D. H.; Parsons, G. N. Mechanisms for
hydrophilic/hydrophobic wetting transitions on cellulose cotton fibers coated using
Al2O3 atomic layer deposition. J. Vac. Sci. Technol. A 2012, 30 (1) DOI:
10.1116/1.3671942.

(21) Lee, K.; Losego, M. D.; Kim, D. H.; Parsons, G. N. High performance photocatalytic
metal oxide synthetic bi-component nanosheets formed by atomic layer deposition.
Mater. Horiz. 2014, 1 (4), 419–423.

245

You might also like