14.10.2016 Views

resolver

resolver

resolver

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Controlled Modification of Metal-Organic<br />

Frameworks at Metal Sites:<br />

Local Environment Study and Properties<br />

Dissertation<br />

Wenhua Zhang


Controlled Modification of Metal-Organic<br />

Frameworks at Metal Sites:<br />

Local Environment Study and Properties<br />

Dissertation<br />

to obtain the doctorate Dr. rer. nat.<br />

of the Faculty of Chemistry and Biochemistry<br />

Ruhr‐University Bochum<br />

Submitted by Wenhua Zhang, M. Sc.<br />

July 2016


This thesis is based on the work performed during the time of October<br />

2012 and July 2016 under the supervision of Prof. Dr. Roland A. Fischer<br />

at the Chair of Inorganic Chemistry II, Organometallic & Materials, of the<br />

Ruhr‐University Bochum.<br />

1 st referee: Prof. Dr. Roland A. Fischer<br />

2 nd referee: Prof. Dr. Nils Metzler‐Nolte<br />

Herein I declare that I have written this thesis independently and<br />

without unauthorized help. Further, I assure that I have used no other<br />

sources, auxiliary means or quotes than those stated. I further declare<br />

that I have not submitted this thesis in this or in a similar form to any<br />

other university or college. Besides, I declare that I have not already<br />

undertaken an unsuccessful attempt to obtain a doctorate from another<br />

college or university.<br />

Wenhua Zhang<br />

July 2016


Acknowledgements<br />

First of all, I sincerely appreciate my supervisor<br />

Prof. Dr. Roland A. Fischer<br />

for giving me the opportunity to study in the group of Inorganic Chemistry II at Ruhr-<br />

University Bochum. I am grateful to him for affording me guidance and motivation on the<br />

scientific research. Besides, I thank his support for allowing me to attend the international<br />

conferences and visit the world-class research group, which open my mind and give me<br />

much inspiration in scientific field as well.<br />

I also thank Prof. Dr. Nils Metzler-Nolte for accepting to be the co‐referee of this<br />

dissertation.<br />

Further I would like to thank Dr. Zhenlan Fang for bringing me to the topic of Cu-based<br />

DEMOFs and her help on this topic.<br />

I address very special thanks to Dr. Olesia Halbherr for introducing me to the world of the<br />

challenging project on Ru-based (DE)MOFs and for having been advising me during my<br />

study on this topic. Thanks very much for your patience, your assistance and your<br />

knowledge on the topic as well as your invaluable guidance on writing reports. I cannot<br />

make so much progress without you.<br />

Furthermore, I would like to thank the following persons without whom this work is not<br />

able to be accomplished:<br />

Dr. Kira Khaletskaya, Dr. Min Tu, Dr. Harish Paralla, Suttipong Wannapaiboon,<br />

Christoph Rösler, and Inke Schwedler for collecting powder X-ray diffraction<br />

patterns.<br />

Dr. Arik Puls, Dr. Kerstin Freitag, Dr. Sebastian Henke and Manuela Winter for<br />

conducting single‐crystal X‐ray experiments and helping me solve the singlecrystal<br />

structure.<br />

Andreas Schneemann for teaching me to perform the TG measurements.<br />

Dr. Yuemin Wang (Karlsruhe Institute of Technology), Max Kauer, and Penghu Guo<br />

(Laboratory of Industrial Chemistry, Prof. Dr. Martin Muhler, Ruhr‐University


Bochum) for the study of XPS and UHV-FTIR and providing the corresponding<br />

figures.<br />

Dr. Bauke Albada and Martin strack for carrying out the HPLC measurements<br />

(Chair of Inorganic Chemistry I - Bioinorganic Chemistry, Prof. Dr. Nils Metzler‐<br />

Nolte, Ruhr‐University Bochum).<br />

Noushin Arshadi (group of Prof. Dr. Martin Muhler, Ruhr‐University Bochum) for<br />

collecting the N2 sorption isotherms.<br />

Dr. Rolf Neuser (Faculty of Geosciences, Institute of Geology, Mineralogy and<br />

Geophysics, Ruhr‐University Bochum) for the gold coating on the non-conducting<br />

SEM samples.<br />

Dr. Kira Khaletskaya and Stefan Cwik for conducting SEM-EDX measurements.<br />

Dr. Christian Wiktor for collecting TEM-EDX spectra.<br />

Dr. Christian Sternemann (and Andreas Schneemann, Suttipong Wannapaiboon)<br />

for helping me to gather X‐ray diffraction patterns at the DELTA facility. Ralph<br />

Wagner (and Andreas Schneemann) for assistance of performing XAS on the Ru-<br />

(DE)MOF samples and for introducing me to the study of XANES.<br />

Karin Bartholomäus for elemental analysis.<br />

Prof. Dr. Jeffrey R. Long for the chance to have a research stay in his group, at the<br />

Department of Chemistry, University of California, Berkeley (USA). The whole<br />

group for an inspiring communication and very nice time during my stay. I address<br />

many thanks to Miguel Gonzalez for guiding me on conducting the catalytic<br />

reaction in Chapter 4 and assistance on evaluating the data. Last but not the least,<br />

Dianne Xiao, Douglas Reed, Julia Oktawiec for their great assistance with sorption<br />

studies and help on collecting sorption isotherms.<br />

Dr. Francesc X. Llabrés i Xamena (Institute of Chemical Technologies, Polytechnical<br />

University of Valencia, Spain) for the collaboration and Konstantin Epp (group of<br />

the Prof. Dr. Roland A. Fischer in Technical University of Munich) for performing<br />

the catalytic reaction in Chapter 4.<br />

Zhihao Chen and Majd Al-Naji (group of Prof. Dr. Roger Gläser, University of<br />

Leipzig) for carrying out the catalytic reaction in Chapter 5 and the assistance on<br />

study the catalytic activity.<br />

I also want to acknowledge the financial support from China Scholarship Council to my<br />

whole Ph.D study and Research School Plus at Ruhr-University Bochum for the funding of


attending conferences, my research stay at UC Berkeley and my invitation of Dianne Xiao<br />

(from UC Berkeley) for the short visit in ACII group.<br />

I am thankful to the master students who work with me during my Ph.D study: Marco<br />

Rehosek, Sebastian Kunze, and Konstantin Epp.<br />

I am grateful to all ACII members for a nice working atmosphere, the great team spirit and<br />

inspiring communications, especially the former and present MOF family. The memorable<br />

and wonderful time we spend together in the MOF 2014 conference and the trip in Japan<br />

with our MOF guest Takashi Toyao. I express my many thanks to Dr. Christian Wiktor on<br />

teaching me skills for my first poster. Also further thanks to Dr. Chuanqiang Li, Suttipong<br />

Wannapaiboon, Andreas Schneemann, and Konstantin Epp for the kind accompany<br />

during XAS data collecting in DELTA facility.<br />

Besides, I address many thanks to my lab neighbor Dr. Raghavender Medishetty and Dr.<br />

Arik Puls for the abundant and fruitful discussions.<br />

I would like to express my thanks to Dr. Mariusz Molon for affording technical help on my<br />

computer, and OM group members for sharing the experience on air-free synthesis.<br />

I am grateful to Uschi Hermann for her kindness and great help in the lab.<br />

Furthermore, I would like to express my thanks to Jana Weiing, Andreas Schneemann,<br />

Jiyeon Kim, Inke Schwedler and Dr. Christian Wiktor for the great time during ACS fall<br />

meeting and the nice trip in Boston.<br />

I would like to express my warm thanks to Sabine Pankau and Jacinta Essling for their<br />

great help with administrative issues.<br />

Many thanks to my office colleagues for nice working atmosphere, giving me suggestions<br />

and help on the daily life and as translators: Sarah Karle, Mathies Evers, Stefan Cwik,<br />

Maximilian Gebhard, Inke Schwedler, Richard O'Donoghue, Jiyeon Kim, Andreas<br />

Schneemann, Dr. Christian Wiktor, Dr. Arik Puls, and Dr. Sun Ja Kim.<br />

I am also thankful to Dr. Ke Xu for helping me go through a bunch of documents to settle<br />

down in Germany.<br />

Last but the most important, I would like to deeply thank my parents Denggao Zhang and<br />

Chunying Wu, for giving me the freedom and support to study abroad and their endless<br />

love and motivation during my studies. Also thank my siblings, Wenjing and Wenbo


Zhang, for their continuous support and care. My love and deep appreciation to Zhihao<br />

Chen for always being at my side. His motivation and continuous care help me recover<br />

from the sadness of homesick and overcome the challenging in my life and study. All these<br />

would not happen without them.


Table of Contents<br />

Table of Contents ................................................................................................................................................. I<br />

Abbreviations ....................................................................................................................................................... V<br />

1 Motivation .................................................................................................................................................... 1<br />

2 General introduction ............................................................................................................................... 3<br />

2.1 Origin of MOFs - Coordination polymers ............................................................................... 3<br />

2.2 Definition, Chemistry and features of MOFs ........................................................................ 7<br />

2.2.1 Definition and nomenclature of MOFs ................................................................ 7<br />

2.2.2 MOF features .................................................................................................................. 8<br />

2.3 Synthetic approaches, modification and applications of MOFs ................................ 14<br />

2.3.1 Traditional synthesis approaches ...................................................................... 14<br />

2.3.2 Reticular synthesis ................................................................................................... 15<br />

2.3.3 Way to the modification of MOFs ....................................................................... 16<br />

2.3.4 Application of MOFs ................................................................................................. 23<br />

3 Controlled secondary building unit approach to the formation of the<br />

isostructural Ru II,II and Ru II,III analogs of [M3(BTC)2]n ........................................................... 26<br />

3.1 Preparation and investigation of the Ru II,III -analogs of [M3(BTC)2]n ...................... 28<br />

3.1.1 The family of [M3(BTC)2]n...................................................................................... 28<br />

3.1.2 Background and the state-of-the-art in research on Ru-analogs of<br />

[M3(BTC)2]n.................................................................................................................. 29<br />

3.1.3 [Ru3(BTC)2Yy]n·G g obtained by CSA using [Ru2(OOCR)4X] and<br />

[Ru2(OOCCH3)4]A: synthesis and characterization ..................................... 31<br />

3.1.4 Pre-activation by solvent exchange ................................................................... 37<br />

3.1.5 Study of [Ru3(BTC)2Yy]n·G g after solvent exchange .................................... 38<br />

3.1.6 Summary ....................................................................................................................... 48<br />

3.2 Elaboration of Ru II,II analog of [M3(BTC)2]n........................................................................ 50


II<br />

3.2.1 Synthesis and characterization of SBU-e and Ru-MOF 5.......................... 50<br />

3.2.2 Investigation on the Ru oxidation state and porosity of Ru-MOF 5 .... 53<br />

3.2.3 Conclusions .................................................................................................................. 55<br />

4 Linker-based MOF solid solutions: Defect-Engineered MOFs (DEMOFs) .................... 56<br />

4.1 Introduction .................................................................................................................................... 58<br />

4.1.1 Solid solutions ............................................................................................................ 58<br />

4.1.2 Defects in MOFs and its “engineering”............................................................. 60<br />

4.2 Ruthenium Metal-Organic Frameworks featuring Different Defect Types ......... 65<br />

4.2.1 Synthesis and characterization of the Ru-DEMOF materials ................. 65<br />

4.2.2 Porosity of Ru-DEMOF samples and the confirmation of the DL<br />

incorporation .............................................................................................................. 76<br />

4.2.3 XANES, XPS and UHV-FTIR studies: ruthenium oxidation state<br />

variation as indication of defect type formation ......................................... 79<br />

4.2.4 CO2, CO and H2 sorption properties of Ru-DEMOFs ................................... 91<br />

4.2.5 Catalytic test reactions ......................................................................................... 100<br />

4.2.6 Conclusions ............................................................................................................... 106<br />

4.3 Defects Engineering in [Cu3(BTC)2]n: Effect of the synthetic parameters ......... 108<br />

4.3.1 Preparation and characterization of Cu-DEMOF samples<br />

[Cu3(BTC)2-x(ip)x]n.................................................................................................. 109<br />

4.3.2 Composition and porosity of the prepared Cu-DEMOFs ....................... 116<br />

4.3.3 Oxidation state(s) of metal-sites in prepared Cu-DEMOFs .................. 119<br />

4.3.4 Conclusions ............................................................................................................... 121<br />

5 Simultaneous introduction of various palladium active sites into MOF via onepot<br />

synthesis: Pd@[Cu3-xPdx(BTC)2]n ......................................................................................... 122<br />

5.1 Introduction on the selection of metal type in MOFs ................................................. 123<br />

5.2 Preparation and Structure of the Cu/Pd-BTC_1-3 ....................................................... 126<br />

5.3 Compositional characterization and sorption properties of the prepared<br />

Cu/Pd-BTC_1-3 ........................................................................................................................... 130


III<br />

5.4 Synthesis, compositional characterization and sorption properties of<br />

Cu/Pd-BTC_4 ............................................................................................................................... 136<br />

5.5 Catalytic test reaction .............................................................................................................. 143<br />

5.6 Conclusions ................................................................................................................................... 146<br />

6 Summary and Outlook ...................................................................................................................... 147<br />

7 Experimental Section ........................................................................................................................ 150<br />

7.1 General methods ........................................................................................................................ 150<br />

7.1.1 X-ray Diffraction (XRD) ....................................................................................... 150<br />

7.1.2 FT-IR spectroscopy ................................................................................................ 151<br />

7.1.3 Nuclear Magnetic Resonance (NMR) Spectroscopy ................................ 153<br />

7.1.4 Thermogravimetric analyses (TGA)............................................................... 154<br />

7.1.5 Scanning Electron Microscopy (SEM) and Energy-dispersive X-ray<br />

spectroscopy (EDX)............................................................................................... 154<br />

7.1.6 Transmission electron microscope (TEM) and EDX ............................... 154<br />

7.1.7 Elemental Analysis / Atomic absorption spectroscopic (AAS) .......... 154<br />

7.1.8 Gas Sorption ............................................................................................................. 155<br />

7.1.9 X-ray absorption near edge structure (XANES)........................................ 157<br />

7.1.10 X-ray photoelectron spectroscopy (XPS)..................................................... 158<br />

7.1.11 High Performance Liquid Chromatography (HPLC) ............................... 158<br />

7.1.12 Catalytic studies ...................................................................................................... 159<br />

7.2 Experimental data on chapter 3........................................................................................... 160<br />

7.2.1 Synthesis of the ruthenium precursors ([Ru2(OOCR)4X] and<br />

[Ru2(OOCCH3)4]A).................................................................................................. 160<br />

7.2.2 Synthesis of Ru-MOFs [Ru3(BTC)2Xx]·Gg ...................................................... 168<br />

7.3 Experimental data on chapter 4........................................................................................... 173<br />

7.3.1 Synthesis of Ru-DEMOF samples (1a-1d, 2a-2d, 3a-3d, 4a-4c) .......... 173<br />

7.3.2 Catalytic reactions using Ru-DEMOFs as catalysts .................................. 177<br />

7.3.3 Synthesis of [Cu3(BTC)2]n (Cu-BTC)............................................................... 177


IV<br />

7.3.4 Synthesis of Cu-DEMOF samples (D1-8)...................................................... 178<br />

7.4 Experimental data on chapter 5........................................................................................... 186<br />

7.4.1 Synthesis of [Cu3-XPdx(BTC)2]n ......................................................................... 186<br />

7.4.2 Hydrogenation of PNP to PAP ........................................................................... 190<br />

7.5 Supplementary details in Outlook ...................................................................................... 192<br />

8 Bibliography.......................................................................................................................................... 194<br />

Appendix ........................................................................................................................................................... 205<br />

List of Publications<br />

List of Presentations<br />

Curriculum Vitae


V<br />

Abbreviations<br />

4-btapa<br />

5-Br-ip<br />

5-Br-ipH2<br />

5-NH2-ip<br />

5-NH2-ipH2<br />

5-OH-ip<br />

5-OH-ipH2<br />

AAS<br />

AcO<br />

AcOH<br />

BBR<br />

bdc<br />

BET<br />

bpy<br />

BTC<br />

BTT<br />

CP<br />

CPO<br />

CSA<br />

CUS<br />

DEMOFs<br />

DL<br />

DMF<br />

1,3,5-benzene tricarboxylic acid tris[N-(4-pyridyl)amide]<br />

5-bromoisophthalate<br />

5-bromoisophthalic acid<br />

5-aminoisophthalate<br />

5-aminoisophthalic acid<br />

5-hydroxyisophthalate<br />

5-hydroxyisophthalic acid<br />

atomic absorption spectrascopy<br />

acetate<br />

acetic acid<br />

building block replacement<br />

1,4-benzenddicarboxylate<br />

Brunauer-Emmett-Teller<br />

bipyridine<br />

1,3,5-benzenetricarboxylate<br />

1,3,5-benzenetristetrazolate<br />

coordination polymer<br />

coordination polymer from Oslo<br />

controlled secondary building unit approach<br />

coordinatively-unsaturated metal sites<br />

defect-engineered metal-organic framworks<br />

defect linker<br />

N,N-dimethylformamid


VI<br />

DMSO<br />

dobdc<br />

e.g.<br />

EA<br />

EDX<br />

EtOH<br />

EXAFS<br />

FT-IR<br />

GC<br />

H2ip<br />

H2pydc<br />

H3BTC<br />

HKUST<br />

HML<br />

HPLC<br />

i.e.<br />

IML<br />

ip<br />

IRMOF<br />

IUPAC<br />

MC<br />

mCUS<br />

MeOH<br />

MIL<br />

MOF<br />

dimethyl sulfoxide<br />

2,5-dioxido-1,4-benzenedicarboxylate<br />

exempli gratia (from latin = for example)<br />

elemental analysis<br />

energy-dispersive X-ray spectroscopy<br />

ethanol<br />

extended X-ray absorption fine structure<br />

fourier transformation infrared (spectroscopy)<br />

Gas chromatography<br />

isophthalic acid<br />

pyridine-3,5-dicarboxylic acid<br />

1,3,5-benzenetricarboxylic acid<br />

Hong Kong Unversity of Science & Technology<br />

heterostructural mixed-linker<br />

High performance liquid chromatography<br />

id est (from Latin = it is, namely)<br />

isostructural mixed-linker<br />

isophthalic<br />

isoreticular metal-organic framework<br />

International Union of Pure and Applied Chemistry<br />

mixed-component<br />

modified coordinatively-unsaturated metal sites<br />

methanol<br />

Matériaux de l′Institut Lavoisier<br />

metal-organic framework


VII<br />

MTV<br />

Mw<br />

NMR<br />

NP<br />

PAP<br />

PivO<br />

PivOH<br />

PNP<br />

PSM<br />

PW<br />

PXRD<br />

pydc<br />

pyz<br />

r.t.<br />

redox<br />

rion<br />

SALE<br />

SBET<br />

SBU<br />

SEM<br />

STP<br />

TEM<br />

TGA<br />

THF<br />

TML<br />

multivariate<br />

molecular weight<br />

nuclear magnetic resonance<br />

nano-particle<br />

p-aminophenol<br />

pivalate<br />

pivalic acid<br />

p-nitrophenol<br />

post-synthetic method<br />

paddlewheel<br />

powder X-ray diffraction<br />

pyridine-3,5-dicarboxylate<br />

pyrazine<br />

room temperature<br />

reduction-oxidation<br />

ionic radii<br />

solvent-assisted linker exchange<br />

BET surface area<br />

secondary building unit<br />

scanning electron microscopy<br />

standard temperature and pressure<br />

transmission electron microscopy<br />

thermal gravimetric analysis<br />

tetrahydrofuran<br />

truncated mixed-linker


VIII<br />

TOF<br />

UHV<br />

UiO<br />

UMCM<br />

vs<br />

WCA<br />

XANES<br />

XAS<br />

XPS<br />

ZIF<br />

turnover frequency<br />

ultra high vacuum<br />

Universitetet i Oslo<br />

University of Michigan Crystalline Material<br />

versus<br />

weakly coordinating anion<br />

X-ray absorption near edge structure<br />

X-ray absorption sepctroscopy<br />

X-ray photoelectron spectroscopy<br />

zeolitic imidazolate framework


1 Motivation<br />

Looking through the essential fields in nature and in our daily life, one can find that porous<br />

materials play an important role. For example, porous rocks can store water, petroleum<br />

and natural gas. Activated carbon as a prominent example of porous materials have been<br />

widely used in various applications involving purification of gas, gold and water, as well<br />

as filters in gas masks and respirators, etc. Another famous example is zeolites, which are<br />

broadly used as ion-exchange beds in domestic and commercial water purification and<br />

softening, catalysts in oil tracking, etc.<br />

As zeolite-like architectures, metal-organic frameworks (MOFs) represent a young class<br />

of hybrid inorganic-organic crystalline porous materials, formed from organic linkers and<br />

inorganic build blocks (metal nodes). Ability to tune the inorganic building blocks as well<br />

as the diverse characteristics of the organic moieties featured in MOFs ensures its great<br />

advantage in comparison with the conventional porous materials (e.g. zeolites and<br />

activated carbons), which attracted huge attentions over last decades. Nowadays, MOFs<br />

are also extensively investigated as porous materials promising for gas<br />

storage/separation, sensing, drug delivery and catalysis, etc.<br />

Various synthetic approaches applied to MOFs afford to achieve their rich diversity, high<br />

level of complexity and functionality. For example, reticular synthesis (versatile design of<br />

functionalized organic linkers), design of mixed-component MOFs via post-synthetic<br />

modification (including metal and linker exchange), or mixed-linker/metal copolymerization<br />

has attracted huge attention with this regard. To note, the coordinatively<br />

unsaturated metal sites (CUS) in MOFs play a key role in many applications, especially in<br />

enhanced catalytic activity and effective gas sorption/separation. Hence, increased<br />

complexity by means of mixed-component MOFs can be combined with the generation of<br />

defects around the metal centers, which could lead to more open metal sites, or in the case<br />

of clustering of the point defects, mesoporous MOFs can be obtained. Thus, it comes to the<br />

“so-called” defect-engineering MOFs (DEMOFs). This manner of advanced modification of<br />

MOFs holds enormous potential to optimize the materials properties beyond the


2 Chapter 1<br />

limitations of the (non-doped) parent MOFs. In particular, increased sorption capacity,<br />

selectivity and enhanced catalytic activity could be expected. Up to date, only limited<br />

reports have been dealing with the studies on DEMOFs and their application.<br />

Among the MOFs featuring CUSs and showing good performance on applications, HKUST-<br />

1 ([Cu3(BTC)2]n) is one of the widely and well investigated MOFs. Besides, its mixedvalence<br />

structural analogue [Ru3(BTC)2Yy]n (Y = counter-ions, 0≤y≤1.5) exhibits sufficient<br />

chemical and thermal stability, and can offer rich photo/redox chemistry. Hence, in this<br />

thesis those two kinds of MOFs with the general formula [M3(BTC)2]n have been chosen<br />

as candidates on the study of defect-engineering.<br />

However, before starting with the exploration of such intriguing DEMOFs, the formation<br />

of both “intrinsic” and “intentional” defects in MOFs should be wisely taken into account.<br />

Especially, in the case of [Ru3(BTC)2Yy]n, due to the kinetic reason, its formation should be<br />

carefully monitored. Hence, Chapter 3 in this thesis mainly focus on the parent<br />

[Ru3(BTC)2Yy]n for the study of possible intrinsic defects, optimized exclusion of the guest<br />

molecules and elaboration of mono-valence Ru analog of HKUST-1. Further, intentional<br />

introduction of defects to [Cu3(BTC)2]n and [Ru3(BTC)2Yy]n respectively by linker and/or<br />

metal doping will be worth investigating. Modifications of the metal sites, generation of<br />

(additional) accessible coordination sites as well as the complexity of the developed<br />

DEMOFs will be studied in detail in Chapters 4 and 5. Moreover, the impact of the defects<br />

engineering on gas sorption and catalysis properties of the resulting DEMOFs will be<br />

explored. It is expected that these study show a comprehensive picture on the DEMOF<br />

derivatives of the selected M-BTC structure and give us deep insights on the defectengineering<br />

of other MOFs.


2 General introduction<br />

The aim of this chapter is to give a general introduction on the research progress of metalorganic<br />

frameworks (MOFs) including the origin of MOFs, their chemistry, structural<br />

features, synthetic approaches, modification on the structure as well as their applications.<br />

2.1 Origin of MOFs - Coordination polymers<br />

In 1833, J. J. Berzelius firstly use the term “polymer” to describe any compound that could<br />

be formulated as consisting of multiple units of a basic building block. [1] Later, Werner for<br />

the first time, proposed the correct structure [Co(NH3)6]Cl3 for coordination compounds<br />

containing complex ions in 1893, which developed the groundwork for the study of<br />

coordination polymers (CPs). In 1916, to define dimers and trimers of various cobalt(II)<br />

ammine nitrates, the term “CPs” was firstly employed by Y. Shibata [2] and has been<br />

continuously used in the scientific literature since the 1950's. Moreover, the first review<br />

on CPs was published in 1964. [3] The International Union of Pure and Applied Chemistry<br />

(IUPAC) Red Book of inorganic nomenclature from 2005 gives the following definition of<br />

coordination compounds: “A coordination compound is any compound that contains a<br />

coordination entity. A coordination entity is an ion or neutral molecule that is composed of<br />

a central atom, usually that of a metal, to which is attached a surrounding array of atoms<br />

or groups of atoms, each of which is called a ligand”. [4] Consequently, CPs can be<br />

conceptually considered as coordination compounds with extended arrays structures of<br />

one, two or three dimensions (Figure 2.1). [5] Typically, CP consists of two central<br />

components: metal ions (serving as connectors) and ligands (serving as linkers). Apart<br />

from these two components, blocking ligands, counter-anions, non-bonding guests or<br />

template molecules can be also included (Figure 2.2). The number and orientation of the<br />

binding sites (i.e. coordination numbers and geometries) are the important<br />

characteristics of the connectors and linkers. Their combination in numerous ways via a<br />

self-assembly enables the formation of practically an infinite number of CPs (Figure 2.1).


4 Chapter 2<br />

For the synthesis of CPs, transition metals and lanthanides are often utilized as connectors.<br />

According to the type of the metal and its oxidation state, coordination numbers can vary<br />

from 2 to 7, resulting in a variety of coordination geometries such as linear, T- or Y-<br />

shaped, tetrahedral, square-planar, square-pyramidal, trigonal-bipyramidal, octahedral,<br />

trigonal-prismatic, pentagonal-bipyramidal and the corresponding distorted forms<br />

(Figure 2.2). In addition, metal-complexes might be also used as connectors instead of<br />

(naked) metal-ions. It has an advantage of offering the bond angles control and restricting<br />

the number of coordination sites via “ligand-regulation” of a connector, where chelating<br />

or macrocyclic ligands directly bound to a metal connector block the redundant sites and<br />

the linkers are left free for specific sites. [6]<br />

Figure 2.1. Schematic demonstration of the CP construction via the self-assembly of connector<br />

(metal nodes) and the simplest linker (ligands) in 1-3D.<br />

Figure 2.2. Components of CPs. Inspired by the report from Kitagawa, et al. [7]


Chapter 2 5<br />

On the other hand, linkers provide various connecting sites with tuned binding strength<br />

and directionality. Linkers utilized in the CP construction may be halides (F, Cl, Br and I),<br />

CN - and SCN - ions, [8-9] cyanometallate anions ([M(CN)x] n- ) as well as neutral, anionic and<br />

cationic organic ligands. Among all these ligands, halides as the smallest and simplest one<br />

are broadly employed to form quasi-1D halogen-bridged mixed-valence compounds (MX<br />

chains), which feature extensive physical properties. [10] Furthermore, halides can be well<br />

incorporated along with neutral organic ligands in the coordination frameworks. [11]<br />

Cyanometallate anions could perform various geometries, similar to the metal ions with<br />

coordination number from 2 to 7 mentioned above. With respect to the neutral organic<br />

ligands, pyrazine (pyz) and 4, 4’-bpy (bpy = bipyridine) are the most often used in the<br />

reports. [12] As to the anionic ligands, di-, tri-, tetra- and hexacarboxylate molecules are<br />

typically employed. [13-15] On the contrary, CPs with cationic ligands are not very common<br />

due to their low affinity for cationic metal ions. [16-17]<br />

The bonding interactions between connectors and linkers in the CPs are mainly<br />

coordination bonds. Additional weak interactions such as hydrogen and metal-metal<br />

bonds, π- π and CH- π, electrostatic and van der Waals interactions can also participate to<br />

construct the structure of CPs. Consequently, both robust and flexible networks could be<br />

made. Many CPs studied in the earlier stage (1 st generation) feature open voids, cavities<br />

or channels which are usually occupied by the solvent molecules or counter-anions. Such<br />

guests might be exchanged post-synthetically against other ions or solvents, what is of<br />

highly interest to study (host-guest chemistry, ion-exchange, etc.). [18-19] However,<br />

complete removal of the guests molecules in this type of polymers is accompanied with<br />

an irreversible structure collapse, what considerably restrict their application in the fields<br />

of gas storage and separation as a result of low structural rigidity and absence of<br />

permanent porosity. On the other hand, structural interpenetrations, in which the voids<br />

constructed by one framework are occupied by one or more other independent<br />

frameworks, [20] frequently happens in chemistry of CPs and might facilitate formation of<br />

robust structures. Providing structural rigidity, this way, however, often leads in<br />

considerable reduction of the pore sizes and precludes creation of highly porous<br />

frameworks. Therefore, in spite of the abundant study on this class of CPs (1 st generation)<br />

in the early stage (before 1990s), synthesis of coordination networks exhibiting<br />

permanent porosity was still an open question until the exploration of [Co2(4,4′-<br />

bpy)3(NO3)4]n(H2O)4 in 1997 by S. Kitagawa. [21] This compound featuring channeling


6 Chapter 2<br />

cavities is able to reversibly adsorb CH4, N2, and O2 in the pressure range of 1–36 atm<br />

without deformation of the crystal framework.


Chapter 2 7<br />

2.2 Definition, Chemistry and features of MOFs<br />

2.2.1 Definition and nomenclature of MOFs<br />

The IUPAC group recommends the following definition of MOFs: “A metal–organic<br />

framework, abbreviated to MOF, is a coordination network with organic ligands containing<br />

potential voids.” Moreover, the hierarchical terminology is also recommended, where the<br />

most general term is CP, followed by the coordination networks acting as a subset of CPs<br />

and MOFs, a further subset of coordination networks (Figure 2.3). [5] It is worth<br />

mentioning that other terms like “porous CPs” or “porous coordination network”, which<br />

have basically the same meaning as MOFs, are also present in many scientific publications.<br />

In addition, another term “hybrid inorganic-organic material” was occasionally used for<br />

MOFs, however is not recommended by the IUPAC group due to its imprecise description.<br />

Owing to the cumbersome work, the systematic nomenclature of MOFs has not been given<br />

yet. Nevertheless, assigning to some conceptually important compounds the trivial names<br />

or nicknames in accordance with place of their origin followed by a number, such as<br />

HKUST-1 (Hong Kong University of Science and Technology), MIL-53 (Matériaux de<br />

l′Institut Lavoisier) and UiO-66 (Universitetet i Oslo), is also acceptable.<br />

Figure 2.3. The hierarchical terminology of the CPs, coordination networks and MOFs based on<br />

the recommendations of the IUPAC group.


8 Chapter 2<br />

2.2.2 MOF features<br />

Advantages of MOFs combining the features of CPs and porous solids<br />

MOFs (i.e. porous coordination networks) could be considered like a combination of both<br />

CPs and porous solids. As a subclass of CPs, MOFs feature even more diverse and robust<br />

structures. The construction of MOFs is commonly realized by the connection of “rigid”<br />

secondary building units (SBUs) rather than the simple linking of node and spacer in<br />

coordination networks where usually single atoms are joined by ditopic linkers. On the<br />

other hand, the feature of being microporous compounds, which is very important<br />

motivation to design MOFs, should not be underrated. Classical porous solids including<br />

zeolites, silica, alumina, carbon-based materials, etc. are known in various research fields<br />

of not only chemistry but also physical and material science due to their versatile<br />

application related to separation, storage and heterogeneous catalysis. For example,<br />

zeolites can be effectively utilized as molecular sieves for adsorbent of gases and liquids,<br />

builders for the production of laundry detergents, etc. Moreover, the usage of activated<br />

carbon in air filters and gas masks is seen in our daily life. MOFs as a novel class of porous<br />

solids are expected to behave in a similar manner as other mentioned porous solids. In<br />

fact, they exhibit uniform pores or open channels which might be preserved upon careful<br />

removal of the guest molecules. Compared to the CPs of 1 st generation, this essential<br />

characteristic allows MOFs to hold huge potential for a wide variety of applications such<br />

as gas sorption/separation, [22-23] gas storage [24-26] and catalysis, [27-29] for which previously<br />

only conventional purely inorganic (e.g. zeolites, alumina, silica) or purely organic porous<br />

solids (e.g. activated carbons) were suitable.<br />

Features highlight 1 –Structure diversity and porosity<br />

As mentioned above, MOFs are constructed from metal nodes (also known as inorganic<br />

building blocks or SBUs) and multidentate organic linkers (e.g. carboxylates, [13, 30]<br />

azolates, [31-33] etc.). One of the common clusters, the octahedral Zn4O(CO2)6 cluster(Figure<br />

2.4) composed of four ZnO4 tetrahedra with a common vertex and six carboxylate C atoms<br />

can be as SBUs and further joined together by the benzene links, which leads to a 3D MOF<br />

Zn4O(bdc)3∙(DMF)8(C6H5Cl) (known as MOF-5, bdc = 1,4-benzenddicarboxylate, DMF =<br />

N,N-Dimethylformamid), where the vertices are the octahedral SBUs and the edges are<br />

the benzene struts. [18] The obtained framework exhibit high thermal stability(300 °C) and<br />

porosity (Brunauer-Emmett-Teller (BET) surface area well above 3500 m 2 /g). The same


Chapter 2 9<br />

inorganic SBU can be connected by distinct ditopic linkers to afford various materials with<br />

the same structural topology (so-called IRMOFs) featuring predetermined cavity size and<br />

functions (Figure2.5).[25, 34]<br />

Figure 2.4. Construction of the MOF-5 framework. Top, the Zn 4 O(CO 2 ) 6 cluster. Bottom, one of the<br />

cavities in the MOF-5 framework. Reprinted by permission from Macmillan Publishers Ltd:<br />

Nature, 402, 276-279, copyright 1999. [18]<br />

Consideration on the chemical and geometric attributes of the essential SBUs and linkers can<br />

afford prediction of the framework topology, and subsequently result in the design and targeted<br />

synthesis of a new class of porous materials featuring robust structures and high porosity.<br />

Transition-metal carboxylate clusters may serve as SBUs with different points of extension<br />

varying from 3 to 66. [30] In addition to the typical octahedral zinc acetate cluster (Zn 4 O(CO 2 ) 6 ) as<br />

SBU, one of the other metal cluster commonly employed as a build blocks in the MOF synthesis is<br />

square bimetallic PW (M 2 (CO 2 ) 4 ) (Figure 2.6), in which four carboxylates are connected to two<br />

metal centers. Each metal center is often capped by labile solvent molecules (such as H 2 O). A<br />

typical example composed of such building blocks is the 3D porous framework [Cu 3 (BTC) 2 ] n (also<br />

known as HKUST-1 [35] or MOF-199 [36] ) reported by Chui. et al. in 1999.


10 Chapter 2<br />

Figure 2.5. A series of isoreticular MOFs (IRMOFs). [34] Reprinted from Micropor. Mesopor.<br />

Mater.,73, J. L. C. Rowsell, O. M. Yaghi, Metal–organic frameworks: a new class of porous materials,<br />

3‐14. Copyright 2004, with permission from Elsevier.<br />

Figure 2.6. Square PW cluster (M 2 (CO 2 ) 4 )with two terminal ligand sites (left) and the structure of<br />

HKUST-1.<br />

Features highlight 2-Flexibility


Chapter 2 11<br />

Figure 2.7. a, Three classes of host materials categorized according to attributes of softness,<br />

hardness (rigidity) and regularity. The overlapping zone of the two stages indicates materials<br />

belonging to either of the ends. b, Classification of porous CPs into three categories. Reprinted by<br />

permission from Macmillan Publishers Ltd: Nature Chemistry, 1, 695–704, copyright 2009. [37]<br />

Besides features mentioned above (structural versatility and permanently porosity),<br />

MOFs might also be flexible. Kitagawa et. al classified porous CPs into three generations<br />

(Figure 2.7). [37-38] The 1 st generation, which have been earlier mentioned in this chapter,<br />

represents porous coordination networks, which are comparably labile and collapse upon<br />

removal of the guests from the pores. The 2 nd generation includes robust frameworks<br />

exhibiting permanent porosity (reversible adsorption/desorption of guest molecules)<br />

with complete preservation of the crystalline order. The 3 rd generation, identified as<br />

flexible or dynamic porous frameworks, is able to undergo defined (and reversible) phase<br />

transitions as a respond to such external stimuli as thermal, [39] , guest molecule<br />

adsorption/desorption, [40] mechanical stress, [41] or light. [42] However, such structural<br />

transitions do not lead to breaking of coordination bonds and the internal connectivity of


12 Chapter 2<br />

the material is retained. Materials undergoing such kinds of phase transitions are often<br />

called soft porous crystals. [37] The effect of large framework flexibility induced by specific<br />

host-guest interaction is also known as “breathing”. [43] One of the best-studied flexible<br />

systems is the MIL-53 system ([M(bdc)(OH)]n, M 3+ =Al 3+ , [44] Fe 3+ , [45] Cr 3+ , [46] Ga 3+ , [47]<br />

In 3+ , [48] Sc 3+ , [49] ). These frameworks demonstrate different crystal phases according to the<br />

activation state and the incorporation of guest molecules. Thus, after activation MIL-53<br />

reveals a high-temperature phase with large pores (lp), which can shrinks to a narrowpore<br />

phase (np) upon the adsorption of low amounts of polar molecules (Figure 2.8).<br />

Moreover, a closed-pore phase was observed after activation of MIL-53(Fe) [45] and MIL-<br />

53(Sc). [49] In addition to these famous MILs-53, some M2-PW based pillared-layered MOFs<br />

also exhibit certain degree of flexibility upon guest adsorption. [50-52]<br />

Figure 2.8.The different forms of MIL-53: (a) as synthesized (as); disordered terephthalic acid<br />

molecules lie within the tunnels; (b) high temperature (open), lp; (c) room temperature hydrated<br />

form (hydr.), np. Note the changes in the cell parameters during the thermal treatments<br />

Reproduced from Chem. Soc. Rev., 2008, 37, 191-214 with permission of The Royal Society of<br />

Chemistry. [53]<br />

Features highlight 3-Good thermal stability even good chemical stability<br />

Thermal stability of MOFs ranges from 250 °C to 500 °C. [25, 31, 54-55] . Chemically stable<br />

MOFs, however, are challenging to be made due to their susceptibility to linkdisplacement<br />

reactions under the treatment with solvents over extended periods of time.<br />

Zeolitic imidazolate framework ZIFs, in particular ZIF-8 (Zn(MIm)2, MIm = 2-<br />

methylimidazolate) is one of the prominent examples which reveals exceptional chemical<br />

stability. [31] Remarkably, the structure integrity is retained after immersion ZIF-8 into<br />

boiling methanol, benzene, or water for up to 7 days. Besides, its treatment in


Chapter 2 13<br />

concentrated sodium hydroxide at 100°C for 24 hours does not alter the framework.<br />

Another well-known example exhibiting high chemical stability is MOFs based on the<br />

Zr(IV) cuboctahedral SBUs. In fact, high acid (HCl, pH = 1) and base resistance (NaOH, pH<br />

= 14) of UiO-66 (Zr6O4(OH)4(bdc)6) and its NO2 - and Br - functionalized derivatives was<br />

observed. [54-55] Furthermore, a structure of pyrazolate-bridged MOF (Ni3(BTP)2; BTP3 - =<br />

4,4’,4’’-(benzene-1,3,5-triyl)tris(pyrazol-1-ide)) was fully preserved after treatment in<br />

aqueous solutions within a wide pH value range (from 2 to 14) at 100 °C for 2 weeks. [56]<br />

Thus, such MOFs featuring high chemical stability are promising to afford benefit of<br />

enhancing their performance in carbon dioxide capture from humid flue gas, pave the way<br />

of their applications to water-containing processes, etc.<br />

To summarize, characteristic features of MOFs include (i) well-ordered porous structures<br />

and designable channel surface functionalities, (ii) flexible, dynamic behavior in response<br />

to guest molecules, and (iii) good thermal stability and (in some cases) also chemical<br />

stability. It can be expected that the diverse chemistry of MOFs (i.e. huge variety of<br />

inorganic SBUs and organic linkers bearing distinct functionalities) can result in a wide<br />

range of different framework structures as well as in deliberate tuning of their properties,<br />

which eventually leads to their applications in various fields.


14 Chapter 2<br />

2.3 Synthetic approaches, modification and applications of MOFs<br />

2.3.1 Traditional synthesis approaches<br />

Figure 2.9. Overview of the synthetic methods applied to MOFs. Reprinted with permission from<br />

N. Stock and S. Biswas, Chem. Rev., 2012, 112, 933-969. Copyright (2012) American Chemical<br />

Society. [57]<br />

As has been earlier mentioned, self-assembly of various metal ions and organic linkers<br />

leads to a large variety of networks featuring high porosity. However, it should be noted,<br />

that MOFs formation is also strongly affected by diverse synthetic parameters and<br />

conditions, such as nature of metal-precursors along with structural characteristics of<br />

linkers, concentration and stoichiometry of the employed reagents, usage of modulator,<br />

kind of solvent(s) and pH of the reaction solution), heating source, reaction temperature,<br />

time, pressure, etc. Figure 2.9 summarizes various synthetic approaches reported so far<br />

for MOFs. Conventional methods imply usage of either room or elevated temperatures<br />

and solvothermal/hydrothermal conditions (i.e. closed vessels under autogenous<br />

pressure above the boiling point of the solvent [58] ) (Figure 2.9). In addition to typical<br />

electric heating, other synthetic approaches such as microwave heating, electrochemical,<br />

mechanochemical, ultrasonic methods have been also emerging(Figure 2.9) [57, 59] and<br />

resulted in materials with various particle sizes and properties. High-throughput methods


Chapter 2 15<br />

afford a good way to study synthesis of MOFs in a systematic way and under well-defined<br />

conditions. Scaling-up of the synthesis on industrial-scale has been also achieved for<br />

several MOFs. In particular, BASF produces Cu-BTC (Basolite ® C300), Fe-BTC (Basolite ®<br />

F300), MIL-53 (Basolite ® A100), and ZIF-8 (Basolite ® Z1200), which are commercial<br />

available through Sigma-Aldrich.<br />

2.3.2 Reticular synthesis<br />

Reticular chemistry is a commonly used synthesis approach in the earlier stage for the<br />

versatile design of expansion of MOF structure to generate ultrahigh porosity and large<br />

pore openings. [60] Isoreticular principle allows the vary of size and nature of a structure<br />

while the underlying topology is kept. After the exploration of MOF-5, Yaghi et al. obtained<br />

a series of isoreticular MOFs on the basis of MOF-5 (Figure 2.5). Indeed, one of these<br />

compounds, namely isoreticular MOF–6 (IRMOF-6, Figure 2.5.6), displays rather high<br />

methane storage capacity (155 cm 3 (STP) / cm 3 ] at 298 K and 36 atm.<br />

Figure 2.10. Isoreticular expansion of HKUST-1. From H. Furukawa, K. E. Cordova, M. O’Keeffe<br />

and O. M. Yaghi, Science, 2013, 341. [66] Reprinted with permission from AAAS.


16 Chapter 2<br />

Expansion of this structures (i.e. isoreticular series of HKUST-1) could be obtained via<br />

employing other tritopic linkers of various lengths such as TATB 3- (4,4’,4’’-(s-triazine-<br />

2,4,6-triyl-tribenzoate) [61] and BBC 3- (4,4’,4’’-(benzene-1,3,5-triyl-tris(benzene-4,1-<br />

diyl))tribenzoate), for example (Figure 2.10). [62] In facts, the cell volume of the largest<br />

reported member of this series (MOF-399 ([Cu3(BBC)2]n)) is 17.4 times than that of<br />

HKUST-1. Moreover, the isoreticular construction of MOFs to enlarge the pores is also<br />

implemented through the expansion of the original phenylene unit of MOF-74-M [63] or<br />

CPO-27-M (M2(dobdc), dobdc = 2,5-dioxido-1,4-benzenedicarboxylate; M = Mg, Mn, Fe,<br />

Co, Ni, Zn, etc.). [64-65]<br />

Synthesis of MOFs, especially reticular synthesis, has earlier mainly focused on the<br />

preparation of new frameworks with new topologies and open structures with high<br />

porosity which are aimed for the gas storage capacities such as hydrogen, methane and<br />

carbon dioxide. [25, 60] Nowadays, the applications of MOFs have been more widely<br />

explored, including some fields of physics and biology. Consequently, modification of<br />

MOFs towards targeted applications has been more widely developed as well.<br />

2.3.3 Way to the modification of MOFs<br />

Beyond the traditional synthesis approaches, controlled modifications of MOFs at the<br />

organic linkers and/or at the metal nodes, could be mainly realized in several approaches:<br />

2.3.3.1 Direct decoration on the utilized linkers<br />

Ligands are often decorated with various organic groups to provide guest-accessible<br />

functional sides. [67-69] For example, Kitagawa et al. reported [Cd(4-<br />

btapa)2(NO3)2]n∙6H2O∙2DMF (4-btapa = 1,3,5-benzene tricarboxylic acid tris[N-(4-<br />

pyridyl)amide]) framework with amide-functionalized linker, where the highly ordered<br />

amide groups play an important role in the interaction with the guest molecules. [68]<br />

Moreover, the material shows selective base catalytic activity in the Knoevenagel<br />

condensation reaction. Furthermore, introduction of the thioether side CH3SCH3CH3Schain<br />

to the 2,5-position of 1,4-bdc and its co-assembly with Zn(II) ions lead to porous<br />

cubic network [Zn4O(L)3], whose topology is the same as MOF-5. [69] The obtained material<br />

exhibits a notable sensing response to nitrobenzene in the form of fluorescence<br />

quenching. Besides, it is able to absorb HgCl2 from the ethanol solution at low


Chapter 2 17<br />

concentrations (e.g. 84 mg/L). The pre-functionalization method is, however, somehow<br />

limited because certain kinds of groups favor to coordinate to metal ions, which result<br />

into frameworks with completely blocked organic sides.<br />

2.3.3.2 Introduction of open metal sites<br />

Open metal sites are CUS where an actual coordination number of the metal center is<br />

lower than typically expected for its oxidation state, charge or actual ligand field. Such<br />

CUSs can serve as binding sites for certain guest molecules (hydrogen, methane, etc.) as<br />

well as catalytic active sites for particular reactions under Lewis acidic conditions. The<br />

most common method to obtain CUSs is to activate (i.e. heat under vacuum) MOFs to<br />

remove the metal-bound volatile species (e.g. water, DMF, methanol, etc.). Indeed, this is<br />

often a case for several well-known MOFs such as HKUST-1 ([Cu3(BTC)2]n) and other<br />

members of this family, as well as MOF-74 ([(M2(dobdc)H2O]n) (Figure 2.11). [70-71]<br />

Figure 2.11. Representation of the crystal structure of HKUST-1 as well as the generation of Cu-<br />

CUS at the PW units upon activation. Green, Cu atoms; red, O atoms; gray, C atoms. Hydrogen<br />

atoms are omitted for clarity.<br />

Apart from this conventional method, the unsaturated metal centers can be introduced by<br />

employing similar metal chelating dicarboxylates, [72-74] 2,2’-bipyridine-5,5’-dicarboxylate


18 Chapter 2<br />

(H2BPDC) [75] or other bridging ligands, [76] which are not part of the inorganic building<br />

blocks of a given framework. Moreover, MOFs with open metal sites featuring diverse or<br />

enhance property can be achieved by using other metal ions in place of those metal<br />

precursors used in the known MOFs. By directly employing metal precursors containing<br />

another metal-ions and the same linkers during the synthesis of MOFs, the formation of<br />

the isostructural MOFs is expected. For example, analogs of HKUST-1 with the general<br />

formula [M3(BTC)2]n where M =Zn, [77] Cr, [78] Ni, [79] Fe, [80] Mo, [81] and Ru [82] etc. were<br />

reported by serveal groups after the appearance of [Cu3(BTC)2]n. [35] These analogs of<br />

HKUST-1 behave rather different properties such as in gas sorption/selectivity. [78,<br />

83] Similarly, by varying the metal in the infinite inorganic rod-type SBUs utilized for the<br />

construction of MOF-74,([Zn2(dobdc)]n) [63] a series of isostructural M-MOF-74 with<br />

various divalent metal ions like Mg, Co, Ni, Mn, Cd, Cu and Fe were described. [84-88] Other<br />

families of the homologous structures prepared using the same principle are M-BTT (BTT<br />

= 1,3,5-benzenetristetrazolate, M = Mn, Fe, Co, Cu, Cd), [89-92] for example.<br />

2.3.3.3 Post-synthetic modification and beyond<br />

Post-synthetic method (PSM), where the modification of ligands or metal containing<br />

nodes is implemented after the formation of MOFs (Figure 2.12), offers an effective way<br />

to overcome the potential for functional-group interference during MOF assembly.<br />

Several research groups, in particular the group of Cohen, have investigated the<br />

modification of amino and aldehyde groups in MOFs via the formation of new covalent<br />

bonds (Figure 2.12 top). [93-100] Still, employing PSM, one can face the problem of the pore<br />

volume decrease (caused by the functionalization reaction), which can lead to the<br />

reduction of gas absorption capacity. Therefore, an optimized method called “postsynthetic<br />

deprotection” was developed (Figure 2.12 bottom). [101] In this case, an organic<br />

linker bearing protected or “masked” functional group is incorporated into a MOF under<br />

standard solvothermal conditions, and the protecting group is subsequently removed by<br />

a thermal [102-103] or light [98, 104] induced deprotection reaction to reveal the desired<br />

functionality. One of the earliest example was presented by Kitagawa with the term of<br />

“protection-complexation-deprotection” process. [102] During the protection step,<br />

hydroxyl group of the 2,5-dihydroxyterephthalic acid (H2dhybdc) was protected by acetyl<br />

group via acetylation to form the 2,5-diacetoxyterephthalic acid (H2dacobdc). In the<br />

complexation/deprotection step, a MOF with a layered-pillared structure was obtained by


Chapter 2 19<br />

the rection of H2dacobdc, bipyridine and Zn(II) in DMF, with the simultaneous removal of<br />

the acetoxyl groups completely. This strategy was found to be useful for preventing<br />

interpenetration, introducing functionality, and controlling pore diameter. [105]<br />

Figure 2.12. A general scheme illustrating the concept of post-synthetic modification (PSM) on<br />

organic likers or metal-containing nodes of MOFs. Top, covalent PSM; Middle, dative PSM; Bottom,<br />

PSD. Reprinted with permission from S. M. Cohen, Chem. Rev., 2012, 112, 970-1000. [101] . Copyright<br />

(2012) American Chemical Society.<br />

Moreover, dative PSM (Figure 2.12 middle) involves heterogeneous chemical reactions to<br />

functionalize preassembled MOF structures via modification of metal-containing nodes<br />

(SBUs) [106-107] or organic linkers [108] can be reached via the formation of dative (i.e. metalligand)<br />

bond. [101] One of the first representative studies on dative PSM was reported by<br />

Hupp et al. [106] It was found that axially bound DMF solvent molecules in the PW SBUs<br />

could be fully removed from the Zn(II) PW-derived MOF by heating under vacuum.<br />

Subsequently the desolvated MOF was treated with pyridine derivatives in CH2Cl2, leading<br />

to materials in which the axial sites left vacant on the SBUs can be bound with pyridine<br />

derivatives. Those obtained MOF materials demonstrated dependent H2 uptake on the<br />

pyridine species coordinated to the SBUs, indicating properties could be<br />

modulated/optimized by dative PSM. Férey and co-workers reported amine grafting on


20 Chapter 2<br />

Cr III CUSs of MIL-101, displaying enhanced activities in the Knoevenagel condensation of<br />

benzaldehyde and ethyl cyanoacetate.<br />

Beyond the above mentioned PSM approaches, numerous conceptually different postsynthesis<br />

routes have been emerging. Intriguing building block replacement (BBR), which<br />

involves replacement of key structural components of the MOF, opens up a very broad<br />

strategy for the synthesis of functionalized isostructural MOFs featuring gradient<br />

compositions and chemical properties. Solvent-assisted linker exchange (SALE), [109-113]<br />

which is also termed as “stepwise synthesis”, [114] , “bridging linker replacement”, [114] “postsynthetic<br />

ligand exchange”, [115] isomorphous ligand replacement, [116] , is one of the<br />

important method to achieve this kind of modification on MOFs. Conceptually, by the<br />

reaction of a parent MOF and a solution of a second linker, a daughter MOF retaining the<br />

parent MOF topology can be obtained. Besides, partial substitution of the metal ions in a<br />

given framework with other metal ions of similar size and coordination chemistry could<br />

be accomplished by the so-called “post-synthetic metal ion exchange”, [117] (also known as<br />

“transmetalation”, [118] “metal metathesis” [119] or “metal-ion exchange” [120-121] ). Kahr et al.<br />

obtained a series of mixed-metal material Mg/Ni-MOF-74 with various degrees of Ni 2+<br />

incorporation by the post-treatment of Mg-MOF-74 and Ni 2+ solution with weak acid. [122]<br />

These materials showed increased stability and improved accessible porosity compared<br />

with the parent Mg-MOF-74 and Ni-MOF-74.<br />

2.3.3.4 Mixed-component co-assembly MOFs<br />

In addition of the post-synthetic modification on MOFs, co-assembly mixed-component<br />

(MC) MOFs which is related to the co-crystallizing of linker or metal-nodes mixtures for<br />

the introduction of functionalization should not be overlooked. Several research groups<br />

reported simultaneous utilizing two or more functionalized linkers during synthesis<br />

yielding to the formation of mixed-linker MOFs (MIXMOFs) or multivariate MOFs (MTV-<br />

MOFs). [123-128] In fact, Yaghi et al. revealed that a number of linkers with distinct functional<br />

groups (up to eight) can be incorporated into MOF-5 within one phase (Figure 2.13). [123]<br />

Curiously, the properties of MTV-MOFs are not just simply a linear combination of those<br />

of the pure components. For instance, MTV-MOF-5-EHI, as a member of this series,<br />

demonstrates strikingly better selectivity for CO2/CO in comparison with its best samelink<br />

counterparts. Moreover, partial metal substitution can be achieved by directly<br />

copolymerization via mixed-metal solid solutions, [126, 129-135] where more than one type of


Chapter 2 21<br />

metal-precursors are employed to react with the organic linkers during the synthesis<br />

(Figure 2.14). For instance, partial substitution of Cu 2+ by Ru 3+[134] and Zn 2+[133] in the PWs<br />

of HKUST-1 were obtained by mixing the Cu-salt with the corresponding Ru-salt and Znsalt<br />

directly during the synthesis, respectively. Orcajo et al. prepared Co-doped MOF-5 via<br />

co-assembly of Co(NO3)2·6H2O, Zn(NO3)2·4H2O and H2bdc during the synthesis. This<br />

obtained material exhibited enhanced adsorption capacities for H2, CO2, and CH4 at high<br />

pressure in comparison with Co-free MOF-5.<br />

Figure 2.13. Schematic representation of MTV-MOF-5 structures containing up to eight different<br />

functionalities distributed in one phase. From H. Deng, C. J. Doonan, H. Furukawa, R. B. Ferreira, J.<br />

Towne, C. B. Knobler, B. Wang and O. M. Yaghi, Science, 2010, 327, 846-850. [123] Reprinted with<br />

permission from AAAS.


22 Chapter 2<br />

Figure 2.14. Schematic presentation of the metal-based solid solution approach by mixing two<br />

different metal centers during the synthesis. Inspired by Burrows. [126]<br />

2.3.3.5 Defect-engineered MOFs<br />

On the basis of the concept of mixed-component MOFs, a very new approach to introduce<br />

functionalized MOFs has been reported very recently. [136] This is related to generation of<br />

so-called “defect-engineering MOFs” (DEMOFs), where in most cases the CUSs are<br />

modified or generated through the incorporation of additional “defect” linkers,<br />

modulators or through post-synthetic treatment. [137] The distinction of DEMOFs mainly<br />

related to the degree of deviation of the local structure as well as the long-ranging<br />

ordering of a variable fraction of the linkers and metal nodes in comparison with the<br />

parent MOF. [138-143] However, this kind of deviation barely can be observed from MC-<br />

MOFs. Only intrinsic defect may be present in certain conditions. [144-145] Besides, another<br />

important features of DEMOFs which should not be overlooked is the correlation and<br />

clustering of defect sites. Goodwin et. al used a combination of measurements and<br />

demonstrated the controlled introduction of nanoscale correlated defects within a<br />

hafnium terephthalate MOF UiO-66. [146] In particular, Fischer et al. reported the<br />

incorporation of pyridine-3,5-dicarboxylate (pydc) into the HKUST-1 and its Ru analog<br />

(Ru-MOF). [138-139] Remarkably, doping pydc resulted in the formation of mCUSs in the<br />

obtained Ru-DEMOFs, in which along with the Ru 2+ -/Ru 3+ -species typically found in the<br />

parent Ru-MOF framework, reduced Ru δ+ -species (0 < δ < 2) were generated. Moreover,<br />

such Ru-mCUSs appeared to be responsible for the activity of Ru-DEMOFs in the CO<br />

sorption as well as the catalytic hydrogenation of olefins. [138] Furthermore, doping pydc<br />

or other similar defect linker (DL) such as 5-hydroxyisophthalic acid (5-OH-ip) into<br />

HKUST-1(Cu-MOF), the creation of mesopores in the final Cu-DEMOFs was observed<br />

dependent on the utilized Cu-precursors and the incorporated level of DL. [139] Due to


Chapter 2 23<br />

appealing perspectives of this method on CUSs modification/introduction, this method is<br />

expected to receive much attention in next years. This kind of modification of MOFs are<br />

also involved in this dissertation. Further details will come in Chapter 4 and 5.<br />

2.3.4 Application of MOFs<br />

What have been addressed above show rather a broad picture demonstrating that MOFs<br />

are attractive porous materials with a huge diversity of structures, functionalities, and<br />

facile modification. All these make MOFs very promising for a range of applications<br />

(Figure 2.15), [147] especially in gas storage, selective adsorption and separation.<br />

Hydrogen and methane are good candidates for on-board fuel. A tank loaded with porous<br />

absorbent allows gas to be stored at much lower pressure in comparison with the<br />

identical tank without an adsorbent. Due to the high porosity and well-defined structures,<br />

MOFs has gained significant attention being a new class of adsorbents. A lot of MOFs have<br />

been evaluated for gas storage such as hydrogen and methane, [24, 26] which provides a<br />

safer and more economical gas storage method. For example, MOF-5 has shown a rapid<br />

and fully-reversible H2 storage density (66 g L -1 at 77 K, 100 bar), [148] which is very close<br />

to the value observed for liquid hydrogen (71 g L -1 at 20.4 K, 1 bar). Later, it was found<br />

that most of the MOFs like MOF-5 demonstrate poor performance at 298 K due to rather<br />

weak interactions between H2 and the framework. However, remarkable enhancement of<br />

H2 adsorption enthalpy can be achieved by designing MOFs with CUSs, [65, 70, 89]<br />

catenation/interpenetration [149] or MOFs with heavy transition metals (e.g. Pd) which can<br />

afford spillover effect for hydrogen storage. [150] For instance, respectively designed<br />

Mn3[(Mn4Cl)3(BTT)8]2 bearing CUSs exhibits H2 uptake of 1.49 total wt% and 12.1 g L -1 at<br />

298 K and 90 bar, [89] what is 77% greater than that of compressed H2 under the same<br />

conditions. Furthermore, HKUST-1 and Ni2(dobdc) have also demonstrated high total<br />

volumetric methane uptakes at 35 bar (225 and 235 v/v, respectively). [26] Significantly,<br />

PCN-14 (Cu2(adip) (adip 4- = 5,5’-(9,10-anthracenediyl)di-isophthalate) containing Cu-<br />

CUSs is also one of the best reported so far MOFs for methane storage (230 v/v, 290K, 35<br />

bar). Note, that 35 bar is the maximum pressure achievable by most inexpensive singlestage<br />

compressors, [151] which is also a widely used standard pressure for evaluating<br />

adsorbents for adsorbed natural gas storage. Both rigid and flexible MOFs as well as MOF<br />

membranes can be utilized as adsorbents for selective gas adsorption on the basis of


24 Chapter 2<br />

adsorbate-surface interactions and/or size-exclusion (molecular sieving effect), such as<br />

selective adsorption of N2/O2,CO2/N2, CO2/H2, etc. [23, 152-156] Moreover, separation of<br />

alkane isotherms from natural gas is also able to be achieved by suitable design of<br />

MOFs. [22-23, 157-158]<br />

Figure 2.15. Widespread potential applications of MOFs. Reproduced from S. Chaemchuen, N. A.<br />

Kabir, K. Zhou and F. Verpoort, Chem. Soc. Rev., 2013, 42, 9304-9332, with permission of The Royal<br />

Society of Chemistry. [159]<br />

Another widely investigated direction of MOFs applications is catalysis. Size- and shapeselective<br />

catalytic behavior could be achieved by utilizing MOFs with suitable porosity and<br />

functional groups/sites. [68, 160] For instance, functionalized 3D MOF with amide groups,<br />

[Cd(4-btapa)2(NO3)2]n∙6H2O∙2DMF, demonstrates catalytic selectivity in the Knoevenagel<br />

condensation reaction due to the relationship between the size of the reactants and the<br />

pore window of the host. [68] Significantly, the presence of catalytically active transitionmetal<br />

centers (loaded nano-particles (NPs), grafted metal complex, or active metal sites)<br />

as well as the organic functional sites in MOFs also enable this kind of porous materials to<br />

be catalysts in a variety of reactions. [27-29, 161] It is typical, that CUSs in MOFs serve as<br />

Lewis-acid catalytic sites to speed up the reactions run under Lewis-acid conditions. [161]<br />

For example, both MIL-101([Cr3F(H2O)2O(bdc)3]) and HKUST-1(Cu3(BTC)2)n) featuring<br />

exposed CUSs have been reported as good catalysts for cyanosilylation of aldehydes. [162-<br />

163] Besides, Lin, Kim and Kaskel demonstrated applications of homochiral MOFs in


Chapter 2 25<br />

asymmetric catalysis. [164-166] Very recently, the application of MOFs as biomimetic<br />

catalysts has been reviewed as well. [167-168] Some researchers also exploit the advances of<br />

MOFs as photocatalysis. [169-170] Apart from these two highlighted fields, applications of<br />

MOFs are certainly studied also in magnetism (ferromagnetic, antiferromagnetic,<br />

ferromagnetic, frustration and canting properties), [171-173] proton [174-178] and electrical [179-<br />

181] conductivity, luminescence and sensors chemistry, [182-186] drug storage and<br />

delivery. [187-188] Thus, in the near future breakthroughs with the continual developments<br />

in these and new applications are expected.


3 Controlled secondary building unit approach to the<br />

formation of the isostructural Ru II,II and Ru II,III analogs of<br />

[M3(BTC)2]n<br />

Abstract<br />

Controlled secondary building unit approach (CSA) was employed to derive a series of<br />

ruthenium metal-organic frameworks (MOFs) of the general formula [Ru3(BTC)2Yy]n·G g<br />

(BTC = 1,3,5-benzenetricarboxylate; Y = counter-anion, G = guest molecules, 0 ≤ y ≤ 1.5),


Chapter 3 27<br />

which are structural analogs of [M3(BTC)2]n (M = Cu, Zn, Ni, Cr, Mo). As Ru-precursors for<br />

CSA mixed-valence [Ru2 II,III (OOCR)4X] and [Ru2 II,III (OOCCH3)4]A compounds, with various<br />

nature of the anions X and A (strong coordinating X (like Cl - ) and weakly coordinating A<br />

(like [BF4] - or [BPh4] - )) as well as alkyl groups at the carboxylate ligand R (R = CH3 or<br />

C(CH3)3 ), were employed. Additionally, [Ru2 II,II (OOCR)4] complexes without any counterions<br />

were tested as Ru-source for Ru-MOF preparation. Systematic studies were<br />

conducted and five phase-pure Ru-MOFs were obtained. The set of characterization data<br />

support indicated the analytical compositions and the structural analogy of the prepared<br />

Ru-MOFs with the [M3(BTC)2]n family. The oxidation state of the Ru-sites in the obtained<br />

solids was studied by the X-ray absorption near edge structure (XANES) spectroscopy.<br />

The chosen Ru-precursors, optimized synthetic and activation protocols allowed<br />

improvement of the overall crystallinity, purity (in terms of residual solvent molecules)<br />

and porosity of the Ru-MOF materials.


28 Chapter 3<br />

3.1 Preparation and investigation of the Ru II,III -analogs of [M3(BTC)2]n *<br />

3.1.1 The family of [M3(BTC)2]n<br />

Figure 3.1. 3D crystal structure of [Cu 3 (BTC) 2 ] n (Cu-HKUST-1) viewed along the (100) direction.<br />

Green, red and gray balls stand for Cu, O and C atoms, respectively. Hydrogen atoms are omitted<br />

for clarity.<br />

[Cu3(BTC)2]n (also known as HKUST-1 [35] or MOF-199 [36] , BTC = 1,3,5-<br />

benzenetricarboxylate), first reported by Chui et al. in 1999, represents one of the most<br />

well investigated MOFs synthesized at the early stages of MOF chemistry. The 3D<br />

structure of [Cu3(BTC)2]n, with the space group Fm-3m, features paddlewheel (PW)<br />

dicopper(II) tetracarboxylate building blocks where Cu II -ions adopt pseudo-octahedral<br />

coordination geometry with the axial sites occupied by H2O (Figure 3.1). These<br />

coordinated water molecules H2O can be easily removed by heating the samples under<br />

vacuum (i.e., activation) leading to exposed Cu II -ions, which subsequently could serve as<br />

active Lewis acid sites in catalytic reactions (Figure 3.2). [189-190] Moreover, the interaction<br />

* The main results of this part are covered in the following publication: W. Zhang, O. Kozachuk, R. A. Fischer,<br />

et al. Eur. J. Inorg. Chem. 2015, 3913–3920.


Chapter 3 29<br />

between gas molecules and such coordinatively unsaturated sites (CUS) center often<br />

enhance the general adsorption properties of the MOF materials. [191-192]<br />

Figure 3.2. Modification of the Cu-sites in Cu-HKUST-1 after activation (and i.e., formation of Cu-<br />

CUS) and substrate coordination. Green, red, gray and violet balls represent Cu, O, C atoms and<br />

substrate (either gas or liquid molecules), respectively.<br />

Given the presence of metal CUSs and the permanent porosity, a number of isostructural<br />

[M3(BTC)2]n frameworks with various 3d and 4d metal-ions (M = Mo, [81, 83] Cr, [78, 83] Fe, [80]<br />

Ni, [79, 83] Ru [82, 138] and Zn [77] ) were reported later. Among them, [Cr3(BTC)2]n featuring Cr II<br />

open metal sites exhibited reversible and highly-selective binding of O2. More<br />

interestingly, among this formally homologous series as suggested by the (simplified)<br />

general formula [M3(BTC)2]n, the Fe and Ru analogs (abbreviated hereafter as Fe-MOF and<br />

Ru-MOF) were found to possess mixed-valence M II,III PW units, which might be of great<br />

use due to their potential photo and redox properties. [193] However, Fe-MOF was reported<br />

with no measurable porosity while Ru-MOF presented quite high porosity as well as good<br />

thermal and chemical stability. In fact, as a consequence of the mixed-valence state, the<br />

actual chemical composition and microstructure of these (Fe, Ru)-MOFs turned out to be<br />

more sophisticated. Counter-ions (X) and strongly bound guest molecules (G) seem to be<br />

notoriously present even after activation. The incorporation of these components as a<br />

function of synthetic conditions must be taken into account. Thus, the empirical formula<br />

of these materials should be more precisely written as [M3(BTC)2Yy]n·G g.<br />

3.1.2 Background and the state-of-the-art in research on Ru-analogs of<br />

[M3(BTC)2]n<br />

Looking through all the reports, the parent material [Cu3(BTC)2]n (HKUST-1) can be<br />

obtained through solvothermal, [35] electrochemical [194] , microwave syntheses [195] and<br />

others [196-197] . However, the synthesis of the Ru-analog is far less investigated. [82, 193] The


30 Chapter 3<br />

formula of [Ru3 II,III (BTC)2Cl1.5]n was initially assigned and obtained either from RuCl3 or<br />

[Ru2(OOCCH3)4Cl]n. The powder X-ray diffraction (PXRD) data support an analogous<br />

structure as [Cu3 II,II (BTC)2]n. Element-analytical, magnetic, X-ray photoelectron<br />

spectroscopy (XPS) and probe molecule adsorption combined with IR spectroscopy<br />

pointed to a mixed-valence Ru2 II,III PW units as the framework’s nodes. The Cl - counterions<br />

are most likely strongly coordinated to the Ru-centers to balance the overall charge<br />

of the framework. This material shows both good chemical and thermal stability;<br />

however, the Ru-centers are partly occupied and coordinated Cl cannot be removed by<br />

standard activation protocols. Hence, the number of active sites for catalysis and gas<br />

sorption is obviously reduced. Later, another analog described as [Ru3(BTC)2][BTC]0.5 was<br />

reported by Dincă et al. [83] The authors employed [Ru2(OOCC(CH3)3)4(H2O)Cl] as<br />

ruthenium source. The absence of Cl - in the samples of [Ru3(BTC)2][BTC]0.5 was postulated<br />

based on elemental analysis data only, and extra framework BTC serving as counter-ion<br />

was suggested. This solid showed better Brunauer-Emmett-Teller (BET) surface area<br />

values compared to the [Ru3 II,III (BTC)2Y1.5]n prepared first in our group. However, the<br />

presence of guest molecules such as trimethyl-acetate or acetate (arising from CSA and/or<br />

synthetic conditions) in the structure still could not be unambiguously determined or<br />

ruled out. Thus, modification of the starting Ru-sources used for the synthesis seems to<br />

be interesting for tuning the metal sites (i.e., oxidation state and CUS), composition and<br />

stability of the HKUST-1 analogous Ru-MOFs, which in turn can have a profound impact<br />

on other properties such as porosity and catalysis, for instance.<br />

Previously [Ru3 II,III (BTC)2Y1.5]n has been studied where the mixed-valence Ru II,III state was<br />

characterized in particular by CO and CO2 adsorption monitored by the FT-IR<br />

spectroscopy at low temperature under ultra high vacuum (UHV) conditions. [198] Also, the<br />

co-assembly of BTC with pyridine-3,5-dicarboxylate (pydc) to yield mixed-component<br />

(linker-based) solid solution type samples of the formula [Ru3(BTC)2-x(pydc)xYy]n·G g was<br />

achieved. These so-called “defect engineered” Ru-MOFs showed interesting new<br />

properties (e.g. dissociative chemisorption of CO2). [138] Motivated by these promising<br />

results, it would be of great interest to investigate the synthesis of the parent Ru-MOF in<br />

more systematic fashion, dependence on Ru-precursors featured with diruthenium PW<br />

units (e.g. applying CSA), variation of the reaction conditions, work-up and activation in<br />

order to optimize the sample quality.


Chapter 3 31<br />

In this Chapter CSA is employed to systematically study the factors of synthetic conditions<br />

affecting the overall structure, composition and properties (in particular, thermal<br />

stability, porosity, etc.) of the Ru-based HKUST-1 analogous compounds (Scheme 7.3).<br />

Various Ru precursors, namely [Ru2(OOCR)4X] and [Ru2(OOCCH3)4]A, in which R = CH3 or<br />

C(CH3)3; X = Cl; A = BF4 or BPh4, were adopted as starting materials in order to modify the<br />

properties of the obtained Ru-MOF samples while maintaining the isoreticular<br />

relationship to HKUST-1. Switching from the -CH3 to bulkier R-group (e.g. -C(CH3)3)<br />

should increase the solubility of the polymeric ruthenium precursors, and, subsequently,<br />

would influence the kinetics of the MOF synthesis (better/faster nucleation and growth).<br />

On the other hand, employing weakly coordinating anions (WCA), like [BF4] - or [BPh4] - ,<br />

may allow modulation of the CUS within the Ru-MOF. Utilizing these counter-ions, A<br />

avoids the formation of a polymeric (insoluble) structure of the Ru precursors on one<br />

hand. On the other hand, it is anticipated that A being less prone to be incorporated into<br />

the framework via coordination to the Ru II,III -centers and, thus, may be removed by<br />

washing and activation of the samples. Unless, note, that H2O is always around in the<br />

solvothermal reactions and formation of -OH and its incorporation as component X must<br />

be also considered even in case of using WCA A, such as BF4 or BPh4.<br />

3.1.3 [Ru3(BTC)2Yy]n·G g obtained by CSA using [Ru2(OOCR)4X] and<br />

[Ru2(OOCCH3)4]A: synthesis and characterization<br />

3.1.3.1 CSA-precursors<br />

The Ru-precursors ([Ru2(OOCR)4X] and [Ru2(OOCR)4]A were prepared according to the<br />

literature (Scheme 7.1). [199-202] Characterization of all precursors confirmed the<br />

compositions and structures being expectedly analogous to the reported compounds. All<br />

structures consist of Ru2-clusters coordinated by four carboxylic groups. Thus, this kind<br />

of binuclear units (PWs) are similar to the building units in the [Cu3(BTC)2]n framework,<br />

affording the idea to construct the isostructural [Ru3(BTC)2]n by means of the CSA. The<br />

PXRD patterns of [Ru2(OOCCH3)4Cl] (SBU-a, R = CH3) showed the phase-purity of the<br />

compound (Figure 7.3). [203] Furthermore, following the well documented literature<br />

method, single crystals of [Ru2(OOCC(CH3)3)4(H2O)Cl](CH3OH) (SBU-b, R = C(CH3)3) were<br />

obtained. The cell parameters and other crystal structure data of SBU-b are presented in<br />

Table 7.1 and Figure 7.5. To note, the obtained crystallographic data are of superior


Intensity, a. u.<br />

32 Chapter 3<br />

quality as compared to the literature reference. Complementary characterizations such as<br />

1 H-NMR and PXRD reveal also high purity of SBU-b (Figures 7.6 and 7.7). In addition,<br />

[Ru2(OOCCH3)4(THF)2](BF4) (SBU-c) were obtained as the microcrystalline powder and<br />

was fully characterized (Figure 7.8). The synthesis of [Ru2(OOCCH3)4(H2O)2](BPh4) (SBUd)<br />

reported in the literature lead to the immediate decomposition of the product after<br />

filtration. [204] However, when we employed [Ru2(OOCCH3)4(H2O)2]2(SO4) instead of<br />

[Ru2(OOCCH3)4Cl] [199] as Ru-precursor, our target [Ru2(OOCCH3)4(H2O)2](BPh4) was<br />

successfully obtained (Figure 7.9).<br />

3.1.3.2 Synthesis, activation and thermal properties of Ru-MOF 1-4<br />

4<br />

3<br />

2<br />

1<br />

Cu-BTC_sim<br />

10 20 30 40 50<br />

2, degree<br />

Figure 3.3. PXRD patterns of the as-synthesized samples 1-4 as well as the simulated PXRD<br />

patterns of the reported [Cu 3 (BTC) 2 ](Cu-BTC_sim).


Chapter 3 33<br />

All Ru-MOFs 1-4 of the general formula [Ru3(BTC)2Yy]n·G g (Y = Cl, F, OH…; G = H2O,<br />

CH3COOH, (CH3)3CCOOH, H3BTC,…; 0 ≤ y ≤ 1.5) were obtained employing [Ru2(OOCR)4X]<br />

or [Ru2(OOCCH3)4]A precursors under solvothermal conditions in a similar way as we<br />

reported earlier. [82] The subsequent activation (i.e., removing of the incorporated<br />

solvent/guest molecules) was conducted at 150 °C for 24 h under dynamic vacuum (ca.<br />

10 -3 mbar). As revealed by the PXRD patterns (Figure 3.3), the prepared Ru-MOFs 1-4 are<br />

phase-pure crystalline solids that are isostructural to [Cu3(BTC)2]n. Ru-MOFs 2 and 4<br />

showed better crystallinity compared to the samples 1 and 3. The coordination of the<br />

carboxylic groups of framework BTC in the activated Ru-MOF samples is indicated by the<br />

strong-intensity bands at 1429 cm -1 and 1359 cm -1 in the FT-IR spectra (Figure 3.4).<br />

Notably, the Ru-MOFs obtained from [Ru2(OOCCH3)4]A, in which Y = [BF4] - and [BPh4] - ,<br />

show similar IR spectra with the other two samples. In other words, the incorporation of<br />

[BF4] - counter-ion in the structure of 3 can be ruled out, as no characteristic ν(B-F) band<br />

was observed at 1073 cm -1 . It is not possible to unambiguously determine the presence of<br />

[BPh4] - based only on the IR spectra because of the overlapping vibrations of [BPh4] - and<br />

BTC. Further analytical evidences in this respect will be discussed later.<br />

Figure 3.4. IR spectra of the activated Ru-MOFs 1-4.


34 Chapter 3<br />

After activating the Ru-MOFs, thermal gravimetric analysis (TGA) was performed<br />

demonstrating stability of the samples 1-4 up to 200 °C. Interestingly, Ru-MOF 4 is the<br />

most stable among the samples of discussed series and can be heated up to 250 °C without<br />

decomposition (Figure 3.5). Both as-synthesized and activated phases of the sample 1<br />

showed weight loss around 100 °C, which can be attributed to either a) loss of the residual<br />

coordinated solvent within the framework or b) adsorbed molecules such as water during<br />

transfer of the activated sample from the glovebox to the instrument. Ru-MOF 2 shows a<br />

slow weight decrease at low temperature range (30-150 °C), which is similar to the<br />

sample 1. Noteworthy, no similar observation was found in the curves of 3 and 4, meaning<br />

there are less residual solvent molecules trapped as compared to the other two samples<br />

1 and 2. Thus, it suggests that the pores of the frameworks 3 and 4 are more accessible<br />

and activation (desorption of the guest molecules) is facilitated.<br />

Figure 3.5. TG curves of the activated Ru-MOFs 1-4.


Intensity, a. u.<br />

Chapter 3 35<br />

3.1.3.3 Acid digestion and composition of the obtained Ru-MOFs 1-4<br />

4_ex<br />

4<br />

3_ex<br />

H(BTC)<br />

H(BPh 4<br />

)<br />

H(AcO)<br />

3<br />

2_ex<br />

H(PivO)<br />

2<br />

1_ex<br />

1<br />

10 8 6 4 2 0<br />

, ppm<br />

Figure 3.6. NMR spectra (measured in DCl/DMSO-d 6 ) of the Ru-MOF samples before (1-4) and<br />

after solvent exchange (1-4_ex) . All spectra were normalized by dividing the intensity of H(BTC),<br />

respectively. PivO = pivalate, AcO = acetate. Due to the sensitivity of water peak to the solution<br />

conditions, the proton resonance of H 3 O + varies here in the range of 4-6 ppm, dependent on the<br />

concentration of DCl in the total solution.<br />

To get more insight into the microstructures (i.e. analytical purity) and composition of the<br />

materials 1-4, 1 H-NMR analysis of the acid-digested (in DMSO-d6/DCl mixture) Ru-MOF<br />

samples was performed. Thus, the 1 H-NMR spectra revealed, that besides the signals at δ<br />

= 8.61 ppm attributed to the aromatic protons of BTC, one additional resonance at δ = 1.88<br />

ppm is seen and can be assigned to the protons of the residual acetic acid (AcOH) for all<br />

four Ru-MOFs samples (Figure 3.6). For Ru-MOF 2, the spectrum shows also an additional<br />

signal at δ = 1.05 ppm, which corresponds to the protons of the residual pivalic acid<br />

(PivOH) (Figure 3.7). The spectra suggest that the larger R-group (-C(CH3)3 instead of -<br />

CH3) in the Ru-SBU leads to more impurity in the synthesis of MOFs in addition to the


36 Chapter 3<br />

presence of the residual acetic acid in Ru-MOF 1, 3 and 4. Note, acetic acid has been a<br />

major component of the solvent medium during the synthesis. In the NMR spectrum of the<br />

sample 4 (Figure 7.11), signals assigned to trace amount of [BPh4] - were detected. This<br />

residue can be removed completely after washing and solvent exchange (for further<br />

explanation see the next section). The integrations of these signals for each Ru-MOF are<br />

listed in Table 3.1. In summary, Ru-MOF 3 and 4 contain lower amounts of the residual<br />

acetic acid. This again brings to the conclusion that these two samples are analytically<br />

purer as compared to the 1 and 2. Moreover, samples of 3 and 4 could be more<br />

quantitatively digested by the employed protocol in the same amount of solvent<br />

DCl/DMSO-d6 mixture, providing, therefore, more intense signals in the NMR spectra for<br />

each Ru-MOF sample.<br />

SBU-b<br />

2<br />

H(BTC)<br />

H(AcO) H(PivO)<br />

10 8 6 4 2 0<br />

, ppm<br />

Figure 3.7. Acid-digested (DCl) 1 H-NMR spectra of SBU-b (top) and 2 (bottom) (solvent: DMSOd<br />

6 ). The tiny peaks at around 0.8 and 1.2 ppm are due to H-grease.


Chapter 3 37<br />

Table 3.1. Comparison of the integrals of the 1 H NMR signals in the samples 1-4 (before solvent<br />

exchange) in comparison with 1-4_ex (after solvent exchange) to determine the linker to guest<br />

molecular ratios (BTC = 1, 3, 5-benzentricarboxylate, AcO = acetate, PivO= pivalate).<br />

Ru-MOFs H(BTC) : H(AcO) H(BTC) : H(PivO)<br />

1 1 : 2.4 -----<br />

1_ex 1 : 0.8 -----<br />

2 1 : 1.5 1 : 1.5<br />

2_ex 1 : 1.2 1 : 0.8<br />

3 1 : 1.6 -----<br />

3_ex 1 : 0.9 -----<br />

4 1 : 0.7 -----<br />

4_ex 1 : 0.7 -----<br />

3.1.4 Pre-activation by solvent exchange<br />

The presence of high amounts of acetic acid and pivalic acid arising from both, the Ru<br />

precursor and the reaction solvent mixture (which contains excess of acetic acid)<br />

apparently leads to partly pore-filling with the residual solvent (acid), which<br />

consequently may considerably affect or restrict the Ru-MOF properties. As revealed,<br />

these components cannot be removed completely regardless of the applied activation<br />

procedure (i.e. 150 °C for 24 h under dynamic vacuum (ca. 10 -3 )). Our attempts to slightly<br />

modify the reaction conditions, namely by reducing the quantity of acetic acid and/or<br />

using acetic acid free conditions yielded either amorphous products and/or prevented the<br />

substitution of RCCOOH in Ru-SBUs, affording just non-reacted starting materials (Figure<br />

7.12). Noteworthy, the reactants concentration in synthesis of the Ru-MOF 1 is quite<br />

crucial and attempting to vary the amounts of solvent with the same molar ratio of Ruprecursor<br />

and H3BTC was not so successful to get crystalline materials. Therefore, we<br />

conducted additional experimental studies directed to improve the activation procedure<br />

(i.e., minimize the amount of the strongly retained guest molecules) and, thus, improve<br />

the sample quality and access to the pores and open metal sites. For this reason we<br />

introduced an additional step to the activation procedure earlier mentioned. [138] Namely,


Intensity, a. u.<br />

38 Chapter 3<br />

the as-synthesized Ru-MOF samples were soaked in a large amount of (coordinatively)<br />

more inert solvent (water or ethanol, 80-100 ml) and stirred (for an average of 4 h) in<br />

order to wash out or replace strongly coordinated/adsorbed acetic or pivalic acid.<br />

Subsequently, the solvent was replaced by fresh portion and the samples were left<br />

standing overnight. This “solvent exchange” was repeated for each sample three times.<br />

Afterwards the samples were collected and dried under ambient conditions. The washed<br />

and dried samples (1-4_ex) were then “typically” treated again according to the<br />

established activation procedure for Ru-MOF: heating at 150 °C for 24 h under dynamic<br />

vacuum (ca. 10 -3 bar) in order to finally remove all adsorbed solvent and residual guests.<br />

3.1.5 Study of [Ru3(BTC)2Yy]n·G g after solvent exchange<br />

3.1.5.1 Phase purity and stability<br />

4-ex<br />

3-ex<br />

2-ex<br />

1-ex<br />

Cu-BTC_sim<br />

10 20 30 40 50<br />

2, degree<br />

Figure 3.8. PXRD patterns of the Ru-MOF samples after solvent exchange (1-4_ex) in comparison<br />

with the simulated patterns of Cu-HKUST-1 (Cu-BTC_sim).


Chapter 3 39<br />

Figure 3.9. IR spectra of the activated Ru-MOFs before (1-4) and after “solvent exchange” (1-<br />

4_ex).<br />

The PXRD patterns of the samples 1-4_ex match well with those of the [Cu3(BTC)2]n,<br />

suggesting fully preserved structural integrity of the Ru-frameworks after employed<br />

solvent exchange procedure (Figure 3.8). The IR spectra of the 1-4_ex further substantiate<br />

that the structure of the Ru-MOFs is retained (Figure 3.9). Both reveal the chemical<br />

stability of the 1-4 after solvent exchange. As expected, the amount of residual guest<br />

molecules (in this case, acetic acid, etc.) hosted by the frameworks of 1, 2 and 3 decreases<br />

considerably after applying an additional “solvent exchange”, which is clearly indicated<br />

by the comparison of the 1 H-NMR data of the digested and activated Ru-MOFs samples<br />

before (1-3) and after solvent exchange (1-3_ex) (Table 3.1 and Figure 3.6). In fact, the<br />

integration of the signals attributed to AcOH (δ = 1.88-1.93 ppm) in 1_ex and 1<br />

corresponds to 0.8:1 vs. 2.4:1. Notably, pivalic acid in 2 was removed even to a greater


40 Chapter 3<br />

extent than acetic acid (Table 3.1). At the same time, application of solvent exchange to<br />

Ru-MOF 4, containing the lowest amount of the guest molecules, does not have significant<br />

effect. The integration of the signals due to acetate protons reveals the same values,<br />

suggesting the amount of the acetate before and after exchanging is constant. Comparing<br />

all 1-4_ex samples, Ru-MOF 4 is the most promising one, with the lowest amount of guest<br />

molecules G. In addition, the tiny amount of residual [BPh4] - in 4 was removed after<br />

solvent exchange (Figure 3.6), as confirmed by the absence of the corresponding proton<br />

signals at 7.3 ppm.<br />

Figure 3.10. TG curves of the Ru-MOFs obtained from various Ru-SBUs before (1-4) and after (1-<br />

4_ex) solvent exchange procedure.<br />

From the TGA 1-4_ex samples (Figure 3.10) it is evident that these materials still possess<br />

good thermal stability after utilized solvent exchange and do not decompose until 200 °C.<br />

The shape of the TG curves of the 1_ex and 3_ex does not change after solvent exchange<br />

and suggest a decomposition temperature of 200 °C and 250 °C, respectively (Figure 3.10).<br />

In contrast, of the thermal stability of Ru-MOF 2 substantially decreases after solvent


Chapter 3 41<br />

exchange. This might stem from the presence of at least two different guest molecules<br />

(acetic acid and pivalic acid), meaning there are more molecules which sustain the whole<br />

framework in some extent through (weak) coordination compared to 1 and 3. Hence, the<br />

stability considerably decreases after solvent exchange while maintaining the structure<br />

(Figure 3.10). The same behavior is observed for Ru-MOF sample 4, most likely due to the<br />

removal of [BPh4] - counter-ions, which previously stabilized the framework (Figure 3.10).<br />

3.1.5.2 Composition determination<br />

To quantitatively determine the composition of Ru-MOFs 1-4_ex elemental analysis (EA)<br />

and Energy-dispersive X-ray spectroscopy (EDX) were subsequently performed. From<br />

our earlier reports, [82, 138] EA results indicated the molar ratio of Ru : Cl being 3 : 1.5. The<br />

XPS studies showed two different Ru-species (Ru III and Ru II ) as well as the presence of<br />

chlorine, suggesting Cl - being the major couter-ion to balance the charge of the<br />

framework. [82] As revealed in the present studies, the molar ratio of Ru : Cl in 1_ex slightly<br />

differs from the sample before solvent exchange (Ru : Cl = 3 : 1.5) [138] and equals to ~ 3 :<br />

1 (Table 3.2). The presence of chlorine was additionally confirmed by the EDX elemental<br />

mapping of Ru and Cl (Figure 3.11). It should be noted, that the analytically determined<br />

Ru-content (weight percentage by AAS) in the exchanged materials (1_ex) decreased from<br />

36.29 % to 30.45 %. [82, 138] Nevertheless, all characterizations proved preservation of<br />

crystallinity, porosity as well as thermal stability (by TGA). Based on these observations,<br />

together with the preserved BET surface area (SBET) observed after solvent exchange (see<br />

the details for next section on porosity), creation of the internal “structural defects” (i.e.<br />

missing of Ru-nodes) in the solvent exchanged sample 1_ex could be expected. As<br />

mentioned earlier, acetic acid is present in the discussed Ru-MOFs. The residual acetic<br />

acid (or acetate) could be included into frameworks mainly in three ways: i) it can be<br />

deprotonated and act as the counter-ion (coordinated to the Ru-center or sitting around<br />

the PW units) to compensate the charge of the [Ru2] 5+ units (Figure 3.12a); ii) inherent<br />

structural defects due to the incomplete substitution by the BTC within used Ru-SBUs<br />

(Figure 3.12b and c); iii) the acetic acid molecules (without deprotonation) might be<br />

accommodated within the pores of the frameworks. Activation procedure of the assynthesized<br />

samples could hardly direct remove the residue of the acetic acid in the<br />

framework. Nonetheless, application of additional solvent exchange method results in<br />

partly elimination of the residual acetic acid, and might consequently create the missing


42 Chapter 3<br />

Ru-nodes. Therefore, relatively reduced Ru-content is observed in the sample 1_ex. This<br />

kind of observation is mostly likely related to two possibilities: i) the strong stirring<br />

during the washing procedure; ii) slower substitution kinetics on the acetate coordinated<br />

Ru2 II,III -PW centers in comparison with Cu2-PW centers (rapid substitution kineties).<br />

Table 3.2. The obtained elemental analysis data and the calculated formula of 1-4_ex after solvent<br />

exchange. BTC = 1,3,5-benzenetricarboxylate, H 3 BTC = 1,3,5-benzenetricarboxylic acid, AcO =<br />

acetate, AcOH = acetic acid, PivOH =pivalic acid.<br />

Samples and Formula wt% C wt% H wt% Cl wt% F wt% Ru<br />

1_ex<br />

[Ru 3 (BTC) 2 Cl(AcO) 0.5 ] n·(AcOH) 1.5 (<br />

H 3 BTC) 0.1 (H 2 O) 4<br />

2-ex<br />

[Ru 3 (BTC) 2 Cl 1.2 (OH) 0.3 ] n·(H 3 BTC)<br />

0.15(AcOH) 2.4 (PivOH) 0.45<br />

3_ex<br />

[Ru 3 (BTC) 2 F 0.7 (OH) 0.8 ] n·(AcOH) 2.4<br />

(H 3 BTC) 0.5 (H 2 O) 3<br />

4_ex<br />

[Ru 3 (BTC) 2 (OH) 1.5 ] n·(H 3 BTC) 0.8 (A<br />

cOH) 1.4 (H 2 O) 3<br />

Found 27.26 1.65 3.41 - 30.45<br />

Cal. 28.5 2.31 3.67 - 31.40<br />

Found 31.57 2.11 4.16 - 29.27<br />

Cal. 32.14 2.18 4.31 - 30.73<br />

Found 29.28 1.28 - 1.09 24.18<br />

Cal. 31.30 2.44 - 1.27 29.90<br />

Found 32.15 1.60 - - 26.64<br />

Cal. 32.05 2.30 - - 28.90<br />

Figure 3.11. SEM image and the EDX elemental Ru- and Cl-map of the 1_ex sample.


Chapter 3 43<br />

Figure 3.12. Possible coordination of the acetic acid in the obtained Ru-BTC structures.<br />

In case of Ru-MOF 2, EDX elemental mapping images (Figure 3.13) also indicate the<br />

presence of chlorine. Besides, EA results indicated the Ru : Cl molar ratio to be ~ 3 : 1.2,<br />

suggesting Cl - serving here also as a major counter-ion to balance the charge of the [Ru2] 5+<br />

PW units, in analogy to 1. The EDX, EA results combined with the 1 H-NMR data of 2_ex<br />

match well with a formula [Ru3(BTC)2Cl1.2(OH)0.3]n·(H3BTC)0.15(AcOH)2.4·(PivOH)0.45,<br />

which includes a small amount of unreacted H3BTC, possibly residing in the pores rather<br />

than being deprotonated (as counter-ions). Although Ru-MOF 2_ex features higher<br />

crystallinity than other Ru-MOFs obtained from various Ru-SBUs, changing the R groups<br />

to -C(CH3)3 in the starting ruthenium precursor makes the whole system more<br />

complicated by introducing a new kind of impurity (pivalic acid) with respect to the<br />

analytical purity.<br />

Figure 3.13. SEM image and the EDX elemental maps for Ru and Cl of the 2_ex sample.<br />

In case of sample 3, the Ru : F molar ratio calculated from the EA data of the 3_ex is ~3 :<br />

0.7, suggesting the presence of strongly coordinated F - counter-ions. However, as the F-<br />

amount is rather low, the distribution of F could not be mapped via EDX (Figure 3.14).<br />

Still, this agrees with the absence of BF4 (monitored by IR) and can be explained by in situ


44 Chapter 3<br />

decomposition of [BF4] - into F - and BF3 during the MOF synthesis. As expected, lower<br />

amounts of counter-ions were observed compared to 1 and 2, when adopting WCAs as<br />

counter-ions in the starting Ru-SBU. Although, there are guest molecules in the material,<br />

the characterization of the Ru-solids after applied solvent exchange procedure indicates<br />

a better purity control (i.e. decrease of the types and amounts of guest molecules) than in<br />

case of 2.<br />

Figure 3.14. SEM image and the EDX elemental maps for Ru and F of the 3_ex sample.<br />

Based on the EA and NMR analyses as well as the assumption of the presence of [OH] - to<br />

help the charge balance, the empirical formula of Ru-MOF 4_ex matches well with<br />

[Ru3(BTC)2(OH)1.5]n·(H3BTC)0.8(AcOH)1.4(H2O)3. In this case our data support the<br />

conclusion that the strongly coordinating counter-ion Y of the final mixed-valence Ru-<br />

MOF may not stem from the employed Ru-SBU.<br />

3.1.5.3 Porosity comparison<br />

In order to compare the porosity of HKUST-1 and its Ru-analogs more reasonably due to<br />

the difference on the metal centers, SBET were calculated in the unit of m 2 /mmol according<br />

to the values in m 2 /g unit and the corresponding molecular weight(Mw) of the samples. In<br />

comparison with the experimental SBET value of HKUST-1 (1049 m 2 /mmol), the overall<br />

SBET of the activated Ru-MOF solids, especially, sample 1_ex and 4_ex is measured as 964<br />

and 941 m 2 /mmol, respectively, which suggest the good porosity of these two Ru-MOFs<br />

(Table 3.3 and Figure 3.15). In other words, the porosity of the Ru-MOFs is not affected<br />

after “solvent exchange”. As was indicated by the NMR studies mentioned earlier,<br />

adopting SBU-b as the Ru-precursor revealed a decreased amount of residual AcOH in the<br />

final framework. However, at the same time another kind of residue was introduced and<br />

the total amount of guest molecules do not decrease much after applying the solvent<br />

exchange procedure. Consequently, the BET surface area of the 2_ex amounts 718


Chapter 3 45<br />

m 2 /mmol (based on the N2 isotherm) and does not differ much in contrast to the sample<br />

before “solvent exchange” (i.e. sample 2). Notably, the thermal stability of 2_ex also<br />

slightly decreased after solvent exchange like SBET. Hence, the overall quality of this Ru-<br />

MOF analog is less convincing under the circumstances of this study. Interestingly, in case<br />

of 3_ex, however, an increase of the SBET has been observed from the N2 sorption<br />

isotherms (850 vs. 683 m 2 /mmol) (Table 3.3). The same increase of SBET has also been<br />

observed for the sample 4_ex (941 m 2 /mmol) in comparison with the value before (916<br />

m 2 /mmol), where the residual [BPh4] - has been removed via solvent exchange procedure<br />

although the amount of acetic acid does not differ. Thus, in particular for Ru-MOFs 1 and<br />

3, solvent exchange procedure proved to be a good method to reduce the amount of<br />

residual acid molecules. In both cases the thermal stability, crystallinity and porosity were<br />

preserved. For the material 4, solvent exchange is helpful to wash out the residual [BPh4] -<br />

counter anions to give access to the metal-sites, while keeping the thermal stability and<br />

porosity.<br />

Table 3.3. Comparison of the BET surface area (S BET ) of the Ru-MOFs before (1-4) and after<br />

solvent exchange (1-4_ex) in the unit of m 2 /g and m 2 /mmol. S BET in the unit of m 2 /mmol is<br />

calculated by S BET (m 2 /g) and the corresponding molecular weight (M w ) determined from both EA<br />

and 1 H NMR results. M w of 1-4_ex is obtained from the formula given in Table 3.2. The M w of<br />

sample 1-4 is derived from the difference of the guest molecules between the samples before and<br />

after solvent exchange, namely: M w (1) = M w (1_ex + 3.2 AcOH); M w (2) = M w (2_ex + 0.6 AcOH +<br />

0.2 PivOH); M w (3) = M w (3_ex + 1.4 AcOH); M w (4) = M w (4_ex + 0.13 BPh 4 ).<br />

samples S BET (m 2 /g) S BET (m 2 /mmol)<br />

1 958 1108<br />

1_ex 998 964<br />

2 851 888<br />

2_ex 728 718<br />

3 604 683<br />

3_ex 812 850<br />

4 840 916<br />

4_ex 897 941<br />

[Cu 3 (BTC) 2 ] n<br />

* theoretical accessible surface area.<br />

1734 [83]<br />

2153* [205]<br />

1049<br />

1301


deriv. normalized E)<br />

46 Chapter 3<br />

Figure 3.15. N 2 adsorption (solid symbols) and desorption (open symbols) isotherms (recorded<br />

at 77 K) of the Ru-MOFs obtained from different SBUs. 1-4: before solvent exchange; 1-4_ex: after<br />

solvent exchange. Squares: 1 and 1_ex; triangles: 2 and 2_ex; diamonds: 3 and 3_ex; circles: 4 and<br />

4_ex.<br />

3.1.5.4 Oxidation state study<br />

SBU-a<br />

RuCl 3<br />

SBU-b<br />

SBU-c<br />

Ru<br />

SBU-d<br />

RuO 2<br />

22100 22120 22140 22160<br />

energy, eV<br />

Figure 3.16. XANES spectra (derivative normalized absorption vs. energy) of the Ru-SBUs in<br />

comparison with Ru-standard samples (Ru, RuCl 3 and RuO 2 ).


deriv. normalized E)<br />

Chapter 3 47<br />

In order to gather information on the valence of the Ru-sites in the Ru-MOFs, comparative<br />

X-ray absorption spectroscopy (XAS) measurements were subsequently carried out. The<br />

derivative of the edge-jump energy positions of Ru (22117 eV) and RuO2 (22131 eV) can<br />

be used as standards, representing the oxidation states +0 and +4, respectively. On the<br />

other hand, according to the literature, the oxidation state of Ru-center in SBU-a reveals<br />

+2.5. [203] Hence, the same position of derived edge-jump energy (ca. 22124 eV) for SBU-a<br />

- -d clearly indicates the oxidation state of Ru-centre to be +2.5 (Figure 3.16). To note, Ru-<br />

MOF sample 1_ex with the edge jump position at 22126 eV matches well with its mixed<br />

oxidation state documented in the literature by UHV-IR technique with CO probe. [82] All<br />

the other Ru-MOFs samples (2-4_ex) exhibit their edge-jump at the range of 22123-22128<br />

eV, suggesting the mixed oxidation state of the Ru-centers in their structure as well<br />

(Figure 3.17). The study on deviation of local coordination environments within these Ru-<br />

MOFs due to the change of the counter-ions (in particular by Extended X-Ray Absorption<br />

Fine Structure (EXAFS)) is one of our further direction.<br />

Ru<br />

Ru NP<br />

3_ex<br />

4_ex<br />

RuO 2<br />

2_ex<br />

1_ex<br />

22100 22120 22140 22160<br />

energy, eV<br />

Figure 3.17. XANES spectra (derivative normalized absorption vs. energy) of Ru-MOFs obtained<br />

from various Ru-SBUs (1-4_ex) in comparison with Ru-standard samples (Ru, Ru NP and RuO 2 ).


48 Chapter 3<br />

3.1.6 Summary<br />

Applying CSA, a series of Ru-MOFs [Ru3(BTC)2Yy]n·G g were synthesized under<br />

solvothermal conditions and fully characterized(Table 3.4). Varying the R-group and<br />

nature of the counter-ion in the Ru-SBU ([Ru2(OOCR)4X] and [Ru2(OOCCH3)4]A), tuning<br />

(up to a certain extent) the local environment of the metal sites while retaining the same<br />

structure were achieved. Changing the R-group at the carboxylate in the Ru-SBU from -<br />

CH3 to bulkier –Rs, for example -C(CH3)3, leads to a striking improvement of the<br />

crystallinity of the final MOFs. However, considering the MOF compositions, the whole<br />

picture gets complicated, as residual guest molecules/species G could stem from both<br />

different R-groups in the Ru-SBU ([Ru2(OOCR)4X]) and solvents (including used<br />

carboxylic acids).<br />

Table 3.4. Summary of the preparation, composition and S BET in [Ru 3 (BTC) 2 Y y ] n·G g (1-4_ex). BTC<br />

= 1,3,5-benzenetricarboxylate, H 3 BTC = 1,3,5-benzenetricarboxylic acid, AcO = acetate, AcOH =<br />

acetic acid, PivOH =pivalic acid.<br />

sample<br />

employed SBUs for synthesis<br />

Counter-ions<br />

Y<br />

guest<br />

molecule G<br />

S BET<br />

(m 2 /mmol)<br />

1_ex<br />

SBU-a<br />

[Ru 2 (OOCCH 3 ) 4 Cl] n<br />

Cl, AcO<br />

1.5AcOH,<br />

0.1H 3 BTC,<br />

4H 2 O<br />

963<br />

2_ex<br />

SBU-b<br />

[Ru 2 (OOCC(CH 3 ) 3 ) 4 (H 2 O)Cl](CH 3 OH)<br />

Cl, OH<br />

2.4AcOH,<br />

0.15H 3 BTC,<br />

0.45PivOH<br />

718<br />

3_ex<br />

SBU-c<br />

[Ru 2 (OOCCH 3 ) 4 (THF) 2 ](BF 4 )<br />

F, OH<br />

2.4AcOH,<br />

0.5H 3 BTC,<br />

3H 2 O<br />

850<br />

4_ex<br />

SBU-d<br />

[Ru 2 (OOCCH 3 ) 4 (H 2 O) 2 ](BPh 4 )<br />

OH<br />

1.4AcOH,<br />

0.8H 3 BTC,<br />

3H 2 O<br />

941<br />

Besides this, the counter-ion Cl - is still present to compensate the charge of the [Ru2] 5+ PW<br />

units. Therefore, keeping the same R-group (-CH3) in the Ru-SBU, the Cl - counter-ions<br />

were replaced by WCAs. As expected, the ruthenium centers were modified through<br />

introducing more weakly coordinating anions. The more coordinatively active the<br />

counter-ions in the Ru-SBU are, the easier they coordinate to the metal centers in the final<br />

MOF structure. In spite of the fact, that BF4 was absent in the structure of 3_ex, F - was<br />

found in the material by EDX and EA. Therefore, Ru-SBUs with more stable WCA, like BPh4,


Chapter 3 49<br />

were investigated as well. Remarkably, by introducing [BPh4] - to the Ru-SBUs, MOF<br />

materials that do not include BPh4 were successfully obtained. Interestingly, the amount<br />

of residual included guest molecules G in the modified MOFs (3_ex and 4_ex) decreases<br />

compared to respective Ru-MOFs obtained starting from [Ru2(OOCR)4Cl]. In particular,<br />

sample 4_ex which contains the lowest amount of the guest molecules (acetic acid), is the<br />

first material obtained in the series of Ru-analogs of [Cu3(BTC)2]n that includes no strongly<br />

coordinating counter-ions X originating from the used Ru-SBUs.<br />

Solvent exchange procedure was introduced to achieve better removal of the guest<br />

molecules without damaging the framework integrity after activation. As expected, after<br />

applying such improved activation procedure, the counter-ions [BPh4] - and most of the<br />

guest and solvent molecules (like acetic acid) were washed out of the material while the<br />

crystallinity and thermal stability are retained. These studies are particularly crucial for<br />

the future investigation and elaboration of these materials for targeted applications, such<br />

as in selective sorption and catalysis. In summary, the synthesis of the Ru-MOF is still<br />

difficult to optimize further. Nevertheless, the investigations herein afford better<br />

understanding and directions towards not only HKUST-1 analogs but also other MOFs<br />

with PW SBU.


50 Chapter 3<br />

3.2 Elaboration of Ru II,II analog of [M3(BTC)2]n<br />

The results summarized in this Chapter 3.2 are initial studies directed to solve the<br />

remaining issues mentioned in Chapter 3.1 and to target Ru II, II analog of [M3(BTC)2]n.<br />

The study mentioned earlier (in Chapter 3.1) involving the investigation of Ru II,III analogs<br />

of HKUST-1 shows rather a complicated picture. As described, the residual acids and the<br />

inevitable counter-ions in the frameworks, to some extent, obstruct the porosity of the<br />

Ru-MOFs. One can imagine that the catalytic activity and the sorption properties of these<br />

materials can be definitely enhanced if the CUSs could be fully available without the<br />

additional coordination (“blocking”) by the counter-ions Y (Cl - , F - , OH - or AcO - , etc.). The<br />

pre-activation procedure (i.e. solvent exchange) helps largely to decrease the<br />

concentration of the incorporated guest molecules such as acetic acid. Getting rid of the<br />

acetic acid as well as the counter-ions are indeed of great importance to elaborate Ru II,II<br />

analog of [M3(BTC)2]n in which the Ru-centers would be the same (in terms of oxidation<br />

state and coordination geometry) as Cu sites in Cu-HKUST-1. One approach to improve<br />

the situation could be the strict exclusion of any additional carboxylic acid as component<br />

of the solvothermal reaction medium. Acid catalysis seems to be needed to facilitate the<br />

coordination equilibria and kinetics of the substitution chemistry at the Ru-centres.<br />

However, every attempt in this direction has not been successful so far (Figure 7.12). On<br />

the other hand, the requirement of avoiding the substitution inert mixed valence Ru II,III -<br />

SBUs for CSA and moving to the Ru II,II -SBUs as well as the strict exclusion of redoxchemistry<br />

(inert conditions) together with any source of coordinating counter-ions Y is<br />

another way to sort out the complexity of the coordination environment (i.e. mixed<br />

valences of Ru centers, residual acids from the starting reactant-solvent and/or the<br />

starting precursors) of the Ru-analogs of HKUST-1.<br />

3.2.1 Synthesis and characterization of SBU-e and Ru-MOF 5<br />

Ru II,II -SBU ([Ru2(OOCCH3)4],SBU-e) was obtained from a blue solution of ruthenium(III)<br />

chloride according to the reported method. [206] The cell parameters obtained from the<br />

single crystal XRD measurements of [Ru2(OOCCH3)4](THF)2 (crystallized from the hot<br />

THF solution) are in a good agreement with the corresponding values communicated in<br />

the literature (Table 7.2). [207] As appeared, prepared complex is air sensitive and loses its<br />

crystallinity after solvent de-coordination. Nonetheless, the FT-IR bands of the symmetric


Intensity, a. u.<br />

Chapter 3 51<br />

and asymmetric COO - vibrations at 1425 and 1541 cm -1 , respectively, match well with the<br />

reported values (Figure 7.10). [82, 208]<br />

Cu-BTC_sim<br />

Cu-BTC_ht<br />

5<br />

Ru II, III -BTC<br />

8 12 16 20 24 28<br />

2, degree<br />

Figure 3.18. PXRD patterns of the activated Ru II,III -BTC 1 and sample 5 in comparison with<br />

simulated patterns of the as-synthesized (Cu-BTC_sim) and activated Cu-BTC (Cu-BTC_ht).<br />

Consequently, degassed solvents (H2O and acetic acid) and inert conditions (Ar<br />

atmosphere) were utilized to synthesize Ru-MOF 5. The PXRD patterns of the activated<br />

mixed-valence Ru II,III -BTC sample obtained in our earlier reports [82, 208] exhibit a little shift<br />

of the peaks in comparison with the simulated PXRD patterns of the Cu-BTC_sim, although<br />

the overall structural integrity is fully preserved. This slight shift stem mainly from the<br />

substitution of the copper-ions by ruthenium-ions (which are of different radii) at the PW<br />

SBUs of the MOF. The PXRD patterns of the prepared here activated sample 5 are in a good<br />

accordance with respective PXRD data of both reported earlier Ru II,III -BTC (Ru-MOF 1)<br />

and Cu-BTC_sim(Figure 3.18), providing a good indication for the phase-purity of the<br />

prepared solid. Furthermore, the IR spectra of the synthesized sample 5 matches well


52 Chapter 3<br />

with that of the Ru II,III -BTC (Figure 3.19). The bands corresponding to the νs(COO) and<br />

νas(COO) vibrations appear at 1359 cm -1 and 1429cm -1 , respectively, suggesting the<br />

coordination of the carboxylate groups of framework BTC. However, presence of<br />

unreacted acetic acid and H3BTC cannot be strictly excluded from IR, as the bands<br />

stemming from C=O stretching in Ru-MOF samples is too small to be distinguished. The<br />

1 H-NMR spectrum of acid-digest sample 5 displays peaks at 8.60 and 1.86 ppm (Figure<br />

7.13). The former resonance is originated from the aromatic protons of BTC. However, the<br />

latter one can be attributed to the protons of acetate, suggesting the presence of AcO(H)<br />

in the obtained sample.<br />

H 3<br />

BTC<br />

(C=O)<br />

Ru II,II<br />

2 (OOCCH 3 ) 4<br />

5<br />

Ru II, III -BTC<br />

4000 3500 3000 2500 2000 1500 1000 500<br />

wavenumber, cm -1<br />

Figure 3.19. IR spectra of the Ru-BTC 5 in comparison with Ru II,III -BTC 1, [Ru 2 (OOCCH 3 ) 4 ] and<br />

H 3 BTC. The vertical dash line corresponds to the position of ν(C=O) bands.


Chapter 3 53<br />

3.2.2 Investigation on the Ru oxidation state and porosity of Ru-MOF 5<br />

To get more information on the oxidation state of Ru-centers in the obtained sample 5,<br />

XANES spectra on sample 5, Ru II,III -BTC 1, used Ru-precursors (SBU-a and -e) and RuCl3<br />

were subsequently recorded. Due to the differences in position of edge jump different<br />

valences of Ru, the oxidation state of Ru-centers in the MOF structure can be concluded<br />

from the position of the highest peak in the plot of derivative normalized absorption vs.<br />

energy. As illustrated by the Figure 3.20, Ru II,III -BTC 1 shows an edge jump position<br />

between RuCl3 and its starting precursor SBU-a, indicating its mixed valence state, as<br />

documented earlier. [82] To remark, only Ru II was found in the SBU-e. [207] Thus, the same<br />

edge jump position observed in the XANES spectra of the sample 5 suggests the presence<br />

of Ru II -center in the framework structure (Figure 3.20). Consequently, the obtained solid<br />

5 should feature the exact Ru2-PWs as that in HKUST-1 without any counter-ions around.<br />

In the fact, EA results do not reveal the presence of any Cl in the sample Ru-MOF 5, ruling<br />

out the existence of Cl - and RuCl3 in both SBU-e employed during the synthesis and the<br />

obtained solid Ru-MOF 5. A formula of [Ru3(BTC)2]n∙(AcOH)2.3 match well with the<br />

obtained EA results, which could be another sign of the absence of counter-ions. In<br />

comparison with Ru II,III -BTC (1-4) described in Chapter 3.1, the obtained framework turns<br />

to be simpler in the absence any counter-ions, although the presence of acetic acid cannot<br />

be avoided. Furthermore, N2 sorption isotherms collected at 77 K for Ru-MOF 5<br />

demonstrate type I isotherm (Figure 3.21), indicating the microporosity of this material.<br />

Given the greater bulk density of Ru-BTC than Cu-BTC and in order to have a better<br />

comparison with HKUST-1, SBET is calculated in the unit of m 2 /mmol. Thus, SBET of Ru-MOF<br />

5 amounts to 1173 m 2 /mmol, which is quite comparable with the reported value of Cu-<br />

BTC (1049 m 2 /mmol), [83] indicating the same accessibility of Ru-centers as Cu-sites.<br />

Moreover, the axial positions of the diruthemium PW units in the synthesized Ru II,II -BTC<br />

(5) are expected to be open compared to the Ru II,III -BTC, where the presence of counterions<br />

(Cl - , OH - , AcO - ) is needed to balance the charge of [Ru2] 5+ units. The considerably<br />

higher BET surface area (SBET) of Ru II,II -BTC 5 (1371 m 2 /g) than that respective values<br />

reported for Ru II,III -BTC (704-1180 m 2 /g) [82-83, 138, 208] further supports an assumption on<br />

availability of Ru-CUSs in Ru II,II -BTC 5.


54 Chapter 3<br />

Figure 3.20. (a) XANES spectra of the Ru II,III -BTC 1 and Ru II,II -BTC (5) in comparison with<br />

correspondingly employed Ru-precursors (SBU-a and –e) and the commonly used reference<br />

(RuCl 3 ); (b) XANES spectra of Ru II,II -BTC (5) and the employed precursor Ru II,II -SBU<br />

([Ru 2 (OOCCH 3 ) 4 ],SBU-e). The vertical line stands for the same position of edge jump.


V, cm 3 /g<br />

Chapter 3 55<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

/ Ru II, II -BTC<br />

/ Ru II, III -BTC<br />

50<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

relative pressure, p/p 0<br />

Figure 3.21. N 2 adsorption (solid symbols) and desorption (open symbols) isotherms recorded at<br />

77K for the activated Ru II,III -BTC 1 and Ru II,II -BTC (5). Violet circle- Ru II,II -BTC; orange triangle-<br />

Ru II,III -BTC.<br />

3.2.3 Conclusions<br />

Moving from the usage of the mixed-valence Ru II,III -SBUs to the mono-valence Ru II,II -SBU<br />

in CSA, a Ru-analog ([Ru3(BTC)2]n∙(AcOH)2.3) of [Cu3(BTC)2]n has been obtained and<br />

characterized. Interestingly, XANES spectra suggest Ru II,II sites being present in the<br />

structure of the obtained sample 5. Moreover, the BET surface area of this material is the<br />

highest (1173 m 2 /mmol) among of all reported so far Ru-BTC MOFs and quite comparable<br />

with that of HKUST-1 (1049 m 2 /mmol). Thus, Ru II,II -BTC hold a huge perspectives for<br />

various applications in sorption and catalysis. Even more, obtained results are important<br />

for the following studies on related but more complex defect-engineered Ru-MOFs (see<br />

Chapter 4), and elaborated here Ru II,II -SBU is a good candidate for the synthesis of Rubased<br />

MOFs.


4 Linker-based MOF solid solutions: Defect-Engineered MOFs<br />

(DEMOFs) †<br />

Abstract: Employing the mixed-linker solid solution approach, various functionalized ditopic<br />

isophthalate (ip) defect generating linkers denoted as 5-X-ipH2, with X = OH (1), H<br />

(2), NH2 (3), Br (4) have been introduced into [Cu3(BTC)2] (HKUST-1) as well as its mixedvalence<br />

ruthenium-analogue to yield DEMOFs (DE = “defect-engineered”) of the general<br />

empirical formula [M3(BTC)2-x(5-X-ip)xYy]n (M = Ru or Cu, Y = counter-ions, 0 ≤ y ≤ 1.5).<br />

The framework incorporation of 5-X-ip into DEMOFs has been confirmed by a number of<br />

† Most work of this chapter has been covered in the accepted manuscript: W. Zhang, M. Kauer, et al. R. A.<br />

Fischer, Chem. Eur. J. 2016.


Chapter 4 57<br />

techniques. Interestingly, Ru-DEMOF (1c) with 32% framework incorporated 5-OH-ip<br />

reveals the highest BET surface area (1300 m 2 /g, N2 adsorption, 77 K) among all related<br />

samples (including the parent framework [Ru3(BTC)2Yy]n). The characterization data are<br />

consistent with two kinds of structural defects induced by 5-X-ip framework<br />

incorporation: type A, modified paddlewheel nodes featuring reduced metal sites and<br />

type B, missing nodes leading to enhanced porosity. Their relative abundances depend on<br />

the choice of the functional group X in the DLs and its doping level. The defects A and B in<br />

Ru-DEMOFs appear also to play a key role in sorption of small molecules (i.e., CO2, CO, H2)<br />

as well as for the catalytic properties of the samples (i.e., ethylene dimerization and Paal-<br />

Knorr reaction). In addition, synthetic parameters such as anions from the metal salts and<br />

the solvent can also influence the formation of Cu-DEMOFs as well as the generation of<br />

defects.


58 Chapter 4<br />

4.1 Introduction<br />

4.1.1 Solid solutions<br />

Among the well-known mixed-component MOFs, the class of MOF solid solutions (also<br />

known as molecular substitutional alloys) attract huge attention. It implies MOFs where to<br />

two or more types of the related components (either linkers, metals, or both) are<br />

randomly mixed within single-phased framework. Remarkably, this often allows the fine<br />

modulation of the MOF’s functionalities and properties. Considering the nature of the<br />

mixed components, MOF solid solutions could be sub-divided mainly into two groups:<br />

linker-based and metal-based MOF solid solutions. The first group covers MOF solids<br />

where two or more homologous linkers (usually of similar size and functionality) are<br />

integrated in various proportions while the structural integrity of MOF is preserved. The<br />

metal-based MOF solid solutions will be discussed in details in Chapter 5 and involve<br />

mixed-metal frameworks where ions of one metal type is partly isomorphous substituted<br />

by the other metal ions.<br />

Figure 4.1. Mixed-linker approaches to the formation of MOF solid solutions. Magenta balls, metal<br />

nodes; cyan sticks, parent linker; green sticks, isostructural mixed-linker; wine short sticks,<br />

truncated mixed-linker; violet sticks, heterostructural mixed-linker. Inspired by the reports of<br />

Bunck [127] and Fang et al. [137]<br />

Approaches for the synthesis of linker-based MOF solid solutions go basically into three<br />

directions depending on the utilized linker mixtures (Figure 4.1): [127]<br />

i) isostructural mixed-linker (IML) approach where mixed linkers feature identical linking<br />

geometry. [123-124, 209-212]


Chapter 4 59<br />

This is the most straightforward way and commonly used to incorporate some specific<br />

functionality, decorate the walls of a framework with particular (active) organic groups .<br />

Thus, Yaghi’s multivariate MOFs (MTV-MOF-5) are typical representatives of such solid<br />

solutions. [123] In fact, by mixing the terephthalic acid and its derivatives in various<br />

combinations, eighteen single phased MTV-MOFs-5 were prepared and, remarkably,<br />

demonstrated the enhancement of H2 storage capacity as well as adsorption selectivity for<br />

CO2 over CO. Furthermore, Baiker et al. reported mixed-linker amine-functionalized MOFs<br />

(MIXMOFs) with the general formula Zn4O(bdc)x(abdc)3–x (bdc = benzene-1,4-<br />

dicarboxylate; abdc = 2-aminobenzene-1,4-dicarboxylate). Notably, obtained solid<br />

solution materials could be used as both nucleophilic catalysts [124] and support to<br />

immobilize Pd-species for Pd-catalyzed cross-coupling reactions. [212]<br />

ii) truncated mixed-linker (TML) approach, in which one linker is truncated in comparison<br />

with the other.<br />

Figure 4.2. Schematic illustration of the formation of the MOF-5 analogs: spng-MOF-5 and pmg-<br />

MOF-5. Reprinted with permission from K. M. Choi, H. J. Jeon, J. K. Kang and O. M. Yaghi, J. Am.<br />

Chem. Soc., 2011, 133, 11920-11923. Copyright (2011) American Chemical Society. [215]<br />

This strategy is strongly related to the modulation approach, whereby a monocarboxylic<br />

acid is commonly added during the MOF synthesis to influence crystal growth and<br />

morphology. For instance, Kitagawa et al. used acetic acid as a modulator to directly affect<br />

the coordination equilibrium, which in turn controls the crystal growth of nanosized


60 Chapter 4<br />

framework [Cu2(ndc)2(dabco)]n (ndc=1,4-naphthalene dicarboxylate; dabco=1,4-<br />

diazabicyclo[2.2.2]octane). [213] Subsequently, similar morphological control of<br />

[Cu3(BTC)2]n (BTC = 1,3,5-benzenetricarboxylate) from octahedral to cuboctahedral to<br />

cubic by addition of lauric acid has been achieved. [214] Moreover, by utilizing various<br />

amounts of dodecyloxybenzoic acid, Yaghi and co-workers have modified the crystals of<br />

MOF-5 to develop isostructural sponge-/pomegranate- like crystal analogs featuring<br />

meso- and macropores, respectively(Figure 4.2). [215]<br />

iii) heterostructural mixed-linker (HML) strategy, where the employed mixed linkers bear<br />

different linkage geometries. [142, 216-219]<br />

UMCM-1 (University of Michigan Crystalline Material-1) reported by Matzger et al. is a<br />

first rigorous example derived following this approach by the usage of dicarboxylate and<br />

tricarboxylate linkers. This MOF contains 3.1 nm mesoporous hexagonal channels<br />

surrounded by 1.4 nm microporous cages and exhibits exceptionally high surface area. [217]<br />

Thus, variation of the relative lengths of the linkers mixed (di- and trifunctional linkers)<br />

may result frameworks of different topologies. [218] In this manner, HML strategy has<br />

paved up a way to unavailable topologies and larger pore sizes which are not easily<br />

achieved in single-component frameworks. However, careful control over MOF synthesis<br />

to avoid the formation of physical mixtures of different phases needs to be considered.<br />

The latter two synthetic strategies, especially TML approach, can be successfully<br />

employed to prepare defect-engineered MOFs (DEMOFs), which will be described in this<br />

Chapter below.<br />

4.1.2 Defects in MOFs and its “engineering”<br />

It has been recognized that MOFs, similar to other (crystalline) solid-state materials, are<br />

typically not perfect crystals with infinite periodic repetition or ordering of the identical<br />

groups of atoms in space. In fact, defects of various types and length scales (even more<br />

generally: structural heterogeneity) cannot be rigorously avoided during synthesis and<br />

crystal growth and often play an important role for the material properties. [220-222] The<br />

diversity of MOF structures connected with the issues of intrinsic and intentional<br />

structural defects, non-stoichiometry and heterogeneity suggests a much broader<br />

approach for understanding and tailoring their chemical and physical properties. [137] In<br />

fact, Ravon and co-workers has reported MOFs featuring structural defects in which acidic


Chapter 4 61<br />

active catalytic centers were generated through fast precipitation and partial substitution<br />

of dicarboxylic by mono-carboxylic acid linkers. Interestingly, the obtained in this manner<br />

defect-engineered Zn-MOFs show shape selectivity in catalyzing alkylation of large<br />

aromatics molecules. [144] Furthermore, by using CF3COOH and HCl during the synthesis,<br />

Vermoortele et al. demonstrated that UiO-66(Zr) can be modified via partial substitution<br />

of terephthalates by trifluoroacetate. Interestingly, subsequent thermal activation lead to<br />

post-synthetic removal of the trifluoroacetate groups in addition to the dehydroxylation<br />

of the hexanuclear Zr-clusters. All together, this yielded more open defect framework with<br />

a large number of open sites. Remarkably, produced defects at metal sites enhanced UiO-<br />

66(Zr) activity in several Lewis acid catalyzed reactions(Figure 4.3). [141] As a result, better<br />

catalytic performance in several Lewis-acid catalyzed reactions was observed for the<br />

DEMOFs. Controlled introduction and characterization of specific defects into MOFs is<br />

quite a challenge including the comparison with the parent (more or less) “defect-free”<br />

reference samples.<br />

Figure 4.3. The generation of defects in UiO-66(Zr). Reprinted with permission from F.<br />

Vermoortele, B. Bueken, G. Le Bars, B. Van de Voorde, M. Vandichel, K. Houthoofd, A. Vimont, M.<br />

Daturi, M. Waroquier, V. Van Speybroeck, C. Kirschhock and D. E. De Vos, J. Am. Chem. Soc., 2013,<br />

135, 11465-11468. Copyright (2013) American Chemical Society. [141]<br />

[Cu3(BTC)2]n (also known as HKUST-1) [35] as well as the isostructural family [M3(BTC)2]n<br />

(M = Mo, [81, 83] Cr, [78, 83] Ni, [79, 83] Ru, [82-83, 198] Zn, [77, 83] , BTC = benzene-1,3,5-tricarboxylate)<br />

attracted considerable attention over last years, mainly due to presence of coordinatively<br />

unsaturated metal-sites (M-CUS), as has been already introduced in Chapter 3. Intentional<br />

defect generation in [Cu3(BTC)2]n was initially reported by Baiker et al. who studied


62 Chapter 4<br />

partial replacing of BTC by the homologous pyridine-3,5-dicarboxylate (pydc) linker and<br />

revealed enhanced catalytic activity of the modified MOF in the oxidation of benzene<br />

derivatives related to modified M-CUS at the nodes. [136] Subsequently, studies conducted<br />

at our group demonstrated that not only pydc, but also a series of other defect generating<br />

linkers can be framework incorporated in case of [Cu3(BTC)2]n. [139] Interestingly,<br />

functionalized mesopores along with the M-CUS appeared to be generated in such defectengineered<br />

MOFs (DEMOFs) (Figure 4.4).<br />

Figure 4.4. The modulation of the defect structure on the micro- and mesoscales by doping of the<br />

framework with DL. The blue and short red sticks represent perfect and defective linkers. The<br />

yellow and black balls represent perfect and defective metal sites, respectively. The green<br />

highlighted unit indicates parent micropores. Reprinted with permission from Z. Fang, J. P.<br />

Dürholt, M. Kauer, W. Zhang, C. Lochenie, B. Jee, B. Albada, N. Metzler-Nolte, A. Pöppl, B. Weber, M.<br />

Muhler, Y. Wang, R. Schmid and R. A. Fischer, J. Am. Chem. Soc., 2014, 136, 9627-9636. Copyright<br />

(2014) American Chemical Society. [139]<br />

In parallel, Hupp et al. gave an account of the enhanced porosity by introducing<br />

isophthalate (ip) into [Cu3(BTC)2]n. [140] The concept of defect engineering by a mixed<br />

component approach using fragmented linkers together with the parent one was also<br />

expanded to the isostructural Ru-analogue. [138] Ru-DEMOFs [Ru3(BTC)2-x(pydc)xYy]n (Y =<br />

counter-ion, Cl - , OH - , OAc, etc.; 0 < y≤ 1.5) were obtained through direct mixing of H2pydc<br />

and H3BTC in one-step solvothermal synthesis. The generation of reduced Ru δ + (0


Chapter 4 63<br />

compared with the parent material, the actual compositions of these Ru-DEMOFs are very<br />

difficult to establish and are not entirely clear. This stems from the structural complexity<br />

of such multi‐component solid solution MOFs and a range of possible compositional<br />

variations, which are not mutually exclusive (counter-ions, residual synthesis<br />

components such as linkers, coordination modulators etc.).<br />

Figure 4.5 Design concept of the DEMOFs applying the mixed component/fragmented linker solid<br />

solution approach and sketch of the two possible, most important paddlewheel-node related<br />

defects.<br />

Figure 4.5 summarizes the concept of defects introduction into mixed-valence<br />

[M3(BTC)2Yy]n by using divalent isophthalate linkers as fragmented variants of trivalent<br />

BTC. Mainly, two types of defects need to be distinguished: Type A refers to structurally<br />

and electronically modified paddlewheel units (missing one carboxylate ligator function),<br />

in which (at least) one out of the four bridging carboxyl groups from the parent BTC linker<br />

in the regular paddlewheels has been partly substituted by the (neutral) functional group<br />

of the DL. Consequently, for charge compensation, the metal sites could be either partly<br />

reduced (mixed-valence state) or may carry additional anionic (mono-valent) ligands<br />

(Y) [137-138] However, apart from such type A, defects of type B could be formed, where<br />

whole paddlewheel nodes are eliminated from the structure, as it has been demonstrated


64 Chapter 4<br />

in the case of ip introduction into [Cu3(BTC)2]n. [140] It is conceivable, that such missing<br />

node defects could be formed in a correlated fashion during the MOF crystal growth<br />

process and in such a case larger areas of missing nodes may be generated to yield<br />

mesopores.


Chapter 4 65<br />

4.2 Ruthenium Metal-Organic Frameworks featuring Different Defect Types<br />

Being mainly studied in Cu-HKUST-1, a comprehensive understanding of the nature and<br />

formation of different types of isolated local and larger scale correlated defect sites in<br />

DEMOFs of the [M3(BTC)2]n family and MOFs in general is very limited and this research<br />

is still at its infancy. [137] The work in this chapter is a follow-up of the previous<br />

investigations on defect engineering in [Ru3(BTC)2Yy]n using H2pydc as (single) defect<br />

generating linker. [138] Similar to the previous more elaborate study on defect-engineered<br />

Cu-HKUST-1 [139] , a series of 5-X-isophthalic acids (5-X-ipH2, X = OH, H, NH2, Br) are<br />

selected and mixed them with H3BTC to yield the Ru-DEMOFs of the general formula<br />

[Ru3(BTC)2-x(5-X-ip)xYy]n (Figure 4.5). The choice of the functional group X in DL is on the<br />

basis of two aspects: i) modulating the coordination binding strength between linkers and<br />

Ru-nodes, namely, changing one of the carboxylate group in parent H3BTC linker to<br />

weaker binding groups (OH, NH2) or even inert coordination groups (H, Br); ii) labeling<br />

DL with easily tracked elements such as Br, NH2, etc. for analytical purpose. Thus, a<br />

number of complementary characterization methods are applied to obtain compositional<br />

and microstructure data of the materials. The aim is to at least qualitatively assess the<br />

tendency of type A and type B defect formation as a function of synthetic conditions and<br />

the chosen 5-X-ip linker. Several catalytic reactions, such as ethylene dimerization and<br />

Paal-Knorr reaction (usually running under Lewis acidic conditions) as well as low<br />

temperature CO2→CO dissociative chemisorption in the dark under UHV conditions, have<br />

been chosen as test reactions to monitor the influence of defects on the performance of<br />

the Ru-DEMOFs.<br />

4.2.1 Synthesis and characterization of the Ru-DEMOF materials<br />

In the following, a comprehensive account on the conducted experiments, obtained<br />

analytical and characterization data is provided for elucidation of the defect structure of<br />

Ru-DEMOFs of the general formula [Ru3(BTC)2-x(5-X-ip)xYy]n. The main goal of the<br />

presented study is to substantiate the hypothesis of type A and type B defect formation in<br />

these systems as introduced above (Figure 4.5).


66 Chapter 4<br />

4.2.1.1 Crystallinity and stability of the obtained samples<br />

Figure 4.6. PXRD patterns of the solvated Ru-DEMOF materials 1a-1d, 2a-2d, 3a-3d, 4a-4c<br />

compared with the solvated parent Ru-MOF, respectively. 1a-1d: DL is 5-OH-ip; from 1a to 1d,<br />

feeding level is increasing: 10%, 20%, 30%, 50%, respectively. 2a-2d: DL is ip; from 2a-2d,<br />

feeding level is increasing: 10%, 20%, 30%, 50%, respectively. 3a-3d: DL is 5-NH 2 -ip, from 3a to<br />

3d, feeding level is increasing: 10%, 20%, 30%, 50%, respectively. 4a-4c: DL is 5-Br-ip, from 4a<br />

to 4c, the feeding level is increasing: 10%, 20%, 30%, respectively. Vertical lines at 42.1°and 44°<br />

stand for the respective reflections of (001) and (101) faces of hexagonal close‐packed metallic<br />

Ru-particles or Ru-NPs.<br />

Parent Ru-MOF in this chapter for comparison was prepared according to the synthetic<br />

method for Ru-MOF 1_ex in Chapter 3. All Ru-DEMOF samples were obtained according<br />

to the previously published procedure, [138] using [Ru2(OOCCH3)4Cl]n, H3BTC and defect<br />

generating 5-X-ipH2 as starting materials in the acetic acid/water solution under<br />

solvothermal conditions. The sample-numbering scheme is based on the framework


Chapter 4 67<br />

incorporated 5-X-ip linker: X= OH (1), H (2), NH2 (3) or Br (4). Further labeling (a, b, c<br />

etc.) refers to various levels (molar%) of DL incorporation. The PXRD patterns of the assynthesized<br />

and activated samples indicate that all materials (except 2d) are crystalline<br />

and isostructural with the parent Ru-MOF that is obtained employing the same method<br />

with H3BTC, only (Figures 4.6 and 4.7). It should be noted that Ru 2+/3+ ions can be possibly<br />

reduced to eventually yield metallic ruthenium. Therefore, the solvothermal synthesis<br />

Figure 4.7. PXRD patterns of the activated Ru-DEMOF materials 1a-1d, 2a-2d, 3a-3d, 4a-4c<br />

compared with the solvated parent Ru-MOF, respectively. 1a-1d: DL is 5-OH-ip; from 1a to 1d,<br />

feeding level is increasing: 10%, 20%, 30%, 50%, respectively. 2a-2d: DL is ip; from 2a-2d,<br />

feeding level is increasing: 10%, 20%, 30%, 50%, respectively. 3a-3d: DL is 5-NH 2 -ip, from 3a to<br />

3d, feeding level is increasing: 10%, 20%, 30%, 50%, respectively. 4a-4c: DL is 5-Br-ip, from 4a<br />

to 4c, the feeding level is increasing: 10%, 20%, 30%, respectively. Vertical lines at 42.1°and 44°<br />

stand for the respective reflections of (001) and (101) faces of hexagonal close‐packed metallic<br />

Ru-particles or Ru-NPs.<br />

conditions have to be carefully monitored. However, no PXRD-peaks in the 2θ range of 40-<br />

45° were found, neither for the as-synthesized nor for the activated samples, thus


68 Chapter 4<br />

suggesting the absence of a substantial amount of Ru 0 nano-particles (Ru-NPs).<br />

Reflections at 42.1°and 44°are usually assigned to (001) and (101) faces of hexagonal<br />

close‐packed metallic Ru-particles or Ru-NPs. [223] Further data to rule out Ru 0 , XANES in<br />

particular, will be described below. The corresponding results provide additional<br />

evidence for the absence of Ru/RuOx‐NPs in the obtained Ru-DEMOFs. With the exception<br />

of 2d, thermal gravimetric analysis (TGA) shows no decomposition of the prepared<br />

materials for temperatures below 250 °C, suggesting that thermal stability of Ru-DEMOFs<br />

is preserved as compared to the parent MOF (Figure 4.8).<br />

Figure 4.8. TG curves of the prepared Ru-DEMOFs 1a-1d (with 5-OH-ip DL), 2a-2c (with ip DL),<br />

3a-3d (with 5-NH 2 -ip DL), 4a-4c (with 5-Br-ip DL) in comparison with the parent Ru-MOF,<br />

respectively.


Chapter 4 69<br />

4.2.1.2 The incorporation amount of DLs as well as the composition of the<br />

materials<br />

The incorporation and quantification of DLs have been initially testified by NMR<br />

spectroscopy of the digested samples after activation and the assessment of porosity<br />

(Table 4.1). Thus, in the 1 H-NMR spectra of the samples 1a-d, appearance of two<br />

additional peaks (compared with the parent Ru-MOF) at 7.51 ppm and 7.89 ppm is clearly<br />

seen (Figure 4.9). These two resonances can be assigned to the aromatic protons of 5-OHip<br />

linkers. Expectedly, upon increasing the feeding amount of the DL (5-OH-ipH2), the<br />

intensities of its characteristic resonances also increases in the digested sample solution.<br />

Figure 4.9. 1 H-NMR spectra of the Ru-DEMOFs (1a-1d) digested in DCl/DMSO-d 6 mixture. DL is<br />

5-OH-ip. H 3 O + comes from H 2 O and HCl contained in DCl. Due to the sensitivity of H 2 O peak to the<br />

solution conditions, the proton resonance of H 3 O + varies here in the range of 4-6 ppm, dependent<br />

on pH value of the total solution.


70 Chapter 4<br />

Table 4.1. The molar fractions (%) of the DLs used for the synthesis and observed in the final Ru-<br />

DEMOFs in respect to the total amount of linkers (H 3 BTC + 5-X-ipH 2 = 100%). The values were<br />

obtained from the integrals of proton peaks in 1 H-NMR spectra of the acid-digest sample. For the<br />

samples 1a-1d DL is 5-OH-ip; for 2a-2c – ip; for 4a-4c – 5-Br-ip, respectively.<br />

Sample<br />

Feeding<br />

(used for synthesis)<br />

Obtained<br />

1a 10% 8%<br />

1b 20% 20%<br />

1c 30% 32%<br />

1d 50% 37%<br />

2a 10% 15%<br />

2b 20% 28%<br />

2c 30% 47%<br />

4a 10% 17%<br />

4b 20% 25%<br />

4c 30% 42%<br />

A similar scenario was observed also for the samples with ip DL (series 2a-c). Resonances<br />

corresponding to the protons of ip are seen at 8.4 ppm, 8.1 ppm and 7.6 ppm (Figure 4.10).<br />

The framework incorporation amount of ip in 2a-c increases upon increasing the<br />

concentration of the DL in the reactant solutions. However, the incorporated amount of ip<br />

was found to be higher than expected from the H3BTC:H2ip feeding ratio, which is in<br />

contrast with that in the case of Ru-DEMOF series 1 where the incorporation of 5-OH-ip<br />

stay the same (or a little less) than the feeding concentration in the starting reactants. This<br />

result might be attributed to the slightly different coordination equilibria that correlate<br />

with the Brønsted acid dissociation constants of the used linkers: pKa (H3BTC) = 3.12,<br />

3.89, 4.70 and pKa (H2ip) = 3.46, 4.46. [224] However, the pKa differences are small and a<br />

conclusive explanation based on pKa is not straightforward. Dincă et al. argued that<br />

deprotonated BTC could serve as a counter-ion to provide the charge balance of [Ru2] 5+<br />

units in Ru-analogue of Cu-HKUST-1. [83] From our earlier investigation we concluded that<br />

H3BTC is present in the pores of [Ru3(BTC)2Yy]n in the form of fully protonated species,<br />

rather than as deprotonated counter-ions. [208] Both studies indicate that the linker used<br />

for the synthesis could still be present in the obtained materials. Hence, it is conceivable<br />

that the defect generating H2ip linker can also reside either as a counter-ion or as a guest


Chapter 4 71<br />

molecule in the pores of the Ru-DEMOFs (2a and 2b). On the basis of these arguments it<br />

can be understood why the amount of incorporated ip can be higher than expected from<br />

the H2ip:H3BTC feeding ratio. Indeed, very weak IR-bands in the range of ν(C=O), ν(C=C)<br />

or νas(COO) vibrations (1736-1521 cm -1 ) have been observed in case of samples 2a-2d.<br />

However, the small separation of the ν(C=O) bands of non-coordinated acetic acid (1715-<br />

1640 cm -1 ), [225] protonated/non-coordinated H2ip (1670 cm -1 ) and H3BTC (1680 cm -1 )<br />

(see Figure 4.11.2) makes it hard to unambiguously determine the exact origin of these<br />

weak vibrations.<br />

Figure 4.10. 1 H-NMR spectra of the Ru-DEMOFs (2a-2c) digested in DCl/DMSO-d 6 mixture. DL is<br />

ip. H 3 O + comes from H 2 O and HCl contained in DCl. Due to the sensitivity of H 2 O peak to the<br />

solution conditions, the proton resonance of H 3 O + varies here along with the variation of pH value<br />

of the total solution.


72 Chapter 4<br />

Figure 4.11. IR spectra of the prepared Ru-DEMOFs 1a-1d (with 5-OH-ip DL), 2a-2d (with ip DL),<br />

3a-3d (with 5-NH 2 -ip DL), 4a-4c (with 5-Br-ip DL) in comparison with the parent Ru-MOF and<br />

the corresponding DL used in the synthesis, respectively. The vertical dash lines in Figure 4.11.3<br />

indicate position of the vibrations of N-H and C-N bonds.<br />

Here it should be pointed out that in all 1 H-NMR spectra of the digested Ru-DEMOF<br />

samples the resonance at ca. 1.91 ppm, originated from the –CH3 groups of acetate, has<br />

been detected. This is in line with our previous investigations on the parent single-linker<br />

Ru-MOFs obtained from various Ru-precursors and the Ru-DEMOFs with pydc DL, where<br />

the same resonance due to acetate in the digested samples with the molar ratio of<br />

n(AcO):n(BTC) in the range from 0.7 to1.8 was always observed as well. [138, 208] . We<br />

anticipate three main origins of the presence of acetate in the discussed Ru-MOFs:<br />

i) it may act as a counter-ion to compensate the overall charge of the framework (Figure<br />

4.12a);


Chapter 4 73<br />

ii) owing to the competition with BTC and 5-X-ip, it could be residual/non-substituted and<br />

paddle wheel coordinated acetate-groups of the used Ru-precursor ([Ru2(OOCCH3)2Cl]n)<br />

(Figure 4.12b);<br />

Figure 4.12. Schematic representation for the two possible origins of acetate in the Ru-DEMOFs.<br />

iii) finally, it might reside as AcOH in the pores of the framework through weak<br />

interactions and, therefore, cannot be easily removed by washing and subsequent thermal<br />

activation (i.e., heating at 150 °C under dynamic vacuum).<br />

To recall, the glacial acetic acid/water mixture (AcOH:H2O = 0.05:1) was found to be<br />

essential for a successful synthesis and was always used for all reported Ru-(DE)MOFs<br />

(see Experimental Section). Besides, employed linkers may also serve as guest molecules<br />

as has been mentioned above. Attempts to evaluate the presence of acetic acid and linkers<br />

from the IR spectra is not quite feasible due to the overlapping of the bands of νas(COO)<br />

and νs(COO) (coordinated in a bidentate-bridging mode and free-coordinated carboxylate<br />

group) [226] as well as low intensity of the respective bands. Although one can monitor<br />

acetate groups in IR-spectroscopy by employing F-labled acetate (eg. perfluoro-acetate<br />

instead of acetate) for synthesis, [227] the crucial synthesis parameters of Ru-MOFs make<br />

this attempt impossible. Earlier effort of using various acids as the reaction solvent failed<br />

to form crystalline parent Ru-MOF. [208] Consequently, the presence of acetate (mainly)<br />

and non-reacted free linkers used in Ru-DEMOFs raises questions about the actual sample<br />

compositions and local structures and, thus, causes more complications with ambiguous<br />

interpretation of the analytical data than in case of the homologous Cu-DEMOFs. [139] The<br />

compositions of the Ru-DEMOFs samples discussed in this work have been derived by


74 Chapter 4<br />

taking into account the complete set of experimental results and a full account is given in<br />

Tables 4.2 and 4.3).<br />

Table 4.2. Elemental analysis results (calculated and found) of the activated Ru-DEMOFs (1a, 1c<br />

and 1d) in comparison with the parent Ru-MOF. Molar ratio of Ru : Cl calculated based on the<br />

found mass percentages are also given below.<br />

Sample<br />

C, wt % H, wt % Ru, wt % Cl, wt %<br />

cal. found cal. found cal. found cal. found<br />

Ru : Cl<br />

parent 28.5 27.26 2.31 1.65 31.40 30.45 3.67 3.41 3: 1<br />

1a 29.3 28.4 2.28 1.49 31.55 30.08 3.69 3.47 3: 1<br />

1c 31.67 31.26 1.89 1.37 32.03 31.08 3.37 3.27 3: 0.9<br />

1d 30.26 29.43 2.06 1.59 31.36 30.41 3.30 3.35 3: 0.9<br />

Table 4.3. Proposed formula* and the molecule weight of the activated Ru-DEMOFs (1a, 1c and<br />

1d) as well as the parent Ru-MOF.<br />

Sample<br />

Parent Ru-<br />

MOF 1_ex<br />

1a<br />

1c<br />

1d<br />

Formula<br />

[Ru 3 (C 9 H 3 O 6 ) 2 Cl(CH 3 COO) 0.5 ]·(CH 3 COOH) 1.5 (C 9 H 6 O 6 ) 0.1 (H 2 O) 4<br />

([Ru 3 (BTC) 2 Cl(AcO) 0.5 ]·(AcOH) 1.5 (H 3 BTC) 0.1 (H 2 O) 4 )<br />

[Ru 3 (C 9 H 3 O 6 ) 1.84 (C 8 H 3 O 5 ) 0.16 Cl(OH) 0.5 ]·(CH 3 COOH) 2.8 (H 2 O) 2<br />

([Ru 3 (BTC) 1.84 (5-OH-ip) 0.16 Cl(OH) 0.5 ]·(AcOH) 2.8 (H 2 O) 2 )<br />

[Ru 3 (C 9 H 3 O 6 ) 1.36 (C 8 H 3 O 5 ) 0.64 Cl 0.9 (OH) 0.6 ]·(CH 3 COOH) 2.2 (C 8 H 6 O 5 ) 0.4<br />

([Ru 3 (BTC) 1.36 (5-OH-ip) 0.64 Cl 0.9 (OH) 0.6 ]·(AcOH) 2.2 (5-OH-ipH 2 ) 0.4 )<br />

[Ru 3 (C 9 H 3 O 6 ) 1.26 (C 8 H 3 O 5 ) 0.74 Cl 0.9 (OH) 0.6 ]·(CH 3 COOH) 1.3 (C 9 H 6 O 6 ) 0.5 (H 2 O) 2.5<br />

([Ru 3 (BTC) 1.26 (5-OH-ip) 0.74 Cl 0.9 (OH) 0.6 ]·(AcOH) 1.3 (H 3 BTC) 0.5 (H 2 O) 2.5 )<br />

Molecular<br />

weight<br />

965.6<br />

961.1<br />

946.6<br />

982.6<br />

* The proposed formula of all the samples above have been calculated according to the facts and<br />

considerations listed below:<br />

<br />

<br />

Parent (BTC) to DL (5-OH-ip) ratios have been calculated based on the 1 H-NMR spectra (i.e.,<br />

integration of the linker-related signals) of the respective acid-digested samples.<br />

The parent Ru-MOF and reported Ru-DEMOFs are isostructural. At this point it is hard to<br />

quantitatively estimate contribution from the defects B (missing metal-nodes), therefore, for initial<br />

evaluations assumption that the general composition of the prepared solids is [Ru3(L)2Yy]n (L = total<br />

linker amount) was employed. Thus, the Ru : L molar ratio can be taken as a fixed value – 3 : 2.<br />

Further, having in mind there is generally 3 molar equivalents of Ru per 1 molecular unit of MOF,


Chapter 4 75<br />

the Cl-amount was subsequently calculated based on the measured EA values and respective Ru :<br />

Cl ratios.<br />

<br />

<br />

As observed by the 1 H-NMR, TG and IR analyses, the contribution of the counter-ions and guest<br />

molecules are mainly from Cl, OH, acetate, H2O and used linkers. In analogy to the previously<br />

reported parent Ru-MOF [208] and the Ru-DEMOF with pydc DL, [138] in the currently discussed Ru-<br />

DEMOFs acetic acid is also present, which is clearly evidenced by the 1 H-NMR spectra of the<br />

digested samples. To note, the amount of the acetic acid can be considerably reduced by additional<br />

solvent exchange procedure, namely, soaking of the activated samples in a large amount of water<br />

while intense stirring to exchange the residue acetic acid. However, the risk of leaching out of Ru<br />

needs also to be taken into account here. [208] Hence, our current studies dealt with the obtained<br />

powders applying gently water soak as solvent exchange procedure to decrease the residue of the<br />

acetic acid. And the overall concentration of the acetic acid decreased in comparison with our<br />

previously report on the Ru-DEMOF with pydc DL in spite of the slightly decrease of Ru content.<br />

In addition of the multiple possible status of acetic acid, starting reactant (i.e. linkers) can be also<br />

accommodated into the framework as guest molecules or served as counter-ions, since based on<br />

the NMR data the amount of the DL in some cases appeared to be higher than its respective feeding<br />

amount. Besides, minor quantities of the coordinated H2O might be occluded / resided in the pores<br />

as well, as the TG curves of the activated samples show the slight decrease in the range of 100-250<br />

°C.<br />

Figure 4.13. EDX spectra of the Ru-DEMOFs 4a-4c (with 5-Br-ip DL). The detection of Br<br />

clearly indicates the presence of the employed DL.


76 Chapter 4<br />

As quantitative digest of the samples 3a-d is not sufficient, detection and determination<br />

of the DL contents (5-NH2-ip) from 1 H-NMR measurements was not possible.<br />

Nevertheless, the IR spectra of these samples provide a qualitative evidence for the 5-NH2-<br />

ip incorporation. Indeed, the observed bands at 1651 cm -1 and 1264 cm -1 stem from δ(N-<br />

H) and ν(C-N) vibrations of the 5-NH2-ip (Figure 4.11). In case of the 5-Br-ip (samples 4ac),<br />

the gradual increase of intensities of the respective signals in both 1 H-NMR and TEM-<br />

EDX spectra suggests the presence of DL in these Ru-DEMOFs (see Figures 4.13 and 4.14).<br />

Figure 4.14. 1 H-NMR spectra of the Ru-DEMOFs (4a-4c) digested in DCl/DMSO-d 6 mixture.<br />

DL is 5-Br-ip. H 3 O + comes from H 2 O and HCl contained in DCl. Due to the sensitivity of H 2 O<br />

peak to the solution conditions, the proton resonance of H 3 O + varies here, dependent on pH<br />

value of the total solution.<br />

4.2.2 Porosity of Ru-DEMOF samples and the confirmation of the DL incorporation<br />

To support further the framework incorporation of the DL and rule out its extra<br />

framework inclusion within the pores in a high quantity, we have carried out N2 sorption<br />

measurements for the activated samples. The N2 sorption isotherms at 77 K for all samples


Chapter 4 77<br />

Figure 4.15. N 2 sorption isotherms collected at 77 K for Ru-DEMOF samples 1a-1d, 2a-2c, 3a-3d,<br />

4a-4c. 1a-1d: DL is 5-OH-ip, incorporation amount is 8%, 20%, 32% and 37%, respectively; 2a-<br />

2c: DL is ip, incorporation amount is 15%,28% and 47%, respectively; 3a-3d: DL is 5-NH 2 -ip; 4a-<br />

4c: DL is 5-Br-ip, incorporation amount is 17%, 25% and 42%, respectively. Closed and open<br />

symbols represent the adsorption and desorption isotherms, respectively. Black circles – parent<br />

Ru-MOF; blue triangles – a samples of the respective series 1-4; dark cyan diamonds – b samples<br />

of the respective series 1-4; magenta squares – c samples of the respective series 1-4; dark yellow<br />

stars – d samples of the respective series 1-4.<br />

(1a-d, 2a-d, 3a-d and 4a-c) reveal type I isotherm without any hysteresis loop (Figure<br />

4.15), indicating that all Ru-DEMOFs are microporous materials. [228] In general, while<br />

comparing with the Brunauer-Emmett-Teller (BET) surface area (SBET) of the parent Ru-<br />

MOF (998 m 2 /g), one can consider Ru-DEMOFs being analogous when the value (SBET) is<br />

in/above the range of the parent one. In other words, the considerably high SBET of the<br />

discussed Ru-DEMOFs rules out substantial guests occlusion and pore blocking such as by<br />

non-reacted DLs, acetate or Ru/RuOx‐NPs. Interestingly, when 5-OH-ipH2 is used as DL,<br />

SBET of the Ru-DEMOFs gradually increases until 5-OH-ip is incorporated up to 32% (1c)


78 Chapter 4<br />

(Figure 4.16). Notably, the SBET of this sample 1c is the highest among all [Ru3(L)2Yy]n (L =<br />

linker) isostructural MOFs (including the single-linker and other mixed-linker Ru-<br />

DEMOFs reported so far). [82, 138] This indicates that the 5-OH-ip is a good candidate for<br />

defects engineering in Ru-MOFs.<br />

Figure 4.16. The variation of the BET surface area (S BET ) derived based on measured N 2 sorption<br />

isotherms (77 K) upon increase of the DL doping. 'Parent' stands here for parent single-linker Ru-<br />

MOF (with only BTC). In the samples 1a-1d, 5-OH-ip serves as a DL; 2a-2c – ip; 3a-3d – 5-NH 2 -<br />

ip; 4a-4c – 5-Br-ip, respectively. The values of S BET have been summarized in Table 7.4 in Chapter<br />

7.<br />

In contrast with the Ru-DEMOFs series 1 with 5-OH-ip DL, the integration of the ip DL<br />

generally results in lower porosities, as demonstrated by the results for samples 2a-d.<br />

However, the SBET of samples 2a and 2b are still higher than the SBET of the parent Ru-MOF.<br />

Nonetheless, upon increasing the ip content, the surface area decreases. Thus, considering<br />

porosity characteristics, the optimum incorporation extent of DL in these series stays at<br />

28% (sample 2b). In case of 5-NH2-ip DL, the porosity of the Ru-DEMOFs (3a-c) is fully<br />

maintained, while increasing of the feeding level to 50% leads to porosity decrease<br />

instead (sample 3d, SBET = 781 m 2 /g). These observations in 2a-d and 3a-d series are<br />

probably due to the accommodation of unreacted 5-X-ipH2 in the pores. Along with the


Chapter 4 79<br />

increasing feeding of 5-X-ipH2 the competition between BTC and 5-X-ip increases.<br />

Subsequently, it is more difficult to rule out the presence of unreacted 5-X-ipH2 as the<br />

reason of the decreasing SBET. In the same line of reasoning, the SBET of the samples 4a-c<br />

dramatically decreased from 996 to 693 m 2 /g. Hence, the optimized incorporation of 5-<br />

Br-ip can only reach up to 17% (4a). Notably, in our previous studies on the defectengineered<br />

Cu-HKUST-1, the generation of mesopores was observed. [139] Curiously, in the<br />

present case of the homologous Ru-DEMOFs, an opposite trend was observed, which<br />

might be attributed to the preference of isolated defects rather than to the correlated<br />

defects in the framework (Figure 4.17). During the formation of Ru-DEMOFs, the<br />

substitution of [Ru]-L + DL ⇌ [Ru]-DL +L (L = AcO/BTC, DL = 5-X-ip) is kinetically<br />

hindered by higher activation energy in comparison to that in the case of Cu-DEMOFs.<br />

Consequently, the substitution is possibly much slower and lead to only isolated defects.<br />

Figure 4.17. Schematic representation of point defects and correlated defects in DEMOFs. Figure<br />

is provided by Dr. O. Halbherr.<br />

4.2.3 XANES, XPS and UHV-FTIR studies: ruthenium oxidation state variation as<br />

indication of defect type formation<br />

As earlier mentioned, in our previous studies on doping of the Ru-MOF with defect pydc<br />

linker, we have assigned the introduced local defects, according to Baiker et al., [136] as<br />

modified Ru-paddlewheels with three coordinating BTC and one pydc linker, in which the<br />

pyridyl-N site acts as a weakly binding ligator-site over the Ru-dimer. Such structural<br />

irregularities are associated with (partially) missing linker function (i.e., carboxylate) and<br />

consequently steric and electronic modification of the Ru-SBUs (see Figure 4.5).<br />

Additionally, partial ruthenium reduction has been observed. Namely, apart from the<br />

expected Ru 2+ and Ru 3+ (as in the parent Ru-MOF), we revealed also distinctly reduced


80 Chapter 4<br />

framework Ru-species (Ru δ+ , 0


Chapter 4 81<br />

XANES spectra (Figure 4.18) show deviation of the edge jump derivative for the samples<br />

1a-d from that one of the parent Ru-MOF. For 1a-c, the position of edge jump gradually<br />

shifts to the lower energies, suggesting reduced Ru-sites with oxidation states lower than<br />

2+ or 3+. We relate this observation to the generation of defects A in the samples 1a-c.<br />

However, the position of the edge jump in case of the sample 1d moved again to higher<br />

energies, illustrating that the amount of reduced Ru-species is now decreased with<br />

respect to samples 1a-c. The different tendency suggests that defects B may also be<br />

created when 5-OH-ip linker was incorporated into the framework at ratios higher than<br />

32%. The same trend has been observed for the 5-NH2-ip linker (samples 3a-c). Indeed,<br />

the XANES spectra of the samples 3a-b suggest again the presence of reduced Ru δ+<br />

(0


82 Chapter 4<br />

Figure 4.20. XANES spectra of the parent Ru-MOF and Ru-DEMOF sample 4a (with 17% 5-Br-ip<br />

DL, derivative of the normalized intensity vs. Energy). Black: parent Ru-MOF; blue: 4a.<br />

Figure 4.21. XANES spectra (derivative of the normalized intensity vs. Energy) of the Ru-DEMOF<br />

materials 2a and 2b compared with the parent Ru-MOF. Black: parent Ru-MOF; blue: 2a (15% ip);<br />

dark cyan: 2b (28% ip)..


Chapter 4 83<br />

Interestingly, the position of the edge jump in case of the ip-doped samples 2a-b does not<br />

change compared with that in the parent Ru-MOF (Figure 4.21). The Ru oxidation state is<br />

not affected by the incorporation of ip, which could be taken as indication of a<br />

predominance of defects B for samples 2a-b. When the functional group of the DLs<br />

changes from small, coordinative inert H to the somewhat larger and coordinative more<br />

suitable ligator-sites like -OH, -NH2 or -Br, the defects A are favored in case of relatively<br />

low doping level. However, upon rising the doping level, defects B are increasingly created<br />

along with some remaining defects A, resulting in less pronounced variation of the<br />

oxidation state of Ru in the defect-engineered samples compared with the parent material.<br />

Figure 4.22. XANES spectra (normalized absorption vs. Energy) of the Ru-DEMOF materials (1a-<br />

1d, 2a, 2b, 3a-3c and 4a) compared with the parent Ru-MOF as well as some Ru references. Solid<br />

Black: parent Ru-MOF; solid blue: a samples of the respective series 1-4; solid dark cyan: b<br />

samples of the respective series 1-4; soid magenta: c samples of the respective series 1-4; solid<br />

dark yellow: d samples of the respective series 1-4; dash dot violet: Ru; solid orange: RuO 2 ; short<br />

dash wine: Ru nano-particles(NPs).


84 Chapter 4<br />

It is important to note that the white line in the XANES spectra (normalized absorption vs.<br />

energy, Figure 4.22) of all the Ru-DEMOFs samples as well as the parent Ru-MOF are<br />

strikingly different from those measured for the Ru-NPs, metallic Ru and RuO2. These<br />

observations support PXRD results (see above and Figures 4.6 and 4.7) and allow also to<br />

rule out the presence of other Ru-phases as significant impurities in the Ru-DEMOFs<br />

including parent Ru-MOF. [223] Thus, we assume that it is valid to assign all Ru-species to<br />

the metal-nodes of the frameworks. Further information on the local environment of Ru<br />

can be obtained also from EXAFS. The analysis of the experimental data is, however, a<br />

formidable effort, since many different models for Ru-DEMOFs have to be considered and<br />

a variety of local structures may be present in parallel in the samples as discussed before.<br />

Therefore, a thorough analysis of the EXAFS-data is outside the scope of this dissertation<br />

project.<br />

The different samples fabricated in this study have also been characterized using Highresolution<br />

X-ray photoelectron spectroscopic (XPS) in order to confirm the assignments<br />

based on XANES and for a more quantitative estimation of the relative abundance of the<br />

different Ru-species (Ru 3+ , Ru 2+ and Ru δ+ ) in the Ru-DEMOFs. The samples 1a, 1c, 1d, 2a<br />

and 2b with 5-OH-ip and ip DLs, respectively, have been selected. Remarkably, similar to<br />

the related pydc-doped Ru-DEMOFs we elaborated earlier, [138] reduced Ru δ+ -species have<br />

been found in the Ru-DEMOFs (samples 1a, 1c and 1d) in addition to the Ru 2+ - and Ru 3+ -<br />

species. Two Ru 3d doublets (3d5/2 and 3d3/2) at ca. 281.6 and 285.8 eV as well as at 282.6<br />

and 286.8 eV, attributed to the Ru 2+ - and Ru 3+ -species, respectively, are seen in the<br />

deconvoluted XP spectra of all measured samples (Figure 4.23). In addition, another Ru<br />

3d doublet appears at ca. 280.5 and 284.7 eV in the Ru-DEMOFs samples 1a, 1c and 1d.<br />

On the basis of quantitative estimations of the XP spectra, some variation of the Ru 3+ /Ru 2+<br />

ratio and, in particular, reduction to Ru δ+ in 1a, 1c and 1d can be clearly seen (Table 4.4).<br />

From the parent Ru-MOF to the Ru-DEMOF sample 1a, the ratio of Ru 3+ / Ru 2+ / Ru δ+ varies<br />

from 1 / 1.24 / 0 to 1 / 1.75 / 0.08, indicating that the incorporation of DL induces partial<br />

ruthenium reduction (i.e., relative concentrations of the Ru 2+ and Ru δ+ increase).<br />

Moreover, upon increasing the incorporation level of 5-OH-ip to 32% for 1c (feeding level<br />

30%), the amount of the reduced Ru δ+ -species considerably increases to the highest value<br />

(Ru 3+ / Ru 2+ / Ru δ+ = 1 / 1.69 / 0.59) among all studied samples. Importantly, the Ru δ+ -<br />

related doublet decreases significantly in intensity with further increase of the<br />

incorporation to 37% for 1d (feeding level 50%). This is in contrast to the observation for


Chapter 4 85<br />

Figure 4.23. Deconvoluted XP spectra of the Ru-DEMOFs samples (A)1a (8% of 5-OH-ip), 1c (32%<br />

of 5-OH-ip) and 1d (37% of 5-OH-ip), (B) 2a (15% of ip) and 2b (28% of ip) in Ru 3d and C 1s<br />

regions in comparison with parent Ru-MOF. Original figures are provided by P. Guo.<br />

pydc-doped Ru-DEMOFs, where a continuous rise of the Ru δ+ signals was detected along<br />

with increasing degree of pydc incorporation. [138] The relatively lower concentration of<br />

Ru 2+ and Ru δ+ in 1d (Ru 3+ / Ru 2+ / Ru δ+ = 1 / 1.23 / 0.09), as compared to the respective<br />

ratios estimated for 1c, can be explained by our assumption of simultaneous generation<br />

of both defects A and B with distinct preference of B over A at higher feeding<br />

concentrations of 5-OH-ip. Interestingly, for the ip-doped samples 2a-b no highly reduced


86 Chapter 4<br />

Ru δ+ have been observed in the XP spectra (no Ru3d doublets at ca. 280.5 and 284.7 eV;<br />

Figure 4.23). These data confirm the assignment deduced from XANES data, suggesting<br />

that application of DL with the coordinatively inactive functional group H (2a-b) is likely<br />

to more selectively induce missing node defects B in the Ru-DEMOFs without Ru<br />

reduction (as this is indicative for a type A defect). According to the results of XANES and<br />

XPS, we thus conclude that the Ru-DEMOFs materials with 5-X-ip (X= OH, NH2 and Br)<br />

contain predominantly defects A, especially at moderate incorporation levels. However,<br />

the simultaneous generation of defects B in these samples is also likely as indirectly<br />

deduced from the non-linear dependence of the XANES and XPS data on the doping level.<br />

Table 4.4. The assignments of binding energies as well as the ratio of Ru 3+ / Ru 2+ / Ru δ+ (calculated<br />

from the deconvoluted spectra) in the parent Ru-MOF sample and Ru-DEMOFs (1a, 1c 1d, 2a and<br />

2b). 1a, 1c and 1d - 8%, 32% and 37% of 5-OH-ip, respectively; 2a and 2b - 15% and 28% of ip,<br />

respectively. Data provided by P. Guo.<br />

Sample Binding energy, eV Assignment Ru 3+ / Ru 2+ / Ru δ+ ratio<br />

Parent Ru-MOF<br />

1a<br />

1c<br />

1d<br />

2a<br />

285.86 / 281.66 Ru 2+<br />

286.86 / 282.66 Ru 3+<br />

285.86 / 281.66 Ru 2+<br />

286.86 / 282.66 Ru 3+<br />

284.65 / 280.45 Ru δ+<br />

285.71 /281.51 Ru 2+<br />

286.71 / 282.60 Ru 3+<br />

284.60 / 280.40 Ru δ+<br />

285.72 / 281.52 Ru 2+<br />

286.72 / 282.52 Ru 3+<br />

284.65 / 280.45 Ru δ+<br />

285.86 /281.66 Ru 2+<br />

286.86 / 282.66 Ru 3+<br />

1 / 1.24 / -*<br />

1 / 1.75 / 0.08<br />

1 / 1.69 / 0.59<br />

1 / 1.23 / 0.09<br />

1 / 1.73/ -*<br />

285.88 / 281.68 Ru 2+<br />

2b<br />

286.88 / 282.76 Ru 3+<br />

* no species have been determined.<br />

1 / 1.42 / -*<br />

Consequently, UHV-IR spectroscopic investigations employing CO as a probe molecule<br />

have been carried out on two representative samples 1c and 1d to gain complementary


Chapter 4 87<br />

evidence on the valence states of framework Ru sites and the associated defect types.<br />

Parent Ru-MOF was reported to reveal bands at 2172 and 2127 cm -1 , arising from CO<br />

interactions with Ru 3+ and Ru 2+ , respectively (Figure 4.24). [138] Apart from these two<br />

bands, the spectra of the samples 1c and 1d show additional bands at 2041 and 1998 cm -<br />

1 , which can be attributed to CO bound to reduced Ru δ+ (0


88 Chapter 4<br />

Figure 4.25. UHV-IR spectra of Ru-DEMOF 1c and 1d compared with parent Ru-MOF and Ru-<br />

DEMOF with 30% pydc incorporation upon CO 2 dosing at 90-95 K after annealing the sample at<br />

475-500 K. The bands at 2041 cm -1 and 1993 cm -1 are assigned to vibrations of CO bound to<br />

reduced Ru δ+ -sites and indicate dissociative chemisorption of CO 2 . Original figures are provided<br />

by M. Kauer.<br />

ration (1d). The formation of defect B, namely missing Ru-paddlewheels, is accompanied<br />

by lowering the abundance of Ru 2+ and Ru δ+ metal centers in the framework, which


Chapter 4 89<br />

explains the observed changes of the decreased intensity of Ru 2+ and Ru δ+ in the sample<br />

1d.<br />

In our earlier report on pydc-doped Ru-DEMOFs, we communicated on the low<br />

temperature CO2→CO dissociative chemisorption under UHV conditions in a dark<br />

environment, a feature which is not inherent for the parent Ru-MOF. [138] We concluded<br />

that the CO2→CO reduction might be triggered by the reduced Ru δ+ -sites and the pyridyl<br />

N-sites in the proximity of the modified Ru-centers at the nodes. In order to confirm these<br />

assumptions further, we have studied the same reaction at Ru-DEMOFs that feature<br />

reduced Ru δ+ -sites. As these sites have been created by the incorporation of the 5-OH-ip<br />

linker in 1a-d, we have selected two representative samples, 1c and 1d, both exhibiting<br />

high 5-OH-ip incorporation but quite different levels of Ru 2+ and Ru δ+ . As shown in Figure<br />

4.25, upon CO2 dosing for both the parent Ru-MOF and Ru-DEMOFs 1c, an intense and<br />

broad IR band appears at 2336 cm -1 that is assigned to the asymmetric stretching mode<br />

νas(CO2) of physisorbed CO2 binding linearly at various Ru-sites. [198] The weak band at<br />

2273 cm -1 indicates the presence of a minority CO2 species originating probably from the<br />

adsorption of a small amount of 13 CO2 with an expected isotopic shift of 1.03. Apart from<br />

the CO2-related bands, two other bands at 2041 cm -1 and 1993 cm -1 have been observed<br />

in 1c, that are assigned to vibrations of CO bound to the reduced Ru δ+ -sites, matching our<br />

previous data on pydc-doped Ru-DEMOF. [138] This result indicates dissociative<br />

chemisorption of CO2 to CO (reduction) in Ru-DEMOF with 5-OH-ip DL (1c), whereas the<br />

parent Ru-MOF is essentially inactive. However, in comparison to the extremely high<br />

reactivity of pydc-doped Ru-DEMOFs, in which CO2 is nearly totally converted to CO at<br />

higher concentrations of pydc defective linker (e.g. 30%), [138] the production of CO is<br />

rather limited for 5-OH-ip doped Ru-DEMOFs (Figure 4.25). Interestingly, Ru-DEMOF 1d,<br />

is almost inactive for CO2 dissociation. We correlate this observation to the much less Ru δ+<br />

and in turn to the more significant formation of missing node defects (type B) for 1d as<br />

compared to 1c. Overall, the presented IR data demonstrate that the 5-OH-ip doped Ru-<br />

DEMOF samples show much less reactivity for CO2 activation as compared to pydc doped<br />

samples. This cannot be attributed only to the relatively low concentration of reduced<br />

Ru δ+ . In previous study it was suggested that the CO2 activation being promoted by pydc<br />

linker due to the presence of basic pyridyl N sites in proximity to the reactive Ru δ+ . [138]<br />

This hypothesis is indirectly supported by the properties of samples 1c and 1d.


90 Chapter 4<br />

Table 4.5. Possible defect combinations in the prepared Ru-DEMOFs according to the conducted<br />

XANES, UHV-IR and XPS studies.<br />

DL Sample Assumption I Assumption II<br />

1a defectsA* + B defects A*<br />

5-OH-ip<br />

1c defects A*** + B defects A***<br />

1d defects A* + B defects A* + B<br />

ip<br />

2a<br />

2b<br />

defects B<br />

defects B<br />

3a defects A** + B defects A**<br />

5-NH 2 -ip<br />

3b defects A*** + B defects A***<br />

3c defects A* + B defects A* + B<br />

5-Br-ip 4a defects A + B defects A<br />

* expresses relative defects concentration.<br />

Figure 4.26. Other possible defect fragments in the structure of Ru-DEMOFs. a). Two DLs and two<br />

BTC in a fragment of modified paddlewheel b). Two DLs and two HBTC in a fragment of missing<br />

node. Free carboxylates of HBTC were bound with each other via hydrogen bounds. c). Correlated<br />

modified paddlewheel and missing paddlewheel.<br />

The combined spectroscopic characterization data (XANES, XPS, UHV-FTIR with CO and<br />

CO2 as probes) presented above suggest more or less consistent, but complicated picture<br />

on the abundance of the two kinds of defects. Overall, we assume both types A and B being<br />

simultaneously generated in the Ru-DEMOFs in the case of coordinative weakly binding<br />

ligator-sites at the fragmented linkers. Type A appears to be favored at low incorporation<br />

levels, e.g. 1a (8%) and becomes more abundant up to a certain threshold, e.g. 1c (32%).<br />

Along with a further increase of incorporation level, e.g. 1d (37%), defects of type B


Chapter 4 91<br />

appear to dominate over type A. One should be aware that the possibility of formation and<br />

distribution of various defect types is rather diverse and may not be restrict to those<br />

mentioned here. The arrangement of the two defect types in Ru-DEMOFs (1a, 1c and 1d)<br />

is rather diverse. Still, based on the information on XANES and UHV-IR with CO probing,<br />

we can assume two possibilities:<br />

i) both defects of type A and B might be generated in the Ru-DEMOFs (1a-1d) with the<br />

incorporation of 5-OH-ip. The defects of type A are much more competitive and became<br />

gradually dominant in case of the lower doping (samples 1a (8%) and 1c (32%)). Along<br />

with increase of doping level (1d, 37%) of 5-OH-ip, the dominant effect of defects A is<br />

substituted slowly by the defects B;<br />

ii) the defects of type A are generated and becomes gradually dominant in case of<br />

relatively low doping (samples 1a and 1c). Afterwards, defects B are created<br />

simultaneously when the doping level reaches 37% in case of 1d. Therefore, the influence<br />

caused by the defects A could be partly eliminated. Nevertheless, both assumptions<br />

indicate that the defects A play a major role in Ru-DEMOFs 1c (32% of 5-OH-ip). A<br />

comprehensive table summarizing possible defect combinations in the particular Ru-<br />

DEMOFs (based on the performed analyses) is given in Table 4.5 and Figure 4.26.<br />

Additional characterization and quantitative determination of the defect types requires<br />

additional work and support by theoretical modeling.<br />

4.2.4 CO2, CO and H2 sorption properties of Ru-DEMOFs<br />

The previous characterizations suggest the existence of two types of defects (A and B)<br />

being present in the samples in different absolute and relative amounts as a consequence<br />

of the 5-X-ip framework incorporation. Both are likely to influence the gas sorption<br />

properties of the Ru-DEMOFs. Hence, we studied the adsorption of CO2, CO and H2 at range<br />

of samples. In general, the results obtained and discussed in the following are not very<br />

conclusive as the observed changes in gas uptake as a function of DL incorporation are<br />

small or even close to the error of the measurements. Nevertheless, we present the data<br />

and provide a tentative discussion with respect to the possible counter acting effects of<br />

the two types of defects in the samples.


CO 2<br />

adsorbed, mmol/g<br />

92 Chapter 4<br />

3<br />

parent Ru-MOF<br />

1a<br />

1c<br />

1d<br />

2<br />

1<br />

0<br />

0 200 400 600 800 1000<br />

P, mbar<br />

Figure 4.27. CO 2 isotherms collected at 298 K for the parent Ru-MOF and DEMOF derivatives 1a<br />

(8%), 1c (32%) and 1d (37%) with 5-OH-ip DL. Black circles – parent Ru-MOF; blue triangles –<br />

1a; magenta squares – 1c; dark yellow stars – 1d.<br />

CO2 adsorption isotherms (298 K) of the parent reference Ru-MOF sample and 1a and 1c<br />

display a gradual enhancement of the adsorption capacity in the order Ru-MOF < 1a < 1c<br />

along with rising incorporation degree of DL (Figure 4.27). However, the CO2 uptake of<br />

the sample 1d revealed to be 2.9 mmol/g at 1 bar, and lies within the order Ru-MOF < 1d<br />

< 1c (Table 4.6). According to our studies on different defects in the series 1a-1d, these<br />

results are not unexpected. The increase of the CO2 uptake is attributed to the reduced Rusites<br />

at the type A defects and somewhat lowered uptake of 1d is attributed to less<br />

abundant Ru-sites as a consequence of missing metal-nodes (i.e., type B defects). Hence,<br />

higher incorporation level of DL in the sample 1d does not promote the increase of the<br />

CO2 uptake.


CO uptake, mmol/g<br />

Chapter 4 93<br />

Table 4.6. H 2 (77 K) and CO 2 (298 K) uptakes for the parent Ru-MOF and its defect-engineered<br />

variants 1a, 1c, 1d, 2a, 2b, 3a-3c and 4a at 1000 mbar. 1: DL is 5-OH-ip; 2: DL is ip; 3: DL is 5-<br />

NH 2 -ip; 4: DL is 5-Br-ip.<br />

Sample name CO 2 uptake (mmol/g) H 2 uptake (mmol/g)<br />

Parent Ru-MOF 2.7 6.8<br />

1a 3.2 9.6<br />

1c 3.4 8.9<br />

1d 2.9 7.6<br />

2a 3.0 7.4<br />

2b 2.8 6.7<br />

3a 3.4 8.3<br />

3b 3.2 7.8<br />

3c 2.9 7.4<br />

4a 2.6 7.1<br />

2.5<br />

2.0<br />

parent Ru-MOF<br />

1a<br />

1c<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

0 200 400 600 800 1000<br />

P, mbar<br />

Figure 4.28. CO isotherms measured at 298 K for the parent Ru-MOF and Ru-DEMOF samples 1a<br />

(8%) and 1c (32%) with 5-OH-ip DL. Black circle – parent Ru-MOF; blue triangles – 1a; magenta<br />

squares – 1c.


H 2<br />

adsorbed, mmol/g<br />

94 Chapter 4<br />

Furthermore, CO adsorption of the parent Ru-MOF, 1a and 1c at 298 K also illustrates the<br />

tendency of enhanced uptake in case of the Ru-DEMOF samples (Figure 4.28). The small<br />

difference between the sample 1a and the parent Ru-MOF on CO adsorption is attributed<br />

to the low incorporation degree of the 5-OH-ip DL (8%). Along with the increase of the DL<br />

incorporation (such as in 1c), the CO uptake rises even at very low pressure (ca. 4 mbar).<br />

This increase of the CO capacity of the Ru-DEMOF is probably due to the generation of the<br />

modified paddlewheels of type A, which expose more open sites as one carboxylate<br />

ligator-site of BTC is substituted by the hydroxyl-group of 5-OH-ip. Besides, the average<br />

electronic density is varied as the oxidation state of ruthenium at the CUS is changed. Both<br />

could contribute to the enhancement of the adsorption properties.<br />

10<br />

8<br />

6<br />

4<br />

2<br />

parent Ru-MOF<br />

1a<br />

1c<br />

0<br />

200 400 600 800 1000<br />

P, mbar<br />

Figure 4.29. H 2 isotherms measured at 77 K for the parent Ru-MOF and DEMOFs 1a (8%) and 1c<br />

(32%) with 5-OH-ip DL. Black circles – parent Ru-MOF; blue triangles – 1a; magenta squares – 1c.<br />

The H2 adsorption isotherms (at 77 K) of the parent Ru-MOF and the Ru-DEMOFs 1a and<br />

1c follow nearly the same trend as in case of CO2 and CO adsorption (Figure 4.29). Both<br />

defect-engineered materials 1a and 1c exhibit higher H2 uptake at 1 bar than the parent<br />

intact framework. In fact, there should be two main kinds of adsorption sites in the Ru-


-Q st<br />

, kJ/mol<br />

Chapter 4 95<br />

DEMOFs (1a and 1c): 1) strong binding affinity between Ru-CUSs and hydrogen, namely<br />

Ru 2+/3+ -sites in the regular paddlewheel units and the Ru δ+ -sites in the modified<br />

paddlewheels (in defects A); 2) the relatively weaker interactions (ie. physisorption and<br />

van der Waals induced interactions) between the hydrogen and ligands/pores walls in the<br />

framework. Obviously, due to the lack of Ru δ+ -sites in the parent Ru-MOF, the Ru-DEMOFs<br />

exhibit higher H2 uptake than the parent Ru-MOF. To quantitatively understand the H2<br />

binding affinity of these samples, the isosteric heats of H2 adsorption (–Qst) was calculated<br />

and reveal an order with 1c > 1a > parent Ru-MOF when comparing the respective values<br />

at 1 mmol (H2) / Ru2-paddlewheel (what might be considered as the saturation<br />

adsorption of H2 at the regular paddlewheel) (Figure 4.30). This suggests stronger binding<br />

affinity of H2 for Ru-DEMOF with 32% of incorporated 5-OH-ip (sample 1c), which again<br />

is attributed to the relative higher amount of Ru δ+ -sites in 1c.<br />

8 parent Ru-MOF<br />

1a<br />

1c<br />

7<br />

6<br />

5<br />

0.0<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0<br />

H 2<br />

adsorbed, mmol/Ru 2<br />

pw<br />

Figure 4.30. Isosteric heats of adsorption of 1a and 1c calculated from the H 2 adsorption<br />

isotherms at 77K and 87K. The vertical line stands for the position where one H 2 molecule is<br />

absorbed per Ru 2 -paddlewheel.


CO 2<br />

adsorbed, mmol/g<br />

96 Chapter 4<br />

Moving further to the 5-NH2-ip DL, respective Ru-DEMOFs (3a-3c) display the same<br />

tendencies in their absorption behavior of CO2 and H2 as in their SBET (N2). From the results<br />

of the XANES studies, we have already revealed that defects B are likely to be generated<br />

at lower doping levels 3a-c (5-NH2-ip) as compared to the samples 1a-d (5-OH-ip). This<br />

difference in defect A/B abundance leads to variation of the CO2 and H2 adsorption in case<br />

of the 3a-c series. Even though 3a-c display higher uptake of CO2 and H2 than the parent<br />

Ru-MOF, the uptake of CO2 and H2 does not further increase after the feeding level of 5-<br />

NH2-ip is raised above 10% (3a) (Figures 4.31 and 4.32). Similar to 1d, we assign the<br />

decrease of the uptake of CO2 and H2 at 1 bar observed for 3c as compared with 3a-b to<br />

more abundant type B defects.<br />

3<br />

parent Ru-MOF<br />

3a<br />

3b<br />

3c<br />

2<br />

1<br />

0<br />

0 200 400 600 800 1000<br />

P, mbar<br />

Figure 4.31. CO 2 isotherms collected at 298 K for the parent Ru-MOF and its derivatives 3a, 3b<br />

and 3c with 5-NH 2 -ip DL. Black circles – parent Ru-MOF; blue triangles – 3a; dark cyan diamonds<br />

– 3b; magenta squares – 3c.


CO 2<br />

adsorbed, mmol/g<br />

H 2<br />

adsorbed, mmol/g<br />

Chapter 4 97<br />

8<br />

6<br />

4<br />

2<br />

parent Ru-MOF<br />

3a<br />

3b<br />

3c<br />

0<br />

0 200 400 600 800 1000<br />

P, mbar<br />

Figure 4.32. H 2 isotherms collected at 77 K for the parent Ru-MOF and its derivatives 3a, 3b and<br />

3c with 5-NH 2 -ip DL. Black circles – parent Ru-MOF; blue triangles – 3a; dark cyan diamonds – 3b;<br />

magenta squares – 3c.<br />

3.0<br />

2.5<br />

2.0<br />

parent Ru-MOF<br />

4a<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

0 200 400 600 800 1000<br />

P, mbar<br />

Figure 4.33. CO 2 isotherms measured at 298 K for the parent Ru-MOF and Ru-DEMOF 4a (17%)<br />

with 5-Br-ip DL. Black circles – parent Ru-MOF; blue triangles – 4a.


H 2<br />

adsorbed, mmol/g<br />

98 Chapter 4<br />

8<br />

6<br />

4<br />

2<br />

parent Ru-MOF<br />

4a<br />

0<br />

0 200 400 600 800 1000<br />

P, mbar<br />

Figure 4.34. H 2 isotherms measured at 77 K for the parent Ru-MOF and Ru-DEMOF 4a (17%) with<br />

5-Br-ip DL. Black circles – parent Ru-MOF; blue triangles – 4a.<br />

Another indication of the suggested counter acting effects of the defect sites is the<br />

comparison of the CO2 and H2 uptakes at 2a, 2b and 4a with the parent Ru-MOF (Figures<br />

4.33, 4.34, 4.35 and 4.36). However, the variations are too small to be regarded as<br />

significant. Still, some assumptions can be concluded, which afford us some ideas on the<br />

future design of DEMOFs towards considerable optimization of sorption properties. In<br />

spite of the creation of defects A with Ru δ+ -sites, the rather small change of the H2 and CO2<br />

uptake in the sample of 4a (17% 5-Br-ip) might be contributed to the functional –Brgroups<br />

from the 5-Br-ip linker at the defect-sites. One the one hand, they may create larger<br />

steric hindrances (compared to the other DLs used) blocking, thus, the access for the gas<br />

molecules in spite of the modified paddlewheels with more rooms around Ru-sites. On the<br />

other hand, weaker H2 binding affinity has been found in MOFs composed of ligands with<br />

electron-withdrawing groups. [229-230] Although this influence caused from the interaction<br />

between H2 and ligands is smaller in comparison with the strong one of H2-M sites, it can<br />

be one reason of the observation study in the sample 4a.


CO 2<br />

adsorbed, mmol/g<br />

Chapter 4 99<br />

3.5<br />

3.0<br />

2.5<br />

parent Ru-MOF<br />

2a<br />

2b<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

0 200 400 600 800 1000<br />

P, mbar<br />

Figure 4.35. CO 2 isotherms collected at 298 K for the parent Ru-MOF and DEMOF derivatives 2a<br />

(15%) and 2b (28%) with ip DL. Black circles – parent Ru-MOF; blue triangles – 2a; dark cyan<br />

diamonds – 2b.<br />

The CO2 and H2 adsorption isotherms of 2a and 2b show that 2a exhibits higher uptake.<br />

The adsorption capacity of the sample 2b revealed to be a little lower compared to 2a<br />

while it is still the same as with the parent Ru-MOF (Figures 4.35 and 4.36). Although the<br />

defects B in the Ru-DEMOFs 2a and 2b do not induce ruthenium reduction, vacant-sites<br />

caused from the creation of defects B (missing of Ru-paddlewheels) could still<br />

accommodate gas molecules. In addition, the vacant-sites in the defects B make the gas<br />

molecules access to the surrounding CUS easier. This could be the positive effect of the<br />

defects B in the Ru-DEMOFs with ip DL. On the other hand, the generation of defects B<br />

leads to the partial missing of the di-ruthenium PW units and, hence, the overall CUSs in<br />

Ru-DEMOFs decrease. In the case of Ru-DEMOF 2a, the positive effect of the defects B is<br />

rather dominant and results in the higher H2 and CO2 capacity. But when it comes to 2b,<br />

the negative effect of the defects B eliminates the positive influence on the gas uptake,<br />

which leads to the observation we see in Figures 4.35 and 4.36. The Ru-DEMOFs 1a and<br />

1c featuring higher gas (H2 and CO2) uptake in comparison with 2a (Table 4.6), implying


H 2<br />

adsorbed, mmol/g<br />

100 Chapter 4<br />

that the enhancement of sorption of small molecules (CO2 and H2) attributed to defects A<br />

is stronger than the influence caused from defects B.<br />

8<br />

6<br />

4<br />

2<br />

parent Ru-MOF<br />

2a<br />

2b<br />

0<br />

0 200 400 600 800 1000<br />

P, mbar<br />

Figure 4.36. H 2 isotherms measured at 77 K for the parent Ru-MOF and its derivatives 2a (15%)<br />

and 2b (28%) with ip DL. Black circles – parent Ru-MOF; blue triangles – 2a; dark cyan diamonds<br />

– 2b.<br />

4.2.5 Catalytic test reactions<br />

Ethylene dimerization utilizing MOF supported catalysts has recently attracted a lot of<br />

attention due to its modifiable characteristic under molecular level. [231-234] Most of the<br />

reported MOF catalysts bear isolated single-metal (e.g., Ni, Ir) active-sites. Ru-MOF<br />

materials as heterogeneous catalysts for the ethylene dimerization have not been studied<br />

yet, although chemisorbed Ru-complex on de-aluminated zeolite Y and Ru-complex itself<br />

catalyzing ethylene to butene have been already reported. [235-237] It is known that the<br />

redox activity of Rh in RhCa-X Zeolite plays an important role on catalyzing ethylene<br />

dimerization. [238] Having distinct Ru-CUS available in the Ru-DEMOFs (i.e., different Ru n+ -<br />

species being potentially involved into the redox-process(es) of the reaction flow),


Chapter 4 101<br />

samples 1a (8% of 5-OH-ip) and 1c (32% of 5-OH-ip) have been selected for testing them<br />

as catalysts for ethylene dimerization. In addition, parent Ru-MOF has been used as a<br />

reference. Thus, activated Ru-DEMOFs were loaded into a reactor to catalyze ethylene<br />

Table 4.7. The conversion of ethylene to butene in toluene under different conditions.<br />

Entry Catalyst (mg) Temperature ( °C) Time (h) Additive TOF (h -1 )<br />

1 1c (2.5) 80 24 ---- 0.92<br />

2 1c (2) 80 1 Et 2 AlCl 4.38<br />

3 1c (2) 80 1 ---- 0.88<br />

4 1c (2.4) 26 2 Et 2 AlCl 2.15<br />

5 1c (2.1) 80 2 Et 2 AlCl 4.36<br />

6 1a (2) 80 2 Et 2 AlCl 2.3<br />

7 Parent Ru-MOF (1.8) 80 2 Et 2 AlCl 2.05<br />

dimerization in toluene at different temperatures under the pressure of 800 psi (≈ 55 bar).<br />

In general, all reactions led to the formation of butene without producing any other α-<br />

olefins (Tables 4.7). Addition of Et2AlCl as co-catalyst accelerates the conversion (Table<br />

4.7, entry 2 vs. entry 3). When the reaction has been conducted at 26 °C, lower yield has<br />

been obtained compared to those at 80 °C (Table 4.7, entry 4 vs. entry 5). However,<br />

reactions conducted for 1 h and 24 h give almost the same TOF (0.88 vs. 0.92 h -1 ),<br />

suggesting that the Ru-DEMOF 1c does not suffer the deactivation as a function of onstream<br />

time. Therefore, parent Ru-MOF and Ru-DEMOFs 1a and 1c have been employed<br />

to study the dimerization of ethylene in toluene under optimized condition (800 psi, 80<br />

°C, 2 h) in the presence of Et2AlCl. The obtained product was analyzed by gas<br />

chromatography. As it can be seen from entry 5-7 in Table 4.7 and Figure 4.37, the TOF<br />

value of 1a and 1c increase along with increase of the incorporation level of DL in<br />

comparison with the parent Ru-MOF. Since the contents of the reduced Ru δ+ -sites in 1c is<br />

the highest among the 1a-d series (according to our study above), there should be more<br />

M-CUSs in 1c and therefore it affords the chance to enhance the catalytic activity in the<br />

reaction. Indeed, the transformation of ethylene to butene is the highest (TOF = 4.36 h -1 )<br />

when employing 1c as a catalyst. This value is still much lower than that using zeolite<br />

supported Ru-complex catalysts in the presence of H2 (4.36 vs. 216 h -1 ). However, it is<br />

higher than the TOF value (1 × 10-4 s -1 = 0.36 h -1 ) in the reported reaction without the


102 Chapter 4<br />

presence of H2. [236] Distinction between butene isomers, optimization of the reaction<br />

conditions using Ru-DEMOF materials (such as usage of H2) as well as clarifying the<br />

mechanism of this reaction are outside the scope of this thesis. Nevertheless, the<br />

preliminary results give a rather clear hint that obtained Ru-materials could be utilized as<br />

catalyst for the dimerization of ethylene. Due to the diverse modification of MOF materials<br />

themselves, our current investigations on the ethylene dimerization using Ru-DEMOFs<br />

can afford a platform for further studies on MOFs as catalysts in this field.<br />

Figure 4.37. The tendency of TOF value for parent Ru-MOF and Ru-DEMOF samples 1a and 1c<br />

utilized as catalysts for transformation of ethylene into toluene (800 psi, 80 °C, 2h) in the presence<br />

of Et 2 AlCl (0.81 ml, 1M in heptane). TOF (turnover frequency) = mol product / (mol Metal<br />

(catalyst)*h).<br />

Paal-Knorr pyrrole synthesis has attracted great attention due to the huge synthetic<br />

variety of pyrroles and their derivatives, which are key intermediates for various<br />

pharmaceutical drugs. [239-240] The reaction is typically performed under protic or Lewis<br />

acidic conditions, [241-243] using primary amines and diketones. Ruthenium complex as<br />

catalysts for pyrrole synthesis was rather rare, [244] where the usage of formate salts such<br />

as sodium formate as activators is required. Due to the ability of MOFs on tuning the pore<br />

size and modifying the diverse structure, MOFs have explored to be as catalysts on various<br />

reaction. [29] Phan et al. reported IRMOF-3 as heterogeneous catalyst for the Paal–Knorr


Chapter 4 103<br />

reaction of benzyl amine with 2,5-hexanedione, where it was mentioned that the catalytic<br />

sites might be related to the structure defects. As we know from earlier reports, parent<br />

Ru-MOF ([Ru3(BTC)2Y1.5]n) is constructed from Ru2-paddlewheels, in which the axial Rupositions<br />

are partly occupied by strongly binding Y or by other weakly guest molecules.<br />

After thermal treatment, a fraction of the axial positions can be exposed as open Lewis<br />

acidic sites (useful as catalytic site, for instance). With regard to the structure of the Ru-<br />

DEMOFs discussed in this contribution, two kinds of defects are present. The modified<br />

paddlewheel units (defects A) could provide the materials with additional sites around<br />

the partly reduced Ru-centers. On the other hand, the eliminated entire Ru-paddlewheel<br />

clusters (defects B) can form vacant-sites to play a role on affecting the adsorption<br />

properties, especially when the functional X-group at the defect generating linker is as<br />

small as H. Hence, Ru-DEMOFs 1a-1c, 2a and 2b have been utilized for catalyzing the<br />

reaction of 2,5-hexadione and aniline in toluene at 90 °C to form dimethyl-phenyl-1Hpyrrole<br />

(Scheme 4.1).<br />

Scheme 4.1. Scheme of the Paal-Knorr reaction.<br />

It is interesting to note that all of the selected Ru-DEMOFs show apparently enhanced<br />

conversion of aniline to the pyrrole within the initial 4 hours in comparison with the<br />

parent Ru-MOF (Figure 4.38). Among the used materials, Ru-DEMOFs 1a and 2a exhibit<br />

the highest catalytic activity, although the other Ru-DEMOFs (1b, 1c, 1d and 2b)<br />

incorporate more DL (Table 4.8). The steric hindrances and distinct defect types (A vs. B)<br />

could be the main reason of this phenomenon. When Ru-DEMOF samples with 5-OH-ip DL<br />

(1a-d) were employed as catalysts, the yield of pyrrole has been increased from 48%<br />

(using parent Ru-MOF as the catalyst) to 58% (1d), 65% (1c), 73% (1b) and 77% (1a),<br />

respectively. This is assigned to the presence of the reduced Ru-centers (defects A) in<br />

these materials. Reduced, softer binding Ru δ+ -sites can be more easily coordinated to the<br />

O atom of the carbonyl group, which favors the nucleophilic attack from the lone-pair of<br />

the amino group of aniline. However, when the incorporation of 5-OH-ip increases, the<br />

gradual dominance of the defects B in these materials probably eliminates part of the


104 Chapter 4<br />

reactive metal centers, thus, affecting the catalytic activity of the 1b-d samples. On the<br />

other hand, Ru-DEMOF 2a displays higher yield of pyrrole compared to the parent Ru-<br />

MOF (82% vs. 48%), regardless of the partial absence of the metal centers (defects B)<br />

(Figure 4.39). This could be a result of the smaller steric hindrance surrounding the Ru-<br />

CUSs while employing the H2ip DL. When the contents of the incorporated DLs are rather<br />

small (i.e., as in 2a), the negative influence caused by missing metal centers is relatively<br />

insignificant compared with the positive effect of the smaller steric hindrance. Therefore,<br />

2a Ru-MOF material shows as high catalytic activity in present reaction as 1a. Similar<br />

observations have been found for the samples 1b-c because of the somewhat larger steric<br />

hindrance of OH-groups in spite of the type A defect in these materials. To note, using<br />

IRMOF-3 as catalysts in the same reaction, 96% conversion after 60 min was achieved.<br />

However, this could be attributed to the higher regent ratio (aniline: 2,5-hexanedione =<br />

1:1.7). [245] Nevertheless, the current study using Ru-DEMOFs as catalysts give us insight<br />

into the influence of catalytic activity caused by different defect types. For further<br />

investigation, it would also be interesting to optimize the catalytic activity by improving<br />

the regent ratio.<br />

Figure 4.38. Time-yield plot of Paal-Knorr synthesis of aniline reacting with 2,5-hexadione, giving<br />

the pyrrole employing Ru-DEMOFs 1a-1d (with 8%, 20%, 32% and 37% 5-OH-ip incorporation,<br />

respectively) as catalysts in comparison with the parent Ru-MOF(parent). Blank stands for the<br />

experiment where no catalyst was added into the reaction mixture.


Yield, %<br />

Chapter 4 105<br />

Table 4.8. Comparison of the yields of pyrrole and TOF values after 4 h in the Paal-Knorr reaction<br />

for different Ru-DEMOF samples as well as the parent Ru-MOF (used as a reference).<br />

Sample Ru content (mol %) yield at 4h, % TOF(h -1 )*<br />

Parent Ru-MOF 2 48 6.17<br />

1a 2 77 9.83<br />

1b 2 73 9.25<br />

1c 2 65 8.17<br />

1d 2 58 7.25<br />

2a 2 82 10.36<br />

2b 2 57 7.09<br />

blank - 28 -<br />

* TOF value was calculated by mol of product (mol of Ru) -1 h -1 .<br />

100<br />

80<br />

60<br />

40<br />

20<br />

2a<br />

2b<br />

parent Ru-MOF<br />

blank<br />

0<br />

0 4 8 12 16 20 24<br />

Time, h<br />

Figure 4.39. Time-yield plot of Paal-Knorr synthesis of aniline reacting with 2,5-hexadione, giving<br />

the pyrrol using Ru-DEMOFs 2a (15% ip incorporation) and 2b (28% ip incorporation) as<br />

catalysts in comparison with the parent Ru-MOF. Blank stands for the experiment where on<br />

catalyst was added into the reaction mixture.


106 Chapter 4<br />

4.2.6 Conclusions<br />

Applying the solid solution approach, a range of Ru-DEMOFs ([Ru3(BTC)2-x(5-X-ip)xYy]n)<br />

isostructural to HKUST-1 were obtained with the framework incorporated DLs 5-OH-ip<br />

(1a-d), ip (2a and 2b), 5-NH2-ip (3a-c) and 5-Br-ip (4a)). Comparably high SBET have been<br />

measured for Ru-DEMOF samples (947-1302 m 2 /g) when considering the parent Ru-MOF<br />

material [Ru3(BTC)2Y1.5]n (704-998 m 2 /g). [82, 208] The highest level of DL incorporation<br />

appeared to be 37% (1d, 5-OH-ip). The data are consistent with two kinds of defects (A<br />

and B) being introduced to the Ru-DEMOFs depending on the nature of the functional<br />

groups (X) of the DLs 5-X-ip (see Figure 4.5). DLs with steric less demanding X-groups of<br />

probably still existing, but weaker coordinative binding properties as a carboxylate ligator<br />

are likely to favor type A defects, defined as modified paddlewheel nodes exhibiting<br />

reduced Ru δ+ -sites. Along with the increasing of the DL contents, defects of type B, i.e.<br />

missing node defects, can be created simultaneously in the DEMOFs. When the functional<br />

groups in the DL 5-X-ip are much smaller than carboxylate and non-coordinating X (such<br />

as H), the defects of type B, defined as missing node defects, apparently become more<br />

dominant even in case of low doping level (case of 2a and 2b). The more or less<br />

simultaneous presence of two kinds of defects leads to the synergetic and as well<br />

counteracting effects on the porosity, sorption and catalytic properties. In fact, Ru-DEMOF<br />

1c (32% 5-OH-ip incorporation), in which defects of type A are more dominant, reveals<br />

the highest BET surface area (1302 m 2 /g) among the all samples, including parent Ru-<br />

MOF and its DEMOF derivatives reported so far. [138] Concerning the enhanced gas sorption<br />

properties, type A defects are more important than those effects related to the defects of<br />

type B, as reduced and more accessible Ru δ+ -sites have been produced in defect type A.<br />

When it comes to the catalytic activity, Ru δ+ -sites in Ru-DEMOF 1c are suggested to be<br />

responsible for the significant increase of TOF value of ethylene dimerization, as clearly<br />

seen while comparing with the parent Ru-MOF and 1a, in which no or less Ru δ+ -sites are<br />

present. However, in case of Paal-Knorr pyrrole reaction, the possible steric hindrance<br />

between the functional groups (OH) at the DL and 2,5-hexadione substrate is suggested<br />

to be responsible for the reduced activity of the Ru-DEMOFs 1a-d, although they all<br />

feature enhanced conversion compared with the parent Ru-MOF. All in all, the described<br />

Ru-DEMOFs obtained via solid solution approach employing various DLs turned out to be<br />

rather complex materials to be studied and characterized. Nevertheless, the observed<br />

spectroscopic evidences as well as effects on sorption properties and catalytic


Chapter 4 107<br />

performance reasonably fit into a model of the two types of defects, modified and missing<br />

nodes (see Figure 4.5), respectively.


108 Chapter 4<br />

4.3 Defects Engineering in [Cu3(BTC)2]n: Effect of the synthetic parameters<br />

To date, the introduction of defects into the [M3(BTC)2]n structure has been achieved via<br />

mixed-linker solid solutions approach for frameworks where M = Cu or Ru. [136, 138-140] By<br />

varying the DLs, simultaneously mixed with H3BTC during the synthesis, DEMOFs with a<br />

general formula [M3(BTC)2-x(DL)x]n could be obtained. Interestingly, in spite of the very<br />

similar structural long range order of the DEMOFs, local point defects can influence the<br />

metal environment at the PW and also enhance structural properties such as porosity.<br />

Different types of defect A and B in Ru-DEMOFs as well as their influence on the sorption<br />

and catalytic properties have already been described in Chapter 4.2. More interestingly,<br />

incorporation of pydc into [M3(BTC)2]n, the generation of reduced metal sites was found<br />

in both the formed Ru-DEMOFs ([Ru3(BTC)2-x(pydc)xYy]n) [138] and Cu-DEMOFs<br />

([Cu3(BTC)2-x(pydc)x]n) [139] However, only in the former case, CO2 → CO dissociative<br />

chemisorption (“reduction”) at low temperature (90 K) under UHV conditions was<br />

achieved. Utilizing 5-OH-ip as DL during the MOF synthesis, the mesopores prefer to be<br />

generated in Cu-DEMOFs rather than in Ru-DEMOFs. Moreover, even the modification of<br />

the doping level of the same DL in Cu-DEMOF can control the formation of the framework<br />

with micropores or mesopores. [139] Recently, Hupp et al. reported the formation of<br />

missing Cu 2+ node defects in the obtained Cu-DEMOF with ip incorporation, [140] while<br />

defect type A (modified PW units) related Cu + formation was observed in the Cu-DEMOF<br />

with pydc or other 5-X-ip(X = OH, CN, NO2) incorporation. [139] With regard to these<br />

accounts, it is essential to understand the structural complexity of DEMOFs, which will be<br />

beneficial to tailor their advanced properties. Therefore, in comparison with rather<br />

complicated Ru-(DE)MOFs system and following the description of Cu-DEMOFs reported<br />

by Hupp et al. and Fang et al., the relative simple [Cu3(BTC)2]n has been chosen as a<br />

candidate here to be incorporated by ip DL under different synthetic conditions (e.g.<br />

solvent, nature of the metal salts). Systematically study on the formation of M-DEMOFs<br />

variants of [M3(BTC)2]n structure, namely, types of local defects (i.e. the generation of<br />

reduced metal-sites or not), their origin as well as the dependence on variation of the<br />

synthetic conditions will be investigated.


Intensity, a. u.<br />

Chapter 4 109<br />

4.3.1 Preparation and characterization of Cu-DEMOF samples [Cu3(BTC)2-x(ip)x]n<br />

All samples D1-8 have been prepared via mixed-linker soild solution approach with<br />

distinct relative H3BTC : H2ip molar ratios under solvothermal conditions (80 °C, 20 h).<br />

For each sample, various Cu-salts and solvents (DMF or EtOH) have been employed.<br />

Samples D5 and D6 are prepared following reported method. [140]<br />

4.3.1.1 Solids obtained using Cu(BF4)2∙6H2O as metal precursor<br />

D4<br />

D3<br />

D2<br />

D1<br />

Cu-BTC_sim.<br />

5 10 15 20 25 30<br />

2, degree<br />

Figure 4.40. PXRD patterns of the activated Cu-DEMOF samples D1-4 in comparison with the<br />

patterns simulated from the single-crystal XRD data of the reported [Cu 3 (BTC) 2 ] n (Cu-BTC_sim).<br />

Metal source: Cu(BF 4 ) 2 6H 2 O. Solvent used for synthesis of D1 and D2 (blue patterns): DMF; for<br />

D3 and D4 (red patterns): EtOH.<br />

PXRD patterns of all activated samples D1-4 (Figure 4.40) match well with the simulated<br />

patterns of [Cu3(BTC)2]n (Cu-BTC_sim), suggesting that they all are all phase-pure,<br />

crystalline and isostructural with the non-doped parent Cu-BTC. To note, the small double<br />

reflection of the as-synthesized sample D4_as at around 9.4°and reflections at 5.7°for the


Intensity, a. u.<br />

110 Chapter 4<br />

as-synthesized samples D2-D4_as (Figure 4.41) disappear after samples activation (i.e.<br />

heating under vacuum). Thus, the origin of these reflections might come from coordinated<br />

solvent molecules, which under employed activation conditions are subsequently<br />

removed. FT-IR spectra of all activated Cu-DEMOFs samples do not vary from the FT-IR<br />

spectrum of the parent Cu-BTC, displaying also overlapping of the characteristic bands of<br />

ip and BTC (Figure 4.42). However, the typical νs(COO) and νas(COO) bands of<br />

carboxylates are observed at around 1442 cm -1 and 1366 cm -1 , respectively. Moreover, the<br />

absence of [BF4] - (from the used Cu-salts) is confirmed, as no ν(B–F) band was found at<br />

1073 cm –1 . [208]<br />

D4_as<br />

D3_as<br />

D2_as<br />

D1_as<br />

Cu-BTC_sim.<br />

5 10 15 20 25 30<br />

2, degree<br />

Figure 4.41. PXRD patterns of the as-synthesized Cu-DEMOF samples D1-4_as in comparison<br />

with the patterns simulated from the single-crystal XRD data of the reported [Cu 3 (BTC) 2 ] n (Cu-<br />

BTC_sim). Metal source: Cu(BF 4 ) 2 6H 2 O. Solvent used for synthesis of D1 and D2 (blue patterns):<br />

DMF; for D3 and D4 (red patterns): EtOH.


Chapter 4 111<br />

H 3<br />

BTC<br />

H 2<br />

ip<br />

D4<br />

D3<br />

D2<br />

D1<br />

Cu-BTC<br />

4000 3500 3000 2500 2000 1500 1000 500<br />

wavenumber, cm -1<br />

H 3<br />

BTC<br />

H 2<br />

ip<br />

D4<br />

D3<br />

D2<br />

D1<br />

Cu-BTC<br />

2000 1500 1000 500<br />

wavenumber, cm -1<br />

Figure 4.42. IR spectra of the activated samples D1-D4 in comparison with the spectra of the<br />

parent Cu-BTC, H 3 BTC and H 2 ip. Bottom figure represents selected region from 2000 cm -1 to 350<br />

cm -1 . Metal source: Cu(BF 4 ) 2 6H 2 O. Solvent used for synthesis of the D1 and D2 (blue): DMF; for<br />

the D3 and D4 (red): EtOH.


weight loss, %<br />

112 Chapter 4<br />

100<br />

Cu-BTC<br />

80<br />

60<br />

D3<br />

40<br />

D1<br />

D2<br />

D4<br />

20<br />

100 200 300 400 500 600<br />

T, C<br />

Figure 4.43. TG curves of the prepared Cu-DEMOFs D1-4 in comparison with the parent Cu-BTC.<br />

Metal source: Cu(BF 4 ) 2 6H 2 O. Solvent used for synthesis of the D1 and D2 (blue): DMF; for the D3<br />

and D4 (red): EtOH.<br />

The thermal stability of the all obtained Cu-DEMOF samples D1-4 (Figure 4.43) is<br />

considerable preserved (decomposition temperature equals ca. 300 °C). A little decrease<br />

of thermal stability in comparison with the parent Cu-BTC (decomposition temperature<br />

is ca. 330 °C) might be attributed to the incorporation of ip DL into the MOF structures,<br />

where formation of the defects at Cu2-PW is expected (Figure 4.5). For example, Cu-COO<br />

bonds can be partially substituted by the weak interactions between benzene ring of ip<br />

and Cu-centers in case of defects of type A (modified paddlewheel). In other case (defects<br />

of type B), missing of complete nodes of the Cu2-PWs is probably to happen as well.<br />

Consequently, presence of such defects (either A, B, or both of them) may cause the slight<br />

decrease of the thermal stability of the final DEMOFs.<br />

4.3.1.2 Cu-DEMOFs obtained using Cu(NO3)2·3H2O as metal precursor<br />

The phase purity and crystallinity of all samples D5-8 have been confirmed by the PXRD<br />

analysis of both as-synthesized and activated solids (Figures 4.44 and 4.45). Besides,


Intensity, a. u.<br />

Intensity, a.u.<br />

Chapter 4 113<br />

D8_as<br />

D7_as<br />

D6_as<br />

D5_as<br />

Cu-BTC_sim.<br />

5 10 15 20 25 30<br />

2, degree<br />

Figure 4.44. PXRD patterns of the as-synthesized Cu-DEMOF samples D5-8_as in comparison<br />

with the patterns simulated from the single-crystal XRD data of the reported [Cu 3 (BTC) 2 ] n (Cu-<br />

BTC_sim). Metal source: Cu(NO 3 ) 2·3H 2 O. Solvent employed for the synthesis of the D5 and D6<br />

(olive patterns): DMF plus HBF 4 ; for D7 and D8 (magenta patterns): only DMF.<br />

D8<br />

D7<br />

D6<br />

D5<br />

Cu-BTC_sim.<br />

5 10 15 20 25<br />

2, degree<br />

Figure 4.45. PXRD patterns of the activated Cu-DEMOF samples D5-8 in comparison with the<br />

patterns simulated from the single-crystal XRD data of the reported [Cu 3 (BTC) 2 ] n (Cu-BTC_sim).<br />

Metal source: Cu(NO 3 ) 2·3H 2 O.


weight loss, %<br />

114 Chapter 4<br />

recorded PXRD patterns are in a good agreement with the patterns simulated from the singlecrystal<br />

XRD data of the reported [Cu 3 (BTC) 2 ] n , suggesting all prepared Cu-MOFs being<br />

isostructural analogs of the parent Cu-BTC. These observations indicate that Cu-DEMOFs with ip<br />

DL incorporation can be prepared employing various Cu-precursors. However, only in the case of<br />

using DMF as a reaction medium, phase-pure solids have been obtained. Similarly, synthesis of the<br />

D7 and D8 in EtOH instead of DMF yielded precipitates with additional unknown phases (Figure<br />

7.23).<br />

Like in the case of Cu-DEMOFs D1-4, IR spectra of the samples (D5-8) obtained from the<br />

reaction of Cu(NO3)2·3H2O and mixed linkers do not differ from the parent Cu-BTC (Figure<br />

4.46). The presence of the residues from the starting metal precursors can be ruled out,<br />

as no bands of [NO3] - (845-815 cm -1 ) is observed in the spectra of the prepared DEMOF<br />

solids, which supports the conclusions made above on their phase purity. TGA curves of<br />

all samples D5-8 show a slight decrease of thermal stability (Figure 4.47), in analogy to<br />

the other four Cu-DEMOFs D1-4 obtained using Cu(BF4)2 6H2O as metal precursor,<br />

suggesting no significant influence of the metal source on the thermal stability of the<br />

resulting DEMOFs.<br />

100<br />

80<br />

D8<br />

Cu-BTC<br />

60<br />

D6<br />

40<br />

D5<br />

D7<br />

20<br />

100 200 300 400 500 600<br />

T, C<br />

Figure 4.46. TG curves of the Cu-DEMOFs D5-8 in comparison with the parent Cu-BTC. Metal<br />

source: Cu(NO 3 ) 2·3H 2 O. Solvent used for preparation of the D5 and D6 (olive): DMF plus HBF 4 ; for<br />

D7 and D8 (magenta): only DMF.


Chapter 4 115<br />

H 3<br />

BTC<br />

H 2<br />

ip<br />

D8<br />

D7<br />

D6<br />

D5<br />

Cu-BTC<br />

4000 3500 3000 2500 2000 1500 1000 500<br />

wavenumber, cm -1<br />

H 3<br />

BTC<br />

H 2<br />

ip<br />

D8<br />

D7<br />

D6<br />

D5<br />

Cu-BTC<br />

2000 1750 1500 1250 1000 750 500<br />

wavenumber, cm -1<br />

Figure 4.47. IR spectra of the activated samples D5-D8 in comparison with the parent Cu-BTC as<br />

well as used linkers H 3 BTC and H 2 ip. Right figure represents selected region from 2000 cm -1 to<br />

350 cm -1 . Metal source: Cu(NO 3 ) 2·3H 2 O. Solvent used for preparation of the D5 and D6 (olive):<br />

DMF plus HBF 4 ; for D7 and D8 (magenta): only DMF.


116 Chapter 4<br />

4.3.2 Composition and porosity of the prepared Cu-DEMOFs<br />

In order to determine the presence and amount of the framework incorporated DL ip in<br />

the Cu-DEMOF solids, 1 H-NMR spectroscopic measurements have been performed on all<br />

activated samples (which were previously acid-digested). Apart from the resonance at ca.<br />

8.6 ppm, which originates from the aromatic protons of BTC, three additional peaks<br />

appear at ca. 7.6, 8.1 and 8.4 ppm in all Cu-DEMOF samples D1-8 (Figures 7.15-7.22),<br />

which could be attributed to the aromatic protons of ip DL. The ratio of incorporated ip to<br />

the total linkers-amount (BTC+ ip) in the obtained solids have been estimated from the<br />

integrals of the proton resonances in 1 H-NMR spectra and are listed in Table 4.9.<br />

Table 4.9. The incorporation of ip (calculated from the integration of proton-signals in 1 H-NMR<br />

spectra) and BET surface area (estimated from the N 2 sorption experiments (at 77 K)) of the<br />

obtained Cu-DEMOFs.<br />

Samples<br />

Metal source<br />

Synthesis<br />

solvent<br />

Molar ratio of ip : (BTC + ip)<br />

feeding<br />

incorporation<br />

(obtained)<br />

S BET (m 2 /g)<br />

D1 Cu(BF 4 ) 2·6H 2 O DMF 1/2 20% 1833<br />

D2 Cu(BF 4 ) 2·6H 2 O DMF 2/3 25% 2030<br />

D3 Cu(BF 4 ) 2·6H 2 O EtOH 1/2 12% 1931<br />

D4 Cu(BF 4 ) 2·6H 2 O EtOH 2/3 17% 1973<br />

D5 Cu(NO 3 ) 2·3H 2 O DMF+HBF 4 1/2 23%, 28%* 1841, -*<br />

D6 Cu(NO 3 ) 2·3H 2 O DMF+HBF 4 2/3 33%, 33%* 1305, 2030*<br />

D7 Cu(NO 3 ) 2·3H 2 O DMF 1/2 21% 1818<br />

D8 Cu(NO 3 ) 2·3H 2 O DMF 2/3 30% 1673<br />

* reported values in the literature. [140]<br />

On the whole, it might be concluded, that both metal salts that have been tested in present<br />

studies could be successfully utilized to introduce ip DL into Cu-BTC. Curiously, addition<br />

of HBF4 increases the amount of the incorporated ip (in about 2-3%) when Cu(NO3)2·3H2O<br />

is utilized. However, on the basis of the performed quantitative calculations and taking<br />

into account an error of integrals of the 1 H-NMR signals, this difference might be ignored.<br />

Hence, the incorporation amount of ip in the reproduced sample D5 can be also<br />

considered to be same as the reported values (Table 4.9). [140] In the case of usage<br />

Cu(BF4)2·6H2O, the reaction in DMF yields to higher extent of ip incorporation (about 8%


V, cm 3 /g<br />

Chapter 4 117<br />

more) in comparison with syntheses performed in EtOH. Thus, the protic EtOH solvent<br />

probably slows down to a certain degree the BTC↔ip substitution rate. Therefore, it is<br />

not surprising that the above-mentioned samples prepared from Cu(NO3)2·3H2O using<br />

EtOH instead of DMF are not phase-pure solids (Figure 7.23).<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

/ Cu-BTC<br />

/ D1<br />

/ D2<br />

/ D3<br />

/ D4<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

relative pressure, p/p 0<br />

Figure 4.48. N 2 sorption isotherms collected at 77 K for the Cu-DEMOF samples D1-4 in<br />

comparison with the parent Cu-BTC. Closed and opened symbols represent the adsorption and<br />

desorption isotherms, respectively. Black circles - Cu-BTC; blue triangles - D1; blue stars - D2; red<br />

diamonds - D3; red squares - D4. Metal source: Cu(BF 4 ) 2 ·6H 2 O. Solvent employed for the synthesis<br />

of the D1 and D2 (blue): DMF; for D3 and D4 (red): EtOH.<br />

All N2 sorption isotherms recorded at 77 K for Cu-DEMOFs D1-8 as well as the parent Cu-<br />

BTC show type I isotherms without any hysteresis loop (Figures 4.48, 4.49), suggesting<br />

their microporosity. To recall and opposed to these observations, in the Cu-DEMOFs with<br />

pydc or 5-X-ip DLs (X = NO2, CN, or OH), the generation of mesopores were reported. [139]<br />

This different behavior indicates, that the defects in Cu-DEMOFs prepared in current work<br />

tend to be isolated rather than correlated as in earlier reported homologous frameworks<br />

(Figure 4.17). Interestingly, all Cu-DEMOFs synthesized from Cu(BF4)2·6H2O (D1-4) show


V, cm 3 /g<br />

118 Chapter 4<br />

500<br />

400<br />

300<br />

200<br />

100<br />

/ Cu-BTC<br />

/ D5<br />

/ D6<br />

/ D7<br />

/ D8<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

relative pressure, p/p 0<br />

Figure 4.49. N 2 sorption isotherms collected at 77 K for the Cu-DEMOF samples D5-8 in<br />

comparison with the parent Cu-BTC. Closed and opened symbols represent the adsorption and<br />

desorption isotherms, respectively. Black circles - Cu-BTC; olive triangles - D5; olive stars - D6;<br />

magenta diamonds - D7; magenta squares - D8. Metal source: Cu(NO 3 ) 2 3H 2 O. Solvent utilized for<br />

the synthesis of the D5 and D6 (olive): DMF +HBF 4 ; for D7 and D8 (magenta): DMF.<br />

increased SBET in comparison with the parent Cu-BTC (1561 m 2 /g) (Table 4.9). In fact, the<br />

porosity of these samples rises along with the increase of incorporated amount of ip, no<br />

matter whether DMF or EtOH has been used. Thus, enhanced porosity provides indirect<br />

indications on absence of unreacted H2ip and proves the in-framework incorporation of<br />

ip in these solids. Remarkably, Cu-DEMOF sample D2 (25% of ip incorporation) reveals<br />

the highest SBET (2030 m 2 /g) among all the samples. Furthermore, solids D6, repeated<br />

following communicated earlier method, show a little lower SBET than the reported values<br />

(2030 m 2 /g). [140] Unfortunately, there is no exact reported SBET value for D5. Only a<br />

general increase trend of SBET value along with the incorporation amount of ip was given<br />

in the earlier report. However, in current study, incorporation of higher quantity of ip, as<br />

in the case of sample D6, leads to the decreased porosity instead in comparison with D5.<br />

Similar trend has been traced for the solids D7 and D8, where the latter sample reveals a<br />

little lower SBET value than the former. Consequently, it might be possible that some guest


Chapter 4 119<br />

(e.g. H2ip, H3BTC, etc.) inclusion happens in the case of D6 and D8. Thus, the highest<br />

incorporation degree of ip in Cu-BTC should be around 25% (D2 or D5). Although utilizing<br />

Cu(NO3)2·3H2O as metal-precursor, the samples with the same incorporation level of ip<br />

could be obtained (D5, 23% ip), the usage of nitrate salt results solids with lower porosity.<br />

Attempts to incorporate even more ip most probably would lead to the guest inclusion in<br />

the pores of DEMOFs. Due to less nucleophilic and basic character of “inert” [BF4] - anion<br />

in comparison with [NO3] - , employing Cu(BF4)2·6H2O probably facilitates the selfassembly<br />

between the metal clusters and the linkers during synthesis and could avoid the<br />

guest inclusion<br />

4.3.3 Oxidation state(s) of metal-sites in prepared Cu-DEMOFs<br />

In the earlier report from Hupp and co-workers, the generation of missing Cu 2+ -nodes was<br />

proposed. [140] In order to gain a deeper understanding of the defects generated in Cu-<br />

DEMOFs under different synthetic conditions, D1 and D2 samples, which are synthesized<br />

from different Cu-salts (Cu(BF4)2·6H2O) and feature high porosity as well as distinct<br />

incorporation level of ip, have been chosen as candidates for CO probing monitored by<br />

UHV-IR spectroscopy. It is interesting to investigate if there is any presence of Cu + in the<br />

obtained solids. Thus, the bands observed at 2178 cm -1 for D1 and D2 appear at the same<br />

position as in the spectra of the parent Cu-BTC and correspond to CO bound to Cu 2+ -sites<br />

in the PW through electrostatic and σ-donation interactions (Figure 4.50). [246] Besides,<br />

additional intense bands at 2120 cm -1 due to C-O stretching vibrations that are associated<br />

with Cu + - species are seen in spectra of both D1 and D2. The appearance of Cu + - species<br />

in sample D1 and D2 can origin from several possibilities:<br />

i) “intrinsic” defects in the framework of Cu-BTC and its analogs. It is known that intrinsic<br />

defects can be present to some extent in the MOF solids. [137, 247-249] Especially, intrinsic<br />

defects in HKUST-1 has been described. [249-250] . The activation procedure (i.e. thermal<br />

treatment under vacuum) lead to some decarboxylation and “missing” COO-sites, [249]<br />

therefore, the generation of Cu + was observed. It is possible that during the thermal<br />

treatment of D1 and D2, such intrinsic defects are formed as well.<br />

ii) “intentional” defects


120 Chapter 4<br />

Figure 4.50. From left to right: UHV-IR spectra of the parent Cu-BTC and Cu-DEMOFs D1 (20% of<br />

ip) and D2 (25% of ip) upon CO dosing (p(CO): 1×10 -6 - 3×10 -4 mbar) collected at 92-98 K after<br />

annealing at 480 K. The low intensity absorption bands at 2127 cm -1 in the parent Cu-BTC indicate<br />

some inherent defects that cannot be avoided during the crystal formation. Figure has been<br />

provided by M. Kauer.<br />

As illustrated in Figure 4.5, two representative types of defects are expected in the<br />

formation of DEMOFs. The generation of defect type A (i.e. modified PWs with the creation<br />

of the reduced Cu-sites) can be also expected in the solids D1 and D2 due to the utilization<br />

of reducing agents (DMF). Although Cu + related CO bands (2127 cm -1 ) appear in parent<br />

HKUST-1, the relative intensity to Cu 2+ related CO bands (2178 cm -1 ) is lower. However,<br />

in both D1 and D2 the CO bands at 2120 cm -1 attributed to Cu + -CO interaction turn to be<br />

dominant in comparison with the bands at 2178 cm -1 . Thus it gives us a hint that presence<br />

of defect type A can be one of the possibilities. Notably, intensity of both Cu 2+ - (2178 cm -<br />

1 ) and Cu + -related bands (2120cm -1 ) decreases from D1 (20% ip) to D2 (25%). The<br />

similar scenario has been also found for the 5-OH-ip incorporated Ru-DEMOF samples<br />

described in Chapter 4.2, suggesting thus simultaneous formation of defects B upon higher<br />

doping of DL, namely missing Cu-PWs. Consequently, lower abundance of Cu 2+ and Cu +<br />

metal centers in the D2 frameworks expected, explaining observed changes of the<br />

decreased intensity of the bands related to Cu 2+ and Cu + . Cu-DEMOFs with ip (obtained<br />

also from Cu(NO3)2·3H2O) reported earlier by Hupp et al. suggest generation of only<br />

defects type B by simulation of defects. Thus, it might be possible that anions of the<br />

utilized metal precursors play an important role on intentional defects formation in Cu-


Chapter 4 121<br />

DEMOFs. Moreover, as the generation of only defect type B was observed in Ru-DEMOF<br />

analogs of HKUST-1 with ip DL, the effect from different metal ions (Cu, Ru) on the<br />

formation of intentional defects in M-DEMOFs should also be taken into account.<br />

4.3.4 Conclusions<br />

Adopting Cu(NO3)2∙3H2O and Cu(BF4)2∙6H2O as metal sources and different feeding of<br />

H2ip DL results in crystalline Cu-DEMOFs, which are isostructural analogs of Cu-BTC. All<br />

obtained Cu-DEMOF solids show good thermal stability, which appeared to be not affected<br />

by the employed reactants (i.e. metal salts and solvents). Based on the composition and<br />

porosity, DMF found to be a better reaction solvent than EtOH leading to higher<br />

incorporation degree of ip and higher porosity of the resulting DEMOFs (for the given<br />

feeding). However, when the feeding of ip is increased, the usage of Cu(NO3)2∙3H2O as<br />

precursor should be carefully monitored owing to accompanied effects of guest(s)<br />

inclusion. Nevertheless, current studies demonstrate the highest ip framework<br />

incorporation into Cu-BTC could reach up to 25%.<br />

More importantly, UHV-IR spectra recorded upon CO dosing suggest generation of Cu + in<br />

the prepared Cu-DEMOF samples D1 and D2. Two important caused factors are<br />

concluded: one is related with the intrinsic defects induced by thermal activation and the<br />

other is intentional defects (i.e. modified PWs). In the case of intentional defects, usage of<br />

reducing agents, modifying the metal precursors including the cations and anions could<br />

be the driven force to generate reduced metal-sites. In one word, even in the “simple<br />

looking” HKUST-1 (in comparison with its Ru analogs), the defect-engineering is<br />

complicated and carefully consideration is required for the elaboration.


5 Simultaneous introduction of various palladium active sites<br />

into MOF via one-pot synthesis: Pd@[Cu3-xPdx(BTC)2]n ‡<br />

Abstract<br />

Simultaneous structural incorporation of palladium within Pd-Pd and/or Pd-Cu<br />

paddlewheels as framework-nodes and Pd nanoparticles (NPs) dispersion into MOF have<br />

been achieved for the first time via one-pot synthesis. The substitution of Cu 2+ by Pd 2+<br />

within polymer carcass as well as the loading of Pd NPs have been confirmed in particular<br />

by XPS. Solids featuring such both Pd-sites show enhanced activity (compared to the<br />

parent “Pd-free” Cu-BTC) in the hydrogenation of p-nitrophenol (PNP) to p-aminophenol<br />

(PAP) using NaBH4 as a reducing agent.<br />

‡ Work in this chapter is covered in the manuscript submitted to Dalton Trans.


Chapter 5 123<br />

5.1 Introduction on the selection of metal type in MOFs<br />

High structural and compositional design ability of metal-organic frameworks (MOFs)<br />

and, hence, the ability of tailoring their properties made MOFs to be among the most<br />

topical materials over last decades. Along with linker(s) modification [64, 123, 251] and<br />

guest(s) inclusion, [252-253] variation of the metal center(s) is also a powerful tool to fine<br />

tune MOFs functionalities. In fact, both control over coordinatively unsaturated metal<br />

sites (CUS) (e.g., via “defects-engineering”) [137] and partial metal substitution (e.g., via<br />

mixed-metal “solid solution” approach) could considerably enhance MOFs activity,<br />

especially in catalysis [138, 141, 254] and selective gas sorption/separation. [255-256]<br />

Figure 5.1. Partial metal (Cu vs Mn, Zn, Co) substitution in Cu-BTC. Reprinted with permission<br />

from D. F. Sava Gallis, M. V. Parkes, J. A. Greathouse, X. Zhang and T. M. Nenoff, Chem. Mater., 2015,<br />

27, 2018-2025. Copyright (2015) American Chemical Society. [256]<br />

Combination of distinct metal-ions, which are closely related in coordination chemistry<br />

and have similar effective ionic radii (rion), like Cu 2+ (73 pm) / Zn 2+ (74 pm) couple [257] for<br />

instance, within single MOF has been reported for several structural types. [131, 255, 258] In<br />

fact, partial substitution of Cu 2+ by Zn 2+ and some other metals of 3d-row, namely Co, Fe<br />

and Mn, in HKUST-1 ([Cu3(BTC)2]n, BTC = benzene-1,3,5-tricarboxylate) has been recently<br />

reported (Figure 5.1). [133, 256] Curiously, in later case doping with second metal caused<br />

enhanced selective sorption of O2 vs N2. [256, 259] What is more challenging however, is to<br />

take benefits of solid solution approach to introduce more distinct metals of 4d- [134] or 5drow<br />

[260] as framework-nodes. In particular, metals of the platinum group are highly<br />

attractive as catalytic- / strong sorption-centers. Besides, square planar coordination is


124 Chapter 5<br />

preferred like Cu 2+ . [261-263],[264] Still, due to kinetic reasons, these metal ions are difficult to<br />

be crystallized within 3D structures [263-264] and, therefore, almost neglected in MOF<br />

field. [265] Reports on MOFs featuring Pd-Pd, Pd-M PW nodes could not be even found.<br />

Figure 5.2. (a) Structural and (b) schematic representations of the synthesis of the cuboctahedral<br />

bimetallic MOPs. Reproduced from J. M. Teo, C. J. Coghlan, J. D. Evans, E. Tsivion, M. Head-Gordon,<br />

C. J. Sumby and C. J. Doonan, Chem. Commun., 2016, 52, 276-279, with permission of The Royal<br />

Society of Chemistry. [264]<br />

Recent years the study of Ru-based MOFs has been mainly focused on the synthesis of Ruanalogues<br />

of HKUST-1, which were successfully obtained in our group. [82, 198] Following<br />

the previous studies, in this work the attention was turned also to palladium, which is well<br />

known to facilitate dissociation of molecular hydrogen (spill-over effect). [266-268]<br />

Furthermore, bimetallic Pd/M-paddlewheel (PW) complexes have been lately found to be<br />

promising economical catalysts in the intramolecular benzylic C-H amination. [269]<br />

Interestingly, to date immobilization of Pd-nanoparticles [270-273] and decoration of MOFs<br />

interior with Pd-complexes [274-275] have been investigated a lot. Moreover, porous Pd 2+ -<br />

M 2+ (M = Cu, Ni, Zn) metal-organic polyhedra have been very recently described (Figure<br />

5.2). [264] Regardless obvious interest, studies on integration of palladium as a framework<br />

node into a MOF are, however, in its infancy. [265, 276-277] Hence, considering its structural<br />

peculiarities (i.e., PW SBUs and CUSs) [35] and primary experimentally studies on [Cu3-<br />

xRux(BTC)2]n including Ru 3+ (rion = 68 pm), [257] it would be of interest to choose<br />

[Cu3(BTC)2]n (Cu-BTC) as a matrix for in-framework incorporation of Pd 2+ (rion = 86


Chapter 5 125<br />

pm) [257] via solid solution approach. However, a rigorous exclusion of reducing conditions<br />

during MOF synthesis is difficult. Even technical activation may cause reduction at some<br />

metal sites due to decarboxylation. [250] Therefore, the standard protocol for HKUST-1<br />

synthesis is, intentionally, not changed in this stage.<br />

On the other hand, loading of Pd 0 NPs onto MOF is widely investigated as gas<br />

adsorbents [273, 278-279] and catalysts in hydrogenation, [271-273, 280-281] cross coupling [272, 282]<br />

reaction and so on. [283] To note, one of the most common approaches to obtain Pd 0 @MOFs<br />

is using H2 to reduce Pd 2+ -precursor loading MOFs that are prepared via solution<br />

impregnation by soluble Pd 2+ species. In some cases, trace amount of Pd 2+ can be still<br />

observed after reduction. [282] Actually, according to the concept of post-synthetic metalion<br />

exchange, [117, 121, 284-285] Pd 2+ substitution in the metal nodes of MOFs could probably<br />

happen under this impregnation. Interestingly, the discrimination between the two Pdsites<br />

(from extra-framework loading or in-framework incorporation) is disregarded to<br />

some extent in the current literatures. Given both facts, herein an appropriate one-pot<br />

synthesis has been selected to study MOFs with simultaneous introduction of both Pd 2+ -<br />

nodes and Pd 0 NPs in one step and test their catalytic activity.


126 Chapter 5<br />

5.2 Preparation and Structure of the Cu/Pd-BTC_1-3<br />

Pd@[Cu3-xPdx(BTC)2]n (Cu/Pd-BTC_1-3) materials have been synthesized under<br />

solvothermal conditions by mixing corresponding metal salts (Pd(OOCCH3)2 and<br />

Cu(NO3)2·3H2O) with H3BTC directly in the starting reactions (see Chapter 7 on<br />

experimental details). All Cu/Pd-BTC_1-3 samples (both as-synthesized and dried<br />

phases) are crystalline solids and isostructural with the parent Cu-BTC (Figures 5.3 and<br />

5.4), indicating the preserved structure integrity after the Pd-sites introduction. Recorded<br />

Figure 5.3. PXRD patterns of the activated Pd@[Cu 3-X Pd x (BTC) 2 ] n (Cu/Pd-BTC_1-3) in<br />

comparison with the activated non-doped Cu-BTC. The vertical dash lines correspond to the peak<br />

position of the (111) planes of the face-centered cubic (fcc) Pd-structure.<br />

PXRD patterns and accordingly calculated cell parameters of Cu/Pd-BTC_1-3 match very<br />

well with the respective simulated data of the reported [Cu3(BTC)2]n single-crystal<br />

structure [35] (Figure 5.5, Table 5.1). Remarkably, intensity of the reflections at ca. 6.67,<br />

9.40 and 13.34° (2θ) assigned to the (200), (220) and (400) planes, respectively, varies<br />

(Figure 5.3). This indicates certain changes of electronic density compared to the nondoped<br />

Cu-analogue (Cu-BTC). The main contribution to the reflection of the (200) and<br />

(220) planes is due to the metal-nodes (Figure 5.6). Hence, such difference in intensities<br />

should primary stem from the partial substitution of Cu 2+ by Pd 2+ within the M2-<br />

paddlewheel units of MOFs. Thus, it indicates Pd 2+ framework incorporation. Considering<br />

the presence of Pd NPs in the discussed Cu/Pd-BTC_1-3, small reflections at 40.02°<br />

resulting from (111) planes of the face-centered cubic structure of Pd [286] are observed


Chapter 5 127<br />

(Figure 5.3). This is probably due to the rather small particle size of these Pd NPs and,<br />

therefore, hard to be detected by PXRD technique. Finally, it should be pointed out, that<br />

employing equal molar amounts of Cu- and Pd-salts or only Pd-precursor in the starting<br />

reaction mixtures, formation of only metal/metal-oxide products has been observed<br />

(Figures 7.25 and 7.26).<br />

Figure 5.4. PXRD patterns of the as-synthesized Pd@[Cu 3-X Pd x (BTC) 2 ] n in comparison with<br />

simulated patterns of the non-doped Cu-BTC.<br />

Table 5.1. Calculated cell parameters of the Cu/Pd-BTC_1-3 (TOPAS academic software and<br />

Pawley method [287] have been employed) in comparison with that obtained from the single crystal<br />

data of the reported [Cu 3 (BTC) 2 ] n . [35]<br />

Sample a (Å) V (Å 3 ) Space group R wp R exp<br />

[Cu 3 (BTC) 2 ] [35] n 26.343 18280 Fm-3m - -<br />

Cu/Pd-BTC_1 26.308 18208 Fm-3m 6.018 4.79<br />

Cu/Pd-BTC_2 26.305 18203 Fm-3m 6.618 5.17<br />

Cu/Pd-BTC_3 26.293 18178 Fm-3m 7.201 6.09


128 Chapter 5<br />

Figure 5.5. Pawley Fits of the samples Cu/Pd-BTC_1-3. Blue, olive and magenta lines represent<br />

the fitted patterns of Cu/Pd-BTC_1, 2 and 3, respectively. Black ticks - theoretical positions of the<br />

Bragg reflections; black crosses - the experimental data; violet lines - the difference between fitted<br />

and measured patterns.


Chapter 5 129<br />

Figure 5.6. Crystal structure of the Cu-HKSUT-1 ([Cu 3 (BTC) 2 ] n ) viewed along the plane (200) (a),<br />

plane (220) (b )and plane (222) (c).


130 Chapter 5<br />

5.3 Compositional characterization and sorption properties of the prepared<br />

Cu/Pd-BTC_1-3<br />

AAS analysis of the activated Cu/Pd-BTC_1-3 solids confirms quantitative incorporation<br />

of palladium. Thus, the Pd-contents and Pd : Cu ratios in the final Cu/Pd-BTC_1-3 samples<br />

remain almost unchanged with respect to the taken feeding ratios of the metal-precursors<br />

(Table 5.2). Further, SEM-EDX elemental mapping demonstrates quite homogeneous<br />

Cu/Pd distribution within obtained Cu/Pd-BTC solids (Figure 5.7).<br />

Table 5.2. Feeding and observed contents of Pd in respect to the total metal amount in Pd@[Cu 3-<br />

xPd x (BTC) 2 ] n (Cu/Pd-BTC_1-3). The incorporation Pd-percentages were calculated based on the<br />

AAS results.<br />

Sample feeding (molar%) obtained (molar%) obtained wt%<br />

Cu/Pd-BTC_1 10 9 3.9<br />

Cu/Pd-BTC_2 15 14 6.2<br />

Cu/Pd-BTC_3 20 20 9.2<br />

Figure 5.7. EDX elemental mapping of the Cu/Pd-BTC_3 solid (red – Cu; green – Pd).<br />

Incorporation of Pd 2+ into the framework as well as the Pd 0 NPs loading has been<br />

subsequently corroborated by high-resolution XPS. Figure 5.8 shows the Pd 3d core-level<br />

spectra of the Cu/Pd-BTC_1 and Cu/Pd-BTC_3 (representative samples with relatively<br />

low and high Pd-contents). Three Pd 3d doublets (3d5/2 and 3d3/2) are resolved in the<br />

deconvoluted spectra. The doublet located at 337.9 and 343.2 eV (Pd2) is characteristic<br />

for Pd 2+ species (Table 5.3). [288-289] In addition, the higher binding energies observed at<br />

338.9 and 344.2 eV for Pd1 reveal the presence of an electronically modified Pd 2+ species.<br />

Moreover, we could unambiguously rule out an assignment of both Pd 2+ species to PdO


Chapter 5 131<br />

NPs, as no typical O 1s peak at about 530 eV has been detected in the corresponding<br />

spectra, which showed only one O 1s peak at 531.8 eV originating from the carboxylate (-<br />

COO) groups in the frameworks. Therefore, both found Pd 2+ species (Pd1 and Pd2) very<br />

likely are associated with the formation of Pd-Pd and Cu-Pd framework-nodes. The<br />

doublet at ca. 335.8 and 341.0 eV is attributed to metallic Pd 0 species (Pd3), indicating the<br />

existence of Pd metallic NPs in these samples. Besides, the amount of Pd 0 NPs in the total<br />

Pd species (Pd 2+ and Pd 0 ) increases (from 33 % for Cu/Pd-BTC_1 to 44 % for Cu/Pd-<br />

BTC_3) along with the doping increase of Pd(II) acetate. TEM characterization has not<br />

been performed because of the beam sensitivity of HKUST-1, which does not allow<br />

characterization of the "pristine" Pd@[Cu3-xPdx(BTC)2]n sample as well.<br />

Figure 5.8. Pd 3d region of the deconvoluted XP spectra of the activated solids Cu/Pd-BTC_1 and<br />

3. Figure is provided by Dr. Y. Wang.


132 Chapter 5<br />

Table 5.3. The assignment of binding energies as well as the determined molar fraction of Pd 2+<br />

and Pd 0 in Cu/Pd-BTC_1 and _3.<br />

Sample<br />

Binding energy, eV<br />

Molar fraction (%)<br />

Assignment<br />

Pd 3d 3/2 Pd 3d 5/2 Pd 2+ : (Pd 2+ +Pd 0 ) Pd 0 : (Pd 2+ +Pd 0 )<br />

Cu/Pd-BTC_1<br />

Cu/Pd-BTC_3<br />

344.2 338.9 Pd 2+ (Pd1)<br />

343.2 337.9 Pd 2+ (Pd2)<br />

340.9 335.6 Pd (Pd3)<br />

344.2 338.9 Pd 2+ (Pd1)<br />

343.2 337.9 Pd 2+ (Pd2)<br />

341.2 335.9 Pd (Pd3)<br />

67 33<br />

56 44<br />

Figure 5.9. IR spectra of the activated samples Cu/Pd-BTC_1-3 in comparison with the parent<br />

Cu-BTC as well as H 3 BTC and Pd-acetate.<br />

Furthermore, any vibrations stemming neither from the non-reacted H3BTC linker nor<br />

used metal-precursors (i.e., acetate-, nitrate-groups) could be revealed in the FT-IR


Chapter 5 133<br />

spectra of all prepared Cu/Pd-BTC solids (Figure 5.9). Moreover, 1 H-NMR spectra of the<br />

digested Cu/Pd-BTC_1-3 samples show presence of only BTC-linker (Figure 5.10). Thus,<br />

absence of the resonances at about 1.9 ppm stemming from the acetate protons fully<br />

supports IR results ruling out any residuals of the employed starting Pd-reactant. Hence,<br />

the presence of Pd 2+ -sites should originate from the in-framework nodes rather than Pd 2+ -<br />

precursor loading.<br />

Figure 5.10. 1 H-NMR spectra of the digested activated Cu/Pd-BTC_1-3 samples. The vertical dash<br />

line stands for the position where the peak of CH 3 -protons from the acetic acid commonly appears.<br />

TGA suggests that thermal stability of the obtained Cu/Pd-BTC_1-3 solids is fully<br />

preserved (in comparison with the parent Cu-BTC) [35] with the decomposition<br />

temperatures close to 300 °C (Figure 5.11). The N2 (77 K) sorption isotherms recorded<br />

for obtained samples reveal type I isotherm (Figure 5.12), confirming their permanent<br />

microporosity. Brunauer-Emmett-Teller (BET) surface area of Cu/Pd-BTC_1 and _2<br />

increase a little in comparison with parent Cu-BTC (prepared under similar conditions)<br />

(Table 5.4), indicating that the presence of Pd 2+ incorporation as metal nodes in the


134 Chapter 5<br />

framework are dominant. When it turns to Cu/Pd-BTC_3, the increasing of Pd 0 NPs<br />

species results in a slight lower BET surface area in comparison with the others samples.<br />

Figure 5.11. TG curves of the prepared Pd-doped solids Cu/Pd-BTC_1-3 in comparison with the<br />

“non-doped” material Cu-BTC.<br />

Table 5.4. BET surface area (S BET ) of Cu/Pd-BTC_1-3 in comparison with the “non-doped” Cu-<br />

BTC.<br />

Sample S BET (m 2 /g) S BET (m 2 /mmol)<br />

Cu-BTC 1561 944<br />

Cu/Pd-BTC_1 (9 mol% of Pd) 1594 982<br />

Cu/Pd-BTC_2 (14 mol% of Pd) 1750 1090<br />

Cu/Pd-BTC_3 (20 mol% of Pd) 1112 700


Chapter 5 135<br />

Figure 5.12. N 2 sorption isotherms collected at 77 K for the non-doped [Cu 3 (BTC) 2 ] n (Cu-BTC) and<br />

Pd-doped solids Pd@[Cu 3-x Pd x (BTC) 2 ] n (Cu/Pd-BTC_1-3).


136 Chapter 5<br />

5.4 Synthesis, compositional characterization and sorption properties of Cu/Pd-<br />

BTC_4<br />

Before the further investigation on catalytic reaction, it is important to prepare analogous<br />

samples featuring dominant extra-framework Pd 0 NPs loading as well. As known from the<br />

earlier experiments, employing higher Pd(II) acetate (more than 20%) in the starting<br />

reaction mixture, will only lead to the formation of metal/metal-oxide products.<br />

Interestingly, PdCl2 exhibits a rather distinct coordination mode with Pd(II) acetate. [290-<br />

292] Besides, longer reaction time in solvothermal conditions could facilitate the creation<br />

of Pd NPs to some extent. Thus, it is expected that higher Pd 0 NPs dispersion can be<br />

reached in Cu/Pd-BTC_4 by altering the Pd-precursor to PdCl2 and extending the reaction<br />

time.<br />

Figure 5.13. PXRD patterns of the as-synthesized (_as) and activated (_ht) Cu/Pd-BTC_4 in<br />

comparison with the patterns simulated from the single crystal XRD data of the reported<br />

[Cu 3 (BTC) 2 ] n (Cu-BTC_sim). [35]


Chapter 5 137<br />

The phase purity and good crystallinity of Cu/Pd-BTC_4 have been proved by the<br />

comparison of the PXRD patterns (Figures 5.13) before and after activation with those<br />

simulated from the single crystal XRD data of the reported [Cu3(BTC)2]n (Cu-BTC_sim). [35]<br />

Notably, the PXRD patterns of the Cu/Pd-BTC_4 sample display higher relative intensity<br />

of the (222) reflection, which is different with the intensity of that in the parent Cu-BTC.<br />

This suggests certain electronic modification on the metal nodes in the framework (e.g.<br />

Pd 2+ substitution of Cu 2+ -sites). However, the presence of Pd NPs could not be precisely<br />

determined by PXRD. No apparent reflections at 40.02 and 46.59 °originating from (111)<br />

and (200) planes of the face-centered cubic structure of Pd are found in the PXRD patterns<br />

of the Cu/Pd-BTC_4. [286] This can be due to the rather small particle size of NPs.<br />

Cu/Pd-BTC_4<br />

Cu-BTC<br />

H 3<br />

BTC<br />

4000 3500 3000 2500 2000 1500 1000 500<br />

wavenumber, cm -1<br />

Figure 5.14. IR spectra of the Cu/Pd-BTC_4 solid in comparison with the parent Cu-BTC and the<br />

starting linker H 3 BTC.<br />

In addition, no vibrations stemming from the non-reacted H3BTC linker or employed<br />

metal-precursors (i.e. nitrate groups) could be observed in the FT-IR spectra of the


138 Chapter 5<br />

obtained Pd-doped solid (Figure 5.14), indicating the absence of nitrate from the<br />

employed for the synthesis of Cu/Pd-BTC_4 Cu-salt. The absence of Cl - in the Cu/Pd-<br />

BTC_4 sample is hard to rule out based only on the IR and 1 H-NMR data. Further<br />

characterizations, namely SEM-EDX and XPS, will addresses this issue.<br />

Figure 5.15. SEM images of the Cu/Pd-BTC samples.<br />

Figure 5.16. EDX elemental mapping of the Cu/Pd-btc_4. Red: Cu-mapping; green: Pd-mapping.<br />

Curiously, SEM image of the sample Cu/Pd-BTC_4 displays that its particles have welldefined<br />

octahedral shape and are visibly bigger (ca. 20 μm) compared to the other solid<br />

(Cu/Pd-BTC_3) (Figure 5.15). Thus, variation of the Pd-precursor apparently influences<br />

extent of palladium inclusion and morphology of the Cu/Pd-MOF particles (as also<br />

reflected in the PXRD, Figures 5.13): Pd(II) acetate favors quantitative palladium<br />

incorporation (samples Cu/Pd-BTC_1-3), while PdCl2 somehow decelerates MOF growth,<br />

leading therefore to larger crystals (sample Cu/Pd-BTC_4). SEM-EDX mapping of the<br />

Cu/Pd-BTC_4 solid shows the homogeneous distribution of Pd and Cu in the particles of<br />

framework (Figure 5.16). In addition, the formation of some Pd species where there is no<br />

contribution of Cu has also been observed. These Pd species could be assigned to Pd-Pd<br />

PWs or just Pd NPs. Due to the beam stability of Cu/Pd-BTC under high electron voltage,


Chapter 5 139<br />

it is hard to distinguish these two possibility. To note, no Cl Kα peak could be observed in<br />

the EDX spectrum of the Cu/Pd-BTC_4 (Figures 5.17), suggesting absence (at least in high<br />

concentration) of Cl - or PdCl2.<br />

Figure 5.17. EDX spectrum of the sample Cu/Pd-btc_4 with the assignments of Cl Kα and Cl Kβ<br />

(green lines).<br />

Deconvoluted XP spectra from the Pd 3d region scan has been gathered for the sample<br />

Cu/Pd-BTC_4 (Figure 5.18). In addition to the Pd 3d doublet (Pd 3d3/2 and Pd 3d5/2) of<br />

Pd 2+ at 343.6 and 338.4 eV as well as 342.4 and 337.2 eV (Table 5.5), [288-289] another<br />

doublet appears at 340.9 and 335.7 eV, which can be attributed to Pd 0 in metallic<br />

palladium. Compared with Cu/Pd-BTC_1-3, this sample reveals dominant Pd 0 than in<br />

case of Cu/Pd-BTC_1-3. The concentration of Pd 0 is determined to amount to 70 %.<br />

Likewise, in Cu/Pd-BTC_1-3, only the O 1s peak at 531.8 eV stemming from COO- has<br />

been revealed. Hence, the presence of PdO can be ruled out due to the absence of<br />

corresponding O 1s peak at about 530 eV. Moreover, the peak of Cl 2p in the survey scan<br />

of Cu/Pd-BTC_4 cannot be observed (Figure 5.19), which together with SEM result<br />

proves that the Pd 2+ in the solid is attributed from Pd 2+ /M 2+ -node in the framework rather<br />

than Pd 2+ -precursor or PdO loading on the framework.


140 Chapter 5<br />

Figure 5.18. Deconvoluted XP spectra of the Cu/Pd-BTC_ 4 in the Pd 3d region. Figure is provided<br />

by Dr. Y. Wang.<br />

Figure 5.19. XPS survey scan of the activated sample Cu/Pd-BTC_4. Figure is provided by P. Guo.


weight loss, %<br />

Chapter 5 141<br />

Table 5.5. The assignment of binding energies as well as the determined molar fraction of Pd 2+<br />

and Pd 0 in Cu/Pd-BTC_4.<br />

sample<br />

Molar fraction (%)<br />

Binding energy, eV<br />

Assignment<br />

Pd 3d 3/2 Pd 3d 5/2 Pd 2+ : (Pd 2+ +Pd 0 ) Pd 0 : (Pd 2+ +Pd 0 )<br />

Cu/Pd-BTC_4<br />

343.6 338.4 Pd 2+ (Pd1)<br />

342.4 337.2 Pd 2+ (Pd2)<br />

340.9 335.7 Pd 0 (Pd3)<br />

30 70<br />

100<br />

90<br />

80<br />

Cu-BTC<br />

Cu/Pd-BTC_3<br />

Cu/Pd-BTC_4<br />

70<br />

60<br />

50<br />

40<br />

100 200 300 400 500 600<br />

T, C<br />

Figure 5.20. TG curves of the activated Cu/Pd-BTC_4 and _5 solids in comparison with the parent<br />

Cu-BTC and Cu/Pd-BTC_3.<br />

In order to gain information on the quantitative incorporation of palladium, AAS analysis<br />

of the activated Cu/Pd-BTC_4 solids has been performed. Thus, with respect to the taken<br />

feeding ratios of the metal-precursors (20%), amount of the incorporated Pd in Cu/Pd-<br />

BTC_4 sample is 12%, which is a little lower than that in Cu/Pd-BTC_3 (20%). TGA of the<br />

Cu/Pd-BTC_4 (Figure 5.20) demonstrates the preserved thermal stability after palladium<br />

introduction and show the framework decomposition after 300 °C. Interestingly, N2


V, cm 3 /g<br />

142 Chapter 5<br />

sorption isotherms recorded at 77K for Cu/Pd-BTC_4 sample show type I isotherms<br />

(Figure 5.21), suggesting the microporosity. This is in analogy to other Cu/Pd-BTC<br />

samples described above. The higher BET surface area of Cu/Pd-BTC_4 (Table 5.6) than<br />

Cu/Pd-BTC_1-3 points to external Pd-NPs rather than included particles into the volumes<br />

of the framework. Besides, from the above described better crystallinity revealed in PXRD<br />

patterns as well as the well-shaped of the particles in SEM image, it is not surprised that<br />

this sample prepared from PdCl2 exhibit higher SBET.<br />

500<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

/ Cu-BTC<br />

/ Cu/Pd-BTC_3<br />

/ Cu/Pd-BTC_4<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

relative pressure, p/p 0<br />

Figure 5.21. N 2 sorption isotherms collected at 77 K for the non-doped [Cu 3 (BTC) 2 ] n (Cu-BTC) and<br />

solids [Cu 3-x Pd x (BTC) 2 ] n (Cu/Pd-BTC_3-4).<br />

Table 5.6. BET surface areas calculated from N 2 sorption isotherms (77 K) of Cu/Pd-BTC samples<br />

in comparison with the parent Cu-BTC.<br />

Sample<br />

BET surface area (m 2 /g)<br />

Cu-BTC 1561<br />

Cu/Pd-BTC_3 1112<br />

Cu/Pd-BTC_4 2000


Chapter 5 143<br />

5.5 Catalytic test reaction<br />

Both Pd 2+ -containing MOF ([Pd(2-pymo)2]n, 2-pymo = 2-pyrimidinolate) [276-277, 293] and<br />

Pd 0 @MOF [272-273, 281-282] have been reported as good reusable catalysts in typical<br />

palladium-catalyzed reactions such as Suzuki C–C couplings and hydrogenation. Owing to<br />

both Pd 0 NPs and potential Pd 2+ -CUS (that should be available after thermal treatment<br />

and specifically promote H2 splitting) in obtained Pd-doped Cu/Pd-BTC materials, it was<br />

of interest to test their catalytic activity.<br />

Figure 5.22. Concentration of PNP as a function of time for non-doped Cu-BTC and Pd containing<br />

Cu/Pd-BTC_1-4 as catalysts in the aqueous-phase hydrogenation of PNP with NaBH 4 to PAP. The<br />

insert shows the initial reaction rate for Cu/Pd-BTC_1-4 normalized to the amount of Pd in the<br />

catalyst, respectively. Reaction conditions: c PNP = 0.18 mmol dm −3 , c NaBH4 = 0.60 mmol dm −3 , T =<br />

298 K, stirring speed = 1300 min −1 , m catalyst = 5.0 mg, ambient pressure. Figure is provided by Z.<br />

Chen.<br />

As a test reaction, the aqueous-phase hydrogenation of PNP to PAP using NaBH4 as a<br />

reducing agent has been chosen. It can be conducted at room temperature and ambient<br />

pressure within reaction times < 10 min using a simple on-line UV-Vis spectroscopy to<br />

monitor the progress of the reaction. [281, 294-295] The concentration of PNP as a function of


144 Chapter 5<br />

time as well as the initial reaction rate normalized to the amount of Pd in the catalyst for<br />

Cu/Pd-BTC_1-4 are depicted in Figure 5.22. Before addition of the solid catalyst, i.e.,<br />

within the first 0.5 min of the experiment, no PNP conversion was observed. Complete<br />

conversion of PNP to PAP was reached within 2 min of reaction time using Cu/Pd-BTC_1-<br />

4, whereas the conversion of PNP remained incomplete for the Pd-free Cu-BTC. This<br />

incomplete conversion is known for catalysts with low activity and results from<br />

unproductive NaBH4 decomposition to hydrogen. [294] Evidently, the Pd-containing Cu-<br />

BTC catalysts are significantly more active than the Pd-free Cu-BTC. Interestingly, the<br />

initial reaction rate related to the Pd amount for Cu/Pd-BTC_1 and 2 (i.e., 4.0×10 -3 and<br />

2.0×10 -3 dm -3 min -1 ) is significantly higher than that achieved over a metallic Pd catalyst<br />

supported on an Al-containing mesocellular silica foam (1.0×10 -3 dm -3 min -1 ) under the<br />

same reaction conditions. [296] Note that the activity of Cu/Pd-BTC_1 and 2 is so high that<br />

the difference in their catalytic activity cannot be distinguished under the reaction<br />

conditions used.<br />

Figure 5.23. PXRD patterns of the sample with 14% of Pd-doping after the hydrogenation of PNP<br />

to PAP (Cu/Pd-BTC_2_cat) in comparison with that one before the reaction (Cu/Pd-BTC_2). The<br />

diamond symbol represents the reflection position of Cu 2 O, which could be present due to the<br />

reduction by NaBH 4 .


Chapter 5 145<br />

Surprisingly, the normalized initial reaction rate decreases with increasing overall Pd<br />

content from 4×10 -3 dm -3 min -1 for Cu/Pd-BTC_1 to 1.0×10 -3 dm -3 min -1 for Cu/Pd-BTC_3,<br />

respectively (insert in Figure 5.22). This can be explained by considering the decreasing<br />

concentration of Pd 2+ species with increasing Pd content from Cu/Pd-BTC_1 to Cu/Pd-<br />

BTC_3 as observed by XPS (Table S3). Especially, Cu/Pd-BTC_4 containing the lowest<br />

concentration of Pd 2+ species (70%) display the lowest initial reaction rate in comparison<br />

with the other three Cu/Pd-BTC samples. It may, therefore, be concluded that Pd 2+<br />

species in the catalyst are catalytically active and that they are dominantly responsible for<br />

the observed high catalytic activity. Moreover, it has been reported that Pd 2+ can act as<br />

active metal sites in olefin hydrogenations. [276-277, 293] To note, PXRD patterns of the<br />

Cu/Pd-BTC samples before and after reaction indicate that the structure of the<br />

framework stays unchanged (Figure 5.23). Finally, according to the ICP-OES analysis of<br />

the solution after reaction (Cu/Pd-BTC_2 as selective catalysts), leaching of palladium is<br />

negligible (i.e., 0.01mg/L), indicating the stability of the used MOF catalysts as well as the<br />

heterogeneous nature of the catalytic reaction.


146 Chapter 5<br />

5.6 Conclusions<br />

In summary, via one-pot synthesis crystalline porous Pd@[Cu3-xPdx(BTC)2]n MOFs with<br />

various doping levels of Pd have been successfully obtained. On the basis of entire<br />

experimental data, structural incorporation of Pd 2+ serving as frameworks-nodes (within<br />

Cu-Pd or/and Pd-Pd paddlewheels) as well as Pd 0 NPs dispersion in the framework of<br />

HKUST-1 are concluded. To best of our knowledge, it is the singular case of MOFs bearing<br />

both Pd 2+ -paddlewheels and Pd NPs known so far. Curiously, certain dependency of the<br />

crystals morphology of Cu/Pd-BTC and extent of relative ratio of Pd-NPs and Pd 2+ -nodes<br />

on the employed Pd-precursors could be traced. Moreover, distinct Pd sites, especially<br />

Pd 2+ /M-CUS considerably enhances MOFs performance in the aqueous-phase<br />

hydrogenation of PNP to PAP comparing to the “Pd-free” Cu-BTC catalysts, which affords<br />

a perspective way to tune their catalytic activity. Furthermore, a synthesis of HKUST-1<br />

analogs excluding Pd 0 formation and featuring exclusive Pd 2+ /Cu 2+ mixed-metal PWs<br />

would lead to highly active catalysts.


6 Summary and Outlook<br />

Among all MOF’s features, one of the most important and interesting characteristics is the<br />

presence of coordinatively unsaturated metal sites (CUS). Due to much stronger affinity<br />

of many “bare” metal-ions towards some gases (e.g. H2, CO2, CH4) and organic molecules,<br />

design of CUSs within MOFs has been a key factor in enhancing MOFs performance in a<br />

variety of applications such as gas storage, gas selectivity and catalysis.<br />

HKUST-1 ([Cu3(BTC)2]n, Cu-BTC) represents one of the well-known and investigated<br />

MOFs where removal of the coordinated water molecules and, thus, generation of CUSs<br />

could be easily achieved upon heating under vacuum. Moreover, [Ru3(BTC)2Yy]n (Ru-<br />

BTC), which is mixed-valence ruthenium structural analog of HKUST-1, has been<br />

elaborated. Interestingly, this ruthenium MOF exhibits sufficient chemical as well as<br />

thermal stability, and offers rich photo/redox chemistry. Hence, in this dissertation Cuand<br />

Ru-MOFs of [M3(BTC)2Yy]n family (M = Cu, Ru, Mo, Cr, Ni, Fe; x = 0 or 1.5, respectively)<br />

have been chosen as candidates to study the modification of MOF’s local structures (via<br />

modification of CUSs, defects engineering, for instance) and its impact on their properties.<br />

First of all, controlled secondary building units approach (CSA) has been employed to<br />

study the (optimized) formation of isostructural Ru II,II and Ru II,III analogs of HKUST-1.<br />

Notably, compared to the Cu-BTC, the composition and local structure of the Ru-BTC<br />

([Ru3(BTC)2Yy]n·G g) turn to be more complex due to the mixed-valence oxidation state of<br />

Ru-centers. The last requires presence of additional counter-ions X from the employed<br />

Ru-SBUs to compensate the charge of [Ru2] 5+ PW units. Besides, some crucial synthetic<br />

parameters like solvents mixture (H2O: AcOH = 1:0.05) lead to the presence of residual<br />

acetic acid/acetate (as counter-ions X, guest molecules/species G or both) in the final<br />

frameworks. By utilizing distinct Ru II,III -SBUs (either [Ru2 II,III (OOCR)4X] or<br />

[Ru2 II,III (OOCCH3)4]A) with various nature of counter-ions X and A and alkyl groups -R, the<br />

local environment of the metal sites in the obtained isostructural MOF solids could be<br />

tuned to a certain extent. Thus, employment of the [Ru2 II,III (OOCCH3)4]A (where A is a<br />

weakly coordinating anion) instead of [Ru2 II,III (OOCR)4X] (where X is a strongly


148 Chapter 6<br />

coordinating anion), facilitates to decrease the amount of coordinated counter-ions in the<br />

framework of Ru II,III -BTC. Additionally, solvent exchange procedure is another way to<br />

remove residual AcO(H), which is occluded within the framework’s pores or strongly<br />

coordinated to the Ru-sites. Furthermore, the usage of the mono-valence Ru II,II -SBU<br />

([Ru2 II,II (OOCCH3)4]) without any counter-ions has turned to be an optimized way to<br />

prepare highly porous Ru II,II -BTC MOF featuring more accessible Ru-CUSs in comparison<br />

with its mixed-valence Ru II,III -BTC variant.<br />

Secondly, modification of the metal PWs in [M3(BTC)2Yy]n has been accomplished by inframework<br />

incorporation of a scope of di-topic isophthalate (ip) defect generating linkers<br />

DLs (5-X-ip, X = OH, H, NH2 or Br). Remarkably, two kinds of structural defects have been<br />

found in the obtained crystalline and isostructural DEMOFs. One type (A) represents<br />

modified PW nodes featuring reduced metal sites, while the other one (type B) is<br />

associated with the missing PW nodes. In fact, the abundances of these two defect types<br />

depend on the choice of the functional group X in the DLs, doping level of DL as well as the<br />

chosen MOF matrix (i.e., Cu-BTC or Ru-BTC). Moreover, the defects of type A and B in Ru-<br />

DEMOFs seem to play a key role in sorption of small molecules (i.e., CO2, CO, H2) and<br />

catalytic activity (i.e., in ethylene dimerization and Paal-Knorr reaction). Thus,<br />

investigations performed on the Ru-DEMOF and Cu-DEMOF solid solutions have shown<br />

rather a complex picture. Still, the studies on the defects characterization have pave up<br />

the way to more deep understanding of the nature of created defects and defects<br />

engineering in MOFs in general.<br />

Apart from the mixed-linker solid solution approach, mixing of distinct metal-ions via onepot<br />

synthesis has also been utilized to successfully obtain Pd-doped solids Pd@[Cu3-<br />

xPdx(BTC)2]n. To note, simultaneously structural incorporation of palladium within Pd-Pd<br />

and/or Pd-Cu PWs as framework-nodes and Pd nanoparticles (NPs) dispersion has been<br />

achieved for the first time. Interestingly, employing prepared Pd-doped Cu/Pd-BTC as<br />

catalysts revealed their enhanced activity (in comparison with the intact Cu-BTC), namely<br />

in aqueous-phase hydrogenation of p-nitrophenol (PNP) to p-aminophenol (PAP) using<br />

NaBH4 as a hydrogen source. Furthermore, the stability and heterogeneous nature of the<br />

catalysts have been confirmed.<br />

All in all, the studies presented in this dissertation provide a deep insight into the<br />

formation and structure of the complex Ru-BTC. Moreover, creation of various types of<br />

structural defects has been gained and analyzed in its defects-engineered variants (Ru-


Chapter 6 149<br />

DEMOFs). These results are essential for further studies on elaboration of the related<br />

DEMOFs. For example, one may avoid many complications by means of employing Ru II,II<br />

SBUs (instead of mixed-valence Ru II,III -SBUs) to form respectively Ru II,II -MOF, which could<br />

by subsequently used as a matrix for defects engineering. From the other side, the doping<br />

of Pd 2+ into less “complicated” Cu-BTC (because of more substitution labile Cu 2+<br />

precursors) has open a way to exploit more functionality of the resulting MOFs. What<br />

more challenging is, to strictly exclude any reducing parameters during synthesis and<br />

prepare a Pd 0 free Pd 2+ /Cu 2+ mixed-metal analog of HKUST-1. Due to the specific<br />

palladium features (like ability to split H2, etc.), this mixed-metal Cu/Pd-BTC materials<br />

including the defect engineering should hold huge promises for many applications. For<br />

instance, gas sorption (e.g. hydrogen uptake), conductivity (TCNQ loading on Cu/Pd-BTC<br />

to modify the electron density at the metal sites, TCNQ = tetracyanoquinodimethane),<br />

catalysis (e.g. using Cu/Pd-BTC as catalysts in alcohol oxidation, cross coupling reaction,<br />

etc.) are topics worth being studied. Last but not the least, phase pure mixed-component<br />

MOFs, structural analogs of HKUST-1, incorporating both Zn 2+ and Cu 2+ ions, H3BTC and<br />

5-nitroisophthalic acid defect linker have been obtained as well (See chapter 7.5).<br />

However, MOFs with only quite low concentration of the incorporated Zn 2+ (about 3%)<br />

have been obtained during this initial study. In a long run, such combination of mixedmetal<br />

and defect linker within single-phased framework could also be very attractive for<br />

advanced properties owing to the more factors (“tuning keys”) influencing modification<br />

on the metal sites.


7 Experimental Section<br />

In this Chapter various experimental and analytical procedures are given. In the first part,<br />

general analytical procedures and methods broadly used for characterization of the<br />

prepared MOF materials in this work are provided. In the second part, synthetic<br />

procedures as well as supplementary analytical details are described for each Chapter.<br />

7.1 General methods<br />

7.1.1 X-ray Diffraction (XRD)<br />

Powder XRD (PXRD) for all the bulk as-synthesized samples and most of the activated<br />

samples were performed by Dr. M. Tu, S. Wannapaiboon, and W. Zhang on an X’ Pert PRO<br />

PANalytical equipment (Bragg–Brentano geometry with automatic divergence slits,<br />

position sensitive detector, continuous mode, room temperature, Cu-Kα radiation(λ =<br />

1.54178 Å), Ni-filter, in the range of 2θ = 5–50, step size 0.01). Activated sample were<br />

prepared on a silicon wafer in an Ar-filled glovebox right before the measurement starts.<br />

For the measurement of activated samples (Ru-DEMOFs series 1 and 3), the data were<br />

collected by W. Zhang with the assistance of Dr. C. Sternemann, A. Schneemann, and S.<br />

Wannapaiboon at beam line BL9 of the synchrotron radiation facility DELTA at a<br />

wavelength of 0.4592 Å using a two-dimensional MAR345 image plate detector. The<br />

sample was filled into standard capillaries (0.5-mm diameter) in an Ar-filled glovebox and<br />

measured. The data were integrated using the program package Fit2D [297] and<br />

transformed to the CuKα radiation used for all other PXRD measurements on reported<br />

materials for the convenient comparison. PXRD of some other activated samples were<br />

collected by Dr. K. Khaletskaya, C. Rösler, and I. Schwedler in a D8-Advance Bruker AXS<br />

diffractometer with CuKα radiation at 25 °C (Göbel mirror; θ–2θscan; 2θ= 5–50°; step size<br />

= 0.0142 (2θ); position sensitive detector; α-Al2O3 as external standard. The sample was<br />

filled into 0.7 mm diameter standard capillaries in an Ar-filled glovebox and measured in<br />

Debye-Scherrer geometry. All the data were interpreted/analyzed by W. Zhang.


Chapter 7 151<br />

Single crystal XRD data for Ru-SBU were collected on an Oxford Supernova with Atlas<br />

detector (Cu-Kα 1.54184 Å) by Dr. K. Freitag and M. Winter. The crystals were coated with<br />

a perfluoropolyether, collected with a glass fiber loop under a microscope and mounted<br />

in the nitrogen cold gas stream of the diffractometer. For the air-sensitive crystals, they<br />

were transferred from a Schlenk tube in an argon stream onto a specimen holder with<br />

perfluoro-polyalkylether before being picked up with the glass fiber loop. The structural<br />

solution and refinement were done M. Winter and W. Zhang (with the assistance of, Dr. A.<br />

Puls, and Dr. S. Henke) using the program suite Olex2 [298] with the executables SHELXS97<br />

and SHELXL97. [299] For refinement the full-matrix least squares on F2 method was<br />

employed. All non-hydrogen atoms were refined anisotropically while the hydrogen<br />

atoms were calculated and refined with isotropic character.<br />

7.1.2 FT-IR spectroscopy<br />

Conventional FTIR spectra were collected by W. Zhang on a Bruker Alpha FTIR instrument<br />

in the ATR geometry with a diamond ATR unit in the range 400–4000 cm -1 inside a<br />

LABStar MB 10 compact MBraun glove-box (argon atmosphere).<br />

Ultra High Vacuum (UHV) FT-IR measurements were performed using a novel UHV-FTIRS<br />

apparatus (state‐of‐the‐art vacuum IR spectrometer (Bruker, VERTEX 80v) coupled to a<br />

novel UHV system (Prevac)) [300] by our collaboration partners, M. Kauer and P. Guo (the<br />

group of Dr. Y. Wang,) in the Laboratory of Industrial Chemistry (Prof. Dr. M. Muhler),<br />

Ruhr‐University Bochum. The sample transfer to the instrument, data acquisition and<br />

interpretation were carried out by M. Kauer and P. Guo. The discussion of the selected<br />

data and the experimental/synthetic work documented in the thesis were done by W.<br />

Zhang. The powder samples were first pressed into a stainless steel grid (0.5 x 0.5 cm)<br />

covered by gold and then mounted on a sample holder, which was specially designed for<br />

the FTIR transmission measurements under UHV conditions. The grid was cleaned by<br />

heating up to 850 K to remove all contaminants formed on it during preparation. The base<br />

pressure in the measurement chamber was 5 x 10 -11 mbar. The optical path inside the IR<br />

spectrometer and the space between the spectrometer and UHV chamber were also<br />

evacuated to avoid atmospheric moisture adsorption resulting in a high sensitivity and<br />

stability. The MOF samples were cleaned in the UHV chamber by heating to 500 K in order<br />

to remove the contaminants involved during synthesis and all the adsorbed species such


152 Chapter 7<br />

as water and hydroxyl groups. Prior to each exposure, a spectrum of the clean sample was<br />

recorded to be as a background reference. The exposure of the sample to CO and CO2 was<br />

carried out by backfilling the measurement chamber through a leak valve. All UHV-FTIR<br />

spectra were collected with 512 scans at a resolution of 4 cm -1 in transmission mode. The<br />

figures below show the UHV-FTIR apparatus (Figure 7.1) and the schematic<br />

representation of the whole optical path of the IR beam in both reflection and<br />

transmission geometries (Figure 7.2).<br />

Figure 7.1. The UHV‐FTIRS apparatus in (a) perspective view and (b) top view. The labeled<br />

components are: (1) load‐lock, (2) distribution, (3) magazine, (4) measurement (FTIR) chambers,<br />

(5) sample manipulator, (6) vacuum FTIR spectrometer, (7) preparation chamber (planned).<br />

Reprinted with permission from Y. Wang, A. Glenz, M. Muhler, C. Wöll, Rev. Sci. Instrum. 2009, 80,<br />

113108‐6, with the permission of AIP Publishing. [300]


Chapter 7 153<br />

Figure 7.2. Schematic representation of different types of IR measurements: (a) reflectionabsorption<br />

at grazing incidence, (b) transmission. Reprinted with permission from Y. Wang, A.<br />

Glenz, M. Muhler, C. Wöll, Rev. Sci. Instrum. 2009, 80, 113108‐6, with the permission of AIP<br />

Publishing. [300]<br />

7.1.3 Nuclear Magnetic Resonance (NMR) Spectroscopy<br />

Liquid phase 1 H-NMR spectra for the digested activated MOFs were measured and<br />

analyzed by W. Zhang on a Bruker Avance DPX-200 spectrometer at 293 K using TMS as


154 Chapter 7<br />

internal standards and normalizing the signals to the DMSO-d6 signal. The samples were<br />

digested in 0.5 ml DMSO-d6 and 0.1 ml DCl (38 wt %)/D2O(99 %) in a NMR tube. Spectra<br />

for Ru-(DE)MOFs were collected with 512 scans and for the other samples with 128 scans.<br />

7.1.4 Thermogravimetric analyses (TGA)<br />

The thermogravimetric analyses (TGA) data were collected and analyzed by W. Zhang<br />

using a TG/DSC NETZSCH STA 409 PC instrument at a heating rate of 5 K min -1 in a<br />

temperature range from 30–600 °C at atmospheric pressure under flowing N2 (N2<br />

,99.999%; gas flow, 20 ml min -1 ). The sample weight is in a range of 7-12 mg. Activated<br />

sample was filled in a crucible with a vial in an Ar-filled glove-box and immediately used<br />

for the measurement.<br />

7.1.5 Scanning Electron Microscopy (SEM) and Energy-dispersive X-ray<br />

spectroscopy (EDX)<br />

SEM images and EDX mapping were recorded from an FEI ESEM Dual Beam TM Quanta 3D<br />

FEG microscope by Dr. K. Khaletskaya and S. Cwik. In order to increase the conductivity<br />

of MOF samples, they were coated with gold by Dr. R. Neuser prior to being loaded in the<br />

instrument. Sample preparation for the SEM measurement and analysis were done by W.<br />

Zhang.<br />

7.1.6 Transmission electron microscope (TEM) and EDX<br />

Bright filed transmission electron microscope (BFTEM) images and EDX spectra for the<br />

composition determination of Ru-DEMOFs were recorded on a Tecnai G 2 F20 equipped<br />

with a Schottky field emission gun operated at an acceleration voltage of 200 kV by Dr. C.<br />

Wiktor at the department of Mechanical Engineering, Ruhr-University Bochum. Data<br />

analysis was done by W. Zhang.<br />

7.1.7 Elemental Analysis / Atomic absorption spectroscopic (AAS)<br />

EA / AAS analysis data for Ru-(DE)MOFs were obtained from the Mikroanalytisches<br />

Laboratorium Kolbe in Mülheim an der Ruhr (http://www.mikro-lab.de/index.html).<br />

Samples were filled into finger schlenks in an Ar-filled glove-box by W. Zhang and


Chapter 7 155<br />

measured under Ar. Atomic absorption spectroscopic (AAS) analysis for mixed-metal<br />

materials to determine Cu, and Pd contents were performed in the Microanalytical<br />

Laboratory of the Department of Analytical Chemistry at the Ruhr‐University Bochum by<br />

K. Bartholomäus. An AAS apparatus by Vario of type 6 (1998) was employed. Data<br />

interpretation was done by W. Zhang.<br />

7.1.8 Gas Sorption<br />

Experiments on sorption of CO2 (298 K) and H2 (77 K) as well as CO (298K) was performed<br />

with assistance of D. J. Xiao and D. Reed (the group of Prof. J. R. Long) at the Department<br />

of Chemistry in the University of California, Berkeley (USA). Part of the N2 sorption data<br />

were collected and interpreted by N. Arshadi at the laboratory of Industrial Chemistry<br />

department (Prof. Dr. M. Muhler) using a Quantachrome Autosorp-1 MP instrument. Most<br />

of N2 sorption data were collected and analyzed by W. Zhang.<br />

CO adsorption (298 K) and the majority of N2 (99.999%) sorption (77 K) measurements<br />

were performed using a Micromeritics 3Flex instrument. The N2 sorption isotherms of Ru-<br />

DEMOFs with defect 5-NH2-ip linker (3a-3d) samples were collected using a<br />

Quantachrome Autosorp-1 MP instrument, optimized protocols and N2 of 99.9995 %<br />

purity. Sorption isotherms of CO2 (298 K) and H2 (77 K) were conducted on a<br />

Micromeritics ASAP 2020 gas adsorption analyzer. The sample cells were cooled with a<br />

liquid nitrogen bath (77 K) or a liquid argon bath (87 K). For room temperature<br />

measurements (298 K) the sample cells were tempered in a water bath.<br />

Calculation of specific surface area and the isosteric heat of adsorption<br />

A method generated from BET theory, which was developed by Brunauer, Emmett, and<br />

Teller in 1938 as an extension of Langmuirs monolayer adsorption to a very simple model<br />

of multilayer physisorption, is commonly employed to calculate the specific surface area<br />

of a porous solid. This area value is usually determined from the data of physisorption<br />

nitrogen isotherms recorded at 77 K.<br />

The BET theory can be derived similarly to the Langmuir theory, but with the<br />

consideration of multilayered gas molecule adsorption, where it is not required for a layer<br />

to be completed before an upper layer formation starts. For the development of the BET<br />

method five assumptions have been made: i) The sample surface is homogeneous; ii)


156 Chapter 7<br />

Uppermost layer is in equilibrium with vapor phase. iii) The heat of adsorption at all sites<br />

is constant and equal to the heat of condensation from the second layer on; iv) There is no<br />

lateral interactions between molecules. v) At saturation pressure, the number of layers<br />

has infinite thickness. Hence, the BET equation could be derived:<br />

p<br />

n a (p 0 − p) = 1<br />

n a m C<br />

+<br />

(C − 1)<br />

n m a C<br />

Here n a is the amount adsorbed, p is the pressure, p0 is the saturation pressure of the<br />

adsorbed gas at a given temperature, n m<br />

a<br />

is the monolayer capacity and C an empirical<br />

constant. The value of C can be associated to the magnitude of adsorbent-adsorbate<br />

interactions. This equation is an adsorption isotherm and can be plotted as a BET plot<br />

where 1/[n a (p0/p-1)] is the y-axis and p/p0 on the x-axis. The linear relationship of the<br />

BET plot is kept only in a pressure region of p/p0= 0.05 - 0.30. For microporous materials<br />

such as MOFs, a even smaller pressure range (p/p0 = 0.02 - 0.10) is usually applied.<br />

Subsequently, the monolayer capacity n m<br />

a<br />

can be derived from the slope of the BET plot.<br />

Consequently, the specific surface area SBET can be calculated from the following equation:<br />

SBET = n m a N A a m /m<br />

p<br />

p 0<br />

where NA is the Avogadro constant and am is molecular cross-section area occupied by a<br />

single adsorbate molecule in the complete monolayer and m the mass of adsorbent. Still,<br />

we should note that the BET theory is developed on basis of the oversimplified model. In<br />

fact, SBET is greatly affected by the nature of the adsorbent as well as the strength of<br />

adsorbate-adsorbent interactions. Microporous solids including MOFs, the pore surface of<br />

which are rather heterogeneous, the above mentioned assumptions are not valid. The<br />

obtained value of the BET surface area (SBET) can be widely used to compare the porosity<br />

of different porous materials and should not be considered as a precise absolute value.<br />

The adsorption isotherms of H2 adsorption at 77 K and 87 K were independently fit with<br />

triple site Langmuir model<br />

n = q sat,Ab A p<br />

1 + b A p + q sat,Bb B p<br />

1 + b B p + q sat,Cb C p<br />

1 + b C p<br />

where n is the molar loading of adsorbate (mmol/g), qsat is the saturation loading of site<br />

A/B/C (mmol/g), b is the Langmuir parameter for A/B/C (bar -1 ), and p the pressure(bar).<br />

After the triple-site Langmuir fits, the exact pressures correspond to the same amount


Chapter 7 157<br />

adsorbed at different temperatures can be plotted. Subsequently, the Clausius−Clapeyron<br />

equation,<br />

−Q st = RT 2 ( ∂lnp<br />

∂T ) n<br />

was used to calculated the isosteric heats of adsorption(-Qst) as a function of the amount<br />

of H2 adsorbed (n). -Qst is a value represent for the average binding energy of an adorbate<br />

at specific surface coverage.<br />

7.1.9 X-ray absorption near edge structure (XANES)<br />

Data acquisition and analysis for XANES were done at beam line BL8 of the synchrotron<br />

radiation facility DELTA by W. Zhang with the assistance of R. Wagner (and A.<br />

Schneemann).<br />

Simplified theory behind XANES<br />

XANES, also known as near edge X-ray absorption fine structure (NEXAFS), is a kind of<br />

absorption spectroscopy showing the features in the X-ray absorption spectra (XAS) of<br />

condensed matter because of the photoabsorption cross section for electronic transitions<br />

from an atomic core level to final states in the energy region of 50-100 eV above the<br />

selected atomic core level ionization energy. X-rays are ionizing electromagnetic radiation<br />

which afford sufficient energy to excite a core electron of an atom to an excited state (an<br />

empty below the ionization threshold), or to the continuum which is above the ionization<br />

threshold. A core hole is the space that a core electron occupied before it absorbs an X-ray<br />

photon and ejected from its core shell. Subsequently, another atomic electron drops down<br />

in to fill the core hole and release energies either through Auger electron ejection or X-ray<br />

Fluorescence.<br />

When the energy of X-ray radiation is scanning through the binding energy region of a<br />

core shell, there will be a sudden appearance of absorption increase, corresponding to<br />

absorption of the X-ray photon by a specific type of core electrons (eg. 1s, 2s, 2p electrons,<br />

etc.). This gives rise to a so-called absorption edge (K, L, M edge, etc.) in the XAS spectrum<br />

due to its vertical appearance. The binding energies of different core electrons are distinct.<br />

Atoms with a higher oxidation state need more energetic X-ray to excite its core level due


158 Chapter 7<br />

to the more tightly bound electron states. Consequently, the absorption edge energy<br />

increases along with the increase of oxidation-state of the absorption site.<br />

Sample measurements details<br />

The collection of XANES spectra was performed at the Ru-K edge (22117 eV) using a Si<br />

(311) monochromator at Beamline BL8 of the synchrotron radiation facility DELTA, TU<br />

Dortmund. The samples were filled into 1 mm caplillaries in an inert Ar atmosphere glovebox<br />

before the measurement. The set-up was acquired using 15cm Ionchamber filled with<br />

Ar as I0, a 15 cm Ionchamber filled with Xe as I1 and a 30 cm Ionchamber filled with Xe as<br />

I2. Approx 30 mm above the sample was the PIN-Diode capturing a relatively wide solid<br />

angle of fluorescence radiation. The energy was calibrated by measuring a metallic Ru foil<br />

as reference simultaneously to each sample scan. The data processing and analysis were<br />

done using the program ATHENA. [301]<br />

7.1.10 X-ray photoelectron spectroscopy (XPS)<br />

XPS spectra were recorded and analyzed by our collaboration partners, P. Guo and Dr. Y.<br />

Wang in the Laboratory of Industrial Chemistry (Prof. Dr. M. Muhler), Ruhr‐University<br />

Bochum. The discussion of the selected data was done by W. Zhang.<br />

XPS measurements were performed in a UHV setup equipped with a high-resolution<br />

Gammadata-Scienta SES 2002 analyzer. A monochromatic Al Kα X-ray source (energy<br />

1486.6 eV) was used as incident radiation. The analyzer slit width was set at 0.3 mm and<br />

the pass energy was fixed at 200 eV for all the measurements. The overall energy<br />

resolution was better than 0.5 eV. A flood gun was used to compensate for the charging<br />

effects. All spectra reported here are calibrated to the C 1s corelevel binding energy at 285<br />

eV. The XP spectra were deconvoluted using the CASA XPS program with a mixed Gaussian<br />

−Lorentzian function and Shirley background subtraction.<br />

7.1.11 High Performance Liquid Chromatography (HPLC)<br />

HPLC measurements to determine the ratio of the organic components were conducted<br />

by Dr. H. B. Albada and M. Strack at the Chair of Inorganic Chemistry I-Bioinorganic<br />

Chemistry (Prof. Dr. N. Metzler-Nolte), Ruhr‐University Bochum.<br />

Experimental setup:


Chapter 7 159<br />

Buffers: A: water/MeCN/TFA – 95/5/0.1 (%, v/v/v); B: MeCN/water/TFA – 95/5/0.1 (%,<br />

v/v/v)<br />

Gradient: start with 100 % A for 5 min; then in 15 min to 75 % B (i.e. 5 %/min); steady at<br />

75 % B for 2.5 min; drop to 100 % A in 5 min (i.e. 20 %/min); steady at 100 % A for 2.5<br />

min.<br />

Column: Prontosil, RP C18‐AQ, with pre‐column; dimensions: 250 x 4.6 mm (length x<br />

internal diameter), particle size: 5 μm; pore diameter: 120 Å (Knauer, Batch No. 2362,<br />

Column SN: CG 108, 25VF184PSJ).<br />

Detection: λ = 254 nm.<br />

Equipment: HPLC‐runs were performed on a Knauer HPLC‐machine.<br />

7.1.12 Catalytic studies<br />

Catalytical tests for Paal-Knorr reaction in Chapter 4 were performed in collaboration<br />

with the group of Prof. A. Corma, in particular by K. Epp (TU Munich) and Dr. F. X. Llabrés<br />

i Xamena, in the Institute of Chemical Technologies (Instituto de Tecnología Química, ITQ),<br />

Polytechnical University of Valencia (Spain). Ethylene dimerization in Chapter 4 was<br />

carried out by W. Zhang with the assistance of M. Gonzalez (the group of Prof. J. R. Long)<br />

in the department of Chemistry, University of California, Berkeley (US). Catalytic reaction<br />

for the hydrogenation of PNP to PAP in Chapter 5 was performed by Z. Chen and M. Al-<br />

Naji (Group of Prof. R. Gläser) in the institute of Chemical Technology, University of<br />

Leipzig. Data for the catalysis was analyzed by W. Zhang with the assistance of these<br />

collaborators.<br />

Gas Chromatography measurements for monitoring the product of Paal-Knorr reaction<br />

were performed on an Agilent Technologies 7890A with FID (Flame Ionization Detector)<br />

using a capillary column HP-5 (5% phenylmethylpolysiloxane) of 30 m length and 0.32<br />

mm internal diameter as well as BP20(WAX) of 15 m length and 0.32 mm internal<br />

diameter as another column. Thereby, the products were measured in high dilution using<br />

volatile organic solvents (usually ethanol or acetone).


160 Chapter 7<br />

7.2 Experimental data on chapter 3<br />

All chemicals (RuCl3·xH2O, LiCl, AgSO4, NaBPh4, AgBF4, pivalic acid, benzene-1, 3, 5-<br />

tricarboxylic acid (H3BTC)) and all solvents (CH3COOH, H2O, CH3OH, EtOH, acetic<br />

anhydride, acetone and hexane) were used as commercially received unless otherwise<br />

noted. Tetrahydrofuran (THF) used in the synthesis of SBU-c was catalytically dried,<br />

deoxygenated, and saturated with argon using an automatic solvent purification system<br />

from MBraun.<br />

7.2.1 Synthesis of the ruthenium precursors ([Ru2(OOCR)4X] and<br />

[Ru2(OOCCH3)4]A)<br />

Scheme 7.1. Preparation of the Ru-SBUs (SBU-a to SBU-d) used for MOF syntheses.<br />

[Ru2(OOCCH3)4Cl] (SBU-a)<br />

This compound was prepared using the method reported in the literature with slight<br />

modification. [202] RuCl3·xH2O (0.5 g, 2 mmol) and LiCl (0.5 g, 11.9 mmol) were mixed in<br />

one flask. Then glacial acetic acid (17.5 ml) and acetic anhydride (3.5 ml) were added.<br />

Subsequently, the mixture was refluxed for 24 h. The resulting red-brown precipitate was<br />

collected, washed several times with acetone, and dried under ambient conditions first<br />

and then under dynamic vacuum for at least 4 h (Yield: 0.33 g).


Chapter 7 161<br />

Figure 7.3. PXRD patterns of [Ru 2 (OOCCH 3 ) 4 Cl] SBU-a in comparison with the simulated patterns<br />

reported. [203]<br />

Figure 7.4. Acid-digested (DCl) 1 H-NMR spectrum of [Ru 2 (OOCCH 3 ) 4 Cl] (SBU-a) in the solvent of<br />

DMSO-d 6 .


162 Chapter 7<br />

[Ru2(OOCC(CH3)3)4(H2O)Cl](CH3OH) (SBU-b)<br />

The synthesis was carried out according to the literature. [201] SBU-a (0.3 g, 0.63 mmol)<br />

was dissolved in 60 ml solution of the H2O : CH3OH (1 : 1) mixture. Later, pivalic acid (0.39<br />

g, 3.79 mmol) was added and the resulting mixture was refluxed. The reaction was<br />

finished after 5 hours, when the solution turned to a red-brown color. This solution was<br />

cooled down and left to stay overnight. As a result, red-brown crystals precipitated. The<br />

final product was collected and washed twice with hexane (Yield: 0.37 g).<br />

Figure 7.5. Schematic description of the molecular structure of SBU-b in the solid state obtained<br />

by single crystal structure analysis. Note the half occupied Cl and half occupied O positions at the<br />

apical position of Ru coordination sphere. The oxygen atoms of the methanol molecule in the<br />

asymmetric unit (solvate) were split into two parts due to disorder. Circle with cross: Ru; circle<br />

shaded: Cl; circle shaded bottom right to top left: O; circle with blank: C. Single crystals of SBU-b<br />

were measured at low temperature (101K) while in the literature reference at 295K. The carbon<br />

atoms in CMe 3 were refined isotropically as well as the centrosymmetric structural model with<br />

partial occupation of Cl and O water as in the reference. [201] However, the measurement in the lower<br />

temperature reported here tends to be of better quality.


Chapter 7 163<br />

Table 7.1. Cell parameters of [Ru 2 (OOCC(CH 3 ) 3 ) 4 (H 2 O)Cl](CH 3 OH) (SBU-b) obtained from single<br />

crystal X-ray diffraction and reported in the literature. [201]<br />

Obtained Literature [201]<br />

Cell<br />

parameters<br />

a = 9.256(4)Å,<br />

b = 9.383(4) Å,<br />

c = 9.605(4) Å;<br />

α = 73.712(4)°,<br />

β = 83.681(4)°,<br />

γ = 73.585(4)°;<br />

V = 767.65(6)Å 3<br />

a = 9.353(3)Å,<br />

b = 9.469(6) Å,<br />

c = 9.804(4) Å;<br />

α = 72.57(4)°,<br />

β = 83.21(3)°,<br />

γ = 73.88(4)°;<br />

V = 795.3(7) Å 3<br />

Figure 7.6. Acid-digested (DCl) 1 H-NMR spectrum of [Ru 2 (OOCC(CH 3 ) 3 ) 4 (H 2 O)Cl](CH 3 OH) (SBUb)<br />

in the solvent of DMSO-d 6 .


164 Chapter 7<br />

Figure 7.7. The PXRD patterns of [Ru 2 (OOCC(CH 3 ) 3 ) 4 (H 2 O)Cl](CH 3 OH) (SBU-b) measured at<br />

different states (immediately after synthesis, one day later and two days later). From the PXRD<br />

patterns, we can infer that when the single crystal stays at room temperature for a while, it will<br />

lose the solvents gradually, which leads to the disappearance of certain reflexes.


Chapter 7 165<br />

[Ru2(OOCCH3)4(THF)2]BF4 (SBU-c)<br />

The compound was synthesized under Ar using the method reported in the literature. [200]<br />

SBU-a (100 mg, 0.211 mmol) and AgBF4 (42 mg, 0.211 mmol) were stirred in dry THF (4<br />

ml) with the protection of tin foil at room temperature. A red solution and white<br />

precipitate of AgCl was obtained after 24 h. The solution was filtered and the precipitate<br />

was washed several times with THF until colorless. All the filtrates were collected<br />

together, and the volume was reduced (ca. 4 times) in vacuum and the product was<br />

obtained (Yield: 108 mg).<br />

Figure 7.8. Acid-digested (DCl) liquid 1 H-NMR spectrum of SBU-c (solvent: DMSO-d 6 ).


166 Chapter 7<br />

[Ru2(OOCCH3)4(H2O)2]BPh4 (SBU-d)<br />

This precursor was synthesized in accordance to the literature procedure. [199] SBU-a (0.48<br />

g, 1.0 mmol) and AgSO4 (0.17 g, 0.5 mmol) were stirred in 20 ml of water at 40 °C. A white<br />

powder (AgCl) precipitated after 10 min and was filtered off. The resulting brown solution<br />

was cooled down to 0 °C, and then the aqueous solution of NaBPh4 (0.4 g, 1.2 mmol) was<br />

added drop-wise while stirring. The orange-brown product precipitated immediately.<br />

Subsequently, it was filtered and washed with cold water three times (Yield: 0.64 g). The<br />

final product was dried in vacuum at room temperature and stored in a fridge.<br />

Figure 7.9. 1 H-NMR spectrum of SBU-d (solvent: DMSO-d 6 ).<br />

[Ru2(OOCCH3)4] (SBU-e)<br />

The precursor SBU-e was obtained according to the reported method. [206-207] RuCl3·xH2O<br />

(4 g) together with 20mg PtO2 were put into a Fischer-Porter bottle in an argon-filled<br />

glove-box. Then 50 mg degassed methanol was added. After the whole system was<br />

degassed, the reactor was pressurized with ca. 2 atm hydrogen. The solution was stirred<br />

for 3 hours until it was deep blue. Subsequently, the solution was filtered under argon into<br />

50 ml degassed methanol which contain 4.7 g Li acetate. After heating under reflux for 18


Chapter 7 167<br />

hours, a deep red brown solution and an orange-brown microcrystalline product were<br />

resulted. The hot solution was filtered and the obtained solid was washed with methanol<br />

(3 x 20 ml). The product was then dried at 80°C under vacuum as a brown powder. m.p.<br />

276 °C.<br />

Scheme 7.2. Synthesis procedure of Ru 2 (OOCCH 3 ) 4 (SBU-e).<br />

Figure 7.10. IR spectra of [Ru 2 II,II (OOCCH 3 ) 4 ] (SBU-e) in comparison with [Ru 2 II,III (OOCCH 3 ) 4 Cl]<br />

(SBU-a). The vertical dash line represents the position of vibrations originated from ν as (COO) and<br />

ν s (COO) in [Ru 2 II,II (OOCCH 3 ) 4 ].


168 Chapter 7<br />

[Ru2(OOCCH3)4](THF)2<br />

Single crystal of [Ru2(OOCCH3)4](THF)2 was obtained in accordance to the reported<br />

procedure. [207] In particular, the sample of [Ru2(OOCCH3)4] was dissolved into<br />

tetrahydrofuran (THF) at 74 °C (oil bath) for 10 min with stirring. Subsequently, the stir<br />

was stopped and the solution was slowly cooled down (10 °C/30 min) to room<br />

temperature in a slow stream of Ar. The solution was then stored in a freezer at -20 °C<br />

overnight giving brown crystals.<br />

Table 7.2. Cell parameters of single crystal [Ru II,II 2 (OOCCH 3 ) 4 ](THF) 2 . Single crystal structure was<br />

not given due to the formation of twin crystal.<br />

obtained literature [207]<br />

a = 14.3787 Å, b = 9.6279 Å,<br />

c = 15.4351 Å; α = 90.000°,<br />

β = 90.000°, γ = 90.000°;<br />

V = 2134.5Å 3<br />

c = 14.606(3)Å a = 9.598(2)Å,<br />

b = 15.803(3)Å; α = 90.00°,<br />

β = 90.00°, γ = 90.00°;<br />

V = 2215.4Å 3<br />

7.2.2 Synthesis of Ru-MOFs [Ru3(BTC)2Xx]·Gg<br />

Scheme 7.3. Syntheses scheme of Ru-MOFs (1-4) employing various SBUs. Reaction Conditions<br />

for 1: CH 3 COOH, H 2 O, 160 °C, 72h; 2: CH 3 COOH, H 2 O, 160 °C, 168h; 3: CH 3 COOH, H 2 O, 160 °C, 72h;<br />

4: CH 3 COOH, H 2 O, 120 °C, 72h.


Chapter 7 169<br />

[Ru3(BTC)2Cl(AcO)0.5]n·(AcOH)1.5(H3BTC)0.1(H2O)4 (1_ex)<br />

This sample was obtained according to the method reported by Kozachuk et al. [82] A<br />

sample of SBU-a (170 mg, 0.36 mmol) and H3BTC (101 mg, 0.48 mmol) were placed in a<br />

20 ml Teflon vessel, and 4 ml of H2O and 0.7 ml of CH3COOH were subsequently added.<br />

The vessel was sealed in an autoclave and kept in a preheated oven at 160 °C for 72 h. The<br />

crude product was filtered and washed several times with distilled water and dried under<br />

ambient conditions (Yield: 171 mg). The obtained powder was further activated by<br />

heating at 150 °C for 24 h under dynamic vacuum (ca. 10 -3 mbar).<br />

[Ru3(BTC)2Cl1.2(OH)0.3]n·(H3BTC)0.15(AcOH)2.4(PivOH)0.45 (2_ex)<br />

The sample was prepared in accordance to the reported procedure. [83] A sample of SBUb<br />

(270 mg, 0.42 mmol) and H3BTC (120 mg, 0.57 mmol) were placed in a Teflon vessel,<br />

and then 6 ml of H2O and 0.1 ml of CH3COOH were added. The vessel was sealed in an<br />

autoclave and kept in the preheated oven at 160 °C for 120 h. As a result, a powdered<br />

product was obtained. It was filtered, washed several times with ethanol and dried at<br />

ambient temperature first (Yield: 223 mg). The sample was activated by heating under<br />

dynamic vacuum (ca. 10 -3 mbar) at 150 °C for 24 h.<br />

[Ru3(BTC)2F0.7(OH)0.8]n·(AcOH)1.5(H3BTC)0.2 (3_ex)<br />

A sample of SBU-c (241 mg, 0.36 mmol) and H3BTC (101 mg, 0.48 mmol) were mixed in a<br />

20 ml Teflon vessel with 4 ml of H2O and 0.7 ml of CH3COOH. The vessel was subsequently<br />

sealed in an autoclave and placed in the preheated oven at 160 °C for 72 h. The final brown<br />

product was collected, washed several times with distilled water, and dried under<br />

ambient conditions first (Yield: 265 mg). The sample was further activated by heating at<br />

150 °C for 24 h under dynamic vacuum (ca. 10 -3 mbar).


170 Chapter 7<br />

[Ru3(BTC)2(OH)1.5]n·(H3BTC)0.8(AcOH)1.4(H2O)3 (4_ex)<br />

A sample of SBU-d (286 mg, 0.36 mmol) and H3BTC (101 mg, 0.48 mmol), 4 ml of H2O and<br />

0.7 ml of CH3COOH were mixed in a Teflon vessel. The vessel was subsequently sealed in<br />

an autoclave and placed in the preheated oven at 120 °C for 72 h. The final product was<br />

filtered, washed several times with distilled water, and dried under ambient conditions<br />

first (Yield: 191 mg). The material was further activated by heating at 150 °C for 24 h<br />

under dynamic vacuum (ca. 10 -3 mbar).<br />

Figure 7.11. 1 H-NMR spectra of 4 (before solvent exchange) and 4-ex (after solvent exchange) in<br />

comparison with NaBPh 4 .


Chapter 7 171<br />

Figure 7.12. PXRD patterns of the materials obtained from different solvents with a small amount<br />

of acetic acid or without acetic acid as component in the solvent mixture at the same conditions<br />

(160 °C, 72h, solvothermal, starting with SBU-a and H 3 BTC) in comparison with SBU-a and Ru-<br />

MOF 1. a) 4 ml diglyme and 4 drops HCl, b) 4 ml H 2 O and 0.7 ml HCOOH, c) 4 ml H 2 O and 0.7 ml<br />

CF 3 COOH, d) 4 ml diglyme, e) 4 ml H 2 O and 4 drops HCl; f) pH = 1.89 solution of H 2 O and CF 3 COOH,<br />

10 ml (note: the same pH value with the literature; [82] formula: 4ml H 2 O and 0.7ml CH 3 COOH); g)<br />

pH = 1.89 solution of H 2 O and CF 3 COOH, 24 ml; h) pH = 1.89 solution of H 2 O and CF 3 COOH, 4.7 ml;<br />

i) pH = 1.89 solution of H 2 O and HCl, 4.7 ml; j). 12 ml H 2 O and 161 μl CH 3 COOH (pH = 2.47; note<br />

the same pH value with the literature [83] formula: 12ml H 2 O and 161 μl CH 3 COOH). Note: CF 3 COOH<br />

was employed to trace the incorporation of the carboxylic acid by the CF 3 -group.


172 Chapter 7<br />

[Ru3(BTC)2]n∙(AcOH)2.3 (5)<br />

The preparation of Ru-MOF 5 was similar with Ru-MOF 1 using SBU-e (158 mg, 0.36<br />

mmol) and H3BTC (101 mg. 0.48 mmol) with degassed water and AcOH. All the degassed<br />

reactant was then mixed in a Teflon vessel sitting in a schlenk flask under Argon<br />

atomosphere. The vessel was subsequently quickly sealed in an autoclave and placed in<br />

the preheated oven at 160 °C for 72 h. The crude product was filtered and washed several<br />

times with degassed distilled water and dried under vacuum. Further activation of the<br />

powder was performed by heating at 150 °C for 24 h under dynamic vacuum (ca. 10 -3<br />

mbar).<br />

Figure 7.13. 1 H-NMR spectra of the sample 5 ([Ru 3 (C 9 H 3 O 6 ) 2 ] n·(CH 3 COOH) 2.3 ) digested in<br />

DCl/DMSO-d 6 mixture.


Chapter 7 173<br />

7.3 Experimental data on chapter 4<br />

Ruthenium SBU precursor [Ru2(OOCCH3)4Cl]n [202] as well as the parent single-linker Ru-<br />

MOF [82, 198] were prepared following procedures previously described in the literature and<br />

in Chapter 7.2. All other reagents were available commercially and used without further<br />

purification. Before further manipulations, all activated samples were stored in a glovebox<br />

under inert Ar atmosphere.<br />

7.3.1 Synthesis of Ru-DEMOF samples (1a-1d, 2a-2d, 3a-3d, 4a-4c)<br />

[Ru3(BTC)2-x(5-X-ip)xXy]n (X = counter ion; 0.1 ≤ x ≤ 1; 0 ≤ y ≤ 1.5). The samples were<br />

synthesized according to the reported method [138] applying the so-called controlled SBU<br />

approach. The mixtures of parent linker (1,3,5-benzenetricarboxylic acid, i.e. H3BTC) and<br />

defective linker (5-X-isophthalic acid, i.e. 5-X-ip, X = OH (1), H (2), NH2 (3) and Br (4),<br />

respectively) combined with 1.5 molar equivalents of [Ru2(OOCCH3)4Cl]n (0.36 mmol,170<br />

mg) were placed in the 20 ml Teflon vessels. Subsequently, 4 ml of HPLC grade water and<br />

0.7 ml of glacial acetic acid were added. The molar ratios and amounts of the linkers used<br />

in the starting materials are listed in the Table 7.3. The Teflon reaction vessels were then<br />

placed in the Teflon-lined stainless steel autoclaves. The autoclaves were further sealed<br />

and placed into the pre-heated oven at 433 K for 72 h. Afterwards the autoclaves were<br />

taken out and cooled down at room temperature. The resulting powder products were<br />

collected by centrifugation. Afterwards, they were soaked into HPLC grade water (ca. 20<br />

ml) and the solvent was refreshed by the same water amount every 24 hours for 3 times.<br />

Finally, the solvent was removed and collected powders were dried under air at room<br />

temperature. The activation of the solids was performed by heating at 423 K for 24 h<br />

under dynamic vacuum (ca. 10 -3 mbar).


174 Chapter 7<br />

Table 7.3. Feeding molar ratios and amounts of the linkers used in the syntheses of the Ru-<br />

DEMOF samples.<br />

parent linker (H 3 BTC)<br />

defect linker (5-X-ip)<br />

Sample<br />

H 3 BTC : (H 3 BTC + 5-XipH<br />

2 )<br />

mass<br />

(mg)<br />

5-X-ipH 2 : (H 3 BTC + 5-X-ipH 2 )<br />

mass<br />

(mg)<br />

1a 90% 90.8<br />

10% 9<br />

1b 80% 80.7 20% 18<br />

X = OH<br />

1c 70% 71 30% 26<br />

1d 50% 50.5 50% 43.7<br />

2a 90% 90.8<br />

10% 8<br />

2b 80% 80.7 20% 16<br />

X = H<br />

2c 70% 71 30% 24<br />

2d 50% 50.5 50% 40<br />

3a 90% 90.8<br />

10% 8.7<br />

3b 80% 80.7 20% 17.4<br />

X = NH 2<br />

3c 70% 71 30% 26<br />

3d 50% 50.5 50% 43.5<br />

4a 90% 90.8<br />

10% 12<br />

4b 80% 80.7 X = Br<br />

20% 24<br />

4c 70% 71 30% 35.3


Chapter 7 175<br />

Table 7.4. BET surface area (S BET ) values of the Ru-DEMOFs (1a-1d, 2a-2c, 3a-3d, 4a-4c)<br />

compared with the parent Ru-MOF as well as [Cu 3 (BTC) 2 ] n . S BET (m 2 /mmol) of the parent Ru-MOFs<br />

and Ru-DEMOFs 1a, 1c and 1d was converted from the respective measured S BET (m 2 /g) values<br />

according to the proposed composition in Chapter 3 and 4.<br />

Sample S BET , m 2 /g S BET , m 2 /mmol<br />

parent Ru-MOF 1_ex 998 [208] 964<br />

parent Ru-MOF 5 1371 1173<br />

[Cu 3 (BTC) 2 ] n 1734 [83] 1049<br />

1a 1106 1062<br />

1b 1284 -<br />

1c 1302 1232<br />

1d 947 930<br />

2a 1059 -<br />

2b 1023 -<br />

2c 838 -<br />

2d 574 -<br />

3a 1003 -<br />

3b 985 -<br />

3c 975 -<br />

3d 781 -<br />

4a 996 -<br />

4b 847 -<br />

4c 693 -


176 Chapter 7<br />

For comparison of the obtained S BET values for Ru-MOFs with the homologous Cu-MOFs, it might<br />

be more informative using m 2 /mmol units (due to substantial difference in the molecular weight<br />

of Cu and Ru and the presence of guest molecules). Thus, the S BET of the parent Ru-MOF (964<br />

m 2 /mmol) is quite similar to that of a typical reference sample [Cu 3 (BTC) 2 ] n (1049 m 2 /mmol).<br />

More importantly, the S BET of the Ru-DEMOF solids slightly increases in comparison with both<br />

parent analogues. Due to the complexity of the composition (which has been mentioned in the<br />

section 4.2.1.2 of Chapter 4), EA was not performed for the rest Ru-DEMOFs to determine the<br />

unambiguous composition with counter-ions and guest molecules. Hence, S BET values in the unit<br />

of m 2 /mmol were not calculated for them.<br />

Figure 7.14. XPS survey scan of the parent Ru-MOF and Ru-DEMOFs (1a, 1c, 1d, 2a and 2b). 1a,<br />

1c and 1d - 8%, 32% and 37% 5-OH-ip incorporation, respectively; 2a and 2b - 15% and 28% ip<br />

incorporation, respectively. Figure is provided by P. Guo.


Chapter 7 177<br />

7.3.2 Catalytic reactions using Ru-DEMOFs as catalysts<br />

7.3.2.1 Ethylene dimerization<br />

In a typical run (entry 5), 2.1 mg of the activated Ru-MOF 1c was put in a 10 ml steel<br />

reactor followed by adding of 0.81 ml of Et2AlCl (1M in heptane), 2.1 ml of toluene and<br />

0.09 ml of undecane (0.01 M in toluene). After degassing, the reactor was flushed with<br />

C2H4 (800psi) and placed in a pre-heated oil bath at 80 °C with stirring for 2 hours.<br />

Subsequently, the reaction was stopped and the reactor was cooled down at -7 °C in a<br />

mixture of dry ice and ethylene glycol bath. Cold distilled water was added to quench the<br />

reaction. The organic layer kept cold was then quickly analyzed by gas chromatography<br />

(SRI 8610V GC, 60 m x 0.54 mm internal diameter, 5.0 µm MXT-1 capillary columns) using<br />

undecane signal as a reference. In case of no additive, the volume of toluene was changed<br />

to 2.91 ml. Other parameters are the same as entry 5 described above.<br />

7.3.2.2 Paal-Knorr reaction<br />

In a typical run, 5 mg of the activated Ru-MOF/DEMOFs in 0.5 ml of toluene were<br />

introduced into 125 mg (1.1 mmol) of 2,5-hexadione and 95 mg (1 mmol) of aniline and<br />

were stirred at 90 °C for 24h under air. The reaction was followed by taking aliquots every<br />

hour and analyzing the products by GC-MS. The reactions were performed in closed<br />

(pressurize able) reaction vials.<br />

7.3.3 Synthesis of [Cu3(BTC)2]n (Cu-BTC)<br />

[Cu3(BTC)2]n was prepared from a reported method [35] at a litter lower temperature (120<br />

°C). Specifically, Cu(NO3)2∙3H2O (1.8 mmol, 438 mg) and 1, 3, 5-benzentricarboxylic acid<br />

(H3BTC, 1.0 mmol, 210 mg) were put into a Teflon vessel, and a solution of 12 ml H2O:EtOH<br />

(1:1) was added. The vessel was sealed in an autoclave and kept in a preheated oven at<br />

100 °C for 12 h. The crude product was filtered and washed several times with distilled<br />

water and absolute ethanol, respectively. After dried under ambient conditions the<br />

obtained powder (152 mg) was further activated by heating at 120 °C for 24 h under<br />

dynamic vacuum (ca. 10 -3 mbar).


178 Chapter 7<br />

7.3.4 Synthesis of Cu-DEMOF samples (D1-8)<br />

D1<br />

H3BTC (100mg, 0.48 mmol) and ip (80mg, 0.48 mmol) equaling a 1:1 ratio were mixed<br />

with Cu(BF4)2 6H2O (488 mg, 1.4 mmol) and placed into a 20 ml screw jaw. Subsequently<br />

14 ml N,N-Dimethylformamide (DMF) was added. The screw jaw was then sealed and put<br />

into a pre-heated oven at 80 °C for 20 h. The resulting blue powder products were<br />

collected by centrifugation. Afterwards, they were washed by DMF (30 ml x 3), EtOH (30<br />

ml x 3) and acetone (30 ml x 3) subsequently. The final powder was collected and dried<br />

under air at room temperature. The activation of the solids was performed by heating at<br />

170 °C for 24 h under dynamic vacuum (ca. 10 -3 mbar).<br />

Figure 7.15. 1 H-NMR spectra of the acid-digest D1 sample (solvent: DMSO-d 6 , DCl).


Chapter 7 179<br />

D2<br />

This sample was prepared as D1 using 160mg ip (2 molar equivalent to that of H3BTC)<br />

instead. The resulting blue powder products were separated from the solution by<br />

centrifugation. Afterwards, they were washed by DMF(30 ml x 3), EtOH (30 ml x 3) and<br />

acetone (30 ml x 3) subsequently. Finally, the solvent was removed and the powder was<br />

dried under air at room temperature. The activation of the solids was performed by<br />

heating at 170 °C for 24 h under dynamic vacuum (ca. 10 -3 mbar).<br />

Figure 7.16. 1 H-NMR spectra of the acid-digest D2 sample (solvent: DMSO-d 6 , DCl).


180 Chapter 7<br />

D3<br />

The mixture of H3BTC (100mg, 0.48 mmol), ip (80 mg, 0.48 mmol) and Cu(BF4)2 6H2O<br />

(488 mg, 1.4 mmol) was placed into a 50 ml Teflon vessel. Subsequently, 24 ml EtOH was<br />

added. The Teflon reaction vessel was then placed in the Teflon-lined stainless steel<br />

autoclaves. The autoclaves were further sealed and placed into the pre-heated oven at 80<br />

°C for 20 h. Afterwards the autoclaves were taken out and cooled down at room<br />

temperature. The resulting blue powder products were separated from the solution by<br />

centrifugation. Afterwards, they were washed by EtOH (30 ml x 3) and acetone (30 ml x<br />

3) subsequently. Finally, the solvent was removed and the powder was dried under<br />

ambient atmosphere at room temperature. The activation of the solids was performed by<br />

heating at 120 °C for 24 h under dynamic vacuum (ca. 10 -3 mbar).<br />

Figure 7.17. 1 H-NMR spectra of the acid-digest D3 sample (solvent: DMSO-d 6 , DCl).


Chapter 7 181<br />

D4<br />

This sample was prepared as D3 using 160mg ip (2 molar equivalent to that of H3BTC)<br />

instead. The resulting powder products were collected by centrifugation. Afterwards, they<br />

were washed by EtOH (30 ml x 3) and acetone (30 ml x 3) subsequently. Finally, the<br />

solvent was removed and the powder was dried under ambient atmosphere at room<br />

temperature. The activation of the solids was performed by heating at 120 °C for 24 h<br />

under dynamic vacuum (ca. 10 -3 mbar).<br />

Figure 7.18. 1 H-NMR spectra of the acid-digest D4 sample (solvent: DMSO-d 6 , DCl).


182 Chapter 7<br />

D5<br />

This sample was repeated following the reported method. [140] H3BTC (50mg, 0.24 mmol)<br />

and 1 molar equivalent of ip (40mg, 0.24 mmol) combined with Cu(NO3)2·3H2O (171mg,<br />

0.70 mmol) was placed into a 20ml screw jaw. Subsequently, 7 ml DMF and 0.3 ml conc.<br />

HBF4 was added. The screw jaw was then sealed and put into a pre-heated oven at 80 °C<br />

for 20 h. The resulting blue powder products were collected by centrifugation.<br />

Afterwards, they were washed by DMF(30 ml x 3), EtOH (30 ml x 3) and acetone (30 ml x<br />

3) subsequently. The final powder was collected and dried under air at room temperature.<br />

The activation of the solids was performed by heating at 140 °C for 24 h under dynamic<br />

vacuum (ca. 10 -3 mbar).<br />

Figure 7.19. 1 H-NMR spectra of the acid-digest D5 sample (solvent: DMSO-d 6 , DCl).


Chapter 7 183<br />

D6<br />

This sample was obtained as D5 using 80 mg ip instead. [140] The resulting blue powder<br />

products were collected by centrifugation. Afterwards, they were washed by DMF(30 ml<br />

x 3), EtOH (30 ml x 3) and acetone (30 ml x 3) subsequently. The final powder was<br />

collected and dried under ambient atmosphere at room temperature. The activation of the<br />

solids was performed by heating at 140 °C for 24 h under dynamic vacuum (ca. 10 -3 mbar).<br />

Figure 7.20. 1 H-NMR spectra of the acid-digest D6 sample (solvent: DMSO-d 6 , DCl).


184 Chapter 7<br />

The synthesis of D7 and D8 are basically the same as D5 and D6, respectively. The only<br />

difference is that there is no usage of HBF4 during synthesis.<br />

Figure 7.21. 1 H-NMR spectra of the acid-digest D7 sample (solvent: DMSO-d 6 , DCl).<br />

Figure 7.22. 1 H-NMR spectra of the acid-digest D8 sample (solvent: DMSO-d 6 , DCl).


Chapter 7 185<br />

Figure 7.23. PXRD patterns of as-synthesized samples prepared from similar reactant with the<br />

sample D7 and D8 using EtOH instead of DMF (100 °C, 12h) in comparison with patterns<br />

simulated from [Cu 3 (BTC) 2 ] n (Cu-HKUST-1 sim.).


186 Chapter 7<br />

7.4 Experimental data on chapter 5<br />

All chemicals (Cu(NO3)2∙3H2O, Pd(OOCCH3)2, PdCl2, benzene-1,3,5-tricarboxylic acid<br />

(H3BTC), p-nitrophenol (PNP) (99%), Sodium borohydrate (NaBH4) (98%)) and all<br />

solvents (H2O, ethanol, and acetone) were used as commercially received unless<br />

otherwise noted. Before further manipulations, all activated samples were stored in an<br />

Ar-filled glove-box.<br />

7.4.1 Synthesis of [Cu3-XPdx(BTC)2]n<br />

Cu/Pd-BTC_1-3<br />

For Cu/Pd-BTC_1-3, 2.0 mmol of H3BTC (420 mg) was combined with 1.8 molar<br />

equivalents of a mixture of Cu(NO3)2∙3H2O and Pd acetate and then placed into a Teflon<br />

vessel. Subsequently 15ml of H2O/ethanol (1:1) solution was added. The molar<br />

percentage and amount of each metal salts in the total metal source as well as their weight<br />

used in the starting reactant are for Cu/Pd-BTC_1, Cu(NO3)2∙3H2O (90%, 3.24 mmol, 783<br />

mg) / Pd acetate (10%, 0.36 mmol. 80.8 mg); for Cu/Pd-BTC_2, Cu(NO3)2∙3H2O (85%,<br />

3.06 mmol, 739 mg) / Pd acetate (15%, 0.54 mmol. 121.2 mg); for Cu/Pd-BTC_3,<br />

Cu(NO3)2∙3H2O (80%, 2.88 mmol, 696 mg) / Pd acetate (20%, 0.72 mmol. 161.6 mg). The<br />

Teflon vessel was sealed in an autoclave and kept in the oven at 125 °C for 12h. The<br />

resulting power product was collected by filtration and washed several times with<br />

distilled water and acetone, respectively. After dried under ambient conditions the<br />

obtained powder was further activated by heating at 120 °C for 24 h under dynamic<br />

vacuum (ca. 10 -3 mbar).


Chapter 7 187<br />

Figure 7.24. XPS survey scans of the activated samples Cu/Pd-BTC_1 and 3. Figure is provided<br />

by P. Guo.<br />

Figure 7.25. PXRD patterns of the sample obtained using 50% feeding of Pd(II)-acetate under the<br />

same conditions as [Cu 3-X Pd x (BTC) 2 ] n with 10%, 15% and 20% feeding .


188 Chapter 7<br />

Figure 7.26. PXRD patterns of the sample prepared using only Pd(II)-acetate and H 3 btc under the<br />

same conditions as introduced mixed-metal solids with 10%, 15% and 20% feeding of Pd(II)-<br />

acetate. No Cu-salt was employed.


Chapter 7 189<br />

Cu/Pd-BTC_4<br />

For sample Cu/Pd-BTC_4, 0.48 mmol of H3BTC (101 mg) was combined with 1.5 molar<br />

equivalents of mixture of Cu(NO3)2∙3H2O (80%, 0.58 mmol, 139 mg) / PdCl2 (20%, 0.14<br />

mmol. 25.5 mg) and then placed into a Teflon vessel. Subsequently 6ml of H2O/ethanol<br />

(1:1) solution was added. The Teflon vessel was sealed in an autoclave and kept in the<br />

oven at 90 °C for 16h. The resulting power product was collected by filtration and washed<br />

several times with distilled water and acetone, respectively. After dried under ambient<br />

conditions the obtained powder was further activated by heating at 120 °C for 24 h under<br />

dynamic vacuum (ca. 10 -3 mbar).<br />

Figure 7.27. 1 H-NMR spectra of the acid-digest sample of Cu/Pd-BTC_4 (solvent: DMSO-d 6 , DCl).


190 Chapter 7<br />

7.4.2 Hydrogenation of PNP to PAP<br />

The catalytic aqueous-phase in-situ hydrogenation of PNP to PAP using NaBH4 as a<br />

hydrogen source was carried out in a 100 cm 3 batch reactor at room temperature (298K),<br />

ambient pressure and 1300 min -1 stirring speed. Figure 7.28 depicts the catalytic<br />

experimental setup. In the typical catalytic experiment, 5.0 cm 3 of an aqueous solution of<br />

NaBH4 (0.60 mmol dm −3 ) were added to 45.0 cm 3 of an aqueous solution of PNP (0.18<br />

mmol dm −3 ). The reaction was started by introducing 5.0 mg of the catalyst. PNP<br />

concentration changes were monitored via online UV-vis spectroscopy (AvaSpec-3648,<br />

optical path length: 5 mm) at a wavelength of 400 nm (see Figure 7.29). The initial<br />

reaction rate (r0) was calculated by assuming a zero order reaction from the linear<br />

decrease of PNP concentration after the addition of the catalyst.<br />

Figure 7.28. Experimental setup for aqueous-phase in-situ hydrogenation of PNP to PAP with<br />

NaBH 4 as hydrogen source using an online UV-vis spectrometer. Figure is provided by M. Al-Naji.


Absorbance, a. u.<br />

Chapter 7 191<br />

O 2<br />

N<br />

OH<br />

NaBH 4<br />

catalyst<br />

H 2<br />

N<br />

t reaction<br />

OH<br />

250 300 350 400 450 500<br />

Wavelength, nm<br />

Figure 7.29. UV-vis spectra for increasing reaction time in the aqueous-phase in-situ<br />

hydrogenation of PNP with NaBH 4 to PAP. Figure is provided by M. Al-Naji.


192 Chapter 7<br />

7.5 Supplementary details in Outlook<br />

Synthesis of mixed-component MOF (S1-5, [Cu3-XZnx(BTC)2-y(5-NO2-ip)y]n)<br />

3.6 mmol of a mixture of Cu(NO3)2∙3H2O (80%, 2.88 mmol, 696 mg) and Zn(NO3)2∙6H2O<br />

(20%, 0.72 mmol. 214 mg) was dissolved into a 15 ml solution of H2O/ethanol (1:1) in a<br />

Teflon vessel. Subsequently, mixed-linkers (H3BTC and 5-NO2-ip) in a total molar amount<br />

of 2 mmol (see Table 7.5 for the molar ratio differences for each sample) were added. After<br />

the Teflon vessel was sealed in an autoclave, it was kept in the oven at 125 °C for 12h. The<br />

resulting power product was collected by centrifuge and washed several times with<br />

distilled water and ethanol, respectively. Subsequently, the powder was soaked in the<br />

fresh ethanol for 24 h. After soaking the collected powder was dried under ambient<br />

conditions and further activated by heating at 120 °C for 24 h under dynamic vacuum (ca.<br />

10 -3 mbar).<br />

Table 7.5. Feeding molar ratios and amounts of the linkers used in the syntheses of the S1-5.<br />

samples<br />

H 3 BTC : (H 3 BTC + 5-NO 2 -ipH 2 )<br />

H 3 BTC 5-NO 2 -ipH 2<br />

Mass<br />

(mg)<br />

5-NO 2 -ipH 2 : (H 3 BTC + 5-<br />

NO 2 -ipH 2 )<br />

Mass<br />

(mg)<br />

S1 90% 378 10% 42<br />

S2 80% 336 20% 84<br />

S3 70% 294 30% 127<br />

S4 50% 210 50% 211<br />

S5 0 - 100% 422<br />

Scheme 7.4. Scheme illustration of preparation for mixed-component MOFs.


Chapter 7 193<br />

Figure 7.30. PXRD patterns of activated mixed-component MOFs S1-4 in comparison with parent<br />

Cu-BTC and only Zn-doped Cu/Zn-BTC.<br />

Table 7.6. The feeding molar percentage of Zn 2+ content and defect linker 5-NO 2 -ip during the<br />

synthesis of mixed-component sample S1-4 as well as the obtained values calculated from HPLC<br />

and AAS results.<br />

sample<br />

n(Zn 2+ ): n(Zn 2+ +Cu 2+ )<br />

n(5-NO 2 -ip): n(5-NO 2 -ip + H 3 BTC)<br />

feeding obtained feeding obtained<br />

S1 20% 3% 10% 3%<br />

S2 20% 2% 20% 10%<br />

S3 20% 2% 30% 14%<br />

S4 20% 1% 50% 21%


8 Bibliography<br />

[1] A. J. Ihde, The Development of Modern Chemistry, Harper and Row, 1964.<br />

[2] Y. Shibata, J. Coll. Sci., Imp. Univ. Tokyo 1916, 37, 1-17.<br />

[3] J. C. Bailar, Jr., Preparative Inorg. Reactions 1964, 1, 1-27.<br />

[4] N. G. Connelly, T. Damhus, R. M. Hartshorn, A. T. Hutton, Editors, Nomenclature of<br />

Inorganic Chemistry: IUPAC Recommendations 2005, Royal Society of Chemistry,<br />

2005.<br />

[5] R. Batten Stuart, R. Champness Neil, X.-M. Chen, J. Garcia-Martinez, S. Kitagawa, L.<br />

Öhrström, M. O’Keeffe, M. Paik Suh, J. Reedijk, in Pure Appl. Chem., Vol. 85, 2013, p.<br />

1715.<br />

[6] H. J. Choi, M. P. Suh, Inorg. Chem. 1999, 38, 6309-6312.<br />

[7] S. Kitagawa, R. Kitaura, S.-i. Noro, Angew. Chem. Int. Ed. 2004, 43, 2334-2375.<br />

[8] A. M. A. Ibrahim, E. Siebel, R. D. Fischer, Inorg. Chem. 1998, 37, 3521-3525.<br />

[9] A. J. Blake, N. R. Brooks, N. R. Champness, M. Crew, L. R. Hanton, P. Hubberstey, S.<br />

Parsons, M. Schroder, J. Chem. Soc., Dalton Trans. 1999, 2813-2817.<br />

[10] H. Okamoto, M. Yamashita, Bull. Chem. Soc. Jpn. 1998, 71, 2023-2039.<br />

[11] O. M. Yaghi, G. Li, Angew. Chem. Int. Ed. 1995, 34, 207-209.<br />

[12] B. Moulton, M. J. Zaworotko, Chem. Rev. 2001, 101, 1629-1658.<br />

[13] M. Eddaoudi, D. B. Moler, H. Li, B. Chen, T. M. Reineke, M. O'Keeffe, O. M. Yaghi, Acc.<br />

Chem. Res. 2001, 34, 319-330.<br />

[14] R. Murugavel, D. Krishnamurthy, M. Sathiyendiran, J. Chem. Soc., Dalton Trans.<br />

2002, 34-39.<br />

[15] S. S. Y. Chui, A. Siu, X. Feng, Z. Ying Zhang, T. C. W. Mak, I. D. Williams, Inorg. Chem.<br />

Commun. 2001, 4, 467-470.<br />

[16] J. Zhang, M. M. Matsushita, X. X. Kong, J. Abe, T. Iyoda, J. Am. Chem. Soc. 2001, 123,<br />

12105-12106.<br />

[17] G. J. E. Davidson, S. J. Loeb, Angew. Chem. Int. Ed. 2003, 42, 74-77.<br />

[18] H. Li, M. Eddaoudi, M. O'Keeffe, O. M. Yaghi, Nature 1999, 402, 276-279.<br />

[19] C. Janiak, Angew. Chem. Int. Ed. 1997, 36, 1431-1434.<br />

[20] R. Robson, J. Chem. Soc., Dalton Trans. 2000, 3735-3744.<br />

[21] M. Kondo, T. Yoshitomi, H. Matsuzaka, S. Kitagawa, K. Seki, Angew. Chem. Int. Ed.<br />

1997, 36, 1725-1727.<br />

[22] Z. R. Herm, E. D. Bloch, J. R. Long, Chem. Mater. 2014, 26, 323-338.<br />

[23] J.-R. Li, R. J. Kuppler, H.-C. Zhou, Chem. Soc. Rev. 2009, 38, 1477-1504.<br />

[24] L. J. Murray, M. Dinca, J. R. Long, Chem. Soc. Rev. 2009, 38, 1294-1314.<br />

[25] M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. O'Keeffe, O. M. Yaghi, Science<br />

2002, 295, 469-472.<br />

[26] J. A. Mason, M. Veenstra, J. R. Long, Chem. Sci. 2014, 5, 32-51.<br />

[27] D. Farrusseng, S. Aguado, C. Pinel, Angew. Chem. Int. Ed. 2009, 48, 7502-7513.<br />

[28] A. Corma, H. García, F. X. Llabrés i Xamena, Chem. Rev. 2010, 110, 4606-4655.<br />

[29] J. Liu, L. Chen, H. Cui, J. Zhang, L. Zhang, C.-Y. Su, Chem. Soc. Rev. 2014, 43, 6011-<br />

6061.


Bibliography 195<br />

[30] D. J. Tranchemontagne, J. L. Mendoza-Cortes, M. O'Keeffe, O. M. Yaghi, Chem. Soc.<br />

Rev. 2009, 38, 1257-1283.<br />

[31] K. S. Park, Z. Ni, A. P. Côté, J. Y. Choi, R. Huang, F. J. Uribe-Romo, H. K. Chae, M.<br />

O’Keeffe, O. M. Yaghi, Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 10186-10191.<br />

[32] A. Phan, C. J. Doonan, F. J. Uribe-Romo, C. B. Knobler, M. O’Keeffe, O. M. Yaghi, Acc.<br />

Chem. Res. 2010, 43, 58-67.<br />

[33] J.-P. Zhang, Y.-B. Zhang, J.-B. Lin, X.-M. Chen, Chem. Rev. 2012, 112, 1001-1033.<br />

[34] J. L. C. Rowsell, O. M. Yaghi, Microporous Mesoporous Mater. 2004, 73, 3-14.<br />

[35] S. S. Y. Chui, S. M. F. Lo, J. P. H. Charmant, A. G. Orpen, I. D. Williams, Science 1999,<br />

283, 1148-1150.<br />

[36] D. J. Tranchemontagne, J. R. Hunt, O. M. Yaghi, Tetrahedron 2008, 64, 8553-8557.<br />

[37] S. Horike, S. Shimomura, S. Kitagawa, Nat. Chem. 2009, 1, 695-704.<br />

[38] S. Kitagawa, M. Kondo, Bull. Chem. Soc. Jpn. 1998, 71, 1739-1753.<br />

[39] S. Henke, A. Schneemann, R. A. Fischer, Adv. Funct. Mater. 2013, 23, 5990-5996.<br />

[40] S. Kitagawa, K. Uemura, Chem. Soc. Rev. 2005, 34, 109-119.<br />

[41] I. Beurroies, M. Boulhout, P. L. Llewellyn, B. Kuchta, G. Férey, C. Serre, R. Denoyel,<br />

Angew. Chem. Int. Ed. 2010, 49, 7526-7529.<br />

[42] N. Yanai, T. Uemura, M. Inoue, R. Matsuda, T. Fukushima, M. Tsujimoto, S. Isoda, S.<br />

Kitagawa, J. Am. Chem. Soc. 2012, 134, 4501-4504.<br />

[43] G. Ferey, C. Serre, Chem. Soc. Rev. 2009, 38, 1380-1399.<br />

[44] T. Loiseau, C. Serre, C. Huguenard, G. Fink, F. Taulelle, M. Henry, T. Bataille, G. Férey,<br />

Chem. Eur. J. 2004, 10, 1373-1382.<br />

[45] F. Millange, N. Guillou, R. I. Walton, J.-M. Greneche, I. Margiolaki, G. Ferey, Chem.<br />

Commun. 2008, 4732-4734.<br />

[46] C. Serre, F. Millange, C. Thouvenot, M. Noguès, G. Marsolier, D. Louër, G. Férey, J.<br />

Am. Chem. Soc. 2002, 124, 13519-13526.<br />

[47] C. Volkringer, T. Loiseau, N. Guillou, G. Ferey, E. Elkaim, A. Vimont, Dalton Trans.<br />

2009, 2241-2249.<br />

[48] E. V. Anokhina, M. Vougo-Zanda, X. Wang, A. J. Jacobson, J. Am. Chem. Soc. 2005, 127,<br />

15000-15001.<br />

[49] J. P. S. Mowat, V. R. Seymour, J. M. Griffin, S. P. Thompson, A. M. Z. Slawin, D. Fairen-<br />

Jimenez, T. Duren, S. E. Ashbrook, P. A. Wright, Dalton Trans. 2012, 41, 3937-3941.<br />

[50] H. Chun, D. N. Dybtsev, H. Kim, K. Kim, Chem. Eur. J. 2005, 11, 3521-3529.<br />

[51] K. Uemura, Y. Yamasaki, Y. Komagawa, K. Tanaka, H. Kita, Angew. Chem. Int. Ed.<br />

2007, 46, 6662-6665.<br />

[52] B. Chen, S. Ma, E. J. Hurtado, E. B. Lobkovsky, C. Liang, H. Zhu, S. Dai, Inorg. Chem.<br />

2007, 46, 8705-8709.<br />

[53] G. Ferey, Chem. Soc. Rev. 2008, 37, 191-214.<br />

[54] J. H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga, K. P. Lillerud,<br />

J. Am. Chem. Soc. 2008, 130, 13850-13851.<br />

[55] M. Kandiah, M. H. Nilsen, S. Usseglio, S. Jakobsen, U. Olsbye, M. Tilset, C. Larabi, E.<br />

A. Quadrelli, F. Bonino, K. P. Lillerud, Chem. Mater. 2010, 22, 6632-6640.<br />

[56] V. Colombo, S. Galli, H. J. Choi, G. D. Han, A. Maspero, G. Palmisano, N. Masciocchi, J.<br />

R. Long, Chem. Sci. 2011, 2, 1311-1319.<br />

[57] N. Stock, S. Biswas, Chem. Rev. 2012, 112, 933-969.<br />

[58] A. Rabenau, Angew. Chem. Int. Ed. 1985, 24, 1026-1040.<br />

[59] S. T. Meek, J. A. Greathouse, M. D. Allendorf, Adv. Mater. (Weinheim, Ger.) 2011, 23,<br />

249-267.<br />

[60] O. M. Yaghi, M. O'Keeffe, N. W. Ockwig, H. K. Chae, M. Eddaoudi, J. Kim, Nature 2003,<br />

423, 705-714.


196 Bibliography<br />

[61] S. Ma, D. Sun, M. Ambrogio, J. A. Fillinger, S. Parkin, H.-C. Zhou, J. Am. Chem. Soc.<br />

2007, 129, 1858-1859.<br />

[62] H. Furukawa, Y. B. Go, N. Ko, Y. K. Park, F. J. Uribe-Romo, J. Kim, M. O’Keeffe, O. M.<br />

Yaghi, Inorg. Chem. 2011, 50, 9147-9152.<br />

[63] N. L. Rosi, J. Kim, M. Eddaoudi, B. Chen, M. O'Keeffe, O. M. Yaghi, J. Am. Chem. Soc.<br />

2005, 127, 1504-1518.<br />

[64] H. Deng, S. Grunder, K. E. Cordova, C. Valente, H. Furukawa, M. Hmadeh, F. Gándara,<br />

A. C. Whalley, Z. Liu, S. Asahina, H. Kazumori, M. O’Keeffe, O. Terasaki, J. F. Stoddart,<br />

O. M. Yaghi, Science 2012, 336, 1018-1023.<br />

[65] D. Gygi, E. D. Bloch, J. A. Mason, M. R. Hudson, M. I. Gonzalez, R. L. Siegelman, T. A.<br />

Darwish, W. L. Queen, C. M. Brown, J. R. Long, Chem. Mater. 2016, 28, 1128-1138.<br />

[66] H. Furukawa, K. E. Cordova, M. O’Keeffe, O. M. Yaghi, Science 2013, 341.<br />

[67] R. Babarao, J. Jiang, Langmuir 2008, 24, 6270-6278.<br />

[68] S. Hasegawa, S. Horike, R. Matsuda, S. Furukawa, K. Mochizuki, Y. Kinoshita, S.<br />

Kitagawa, J. Am. Chem. Soc. 2007, 129, 2607-2614.<br />

[69] J. He, K.-K. Yee, Z. Xu, M. Zeller, A. D. Hunter, S. S.-Y. Chui, C.-M. Che, Chem. Mater.<br />

2011, 23, 2940-2947.<br />

[70] M. Dincă, J. R. Long, Angew. Chem. Int. Ed. 2008, 47, 6766-6779.<br />

[71] A. Schneemann, S. Henke, I. Schwedler, R. A. Fischer, ChemPhysChem 2014, 15, 823-<br />

839.<br />

[72] M. E. Kosal, J.-H. Chou, S. R. Wilson, K. S. Suslick, Nat. Mater. 2002, 1, 118-121.<br />

[73] D. W. Smithenry, S. R. Wilson, K. S. Suslick, Inorg. Chem. 2003, 42, 7719-7721.<br />

[74] R. Kitaura, G. Onoyama, H. Sakamoto, R. Matsuda, S.-i. Noro, S. Kitagawa, Angew.<br />

Chem. Int. Ed. 2004, 43, 2684-2687.<br />

[75] K. C. Szeto, K. O. Kongshaug, S. Jakobsen, M. Tilset, K. P. Lillerud, Dalton Trans. 2008,<br />

2054-2060.<br />

[76] S. S. Kaye, J. R. Long, J. Am. Chem. Soc. 2008, 130, 806-807.<br />

[77] J. I. Feldblyum, M. Liu, D. W. Gidley, A. J. Matzger, J. Am. Chem. Soc. 2011, 133, 18257-<br />

18263.<br />

[78] L. J. Murray, M. Dinca, J. Yano, S. Chavan, S. Bordiga, C. M. Brown, J. R. Long, J. Am.<br />

Chem. Soc. 2010, 132, 7856-7857.<br />

[79] P. Maniam, N. Stock, Inorg. Chem. 2011, 50, 5085-5097.<br />

[80] L. Xie, S. Liu, C. Gao, R. Cao, J. Cao, C. Sun, Z. Su, Inorg. Chem. 2007, 46, 7782-7788.<br />

[81] M. Kramer, U. Schwarz, S. Kaskel, J. Mater. Chem. 2006, 16, 2245-2248.<br />

[82] O. Kozachuk, K. Yusenko, H. Noei, Y. Wang, S. Walleck, T. Glaser, R. A. Fischer, Chem.<br />

Commun. 2011, 47, 8509-8511.<br />

[83] C. R. Wade, M. Dinca, Dalton Trans. 2012, 41, 7931-7938.<br />

[84] W. Zhou, H. Wu, T. Yildirim, J. Am. Chem. Soc. 2008, 130, 15268-15269.<br />

[85] E. D. Bloch, L. J. Murray, W. L. Queen, S. Chavan, S. N. Maximoff, J. P. Bigi, R. Krishna,<br />

V. K. Peterson, F. Grandjean, G. J. Long, B. Smit, S. Bordiga, C. M. Brown, J. R. Long, J.<br />

Am. Chem. Soc. 2011, 133, 14814-14822.<br />

[86] M. Díaz-García, M. Sánchez-Sánchez, Microporous Mesoporous Mater. 2014, 190,<br />

248-254.<br />

[87] R. Sanz, F. Martinez, G. Orcajo, L. Wojtas, D. Briones, Dalton Trans. 2013, 42, 2392-<br />

2398.<br />

[88] P. D. C. Dietzel, Y. Morita, R. Blom, H. Fjellvåg, Angew. Chem. Int. Ed. 2005, 44, 6354-<br />

6358.<br />

[89] M. Dincǎ, A. Dailly, Y. Liu, C. M. Brown, D. A. Neumann, J. R. Long, J. Am. Chem. Soc.<br />

2006, 128, 16876-16883.


Bibliography 197<br />

[90] W. Ouellette, K. Darling, A. Prosvirin, K. Whitenack, K. R. Dunbar, J. Zubieta, Dalton<br />

Trans. 2011, 40, 12288-12300.<br />

[91] M. Dincă, W. S. Han, Y. Liu, A. Dailly, C. M. Brown, J. R. Long, Angew. Chem. Int. Ed.<br />

2007, 46, 1419-1422.<br />

[92] K. Sumida, S. Horike, S. S. Kaye, Z. R. Herm, W. L. Queen, C. M. Brown, F. Grandjean,<br />

G. J. Long, A. Dailly, J. R. Long, Chem. Sci. 2010, 1, 184-191.<br />

[93] Z. Wang, S. M. Cohen, J. Am. Chem. Soc. 2007, 129, 12368-12369.<br />

[94] J. S. Costa, P. Gamez, C. A. Black, O. Roubeau, S. J. Teat, J. Reedijk, Eur. J. Inorg. Chem.<br />

2008, 2008, 1551-1554.<br />

[95] T. Haneda, M. Kawano, T. Kawamichi, M. Fujita, J. Am. Chem. Soc. 2008, 130, 1578-<br />

1579.<br />

[96] M. J. Ingleson, J. Perez Barrio, J.-B. Guilbaud, Y. Z. Khimyak, M. J. Rosseinsky, Chem.<br />

Commun. 2008, 2680-2682.<br />

[97] Y.-F. Song, L. Cronin, Angew. Chem. Int. Ed. 2008, 47, 4635-4637.<br />

[98] K. K. Tanabe, C. A. Allen, S. M. Cohen, Angew. Chem. Int. Ed. 2010, 49, 9730-9733.<br />

[99] A. D. Burrows, C. G. Frost, M. F. Mahon, C. Richardson, Angew. Chem. Int. Ed. 2008,<br />

47, 8482-8486.<br />

[100] Z. Wang, S. M. Cohen, Chem. Soc. Rev. 2009, 38, 1315-1329.<br />

[101] S. M. Cohen, Chem. Rev. 2012, 112, 970-1000.<br />

[102] T. Yamada, H. Kitagawa, J. Am. Chem. Soc. 2009, 131, 6312-6313.<br />

[103] D. J. Lun, G. I. N. Waterhouse, S. G. Telfer, J. Am. Chem. Soc. 2011, 133, 5806-5809.<br />

[104] H. Sato, R. Matsuda, K. Sugimoto, M. Takata, S. Kitagawa, Nat. Mater. 2010, 9, 661-<br />

666.<br />

[105] R. K. Deshpande, J. L. Minnaar, S. G. Telfer, Angew. Chem. Int. Ed. 2010, 49, 4598-<br />

4602.<br />

[106] O. K. Farha, K. L. Mulfort, J. T. Hupp, Inorg. Chem. 2008, 47, 10223-10225.<br />

[107] Y. K. Hwang, D.-Y. Hong, J.-S. Chang, S. H. Jhung, Y.-K. Seo, J. Kim, A. Vimont, M.<br />

Daturi, C. Serre, G. Férey, Angew. Chem. Int. Ed. 2008, 47, 4144-4148.<br />

[108] E. D. Bloch, D. Britt, C. Lee, C. J. Doonan, F. J. Uribe-Romo, H. Furukawa, J. R. Long, O.<br />

M. Yaghi, J. Am. Chem. Soc. 2010, 132, 14382-14384.<br />

[109] O. Karagiaridi, M. B. Lalonde, W. Bury, A. A. Sarjeant, O. K. Farha, J. T. Hupp, J. Am.<br />

Chem. Soc. 2012, 134, 18790-18796.<br />

[110] O. Karagiaridi, W. Bury, E. Tylianakis, A. A. Sarjeant, J. T. Hupp, O. K. Farha, Chem.<br />

Mater. 2013, 25, 3499-3503.<br />

[111] O. Karagiaridi, W. Bury, A. A. Sarjeant, C. L. Stern, O. K. Farha, J. T. Hupp, Chem. Sci.<br />

2012, 3, 3256-3260.<br />

[112] N. A. Vermeulen, O. Karagiaridi, A. A. Sarjeant, C. L. Stern, J. T. Hupp, O. K. Farha, J.<br />

F. Stoddart, J. Am. Chem. Soc. 2013, 135, 14916-14919.<br />

[113] W. Bury, D. Fairen-Jimenez, M. B. Lalonde, R. Q. Snurr, O. K. Farha, J. T. Hupp, Chem.<br />

Mater. 2013, 25, 739-744.<br />

[114] B. J. Burnett, P. M. Barron, C. Hu, W. Choe, J. Am. Chem. Soc. 2011, 133, 9984-9987.<br />

[115] M. Kim, J. F. Cahill, Y. Su, K. A. Prather, S. M. Cohen, Chem. Sci. 2012, 3, 126-130.<br />

[116] S. Kim, K. W. Dawson, B. S. Gelfand, J. M. Taylor, G. K. H. Shimizu, J. Am. Chem. Soc.<br />

2013, 135, 963-966.<br />

[117] P. Deria, J. E. Mondloch, O. Karagiaridi, W. Bury, J. T. Hupp, O. K. Farha, Chem. Soc.<br />

Rev. 2014, 43, 5896-5912.<br />

[118] M. Lalonde, W. Bury, O. Karagiaridi, Z. Brown, J. T. Hupp, O. K. Farha, J. Mater. Chem.<br />

A, 2013, 1, 5453-5468.<br />

[119] S. Das, H. Kim, K. Kim, J. Am. Chem. Soc. 2009, 131, 3814-3815.<br />

[120] M. Dincǎ, J. R. Long, J. Am. Chem. Soc. 2007, 129, 11172-11176.


198 Bibliography<br />

[121] T. K. Prasad, D. H. Hong, M. P. Suh, Chem. Eur. J. 2010, 16, 14043-14050.<br />

[122] J. Kahr, R. E. Morris, P. A. Wright, CrystEngComm 2013, 15, 9779-9786.<br />

[123] H. Deng, C. J. Doonan, H. Furukawa, R. B. Ferreira, J. Towne, C. B. Knobler, B. Wang,<br />

O. M. Yaghi, Science 2010, 327, 846-850.<br />

[124] W. Kleist, F. Jutz, M. Maciejewski, A. Baiker, Eur. J. Inorg. Chem. 2009, 2009, 3552-<br />

3561.<br />

[125] S. Marx, W. Kleist, J. Huang, M. Maciejewski, A. Baiker, Dalton Trans. 2010, 39, 3795-<br />

3798.<br />

[126] A. D. Burrows, CrystEngComm 2011, 13, 3623-3642.<br />

[127] D. N. Bunck, W. R. Dichtel, Chem. Eur. J. 2013, 19, 818-827.<br />

[128] S. Henke, A. Schneemann, A. Wütscher, R. A. Fischer, J. Am. Chem. Soc. 2012, 134,<br />

9464-9474.<br />

[129] J. A. Villajos, G. Orcajo, C. Martos, J. Á. Botas, J. Villacañas, G. Calleja, Int. J. Hydrogen<br />

Energy 2015, 40, 5346-5352.<br />

[130] J. A. Botas, G. Calleja, M. Sánchez-Sánchez, M. G. Orcajo, Langmuir 2010, 26, 5300-<br />

5303.<br />

[131] O. Kozachuk, K. Khaletskaya, M. Halbherr, A. Bétard, M. Meilikhov, R. W. Seidel, B.<br />

Jee, A. Pöppl, R. A. Fischer, Eur. J. Inorg. Chem. 2012, 2012, 1688-1695.<br />

[132] O. Kozachuk, M. Meilikhov, K. Yusenko, A. Schneemann, B. Jee, A. V. Kuttatheyil, M.<br />

Bertmer, C. Sternemann, A. Pöppl, R. A. Fischer, Eur. J. Inorg. Chem. 2013, 2013,<br />

4546-4557.<br />

[133] F. Gul-E-Noor, B. Jee, M. Mendt, D. Himsl, A. Pöppl, M. Hartmann, J. Haase, H.<br />

Krautscheid, M. Bertmer, J. Phys. Chem. C 2012, 116, 20866-20873.<br />

[134] M. A. Gotthardt, R. Schoch, S. Wolf, M. Bauer, W. Kleist, Dalton Trans. 2015, 44,<br />

2052-2056.<br />

[135] F. Nouar, M. I. Breeze, B. C. Campo, A. Vimont, G. Clet, M. Daturi, T. Devic, R. I.<br />

Walton, C. Serre, Chem. Commun. (Cambridge, U. K.) 2015, 51, 14458-14461.<br />

[136] S. Marx, W. Kleist, A. Baiker, J. Catal. 2011, 281, 76-87.<br />

[137] Z. Fang, B. Bueken, D. E. De Vos, R. A. Fischer, Angew. Chem. Int. Ed. 2015, 54, 7234-<br />

7254.<br />

[138] O. Kozachuk, I. Luz, F. X. Llabrés i Xamena, H. Noei, M. Kauer, H. B. Albada, E. D.<br />

Bloch, B. Marler, Y. Wang, M. Muhler, R. A. Fischer, Angew. Chem. Int. Ed. 2014, 53,<br />

7058-7062.<br />

[139] Z. Fang, J. P. Dürholt, M. Kauer, W. Zhang, C. Lochenie, B. Jee, B. Albada, N. Metzler-<br />

Nolte, A. Pöppl, B. Weber, M. Muhler, Y. Wang, R. Schmid, R. A. Fischer, J. Am. Chem.<br />

Soc. 2014, 136, 9627-9636.<br />

[140] G. Barin, V. Krungleviciute, O. Gutov, J. T. Hupp, T. Yildirim, O. K. Farha, Inorg. Chem.<br />

2014, 53, 6914-6919.<br />

[141] F. Vermoortele, B. Bueken, G. Le Bars, B. Van de Voorde, M. Vandichel, K. Houthoofd,<br />

A. Vimont, M. Daturi, M. Waroquier, V. Van Speybroeck, C. Kirschhock, D. E. De Vos,<br />

J. Am. Chem. Soc. 2013, 135, 11465-11468.<br />

[142] T.-H. Park, A. J. Hickman, K. Koh, S. Martin, A. G. Wong-Foy, M. S. Sanford, A. J.<br />

Matzger, J. Am. Chem. Soc. 2011, 133, 20138-20141.<br />

[143] O. V. Gutov, M. G. Hevia, E. C. Escudero-Adán, A. Shafir, Inorg. Chem. 2015, 54, 8396-<br />

8400.<br />

[144] U. Ravon, M. Savonnet, S. Aguado, M. E. Domine, E. Janneau, D. Farrusseng,<br />

Microporous Mesoporous Mater. 2010, 129, 319-329.<br />

[145] N. Al-Janabi, X. Fan, F. R. Siperstein, The Journal of Physical Chemistry Letters 2016,<br />

7, 1490-1494.


Bibliography 199<br />

[146] M. J. Cliffe, W. Wan, X. Zou, P. A. Chater, A. K. Kleppe, M. G. Tucker, H. Wilhelm, N. P.<br />

Funnell, F.-X. Coudert, A. L. Goodwin, Nat. Commun. 2014, 5.<br />

[147] R. J. Kuppler, D. J. Timmons, Q.-R. Fang, J.-R. Li, T. A. Makal, M. D. Young, D. Yuan, D.<br />

Zhao, W. Zhuang, H.-C. Zhou, Coord. Chem. Rev. 2009, 253, 3042-3066.<br />

[148] S. S. Kaye, A. Dailly, O. M. Yaghi, J. R. Long, J. Am. Chem. Soc. 2007, 129, 14176-14177.<br />

[149] S. Ma, J. Eckert, P. M. Forster, J. W. Yoon, Y. K. Hwang, J.-S. Chang, C. D. Collier, J. B.<br />

Parise, H.-C. Zhou, J. Am. Chem. Soc. 2008, 130, 15896-15902.<br />

[150] Y. Li, R. T. Yang, J. Am. Chem. Soc. 2006, 128, 8136-8137.<br />

[151] D. F. Quinn, J. A. MacDonald, K. Sosin, Prepr. Pap. - Am. Chem. Soc., Div. Fuel Chem.<br />

1994, 39, 451-455.<br />

[152] K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. R. Herm, T.-H.<br />

Bae, J. R. Long, Chem. Rev. 2012, 112, 724-781.<br />

[153] S. Ma, X.-S. Wang, D. Yuan, H.-C. Zhou, Angew. Chem. Int. Ed. 2008, 47, 4130-4133.<br />

[154] M. Dincǎ, A. F. Yu, J. R. Long, J. Am. Chem. Soc. 2006, 128, 8904-8913.<br />

[155] R. Kitaura, K. Seki, G. Akiyama, S. Kitagawa, Angew. Chem. 2003, 115, 444-447.<br />

[156] Y. E. Cheon, M. P. Suh, Chem. Eur. J. 2008, 14, 3961-3967.<br />

[157] B. Chen, C. Liang, J. Yang, D. S. Contreras, Y. L. Clancy, E. B. Lobkovsky, O. M. Yaghi,<br />

S. Dai, Angew. Chem. Int. Ed. 2006, 45, 1390-1393.<br />

[158] A. Schneemann, E. D. Bloch, S. Henke, P. L. Llewellyn, J. R. Long, R. A. Fischer, Chem.<br />

Eur. J. 2015, 21, 18764-18769.<br />

[159] S. Chaemchuen, N. A. Kabir, K. Zhou, F. Verpoort, Chem. Soc. Rev. 2013, 42, 9304-<br />

9332.<br />

[160] S. Horike, M. Dincǎ, K. Tamaki, J. R. Long, J. Am. Chem. Soc. 2008, 130, 5854-5855.<br />

[161] J. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen, J. T. Hupp, Chem. Soc. Rev.<br />

2009, 38, 1450-1459.<br />

[162] A. Henschel, K. Gedrich, R. Kraehnert, S. Kaskel, Chem. Commun. 2008, 4192-4194.<br />

[163] K. Schlichte, T. Kratzke, S. Kaskel, Microporous Mesoporous Mater. 2004, 73, 81-88.<br />

[164] M. Yoon, R. Srirambalaji, K. Kim, Chem. Rev. 2012, 112, 1196-1231.<br />

[165] G. Nickerl, A. Henschel, R. Grünker, K. Gedrich, S. Kaskel, Chem. Ing. Tech. 2011, 83,<br />

90-103.<br />

[166] L. Ma, C. Abney, W. Lin, Chem. Soc. Rev. 2009, 38, 1248-1256.<br />

[167] M. Zhao, S. Ou, C.-D. Wu, Acc. Chem. Res. 2014, 47, 1199-1207.<br />

[168] Z.-Y. Gu, J. Park, A. Raiff, Z. Wei, H.-C. Zhou, ChemCatChem 2014, 6, 67-75.<br />

[169] J.-L. Wang, C. Wang, W. Lin, ACS catal. 2012, 2, 2630-2640.<br />

[170] H. Garcia, B. Ferrer, in Metal Organic Frameworks as Heterogeneous Catalysts, The<br />

Royal Society of Chemistry, 2013, pp. 365-383.<br />

[171] D. Maspoch, D. Ruiz-Molina, J. Veciana, J. Mater. Chem. 2004, 14, 2713-2723.<br />

[172] E. Pardo, R. Ruiz-Garcia, J. Cano, X. Ottenwaelder, R. Lescouezec, Y. Journaux, F.<br />

Lloret, M. Julve, Dalton Trans. 2008, 2780-2805.<br />

[173] M. Kurmoo, Chem. Soc. Rev. 2009, 38, 1353-1379.<br />

[174] D. Umeyama, S. Horike, M. Inukai, Y. Hijikata, S. Kitagawa, Angew. Chem. Int. Ed.<br />

2011, 50, 11706-11709.<br />

[175] S. Bureekaew, S. Horike, M. Higuchi, M. Mizuno, T. Kawamura, D. Tanaka, N. Yanai,<br />

S. Kitagawa, Nat. Mater. 2009, 8, 831-836.<br />

[176] J. A. Hurd, R. Vaidhyanathan, V. Thangadurai, C. I. Ratcliffe, I. L. Moudrakovski, G. K.<br />

H. Shimizu, Nat. Chem. 2009, 1, 705-710.<br />

[177] A. Shigematsu, T. Yamada, H. Kitagawa, J. Am. Chem. Soc. 2011, 133, 2034-2036.<br />

[178] M. Yoon, K. Suh, S. Natarajan, K. Kim, Angew. Chem. Int. Ed. 2013, 52, 2688-2700.<br />

[179] F. Gándara, F. J. Uribe-Romo, D. K. Britt, H. Furukawa, L. Lei, R. Cheng, X. Duan, M.<br />

O'Keeffe, O. M. Yaghi, Chem. Eur. J. 2012, 18, 10595-10601.


200 Bibliography<br />

[180] Y. Kobayashi, B. Jacobs, M. D. Allendorf, J. R. Long, Chem. Mater. 2010, 22, 4120-<br />

4122.<br />

[181] A. A. Talin, A. Centrone, A. C. Ford, M. E. Foster, V. Stavila, P. Haney, R. A. Kinney, V.<br />

Szalai, F. El Gabaly, H. P. Yoon, F. Léonard, M. D. Allendorf, Science 2014, 343, 66-<br />

69.<br />

[182] C. Janiak, Dalton Trans. 2003, 2781-2804.<br />

[183] M. D. Allendorf, C. A. Bauer, R. K. Bhakta, R. J. T. Houk, Chem. Soc. Rev. 2009, 38,<br />

1330-1352.<br />

[184] B. V. Harbuzaru, A. Corma, F. Rey, P. Atienzar, J. L. Jordá, H. García, D. Ananias, L. D.<br />

Carlos, J. Rocha, Angew. Chem. Int. Ed. 2008, 47, 1080-1083.<br />

[185] M. D. Allendorf, R. J. T. Houk, L. Andruszkiewicz, A. A. Talin, J. Pikarsky, A.<br />

Choudhury, K. A. Gall, P. J. Hesketh, J. Am. Chem. Soc. 2008, 130, 14404-14405.<br />

[186] B. Chen, Y. Yang, F. Zapata, G. Lin, G. Qian, E. B. Lobkovsky, Adv. Mater. (Weinheim,<br />

Ger.) 2007, 19, 1693-1696.<br />

[187] P. Horcajada, C. Serre, G. Maurin, N. A. Ramsahye, F. Balas, M. Vallet-Regí, M.<br />

Sebban, F. Taulelle, G. Férey, J. Am. Chem. Soc. 2008, 130, 6774-6780.<br />

[188] P. Horcajada, C. Serre, M. Vallet-Regí, M. Sebban, F. Taulelle, G. Férey, Angew. Chem.<br />

Int. Ed. 2006, 45, 5974-5978.<br />

[189] K. Brown, S. Zolezzi, P. Aguirre, D. Venegas-Yazigi, V. Paredes-Garcia, R. Baggio, M.<br />

A. Novak, E. Spodine, Dalton Trans. 2009, 1422-1427.<br />

[190] A. Dhakshinamoorthy, M. Alvaro, H. Garcia, J. Catal. 2009, 267, 1-4.<br />

[191] S. Xiang, W. Zhou, J. M. Gallegos, Y. Liu, B. Chen, J. Am. Chem. Soc. 2009, 131, 12415-<br />

12419.<br />

[192] S. Bordiga, L. Regli, F. Bonino, E. Groppo, C. Lamberti, B. Xiao, P. S. Wheatley, R. E.<br />

Morris, A. Zecchina, Phys. Chem. Chem. Phys. 2007, 9, 2676-2685.<br />

[193] D. Y. Lee, E.-K. Kim, C. Y. Shin, D. V. Shinde, W. Lee, N. K. Shrestha, J. K. Lee, S.-H. Han,<br />

RSC Adv. 2014, 4, 12037-12042.<br />

[194] U. Mueller, M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndt, J. Pastre, J. Mater.<br />

Chem. 2006, 16, 626-636.<br />

[195] Y.-K. Seo, G. Hundal, I. T. Jang, Y. K. Hwang, C.-H. Jun, J.-S. Chang, Microporous<br />

Mesoporous Mater. 2009, 119, 331-337.<br />

[196] M. Schlesinger, S. Schulze, M. Hietschold, M. Mehring, Microporous Mesoporous<br />

Mater. 2010, 132, 121-127.<br />

[197] O. Shekhah, H. Wang, S. Kowarik, F. Schreiber, M. Paulus, M. Tolan, C. Sternemann,<br />

F. Evers, D. Zacher, R. A. Fischer, C. Wöll, J. Am. Chem. Soc. 2007, 129, 15118-15119.<br />

[198] H. Noei, O. Kozachuk, S. Amirjalayer, S. Bureekaew, M. Kauer, R. Schmid, B. Marler,<br />

M. Muhler, R. A. Fischer, Y. Wang, J. Phys. Chem. C 2013, 117, 5658-5666.<br />

[199] M. McCann, A. Carvill, C. Cardin, M. Convery, Polyhedron 1993, 12, 1163-1169.<br />

[200] F. A. Urbanos, M. C. Barral, R. Jiménez-Aparicio, Polyhedron 1988, 7, 2597-2600.<br />

[201] M. C. Barral, R. Jimenez-Aparicio, J. L. Priego, E. C. Royer, M. J. Saucedo, F. A.<br />

Urbanos, U. Amador, J. Chem. Soc., Dalton Trans. 1995, 2183-2187.<br />

[202] R. W. Mitchell, A. Spencer, G. Wilkinson, J. Chem. Soc., Dalton Trans. 1973, 846-854.<br />

[203] T. Togano, M. Mukaida, T. Nomura, Bull. Chem. Soc. Jpn. 1980, 53, 2085-2086.<br />

[204] F. D. Cukiernik, A.-M. Giroud-Godquin, P. Maldivi, J.-C. Marchon, Inorg. Chim. Acta<br />

1994, 215, 203-207.<br />

[205] T. Düren, F. Millange, G. Férey, K. S. Walton, R. Q. Snurr, J. Phys. Chem. C 2007, 111,<br />

15350-15356.<br />

[206] D. Rose, G. Wilkinson, J. Chem. Soc. A 1970, 1791-1795.<br />

[207] A. J. Lindsay, G. Wilkinson, M. Motevalli, M. B. Hursthouse, J. Chem. Soc., Dalton<br />

Trans. 1985, 2321-2326.


Bibliography 201<br />

[208] W. Zhang, O. Kozachuk, R. Medishetty, A. Schneemann, R. Wagner, K. Khaletskaya,<br />

K. Epp, R. A. Fischer, Eur. J. Inorg. Chem. 2015, 2015, 3913-3920.<br />

[209] T. Fukushima, S. Horike, Y. Inubushi, K. Nakagawa, Y. Kubota, M. Takata, S.<br />

Kitagawa, Angew. Chem. Int. Ed. 2010, 49, 4820-4824.<br />

[210] S. Horike, Y. Inubushi, T. Hori, T. Fukushima, S. Kitagawa, Chem. Sci. 2012, 3, 116-<br />

120.<br />

[211] M. Lin Foo, S. Horike, T. Fukushima, Y. Hijikata, Y. Kubota, M. Takata, S. Kitagawa,<br />

Dalton Trans. 2012, 41, 13791-13794.<br />

[212] W. Kleist, M. Maciejewski, A. Baiker, Thermochim. Acta 2010, 499, 71-78.<br />

[213] T. Tsuruoka, S. Furukawa, Y. Takashima, K. Yoshida, S. Isoda, S. Kitagawa, Angew.<br />

Chem. Int. Ed. 2009, 48, 4739-4743.<br />

[214] A. Umemura, S. Diring, S. Furukawa, H. Uehara, T. Tsuruoka, S. Kitagawa, J. Am.<br />

Chem. Soc. 2011, 133, 15506-15513.<br />

[215] K. M. Choi, H. J. Jeon, J. K. Kang, O. M. Yaghi, J. Am. Chem. Soc. 2011, 133, 11920-<br />

11923.<br />

[216] K. Koh, A. G. Wong-Foy, A. J. Matzger, J. Am. Chem. Soc. 2009, 131, 4184-4185.<br />

[217] K. Koh, A. G. Wong-Foy, A. J. Matzger, Angew. Chem. Int. Ed. 2008, 47, 677-680.<br />

[218] K. Koh, A. G. Wong-Foy, A. J. Matzger, J. Am. Chem. Soc. 2010, 132, 15005-15010.<br />

[219] Y. Liao, S. K. Yang, K. Koh, A. J. Matzger, J. S. Biteen, Nano Lett. 2012, 12, 3080-3085.<br />

[220] R. A. Beyerlein, C. Choi-Feng, J. B. Hall, B. J. Huggins, G. J. Ray, Top. Catal., 4, 27-42.<br />

[221] C. Baerlocher, D. Xie, L. B. McCusker, S.-J. Hwang, I. Y. Chan, K. Ong, A. W. Burton, S.<br />

I. Zones, Nat. Mater. 2008, 7, 631-635.<br />

[222] B. Bonelli, L. Forni, A. Aloise, J. B. Nagy, G. Fornasari, E. Garrone, A. Gedeon, G.<br />

Giordano, F. Trifirò, Microporous Mesoporous Mater. 2007, 101, 153-160.<br />

[223] F. Schröder, D. Esken, M. Cokoja, M. W. E. van den Berg, O. I. Lebedev, G. Van<br />

Tendeloo, B. Walaszek, G. Buntkowsky, H.-H. Limbach, B. Chaudret, R. A. Fischer, J.<br />

Am. Chem. Soc. 2008, 130, 6119-6130.<br />

[224] E. A. Braude, F. C. Nachod, Editors, Determination of Organic Structures by Physical<br />

Methods, Academic Press, 1955.<br />

[225] F. Genin, F. Quiles, A. Burneau, Phys. Chem. Chem. Phys. 2001, 3, 932-942.<br />

[226] K. Nakamoto, in Infrared and Raman Spectra of Inorganic and Coordination<br />

Compounds, John Wiley & Sons, Inc., 2008, pp. 1-273.<br />

[227] Z. Wang, J. Liu, B. Lukose, Z. Gu, P. G. Weidler, H. Gliemann, T. Heine, C. Wöll, Nano<br />

Lett. 2014, 14, 1526-1529.<br />

[228] K. S. W. Sing, D. H. Everett, R. A. W. Haul, L. Moscou, R. A. Pierotti, J. Rouquerol, T.<br />

Siemieniewska, Pure Appl. Chem. 1985, 57, 603-619.<br />

[229] R. C. Lochan, R. Z. Khaliullin, M. Head-Gordon, Inorg. Chem. 2008, 47, 4032-4044.<br />

[230] R. C. Lochan, M. Head-Gordon, Phys. Chem. Chem. Phys. 2006, 8, 1357-1370.<br />

[231] S. T. Madrahimov, J. R. Gallagher, G. Zhang, Z. Meinhart, S. J. Garibay, M. Delferro, J.<br />

T. Miller, O. K. Farha, J. T. Hupp, S. T. Nguyen, ACS catal. 2015, 5, 6713-6718.<br />

[232] D. Yang, S. O. Odoh, J. Borycz, T. C. Wang, O. K. Farha, J. T. Hupp, C. J. Cramer, L.<br />

Gagliardi, B. C. Gates, ACS catal. 2016, 6, 235-247.<br />

[233] J. Canivet, S. Aguado, Y. Schuurman, D. Farrusseng, J. Am. Chem. Soc. 2013, 135,<br />

4195-4198.<br />

[234] E. D. Metzger, C. K. Brozek, R. J. Comito, M. Dincă, ACS Cent. Sci. 2016.<br />

[235] I. Ogino, B. C. Gates, Chem. Eur. J. 2009, 15, 6827-6837.<br />

[236] I. Ogino, B. C. Gates, J. Am. Chem. Soc. 2008, 130, 13338-13346.<br />

[237] G. Laurenczy, A. E. Merbach, J. Chem. Soc., Chem. Commun. 1993, 187-189.<br />

[238] J. S. Bass, L. Kevan, J. Phys. Chem. 1990, 94, 1483-1489.


202 Bibliography<br />

[239] A. L. Harreus, in Ullmann's Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag<br />

GmbH & Co. KGaA, 2000.<br />

[240] R. J. Sundberg, in Comprehensive Heterocyclic Chemistry II (Ed.: Scriven), Pergamon,<br />

Oxford, 1996, pp. 119-206.<br />

[241] K. Aghapoor, L. Ebadi-Nia, F. Mohsenzadeh, M. Mohebi Morad, Y. Balavar, H. R.<br />

Darabi, J. Organomet. Chem. 2012, 708–709, 25-30.<br />

[242] J. Chen, H. Wu, Z. Zheng, C. Jin, X. Zhang, W. Su, Tetrahedron Lett. 2006, 47, 5383-<br />

5387.<br />

[243] M. Curini, F. Montanari, O. Rosati, E. Lioy, R. Margarita, Tetrahedron Lett. 2003, 44,<br />

3923-3925.<br />

[244] N. D. Schley, G. E. Dobereiner, R. H. Crabtree, Organometallics 2011, 30, 4174-4179.<br />

[245] N. T. S. Phan, T. T. Nguyen, Q. H. Luu, L. T. L. Nguyen, J. Mol. Catal. A: Chem. 2012,<br />

363–364, 178-185.<br />

[246] H. Noei, S. Amirjalayer, M. Müller, X. Zhang, R. Schmid, M. Muhler, R. A. Fischer, Y.<br />

Wang, ChemCatChem 2012, 4, 755-759.<br />

[247] M. Shoaee, M. W. Anderson, M. P. Attfield, Angew. Chem. Int. Ed. 2008, 47, 8525-<br />

8528.<br />

[248] M. Shoaee, J. R. Agger, M. W. Anderson, M. P. Attfield, CrystEngComm 2008, 10, 646-<br />

648.<br />

[249] P. St. Petkov, G. N. Vayssilov, J. Liu, O. Shekhah, Y. Wang, C. Wöll, T. Heine,<br />

ChemPhysChem 2012, 13, 2025-2029.<br />

[250] J. Szanyi, M. Daturi, G. Clet, D. R. Baer, C. H. F. Peden, Phys. Chem. Chem. Phys. 2012,<br />

14, 4383-4390.<br />

[251] A. Schneemann, V. Bon, I. Schwedler, I. Senkovska, S. Kaskel, R. A. Fischer, Chem.<br />

Soc. Rev. 2014, 43, 6062-6096.<br />

[252] D. Yan, G. O. Lloyd, A. Delori, W. Jones, X. Duan, ChemPlusChem 2012, 77, 1112-<br />

1118.<br />

[253] A. Carné-Sánchez, I. Imaz, M. Cano-Sarabia, D. Maspoch, Nat. Chem. 2013, 5, 203-<br />

211.<br />

[254] C. K. Brozek, M. Dincă, J. Am. Chem. Soc. 2013, 135, 12886-12891.<br />

[255] G.-T. Vuong, M.-H. Pham, T.-O. Do, Dalton Trans. 2013, 42, 550-557.<br />

[256] D. F. Sava Gallis, M. V. Parkes, J. A. Greathouse, X. Zhang, T. M. Nenoff, Chem. Mater.<br />

2015, 27, 2018-2025.<br />

[257] R. Shannon, Acta Crystallogr. A 1976, 32, 751-767.<br />

[258] Y. Jiao, C. R. Morelock, N. C. Burtch, W. P. Mounfield, J. T. Hungerford, K. S. Walton,<br />

Ind. Eng. Chem. Res. 2015, 54, 12408-12414.<br />

[259] M. V. Parkes, D. F. Sava Gallis, J. A. Greathouse, T. M. Nenoff, J. Phys. Chem. C 2015,<br />

119, 6556-6567.<br />

[260] Q. Zhang, L. Cao, B. Li, L. Chen, Chem. Sci. 2012, 3, 2708-2715.<br />

[261] N. Y. Kozitsyna, S. E. Nefedov, F. M. Dolgushin, N. V. Cherkashina, M. N. Vargaftik, I.<br />

I. Moiseev, Inorg. Chim. Acta 2006, 359, 2072-2086.<br />

[262] N. S. Akhmadullina, N. V. Cherkashina, N. Y. Kozitsyna, I. P. Stolarov, E. V. Perova, A.<br />

E. Gekhman, S. E. Nefedov, M. N. Vargaftik, I. I. Moiseev, Inorg. Chim. Acta 2009, 362,<br />

1943-1951.<br />

[263] G. Nickerl, U. Stoeck, U. Burkhardt, I. Senkovska, S. Kaskel, J. Mater. Chem. A, 2014,<br />

2, 144-148.<br />

[264] J. M. Teo, C. J. Coghlan, J. D. Evans, E. Tsivion, M. Head-Gordon, C. J. Sumby, C. J.<br />

Doonan, Chem. Commun. 2016, 52, 276-279.<br />

[265] J. A. R. Navarro, E. Barea, J. M. Salas, N. Masciocchi, S. Galli, A. Sironi, C. O. Ania, J. B.<br />

Parra, Inorg. Chem. 2006, 45, 2397-2399.


Bibliography 203<br />

[266] I. Cabria, M. J. López, S. Fraile, J. A. Alonso, J. Phys. Chem. C 2012, 116, 21179-21189.<br />

[267] U. Muschiol, P. K. Schmidt, K. Christmann, Surf. Sci. 1998, 395, 182-204.<br />

[268] M. Lischka, A. Groß, Phys. Rev. B 2002, 65, 075420.<br />

[269] X. Zhang, H. Xu, X. Liu, D. L. Phillips, C. Zhao, Chem. Eur. J. 2016, 22, 7288-7297.<br />

[270] L. Peng, J. Zhang, S. Yang, B. Han, X. Sang, C. Liu, G. Yang, Green Chem. 2015, 17,<br />

4178-4182.<br />

[271] X. Jia, S. Wang, Y. Fan, J. Catal. 2015, 327, 54-57.<br />

[272] B. Gole, U. Sanyal, R. Banerjee, P. S. Mukherjee, Inorg. Chem. 2016, 55, 2345-2354.<br />

[273] M. Sabo, A. Henschel, H. Frode, E. Klemm, S. Kaskel, J. Mater. Chem. 2007, 17, 3827-<br />

3832.<br />

[274] T. Otto, N. N. Jarenwattananon, S. Glöggler, J. W. Brown, A. Melkonian, Y. N. Ertas,<br />

L.-S. Bouchard, Appl. Catal., A 2014, 488, 248-255.<br />

[275] A. Arnanz, M. Pintado-Sierra, A. Corma, M. Iglesias, F. Sánchez, Adv. Synth. Catal.<br />

2012, 354, 1347-1355.<br />

[276] F. X. Llabrés i Xamena, A. Abad, A. Corma, H. Garcia, J. Catal. 2007, 250, 294-298.<br />

[277] S. Schuster, E. Klemm, M. Bauer, Chem. Eur. J. 2012, 18, 15831-15837.<br />

[278] C. Zlotea, R. Campesi, F. Cuevas, E. Leroy, P. Dibandjo, C. Volkringer, T. Loiseau, G.<br />

Férey, M. Latroche, J. Am. Chem. Soc. 2010, 132, 2991-2997.<br />

[279] G. Li, H. Kobayashi, J. M. Taylor, R. Ikeda, Y. Kubota, K. Kato, M. Takata, T. Yamamoto,<br />

S. Toh, S. Matsumura, H. Kitagawa, Nat. Mater. 2014, 13, 802-806.<br />

[280] S. Opelt, S. Türk, E. Dietzsch, A. Henschel, S. Kaskel, E. Klemm, Catal. Commun. 2008,<br />

9, 1286-1290.<br />

[281] C. Wang, H. Zhang, C. Feng, S. Gao, N. Shang, Z. Wang, Catal. Commun. 2015, 72, 29-<br />

32.<br />

[282] B. Yuan, Y. Pan, Y. Li, B. Yin, H. Jiang, Angew. Chem. Int. Ed. 2010, 49, 4054-4058.<br />

[283] M. Gulcan, M. Zahmakiran, S. Özkar, Appl. Catal., B 2014, 147, 394-401.<br />

[284] M. Kim, J. F. Cahill, H. Fei, K. A. Prather, S. M. Cohen, J. Am. Chem. Soc. 2012, 134,<br />

18082-18088.<br />

[285] H. Fei, J. F. Cahill, K. A. Prather, S. M. Cohen, Inorg. Chem. 2013, 52, 4011-4016.<br />

[286] G. H. Jeong, S. H. Kim, M. Kim, D. Choi, J. H. Lee, J.-H. Kim, S.-W. Kim, Chem. Commun.<br />

2011, 47, 12236-12238.<br />

[287] G. Pawley, J. Appl. Crystallogr. 1981, 14, 357-361.<br />

[288] H.-F. Wang, W. E. Kaden, R. Dowler, M. Sterrer, H.-J. Freund, Phys. Chem. Chem. Phys.<br />

2012, 14, 11525-11533.<br />

[289] R. Arrigo, M. E. Schuster, Z. Xie, Y. Yi, G. Wowsnick, L. L. Sun, K. E. Hermann, M.<br />

Friedrich, P. Kast, M. Hävecker, A. Knop-Gericke, R. Schlögl, ACS catal. 2015, 5,<br />

2740-2753.<br />

[290] A. C. Skapski, M. L. Smart, J. Chem. Soc. D 1970, 658b-659.<br />

[291] D. B. Dell'Amico, F. Calderazzo, F. Marchetti, S. Ramello, Angew. Chem. Int. Ed. 1996,<br />

35, 1331-1333.<br />

[292] A. F. Wells, Z. Kristallogr., Kristallgeom., Kristallphys., Kristallchem. 1938, 100, 189-<br />

194.<br />

[293] S. Opelt, V. Krug, J. Sonntag, M. Hunger, E. Klemm, Microporous Mesoporous Mater.<br />

2012, 147, 327-333.<br />

[294] M. Goepel, M. Al-Naji, P. With, G. Wagner, O. Oeckler, D. Enke, R. Gläser, Chem. Eng.<br />

Technol. 2014, 37, 551-554.<br />

[295] J. Sun, Y. Fu, G. He, X. Sun, X. Wang, Cat. Sci. Tec. 2014, 4, 1742-1748.<br />

[296] M. Al-Naji, A. M. Balu, A. Roibu, M. Goepel, W.-D. Einicke, R. Luque, R. Glaser, Cat.<br />

Sci. Tec. 2015, 5, 2085-2091.


204 Bibliography<br />

[297] A. P. Hammersley, S. O. Svensson, M. Hanfland, A. N. Fitch, D. Hausermann, High<br />

Pressure Res. 1996, 14, 235-248.<br />

[298] O. V. Dolomanov, L. J. Bourhis, R. J. Gildea, J. A. K. Howard, H. Puschmann, J. Appl.<br />

Crystallogr. 2009, 42, 339-341.<br />

[299] G. Sheldrick, Acta Crystallogr. A 2008, 64, 112-122.<br />

[300] Y. Wang, A. Glenz, M. Muhler, C. Wöll, Rev. Sci. Instrum. 2009, 80, 113108.<br />

[301] B. Ravel, M. Newville, J. Synchrotron Radiat. 2005, 12, 537-541.


Appendix<br />

List of Publications<br />

Several paragraphs and figures of this thesis have already been published in peerreviewed<br />

journals. These articles are reproduced in part with permission from the related<br />

publishers:<br />

Wenhua Zhang, Olesia Kozachuk, Raghavender Medishetty, Andreas Schneemann,<br />

Ralph Wagner, Kira Khaletskaya, Konstantin Epp, and Roland A. Fischer*. “Controlled<br />

SBU Approaches to Isoreticular Metal-Organic Framework Ruthenium-Analogues of<br />

HKUST-1”, European Journal of Inorganic Chemistry, 2015, 2015, 3913-3920.<br />

Wenhua Zhang, Max Kauer, Olesia Halbherr, Konstantin Epp, Penghu Guo, Miguel I.<br />

Gonzalez, Dianne J. Xiao, Christian Wiktor, Francesc X. LIabrés i Xamena, Christof Wöll,<br />

Yuemin Wang,* Martin Muhler, and Roland A. Fischer*. “Ruthenium Metal-Organic<br />

Frameworks with Different Defect Types: Influence on Porosity, Sorption and Catalytic<br />

Properties”, Chemistry-A European Journal, accepted.<br />

Wenhua Zhang, Zhihao Chen, Majd Al-Naji, Penghu Guo, Stefan Cwik, Olesia Halbherr,<br />

Yuemin Wang, Martin Muhler, Nicole Wilde, Roger Gläser and Roland A. Fischer*.<br />

“Simultaneous introduction of various palladium active sites into MOF via one-pot<br />

synthesis: Pd@[Cu3-xPdx(BTC)2]n”, Dalton Transactions (submitted).<br />

Wenhua Zhang, Kerstin Freitag, Suttipong Wannapaiboon, Roland A. Fischer*.<br />

“Elaboration of a Highly Porous Ru II,II analog of HKUST-1”, Inorganic Chemistry<br />

(submitted).<br />

Wenhua Zhang, Max Kauer, Sebastian Kunze, Stefan Cwik, Yueming Wang, and Roland<br />

A. Fischer*. “Study of synthesis parameters on the defect engineering of HKUST-1”,<br />

European Journal of Inorganic Chemistry (to be submitted).<br />

Zhenlan Fang, Johannes P. Dürholt, Max Kauer, Wenhua Zhang, Charles Lochenie,<br />

Bettina Jee, Bauke Albada, Nils Metzler-Nolte, Andreas Pöppl, Birgit Weber, Martin<br />

Muhler, Yuemin Wang*, Rochus Schmid*, and Roland A. Fischer*. “Structural<br />

Complexity in Metal–Organic Frameworks: Simultaneous Modification of Open Metal<br />

Sites and Hierarchical Porosity by Systematic Doping with Defective Linkers”, Journal<br />

of the American Chemical Society, 2014, 136, 9627-9636.


206 Appendix<br />

List of Presentations<br />

Wenhua Zhang, Max Kauer, Dianne J. Xiao, Ralph Wagner, Konstantin Epp, Olesia<br />

Kozachuk, Yuemin Wang, Roland A. Fischer. “Defect Engineered” mixed valence Ru-<br />

MOFs: study on the influence of defect metal sites.<br />

250th American Chemical Society National Meeting & Exposition<br />

Boston (USA), August 16 - 20, 2015, oral presentation.<br />

Wenhua Zhang, Olesia Kozachuk, Max Kauer, Zhenlan Fang, Yuemin Wang, Roland A.<br />

Fischer. “Defect Engineered” Mixed Valence MOFs.<br />

4th International Conference on Metal-Organic Frameworks and Open Framework<br />

Compounds<br />

Kobe (Japan), September 28 - October 1, 2014, poster presentation.<br />

Wenhua Zhang, Ralph Wagner and Roland A. Fischer. Investigation on the Ru-sites<br />

of the ruthenium Defect-Engineered Metal-Organic Frameworks by XAS.<br />

DELTA user meeting<br />

Dortmund (Germany), November 2015, post presentation.


Appendix 207<br />

Curriculum Vitae<br />

Personal information<br />

Name:<br />

Wenhua Zhang<br />

Date of Birth: 6th Dec. 1987<br />

Place of Birth:<br />

Nationality:<br />

Marital Status<br />

Education<br />

Henan, China<br />

Chinese<br />

unmarried<br />

10. 2012- Present Ruhr-University Bochum, Bochum, Germany<br />

Ph. D. candidate at the Chair of Inorganic Chemistry II<br />

Supervisor: Prof. Roland A. Fischer<br />

06.2015-08.2015 University of California, Berkeley, USA<br />

Visit Scholar in the group of Prof. Jeffrey R. Long at the<br />

Department of Chemistry<br />

09. 2009- 07. 2012 Zhengzhou University, Zhengzhou, China<br />

Master of Science in Inorganic Chemistry, College of Chemistry<br />

and Molecular Engineering<br />

Supervisor: Prof. Guang Yang<br />

Recommended postgraduate candidates exempt from admission<br />

exam<br />

09. 2005- 07. 2009 Zhengzhou University, Zhengzhou, China<br />

Bachelor of Science in Chemistry, Department of Chemistry<br />

Bachelor Thesis Supervisor: Prof. Guang Yang<br />

08. 2001-06. 2005 Yichuan High School, Luoyang, China<br />

Awards and Fellowships<br />

High school education<br />

2012- Present Scholarship from China Scholarship Council (CSC)


208 Appendix<br />

05. 2015 Grant from the RS plus (Ruhr-University Bochum) for the<br />

research (exchange) stay in the group of Prof. Jeffrey. R. Long,<br />

University of California, Berkeley (USA) including the business<br />

trip to ACS fall meeting 2015<br />

02. 2015 Grant from RS Plus (RUB) for PR.INT (International Project) to<br />

invite Dianne J. Xiao from the group of Prof. Jeffrey R. Long to<br />

RUB<br />

07. 2014 Grant from RS Plus (RUB) for funding the business trip to MOF<br />

2014 conference<br />

2009-2012 Scholarship for Outstanding Student of Zhengzhou University

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!