synthesis and solid state studies of cryptophane based

207
SYNTHESIS AND SOLID-STATE STUDIES OF CRYPTOPHANE-BASED MATERIALS A Dissertation submitted to the Faculty of the Graduate School of Arts and Sciences of Georgetown University in partial fulfillment of the requirements for the degree of Doctor of Philosophy In Chemistry By Scott T. Mough, B. S. Washington, DC April 21, 2011

Upload: others

Post on 19-Apr-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

SYNTHESIS AND SOLID-STATE STUDIES OF CRYPTOPHANE-BASED MATERIALS

A Dissertation

submitted to the Faculty of the

Graduate School of Arts and Sciences

of Georgetown University

in partial fulfillment of the requirements for the

degree of

Doctor of Philosophy

In Chemistry

By

Scott T. Mough, B. S.

Washington, DC

April 21, 2011

Page 2: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

ii

SYNTHESIS AND SOLID-STATE STUDIES OF CRYPTOPHANE-BASED MATERIALS

Scott T. Mough, B.S.

Thesis Advisor: K. Travis Holman, Ph.D.

ABSTRACT

Continued political unrest in oil producing countries has led the United States

to search for new sources of energy technology, such as fuel cells and gas storage

materials. Natural gas, which is comprised primarily of methane, is a particularly

attractive energy source in that the United States has considerable natural gas

reserves. Unfortunately, methane gas is typically stored in gas cylinders due to

potential risks of explosion, which has prevented the significant use of natural gas in

small automobiles.

Gas storage on porous solid surfaces is one potential alternative for reversible

storage in gas cylinders. Gas storage has been observed in carbon nanotubes,

graphitic carbon, in metal hydrides, and in coordination polymers. However, gas

storage on carbon materials and current coordination polymers does not achieve the

United States Department of Energy goal of 6.5 mass % of hydrogen gas while metal

hydrides can suffer from poor kinetics of hydride formation.

Page 3: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

iii

We proposed the use of irregularly shaped, functionalized “container

molecules” in metal-organic coordination polymers as new materials. Container

molecules, such as cryptophanes, have been shown to bind and store gases, such as

methane and xenon. Due to their unusual shape, container molecules pack

inefficiently in the solid-state. Specifically, we synthesized and characterized m-

xylyl bridge, exo-functionalized cryptophanes, and examined the solid-state behavior

of the resulting cryptophane based materials.

A single-crystal X-ray diffraction study of exo-functionalized cryptophanes

was performed, and solvent binding within the cryptophane was analyzed in the

solid-state. Aromatic guests were found to adopt a preferred orientation within the

cryptophane, likely due to CH-pi intermolecular interactions. Heating of an tris exo-

ester functionalized cryptophane resulted in the formation of an “imploded”

atropisomer in the solid-state, which was the first to be structurally characterized by

X-ray crystallography. The first cryptophane-based coordination polymer was

synthesized and characterized structurally. This material was found to spontaneously

desolvate, and the initial stages of this desolvation were found to occur in a single-

crystal to single-crystal fashion. The material was found to regenerate its initial

structure when reintroduced to mother liquor. However, this polymer was not found

to adsorb significant quantities of gas.

Page 4: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

iv

ACKNOWLEDGEMENTS

There are so many people and groups that I need to recognize that I may need

to add an extra chapter. First, I must recognize my advisor K. Travis Holman. His

love of science is so pure, so enthusiastic and so infectious that it makes him the ideal

mentor. I would not be where I am without his help and guidance.

The Department of Chemistry has been far more patient and understanding

than I deserve. I thank everyone involved in the department, but I especially thank

my committee for their constructive criticism and Dr. Tong for his helpful pep talks.

If not for YuYe, I would not be completing this thesis.

I have many fellow classmates to recognize for their friendship and help, in

both good times and bad. My group mates Thai Binh Nguyen, Steven Drake, Sayon

Kumalah, and Robert Fairchild have all helped me in ways great and small. I also

need to acknowledge the work of John Goeltz and Katherine Zumberge, who worked

incredibly hard to further this research. The Wolf group was very helpful, and I

would like to thank Rachel Lerebours, Kim Yearick, Brian Reinhardt and Xuefeng

Mei for their friendship and guidance.

I would like to recognize aaiPharma Services for their support and help. I

truly enjoyed my time at AAI, and would like to thank Jim Murtagh and Bob Whittle

for taking me under their wings. I thank TJ Harper and John Burke for cracking me

up in the lab and to Kristie Willoughby for her support.

Page 5: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

v

Finally, I would like to thank my family. My father, John Mough, is the

smartest man that I have ever met. I admire the fact that he always strives to learn, to

grow, and to better himself. My mother, Virginia, has incredible intuition and has

taught me to trust my instincts. My brothers John and Michael are friends that I

know that I can count on when things become difficult, and they know that they can

count on me in the same way. Aaron Brentzel and Deidre Savisky, my brother and

sister from another mother, have always been there to listen and offer unfiltered

advice when needed.

Most of all, I would like to recognize my wife, Rhonda and my children

Nathan and Connor. Rhonda is a beautiful soul, and is the strongest person that I

have ever met. Her determination is a constant source of amazement and inspiration.

At the same time, she is a sensitive, loving, caring wife and mother. I am so blessed

to have met my soul mate, and that I get to spend the rest of my life with her. Nathan

is equally inspirational. He works hard to learn and understand the world around

him. He shows us flashes of a brilliant sense of humor and his caring nature, and he

is loved by all who know him. Connor is our comic relief. His natural brilliance and

charisma bring joy to all around him.

Finally, I would like to thank God for my good fortune. He is truly great and

I will never be able to appreciate all of His gifts.

Page 6: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

vi

TABLE OF CONTENTS

Abstract ......................................................................................................................... ii

Acknowledgements ...................................................................................................... iv

List of Figures .............................................................................................................. xi

List of Schemes and Tables ...................................................................................... xxii

Chapter 1: Introduction ................................................................................................ 1

1.1. Introduction: Zeolites as Functional Materials ................................... 1

1.2. Crystal engineering: Designing alternatives to zeolites ....................... 3

1.2.1. Crystal Engineering of Coordination Polymers ....................... 5

1.3 Supramolecular Chemistry and Host-Guest Binding ........................... 9

1.3.1 Container Molecules: Potentially Useful Supramolecular CP

Ligands................................................................................................ 12

1.4. Cryptophanes ...................................................................................... 17

1.5. Research project: Synthesis and Characterization of Cryptophane

Container-Based Materials ................................................................. 20

1.6 References ........................................................................................... 22

CHAPTER 2: Synthesis, Characterization and Inclusion Compounds of m-xylyl

Bridged Cryptophanes ................................................................................................ 33

2.1. Introduction: Synthesis and characterization of cryptophanes .......... 33

Page 7: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

vii

2.2. Solution vs. Solid-State Analyses of Supramolecular Host-Guest

Complexes .......................................................................................... 36

2.3. History of Crystallographic Analysis of Container Molecule Host-

Guest Complexes ................................................................................ 37

2.4. Syntheses of m-xylyl bridged cryptophanes ....................................... 39

2.5. Crystal Structures of m-xylyl Bridged Cryptophanes ......................... 44

2.5.1 SQUEEZE analysis of m-xylyl Bridged CryptophaneGuest

Structures ................................................................................ 45

2.6. m-xylyl CryptophaneAromatic Guest Complexes ........................... 46

2.6.1. Common Guest Binding Motif: The Importance of CH-

Interactions.............................................................................. 46

2.6.2 CH- Interactions: Weak Intermolecular Forces ................... 47

2.7. m-Xylyl Bridged Cryptophane Host Conformational Changes ............. 49

2.8. Halobenzene guests: Conglomorate vs. Racemate formation ............... 58

2.9. Complete Encapsulation vs. Partial Encapsulation................................ 60

2.10. Nonaromatic Guests ............................................................................. 64

2.10.1. Encapsulation of two guests ................................................. 64

2.10.2. Guest Disorder in Cryptophane Structures ........................... 66

2.10.3. Host Conformation Changes of Cryptophanes Binding

Nonaromatic Guests ................................................................ 67

Page 8: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

viii

2.10.4. Crystal Structures of Host 7: High Symmetry Structures .... 72

2.11. Conclusions ......................................................................................... 74

2.12. Experimental ....................................................................................... 75

2.12.1 General Methods ..................................................................... 75

2.12.2. New Molecule Characterization ............................................. 76

2.12.3. Crystal Structures .................................................................... 82

2.13 References ......................................................................................... 101

CHAPTER 3: Synthesis and Characterization of an “Imploded” Cryptophane

Atropisomer .............................................................................................................. 107

3.1. Introduction to CTB “cup” inversion................................................ 107

3.2. Thermal Analysis of (±)-anti-4THF3THF ......................................... 110

3.3. Consequences of Desolvation: Atropisomerization of (±)-anti-4 ......... 111

3.3.1. 1H NMR of desolvated (±)-anti-4 material ............................. 111

3.3.2. 2D NMR of (±)-imp-4: COSY and ROESY ........................ 113

3.4. Kinetics of atropisomerization by 1H

NMR ...................................... 120

3.5. Single crystal structure of (±)-imp-411CHCl3 ................................. 123

3.6. Conclusions ....................................................................................... 125

3.7. Experimental ..................................................................................... 126

3.7.1. General Methods ................................................................... 126

3.7.2. New Molecule Characterization ........................................... 126

Page 9: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

ix

3.7.3 Kinetics Experiments ............................................................ 127

3.7.4 Crystal Structure ................................................................... 128

3.8 References ......................................................................................... 130

CHAPTER 4: Synthesis and Characterization of Cryptophane-Based Metal-Organic

Polymer (CBMOP): Single Crystal to Single Crystal Partial Desolvation .............. 132

4.1. Introduction: Single Crystal to Single Crystal Processes ............... 132

4.2. A 1-D Cryptophane-Derived Coordination Polymer ........................ 138

4.3. Unit Cell Changes in CBMOP .......................................................... 143

4.3.1. X-Ray Single Crystal Structure of CBMOP-desolvated ...... 145

4.3.2. Explanation of Single Crystal Desolvation ........................... 146

4.4. Desolvation/Resolvation of Bulk CBMOP Observed by Powder X-ray

Diffraction ......................................................................................... 147

4.5. Gas Sorption Study on CBMOP-desolvated ..................................... 151

4.6. Synthesis of a Cryptophane-Dimer ................................................... 152

4.7. Conclusions ....................................................................................... 154

4.8. Experimental ..................................................................................... 155

4.8.1. General Methods ................................................................... 155

4.8.2. Synthesis of New Materials .................................................. 155

4.8.3. Crystal Structures .................................................................. 156

4.8.4. Unit Cell Analysis over time ................................................ 159

Page 10: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

x

4.9 References ......................................................................................... 160

CHAPTER 5: Solid-State Kinetics of Supramolecular Host-Guest Clathrates: ...... 164

5.1. Introduction: Solid-State Kinetics ................................................... 164

5.2. Desolvation of Host-Guest Systems ................................................. 168

5.3. Isothermal TGA Analysis of CTVTHF0.5 ........................................ 169

5.4. Nonisothermal Kinetic Analysis of CTVTHF0.5 .............................. 174

5.5. Nonisothermal DSC Analysis of (±)-anti-8THF ........................... 175

5.6. Discussion of Activation Parameters ................................................ 179

5.7. Conclusions ....................................................................................... 180

5.8. Experimental ..................................................................................... 180

5.8.1. General Methods ................................................................... 180

5.8.2. Sample Preparation ............................................................... 181

5.9 References ......................................................................................... 182

Page 11: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xi

LIST OF FIGURES:

Figure 1.1. Dsolvated ZSM-5. The sustained pores are evident. Red, Oxygen;

Silicon, Pink .......................................................................................... 2

Figure 1.2. Crystal structures of organic carboxylic acids which show how

molecular structure can influence two-dimensional and three-

dimensional solid-state structure. a) 1D-chain structure of 1,4

terephthalic acid. b) Honeycomb structure of trimesic acid. Gray,

Carbon; Red, Oxygen; White, Hydrogen. ............................................. 5

Figure 1.3. Two canonical structures derived from metal cations and organic

linkers. Left: Cubic structure derived from octahedral metal cations

and linear organic linkers. Right: Adamantyl structure derived from

tetrahedral metal cations and linear organic linkers ............................. 6

Figure 1.4. Top. Zn4O cluster and 4 carboxylate CO2- groups leading to cubic

structures. Gray, Carbon; Red, Oxygen; Pink, Zinc, Bottom. Pore

volume and crystal density as a function of ligand length in Yaghi’s

Isoreticular cubic metal-organic frameworks ....................................... 7

Figure 1.5. Rebek’s self-recognizing “tennis ball” dimer. .................................... 10

Figure 1.6. Nine molecules included in Isaacs self-sorting study. ........................ 11

Figure 1.7. Kinetic barrier to decomplexation as a result of constrictive binding in

a container molecule ........................................................................... 13

Figure 1.8. Conformational gating mechanisms in container molecules that enable

Page 12: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xii

“constrictive binding”. ........................................................................ 14

Figure 1.9. Selective [2+2] photodimerization inside a metal-ligand container

molecule. This reaction occurs only in the presence of the container,

and yielded only the cross syn dimer. ................................................. 15

Figure 1.10. Left. General structure of (±)-anti cryptophane. Right. General

structure of syn cryptophane. In most common cryptophanes, A = Z =

OMe. ................................................................................................... 16

Figure 1.11. Descriptions of cryptophane regions. ................................................. 17

Figure 2.1. 1H

NMR of cryptophane E (propyl-bridged). ..................................... 35

Figure 2.2 Left: Alkynyl bridged hemicarcerand, which is defined by the C4

symmetric cavitands that make up the polar regions of the container.

Right. Crystal structure (YETKAJ) describing alkynyl bridged

hemicarcerand. The encapsulated molecule is 1,1,2,2

tetrachloroethane. ................................................................................ 38

Figure 2.3 Structure of Weber’s endo-carboxylate m-xylyl cryptophane.

Hydrogen atoms and guest molecules omitted for clarity. Carbon:

Gray; Oxygen: Red. ............................................................................ 40

Figure 2.4. Left) 1H NMR spectrum of (±)-anti-4. Right) H NMR spectrum of

syn-5. ................................................................................................... 44

Figure 2.5. Determination of distance for CH- interactions. Distance d is taken

from the guest proton to either the closest cryptophane aromatic

Page 13: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xiii

carbon or the centroid defined by the six carbons in the appropriate

ring. Angle is defined as the C-H- angle where C = guest carbon,

H = guest hydrogen and = either the closest cryptophane aromatic

carbon or the centroid defined by the six carbons in the ring. ............ 47

Figure 2.6. Side and Top views of cryptophanearomatic guest complexes. Guest

molecules shown in space-fill form and cryptophanes shown in stick

form. a) (±)-anti-4C6H5NO2 b) (±)-anti-4C6H5CN c) (±)-anti-

4C6H5(CH2)2CH3 d) anti-4C6H5Br1 e) (±)-anti-4C6H5Cl f) 2-

(±)-anti-6m-C6H4(CH3)2 g) (±)-anti-6C6H5Br h) (±)-anti-61,2,4-

C6H3(CH3)3 i) (±)-anti-6o-C6H4(CH3)2 j) (±)-anti-4C6H5I1 k) (±)-

anti-4C6H5(CH2)5CH3 l) syn-5C6H5NO2 m) (±)-anti-6C6H5NO2.

Carbon: Gray; Hydrogen: White; Nitrogen: Blue; Oxygen: Red;

Chlorine: Yellow; Bromine: Yellow. 1

Forms conglomerate structures

(see vida infra). ................................................................................... 50

Figure 2.7. Measured torsion angle in cryptophanes. ........................................... 57

Figure 2.8. Left: CCTB-O-CH2-Cbridge torsion angles where CTB arene does not

engage in CH- interactions (CTB arene-methyl close contacts also

removed). Right: CCTB-O-CH2-Cbridge torsion angles where CTB

arene is engaged in CH- interactions. ............................................... 57

Page 14: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xiv

Figure 2.9. Top: Histogram describing the bimodal distribution of torsion angles

in which host-guest CH- interactions are observed. Bottom:

Histogram describing the distribution of torsion angles in which host-

guest CH- interactions are not observed. .......................................... 60

Figure 2.10. Top Left: (±)-anti-4PhBr looking down the –c axis. Top Right:

(±)-anti-4PhCl looking down the c-axis. Bottom: Interaction of

bromobenzene guest with ester of adjacent cryptophane. The

bromine-oxygen interaction is shown with the dotted red line. Guest

molecules shown in space-fill form and cryptophanes shown in stick

forms. Carbon: Gray; Hydrogen: White; Nitrogen: Blue; Oxygen: Red;

Chlorine: Yellow; Bromine: Yellow .................................................. 62

Figure 2.11. Top Left: Stick representation of (±)-anti-4C6H5(CH2)2CH3 with

spacefilled guest. Top Right: Stick representation of C6H5(CH2)2CH3

guest. Bottom Left: Stick representation of (±)-anti-

4C6H5(CH2)5CH3 with spacefilled guest. Bottom Right: Stick

representation of C6H5(CH2)5CH3 guest. Carbon: Gray; Hydrogen:

White; Oxygen: Red. .......................................................................... 63

Page 15: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xv

Figure 2.12. Left: Two encapsulated NO2Me molecules as packed within (±)-anti-

42NO2Me. Right: Two encapsulated CH2Cl2 molecules as packed

within (±)-anti-42CH2Cl2. ............................................................... 64

Figure 2.13. Left: Model of disordered, encapsulated DMF. Right: Model of

disordered, encapsulated CHCl3 ......................................................... 66

Figure 2.14. ArCTB-O-CH2-Arbridge torsion angles. .................................................. 72

Figure 2.15. Spacefill packing of syn-7 as seen down c-axis. Lattice solvent has

been deleted to show the channels that run between the cryptophane

molecules. ........................................................................................... 74

Figure 3.1. Crown inversion of CTBs. Note that the saddle-twist intermediate is

close in energy to the crown conformation (ΔΔG298 = ~12-16 kJ/mol

higher for saddle-twist intermediate). ............................................... 108

Figure 3.2. Saddle-twist rotation of CTBs. Arrows show one direction of rotation,

but rotation could occur in the other direction. ................................. 109

Figure 3.3. CTB oxide molecule in saddle-twist formation. ............................... 110

Figure 3.4. TGA of cryptophane materials. Top: Thermograms of (±)-anti-

4THF3THF and (±)-anti-4THF as a function of temperature. Note

that the THF molecules in the lattice can be separated from the

encapsulated THF molecule. Bottom: Isothermal thermogram of (±)-

anti-4THF3THF at 85C. This heating profile allows for removal of

only lattice THF molecules. .............................................................. 111

Page 16: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xvi

Figure 3.5. 1H NMR of atropisomerization of (±)-anti-4 over time. .................. 112

Figure 3.6. Left: Model of a “cup-in-cup” cryptophane. Right: Model of a

“saddle-twist-cup” cryptophane. ....................................................... 113

Figure 3.7. COSY of (±)-imp-4 in CDCl3 at 25°C. Top: Full spectrum. Bottom:

Expanded spectrum. .......................................................................... 115

Figure 3.8 COSY of (±)-imp-4 in acetone-d6 at 25°C. ....................................... 116

Figure 3.9. ROESY of (±)-imp-4 in CDCl3 at 25°C.. ......................................... 118

Figure 3.10. 1H NMR of (±)-imp-4 at -55C. ........................................................ 119

Figure 3.11. Kinetics of inflation of (±)-imp-4 to (±)-anti-4 by 1H NMR integration

at 298 K. ............................................................................................ 121

Figure 3.12. Eyring plot derived from isothermal kinetic data. ............................ 122

Figure 3.13. Single crystal structure of (±)-imp-4. The 11 CHCl3 molecules in the

ASU have been removed for clarity. Left: Stick structure of (±)-imp-

4. Note that the saddle-twist CTB points a methoxy group into the

cup-shaped CTB. CTB arene rings are highlighted in blue. Right:

(±)-imp-4 in which the cup-shaped CTB unit is shown in CPK form,

and the saddle-twist CTB unit is shown in stick form. The saddle-

twist CTB arene rings are highlighted in blue. ................................. 124

Figure 3.14. Assigned 1H NMR of (±)-imp-4. ...................................................... 124

Figure 4.1. Single-crystal to single-crystal processes. ........................................ 133

Page 17: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xvii

Figure 4.2. a) [Ni2(C26H52N10)]3[BTC]46C5H5N36H2O bilayer viewed in stick

form (left) and spacefill (right). b) [Ni2(C26H52N10)]3[BTC]44H2O

bilayer viewed in stick form (left) and spacefill (right). ................... 134

Figure 4.3. a) Structure of tetrapyridone crystallized with isovaleric acid. b)

Structure of tetrapyridone after guest exchange with propionic acid.

Note the decrease in the c-axis with the smaller propionic acid. ...... 135

Figure 4.4. Single-crystal to single-crystal syntheses. Top left: Crystal structure

of 1,6 triene. Top Right: Crystal structure of photo-polymerized 1,6

triene. Bottom left: Hydrogen bond mediated assembly of 4,4’

bipyridyl ethylene. Bottom right: [2+2] Photochemical synthesis of

[2.2]paracyclophane. ......................................................................... 136

Figure 4.5. One dimensional chain of CBMOP-solvent as viewed down the b-

axis. Chain propagates along the [1 0 1] direction. Hydrogen atoms,

encapsulated DMF molecules, and solvent molecules omitted for

clarity. CTB arenes have been filled in............................................ 137

Figure 4.6. Coordination environment about metal centers. ............................... 139

Figure 4.7. Overlay of cryptophane in anti-(±)-H39 and CBMOPsolvent viewed

from top of cryptophane. Note the difference in the cryptophane

bridges on the left. While (±)-anti-H39 continues in a nearly straight

path, CBMOP-solvent is rotated approximately 60°. ...................... 140

Page 18: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xviii

Figure 4.8. Thermogravimetric Analysis of CBMOP-solvent. Top. CBMOP

mass as a function of heating at 10°C/min from room temperature to

380°C. Bottom. CBMOP heated at 0.2°C /min from room

temperature to 160°C , then held at 160°C for ~ 800 minutes. ....... 142

Figure 4.9. Unit cell change of CBMOP as a function of time at 25C. ............ 145

Figure 4.10. Changes in c-axis (top) and b-axis (bottom) before and after

desolvation. Green points represent individual CBMOP∙solvent

crystals, except for labeled spot. The orange points below the green

points represent CBMOP∙solvent crystals after desolvation

(CBMOP∙desolvated. ........................................................................ 147

Figure 4.11. Faces of crystal CBMOP. ................................................................ 148

Figure 4.12. Top: Spacefill representations of the single crystal structures of

CBMOPsolvent (top) and CBMOPdesolvated (bottom) as viewed

down the [001] direction. The lattice solvent molecules have been

removed for clarity. Note that the channels collapse upon desolvation.

Bottom: Connolly surface plot of CBMOP-solvent as viewed normal

to the [010] direction. The dark blue highlighted areas are the

channels observed above. The channels form a ladder-like structure

extending along the c-axis, and having rungs that run roughly along

the a-axis. .......................................................................................... 149

Page 19: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xix

Figure 4.13. Calculated powder X-ray diffractograms were determined using the

program LAZY-PULVERIX and the single-crystal x-ray diffraction

data for CBMOP-solvent and CBMOP-desolvated. The room

temperature unit cells for each material were used to more accurately

match the experimental data, which were collected at room

temperature. The powder x-ray diffraction (PXRD) experiment was

performed on one open capillary tube of CBMOP collected over time.

Experimental partial powder X-ray diffractograms (PXRD) of the

desolvation process of CBMOP∙solvent. a) Calculated PXRD pattern

using the unit cell CBMOP∙solvent obtained at room temperature. b)

Calculated PXRD pattern of CBMOP∙solvent that has been arbitrarily

broadened to more accurately reflect the peaks widths. c) Calculated

PXRD of CBMOP∙desolvated at room temperature. d) Experimental

PXRD pattern of CBMOP at t = 0. e) PXRD pattern of CBMOP at t

= 1 day. f) PXRD pattern of CBMOP at t = 6 days. g) PXRD pattern

of CBMOP at t = 8 days. h) PXRD pattern of CBMOP at t = 14 days.

i) PXRD pattern of CBMOP at t = 22 days. j) Experimental PXRD

pattern of the material in (i) after moistening the material with mother

liquor. ................................................................................................ 151

Figure 4.14. Proposed model of desolvation of CBMOF over time. ................... 152

Page 20: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xx

Figure 5.1. Series of xanthenol clathrates studied to relate structure and

desolvation kinetics........................................................................... 169

Figure 5.2. Crystal structure of CTV0.5THF. CTV molecules shown in stick,

THF molecules shown in spacefill form. .......................................... 170

Figure 5.3. Nonisothermal TGA experiments of CTVTHF0.5 as a function of

alpha (extent of desolvation) at 5C/min and 10/min. .................... 171

Figure 5.4. Top: Plot of alpha vs. time for an isothermal TGA experiment for the

desolvation of CTVTHF0.5 at 96.2 C. Bottom: Isothermal TGA data

after application of the three-dimensional diffusion model for the

desolvation of CTVTHF0.5. .............................................................. 172

Figure 5.5. Eyring plot of desolvation of CTVTHF0.5 derived from isothermal

TGA data........................................................................................... 173

Figure 5.6. DSC overlay of desolvation of CTVTHF0.5 at varying heating

rates .................................................................................................. 175

Figure 5.7. Ozawa plot derived from nonisothermal DSC experiments for the

desolvation of CTVTHF0.5 at various heating rates (1-20C/min). .. 176

Figure 5.8. Stucture of (±)-anti-8THF. Left: Side view of complex. Right:

view of complex. Carbon: Gray; Oxygen: Red; Hydrogen: White.. 176

Figure 5.9. DSC overlay of desolvation of (±)-anti-8THF at varying heating

rates. .................................................................................................. 177

Page 21: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xxi

Figure 5.10. Ozawa plot derived from nonisothermal DSC data for the desolvation

of (±)-anti-8THF. ........................................................................... 178

Figure 5.11. Side view of (±)-anti-4THF complex. ........................................... 179

Page 22: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xxii

LIST OF SCHEMES AND TABLES:

Scheme 2.1. Three potential synthetic pathways to cryptophane molecules. In most

cases, E = E’ = OCH3. ........................................................................ 33

Scheme 2.2. Synthesis of m-xylyl bridged cryptophanes. ....................................... 41

Table 1.1. Cryptand binding in water at 298K as a function of size. ................... 12

Table 1.2 Description of symbols used throughout text. .................................... 18

Table 1.3. Free energy of complexation for various cryptophanes in (CDCl2)2 at

300K.................................................................................................... 18

Table 2.1. Molecule Abbreviations ...................................................................... 45

Table 2.2. New m-xylyl bridged cryptophanes .................................................... 48

Table 2.3. CH- Interactions between cryptophanes and aromatic guests. .... 51-53

Table 2.4. X-Ray data for CryptophaneAromatic Guest materials. ............. 54-56

Table 2.5. Encapsulated guest volume in various cryptophane host molecule

materials. ............................................................................................. 65

Table 2.6. X-Ray data for CryptophaneNon-Aromatic Guest materials ...... 69-71

Table 2.7. Torsion angles found in cryptophanesnon-aromatic guest(s)

structures. ............................................................................................ 73

Table 3.1. Comparison of activation parameters for CTB conformers.............. 122

Table 4.1. Unit Cell data for CBMOP∙solvent and CBMOP∙desolvated at RT

and 173K. ................................................................................. 144

Page 23: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

xxiii

Table 5.1. Solid-state rate expressions for several reaction models. ................. 166

Table 5.2. Nonisothermal kinetic models. ......................................................... 167

Table 5.3 Activation parameters for desolvation of CTV·THF0.5 and (±)-anti-

8THF. ............................................................................................. 178

Page 24: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

1

CHAPTER 1: INTRODUCTION

1.1. Introduction: Zeolites as Functional Materials

The development of new functional materials is a burgeoning area at the

interface of materials science and chemistry. Existing materials, such as zeolites,

perform a variety of useful functions exploited by chemists at the academic and

industrial level.1 Zeolites are defined as “microporous crystalline aluminosilicates,

composed of TO4 tetrahedra (T = Si, Al) with O atoms connecting neighboring

tetrahedra (Figure 1.1).”1,2,3

Their structure can be described as corner sharing

tetrahedra, comprised mostly of SiO4 tetrahedra, but having some AlO4 substitution

(Figure 1.1). The substitution of silicon (Si4+

) for aluminum (Al3+

) gives the

frameworks a net negative charge; these anionic frameworks are filled with loosely

held water molecules and charge balancing metal cations. These materials exhibit

permanent porosity, which means that the material can be dehydrated and rehydrated

reversibly without collapse of the framework.4 Permanent porosity is an interesting

and potentially useful property of solid-state materials, and zeolitic materials have

uses in separations, catalysis, and gas storage.5,6

The sustained pores are evident in

Figure 1.1. In zeolites, as with many materials, their myriads of functions are

directly related to their interesting structures. They are commonly used in ion

exchange beds, which is not surprising given the fact that the zeolitic countercations

are loosely held inside their large cavities.7 Some zeolites, such as 4Å molecular

sieves, are used in synthetic laboratories to dehydrate common organic solvents. This

Page 25: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

2

behavior is a specific

example of a more

common ability of

molecular sieves to

separate molecules based

on size and shape.3 This

size and shape selectivity

is commonly exploited in

Zeolite A, which absorbs

and thereby separates

straight chain

hydrocarbons from

aromatic or branched hydrocarbons.5,8 Separation of p-xylene from xylene mixtures

by silicalite is another example of size and shape based selectivity in zeolites.5,8

The

driving forces for these phenomena are the satisfaction of electrostatic and van der

Waals forces within the dehydrated zeolite. More simply, “Nature abhors a vacuum,”

even one on an atomic scale. Ion exchange also plays a major role in zeolite

catalysis, which is one of the most important functions of zeolites. By exchanging

metal cations with protons, one creates a highly acidic zeolite which is used in crude

oil cracking, isomerization reactions and fuel syntheses.9

Figure 1.1. Desolvated ZSM-5. The sustained pores are evident.

Red, Oxygen; Silicon, Pink.

Page 26: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

3

1.2. Crystal engineering: Designing alternatives to zeolites

Though research into zeolites continues, it remains difficult to synthesize and

design new zeolite topologies. Chemists are attempting to create zeolites and zeolite-

like materials for the purpose of making new materials that are useful and interesting.

Some organic solid-state chemists are attempting to design similar materials from

different, modular building blocks.10

This strategy is predicated on the hypothesis

that material structure and function are intimately related, and that the global

structure of a crystalline material can be designed by using molecular symmetry

arguments and known or foreseeable intermolecular interactions (or synthons) in a

bottom-up design. This strategy has often been described in the research literature as

“crystal engineering”,11

and it constitutes a significant area of research for solid-state

organic chemists, materials scientists, and organometallic chemists.12

This

burgeoning field is semi-empirical in nature, as many noncovalent forces (London

dispersion, CH- interactions, - stacking, and dipole-dipole interactions for

example) simultaneously influence the overall structure and packing of a material.

Unfortunately, these weak intermolecular forces are difficult to quantify, thereby

making crystal structure prediction of even the most basic structures a scientific

achievement: Maddox claims “One of the continuing scandals in the physical

sciences is that it remains in general impossible to predict the structure of even the

simplest crystalline solids from a knowledge of their chemical composition.”13

While

Maddox’s statement is true even today, considerable progress has been made in the

Page 27: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

4

design of crystalline structures through the exploitation of predictable strong,

directional solid-state synthons. For example, hydrogen-bond donor-acceptor

systems have been studied extensively, and have been successfully employed in

materials design.14,15

Hydrogen bonds are quite strong (generally considered to be 3-

5 kcal/mol, but have been found to be up to 40 kcal/mol per H-bond in the solid

state), directional (between discreet hydrogen bond donors and acceptor lone pairs),

and may contribute considerably to the overall lattice energy of a crystalline

molecular solid.16,17

For example, a typical O-H

O hydrogen bond brings two

neighboring oxygen atoms to a distance of 2.5-2.8 Å, with an energy of 4-10 kcal/mol

for moderately strong hydrogen bonds.14,16

Materials derived from organic molecules that have relatively strong,

directional intermolecular forces have been studied extensively for their structural

properties. Carboxylic moieties have been observed to form one-dimensional chains,

two-dimensional sheets, and three-dimensional structures, from linear linkages (1,4

benzene dicarboxylic acid),18

triangular linkages (1,3,5 benzene tricarboxylic acid)19

and tetrahedral linkages (adamantane tetracarboxylic acid),20

respectively (Figure

1.2). The adamantyl structure is especially important because it was a designed

structure.

Research continues into finding and exploiting useful, noncovalent

intermolecular (supramolecular) synthons in the solid-state.11

,21

However, materials

that are organized by weak intermolecular forces are often of insufficient strength to

Page 28: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

5

sustain permanent porosity, as their structures often collapse upon desolvation.22

On

the other hand, materials held together by coordinate covalent linkages have proven

to be more robust, and therefore of greater conceivable utility.

Figure 1.2. Crystal structures of organic carboxylic acids which show how molecular structure

can influence two-dimensional and three-dimensional solid-state structure. a) 1D-chain

structure of 1,4 terephthalic acid. b) Honeycomb structure of trimesic acid. Gray, Carbon;

Red, Oxygen; White, Hydrogen.

1.2.1. Crystal Engineering of Coordination Polymers

Recently, crystal engineering research has focused greatly on the ubiquitous

coordination polymer (hereafter CP) or “Metal-Organic Framework (MOF).”23

CPs

are well-defined, stoichiometric assemblies that are composed of metal cations and

coordination ligands that form infinite polymers (1-D) or infinite networks (2-D or 3-

D). In most cases, the metal acts as a node, and the coordination ligand acts as

spacers or bridges that link the nodes (Figure 1.3). Some of these materials can be

described as being “zeolite-like” or as “designer zeolites,” as they often exhibit the

most useful properties of zeolite materials: permanent porosity,24

catalysis,25

ion

exchange,26

and selective sorption of isomers.27

Page 29: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

6

The potential of CPs was first described in detail by Robson, in his seminal

1990 paper “Design and construction of a new class of scaffolding-like materials

comprising infinite polymeric frameworks of 3D-linked molecular rods.” This paper

presented a reappraisal of the zinc cyanide and cadmium cyanide structures and the

synthesis and structure of the diamond-related frameworks [N(CH3)4][CuIZn

II(CN)4]

and CuI[4,4',4'',4'''-tetracyanotetraphenylmethane]BF4.xC6H5NO2.”

28 This work

describes the potential utility of CPs, as well as some potentially interesting structural

targets, including diamondoid and cubic networks. While infinite inorganic

frameworks such as Prussian Blue29

had been known for over 40 years, Robson truly

ushered in a new area of chemistry by describing how and why metal-organic

materials should be pursued. He essentially laid out a blueprint for solid-state

chemists to follow.

Figure 1.3. Two canonical structures derived from metal cations and organic linkers. Left:

Cubic structure derived from octahedral metal cations and linear organic linkers. Right:

Adamantyl structure derived from tetrahedral metal cations and linear organic linkers.

Page 30: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

7

CP research has increased

dramatically over the past several

years; the terms “coordination

polymer” or “metal-organic

framework” was cited in 14 ACS

journal articles in 1995 and 441 times

in 2010. The dramatic increase in

research of CP-based materials is a

result of the recognition of many

advantages derived from using

organic and coordination chemistries.

Organic ligand spacers are versatile,

modular, easily functionalized, and

react in a robust, well-understood

manner. One can, for instance,

change the length of the spacer to

modulate pore size. This has been

beautifully illustrated by Yaghi and

coworkers in their 2002 Science

article “Systematic Design of Pore

Size and Functionality in Isoreticular

Figure 1.4. Top. Zn4O cluster and 4 carboxylate CO2-

groups leading to cubic structures. Gray, Carbon;

Red, Oxygen; Pink, Zinc, Bottom. Pore volume and

crystal density as a function of ligand length in

Yaghi’s Isoreticular cubic metal-organic frameworks.

Page 31: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

8

MOFs and Their Application in Methane Storage”30

(Figure 1.4.). One also can

functionalize the spacer so as to functionalize the pore.31

Coordination chemistry

also has many advantages, including well-defined metal coordination geometries,

potentially labile metal-ligand interactions, and metal-ligand directionality (cis vs.

trans).2

In CPs, predictable metal-ligand or metal cluster-ligand geometries have been

observed, and have been exploited in numerous examples. These predictable units

can be described as solid-state synthons, and are also commonly described as

“Secondary Building Units (hereafter SBUs).32

SBUs provide a geometric construct

for the metal nodes, which may be combined with the ligand spacer of varying

symmetries to form frameworks of predictable topologies. For example, a linear

spacer combined with an octahedral node can form a cubic network, while linear

spacers can combine with square nodes to form two-dimensional square networks. In

general, the resultant CP materials are stronger and more stable to desolvation than

similar hydrogen-bonded materials.

In the years since Robson’s seminal paper, a number of chemists have

synthesized CP materials with varied functional properties: magnetism,33

porosity

and gas storage,34

sensing,35

catalysis22b,36

and separations.37

Yaghi and coworkers

have been quite successful in designing metal carboxylate materials exhibiting

permanent porosity, and some of Yaghi’s MOFs have been shown to have the highest

surface area and lowest density known for crystalline materials. Yaghi, along with

Page 32: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

9

Kitagawa,38

Rosseinsky,39

Long,40

and other41

s have designed materials that

reversibly adsorb gases, such as H2 and CO2, O2 and C2H2. Others, including Lin42

and Hupp,43

have designed CP materials that function as heterogeneous catalysts,

catalyzing asymmetric addition to aldehydes, olefin epoxidation, and hydrogenation

of ketones. Zaworotko44

and Eddaoudi45

have designed spin frustrated CP materials

that form a Kagome’ lattice, while van Koten46

designed a sensing material which

reversibly binds and releases SO2 gas. The result is a relatively new area of research

that is as versatile and has the potential for large-scale industrial use.

1.3 Supramolecular Chemistry and Host-Guest Binding

Donald Cram, Jean-Marie Lehn, and Charles Pedersen shared the Nobel Prize

in Chemistry in 1987 for their contributions to science, specifically in supramolecular

chemistry.47

Supramolecular chemistry can be broadly defined as the study and

understanding of non-covalent interactions between molecules, as well as the

response of molecules to these intermolecular forces.48

Biochemistry has shown that

these relatively weak interactions are often highly important in biological systems

(hydrogen bonding of complimentary DNA strands,49

proper folding of proteins,50

antigen-antibody recognition51

). As in Nature, supramolecular chemists aim to

control and manipulate molecules in space through these non-covalent interactions.

Two major areas of research for supramolecular chemists are molecular

recognition52

and host-guest binding.53

Molecular recognition describes the ability of

Page 33: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

10

an entity (enzyme, molecular host) to be able to discriminate between similar yet

distinct species (ion, molecule). Julius Rebek has successfully designed a series of

cup-shaped molecules (hereafter cavitands) that recognize one another in solution to

form a hydrogen-bonded dimer in solution54

(Figure 1.5). These molecules are

structurally

programmed to

interact with one

another; they are

complimentary in

the same sense that

DNA strands are

complimentary. Rebek

has developed a rich chemistry based on hydrogen-bonded dimeric capsules, which

have been shown to exhibit interesting isomerism,55

accelerate inner-phase

reactions,56

and bind three different guests.57

Lyle Isaacs and coworkers, in the

paper “Self-Sorting: The Exception or the Rule?”,58

have performed an elegant

experiment in which several synthetic systems shown to recognize like molecules

were mixed. Issacs found that despite the chemical similarity of many systems,

thermodynamic self-sorting occurred; this indicates that each of the eight components

in solution (Figure 1.6) exhibited recognition of self.

Figure 1.5. Rebek’s self-recognizing “tennis ball” dimer.

Page 34: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

11

Host-guest binding is similar to molecular recognition, in that both require a

structural matching of at least two components. Guest binding is a specific example

of the more general concept of molecular recognition, in that binding requires a

recognition event as well as an immobilization of a guest by a host. Crown ethers59

and cryptands60

are quintessential examples of molecular hosts that bind metal

cations. Each host has structural features (electron lone pairs) that are selective

toward cations; furthermore, the size and shape of these hosts are adaptable to select

toward cations of a specific size. Table 1.1 describes the binding constant of a series

of cryptands with their preferred group 1 cation.60b

As expected, the selectivity of the

cryptand is intimately related to the size of the host: the larger cations preferentially

bind larger cations. One ethyleneoxy (OCH2CH2) group has dramatic effects on

binding selectivity, showing how small structural changes can impact their binding

properties.

Figure 1.6. Nine molecules included in Isaacs self-sorting study.

Page 35: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

12

Of course, container molecules are a specific class of supramolecular host;

they are capable of binding neutral guest molecules on the basis of size and polarity.

There are few strongly selective structural features in container molecules, mainly the

size of the host portals and the concave nature of the host interior.61

Obviously, van

der Waals forces predominate in container-guest complexes; however, other

interactions (solvophobic effects, ion-dipole, cation-, - interaction) may provide

stability to these complexes.62

1.3.1 Container Molecules: Potentially Useful Supramolecular CP Ligands

While much effort has been exerted in the basic study of CP topology and

supramolecular isomerism (i.e. polymorphism) of CP materials, far less effort has

gone into the incorporation of potentially functional ligands. Many materials have

been derived from simple, aromatic ligands such as terephthalic acid,23b,63

and 4-4’

bipyridine,64

or linear aliphatic ligands such as succinic acid,65

and glutaric acid.66

The incorporation of functional ligands, however, has resulted in more useful

materials (Lin, Hupp). Furthermore, little work has been done to incorporate

macrocyclic ligands. Importantly,

many macrocycles, such as cryptands,

cyclodextrins and cucubiturils, have

well-defined supramolecular functions

Cryptand log Ka Preferred Group 1

Cation (M+)

[2.1.1] 5.5 Li+

(1.36Å)

[2.2.1] 5.4 Na+

(1.90Å)

[2.2.2] 5.4 K+

(2.66Å)

Table 1.1. Cryptand binding in water at 298K

as a function of size.

Page 36: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

13

(atomic or molecular binding, reaction selectivity) that may be exploited in new

macrocycle-based CP materials.

One specific class of macrocycle that has function which may be useful in

CPs are the so-called “container molecules”, which, as the name implies, have the

ability to complex and completely encapsulate small molecular substrates. Container

molecules are hollow molecules with enforcedly rigid cavities that are capable of

binding a variety of smaller molecules based on favorable electrostatic and van der

Waals interactions. Well known supramolecular hosts of this type include

cryptophanes,67

(hemi) carcerands,68

metal-organic polyhedra,69

and hydrogen

bonded self-assembled

capsules.70

Although

container molecules differ

in size, shape and

connectivity, they are all

known to form stable

molecule-within-molecule

complexes.

An intriguing and

potentially useful property

of container molecules is

that they have the capacity

Figure 1.7. Kinetic barrier to decomplexation as a result of

constrictive binding in a container molecule.

Page 37: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

14

Figure 1.8. Conformational gating mechanisms in container molecules that enable

“constrictive binding”.

to strongly hold other species. The binding behavior is not necessarily

thermodynamically driven; the free energy difference between free species and bound

host-guest complexes is generally small (ΔG often < 20 kJ/mol). Instead, the

stability of these container-guest complexes is kinetic in nature. Because of the near

closed-surface nature of these hosts, there are typically high activation barriers to

guest decomplexation. Moreover, even complexes of low intrinsic thermodynamic

stability can exhibit remarkable kinetic stability due to high kinetic barriers

associated with the ingress and egress of guests (Figure 1.7). Donald Cram described

this behavior as “constrictive binding,”61a,71

and defines the constrictive binding

energy as ∆G‡

comp= ∆G‡

decomp- ∆Go. Essentially, for molecules to enter or leave the

cavity, the host must go through a high-energy conformation, behavior that has been

described by Houk as “conformational gating”72

(Figure 1.8).

Binding of gas molecules by supramolecular species is an even more daunting

task than binding of ions or larger neutral molecules, due to the entropic cost of

Page 38: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

15

binding. However, gas encapsulation has been observed in many supramolecules and

Rudkevich has written two reviews related to gas encapsulation in open cavity

species such as α-cyclodextrin, calixarenes, cucubiturils, and “surgically opened”

fullerenes as well as in closed-surface molecules such as carcerands/hemicarcerands,

Rebek’s self-assembled capsules, and cryptophanes.73

Supramolecules that bind gases

effectively generally possess small cavities and are dissolved in solvent molecules

that cannot fit easily into the pore of the supramolecule. This binding was generally

studied in solution by various NMR techniques; few solid-state studies of gas binding

have been reported. While single crystal analyses of gas complexes of clathrate

hydrates, decamethyl-cucubit[5]uril74

and Atwood’s calix[4]arene have been

reported, no single crystal analyses have been performed on container-gas complexes.

Container molecules not

only hold onto molecular species

very strongly, they also strongly

discriminate between similar

molecules on the basis of size and

charge. This binding behavior

may be exploited to perform

interesting chemistry, such as

stereoselective synthesis75

and

stabilization of reactive

Figure 1.9. Selective [2+2] photodimerization inside a

metal-ligand container molecule. This reaction occurs

only in the presence of the container, and yielded only

the cross syn dimer.

Page 39: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

16

intermediates.76

For example, Fujita and coworkers have performed selective [2+2]

photodimerization reactions inside a metal-organic container; one example is

illustrated in Figure 1.9.69b

The photodimerization reaction occurs exclusively inside

Fujita’s polyhedra; no reaction occurs in the bulk solution. Clearly, the container

preorganizes the two substrates in a specific conformation. In contrast, Warmuth and

coworkers have performed the opposite task by protecting reactive intermediates

from reaction by confinement within an unreactive hemicarcerand container.69b,77

Warmuth (and Cram before him) has stabilized and characterized many highly

reactive species, including the antiaromatic cyclobutadiene, benzyne, and carbenes

and nitrenes. The “inner phase” and “outer phase” of a molecular container often

differ dramatically in electronic character; Cram described the inner phase of a

hemicarcerand as a “new phase of matter.”78

Figure 1.10. Left. General structure of (±)-anti cryptophane. Right. General structure of syn

cryptophane. In most common cryptophanes, A = Z = OCH3.

Page 40: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

17

1.4. Cryptophanes

Cryptophanes are one class of container molecules, initially reported in 1981

and intensively studied by the late Andre Collet.79

They are composed of two axially

chiral C3 symmetric cyclotribenzylenes (hereafter CTBs), that are connected to one

another by three spacers. Each cryptophane has two chiral units, resulting in two

diastereomers: the syn form and the anti form (Figure 1.10). The syn diastereomer

has the groups A and Z on the same side of the bridging unit (X), while the anti

diastereomer has the groups A and Z on opposite sides with respect to the bridge.

The syn diastereomer is achiral (C3h) when A = Z, but is chiral (C3) when A and Z are

not the same (Figure 1.10). The anti diastereomer is inherently chiral, with C3

symmetry when A and Z are different and D3 symmetry when A = Z. It is important

to note that unless otherwise stated, cryptophane syntheses result in both anti

enantiomers. Table 1.2 describes the notation for racemic anti-cryptophane

molecules.

Cryptophanes are essentially spheroidal molecules that have a distinct “inner

phase” with “equatorial”

and “polar” regions (Figure

1.11). The “inner phase” is

the interior of the

cryptophane, which has two

concave surfaces from the

Figure 1.11. Descriptions of cryptophane regions.

Page 41: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

18

Symbol Description Example

(±)-anti Racemic mixture of cryptophanes (±)-anti-4

Cryptophane encapsulates following guest molecule (±)-anti-4THF

Table 1.2. Description of symbols used throughout text.

CTB “poles”. The equatorial region consists of the three bridges and three container

openings or pores. These different regions often have truly different chemical and

electronic behavior. Within the inner phase, there are large differences in sterics and

electronics in the polar regions of the container relative to the equatorial region.

Cryptophane molecules have a rich history of host-guest chemistry, with

behavior that is similar to that of other container molecules. A variety of cryptophane

molecules have been synthesized, in which variations have been made to the three

bridges (X, see Figure 1.10), as well as to the CTB subunits (A, Z, Figure 1.10), and

Collet gave the cryptophanes a letter designation based upon their order of synthesis

(Cryptophane A, B, C, etc.). In general, ∆G‡

decomp decreases with an increase in

bridge length as the opening to the container becomes larger and less restrictive to

guest entry. Cryptophanes, like other container molecules, exhibit a strong size

dependence on molecular binding, as shown in Table 1.3.80

Though each

Cryptophane X X’ Y Ideal Guest Ideal Guest

Volume(Å3)

-∆Go (kJ/mol)

A OCH3 OCH3 (CH2)2 Xe 41 20.6

C H OCH3 (CH2)2 CH2Cl2 57 15.4

E

(neutral guests) OCH3 OCH3 (CH2)3 CHCl3 72 15.3

E

(cationic guests) OCH3 OCH3 (CH2)3 N(CH4)4

+ 96 30.9

Table 1.3. Free energy of complexation for various cryptophanes in (CDCl2)2at 300K.65

Page 42: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

19

cryptophane is capable of binding guests of varying size (neutral cryptophane E

guests range from 40 to 95 Å3), it is clear that each cryptophane has a guest of

optimum size that is bound most effectively. It is also important to note that

cryptophane A shows a substantial affinity to bind gases such as xenon (ΔG ~ 20

kJ/mol) and methane (ΔG ~ 12 kJ/mol).81

Cryptophanes’ molecular recognition and binding have been exploited in

several situations. Recently, biofunctionalized cryptophane molecules have been

used as a biosensor.82

Remarkably, xenon encapsulated within the cryptophane

exhibits measurable changes in the Xe NMR chemical shift based on the binding

events on the exterior of the cryptophane. For instance, a biotin-functionalized

cryptophane has been introduced to the protein avidin, and the Xe NMR response of

the biotinylated cryptophane AXe has been shown to shift downfield and broaden

after the functionalized cryptophane has been introduced to avidin (See Table 1.2 for

description). Also, a specifically designed cryptophane molecule ((+)-Cryptophane

C) was used to bind the racemic CHFClBr83

and the difference in relative stability of

the resulting diastereomeric complexes was calculated (ΔΔG~1.1 kJ/mol). The

enantiomeric excess of (+)-CHFClBr was calculated by NMR (ee = 4.3±1%), which

allowed for the experimental determination of the halomethane’s specific optical

rotation.

Page 43: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

20

1.5. Research project: Synthesis and Characterization of Cryptophane

Container-Based Materials

The overall purpose or goal of this research project is multipronged. First,

new exo-carboxylic acid functionalized m-xylyl bridged cryptophane molecules and

materials have been synthesized and characterized, with the ultimate goal of using

these container molecule ligands to synthesize rigid, porous coordination polymers.

Since cryptophanes are spheroidal in shape, the resulting cryptophane-based solid

was hypothesized to pack inefficiently and not be able to generate interpenetrating

coordination networks. The resulting material should be porous in nature.

Furthermore, the network would also be imparted with container molecules that are

capable of size and shape-selective binding. The resulting materials should have

interesting structural features and functionality.

We chose to synthesize the exo-carboxylic acid cryptophane species as metal-

carboxylates can often form well-defined SBUs such as the copper-carboxylate

paddlewheel and the zinc-carboxylate octahedron. Knowing the potential SBUs for

zinc-carboxylates, and that the cryptophane has a C3 axis allows structural predictions

to be made about the potential cryptophane-based materials.

A second goal for this research project was to identify and possibly quantify

the “constrictive binding” that occurs in cryptophane based materials, as it is our

hypothesis that constrictive binding is a universal container behavior, not simply a

solution-based behavior of containers. This study requires that these materials be

Page 44: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

21

studied by thermal techniques, such as differential scanning calorimetry (DSC) and

thermogravimetric analysis (TGA) to determine the kinetics of

desolvation/decomplexation in cryptophane-based materials. Structural information

about the cryptophaneguest complex is important information in this study, as it

confirms the cryptophaneguest stoichiometry as well as the guest(s) conformation

inside the cryptophane host. The material’s structure is also crucial in that it may

identify additional lattice solvent molecules whose desolvation may overlap or

interfere with the included solvent guest.

Page 45: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

22

1.6 References

1. Payra, P.; Dutta, P., Zeolites. In Handbook of Zeolite Science and Technology,

CRC Press: 2003.

2. Cotton, F. A.; Wilkinson, G., Advanced inorganic chemistry. 5th ed.; Wiley:

New York, 1988.; p 283-286.

3. Dyer, A., An introduction to zeolite molecular sieves. J. Wiley: Chichester ;

New York, 1988; p 1-149

4. Nishi, K.; Hidaka, A.; Yokomori, Y., Acta Crystallogr., Sect B: Struct Sci

2005, 61, 160-163.

5. Ciullo, P. A., Industrial Minerals and Their Uses - A Handbook and

Formulary. William Andrew Publishing/Noyes: 1996; p 78-81.

6. Doraiswamy, L. K., Organic Synthesis Engineering. Oxford University Press.

2001, 129-142.

7. Shriver, D. F.; Atkins, P. W.; Langford, C. H., Inorganic Chemistry. 2nd ed.;

W. H. Freeman and Company: New York, 1994.

8. Chiusoli, G. P.; Maitlis, P. M.; Metal-catalysis in industrial organic

processes. RSC Pub.: Cambridge, 2006.

9. Soler-Illia, G. J. d. A. A.; Sanchez, C.; Lebeau, B.; Patarin, J., Chem. Rev.

2002, 102, 4093-4138.

10. Eddaoudi, M.; Moler, D. B.; Li, H.; Chen, B.; Reineke, T. M.; O'Keeffe, M.;

Yaghi, O. M., Acc. Chem. Res. 2001, 34, 319-30.

Page 46: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

23

11. Schmidt, G. M. J., Pure Appl. Chem. 1971, 27, 647.

12. (a) Braga, D.; Grepioni, F.; Desiraju, G. R., Chem. Rev. 1998, 98, 1375-1406;

(b) Moulton, B.; Zaworotko, M. J., Chem. Rev. 2001, 101, 1629-58.

13. Maddox, J., Nature 1988, 335, 201.

14. Etter, M. C., Acc. Chem. Res. 1990, 23, 120-126.

15. Seddon, K. R.; Zaworotko, M. J., Crystal Engineering: The Design and

Application of Functional Solids. Kluwer Academic Publishers: Dordrecht,

1999; Vol. 539, p 505.

16. Steiner, T., Angew. Chem., Int. Ed. 2002, 41, 48-76.

17. Giacovazzo, C., Fundamentals of crystallography. International Union of

Crystallography; Chester, England, 1992, p. 654.

18. Bailey, M.; Brown, C. J., Acta Crystallogr., Sect B: Struct Sci 1967, 22, 387-

391.

19. Duchamp, D. J.; Marsh, R. E., Acta Crystallogr., Sect. B: Struct Sci 1969, 25,

5-19.

20. Ermer, O., J. Am. Chem. Soc. 1988, 110, 3747-3754.

21. (a) Desiraju, G. R., Crystal engineering : the design of organic solids.

Elsevier: Amsterdam ; New York, 1989; p pp. 1-312; (b) Hofmann, D. W.

M.; Kuleshova, L. N.; Antipin, M. Y., Cryst. Growth Des. 2004, 4, 1395-

1402; (c) Zerkowski, J. A.; MacDonald, J. C.; Seto, C. T.; Wierda, D. A.;

Whitesides, G. M., J. Am. Chem. Soc. 1994, 116, 2382-2391.

Page 47: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

24

22. (a) Brunet, P.; Simard, M.; Wuest, J. D., ibid.1997, 119, 2737-2738; (b)

Endo, K.; Koike, T.; Sawaki, T.; Hayashida, O.; Masuda, H.; Aoyama, Y., J.

Am. Chem. Soc. 1997, 119, 4117-4122; (c) Venkataraman, D.; Lee, S.;

Zhang, J.; Moore, J. S., Nature 1994, 371, 591-593.

23. (a) James, S. L., Chem. Soc. Rev. 2003, 32, 276-288; (b) Li, H.; Eddaoudi,

M.; O'Keeffe, M.; Yaghi, O. M., Nature 1999, 402, 276-279.

24. Davis, M. E., ibid.2002, 417, 813-821.

25. Hagen, J., Industrial catalysis : a practical approach. Wiley-VCH: Weinheim

; New York, 1999; p pp. 1-416

26. (a) Dyer, A.; Emms, T. I., J. Mater. Chem. 2005, 15, 5012-5021; (b) Jiang,

M.; Tatsumi, T., J. Phys. Chem. B 1998, 102, 10879-10884.

27. (a) Calero, S.; Smit, B.; Krishna, R., Phys. Chem. Chem. Phys. 2001, 3, 4390-

4398; (b) Huddersman, K.; Klimczyk, M., J. Chem. Soc., Faraday Trans.

1996, 92, 143-147.

28. Hoskins, B. F.; Robson, R., J. Am. Chem. Soc. 1990, 112, 1546-1554.

29. Shriver, D. F., Struct. Bonding 1, 32.

30. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O'Keeffe, M.;

Yaghi, O. M., Science 2002, 295, 469-72.

31. (a) Custelcean, R.; Gorbunova, M. G., J. Am. Chem. Soc. 2005, 127, 16362-

16363; (b) Custelcean, R.; Moyer, B. A.; Bryantsev, V. S.; Hay, B. P., Cryst.

Growth Des. 2005, 6, 555-563; (c) Horike, S.; Matsuda, R.; Tanaka, D.;

Page 48: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

25

Mizuno, M.; Endo, K.; Kitagawa, S., J. Am. Chem. Soc. 2006, 128, 4222-

4223.

32. Reineke, T. M.; Eddaoudi, M.; Moler, D.; O'Keeffe, M.; Yaghi, O. M., J. Am.

Chem. Soc. 2000, 122, 4843-4844.

33. (a) Maspoch, D.; Ruiz-Molina, D.; Veciana, J., J. Mater. Chem. 2004, 14,

2713-2723; (b) Zhao, W.; Song, Y.; Okamura, T.-a.; Fan, J.; Sun, W.-Y.;

Ueyama, N., Inorg. Chem. 2005, 44, 3330-3336.

34. (a) Kepert, C. J.; Rosseinsky, M. J., Chem. Commun 1999, 375-376; (b)

Kondo, M.; Okubo, T.; Asami, A.; Noro, S.-i.; Yoshitomi, T.; Kitagawa, S.;

Ishii, T.; Matsuzaka, H.; Seki, K., Angew. Chem., Int. Ed. 1999, 38, 140-143;

(c) Lin, X.; Blake, A. J.; Wilson, C.; Sun, X. Z.; Champness, N. R.; George,

M. W.; Hubberstey, P.; Mokaya, R.; Schröder, M., J. Am. Chem. Soc. 2006,

128, 10745-10753.

35. Lee, S. J.; Lin, W., J. Am. Chem. Soc. 2002, 124, 4554-4555.

36. (a) Alaerts, L.; Séguin, E.; Poelman, H.; Thibault-Starzyk, F.; Jacobs, P. A.;

De Vos, D. E., Chem--Eur. J. 2006, 12, 7353-7363; (b) Gómez-Lor, B.;

Gutiérrez-Puebla, E.; Iglesias, M.; Monge, M. A.; Ruiz-Valero, C.; Snejko,

N., Chem. Mater. 2005, 17, 2568-2573; (c) Perles, J.; Iglesias, M.; Martín-

Luengo, M.-Á.; Monge, M. Á.; Ruiz-Valero, C.; Snejko, N., Chem. Mater.

2005, 17, 5837-5842.

Page 49: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

26

37. (a) Chen, B.; Liang, C.; Yang, J.; Contreras, D. S.; Clancy, Y. L.; Lobkovsky,

E. B.; Yaghi, O. M.; Dai, S., Angew. Chem., Int. Ed. 2006, 45, 1390-3; (b)

Pan, L.; Olson, D. H.; Ciemnolonski, L. R.; Heddy, R.; Li, J., Angew. Chem.,

Int. Ed. 2006, 45, 616-619.

38. (a) Kubota, Y.; Takata, M.; Matsuda, R.; Kitaura, R.; Kitagawa, S.; Kato, K.;

Sakata, M.; Kobayashi, T. C., ibid.2005, 44, 920-923; (b) Matsuda, R.;

Kitaura, R.; Kitagawa, S.; Kubota, Y.; Belosludov, R. V.; Kobayashi, T. C.;

Sakamoto, H.; Chiba, T.; Takata, M.; Kawazoe, Y.; Mita, Y., Nature 2005,

436, 238-241; (c) Uemura, T.; Kitaura, R.; Ohta, Y.; Nagaoka, M.; Kitagawa,

S., Angew. Chem., Int. Ed. 2006, 45, 4112-4116.

39. Zhao, X.; Xiao, B.; Fletcher, A. J.; Thomas, K. M.; Bradshaw, D.;

Rosseinsky, M. J., Science 2004, 306, 1012-1015.

40. (a) Dinc , M.; Long, J. R., J. Am. Chem. Soc. 2005, 127, 9 -9 ; (b)

Dinc , M.; Yu, A. F.; Long, J. R., J. Am. Chem. Soc. 2006, 128, 8904-8913.

41. (a) Moon, H. R.; Kobayashi, N.; Suh, M. P., Inorg. Chem. 2006, 45, 8672-

8676; (b) Navarro, J. A. R.; Barea, E.; Salas, J. M.; Masciocchi, N.; Galli, S.;

Sironi, A.; Ania, C. O.; Parra, J. B., Inorg. Chem. 2006, 45, 2397-2399; (c)

Pan, L.; Parker, B.; Huang, X.; Olson, D. H.; Lee; Li, J., J. Am. Chem. Soc.

2006, 128, 4180-4181.

42. (a) Hu, A.; Ngo, H. L.; Lin, W., ibid.2003, 125, 11490-11491; (b) Kesanli,

B.; Lin, W., Coord. Chem. Rev. 2003, 246, 305-326.

Page 50: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

27

43. Cho, S.-H.; Ma, B.; Nguyen, S. T.; Hupp, J. T.; Albrecht-Schmitt, T. E.,

Chem. Commun 2006, 2563-2565.

44. (a) Moulton, B.; Lu, J.; Hajndl, R.; Hariharan, S.; Zaworotko, M. J., Angew.

Chem., Int. Ed. 2002, 41, 2821-4; (b) Perry, J. J.; McManus, G. J.;

Zaworotko, M. J., Chem. Commun 2004, 2534-5.

45. Liu, Y.; Kravtsov, V. C.; Beauchamp, D. A.; Eubank, J. F.; Eddaoudi, M., J.

Am. Chem. Soc. 2005, 127, 7266-7267.

46. Albrecht, M.; Lutz, M.; Spek, A. L.; van Koten, G., Nature 2000, 406, 970-

974.

47. (a) Cram, D. J., Angew. Chem., Int. Ed. Engl. 1988, 27, 1009-1020; (b) Lehn,

J.-M., Angew. Chem., Int. Ed. Engl. 1988, 27, 89-112; (c) Pedersen, C. J.,

Angew. Chem., Int. Ed. Engl. 1988, 27, 1021-1027.

48. (a) Lehn, J.-M., Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 4763-4768; (b) Lehn,

J. M., Supramolecular chemistry : concepts and perspectives : . VCH:

Weinheim ; New York, 1995; p p. 1-271

49. Campbell, N. A., Biology. 3rd ed.; Benjamin/Cummings: Redwood City,

Calif., 1993; p 1-1190.

50. Creighton, T. E., Proteins. W. H. Freeman and Company: New York.

51. Fields, B. A.; Goldbaum, F. A.; Dall'Acqua, W.; Malchiodi, E. L.; Cauerhff,

A.; Schwarz, F. P.; Ysern, X.; Poljak, R. J.; Mariuzza, R. A., Biochemistry

1996, 35 (48), 15494-15503.

Page 51: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

28

52. (a) Buckingham, A. D.; Legon, A. C.; Roberts, S. M., Principles of molecular

recognition. 1st ed.; Blackie Academic & Professional: London ; New York,

1993; p 1-200 ; (b) Ikeda, A.; Shinkai, S., Chem. Rev. 1997, 97, 1713-1734;

(c) Kruppa, M.; König, B., Chem. Rev. 2006, 106, 3520-3560; (d) Gale, P.

A., Acc. Chem. Res. 2006, 39, 465-475.

53. (a) Izatt, R. M.; Bradshaw, J. S.; Pawlak, K.; Bruening, R. L.; Tarbet, B. J.,

Chem. Rev. 1992, 92, 1261-1354; (b) Izatt, R. M.; Pawlak, K.; Bradshaw, J.

S.; Bruening, R. L., Chem. Rev. 1995, 95, 2529-2586; (c) Pluth, M. D.;

Raymond, K. N., Chem. Soc. Rev. 2007, 36, 161-171; (d) Vögtle, F.; Weber,

E., Host guest complex chemistry/macrocycles : synthesis, structures,

applications. Springer: Berlin ; New York, 1985; p 421 p.

54. (a) Branda, N.; Wyler, R.; Rebek, J., Jr., Science 1994, 263, 1267-8; (b)

Meissner, R. S.; Rebek, J., Jr.; de Mendoza, J., Science 1995, 270, 1485-8.

55. (a) Shivanyuk, A.; Rebek, J., Jr., J. Am. Chem. Soc. 2002, 124, 12074-5; (b)

Yamanaka, M.; Rebek, J., Jr., Chem. Commun 2004, 1690-1; (c) Yamanaka,

M.; Shivanyuk, A.; Rebek, J., Jr., Proc. Natl. Acad. Sci. U. S. A. 2004, 101,

2669-72.

56. (a) Kang, J.; Rebek, J., Nature 1997, 385, 50-52; (b) Kang, J.; Santamaría, J.;

Hilmersson, G.; Rebek, J., J. Am. Chem. Soc. 1998, 120, 7389-7390.

57. Amaya, T.; Rebek, J., Jr., Chem. Commun 2004, 1802-3.

58. Wu, A.; Isaacs, L., J. Am. Chem. Soc. 2003, 125, 4831-4835.

Page 52: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

29

59. (a) Pedersen, C. J., ibid.1967, 89, 7017-7036; (b) Weber, E.; Patai, S.;

Rappoport, Z., Crown ethers and analogs. Wiley: Chichester West Sussex ;

New York, 1989.

60. (a) Lehn, J. M.; Sauvage, J. P., J. Am. Chem. Soc. 1975, 97, 6700-6707; (b)

Lehn, J. M.; Sauvage, J. P.; Dietrich, B., J. Am. Chem. Soc. 1970, 92, 2916-

2918.

61. (a) Cram, D. J.; Tanner, M. E.; Knobler, C. B., ibid.1991, 113, 7717-7727;

(b) Garcia, C.; Humiliere, D.; Riva, N.; Collet, A.; Dutasta, J.-P., Org.

Biomol. Chem. 2003, 1, 2207-2216.

62. (a) Garel, L.; Lozach, B.; Dutasta, J. P.; Collet, A., J. Am. Chem. Soc. 1993,

115, 11652-11653; (b) Gosse, I.; Dutasta, J.-P.; Perrin, M.; Thozet, A., New

J. Chem. 1999, 23, 545-548; (c) Lee, J. W.; Samal, S.; Selvapalam, N.; Kim,

H.-J.; Kim, K., Acc. Chem. Res. 2003, 36, 621-630; (d) Yoon, J.; J. Cram, D.,

Chem. Commun 1997, 497-498.

63. Reineke, T. M.; Eddaoudi, M.; Fehr, M.; Kelley, D.; Yaghi, O. M., J. Am.

Chem. Soc. 1999, 121, 1651-1657.

64. (a) Subramanian, S.; Zaworotko, M. J., Angew. Chem., Int. Ed. Engl. 1995,

34, 2127-2129; (b) MacGillivray, L. R.; Groeneman, R. H.; Atwood, J. L., J.

Am. Chem. Soc. 1998, 120, 2676-2677.

Page 53: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

30

65. (a) Bowden, T. A.; Milton, H. L.; Slawin, A. M. Z.; Lightfoot, P., Dalton

Trans. 2003, 936-939; (b) Zhou, Z.-H.; Yang, J.-M.; Wan, H.-L., Cryst.

Growth Des. 2005, 5, 1825-1830.

66. Rather, B.; Zaworotko, M. J., Chem. Commun 2003, 830-1.

67. (a) Atwood, J. L.; Lehn, J. M., Comprehensive supramolecular chemistry. 1st

ed.; Pergamon: New York, 1996; Vol. 2; (b) Collet, A., Tetrahedron 1987,

43, 5725-5759.

68. (a) Cram, D. J.; Cram, J. M.; Royal Society of Chemistry (Great Britain),

Container molecules and their guests. Royal Society of Chemistry:

Cambridge, 1994; p 1-223; (b) Jasat, A.; Sherman, J. C., Chem. Rev. 1999,

99, 931-968.

69. (a) Caulder, D. L.; Raymond, K. N., Acc. Chem. Res. 1999, 32, 975-982; (b)

Fujita, M.; Umemoto, K.; Yoshizawa, M.; Fujita, N.; Kusukawa, T.; Biradha,

K., Chem. Commun 2001, 509-518; (c) Leininger, S.; Olenyuk, B.; Stang, P.

J., Chem. Rev. 2000, 100, 853-908.

70. Conn, M. M.; Rebek, J., Jr., ibid.1997, 97, 1647-1668.

71. Quan, M. L. C.; Cram, D. J., J. Am. Chem. Soc. 1991, 113, 2754-2755.

72. Houk, K. N.; Nakamura, K.; Sheu, C.; Keating, A. E., Science 1996, 273, 627-

629.

73. (a) Rudkevich, D. M., Angew. Chem., Int. Ed. 2004, 43, 558-571; (b)

Rudkevich, D. M.; Leontiev, A. V., Aust. J. Chem. 2004, 57, 713-720.

Page 54: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

31

74. Miyahara, Y.; Abe, K.; Inazu, T., Angew. Chem., Int. Ed. 2002, 41, 3020-

3023.

75. (a) Chen, J.; Rebek, J., Jr., Org. Lett. 2002, 4, 327-9; (b) Yoshizawa, M.;

Takeyama, Y.; Okano, T.; Fujita, M., J. Am. Chem. Soc. 2003, 125, 3243-

3247.

76. (a) Warmuth, R.; Yoon, J., Acc. Chem. Res. 2001, 34, 95-105; (b) Cram, D.

J.; Tanner, M. E.; Thomas, R., J. Am. Chem. Soc. 1991, 103, 1048-51.

77. (a) Warmuth, R., Chem. Commun 1998, 59-60; (b) Warmuth, R.; Marvel, M.

A., Angew. Chem., Int. Ed. 2000, 39, 1117-1119.

78. Sherman, J. C.; Cram, D. J., J. Am. Chem. Soc. 1989, 111, 4527-4528.

79. Gabard, J.; Collet, A., J. Chem. Soc., Chem. Commun. 1981, 1137-1139.

80. Holman, K. T. “Cryptophanes: Molecular Containers” In Encyclopedia of

Supramolecular Chemistry; J. A. Atwood, J. W. Steed, Eds., Marcel Dekker:

New York, NY 2004; pp. 340-348. .

81. Garel, L.; Dutasta, J.-P.; Collet, A., Angew. Chem., Int. Ed. Engl. 1993, 32,

1169-1171.

82. (a) Spence, M. M.; Rubin, S. M.; Dimitrov, I. E.; Ruiz, E. J.; Wemmer, D. E.;

Pines, A.; Yao, S. Q.; Tian, F.; Schultz, P. G., Proc. Natl. Acad. Sci. U.S.A.

2001, 98, 10654-10657; (b) Spence, M. M.; Ruiz, E. J.; Rubin, S. M.;

Lowery, T. J.; Winssinger, N.; Schultz, P. G.; Wemmer, D. E.; Pines, A., J.

Am. Chem. Soc. 2004, 126, 15287-15294.

Page 55: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

32

83. Canceill, J.; Lacombe, L.; Collet, A., ibid.1985, 107, 6993-6996.

Page 56: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

33

CHAPTER 2: SYNTHESIS, CHARACTERIZATION AND INCLUSION

COMPOUNDS OF M-XYLYL BRIDGED CRYPTOPHANES

2.1. Introduction: Synthesis and characterization of cryptophanes

There are three published synthetic pathways that have been used to synthesize

cryptophanes: the “two-step method”,1 the “template method”

2 and the so-called

“capping method”3 (Scheme 2.1). Each method has its own relative strengths and

weaknesses; however, all three methods require the formation of cup-like CTB

subunits. The CTB cyclization reaction is generally an acid-catalyzed dehydration and

cyclization of vanillyl alcohol derivatives, with vanillyl alcohol being 3-methoxy-4-

Scheme 2.1. Three potential synthetic pathways to cryptophane molecules. In most cases, E = E’

= OCH3.

Page 57: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

34

hydroxybenzyl alcohol. Three vanillyl groups react regioselectively to make up each

CTB subunit.

The two-step method is the middle pathway shown in Scheme 2.1, and can be

described as initially synthesizing one third of the cryptophane by joining one bridging

subunit with two cyclotribenzylene (CTB) precursors (usually 3-methoxy-4-

hydroxybenzyl alcohol, commonly known as vanillyl alcohol). The second step is the

cryptophane synthesis, in which the cryptophane molecule(s) are synthesized by the

cyclization reaction that forms both CTB subunits by performing six regioselective

electrophilic aromatic substitution reactions. This reaction synthesizes cryptophane

molecules in the fewest number of synthetic steps, and is most useful in the synthesis

of symmetric cryptophanes, which is a cryptophane in which all bridges are identical.

However, the cryptophane cyclization reaction yields are low (<20%) and generally

produce different quantities of syn and anti diastereomers, as a function of bridge

length, and the two-step method usually produces more of the anti diastereomer.

The template method is more challenging and time-consuming synthetically

(Scheme 2.1, top), but is usually higher yielding in the final cyclization step. Also, the

template method can be used to create asymmetric cryptophanes that have different

bridging subunits.4 The template in the so-called template method is an already

synthesized CTB subunit, which holds the three CTB precursors in close proximity to

one another and makes the cryptophane cyclization reaction more successful when

performed under highly dilute conditions.

Page 58: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

35

The capping method (Scheme 2.1, bottom) brings one CTB subunit together

with a second CTB subunit. The advantage of this type of reaction is that it allows for

differentiation of the two cryptophane caps. Also, interesting chemistry can be used to

perform the capping reaction. Shinkai provides one example in which a Pd-pyridine

interaction was used to assemble novel cryptophane.5

The 1H NMR spectra of cryptophane molecules and related cup-shaped CTB

molecules have several characteristic features. The methylene groups at the base of the

[1.1.1]orthocyclophane cups exhibit an unusual NMR splitting (Figure 2.1); the

equatorial proton gives a doublet at approximately 3.5 ppm while the axial proton

gives a doublet at approximately 4.7 ppm.6 The axial protons are in such close

proximity to one another that their electrons repel, resulting in proton deshielding from

mutual electron repulsion. Consequently, Ha resonate significantly downfield (~1.2

ppm) from He. Symmetric cryptophanes (Figure 1.10) are also characterized by two

singlets (or weakly coupling doublets) resulting from arene protons on the CTB, while

Figure 2.1. 1H

NMR of cryptophane E (propyl-bridged).

Page 59: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

36

the remaining protons come from the bridges. The NMR spectrum of a symmetric

cryptophane (cryptophane E, Figure 2.1) has relatively few peaks for such a large

molecule, due to its high (D3) point group symmetry.

The 1H NMR shown in Figure 2.1 was performed in (CDCl2)2, and has two

interesting NMR peaks that are identified as free and bound CHCl3, which were

observed at about 7.3 and 2.8 ppm, respectively. The two NMR peaks indicate that the

chloroform binding is in a slow exchange regime, and the highly upfield shift (~4.5

ppm) of bound chloroform is a result of its position within the highly shielding

cryptophane molecule. This NMR demonstrates the selective guest binding of

cryptophane E for chloroform relative to d2-1,1,2,2 tetrachloroethane, which has two

additional non-hydrogen atoms. Variable temperature 200 MHz 1H NMR experiments

were performed, and the chloroform peaks disappeared above 360K (the fast exchange

chloroform peak could not be obtained due to temperature restraints). The stability

constant of the cryptophane-ECHCl3 in (CDCl2)2 is 470 M-1

.7

2.2. Solution vs. Solid-State Analyses of Supramolecular Host-Guest

Complexes

Most host-guest binding studies to date have been performed in solution. In

solution, one observes the “real life” binding of a guest molecule to its host, as most

biomolecular and molecular hosts are designed to bind guest(s) in solution. Even more

importantly, quantification of the binding constant and the free energy of binding are

Page 60: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

37

possible through solution studies.8 Furthermore, a large number of techniques are

available to study the solution binding of these systems (NMR,9 Isothermal Titration

Calorimetry or ITC,10

and UV/VIS11

).

Though solution-based binding studies are important for the reasons stated

above, many factors can complicate these experiments. In container molecules, the

guest molecules are neutral organic species that are often similar to the solvent. Often,

a noncompeting solvent is employed; however, such a solvent is not always available.

If the bulk solvent competes for binding, it can be impossible to determine the binding

constant for a specific guest. Also, large pore container molecules do not

constrictively bind as effectively as cryptophanes with shorter bridges. The resulting

host-guest complexes, therefore, are not as kinetically stable and the guest/solvent is in

fast exchange with the bulk solvent.

An alternative to NMR binding studies is a crystallographic analyses of host-

guest complexes. One can observe the host conformations sampled in various host-

guest complexes.12

One obtains a “snapshot” of the guest molecule inside of the host,

which allows for an analysis of the host-guest interactions governing the complexation

event. Furthermore, understanding the structure of the host-guest complex may help

explain the solid-state behavior (i.e. desolvation) in host-guest materials.

2.3. History of Crystallographic Analysis of Container Molecule Host-Guest

Complexes

Page 61: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

38

While cryptophanes have existed for 25 years, there are surprisingly few crystal

structures of cryptophanes reported in the Cambridge Structural Database (CSD). In

fact, only fifteen cryptophane structures have been reported.13

This lack of data on

container molecule complexes is not limited to cryptophanes. Another container

molecule, [hemi]carcerands (Figure 2.2), have been extensively studied; however,

there are only 29 [hemi]carcerand crystal structures deposited in the CSD.14

Several

factors may contribute to this low number, including not reporting crystal structures to

the CSD and the relative difficulty in obtaining high quality crystals of fairly large

container molecules.

Figure 2.2. Left: Alkynyl bridged hemicarcerand, which is defined by the C4 symmetric cavitands

that make up the polar regions of the container. Right. Crystal structure (YETKAJ) describing

alkynyl bridged hemicarcerand. The encapsulated molecule is 1,1,2,2 tetrachloroethane.

Page 62: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

39

Cram and coworkers performed a detailed structural analysis of

hemicarceplexes in a paper entitled “Correlations of Structure with Binding Ability

Involving Nine Hemicarcerand Hosts and Twenty-Four Guests,”15

reported in 1997.

This work looks at the effects of different bridges (aliphatic vs. aromatic), different

bowl spanners (OCH2O, O(CH2)2O, O(CH2)3O), and of asymmetric containers on

guest complexation and host conformation. This comprehensive analysis successfully

explained the difference in chemical shift between free and incarcerated guest signals

in 1H NMR as a function of their position in the cavity.

The other major crystallographic study that included hemicarcerands and

cryptophanes was a CCD database analysis of supramolecular systems. Nishio, in

“CH- Interactions as Demonstrated in the Crystal Structures of Host-Guest

Compounds: A Database Study,”16

examined the CH- interactions in a variety of

systems, including calix[4]arenes, cyclodextrins, pseudorotaxanes as well as

cryptophanes. Nishio found a number of short CH- contacts in these systems,

indicative of weak interactions between host and guest.

2.4. Syntheses of m-xylyl bridged cryptophanes

o-, m-, and p-xylyl bridged cryptophanes have been synthesized previously.17

An endo-acid functionalized m-xylyl bridged cryptophane was reported by Weber and

coworkers, and provides an interesting contrast to the exo-functionalized species that

this research targeted. The endo-functionalized m-xylyl cryptophane, seen in Figure

Page 63: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

40

Figure 2.3. Structure of Weber’s endo-carboxylate m-

xylyl cryptophane. Hydrogen atoms and guest

molecules omitted for clarity. Carbon: Gray;

Oxygen: Red.

2.3, points its carboxylate functional groups into the cavity of the cryptophane. The

resulting cryptophane was studied for its metal-cation binding properties, and a crystal

structure of the cryptophane showed that the cryptophane had a crystallographic space

group of P-6, which reflects the idealized point group symmetry of the molecule (C3h).

In contrast, the synthetic targets in this research are exo-functionalized m-xylyl bridged

cryptophanes, which can be used to create three-dimensional materials that are

embedded with container-based molecules. To create an exo-functionalized

cryptophane, the m-xylyl bridges were designed such that the appropriate functional

groups were meta to each of the xylyl groups, which placed the functional groups

normal to the C3 axis of the cryptophane. CPK models strongly suggested that the

functional groups would be unable to

turn into the cryptophane, and that

they would point away from the

cryptophane pore. The resulting exo-

functionalized cryptophane could

then be viewed as a three-fold

organic spacer, similar to trimesic

acid, and plausible materials can be

designed from these cryptophane

spacers, in theory.

Specifically, the exo-acid cryptophane was targeted for its potential in the

Page 64: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

41

Scheme 2.2. Synthesis of m-xylyl bridged cryptophanes.

design of metal-carboxylate materials, or in hydrogen-bonded networks. The exo-

bromobenzene cryptophane was targeted because of the potential for additional

chemistry via cross coupling reactions such as Suzuki, Stille or Negishi coupling

Page 65: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

42

reactions.18

For example, coupling chemistry could generate exo-functionalized

cryptophanes that have interesting functionalization that is quite distant from the

cryptophane pore. Also, the bridging reactions can be used to systematically examine

the effect of functional group position and type on their corresponding solid-state

materials.

Syntheses of various m-xylyl bridged cryptophanes were performed using the

“two-step” method (Scheme 2.2). This method was chosen because the desired

symmetric cryptophanes could be synthesized and isolated in approximately 1-2

weeks. Appropriately functionalized bis(bromomethyl)benzenes19

were reacted with

two equivalents of vanillyl alcohol to yield the corresponding diols in good yields (75-

90%). The diol molecules are essentially a monomer in a cyclic trimerization, as they

correspond to ~1/3 of the cryptophane molecule. The cyclization reaction requires an

acid catalyst, which is typically formic acid. In fact, formic acid acts both as solvent

and catalyst. This reaction is performed at high dilution (~1 mM) to minimize

unwanted polymerization side-reactions.

The cyclization reactions forming exo-ester cryptophanes (±)-anti-4 and syn-5,

m-xylyl bridged (±)-anti-6 and syn-7, and exo-bromo cryptophane (±)-anti-8 were

somewhat successful. The cryptophanes were isolated by flash column

chromatography; followed by recrystallization from the appropriate solvent. The

overall yield of cryptophane molecules is low (1-12%), and these yields were

somewhat lower than reported yields for other cryptophanes synthesized using the two-

Page 66: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

43

step method (2-20%).1 The lower yield for m-xylyl bridged cryptophanes relative to

aliphatic bridged cryptophanes is not surprising, because unwanted electrophilic

aromatic substitution (EAS) reactions can occur on the aromatic bridges. The anti

isomers were higher yielding than the syn isomers, and the syn form of the exo-bromo

cryptophane was never isolated (Scheme 2.2).

The ester-functionalized cryptophanes (±)-anti-4 and syn-5 required a

deprotection step before it could be employed as a coordinating ligand. Acid-catalyzed

Fischer de-esterification was not successful in the synthesis of (±)-anti-H39 or syn-

H310; base-catalyzed hydrolysis with NaOH resulted in an intractable solid. The

deprotection was successfully completed, however, in two steps as shown in Scheme

2.2: Saponification was performed with NMe4OH followed by immediate acidification

with aqueous HCl. The organic base was employed for its improved solubility

properties in organic solvents, as the tristetramethylammonium cryptophane salt is

somewhat soluble in organic solvents. 1H NMR confirmed the formation of (±)-anti-

H39 and syn-H310, most notably through the loss of the ester peak near δ 3.9 ppm.

1H NMR spectra for these cryptophanes are similar and representative spectra

of (±)-anti-4 and syn-5 are shown in Figure 2.4. It is important to note that in the

absence of a chiral shift reagent, NMR alone cannot distinguish between the syn and

anti diastereomers, as these isomers have the same number and type of unique protons.

For cryptophanes 4-10, single-crystal x-ray crystallography was required to determine

the identity of the diastereomers (vide infra).

Page 67: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

44

Figure 2.4. Left) 1H NMR spectrum of (±)-anti-4. Right) H NMR spectrum of syn-5.

2.5. Crystal Structures of m-Xylyl Bridged Cryptophanes

Cryptophanes (±)-anti-4, syn-5, (±)-anti-6, syn-7, (±)-anti-8, (±)-anti-H39, and

syn-H310 were all characterized by single crystal X-ray diffraction, confirming the

stereochemical assignments as either syn or anti diastereomers. In all, 26 single crystal

structural determinations were performed and will be described in the following

sections. See Table 2.1 for a list of abbreviations used in the forthcoming chapters.

In all cases, the cryptophane hosts contain encapsulated guests, extracted from

the solvent, in the solid state. Much attention was directed (22 structures) on the anti-

diastereomer, and these structures will be discussed in detail in the ensuing sections.

The data quality of the syn-H310 sample was extremely poor. The structure was

solved, confirming the presence of the syn-H310 diastereomer, but the refinement was

Page 68: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

45

disagreeable and the details are not formally reported.

2.5.1 SQUEEZE analysis of m-xylyl Bridged CryptophaneGuest Structures

Numerous cryptophaneguest crystal structures were collected; however, the

subroutine SQUEEZE20

was often employed in the solution. SQUEEZE was used to

Table 2.1. Molecule Abbreviations.

model diffuse electron density

in highly disordered regions of

the model. SQUEEZE proved

to be a useful technique for

these systems for several

reasons:

Twenty six

cryptophaneguest crystal

structures were collected, and

the subroutine SQUEEZE was

often employed in the

refinement in order to model

diffuse electron density

associated with highly

disordered regions of the refinement model. Unless explicitly stated, only lattice

Abbreviation Molecule Name Molecule Formula

n-hexPh n-hexylbenzene C6H5(CH2)5CH3

n-prPh n-propylbenzene C6H5(CH2)2CH3

1,2,4PhMe3 1,2,4-trimethylbenzene 1,2,4-C6H3(CH3)3

o-xylene 1,2-dimethylbenzene 1,2-C6H4(CH3)2

m-xylene 1,3-dimethylbenzene 1,3-C6H4(CH3)2

PhI Iodobenzene C6H5I

NO2Ph Nitrobenzene C6H5NO2

PhBr Bromobenzene C6H5Br

PhCN Benzonitrile C6H5CN

PhCl Chlorobenzene C6H5Cl

Et2O Diethyl ether C4H10O

THF Tetrahydrofuran C4H8O

NO2Me Nitromethane CH3NO2

DMSO Dimethyl sulfoxide CH3SO

DMF N,N-Dimethylformamide (CH3)2NCHO

Page 69: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

46

solvent molecule(s) have been modeled using SQUEEZE. Generally, the encapsulated

guests are highly ordered, or can be adequately modeled which is undoubtedly a

consequence of the container effect in the solid state. The following discussion on the

geometrical parameters associated with the host-encapsulated guest interactions

observed in the solid state. Although there are certainly packing effects resulting from

the lattice solvent molecules, it is believed that this effect is minimalized in highly

disordered systems. Where possible, both SQUEEZE refinement models and the non-

SQUEEZEd models are included for direct comparison.

2.6. M-m-xylyl CryptophaneAromatic Guest Complexes

2.6.1. Common Guest Binding Motif: The Importance of CH- Interactions

Guest binding within a host balances both satisfaction of favorable interactions

as well as avoidance of unfavorable interactions. From a strictly steric repulsion

standpoint, one would expect that a bound guest molecule would adopt an orientation

that would minimize host-guest close contacts. A loosely bound guest molecule is also

entropically favored, as the guest molecule would have increased movement in the

host.21

As a result, one would expect that the bound aromatic guest molecule would

reside in the most spacious region of the cryptophane, which is along its C3 axis, where

the likelihood of guest movement is highest. However, the aromatic guest molecules

singularly adopt an orientation away from the C3 axis by approximately 10. A closer

analysis of these structures revealed that the guest molecules tilted from the C3 in order

Page 70: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

47

to maximize the CH- interactions between cryptophane host and aromatic guest

(Figure 2.5). It is also noteworthy that the guests were well ordered, with no evidence

of guest rotation within the cryptophane host. In most cases, the substituted benzenes

were partially protruding from the opening of the cryptophane, which further

prohibited guest rotation in the interior of the cryptophane.

2.6.2 CH- Interactions: Weak Intermolecular Forces

Intermolecular forces exist in a continuum, ranging from weak (van der Waals

forces, induced dipole-induced dipole interactions) to

very strong (Hydrogen bonding, ion-ion electrostatic

interactions).22

CH- interactions are relatively weak,

estimated to be worth 0.5-2.5 kcal/mol per interaction.23

Despite the small energy contribution of these

interactions, reaction selectivity, crystal packing in

organic compounds,24

and DNA and protein structure are

influenced by CH- interactions.25

It is important to define the parameters that are

associated with a CH- interaction. Nishio defined a

number of geometric parameters to define a CH-

interaction in the solid state, including the H- distance

Figure 2.5. Determination of

distance for CH-

interactions. Distance d is

taken from the guest proton to

either the closest cryptophane

aromatic carbon or the

centroid defined by the six

carbons in the appropriate

ring. Angle is defined as the

C-H- angle where C = guest

carbon, H = guest hydrogen

and = either the closest

cryptophane aromatic carbon

or the centroid defined by the

six carbons in the ring.

Page 71: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

48

Table 2.2. New m-xylyl bridged cryptophanes

Molecule L

(±)-anti-4 COOMe

syn-5 COOMe

(±)-anti-6 H

syn-7 H

(±)-anti-8 Br

(±)-anti-H39 COOH

syn-H310 COOH

(3.05Å). However, Nishio’s system is unnecessarily unwieldy; therefore, I defined a

simpler system to measure H- distances and C-H- angles, as seen in Figure 2.5. I

simply measure the distance between a guest proton to each host aromatic carbon as

well as the centroid defined by each of the six aromatic host carbons, and report the

closest contact.

In every cryptophane, multiple CH- interactions were found for each

cryptophaneguest complex (Table 2.3). In thirteen structures (Figure 2.6) with

fourteen total cryptophanes, forty CH- interactions were identified where the H-

distance was on or below 3.05Å. There are 65 hydrogen atoms on the guest molecules

Page 72: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

49

that are available for CH- interaction, meaning that 62% of available guest protons

participate in these interactions. However, only four CTB arene rings can interact with

any given guest. Taking this into account, only 55 hydrogen guest atoms could interact

with the cryptophane CTB arenes; thus 73% of possible CH- interactions are

observed.

The crystallographic data revealed the favored binding motif for aromatic guests;

however, it does not completely explain why this behavior is observed. One

explanation for this guest orientation is that the low temperature at which all structures

were collected (173K) favors enthalpic contributions to guest binding and disfavors the

thermal disorder associated with entropic binding. In effect, the low temperature

structures may simply “freeze” the guest in one conformation. Another possible

explanation for this observation may be that tilting the guest relative to the cryptophane

improves the steric relationship between the guest substituents (i.e. methyl groups,

chloro groups, etc.) and the windows of the cryptophane.

2.7. m-Xylyl Bridged Cryptophane Host Conformational Changes

M-xylyl bridged cryptophanes have rigid [1.1.1] orthocyclophane caps and m-

xylyl linkages. One would surmise that the resultant molecules would have little

flexibility; however, we observe host cryptophanes with more degrees of freedom than

Page 73: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

50

Figure 2.6. Side and Top views of cryptophanearomatic guest complexes. Guest molecules

shown in space-fill form and cryptophanes shown in stick form. a) (±)-anti-4C6H5NO2 b) (±)-

anti-4C6H5CN c) (±)-anti-4C6H5(CH2)2CH3 d) anti-4C6H5Br1 e) (±)-anti-4C6H5Cl f) 2-(±)-

anti-6m-C6H4(CH3)2 g) (±)-anti-6C6H5Br h) (±)-anti-61,2,4-C6H3(CH3)3 i) (±)-anti-6o-

C6H4(CH3)2 j) (±)-anti-4C6H5I1 k) (±)-anti-4C6H5(CH2)5CH3 l) syn-5C6H5NO2 m) (±)-anti-

6C6H5NO2. Carbon: Gray; Hydrogen: White; Nitrogen: Blue; Oxygen: Red; Chlorine: Yellow;

Bromine: Yellow. 1 Forms conglomerate structures (see vida infra).

Page 74: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

51

Table 2.3. CH- Interactions between cryptophanes and aromatic guests.

Host Guest Guest

Volume[a]

(Å3)

d-CH- [b]

(Å)

CH- [c,d]

()

Torsion Ar-O-CH2-Ar [d]

(||)

(±)-

anti-4

n-

hexPh

174[e]

2.74-c

2.81-c

3.00-c

3.06-c

153.0

177.1

158.5

155.1

91.8, 163.8

77.5, 152.4

82.7, 170.5

(±)-

anti-4

n-prPh

126 2.78-c

2.81-c

156.0

147.9

81.0, 147.8

159.4, 161.3

82.3, 155.4

(±)

-anti-6

1,2,4

PhMe3

125 2.58-c

2.74-c

176.9

163.5

84.9, 160.3

86.6, 159.4

74.0, 144.9

(±)-

anti-6

o-

xylene

110 2.71-c

2.79-c

164.3

167.1

84.2, 157.4

73.3, 147.5

83.6, 161.7

(±)-

anti-6[f]

m-

xylene

109 2.64-c

2.73-c

2.94-a

2.60-c

2.80-a

2.85-a

149.4

153.3

158.8

150.0

164.8

164.1

87.5, 170.6

175.0, 94.0

148.6, 170.7

159.4, 157.2

83.5, 176.3

107.4, 167.9

anti-4[g]

PhI

105 2.63-c

2.67-c

2.74-c

2.80-c

153.4

156.1

169.7

177.0

95.6, 145.8

88.7, 166.5

80.4, 167.2

Page 75: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

52

Table 2.3. cont. CH- Interactions between cryptophanes and aromatic guests.

Host Guest Guest

Volume[a]

(Å3)

d-CH- [b]

(Å)

CH- [c,d]

()

Torsion Ar-O-CH2-Ar [d]

(||)

(±)-anti-4

NO2Ph

102 2.60-c

2.87-c

2.98-a

3.38-c

163.2

156.1

166.4

165.0

80.6, 119.1

151.4, 171.2

149.3, 170.1

syn-5 NO2Ph

102 2.77-c

2.96-c

3.16-c

155.7

170.5

163.8

88.8, 165.0

168.9, 152.4

76.4, 163.9

(±)-anti-8 NO2Ph

102 2.68-c

2.79-c

2.85-c

2.89-c

166.7

163.3

167.3

172.6

85.9, 144.8

78.2, 166.5

77.9, 156.3

anti-4[g]

PhBr

101 2.71-c

2.73-c

2.74-c

2.85-c

152.6

156.8

172.2

176.5

147.6, 95.8

90.9, 155.4

79.0, 168.7

(±)-anti-6

PhBr

101 2.62-c

2.69-c

2.77-c

2.88-c

153.0

164.1

175.8

179.5

82.5, 172.6

85.3, 176.9

90.6, 156.3

(±)-anti-4

PhCN

97 2.69-c

2.98-c

2.99-a

3.12-c

170.4

162.1

172.2

176.0

148.9, 171.5

149.0, 171.0

84.6, 109.3

Page 76: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

53

Table 2.3. cont. CH- Interactions between cryptophanes and aromatic guests.

Host Guest Guest

Volume[a]

(Å3)

d-CH- [b]

(Å)

CH- [c,d]

()

Torsion Ar-O-CH2-Ar [d]

(||)

[a] Molecular volume determined by X-SEED. [b] Distance d as defined in Figure 2.5. Label c

denotes the H- distance measured to -ring centroid, while a denotes that the distance was

measured to an atom on the -ring. [c] Angle as defined in Figure 2.5. [d] Italicized values

indicate CH- interactions while underlined values correspond to CH- interactions that are

greater than the 3.05Å value defined as being significant by Nishio. Bolded angles correspond to

angles influenced by guest methyl groups. [e] Guest not completely encapsulated within the

cavity. [f] Two crystallographically distinct cryptophanes. [g] Forms conglomerate structures

(see vida infra).

previously hypothesized from CPK models. Most structural changes in the

cryptophane host can be attributed to the ArCTB-O-CH2-Arbridge linkage (Figure 2.7).

This region of the cryptophane molecule is the only part of the molecule with

significant degrees of rotational freedom, since the vast majority of the molecule is

aromatic and rigidly locked into place. Torsion about these bonds influences how the

m-xylyl bridges turn with respect to the C3 axis; it also affects the angle between the

two caps and their distance from one another. Torsion may also allow closer CH-

contacts between host and guest. Analysis of Ar-O-CH2-Ar torsion angles across all

cryptophanearomatic guest structures reveals some interesting general trends. First,

the distribution of this torsion angle is bimodal, with one grouping between 140 and

(±)-anti-4

PhCl

95 2.78-c

3.08-c

3.09-c

171.7

170.6

157.4

83.0, 114.9

150.4, 170.5

149.5, 171.9

Page 77: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

54

Table 2.4. X-Ray data for CryptophaneAromatic Guest materials.

Host Guest

Lattice

Solvent Sqz?1

S. G.2

a (Å)

b (Å)

c (Å)

α (°)

β (°)

γ (°)

Vol3

(Å3) R1 wR2 GOF

(±)-

anti-4

n-

hexPh

n/a No P21/n

16.503

26.981

16.729

90

99.65

90

7344 0.072 0.177 0.869

(±)-

anti-4

n-prPh

n-prPh

1.5

CH2Cl2

No P-1

13.829

15.207

22.304

74.60

73.37

81.35

4319 0.101 0.280 0.905

(±)-

anti-6

1,2,4

PhMe3

1,2,4

PhMe3

No P21/n

23.323

13.767

24.591

90

112.89

90

7274 0.060 0.180 1.078

(±)-

anti-6

o-

xylene

o-

xylene

No P21/n

23.831

13.504

24.063

90

112.81

90

7138 0.070 0.233 0.978

2_(±)-

anti-6

2_m-

xylene

2_m-

xylene

No P21/n

22.306

22.435

26.475

90

109.78

90

14134 0.061 0.162 0.752

1 SQUEEZE subroutine.

2 Space Group.

3 Unit Cell Volume.

Page 78: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

55

Table 2.4. cont. X-Ray data for CryptophaneAromatic Guest materials

Host Guest

Lattice

Solvent Sqz?1 S. G.

2

a (Å)

b (Å)

c (Å)

α (°)

β (°)

γ (°)

Vol3

(Å3) R1 wR2 GOF

anti-4 PhI PhI

Yes

(PhI)

P21

14.114

17.369

17.861

90

101.38

90

4293 0.044 0.115 1.043

(±)-

anti-4

NO2Ph NO2Ph No P21/c

24.793

12.450

25.788

90

104.13

90

7719 0.103 0.321 0.991

syn-5 NO2Ph 4NO2Ph No P-1

11.627

14.792

27.340

87.84

85.54

82.67

4648 0.051 0.108 0.826

(±)-

anti-8

NO2Ph

2.4

NO2Ph

Yes

(0.4

NO2Ph)

P-1

13.908

14.660

23.346

84.23

78.34

64.22

4191 0.049 0.140 1.005

anti-4 PhBr 2 PhBr

Yes (2

PhBr)

P21

14.133

17.466

17.844

90

101.71

90

4313 0.071 0.175 0.831

1 SQUEEZE subroutine.

2 Space Group.

3 Unit Cell Volume.

Page 79: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

56

Table 2.4. cont. X-Ray data for CryptophaneAromatic Guest materials

Host Guest

Lattice

Solvent Sqz?1 S. G.

2

a (Å)

b (Å)

c (Å)

α (°)

β (°)

γ (°)

Vol3

(Å3) R1 wR2 GOF

(±)-

anti-6

PhBr

2.5

PhBr

No C2/c

25.609

12.116

47.281

90

103.04

90

14292 0.080 0.260 1.077

(±)-

anti-4

PhCN

Et2O

0.5

PhCN

Yes

(C4H10O

0.5

PhCN)

P21/c

24.834

12.538

25.847

90

104.53

90

7790 0.065 0.175 0.787

(±)-

anti-4

PhCl

Et2O

0.5

PhCl

No P21/c

24.890

12.453

25.884

90

104.52

90

7766 0.068 0.197 1.018

1 SQUEEZE subroutine.

2 Space Group.

3 Unit Cell Volume.

180 and a second grouping while the second group represents a nearly eclipsed

conformation. Hyperchem 7.5 was employed to determine the energy difference

between these torsion angles. A simplified fragment that contained a methoxy

functionalized arene and one ester-functionalized m-xylyl group was allowed to sample

Page 80: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

57

various conformations through MM+ molecular dynamics

calculations; geometric optimization was performed

immediately after the molecular dynamics calculation.

Although this model is quite crude (based on gas phase,

discounts CTB cone conformation), the calculations

confirmed that the staggered conformation was lowest in

energy (174.5; 3.92 kcal/mol for torsion angle and 13.00

kcal/mol total energy), and that the eclipsed conformation

was slightly higher in energy (114.0; 4.33 kcal/mol for

torsion angle and 13.85 kcal/mol total energy).26

`

ArOCH2Ar torsion angles were calculated for CTB arenes involved in CH-

interactions, and compared to the torsion angles in which no CH- interactions were

Figure 2.8. Left: CCTB-O-CH2-Cbridge torsion angles where CTB arene does not engage in CH-

interactions (CTB arene-methyl close contacts also removed). Right: CCTB-O-CH2-Cbridge

torsion angles where CTB arene is engaged in CH- interactions.

Figure 2.7. Measured tor-

sion angle in cryptophanes.

Page 81: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

58

observed. The torsion angles observed for ArCTB moieties that do not participate in

and 180, with a mean of 146 and standard deviation (hereafter SD) of 32. An CH-

interactions primarily fall between 150analysis of the torsion angles observed below

120 (82.7 and 76.4) revealed that two of these angles contained CH- distances

(3.06Å, 3.16 Å) that were slightly longer than the 3.05Å cutoff described previously.

Also, two torsion angles below 120 (94.0 and 83.6) were found to have close

contacts (3.01Å and 2.96 Å) between the host arene ring centroids and guest methyl

protons, while two torsion angles below 120 (83.5 and 86.6) were found to have

longer distances between the host arene ring centroids and guest methyl protons

(3.34Å and 3.15Å). Figure 2.8 shows the scatter plot for this data, including the six

data points discussed previously, while Figure 2.9 shows the data in histogram form.

The dramatic bimodal distribution observed in CH- influenced systems may

indicate that the cryptophane host is adapting to its guest in order to maximize host-

guest CH- interactions. The rationale for describing this phenomon is that the host-

guest attraction pulls the CTB arene toward the guest. The bridging unit of the

cryptophane adjusts to allow for closer contact between guest and host; the result is a

reduction in the given torsion angles. The energetic penalty associated with the nearly

eclipsed conformation is apparently compensated by the energetic gain due to the CH-

interactions.

2.8. Halobenzene guests: Conglomorate vs. Racemate formation

Page 82: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

59

Racemic cryptophanes can crystallize from solution as either racemates or

conglomerates (wherein the enantiomers are spontaneously resolved in individual

crystals). Conglomorate formation can be a useful technique to separate enantiomers

using preferential crystallization;27

however, racemic crystals are much more common.

It is interesting to note that the ability of cryptophanes to form host-guest complexes,

with a variety of crystalline structures, generates greater opportunities for

conglomerate formation than do compounds with less propensity to form inclusion

complexes.

It was discovered that cryptophane (±)-anti-4 forms isostructural, conglomerate

crystals (±)-anti-4PhX2.5PhX (X=Br, I) when crystallized by ether precipitation

from PhX solvent. However, ether crystallization of (±)-anti-4 from PhClforms a

racemate crystal, despite the similarity of chlorobenzene, to bromobenzene or

iodobenzene. A comparison of the structural features of the conglomerate (±)-anti-

4PhX (X= Br, I) complexes with the racemate (±)-anti-4PhCl shows that there are

several important differences that may possibly explain their different packing. The

halogen atom of the guest halobenzene makes a close contact with a second host

molecule through the methanolic ester oxygen (Br(1S)-O(4B) = 3.56Å, C(1S)-Br(1S)-

O(4B) = 154.0); I(1S)-O(4B) = 3.48Å, C(1S)-I(1S)-O(4B) = 152.6). This

interaction continues such that the resulting chain of cryptophanes forms a helix

propagating along the b axis. The resultant helix must recognize the appropriate M or

P enantiomer, bringing about the formation of the conglomerate crystal (Figure 2.10).

Page 83: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

60

Interestingly, while (±)-anti-4PhBr

and PhI formed conglomerate structures,

(±)-anti-4PhCl crystallized in the

centrosymmetric space group P21/c. The

(±)-anti-4PhCl cryptophane complexes

form antiparallel 1-D sheets parallel to the

ab plane (Figure 2.9). The chlorobenzene

may be too small to interact effectively with

the next nearest cryptophane in the crystal,

preventing the interaction that likely causes

the conglomerate crystal to form. Also,

attempts to generalize this phenomenon to

similar cryptophanes were unsuccessful, as

the crystal structure of (±)-anti-6PhBr was

also centrosymmetric (C2/c). This was not

unexpected, since the guest’s interaction

with the ester functionality appeared to drive the formation of the conglomerate crystal.

2.9 Complete Encapsulation vs. Partial Encapsulation

One goal for this study was to probe the effective size of the container cavity

crystallographically by crystallizing with continually larger guest molecules. (±)-

Figure 2.9. Top: Histogram describing the

bimodal distribution of torsion angles in

which host-guest CH- interactions are

observed. Bottom: Histogram describing the

distribution of torsion angles in which host-

guest CH- interactions are not observe

Page 84: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

61

anti-4 was dissolved in a mixture of CH2Cl2 and n-propylbenzene and the CH2Cl2 was

slowly evaporated from the solution. The resulting crystal, (±)-anti-

4C6H5(CH2)2CH3 C6H5(CH2)2CH31.5CH2Cl2, revealed that the cryptophane

encapsulated the large aromatic guest, which has a guest size of 126Å3 and contains

nine heavy atoms. The aromatic portion of the guest resides inside the center of the

cryptophane pore (to satisfy CH- interactions), while the alkyl chain is found near the

cryptophane’s opening. The alkyl chain is very nearly equidistant from the two CTB

subunits (C8A-C9S = 4.07Å, O1A-C9S = 3.70Å; C26B-C9S = 3.97Å, O6B-C9S =

3.81Å). The guest’s benzene ring makes only two CH- contacts, and is not found to

be centered inside of the cryptophane. The center of the cryptophane is defined as the

centroid which described by the six inward methylene groups from both CTB subunits

(Ha), and the center of the guest arene is defined as the centroid which described by the

six arene carbon atoms. The distance between the center of the cryptophane and the

center of the arene ring is 1.26Å. The guest adopts this orientation to remain

completely encapsulated within the cryptophane. The alkyl chain of the guest, shown

in Figure 2.11 also adopted a less favored conformation to remain encapsulated within

the cryptophane. The major torsion angle describing the propyl chain were

gauche(C6S-C7S-C8S-C9S = 60.9) rather than the more energetically favored anti

conformation. Not surprisingly, the choice of a larger alkylbenzene, n-hexylbenzene,

resulted in partial guest encapsulation. In the structure (±)-anti-4C6H5(CH2)5CH3, the

Page 85: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

62

cryptophane preferentially binds the aromatic ring while the alkyl chain dangles out of

the cryptophane through its opening. The alkyl chain, no longer confined to the

cryptophane, adopts the anti conformation about the C97-C103, C103-C106, and the

C75-C106 bonds and a gauche conformation about the C98-C97 bond.

The hexylbenzene guest arene was also affected by the partial encapsulation

Figure 2.10. Top Left: (±)-anti-4PhBr looking down the –c axis. Top Right: (±)-anti-

4PhCl looking down the c-axis. Bottom: Interaction of bromobenzene guest with ester of

adjacent cryptophane. The bromine-oxygen interaction is shown with the dotted red line.

Guest molecules shown in space-fill form and cryptophanes shown in stick forms. Carbon:

Gray; Hydrogen: White; Nitrogen: Blue; Oxygen: Red; Chlorine: Yellow; Bromine: Yellow.

Page 86: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

63

event. The arene ring “recenters” itself, since the alkyl chain is no longer contained

within the cryptophane (0.50Å distance between cryptophane center and guest arene

centroid). The arene ring satisfies four CH- interactions, as a function of its centered

position. The volume of the encapsulated portion of the hexylbenzene is hard to

estimate, but falls between 110Å3 (Ph(CH2)2) and 125Å

3 (Ph(CH2)3). The upper limit

for guest size appears to be near 125 Å3 or nine non-hydrogen atoms (see also the

encapsulation of 1,2,4 C6H3Me3).

Figure 2.11. Top Left: Stick representation of (±)-anti-4 C6H5(CH2)2CH3 with spacefilled guest.

Top Right: Stick representation of C6H5(CH2)2CH3 guest. Bottom Left: Stick representation of

(±)-anti-4C6H5(CH2)5CH3 with spacefilled guest. Bottom Right: Stick representation of

C6H5(CH2)5CH3 guest. Carbon: Gray; Hydrogen: White; Oxygen: Red.

Page 87: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

64

2.10. Nonaromatic Guests

M-xylyl bridged cryptophanes also crystallize with a variety of nonaromatic

solvents, including THF, CH2Cl2, CH2BrCl, CH3Cl, DMF, DMSO, NO2Me, Et2O and

acetone. Note that the chemical

nature of these molecules is quite

varied, in terms of size, shape and

polarity. It is quite surprising that

chemically similar hosts can bind

these very different guests. We

observe many interesting features in

these crystals, including guest

disorder resultant from host-guest

size mismatch, encapsulation of two molecules, high symmetry structures, and host

conformational changes.

2.10.1. Encapsulation of two guests

Aromatic guests are comparatively large when looking at the cryptophane host

molecule; one molecule fits inside the pore with no chance to fit a second guest inside.

However, just as the hexylbenzene guest was too large to fit entirely inside of the

cryptophane container, very small organic species are too small to effectively solvate

Figure 2.12. Left: Two encapsulated NO2Me

molecules as packed within (±)-anti-42NO2Me.

Right: Two encapsulated CH2Cl2 molecules as packed

within (±)-anti-42CH2Cl2.

Page 88: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

65

the cavity. As a result, m-xylyl

bridged cryptophanes are forced

to bind two small guests rather

than the one large molecule

previously noted (Figure 2.12).

Table 2.4 delineates which

guests induce coencapsulation

with relation to the overall guest

volume.

Coencapsulated guests,

including NO2Me, CH2Cl2,

CH2BrCl DMSO, and acetone, have several common structural features. The guests

are all small, with molecular volumes ranging from 50Å3 to 68 Å

3; larger guests have

been found to only encapsulate one guest molecule. Also, these guests have a

generally pyramidal or triangular shape, which may allow for a more effective packing

and coencapsulation inside of the cryptophane. The pyramidal bases reside in the

larger equatorial region, while the point resides in the [1.1.1]orthocyclophane cap. On

the other hand, the triangular guests basically stack upon one another (Figure 2.12).

Due to the constrictive nature of the container, the guests are in close contact with one

another and are generally well ordered, as one might expect considering that two

molecules are being held in a relatively small pocket inside the cryptophane.

Table 2.5. Encapsulated guest volume in various

cryptophane host molecule materials.

Host Guest

Guest

Volume

(Å3)

Dielectric

Constant

(±)-anti-4 2 NO2Me 50 35.9

(±)-anti-6 2 NO2Me 50 35.9

(±)-anti-4 2 CH2Cl2 58 9.1

(±)-anti-H39 2 acetone 61 20.7

(±)-anti-4 2 DMSO 68 47

(±)-anti-4 THF 711

7.6

Syn-5 THF 711

7.6

(±)-anti-6 THF 711 7.6

(±)-anti-8 THF 711 7.6

(±)-anti-4 DMF 731 36.7

(±)-anti-4 CHCl3 741

4.8

Syn-7 Et2O 75 4.3

1 Guest exhibits disorder in structure.

Page 89: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

66

The coencapsulated guests

are disparate in terms of their

electronic nature (polarity,

dielectric); however, these guests

generally orient themselves inside of

the cryptophane the same way. The

guests point their C-H groups toward the polar region of the CTB cone, while the

heteroatom region of these guests point toward the equatorial region of the

cryptophane. The guests do this for at least two reasons. First, the guests align

themselves to minimize or completely cancel their electronic dipole. This behavior

lowers the overall energy of the system. Secondly, the electronically “soft” alkyl

region of the guests better match the soft CTB cone than the harder equatorial region.

2.10.2 Guest Disorder in Cryptophane Structures

Employing nonaromatic guest probes allows one to assess guest binding with

guests of a variety of sizes and shapes, rather than the roughly disc shape of substituted

benzenes. It is because of this fact that we observe coencapsulation in the solid state.

However, some guests are too large to have two guests inside of the cryptophane, and

too small to have one guest that fits well inside of the cryptophane cavity. The result is

a guest-host mismatch, which leads to a disordered solvent molecule, residing within

the host. Guest disorder at low temperature is uncommon, in part because

Figure 2.13. Left: Model of disordered,

encapsulated DMF. Right: Model of disordered,

encapsulated CHCl3.

Page 90: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

67

small guests will encapsulate a second molecule and because the cryptophane has the

ability to adapt to its guests. The crystal structure of (±)-anti-4DMF3DMF,

however, exhibited a (presumably dynamically) disordered encapsulated DMF

molecule. DMF has a molecular volume of 73Å3, which is slightly larger than the

largest coencapsulated guest, DMSO (68Å3). The final refinement for the encapsulated

DMF consisted of two half-occupancy DMF molecules that share full occupancy

nitrogen and oxygen heteroatoms (Figure 2.13). The movement of the encapsulated

DMF in the solid state (even at -100C) makes it difficult to determine the ideal

orientation of the guest; ironically, two of the three lattice-included DMF molecules

were well ordered. Similarly, (±)-anti-4CHCl3 complex in crystals of (±)-anti-

4CHCl3·CHCl3·C4H10O displays a spinning CHCl3 molecule (74Å3 volume), in four

orientations.

2.10.3. Host Conformation Changes of Cryptophanes Binding Nonaromatic

Guests

The m-xylyl bridged cryptophanes have so far been shown to adapt to the

guests that they bind. Aromatic guests have been shown to induce changes in the

cryptophanes’ ArCTB-O-CH2-Arbridge (Figures 2.8-2.9), so as to improve CH-

interactions between guest and host. Nonaromatic guests, unlike aromatic guests, do

Page 91: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

68

not have a single energetically favored interaction to influence host-guest binding.

However, cryptophanes preferentially bind one guest or two guests based on their

relative sizes. These m-xylyl bridged cryptophanes obviously bind two small guests

(NO2Me) to best solvate the interior of the cryptophane pore, and bind one slightly

larger guest (CHCl3) because binding two of these guests cannot be accommodated by

the cryptophane host.

Experimentally, we have explored the size of the cryptophane cavity

crystallographically. It is evident that the cryptophane can adapt to relatively large

guests (1,2,4 trimethylbenzene, 9 heavy atoms) and relatively small guests (CH2Cl2, 3

heavy atoms). However, we wanted to quantify how the cryptophanes adapted to

nonaromatic guests. Again, the CCTB-O-C-CAr torsion angles were measured for each

cryptophaneguest structure obtained (Table 2.6), and the host molecule

conformations were separated by coencapsulated guest structures (two guests) and

singly encapsulated guest structures.

The distributions of CCTB-O-C-CAr torsion angles for the 12 crystal structures were

once again bimodal (Figure 2.14), and the ranges of each distribution were similar to

those observed for arene guest (70-120, 140-180). This reinforces the hypothesis

that these are the two lowest energy conformations for the m-xylyl bridges of these

cryptophanes. A comparison of the torsion angle histograms for crypto-

Page 92: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

69

Table 2.6. X-Ray data for CryptophaneNon-Aromatic Guest materials.

Host Guest

Lattice

Solvent Sqz?1

S. G.2

a (Å)

b (Å)

c (Å)

α (°)

β (°)

γ (°)

Vol3

(Å3) R1 wR2 GOF

(±)-

anti-4

THF 3 THF No P21/n

20.233

18.070

23.638

90

108.47

90

8197 0.065 0.179 0.966

(±)-

anti-4

2

NO2Me

2

NO2Me

Yes P21/c

24.936

12.225

25.771

90

103.82

90

7629 0.085 0.243 0.943

(±)-

anti-4

2

CH2Cl2

0.5

CH2Cl2 +

0.6 Et2O

Yes (0.6

Et2O)

P21/c

24.569

12.432

25.777

90

102.98

90

7672 0.085 0.281 1.094

(±)-

anti-4

2

DMSO

4.6

DMSO

Yes (4.6

DMSO)

P-1

13.750

15.962

21.657

76.38

87.95

82.34

4578 0.078 0.213 0.940

(±)-

anti-4

CHCl3

CHCl3 +

Et2O

Yes

(CHCl3 +

Et2O)

P21/n

20.368

17.492

23.115

90

106.04

90

7915 0.069 0.217 1.087

1 SQUEEZE subroutine.

2 Space Group.

3 Unit Cell Volume.

Page 93: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

70

Table 2.6. cont.. X-Ray data for CryptophaneNon-Aromatic Guest materials

Host Guest

Lattice

Solvent Sqz?1 S. G.

2

a (Å)

b (Å)

c (Å)

α (°)

β (°)

γ (°)

Vol3

(Å3) R1 wR2 GOF

(±)-

anti-4

DMF

2 DMF +

Et2O

Yes

(Et2O)

P21/n

20.862

17.512

24.050

90

108.47

90

8334 0.077 0.247 0.994

syn-5 THF

THF +

Et2O

Yes

(THF +

Et2O)

P21/c

12.694

25.783

24.941

90

102.87

90

7958 0.080 0.260 1.077

(±)-

anti-6

THF THF No P-1

10.829

13.917

24.447

102.58

92.36

111.56

3306 0.084 0.231 0.819

(±)-

anti-6

2

NO2Me

4

NO2Me

Yes (2

NO2Me)

P-1

13.080

14.549

20.930

73.12

77.39

83.19

3713 0.069 0.163 0.826

syn-7 Et2O

2 m-

xylene +

0.5 Et2O

Yes(2 m-

xylene +

0.5 Et2O)

P63/m

13.786

13.786

23.058

90

90

120

3795 0.067 0.210 0.954

1 SQUEEZE subroutine.

2 Space Group.

3 Unit Cell Volume.

Page 94: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

71

Table 2.6. cont. X-Ray data for CryptophaneNon-Aromatic Guest materials

Host Guest

Lattice

Solvent Sqz?1 S. G.

2

a (Å)

b (Å)

c (Å)

α (°)

β (°)

γ (°)

Vol3

(Å3) R1 wR2 GOF

(±)-

anti-8

THF n/a No P-1

10.830

13.053

23.711

86.93

89.12

73.30

3206 0.037 0.159 1.050

(±)-

anti-

H39

2

acetone

Acetone

+ Et2O

Yes P-1

11.917

13.012

29.570

90.16

96.84

107.56

4337 0.063 0.191 1.020

1 SQUEEZE subroutine.

2 Space Group.

3 Unit Cell Volume.

guest and cryptophane2guests shows a greater relative number of eclipsed torsion

angles (70-120) vs. anti torsion angles (140-180) when only one guest was

encapsulated. Where only one nonaromatic guest was encapsulated, 43% of the

torsion angles (18 of 42) were eclipsed whereas only 27% (8 of 30) were eclipsed in

cryptophanes where two nonaromatic guests were encapsulated. As noted before, the

container molecule adjusts these torsion angles to adjust to the encapsulated guests,

and does this in a fairly uniform fashion when only one small guest is encapsulated.

This makes intuitive sense, as the container molecule is attempting to maximize its van

der Waals contact with its relatively small guest. Adding a second guest requires an

increase in the cavity size, which is reflected in the data. Only one or two of six

Page 95: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

72

torsion angles are below 120 when two guests have been encapsulated, as there is less

need to shrink the cryptophane’s cavity size.

Figure 2.14. ArCTB-O-CH2-Arbridge torsion angles.

2.10.4. Crystal Structures of Host 7: High Symmetry Structure

Crystal engineers attempt to empirically predict and design crystal structures by using

molecular symmetry and packing arguments. Though the anti and syn m-xylyl bridged

cryptophanes reported here possess high molecular symmetry in solution (D3, C3h), the

anti cryptophanes, as described in previous sections, exhibit substantially lower

symmetry (C1, pseudo-C2) in the solid state. The previous sections showed that hosts

4, 5, 6 and 8 were structurally influenced by their guests. Unlike its anti diastereomer

(±)-anti-6, syn-7 expressed its highest possible molecular symmetry (C3h, -6) in its

crystal structure. Syn-7Et2O2 m-xylene0.5 Et2O crystallized in the high symmetry

02468

10121416

Two Guests One Guest

Page 96: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

73

space group P63/m, where molecules of syn-7 reside on crystallographic -6 positions.

The solvent molecules, being incapable of holding -6 (C3h) symmetry, are obviously

disordered in these structures. The encapsulated ether molecule is disordered over

three positions as a result of its position along the three-fold crystallographic axis

Table 2.7. Torsion angles found in cryptophanesnon-aromatic guest(s) structures.

Host Guest

Torsion Ar-O-CH2-Ar [d]

(||) Host Guest

Torsion Ar-O-CH2-Ar

[d] (||)

(±)-

anti-4

2 NO2Me

149.8, 171.3

146.1, 84.5

167.5, 151.3

(±)-

anti-6

2 NO2Me

90.9, 171.2

172.6, 87.3

161.8, 150.4

(±)-

anti-4

2 CH2Cl2

152.6, 168.3

165.3, 153.6

148.7, 78.3

(±)-

anti-H39

2 C3H6O

162.7, 94.0

148.6, 172.6

76.7, 172.3

(±)-

anti-4

2 DMSO

131.9, 158.3

88.4, 178.1

177.2, 84.2

(±)-

anti-4

THF

84.7, 170.5

164.4, 99.0

173.8, 86.1

syn-5 THF

164.1, 83.2

158.2, 91.3

164.1, 83.2

(±)-

anti-6

THF

161.8, 82.3

82.1, 176.1

156.2, 87.3

(±)-

anti-8 THF

76.9, 178.2

171.8, 85.4

78.1, 169.4

(±)-

anti-4

DMF

87.0, 170.5

168.7, 82.8

162.3, 97.4

(±)-

anti-4

CHCl3

169.6, 88.2

161.0, 100.1

86.5, 165.8

syn-7 Et2O

164.3

Page 97: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

74

(Figure 2.15).

2.11. Conclusions

The two-step synthesis of

several cryptophanes was

described and several m-xylyl

bridged cryptophanes were

characterized in solution by

1H NMR and structurally by

single crystal X-Ray

diffraction. M-xylyl bridged cryptophane molecules were shown to exhibit a rich

binding chemistry in the solid state. A variety of aromatic guests were used as guests,

and a common binding motif was observed in which the guests tilted away from the

cryptophane’s idealized C3 axis. Guest-cryptophane CH- interactions were observed

in these complexes, and this interaction was used to explain guest orientation within

the host. Analysis of CCTB-O-C-CAr torsion angles in these complexes revealed that

cryptophane may adjust to increase CH- contact by decreasing these torsion angles.

This analysis was also performed on cryptophane-nonaromatic complexes, which

confirmed that cryptophanes could adjust to their guests’ based on their size as well.

Crystallization with guests of various size revealed that guests above nine

Figure 2.15. Spacefill packing of syn-7 as seen down c-axis.

Lattice solvent has been deleted to show the channels that

run between the cryptophane molecules.

Page 98: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

75

heavy atoms would be partially encapsulated (with the guest arene preferentially

bound), while small guests (four heavy atoms, or less than 70Å3) are doubly

encapsulated. Guests of intermediate size (70-80Å3), while encapsulated, are often

disordered in the solid state; this suggests a host-guest mismatch with these guests.

Cryptophane syn-7 was found to crystallize in the high symmetry P63/m space group,

which corresponds exactly to syn-7’s C3h idealized molecular symmetry.

2.12. Experimental

2.12.1 General Methods

All reactions were carried out under nitrogen atmosphere. All solvents and

reagents were used without further purification. Flash chromatography was carried out

on silica gel (32-64μm). 1H (300MHz) and

13C (90MHz) NMR were recorded on

Varian Mercury 300NMR. Uncorrected melting points were performed on a Thomas

Hoover capillary melting point apparatus. Single crystal x-ray diffraction was

performed using a Bruker-AXIS SMART diffractometer with CCD area detector

(MoK radiation) at -100 ºC. Lattice parameters were determined from least-square

analysis and the reflection data was integrated using SAINT. Structures were solved

using direct methods and refined by full matrix least-squares based on F2 using X-

SEED.28

Page 99: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

76

2.12.2. New Molecule Characterization

3,5-Bis-(4-hydroxymethyl-2-methoxy-phenoxymethyl)-benzoic acid methyl

ester (1). 3-methoxy-4-hydroxy benzyl alcohol (3.51g, 22.8 mmol) and K2CO3 (3.15

g, 22.8 mmol) were stirred in 50 mL acetone. 3,5-bis(bromomethyl)-benzoic acid

methyl ester (3.25g, 10.1 mmol) was added to the slurry. The mixture was stirred

under nitrogen at room temperature for 4 hours. H2O (500mL) was added and product

extracted with CH2Cl2 (3x40mL). The solution was dried over MgSO4 and filtered.

Solvent was removed en vacuo. The crude product was chromatographed on silica gel,

1.1 Et2O/acetone was used as eluent. The product was isolated as a white solid. The

product was also purified by sonication of crude oil in Et2O. Yield (3.49 g, 79%); m.p.

98C; Rf 0.76 (1:1 Acetone/Et2O); 1H NMR (CDCl3): δ8.00(s, br, 2H, Ar); δ7.69(s, br,

1H, Ar); δ6.89(s, 2H, Ar); δ6.74(s, 4H, Ar); δ5.13(s, 4H, ArCH2O); δ4.55(s, 4H,

CH2OH); δ3.87(s, 3H, COOCH3); δ3.84(s, 6H, OCH3). 13

C NMR (CDCl3): 52.18,

56.74, 63.32, 66.89, 115.01, 120.70, 127.06, 131.48, 132.61, 135.38, 140.61, 142.25,

159.09, 168.48. Anal. % Calcd for C26H28O8: C, 66.66; H, 6.02. Found: C, 66.42; H,

6.24.

{4-[3-(4-Hydroxymethyl-2methoxy-phenoxymethyl)-benzyloxy]-3-

methoxy-phenyl}-methanol (2). 3-methoxy-4-hydroxy benzyl alcohol (1.06g,

6.87mmol) and K2CO3 (1.08 g, 7.78 mmol) were stirred in 30mL MeOH.

,’dibromo-m-xylene (0.736g, 2.79 mmol) was added to the slurry. The mixture was

Page 100: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

77

stirred under nitrogen at room temperature for 1.5 hours. H2O (100 mL) was added

and product extracted with CHCl3 (3 x 30 mL). Solution was dried over MgSO4 and

filtered. Solvent was removed under vacuum. Crude product was chromatographed on

silica gel, with ether as the initial eluent, followed by acetone. Product isolated as a

white solid. Yield (0.729g, 77%). MP=103˚C. 1H NMR (CDCl3): δ 3.88 (s, 6H,

OCH3), 4.60 (s, 4H, ArCH2OH), 5.15 (s, 4H, ArCH2OAr), 6.79 (s, 4H, Ar), 6.93 (s,

2H, Ar), 7.36 (s, 3H, Ar), 7.49 (s, 1H, Ar). Anal. % Calcd for C24H26O6, 70.23; H,

6.38. Found: C, 69.95; H, 6.58.

3,5-Bis-(4-hydroxymethyl-2-methoxy-phenoxymethyl)-bromobenzene (3).

3-methoxy-4-hydroxy benzyl alcohol (5.7 g, 30.9 mmol), 3,5-bis(bromomethyl)-

bromobenzene (3.25g, 10.1 mmol) and K2CO3 (5.1 g, 30.9 mmol) were added to 200

mL 1:1 CH2Cl2/methanol and stirred at room temperature for 24 hours. The reaction

was monitored by TLC in 1:5 hexanes/ ethyl acetate (Rf = 0.33). The solvent was

removed in vacuo and the crude product was extracted with CH2Cl2/H2O. The organic

layer was dried with MgSO4 and the CH2Cl2 was removed in vacuo. The product was

isolated as a white solid. Yield: (4.47 g, 59%). 1H NMR (CDCl3): δ 3.86 (s, 6H,

OCH3), 4.58 (s, 4H, CH2OH), 5.07 (s, 4H, OCH2Ar), 6.76 (s, 4H, Ar), 6.91 (s, 2H, Ar),

7.38 (s, 1H, Ar), 7.50 (s, 2H, Ar); 13

C NMR (CDCl3): δ 150.38, 147.75, 140.35,

135.27,130.20, 125.09, 123.50, 119.85, 114.82, 111.54, 70.87, 65.70, 56.52. Anal. %

Calcd for C24H25O6Br: C, 58.91; H, 5.15. Found: C, 59.11; H, 5.07.

Page 101: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

78

Synthesis of cryptophane ((±)-anti-4 and syn-5). Diol (1) (4.02 g, 8.58

mmol) was dissolved in 75 mL CHCl3. The solution was added dropwise to 2 L

stirring HCOOH. The solution was stirred under nitrogen atmosphere and heated to

50˚C for 3 hours. The solvent was removed en vacuo, and distilled H2O was added to

crude solid. The solid was filtered and redissolved in CH2Cl2. The solution was dried

over MgSO4 and filtered. The solvent removed under reduced pressure. The crude

product was chromatographed on silica gel; 8:1 CH2Cl2/Et2O was used as the eluent.

The cryptophane diastereomers appear as white solid.

(±)-Anti-4. Yield (0.376 g, 10%); m.p. = 202C; Rf = 0.43; 1H NMR (300

MHz, CDCl3, 25C): δ8.07 (s, br, 6H, Ar); δ7.22 (s, br, 3H, Ar); δ6.60 (s, 6H, Ar);

δ6.40 (s, 6H, Ar); δ5.11 (d, 6H, ArCH2O, 2J (H,H) = 13.3Hz); δ4.94 (d, 6H, ArCH2O,

2J (H,H) = 13.0Hz); δ4.58 (d, 6H, ArCH2a,

2J (H,H) = 13.7Hz); δ3.95 (s, 9H, CO2CH3);

δ3.47 (s, 18H, OCH3); δ3.34 (d, 6H, ArCH2e, 2J (H,H) = 13.7Hz).

13C NMR (90MHz,

CDCl3, 25C): 36.27, 52.40, 70.25, 113.56, 114.68, 126.92, 128.85, 131.62, 132.62,

138.64, 146.45, 148.02, 166.76. Anal. % Calcd for C78H72O18 + 4C4H8O: C, 71.19; H,

6.61. Found: C, 70.85; H, 6.47.

Syn-5. Yield (0.361g, 10%); Rf = 0.35 (8:1 CH2Cl2/Et2O). m.p. = 202 C

Page 102: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

79

(dec).. 1

H NMR (300MHz, CDCl3, 25C): δ7.96(s, br, 3H, Ar); δ7.89(s, br, 6H, Ar);

δ6.63 (s, 6H, Ar); δ6.38(s, 6H, Ar); δ5.30(d, 6H, ArCH2O, 2J(H,H) = 13.3Hz); δ4.82(d,

6H, ArCH2O, 2J(H,H) = 13.4Hz); δ4.58(d, 6H, ArCH2a,

2J(H,H) =13.7Hz); δ3.95(s,

9H, CO2CH3); δ3.42(s, 18H, OCH3); δ3.36(d, 6H, ArCH2e, 2J(H,H) = 13.8Hz).

13C

NMR (90MHz, CDCl3, 25C): 34.90, 54.18, 68.77, 115.92, 116.64, 123.13, 130.37,

132.55, 141.81, 154.25, 169.46. Anal. % Calcd for C78H72O18 + 5C6H5NO2: C, 67.81;

H, 5.11; N, 3.66. Found: C, 68.04; H, 4.94; N, 3.31.

Synthesis of cryptophane ((±)-anti-6 and syn-7). Diol (2) (4.13g, 10.0mmol)

was dissolved in 50 mL CHCl3. The solution was added dropwise to 4L stirring

HCOOH. The solution was stirred under a nitrogen atmosphere for 23 hours. The

solvent was removed en vacuo, and distilled H2O was added to crude solid. The solid

was filtered and redissolved in CH2Cl2. The solution was dried over MgSO4 and

filtered. The solvent removed under reduced pressure. The crude product was

chromatographed on silica gel; 10:1 CH2Cl2/Et2O was used as eluent. The cryptophane

diastereomers appear as white solid.

(±)-Anti-6. Yield (0.311g, 8%); Rf = 0.65 (10:1 CH2Cl2/Et2O); m.p. = 204˚C

(dec.). 1H NMR (300MHz, CDCl3, 25C): δ7.84 (d, 9H, Ar,

4J (H,H) = 1Hz); δ7.01

(s, br, Ar, 3H); δ6.62 (s, 6H, Ar); δ6.42 (s, 6H, Ar); δ5.10 (d, 6H, ArCH2O, 2J (H,H) =

12.9Hz); δ4.89 (d, 6H, ArCH2O, 2J (H,H) = 13.0Hz); δ4.58 (d, 6H, ArCH2a,

2J (H,H) =

Page 103: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

80

13.7Hz); δ3.48 (s, 18H, OMe); δ3.34 (d, 6H, ArCH2e, 2J (H,H) = 13.8Hz).

13C NMR

(CDCl3): 36.21, 55.32, 64.48, 113.67, 115.03, 127.59, 128.20, 131.74, 138.07, 140.66,

141.71. Anal. % Calcd for C72H66O12 + 6 CH3NO2: C, 62.90; H, 5.68; N, 5.64.

Found: C, 62.47; H, 6.01; N, 5.27.

Syn-7. Yield (0.041g, 1%); Rf = 0.46 (10:1 CH2Cl2/Et2O); m.p. = 201C

(dec.); 1H NMR (300MHz, CDCl3, 25C): δ7.73(s, br, 3H, Ar); δ6.66 (s, 6H, Ar);

δ6.42(s, 6H, Ar); δ5.26(d, 6H, ArCH2O, 2J (H,H) = 13.0Hz); δ4.79(d, 6H, ArCH2O,

2J

(H,H) = 13.2Hz); δ4.58(d, 6H, ArCH2a, 2J (H,H) = 13.6Hz); δ3.44(s, 18H, OCH3);

δ3.35(d, 6H, ArCH2e, 2J (H,H) = 13.7Hz). Anal. % Calcd for C72H66O12 + C4H10O +

C8H10: C, 77.40; H, 6.65. Found: C, 77.66; H, 6.85.

Synthesis of cryptophane ((±)-anti-8). Diol (3) (6.10g, 12.4 mmol) was

dissolved in 400 mL CHCl3. The solution was added dropwise to 4L stirring HCOOH.

The solution was stirred under a nitrogen atmosphere for 24 hours at room temperature

and was montitored by TLC in 60:1 CH2Cl2/ ether. The formic acid was removed in

vacuo and solid product was dissolved in 100 ml 60:1 CH2Cl2/ ether to yield a dark

brown solution. 400 mL of ether was added to the solution and it turned yellow with a

white precipitate. The precipitate was filtered and dissolved in 100 mL CH2Cl2. The

solution was condensed by removing about 75 mL in vacuo. A pure precipitate formed

and was filtered.

Page 104: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

81

(±)-Anti-8. Yield (0.170 g, 3 %); Rf = 0.72 (40:1 CH2Cl2/Et2O; m.p. = 218C

(dec.). 1H NMR (CDCl3): δ 7.55 (s, 2H, Ar), δ 6.94 (s, 1H, Ar), δ 6.58 (s, 4H, Ar), δ

6.43 (s, 2H, Ar), δ 5.05 ((d, 6H, ArCH2O, 2J (H,H) = 13.1Hz), δ 4.84 (d, 6H, ArCH2O,

2J (H,H) = 13.1Hz), δ 4.58 (d, 6H, ArCH2a,

2J(H,H) = 13.7Hz), δ 3.53 (s, 6H, OCH3), δ

3.35 (d, 6H, ArCH2e, 2J (H,H) = 13.7Hz);

13C NMR (CDCl3): δ 148.21, 146.51,

140.44, 132.85, 131.73, 128.82, 123.11, 122.98, 114.86, 113.64, 70.13, 56.49, 36.51.

Anal. % Calcd for C72H63O12Br3 + C4H8O: C, 68.32 ;H, 5.36. Found: C, 68.68; H,

5.31.

(±)-Anti-H39: Compound (±)-Anti-4 (289 mg, 0.223 mmol) was dissolved in

22 mL DMF. 10 % NMe4OH in H2O (4.1 mL, 5.5 mmol) was added in one portion.

The solution was heated to 80C for 2.5 hours, and reaction was monitored by TLC

(8:1 CH2Cl2/Et2O). The solvent was removed under reduced pressure and the

remaining solid was dissolved in 30 mL of 1:1 H2O/acetone solution. 12M HCl was

added to the solution, and the resultant solid was filtered and washed with distilled

H2O. Crude (±)-Anti-H39 was recrystallized by dissolving in an acetone/methanol

solution and allowing the slow evaporation of acetone. Yield: (252 mg, 87%). M.P.

240C (dec). 1H NMR (300 MHz, DMSO[D6], 298K) 7.96 (s, 6H, Ar-H), 7.33 (s,

3H, Ar-H), 6.90 (s, 6H, Ar-H), 6.70 (s, 6H, Ar-H), 5.12 (d, 6H, Ar-CH2O,

2J(H,H) = 12.6 Hz), 5.02 (d, 6H, Ar-CH2O,

2J(H,H) = 12.6 Hz), 4.60 (d, 6H, , Ar-

Page 105: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

82

CH2a, 2J(H,H) = 12.5 Hz), 3.39 (s, 18H, OCH3).

1H NMR (300 MHz, Acetone[D6],

298K) 8.12 (s, 6H, Ar-H), 7.54 (s, 3H, Ar-H), 7.02 (s, 6H, Ar-H), 6.82 (s, 6H,

Ar-H), 5.15 (s, 12H, Ar-CH2O), 4.73 (d, 6H, Ar-CH2a, 2J(H,H) = 13.3 Hz), 3.56

(s, 18H, OCH3), 3.51 (d, 6H, Ar-CH2e 2J(H,H) = 13.3 Hz). IR (cm

-1, selected bands)

3448, 2927, 1706, 1607, 1509, 1479, 1448, 1398, 1375, 1261, 1214, 1144, 1086, 1027,

948, 883, 852, 774, 740, 618, 530. Anal. calcd. for C75H66O18 + 2 C3H6O: C, 70.94; H,

5.73. Found: C, 70.89 H, 5.93.

Syn-H310: Compound Syn-5 (158 mg, 0.121 mmol) was heated and dissolved

in 11 mL DMF. 10 % NMe4OH in H2O (5.2 mL, 5.5 mmol) was added and solution

was heated for 5 minutes at 80°C. The solution was rotovapped to dryness and the

solid was dissolved in a minimal volume of 1:1 acetone/H2O solution. The solution

was acidified with several drops of 6M HCl, and the pH of the solution was monitored.

The solution was placed in the freezer, and the crystalline solid was filtered. Yield:

(136 mg, 91%). M.P. 240C (dec). 1H NMR (300 MHz, Acetone[D6], 298K) 8.04

(s, 9H, Ar-H), 7.08 (s, 6H, Ar-H), 6.87 (s, 6H, Ar-H), 5.23 (d, 6H, Ar-CH2O,

2J(H,H) = 13.3 Hz), 5.08 (d, 6H, Ar-CH2O,

2J(H,H) = 13.3 Hz), 4.74 (d, 6H, Ar-

CH2a, 2J(H,H) = 13.5 Hz), 3.55 (s, 18H, OCH3), 3.48 (d, 6H, Ar-CH2e

2J(H,H) =

13.5 Hz).

Page 106: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

83

2.12.3. Crystal Structures

(±)-Anti-4THF3THF:

Newly purified (±)-anti-4 was dissolved in THF. After several minutes, colorless

crystals precipitated out of solution. A suitable crystal was analyzed by x-ray single

crystal diffraction. Single crystal structure of (±)-anti-4THF3THF: C94H104O22,

0.48 x 0.45 x 0.40 mm, monoclinic, space group P21/n, a = 20.233(2), b = 18.070(2), c

= 23.638(2) Å, = 108.474(2), V = 8196.6(14) Å3, Z = 4, calcd = 1.285g/cm

3, MoK

radiation, = 0.71073 Å, 2max = 50, scans, 173(2)K, 52817 total reflections, 14430

unique reflections, 7507 reflections with I>2(I) (Rint = 0.0753); absorption correction

SADABS (Tmin = 0.9577, Tmax = 0.9646, = 0.09 mm-1

), structure solution using

SHELX-S, refinement (against F2) with SHELX-97-2, 1090 parameters, 0 restraints,

H atoms placed in calculated positions and refined with a riding model, R1 = 0.0650

(I>2(I)) and wR2 = 0.1785 (all data), residual electron density max./min = 0.55/-0.35

e- Å3, GOF = 0.966.

(±)-Anti-42CH3NO2•2CH3NO2:

Crude (±)-anti-4 was dissolved in nitromethane and recrystallized by slow evaporation.

Single crystal structure of (±)-anti-42C3H6OC3H6O C4H10O: C88H84N4 O26, Mr =

1541.58, 0.30 x 0.20 x 0.10 mm, monoclinic, space group P21/c (no.14), a =

24.9356(16), b = 12.2250(8), c = 25.7708(16) Å, = 103.8210(10)°, V = 7628.5(8) Å3,

Page 107: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

84

Z = 4, calc = 1.34g/cm3, MoK radiation, = 0.71073 Å, 2max = 50°, scans, 186(2)

K, 55552 total reflections, 13427 unique reflections, 6589 reflections with I>2(I)

(Rint = 0.090); absorption correction SADABS (Tmin = 0.8627, Tmax = 1.000, = 0.10

mm-1

), structure solution with SHELX-S, refinement (against F2) using SHELX-97-2,

935 parameters, 0 restraints, H atoms placed in calculated positions on ordered

moieties and refined with a riding model, R1 = 0.0849 (I>2(I)) and wR2 = 0.2432 (all

data), residual electron density min./max. – 0.46/ 1.10 e- /Å3, GOF = 0.943.

SQUEEZE analysis of the unmodeled solvent reveals the solvent-accessible volume to

be 741 Å3 per unit cell, which is occupied by 245 electrons (calculated. 2 NO2CH3 per

asymmetric unit).

Anti-4C6H5I•C6H5I:

Crude (±)-anti-4 was dissolved in iodobenzene and recrystallized by diffusion of ether

antisolvent. C85.2H80O18.6I, Mr = 1528.39, 0.55 x 0.50 x 0.38 mm, monoclinic, space

group P21 (no. 4), a = 14.1142(11), b = 17.3693(13), c = 17.8605(13) Å, =

101.3760(10)°, V = 4292.5(6) Å3, Z = 2, calc = 1.32 g/cm

3, MoK radiation, =

0.71073 Å, 2max = 52°, scans, 173(2) K, 45808 total reflections, 16829 unique

reflections, 14746 reflections with I>2(I) (Rint = 0.033); absorption correction

SADABS (Tmin/Tmax = 0.8415, = 0.80 mm-1

), structure solution with SHELX-S,

refinement (against F2) using SHELX-97-2, 937 parameters, 1 restraints, H atoms

Page 108: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

85

placed in calculated positions on ordered moieties and refined with a riding model, R1

= 0.044 (I>2(I)) and wR2 = 0.1145 (all data), residual electron density min./max. –

0.29/ 0.46 e- /Å3, GOF = 1.043. SQUEEZE analysis of the unmodeled solvent reveals

the solvent-accessible volume to be 984 Å3 per unit cell, which is occupied by 163

electrons (calculated. C6H5I per asymmetric unit).

(±)-Anti-42CH2Cl2•0.5CH2Cl2•0.6C4H10O:

Crude (±)-anti-4 was dissolved in dichloromethane and recrystallized by diffusion of

ether antisolvent. C82.9H83O18.6Cl5, Mr = 1554.14, 0.50 x 0.46 x 0.16 mm, monoclinic,

space group P21/c (no. 14), a = 24.5688(19), b = 12.4315(9), c = 25.7768(19) Å, =

102.9790(10)°, V = 7671.8(10) Å3, Z = 2, calc = 1.35g/cm

3, MoK radiation, =

0.71073 Å, 2max = 56°, scans, 173(2) K, 90988 total reflections, 18410 unique

reflections, 12242 reflections with I>2(I) (Rint = 0.039); absorption correction

SADABS (Tmin = 0.8807, Tmax = 0.9595, = 0.26 mm-1

), structure solution with

SHELX-S, refinement (against F2) using SHELX-97-2, 991 parameters, 0 restraints,

H atoms placed in calculated positions on ordered moieties and refined with a riding

model, R1 = 0.085 (I>2(I)) and wR2 = 0.2812 (all data), residual electron density

min./max. – 0.83/ 2.25 e- /Å3, GOF = 1.094. SQUEEZE analysis of the unmodeled

solvent reveals the solvent-accessible volume to be 590 Å3 per unit cell, which is

occupied by 108 electrons (calculated. 0.6 C4H10O per asymmetric unit).

Page 109: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

86

(±)-Anti-4C6H5(CH2)5CH3 :

Crude (±)-anti-4 was dissolved in phenylhexane and recrystallized by diffusion of

ether antisolvent. C90H90O18, Mr = 1459.69, 0.12 x 0.12 x 0.08 mm, Monoclinic, space

group P21/n (no. 14), a = 16.503(10), b = 26.981(15), c = 16.729(9) Å, =

99.645(11)°, V = 7344(7) Å3, Z = 4, calc = 1.32 g/cm

3, MoK radiation, = 0.71073 Å,

2max = 47.14°, scans, 173(2) K, 45177 total reflections, 10213 unique reflections,

4081 reflections with I>2(I) (Rint = 0.243); absorption correction SADABS (ratio of

Tmin/Tmax = 0.005632, = 0.09 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 984 parameters, H atoms placed in calculated

positions on ordered moieties and refined with a riding model, R1 = 0.0721 (I>2(I))

and wR2 = 0.1769 (all data), residual electron density min./max. –0.35/0.36 e- /Å3,

GOF = 0.869.

(±)-Anti-42C3H6SO•4.6C3H6SO:

Crude (±)-anti-4 was dissolved in DMSO and recrystallized by diffusion of ether

antisolvent. C97.8H111.6O24.6S6.6, Mr = 1892.37, 0.41 x 0.33 x 0.25 mm, triclinic, space

group P-1 (no. 2), a = 13.750(2), b = 15.962(3), c = 21.657(4) Å, = 76.380(3), =

87.950(3), = 82.338(3)°, V = 4578(1) Å3, Z = 2, calc = 1.37g/cm

3, MoK radiation,

= 0.71073 Å, 2max = 50°, scans, 173(2) K, 34670 total reflections, 16019 unique

Page 110: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

87

reflections, 7603 reflections with I>2(I) (Rint = 0.067); absorption correction

SADABS (Tmin = 0.9329, Tmax = 0.9583, = 0.13 mm-1

), structure solution with

SHELX-S, refinement (against F2) using SHELX-97-2, 938 parameters, 0 restraints,

H atoms placed in calculated positions on ordered moieties and refined with a riding

model, R1 = 0.0777 (I>2(I)) and wR2 = 0.2134 (all data), residual electron density

min./max. – 0.54/ 0.82 e- /Å3, GOF = 0.940. SQUEEZE analysis of the unmodeled

solvent reveals the solvent-accessible volume to be 1205 Å3 per unit cell, which is

occupied by 447 electrons (calculated. 4.6 C3H6SO per asymmetric unit).

(±)-Anti-4CHCl3•CHCl3•C4H10O:

Crude (±)-anti-4 was dissolved in chloroform and recrystallized by diffusion of ether

antisolvent. C84H84O19Cl6, Mr = 1610.30, 0.41 x 0.33 x 0.25 mm, monoclinic, space

group P21/n (no. 14), a = 20.368(2), b = 17.492(2), c = 23.115(3) Å, = 106.044(2), V

= 7914.6(16) Å3, Z = 4, calc = 1.29g/cm

3, MoK radiation, = 0.71073 Å, 2max = 50°,

scans, 173(2) K, 57942 total reflections, 13929 unique reflections, 8414 reflections

with I>2(I) (Rint = 0.047); absorption correction SADABS (ratio of Tmin/Tmax =

0.8823, = 0.13 mm-1

), structure solution with SHELX-S, refinement (against F2)

using SHELX-97-2, 927 parameters, 0 restraints, H atoms placed in calculated

positions on ordered moieties and refined with a riding model, R1 = 0.0690 (I>2(I))

and wR2 = 0.2171 (all data), residual electron density min./max. – 1.06/ 0.86 e- /Å3,

Page 111: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

88

GOF = 1.087. SQUEEZE analysis of the unmodeled solvent reveals the solvent-

accessible volume to be 1432 Å3 per unit cell, which is occupied by 383 electrons

(calculated. 1 CHCl3 and 1 C4H10O per asymmetric unit).

(±)-Anti-4C3H7NO•2C3H7NO•C4H10O:

Crude (±)-anti-4 was dissolved in DMF and recrystallized by diffusion of ether

antisolvent. C91H103N3O22, Mr = 1590.83, 0.50 x 0.25 x 0.25 mm, monoclinic, space

group P21/n (no. 14), a = 20.862(4), b = 17.512(3), c = 24.050(4) Å, = 108.469(3)°,

V = 8334(2) Å3, Z = 4, calc = 1.27g/cm

3, MoK radiation, = 0.71073 Å, 2max = 50°,

scans, 173(2) K, 61494 total reflections, 14668 unique reflections, 7700 reflections

with I>2(I) (Rint = 0.069); absorption correction SADABS (Tmin = 0.9581, Tmax =

0.9787, = 0.09 mm-1

), structure solution with SHELX-S, refinement (against F2)

using SHELX-97-2, 980 parameters, 0 restraints, H atoms placed in calculated

positions on ordered moieties and refined with a riding model, R1 = 0.0774 (I>2(I))

and wR2 = 0.2465 (all data), residual electron density min./max. – 0.57/ 1.81 e- /Å3,

GOF = 0.994. The two lattice DMF molecules were modeled; however, the disordered

ether molecule required a SQUEEZE analysis, which revealed the solvent-accessible

volume to be 674 Å3 per unit cell, which is occupied by 181 electrons (calculated. 1

C4H10O per asymmetric unit).

Page 112: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

89

(±)-Anti-4C6H5NO2•0.3C6H5NO2•0.7C4H10O:

Crude (±)-anti-4 was dissolved in nitrobenzene and recrystallized by diffusion of ether

antisolvent. C87.7H83.2N1.3O21.3, Mr = 1496.22, 0.20 x 0.14 x 0.10 mm, monoclinic,

space group P21/c (no. 14), a = 24.793(2), b = 12.4495(11), c = 25.788(2) Å, =

104.130(2)°, V = 7718.7(12) Å3, Z = 4, calc = 1.29g/cm

3, MoK radiation, = 0.71073

Å, 2max = 50°, scans, 173(2) K, 69117 total reflections, 13574 unique reflections,

6188 reflections with I>2(I) (Rint = 0.116); absorption correction SADABS (ratio of

Tmin/Tmax = 0.8729, = 0.09 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 952 parameters, 0 restraints, H atoms placed in

calculated positions on ordered moieties and refined with a riding model, R1 = 0.1030

(I>2(I)) and wR2 = 0.3209 (all data), residual electron density min./max. – 0.56/ 0.80

e- /Å3, GOF = 0.991.

(±)-Anti-4C6H5CN•0.5C6H5CN•C4H10O:

Crude (±)-anti-4 was dissolved in benzonitrile and recrystallized by diffusion of ether

antisolvent. C92.5H89.5N1.5O19, Mr = 1526.23, 0.65 x 0.25 x 0.15 mm, monoclinic, space

group P21/c (no. 14), a = 24.834(3), b = 12.5378(14), c = 25.847(3) Å, =

104.530(2)°, V = 7790.4(15) Å3, Z = 4, calc = 1.30g/cm

3, MoK radiation, = 0.71073

Å, 2max = 50°, scans, 173(2) K, 40701 total reflections, 13700 unique reflections,

4609 reflections with I>2(I) (Rint = 0.095); absorption correction SADABS (Tmin =

Page 113: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

90

0.9435, Tmax = 0.9865, = 0.09 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 946 parameters, 0 restraints, H atoms placed in

calculated positions on ordered moieties and refined with a riding model, R1 = 0.0649

(I>2(I)) and wR2 = 0.1748 (all data), residual electron density min./max. – 0.30/ 0.24

e- /Å3, GOF = 0.787. SQUEEZE analysis of the unmodeled solvent reveals the

solvent-accessible volume to be 961 Å3 per unit cell, which is occupied by 268

electrons (calculated. 0.5 C6H5CN and 1 C4H10O per asymmetric unit).

Anti-4C6H5Br•2C6H5Br:

Crude (±)-anti-4 was dissolved in bromobenzene and recrystallized by diffusion of

ether antisolvent. C96H87O18Br3, Mr = 1768.45, 0.40 x 0.40 x 0.10 mm, monoclinic,

space group P21 (no. 4), a = 14.133(5), b = 17.466(7), c = 17.844(7) Å, =

101.710(7)°, V = 4313(3) Å3, Z = 2, calc = 1.36 g/cm

3, MoK radiation, = 0.71073 Å,

2max = 52°, scans, 173(2) K, 45808 total reflections, 16892 unique reflections,

14746 reflections with I>2(I) (Rint = 0.033); absorption correction SADABS (Tmin =

0.6889, Tmax = 0.9060, = 0.44 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 922 parameters, 1 restraints, H atoms placed in

calculated positions on ordered moieties and refined with a riding model, R1 = 0.071

(I>2(I)) and wR2 = 0.1746 (all data), residual electron density min./max. – 0.32/ 0.47

e- /Å3, GOF = 0.831. SQUEEZE analysis of the unmodeled solvent reveals the

Page 114: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

91

solvent-accessible volume to be 949 Å3 per unit cell, which is occupied by 319

electrons (calculated. 2 C6H5Br per asymmetric unit).

(±)-Anti-4C6H5(CH2)2CH3• C6H5(CH2)2CH3•1.5CH2Cl2:

Crude (±)-anti-4 was dissolved in n-propylbenzene and dichloromethane and

recrystallized by slow evaporation of dichloromethane. C97.5H99O18Cl3, Mr = 1665.21,

0.18 x 0.14 x 0.10 mm, triclinic, space group P-1 (no. 2), a = 13.829(3), b = 15.207(4),

c = 22.304(5) Å, = 74.599(4), = 73.373(5), = 81.353(4)°, V = 4319(2) Å3, Z = 2,

calc = 1.28g/cm3, MoK radiation, = 0.71073 Å, 2max = 50°, scans, 173(2) K,

22816 total reflections, 15122 unique reflections, 5369 reflections with I>2(I) (Rint =

0.086); absorption correction SADABS (ratio of Tmin/Tmax = 0.7833, = 0.18 mm-1

),

structure solution with SHELX-S, refinement (against F2) using SHELX-97-2, 1063

parameters, 0 restraints, H atoms placed in calculated positions on ordered moieties

and refined with a riding model, R1 = 0.1005 (I>2(I)) and wR2 = 0.2801 (all data),

residual electron density min./max. – 0.02/ 0.01 e- /Å3, GOF = 0.905.

(±)-Anti-4C6H5Cl•0.5C6H5Cl•C4H10O:

Crude (±)-anti-4 was dissolved in chlorobenzene and recrystallized by diffusion of

ether antisolvent. C91H89.5O19Cl1.5, Mr = 1540.38, 0.60 x 0.40 x 0.20 mm, monoclinic,

Page 115: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

92

space group P21/c (no. 14), a = 24.890(2), b = 12.4526(11), c = 25.884(2) Å, =

104.519(2)°, V = 7766.4(12) Å3, Z = 4, calc = 1.32g/cm

3, MoK radiation, = 0.71073

Å, 2max = 50°, scans, 173(2) K, 56987 total reflections, 13657 unique reflections,

7777 reflections with I>2(I) (Rint = 0.073); absorption correction SADABS (Tmin =

0.9202, Tmax = 0.9724, = 1.41 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 1000 parameters, 0 restraints, H atoms placed in

calculated positions on ordered moieties and refined with a riding model, R1 = 0.0675

(I>2(I)) and wR2 = 0.1972 (all data), residual electron density min./max. – 0.74/ 0.55

e- /Å3, GOF = 1.018.

Syn-5C6H5NO2•4 C6H5NO2:

Crude syn-5 was dissolved in nitrobenzene and recrystallized by diffusion of ether

antisolvent. C108H97N5O28, Mr = 1912.98, 0.50 x 0.40 x 0.35 mm, triclinic, space group

P-1 (no. 2), a = 11.627(2), b = 14.792(3), c = 27.340(5) Å, = 87.835(4), =

85.537(4), = 82.665(4)°, V = 4647.7(14) Å3, Z = 2, calc = 1.37g/cm

3, MoK radiation,

= 0.71073 Å, 2max = 50°, scans, 173(2) K, 25010 total reflections, 16277 unique

reflections, 7603 reflections with I>2(I) (Rint = 0.041); absorption correction

SADABS (Tmin = 0.9520, Tmax = 0.9660, = 0.10 mm-1

), structure solution with

SHELX-S, refinement (against F2) using SHELX-97-2, 1279 parameters, 0 restraints,

H atoms placed in calculated positions on ordered moieties and refined with a riding

Page 116: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

93

model, R1 = 0.0512 (I>2(I)) and wR2 = 0.1080 (all data), residual electron density

min./max. – 0.24/ 0.25 e- /Å3, GOF = 0.826.

Syn-5THF•THF•C4H10O:

Crude syn-5 was dissolved in THF and recrystallized by diffusion of ether antisolvent.

C90H98O21, Mr = 1515.68, 0.50 x 0.25 x 0.25 mm, monoclinic, space group P21/c

(no.14), a = 12.694(2), b = 25.783(5), c = 24.941(5) Å, = 102.866(4)°, V = 7958(3)

Å3, Z = 4, calc = 1.27g/cm

3, MoK radiation, = 0.71073 Å, 2max = 50°, scans,

173(2) K, 28256 total reflections, 13278 unique reflections, 4685 reflections with

I>2(I) (Rint = 0.070); absorption correction SADABS (Tmin = 0.9567, Tmax = 0.9780,

= 0.09 mm-1

), structure solution with SHELX-S, refinement (against F2) using

SHELX-97-2, 902 parameters, 0 restraints, H atoms placed in calculated positions on

ordered moieties and refined with a riding model, R1 = 0.0616 (I>2(I)) and wR2 =

0.1575 (all data), residual electron density min./max. – 0.41/ 0.55 e- /Å3, GOF = 0.754.

SQUEEZE analysis of the unmodeled solvent reveals the solvent-accessible volume to

be 1229 Å3 per unit cell, which is occupied by 313 electrons (calculated. 1 C4H8O and

1C4H10O per asymmetric unit).

(±)-Anti-6THF•THF:

Crude (±)-anti-6 was dissolved in THF and recrystallized by diffusion of ether

Page 117: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

94

antisolvent. C80H82O14, Mr = 1267.46, 0.60 x 0.10 x 0.10 mm, triclinic, space group P-

1 (no.2), a = 10.8293(17), b = 13.917(2), c = 24.447(4) Å, = 102.578(3), =

92.363(3), = 111.561(3)°, V = 3306.0(9) Å3, Z = 2, calc = 1.27g/cm

3, MoK radiation,

= 0.71073 Å, 2max = 50°, scans, 173(2) K, 23510 total reflections, 11610 unique

reflections, 3697 reflections with I>2(I) (Rint = 0.108); absorption correction

SADABS (Tmin = 0.9501, Tmax = 0.9914, = 0.081 mm-1

), structure solution with

SHELX-S, refinement (against F2) using SHELX-97-2, 814 parameters, 0 restraints,

H atoms placed in calculated positions on ordered moieties and refined with a riding

model, R1 = 0.0836 (I>2(I)) and wR2 = 0.2308 (all data), residual electron density

min./max. – 0.39/ 0.84 e- /Å3, GOF = 0.819.

(±)-Anti-6C6H5Br•2.5C6H5Br:

Crude (±)-anti-6 was dissolved in bromobenzene and recrystallized by diffusion of

ether antisolvent. C87H78.5O12Br2.5, Mr = 1515.84, 0.65 x 0.35 x 0.20 mm, monoclinic,

space group C2/c (no. 15), a = 25.609(2), b = 12.1159(11), c = 47.281(4) Å, =

103.039(2)°, V = 14292(2) Å3, Z = 8, calc = 1.41g/cm

3, MoK radiation, = 0.71073

Å, 2max = 50°, scans, 188(2) K, 57338 total reflections, 12582 unique reflections,

7774 reflections with I>2(I) (Rint = 0.048); absorption correction SADABS (Tmin =

0.4463, Tmax = 0.7562, = 1.48 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 894 parameters, 0 restraints, H atoms placed in

Page 118: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

95

calculated positions on ordered moieties and refined with a riding model, R1 = 0.0801

(I>2(I)) and wR2 = 0.2596 (all data), residual electron density min./max. – 2.20/ 0.80

e- /Å3, GOF = 1.077.

2_(±)-Anti-61,3-C6H4(CH3)2•2-1,3-C6H4(CH3)2:

Crude (±)-anti-6 was dissolved in m-xylene and recrystallized by diffusion of ether

antisolvent. C176H172O24, Mr = 2671.14, 0.60 x 0.45 x 0.20 mm, monoclinic, space

group P21/n (no. 14), a = 22.306(2), b = 25.435(3), c = 26.475(3) Å, = 109.780(2)°,

V = 14134(3) Å3, Z = 4, calc = 1.26g/cm

3, MoK radiation, = 0.71073 Å, 2max = 50°,

scans, 173(2) K, 127872 total reflections, 24895 unique reflections, 7733

reflections with I>2(I) (Rint = 0.112); absorption correction SADABS (Tmin = 0.9522,

Tmax = 0.9837, = 0.08 mm-1

), structure solution with SHELX-S, refinement (against

F2) using SHELX-97-2, 1802 parameters, 0 restraints, H atoms placed in calculated

positions on ordered moieties and refined with a riding model, R1 = 0.0605 (I>2(I))

and wR2 = 0.1620 (all data), residual electron density min./max. – 0.43/ 1.11 e- /Å3,

GOF = 0.752.

(±)-Anti-61,2,4-C6H3(CH3)3•1,2,4C6H3(CH3)3:

Crude (±)-anti-6 was dissolved in warm 1,2,4 trimethylbenzene and recrystallized by

slow cooling. C90H90O12, Mr = 1363.62, 0.60 x 0.50 x 0.25 mm, monoclinic, space

Page 119: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

96

group P21/n (no. 14), a = 23.323(2), b = 13.7672(13), c = 24.591(2) Å, =

112.890(2)°, V = 7274.3(12) Å3, Z = 4, calc = 1.25g/cm

3, MoK radiation, = 0.71073

Å, 2max = 50°, scans, 173(2) K, 52653 total reflections, 12815 unique reflections,

9060 reflections with I>2(I) (Rint = 0.040); absorption correction SADABS (Tmin =

0.9528, Tmax = 0.9799, = 0.08 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 977 parameters, 0 restraints, H atoms placed in

calculated positions on ordered moieties and refined with a riding model, R1 = 0.0597

(I>2(I)) and wR2 = 0.1796 (all data), residual electron density min./max. – 0.38/ 0.74

e- /Å3, GOF = 1.078.

(±)-Anti-61,2 C6H4(CH3)2•1,2 C6H4(CH3)2:

Crude (±)-anti-6 was dissolved in o-xylene and recrystallized by diffusion of ether

antisolvent. C88H86O12, Mr = 1335.65, 0.50 x 0.45 x 0.22 mm, monoclinic, space

group P21/n (no. 14), a = 23.8309(18), b = 13.5043(10), c = 24.0631(18) Å, =

112.8090(10)°, V = 7138.4(9) Å3, Z = 4, calc = 1.24g/cm

3, MoK radiation, =

0.71073 Å, 2max = 50°, scans, 186(2) K, 52178 total reflections, 12562 unique

reflections, 7546 reflections with I>2(I) (Rint = 0.040); absorption correction

SADABS (Tmin = 0.9604, Tmax = 0.9823, = 0.08 mm-1

), structure solution with

SHELX-S, refinement (against F2) using SHELX-97-2, 942 parameters, 0 restraints,

H atoms placed in calculated positions on ordered moieties and refined with a riding

Page 120: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

97

model, R1 = 0.0699 (I>2(I)) and wR2 = 0.2327 (all data), residual electron density

min./max. – 0.33/ 0.504 e- /Å3, GOF = 0.978.

(±)-Anti-62CH3NO2•4CH3NO2:

Crude (±)-anti-6 was dissolved in nitromethane and recrystallized by diffusion of ether

antisolvent. C78H84N6O24, Mr = 1489.51, 0.50 x 0.50 x 0.50 mm, triclinic, space group

P-1 (no.2), a = 13.080(2), b = 14.549(2), c = 20.930(3) Å, = 73.116(2), =

77.392(3), = 83.188(3)°, V = 3712.8(10) Å3, Z = 2, calc = 1.33g/cm

3, MoK radiation,

= 0.71073 Å, 2max = 50°, scans, 173(2) K, 33714 total reflections, 13063 unique

reflections, 5909 reflections with I>2(I) (Rint = 0.070); absorption correction

SADABS (ratio of Tmin/Tmax = 0.7614, = 0.10 mm-1

), structure solution with

SHELX-S, refinement (against F2) using SHELX-97-2, 934 parameters, 0 restraints,

H atoms placed in calculated positions on ordered moieties and refined with a riding

model, R1 = 0.0595 (I>2(I)) and wR2 = 0.1625 (all data), residual electron density

min./max. – 0.25/ 0.41 e- /Å3, GOF = 0.826. Two lattice NO2CH3 molecules per ASU

were modeled as disordered solvent. SQUEEZE analysis of the remaining solvent

reveals the solvent-accessible volume to be 423 Å3 per unit cell, which is occupied by

136 electrons (calculated. 2 NO2CH3 molecules per asymmetric unit).

Syn-7C4H10O•2C8H10•0.5C4H10O:

Page 121: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

98

Crude syn-7 was dissolved in m-xylene and recrystallized by diffusion of ether

antisolvent. C94H101O13.5, Mr =1446.82, 0.50 x 0.50 x 0.20 mm, hexagonal, space

group P63/m (no.), a = 13.786(2), b = 13.786(2), c = 23.058(4) Å, = 90, = 90, =

120°, V = 3795.2(11) Å3, Z = 12, calc = 1.33g/cm

3, MoK radiation, = 0.71073 Å,

2max = 54.16°, scans, 178(2) K, 12264 total reflections, 2293 unique reflections,

1300 reflections with I>2(I) (Rint = 0.062; absorption correction SADABS (Tmin =

0.9609, Tmax = 0.9841, = 0.08 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 166 parameters, 0 restraints, H atoms placed in

calculated positions on ordered moieties and refined with a riding model, R1 = 0.0671

(I>2(I)) and wR2 = 0.2100 (all data), residual electron density min./max. – 0.24/ 0.42

e- /Å3, GOF = 0.954. SQUEEZE analysis reveals the solvent-accessible volume to be

619Å3 per unit cell, which is occupied by 139 electrons (calculated. 2 C8H10 and 0.5

C4H10O molecules per asymmetric unit).

(±)-Anti-8C6H5NO2•2.4C6H5NO2:

Crude (±)-anti-8 was dissolved in nitrobenzene and recrystallized by diffusion of ether

antisolvent. C92.4H83N3.4O18.8Br3, Mr = 1781.60, 0.18 x 0.28 x 0.80 mm, triclinic, space

group P-1 (no. 2), a = 13.9075(10), b = 14.6595(10), c = 23.3456(16) Å, =

82.227(1), = 78.335(1), = 64.221(1)°, V = 4191.4(5) Å3, Z = 2, calc = 1.45g/cm

3,

MoK radiation, = 0.71073 Å, 2max = 56°, scans, 173(2) K, 38413 total

Page 122: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

99

reflections, 19271 unique reflections, 12870 reflections with I>2(I) (Rint = 0.033);

absorption correction SADABS (Tmin = 0.3757, Tmax = 0.7713, = 1.52 mm-1

),

structure solution with SHELX-S, refinement (against F2) using SHELX-97-2, 1033

parameters, 0 restraints, H atoms placed in calculated positions on ordered moieties

and refined with a riding model, R1 = 0.0488 (I>2(I)) and wR2 = 0.1401 (all data),

residual electron density min./max. – 0.99/ 1.83 e- /Å3, GOF = 1.005. SQUEEZE

analysis of the unmodeled solvent reveals the solvent-accessible volume to be 441 Å3

per unit cell, which is occupied by 55 electrons (calculated. 0.4 C6H5NO2 per

asymmetric unit).

(±)-Anti-8THF:

Crude (±)-anti-8 was dissolved in THF and recrystallized by evaporation.

C76H71N3.4O13Br3, Mr = 1432.06, 0.18 x 0.28 x 0.80 mm, triclinic, space group P-1 (no.

2), a = 10.8295(17), b = 13.053(2), c = 23.711(4) Å, = 86.932(2), = 89.123(2), =

73.302(2)°, V = 3205.9(9) Å3, Z = 2, calc = 1.484g/cm

3, MoK radiation, = 0.71073

Å, 2max = 50°, scans, 173(2) K, 23181 total reflections, 11140 unique reflections,

9310 reflections with I>2(I) (Rint = 0.033); absorption correction SADABS (Tmin =

0.6174, Tmax = 1.0000, = 1.96 mm-1

), structure solution with SHELX-S, refinement

(against F2) using SHELX-97-2, 880 parameters, 0 restraints, H atoms placed in

calculated positions on ordered moieties and refined with a riding model, R1 = 0.0367

Page 123: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

100

(I>2(I)) and wR2 = 0.1589 (all data), residual electron density min./max. – 0.34/ 0.49

e- /Å3, GOF = 1.050.

(±)-Anti-H392C3H6OC3H6OC4H10O:

Crude (±)-anti-H39 was dissolved in an acetone/methanol mixture, and recrystallized

by slow evaporation of acetone. Single crystal structure of (±)-anti-

H392C3H6OC3H6O C4H10O: C88H94O23, 0.75 x 0.30 x 0.28 mm, triclinic, space

group P-1, a = 11.9169(13), b = 13.0118(15), c = 29.570(3) Å, = 90.160(2), =

96.840(2), γ = 107.562(2), V = 4336.7(8) Å3, Z = 2, calcd = 1.164g/cm

3, MoK

radiation, = 0.71073 Å, 2max = 56, scans, 173(2)K, 39222 total reflections, 19863

unique reflections, 12374 reflections with I>2(I) (Rint = 0.0325); absorption

correction SADABS (Tmin = 0.9398, Tmax = 0.9769, = 0.084 mm-1

), structure solution

using SHELX-S, refinement (against F2) with SHELX-97-2, 920 parameters, 0

restraints, H atoms placed in calculated positions and refined with a riding model, R1 =

0.0631 (I>2(I)) and wR2 = 0.1912 (all data), residual electron density max./min =

0.376/-0.316 e- Å3, GOF = 1.020. SQUEEZE analysis of the unmodeled solvent

reveals the solvent-accessible volume to be 1232 Å3 per unit cell, which is occupied by

146 electrons (calculated. 2 C3H6O and 1 C4H10O per asymmetric unit).

Page 124: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

101

2.13 References

1. Canceill, J.; Collet, A., J. Chem Soc., Chem. Commun. 1988, 582-584.

2. (a) Gabard, J.; Collet, A., J. Chem. Soc., Chem. Commun. 1981, 1137-1139;

(b) Canceill, J.; Collet, A.; Gottarelli, G.; Palmieri, P., J. Am. Chem. Soc. 1987,

109, 6454-6464; (c) Brotin, T.; Devic, T.; Lesage, A.; Emsley, L.; Collet, A.,

Chem. Eur. J. 2001, 7, 1561-1573.

3. Holman, K. T., unpublished results.

4. (a) Brotin, T.; Dutasta, J.-P., Eur. J. Org. Chem. 2003, 2003, 973-984; (b)

Darzac, M.; Brotin, T.; Rousset-Arzel, L.; Bouchu, D.; Dutasta, J.-P., New J.

Chem. 2004, 28, 502-512; (c) Huber, G.; Brotin, T.; Dubois, L.; Desvaux, H.;

Dutasta, J.-P.; Berthault, P., J. Am. Chem. Soc. 2006, 128, 6239-6246.

5. Zhong, Z.; Ikeda, A.; Shinkai, S.; Sakamoto, S.; Yamaguchi, K., Org. Lett.

2001, 3, 1085-1087.

6. (a) Manville, J. F.; Troughton, G. E., J. Org. Chem. 1973, 38, 4278-4281; (b)

Lesot, P.; Merlet, D.; Sarfati, M.; Courtieu, J.; Zimmermann, H.; Luz, Z., J.

Am. Chem. Soc. 2002, 124, 10071-10082.

7. Canceill, J.; Lacombe, L.; Collet, A., ibid.1986, 108, 4230-4232.

8. (a) Kerckhoffs, J. M. C. A.; ten Cate, M. G. J.; Mateos-Timoneda, M. A.; van

Leeuwen, F. W. B.; Snellink-Ruël, B.; Spek, A. L.; Kooijman, H.; Crego-

Calama, M.; Reinhoudt, D. N., ibid.2005, 127, 12697-12708; (b) Sambrook,

M. R.; Beer, P. D.; Wisner, J. A.; Paul, R. L.; Cowley, A. R.; Szemes, F.; Drew,

Page 125: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

102

M. G. B., J. Am. Chem. Soc. 2005, 127, 2292-2302; (c) Braunschweig, A. B.;

Ronconi, C. M.; Han, J.-Y.; Aricó, F.; Cantrill, S. J.; Stoddart, J. F.; Khan, S. I.;

White, A. J. P.; Williams, D. J., Eur. J. Org. Chem. 2006, 2006, 1857-1866.

9. (a) Zyryanov, G. V.; Rudkevich, D. M., J. Am. Chem. Soc. 2004, 126, 4264-

4270; (b) Zhang, S.; Echegoyen, L., J. Am. Chem. Soc. 2005, 127, 2006-2011;

(c) Nielsen, K. A.; Cho, W.-S.; Lyskawa, J.; Levillain, E.; Lynch, V. M.;

Sessler, J. L.; Jeppesen, J. O., J. Am. Chem. Soc. 2006, 128, 2444-2451.

10. (a) Park, S. J.; Lee, J. W.; Sakamoto, S.; Yamaguchi, K.; Hong, J.-I., Chem.--

Eur. J. 2003, 9, 1768-1774; (b) Corbellini, F.; Mulder, A.; Sartori, A.; Ludden,

M. J. W.; Casnati, A.; Ungaro, R.; Huskens, J.; Crego-Calama, M.; Reinhoudt,

D. N., J. Am. Chem. Soc. 2004, 126, 17050-17058; (c) Sessler, J. L.; Gross, D.

E.; Cho, W.-S.; Lynch, V. M.; Schmidtchen, F. P.; Bates, G. W.; Light, M. E.;

Gale, P. A., J. Am. Chem. Soc. 2006, 128, 12281-12288.

11. (a) Tobey, S. L.; Anslyn, E. V., ibid.2003, 125, 14807-14815; (b) Paisey, S. J.;

Sadler, P. J., Chem. Commun. 2004, 306-307; (c) Correia, I.; Dornyei, A.;

Jakusch, T.; Avecilla, F.; Kiss, T.; Costa Pessoa, J., Eur. J. Inorg. Chem. 2006,

2006, 2819-2830.

12. (a) Mansikkamaki, H.; Nissinen, M.; Schalley, C. A.; Rissanen, K., New J.

Chem. 2003, 27, 88-97; (b) Liu, Y.; Guo, D.-S.; Yang, E.-C.; Zhang, H.-Y.;

Zhao, Y.-L., Eur. J. Org. Chem. 2005, 2005, 162-170; (c) Clark, T. E.; Makha,

M.; Raston, C. L.; Sobolev, A. N., Dalton Trans. 2006, 5449-5453.

Page 126: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

103

13. CCDC Structure Codes, B., BIMXUR (Mough et. al. Angew. Chem. Int. Ed.

2004, 43, 5631-5635), CUSCEY (Canceill et. al. J. Chem. Soc., Chem

Commun. 1985, 361), DIJJUB (Canceill et. al. J. Chem. Soc., Chem. Commun.

1986, 339), IYETUB, IYEVAJ, IYEVEN (Cavagnat et. al. J. Phys. Chem. B.

2004, 108, 5572), JOHGAO, JOHGES (Cram et. al. J. Am. Chem. Soc. 1991,

113, 8909), PICKUH, PICLAO (Renault et. al. Bull. Soc. Chim. Fr. 1993 130,

740), SEDPOG (Canceill et. al. Angew. Chem. Int. Ed. Engl. 1989, 28, 1246),

TUHDAB, TUHDAB01 (Garcia et. al. Bull. Soc. Chim. Fr. 1996, 133, 853),

XABDOU (Roesky et. al. Chem. Eur. J. 2003, 9, 1104.

14. CCDC Structure Codes, G. C. e. a. T. L., 41, 19465), KEHDAC, KEHDAC10

(Tanner et. al. J. Am. Chem. Soc. 1990, 112, 1659), KIRRIM (Quan et. al. J.

Chem. Soc., Chem. Commun. 1991, 660), JILZIN (Sherman et. al. J. Am.

Chem. Soc. 1991, 113, 2194), JUMZAS (Choi et. al. J. Chem. Soc., Chem.

Commun. 1992, 1733), LUXVAB, LUXVEF, LUXVIS (Warmuth et. al. J.

Org. Chem. 2003, 68, 2077), MALHAI (Gibb et. al. Chem. Commun. 2000,

363), NERYIS (Yoon et. al. Chem Commun 1997, 1303), PAQFES (Park et. al.

Chem Commun 1998, 55), PIHZAH, PIHZEL, PIHYEK, PIHYOU, PIHYIO,

PIHYUA (Robbins et. al. J. Am. Chem. Soc. 1994, 116, 111), TENLON,

TENLUT (Helgeson et. al. J. Am. Chem. Soc. 1996, 118, 5590), TUCKUX,

TUCLAE (Yoon et. al. J. Org. Chem. 1996, 61, 9323), VURBUF (Cram et. al.

J. Am. Chem. Soc. 1992, 114, 7765), YETKAJ, YETKEN (Eid Jr. et. al. J.

Page 127: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

104

Am. Chem. Soc. 1994, 116, 8506), YOCRAJ (Helgeson et. al. J. Chem. Soc.,

Chem. Commun. 1995, 307), ZAQMIN (Fraser et. al. J. Org. Chem. 1995, 60,

1207), ZELDID (Byun et. al. J. Chem. Soc., Chem. Commun. 1995, 1825),

ZUFVEB (Byun et. al. J. Chem. Soc., Chem. Commun. 1995, 1947). .

15. Helgeson, R. C.; Knobler, C. B.; Cram, D. J., J. Am. Chem. Soc. 1997, 119,

3229-3244.

16. Takahashi, S.; Miura, H.; Kasai, H.; Okada, S.; Oikawa, H.; Nakanishi, H.,

ibid.2002, 124, 10944-10945.

17. (a) Miura, H.; Yuzawa, S.; Takeda, M.; Takeda, M.; Yoichi, H.; Tomoaki, T.;

Akabori, S., Supramol. Chem. 1996, 8, 53-65; (b) Roesky, C. E. O.; Weber, E.;

Rambusch, T.; Stephan, H.; Gloe, K.; Czugler, M., Chem.– Eur. J. 2003, 9,

1104-1112.

18. (a) Metal-Catalyzed Cross-Coupling Reactions; Diederich, F., Stang, P. J. Eds.;

Wiley-VCH: New York, 1998; (b) Tsuji, J. Palladium Reagents and Catalysts,

Innovations in Organic Synthesis; Wiley: New York, 1995.

19. (a) Kurz, K.; Goebel, M. W., Helv. Chim. Acta 1996, 1967-1979; (b) Guldi, D.

M.; Swartz, A.; Luo, C.; Gómez, R.; Segura, J. L.; Martín, N., J. Am. Chem.

Soc. 2002, 124, 10875-10886.

20. Van Der Sluis, P.; Spek, A. L., Acta Crystallogr. Sect. A: Found. Crystallogr.

1990, 46, 194-201.

Page 128: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

105

21. Garel, L.; Lozach, B.; Dutasta, J. P.; Collet, A., J. Am. Chem. Soc. 1993, 115,

11652-11653.

22. (a) Giacovazzo, C., Fundamentals of crystallography. International Union of

Crystallography, Chester, 1992. (b) Braga, D.; Grepioni, F.; Desiraju, G. R.,

Chem. Rev. 1998, 98, 1375-1406.

23. Nishio, M., Tetrahedron 2005, 61, 6923-6950.

24. (a) Hancock, K. S. B., Chem. Commun. 1998, 1409-1410; (b) Nagahama, S.;

Inoue, K.; Sada, K.; Miyata, M.; Matsumoto, A., Cryst. Growth Des. 2003, 3,

247-256; (c) Nishio, M., CrystEngComm 2004, 6, 130-158; (d) Manimaran,

B.; Lai, L.-J.; Thanasekaran, P.; Wu, J.-Y.; Liao, R.-T.; Tseng, T.-W.; Liu, Y.-

H.; Lee, G.-H.; Peng, S.-M.; Lu, K.-L., Inorg. Chem. 2006, 45, 8070-8077.

25. (a) Brandl, M.; Weiss, M. S.; Jabs, A.; Suhnel, J.; Hilgenfeld, R., J. Mol. Biol.

2001, 307, 357-377; (b) Kinoshita, T.; Miyake, H.; Fujii, T.; Shoji, T.; Goto,

T., Acta Crystallogr. Sect. D: Biol. Crystallogr. 2002, 58, 622-626; (c)

Umezawa, Y.; Nishio, M., Nucleic Acids Res. 2002, 30, 2183-2192.

26. Hyperchem 7.5 was used to estimate energy for a simplified model, which

neglected the CTB groups. MM+ molecular dynamics was performed,

followed by geometry optimization. The model is admittedly crude, but

repeatedly demonstrated that the energy difference between the staggered and

eclipsed conformation was 0.5-1.0 kcal/mol, which is reasonable based on the

functional groups. .

Page 129: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

106

27. Collet, A.; Brienne, M. J.; Jacques, J., Chem. Rev. 1980, 80, 215-230.

28. Barbour, L. J., J. Supramol. Chem. 1, 189-191.

Page 130: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

107

CHAPTER 3: SYNTHESIS AND CHARACTERIZATION OF AN

“IMPLODED” CRYPTOPHANE ATROPISOMER

3.1. Introduction to CTB “cup” inversion

The CTB cups that constitute the cryptophanes are C3 symmetric and axially

chiral when A ≠ Z (Figure 3.1). In CTB molecules, the cup-shaped molecules can

invert and racemize when A ≠ Z (Figure 3.1).1 This racemization has been well

studied in this class of molecules, and kinetic data has been obtained for a variety of C3

symmetric CTB derivatives (Figure 3.1, A ≠ Z).2 It has been observed that molecular

derivation at the A and Z positions has little effect on the racemization kinetics.3 A

common mechanistic pathway has been identified for the conformation change, as seen

in Figure 3.1. The CTB first undergoes a partial inversion to a “saddle-twist”

intermediate, in which one methylene group inverts.4 The “saddle-twist” intermediate

is almost as stable as the cup form of the molecule, with a difference of approximately

12-16 kJ/mol between the two.3,4

The overall activation energy barrier to

atropisomerization is 110-115 kJ//mol,4,5

while the barrier from the saddle-twist

conformer to the cup conformer is 94-102 kJ/mol.4

The saddle-twist CTB conformation has characteristic NMR spectral features

that are dramatically different from the cup conformation. The metastable saddle-twist

conformer undergoes a low-energy rotation, which is shown in Figure 3.2. In the

saddle-twist CTB conformation, one methylene group points in the opposite direction

from the other two. However, the inverted methylene group rotates “up” while the

Page 131: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

108

next methylene group rotates “down” or inverts. This saddle-twist rotation (Figure

3.2) occurs very quickly, such that the rate of rotation at 120K is at least 106

s-1

(and

possibly much greater).4 As a result, the saddle-twist form of an A,Z substituted CTB

displays its time-averaged C3h symmetry on the NMR timescale. The unique doublet

splitting of the axial and equatorial methylene groups in cup-shaped CTB molecules is

not observed in the NMR spectra of saddle-twist CTB molecules and this behavior has

been used to monitor the kinetics and thermodynamic behavior of the saddle-twist/cup

atropisomerization.

The crown form of most CTB molecules is typically the thermodynamically

stable form, and very little or no saddle-twist CTB atropisomer is observed due to its

(±)-anti-4THF·3THF

(±)-anti-4THF

Figure 3.1. Crown inversion of CTBs. Note that the saddle-twist intermediate is close in

energy to the crown conformation (ΔΔG298 = ~12-16 kJ/mol higher for saddle-twist

intermediate).3,4

Page 132: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

109

relative thermodynamic instability. However, one can functionalize the CTB molecule

to destabilize the crown CTB atropisomer, and the result is the observation of both

crown and saddle-twist CTB atropisomers in equilibrium. Changes to the methylene

linkages (functionalization, oxidation, heteroatom substitution) or ortho substitution

with bulky constituents (Figure 3.3) have been found to produce significant saddle-

twist CTB atropisomers.3,6

Figure 3.2. Saddle-twist rotation of CTBs. Arrows show one direction of rotation,

but rotation could occur in the other direction.

Page 133: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

110

3.2. Thermal Analysis of (±)-Anti-

4THF3THF

The material (±)-anti-4THF3THF,

reported formally and completely in Chapter 2,

desolvates spontaneously under ambient

conditions, and single crystals of this material eventually become opaque after being

removed from the mother liquor. Thermogravimetric analysis (TGA) of this material

reveals two distinct but partially overlapping weight losses, which we tentatively

assigned to desolvation of lattice-included THFs followed by desolvation of the

encapsulated THF species (Figure 3.4). Maintaining (±)-anti-4THF3THF at 85C

for fifteen hours allows the mass loss to stabilize, and it corresponds approximately to

the loss of 2 equivalents of THF (calculated: 3 equivalents). The discrepancy between

the calculated and actual mass loss is likely a result of THF loss prior to TGA heating.

The 85-equilibrated material was then heated to 200C, and lost 5.2% of its original

mass (calcd. 5.3%), corresponding to 1 equivalent of THF. This series of experiments

strongly suggests the following: 1. It is possible to completely remove the lattice-

included THF molecules without removing encapsulated THF molecules. 2.

Encapsulated THF molecules are retained more strongly than the lattice-included THF,

suggesting that constrictive binding properties of the cryptophane convey to the solid

Figure 3.3. CTB oxide molecule in

saddle-twist formation.

Page 134: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

111

state, as the encapsulated THF is removed

upon heating to >130C while the lattice

THFs are slowly removed at room

temperature over several days, or over

several hours at 85C (THF boiling point =

65C).7 3. Heating to ~210C results in the

removal of all THF, leaving guest-free

cryptophane material.

3.3. Consequences of Desolvation:

Atropisomerization of (±)-anti-4

3.3.1. 1H NMR of desolvated (±)-anti-4

material

1H NMR analysis was performed on

anti-(±)-4 that had been heated to 210C to confirm that the material had completely

desolvated and to verify the integrity of the compound after TGA analysis. 1H NMR

clearly revealed changes as the spectrum displayed several new peaks in addition to the

peaks corresponding to anti-(±)-4 (Figure 3.5). A careful examination of the new

NMR spectrum revealed that the number of new peaks was approximately twice the

number of peaks in (±)-anti-4. Fortunately, the new species, hereafter (±)-imp-4, could

Figure 3.4. TGA of cryptophane materials.

Top: Thermograms of (±)-anti-4THF3THF

and (±)-anti-4THF as a function of

temperature. Note that the THF molecules in

the lattice can be separated from the

encapsulated THF molecule. Bottom:

Isothermal thermogram of (±)-anti-

4THF3THF at 85C. This heating profile

allows for removal of only lattice THF

molecules.

(±)-anti-4THF

(±)-anti-4THF3THF

Page 135: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

112

be separated from (±)-anti-4 by

preparative TLC (eluent 8:1

CH2Cl2/Et2O), allowing for a 1H

NMR analysis of the new species

without the interference of (±)-

anti-4. Spectroscopic analysis of

this new species is consistent with

a reduction in molecular

symmetry from D3 in (±)-anti-4 to

C3 in (±)-imp-4. Notably, over a

period of several days in solution,

(±)-imp-4 reverts entrirely back

into (±)-anti-4 and the original

spectrum of (±)-anti-4 results (Figure 3.4). From this, we concluded that ()-imp-4

was actually an atropisomer of ()-anti-4, and not a decomposition product. Also, the

atropisomerization was occurring in the solid-state after the cavity had been emptied at

high temperature. The new atropisomer could be separated from the starting material,

allowing for characterization of the new material. However, ()-imp-4 was unstable in

solution and converted completely to (presumably solvent-occupied) ()-anti-4 when

the material was dissolved in an appropriate solvent.

Figure 3.5. 1H NMR of atropisomerization of (±)-anti-4

over time.

Page 136: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

113

Collet described similar behavior after thermal treatment of other

cryptophanes,5b

in which cryptophanes having bridges longer than three carbon units.

Collet observed similar 1H NMR behavior, and described these cryptophanes as being

“in-out” cryptophanes, although it is not entirely clear whether he believes these

cryptophanes exist in a cup-in-cup form or in a saddle-twist-cup form (see Figure 3.6).3

3.3.2. 2D NMR of (±)-imp-4: COSY and ROESY

The reversion of (±)-imp-4 to (±)-anti-4 was monitored over time by 1H NMR.

However, there were several unanswered questions concerning the 1H NMR spectrum

of the C3 symmetric (±)-imp-4 species. First, there appeared to be missing peaks in the

CDCl3 NMR spectrum, corresponding to the methylenic protons of one of the CTB

moieties. The D3 symmetric (±)-anti-4 has four sets of doublets, therefore, the

symmetry reduced C3 symmetric (±)-imp-4 should have eight doublets. However, the

1H NMR spectrum of (±)-imp-4 contained only seven clearly defined doublets in the

CDCl3 spectrum of (±)-imp-4. The

peak integrations and peak shapes in

the 4.5-5.5 ppm region of the NMR

indicated that there was significant

peak overlap. Two dimensional

NMR techniques were therefore

employed to understand the details of

Figure 3.6. Left: Model of a “cup-in-cup”

cryptophane. Right: Model of a “saddle-twist-cup”

cryptophane.

Page 137: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

114

(±)-imp-4’s 1H NMR spectrum. Since COSY (Correlated Overhauser Spectroscopy)

NMR yields information about the coupling connectivity of a species,8 it was

performed on (±)-imp-4 to clarify the position of the overlapped peaks and to provide

information about the coupling found within the (±)-imp-4 spectrum. COSY of (±)-

imp-4 in CDCl3 revealed four coupled sets between the observed seven doublets, which

confirmed that one doublet was hidden in the 1H NMR spectrum of (±)-imp-4 in CDCl3

(Figure 3.7). The COSY spectrum revealed two couplets (9, 5.05 ppm and 10, 4.99

ppm; 8, 5.11 ppm and 11, 4.88 ppm), which are consistent with the coupling between

the diastereotopic benzyl protons of the m-xylyl cryptophane bridge. A third COSY

couplet (16-17, 3.45 ppm and 12, 4.67 ppm), is consistent with coupling between

equatorial and axial CTB methylene protons for a cup-shaped CTB unit. The fourth

COSY couplet, (16-17, 3.45 ppm and 14, 3.58 ppm) revealed the position of the eighth

doublet at 3.58 ppm. This coupled pair must be between equatorial and axial CTB

methylene protons; however, an axial CTB proton had moved significantly upfield

from approximately 4.68 ppm to 3.58 ppm. The interpreted COSY spectrum

confirmed that the 1H NMR spectrum of (±)-imp-4 (Figure 3.7) corresponds to a C3

symmetric atropisomer of (±)-anti-4.

The upfield shift of the axial CTB proton is significant, as this behavior is

observed in the NMR spectra of saddle-twist CTB molecules.4 However, upfield

chemical shifts have also been noted in cryptophane guest molecules, as shown in

Figure 2.3. The close proximity of the CTB arene rings effectively shields the guest

Page 138: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

115

Figure 3.7. COSY of (±)-imp-4 in CDCl3 at 25°C. Top: Full spectrum. Bottom:

Expanded spectrum.

Page 139: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

116

Figure 3.8. COSY of (±)-imp-4 in acetone-d6 at 25°C.

molecules. A CPK model of a cup-in-cup (±)-imp-4 shows that the axial protons of

one cup reside neatly within the second cup, which may also explain the observed

upfield chemical shift.

The proton NMR spectrum of (±)-imp-4 was also acquired in acetone-d6. In

acetone, only six doublets were clearly observed. However, the COSY spectrum of

(±)-imp-4 in acetone-d6 clearly revealed four couplets, revealing the position of the two

missing doublets in the 1H NMR spectrum (Figure 3.8). The COSY coupling between

H8 (5.21 ppm) and H10 (5.05 ppm) and H9 (5.09 ppm) and H11 (5.00 ppm) described

Page 140: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

117

the coupling between the diastereotopic protons on the m-xylyl bridge, while the

coupling between H12 (4.80 ppm) and H14 (3.60 ppm) and H14 and H16 (3.45 ppm)

describe the COSY coupling between axial and equatorial protons on the

cryptophane’s two CTB subunits. Again, one axial proton has shifted significantly

upfield (H14, 3.60 ppm). This spectrum helped to confirm the data acquired in the

CDCl3 COSY analysis. From these experiments, it was concluded that the

atropisomerization seriously affected one set of the CTB axial protons, shifting them

significantly upfield.9 The NMR experiments also confirmed the expected coupling

for the C3 symmetry of (±)-imp-4.

ROESY (Rotating Frame Overhauser Effect Spectroscopy) yields information

about the distance between protons in space,8,10

and ROESY experiments were carried

out to understand the spatial relationships of the protons in (±)-imp-4 in CDCl3. The

goals were to assign the entire NMR spectrum and to determine the structure of (±)-

imp-4 in solution. There were two plausible structures for the imploded conformer

(Figure 3.6): a fully imploded, cup-in-cup cryptophane with molecular symmetry of

C3, or a saddle-twist conformer with a molecular symmetry of C1 (but a time-averaged

C3 symmetry).

The ROESY spectrum of (±)-imp-4 taken in CDCl3 provided a wealth of

information concerning the through-space relationship of (±)-imp-4’s protons. Several

through-space interactions were found, and have been highlighted in Figure 3.9. An

Page 141: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

118

oval shows the cross-peaks between a bridge aromatic peak (1) and two diastereotopic

Figure 3.9. ROESY of (±)-imp-4 in CDCl3 at 25°C.

Page 142: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

119

bridging protons (9-10). Another oval reveals cross-peaks between the same two

diastereotopic bridging protons (9-10) and one CTB peak (4). A third oval shows a

cross-peak between the CTB peaks (4 and 5) and a fourth identifies the cross-peaks

between CTB peak (5) and the unshifted OMe peak (15). The ROESY spectrum

relates all of these peaks in space, and likely corresponds to the unchanged, cup-like

CTB moiety. Two additional cross-peaks are of interest: ovals relate aromatic peaks

(1 and 4) and aromatic peaks (2 and 6). This confirms that CTB peaks 4 and 6 are

nearer to the bridge while CTB peaks are nearer to the methoxy peaks. Unfortunately,

no concrete information concerning the overall structure of (±)-imp-4 could be

obtained.

While the time-averaged C3 symmetry characteristic of saddle-twist rotation in

CTB moieties have been observed as low as 100K, it was hoped that the energy barrier

to saddle-twist rotation would be larger in (±)-imp-4, which would increase the

temperature at which the C1 species would be observed by NMR . A low temperature

1H NMR spectrum of (±)-imp-4 was obtained at -55C in CDCl3 (Figure 3.10). While

Figure 3.10. 1H NMR of (±)-imp-4 at -55C.

Page 143: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

120

no C1 species were observed by NMR, very significant spectral broadening was

observed. Since spectral broadening can occur as a process slows to near the NMR

timescale, this low temperature NMR was evidence that a CTB saddle-twist rotation

was occurring, and that the spectrum was broadening in response to the reduced energy

in the solution.

The NMR experiments (ROESY, COSY, and 1-D 1

H NMR at varied

temperatures) in total strongly suggested that (±)-imp-4 was a time-averaged C3

symmetric conformer of (±)-anti-4. The data, particularly the collapse of the Ha and He

signals for one of the CTB moieties, suggested that the saddle-twist model of (±)-imp-

4, a molecule with an instantaneous C1 symmetry, was the more appropriate model.

The broadening of one methoxy peak and the two Ar-H associated with the

unconventional CTB suggested dynamic behavior on the 1H NMR time scale, while the

upfield shift of one methoxy peak (Δ -0.65 ppm) also indicated some shielding, which

could occur if the methoxy pointed into the cone of a CTB. However, the NMR data is

merely suggestive, but not conclusive, of a cup-saddle-twist conformation for (±)-imp-

4 as opposed to the cup-in-cup structure proposed by Collet for their cryptophanes

exhibiting similar behavior.

3.4. Kinetics of atropisomerization by 1H

NMR

As stated previously, NMR has been used successfully to monitor the

conversion of (±)-imp-4 to (±)-anti-4. A series of experiments were prepared, such

Page 144: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

121

that (±)-imp-4 was purified and dissolved in CDCl3 and the solutions were monitored

by 1H NMR over time at a number of different temperatures. Since the concentration

of the solution is directly proportional to the peak integration, the changes in

concentration were monitored by following the peak area of (±)-imp-4 over time. The

kinetics of atropisomerization at 298K are shown in Figure 3.11. The data reveals a

(pseudo)first order process with respect to (±)-imp-4. At room temperature (298K), the

rate constant for the atropisomerization of (±)-imp-4 to (±)-anti-4 was determined to be

2.3x10-5

sec-1

, which corresponds to a half-life of approximately 8.3 hours (t1/2 = 500

minutes). The rate constant was determined at multiple temperatures (298K to 323K at

5 degree intervals) and this data used to generate an Eyring plot.11

The Eyring plot,

shown in Figure 3.12, was used to calculate the activation parameters for the

conversion of (±)-imp-4 to (±)-anti-4.

Figure 3.11. Kinetics of inflation of (±)-imp-4 to (±)-anti-4 by 1H NMR integration at

298 K.

Page 145: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

122

For comparison, the

experimentally determined

activation parameters for the

crown-to-crown inversion of

CTBs6,7,12

and for the

“explosion” of the pentyl

bridged cryptophane

(Cryptophane “O”)13

are listed alongside the isomerization of (±)-imp-4 in Table 3.1.

The measured for ∆G‡ (±)-imp-4 is much lower than the activation energy barrier

corresponding to the known crown-to-crown atropisomerizations; in fact, the barrier is

much closer to the saddle-to-crown conformational change. Also note that the crown-

to-saddle transformation has a highly negative ΔS‡ value, as do both implosions

studied, while the crown-to-crown transformation has ΔS‡ value near zero. The

kinetics data and derived activation parameters support the saddle-twist model for (±)-

imp-4.

Table 3.1. Comparison of activation parameters for CTB conformers.

ΔG298‡

(kJ/mol)

ΔH‡

(kJ/mol)

ΔS‡

(J/mol K)

(±)-Imp-4 99(3) 70(3)

-98(5)

Implosion:

Cryptophane O13

97 86 -36

CTB crown-to-saddle

atropisomerization3,4

108 95

-44

CTB crown-to-crown

atropisomerization6,7,12

110-115 108-119 -13-(+13)

Figure 3.12. Eyring plot derived from isothermal kinetic data.

Page 146: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

123

3.5. Single crystal structure of (±)-imp-411CHCl3

The results of NMR analysis and kinetic studies strongly suggested that the

structure of (±)-imp-4 contains one cup-shaped CTB and one saddle-twist CTB, and

that the atropisomerization to (±)-anti-4 corresponds to a saddle-twist to cone CTB

transition. However, a crystal structure would supply concrete proof of (±)-imp-4

structure, and provide an explanation for the NMR and kinetics data. Unfortunately,

growing an X-ray diffraction quality crystal of such a molecule would prove to be very

difficult, as the half-life of (±)-imp-4 in solution at 298K was a mere eight hours. In

spite of this, attempts were begun to grow a suitable crystal, as crystals were observed

when chloroform solutions of (±)-imp-4 were quickly evaporated to dryness in vacuo.

Instead of growing crystals at room temperature, attempts were made to grow (±)-imp-

4 crystals in the freezer, as this would dramatically increase the half-life of the unstable

atropisomer.

Single crystals of (±)-imp-411CHCl3 were obtained by vapor diffusion of n-

pentane into a purified solution of (±)-imp-4 in CHCl3 at 253K, where the half-life was

estimated to be ~56 days.14

The X-ray structure indeed confirms that the structure of

(±)-imp-4 corresponds to the saddle-twist within cup conformation, as seen in Figure

3.13. The saddle-twist conformation had historically been difficult to crystallize, and

(±)-imp-4 was the first example of a saddle-twist CTB structure that has not been

Page 147: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

124

Figure 3.13. Single crystal structure of (±)-imp-4. The 11 CHCl3 molecules in the ASU have been

removed for clarity. Left: Stick structure of (±)-imp-4. Note that the saddle-twist CTB points a

methoxy group into the cup-shaped CTB. CTB arene rings are highlighted in blue. Right: (±)-

imp-4 in which the cup-shaped CTB unit is shown in CPK form, and the saddle-twist CTB unit is

shown in stick form. The saddle-twist CTB arene rings are highlighted in blue.

Figure 3.14. Assigned 1H NMR of (±)-imp-4.

Page 148: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

125

destabilized intentionally.8,15

The structure also revealed that the methoxy group of the

saddle-twist CTB subunit points directly into the cavity of the cone CTB subunit.

The structure of (±)-imp-4 helps to explain the anomalous upfield (∆δ~0.65

ppm) shielding of one OMe, which was observed in the 1H NMR spectrum. The

structure of (±)-imp-4 in the crystal structure is C1; in solution, fast saddle-twist

rotation on the NMR timescale gave a C3 averaged spectrum. However, cooling the

solution caused peak broadening to occur, as the saddle-twist rotation was slowed

down to approach the NMR timescale. With the crystal structure in hand, a complete

assignment of the 1H NMR spectrum was completed, and is shown in Figure 3.14.

3.6. Conclusions

The thermal properties of (±)-anti-4THF3THF were studied, and the material

was found to lose THF in two distinct but partially overlapping, separable transitions.

The deconvolution of these mass losses revealed that the THF molecules encapsulated

within the cryptophanes were more thermally stable than the lattice-included THF

molecules. A 1H NMR of the completely guest free material revealed that an

atropisomerization had occurred in the solid-state, yielding an imploded cryptophane

(±)-imp-4. NMR studies and kinetics analyses suggested that the structure of (±)-imp-4

consisted of one cone-shaped CTB subunit and one saddle-twisted subunit, whereas

(±)-anti-4 possess two CTB cone subunits. The saddle-twist structure of the

thermodynamically unstable (±)-imp-4 was confirmed by single-crystal X-ray

Page 149: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

126

crystallography. From the NMR (1D and 2D) data and the crystal structure, the 1H

NMR spectrum for (±)-imp-4 was completely assigned and the anomalous spectral

features were explained.

3.7. Experimental

3.7.1. General Methods

All reactions were carried out under nitrogen atmosphere. All solvents and

reagents were used without further purification. Flash chromatography was carried out

on silica gel (32-64μm). 1H (300MHz) NMR was recorded on Varian Mercury

300NMR. Thermogravimetric analyses were performed using a TA Instruments TGA

2050 under a constant stream of nitrogen gas. Single crystal x-ray diffraction was

performed using a Bruker-AXIS SMART diffractometer with CCD area detector

(MoK radiation) at -100 ºC. Lattice parameters were determined from least-square

analysis and the reflection data was integrated using SAINT. Structures were solved

using direct methods and refined by full matrix least-squares based on F2 using X-

SEED.

3.7.2. New Molecule Characterization

(±)-imp-4. (±)-Anti-4THF, earlier prepared from (±)-Anti-4THF3THF by

heating to 85C for fifteen hours, was heated in a vacuum oven at 180˚C for 16 hours,

either under positive N2 pressure or under vacuum. The material was air cooled

Page 150: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

127

immediately after removal from the oven. (±)-Imp-4 was separated from residual (±)-

anti-4 on Whatman 150Å 1000m thickness silica gel preparatory TLC plates, using

8:1 CH2Cl2/Et2O as eluent. The separation was performed in the freezer to prevent

reconversion during separation. (±)-Imp-4 was desorbed from silica gel with acetone-

d6, which was removed under reduced pressure. (±)-Imp-4 was recrystallized by vapor

diffusion of pentane into a CHCl3 solution at 253K. (±)-Imp-4 appears as a white solid.

Rf = 0.36; 1H NMR(300MHz, CDCl3, 25C): δ8.02 (s, br, 3H, Ar); δ7.94 (s, br, 3H,

Ar); δ7.25 (s, br, 3H, Ar); δ6.83 (s, 3H, Ar); δ6.61 (s, 3H, Ar); δ6.34 (s, 3H, Ar); δ6.07

(s, br, 3H, Ar); δ5.11 (d, 3H, ArCH2O, 2J (H,H) = 13.9Hz); δ5.05 (d, 3H, ArCH2O,

2J

(H,H) = 13.5Hz); δ4.99 (d, 3H, ArCH2O, 2J (H,H) = 13.5Hz); δ4.88 (d, 3H, ArCH2O,

2J (H,H) = 13.9Hz); δ4.67 (d, 3H, ArCH2,

2J (H,H) = 13.5Hz); δ3.95 (s, 9H, CO2CH3);

δ3.58 (d, 3H, ArCH2, 2J (H,H) = 15.8Hz); δ3.52 (s, 9H, OCH3); δ3.48 (d, 6H, ArCH2,

2J (H,H) = 15.7Hz)); δ2.82 (s, 9H, OCH3).

3.7.3 Kinetics Experiments

Mixtures of (±)-anti-4 and (±)-imp-4 were dissolved in CDCl3 and the

concentration of both species was monitored by 1H NMR at five degree intervals from

25 to 50C. Six peaks identified as (±)-imp-4 were monitored over time, and the

change in integrated area was followed. The instrument was locked and shimmed at

the experimental temperature, and the sample was analyzed using a time-sequence

macro. The macro programs the instrument to obtain a 1H NMR spectrum at a given

Page 151: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

128

time interval. The peak corresponding to 2.82 ppm is shown for each experiment.

Experiments were monitored for 1-3 half-lives. The corresponding rate constant

values were calculated from the slope, and used to derive the Eyring plot. The Eyring

plot in Figure 3.12 was used to derive the activation parameters for the reversion from

(±)-imp-4 to (±)-anti-4.

3.7.4 Crystal Structure

(±)-Imp411CHCl3:

Newly purified (±)-imp-4 was dissolved in CHCl3. Diffusion of n-pentane into

the CHCl3 solution was performed at -30C; large single crystals were formed. Single

crystal structure of racemic (±)-imp-411CHCl3: C89H83O36Cl33, 0.56 x 0.32 x 0.12

mm, triclinic, space group P-1, a = 14.094(5), b = 16.361(5), c = 26.247(8) Å, =

103.372(5), = 101.191(5), = 104.318(5), V = 5499(3) Å3, Z = 22, calcd = 1.577

g/cm3, MoK radiation, = 0.71073 Å, 2max = 45, scans, 183(2) K, 31876 total

reflections, 14293 unique reflections, 7834 reflections with I>2(I) (Rint = 0.0835);

absorption correction SADABS (Tmin = 0.6403, Tmax = 0.9024, = 0.874 mm-1

),

structure solution with SHELX-S, refinement (against F2) using SHELX-97-2, 1419

parameters, 0 restraints, H atoms placed in calculated positions on ordered moieties

and refined with a riding model, R1 = 0.1309 (I>2(I)) and wR2 = 0.3400 (all data),

residual electron density max./min. 0.76/ -0.93 e- /Å3, GOF = 1.060. The relatively

high values of the merging and final R factors are directly attributed to the presence of

Page 152: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

129

a large number of highly disordered chloroform molecules and crystal decomposition

of the sample on transfer to the low-temperature stream. Disordered chloroform

molecules are modeled as partial occupancy carbon and chlorine atoms. The

stoichiometry of the crystal was estimated by using the SQUEEZE subroutine of the

program PLATON,16

which estimates the solvent-accessible volume of 2738Å3 (50%

of the unit cell) to be occupied by 1239 electrons (calcd. 10.7 equivalents CHCl3).

Page 153: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

130

3.8 References

1. Lüttringhaus, A.; Peters, K. C., Angew. Chem., Int. Ed. Engl. 1966, 5, 593-594.

2. (a) Collet, A.; Gabard, J., J. Org. Chem. 1980, 45, 5400-5401; (b) Canceill, J.;

Collet, A.; Gottarelli, G., J. Am. Chem. Soc. 1984, 106, 5997-6003; (c)

Canceill, J.; Collet, A.; Gottarelli, G.; Palmieri, P., J. Am. Chem. Soc. 1987,

109, 6454-6464; (d) Malthete, J.; Collet, A., J. Am. Chem. Soc. 1987, 109,

7544-7545.

3. Collet, A., Tetrahedron 1987, 43, 5725-5759.

4. Zimmermann, H.; Tolstoy, P.; Limbach, H.-H.; Poupko, R.; Luz, Z., J. Phys.

Chem. B 2004, 108, 18772-18778.

5. (a) Collet, A.; Jacques, J., Tetrahedron Lett. 1978, 19, 1265-1268; (b) Garcia,

C.; Collet, A., Bull. Chim. Soc. Fr. 1995, 132, 52-58.

6. (a) Anand, N. K.; Cookson, R. C.; Halton, B.; Stevens, I. D. R., J. Am. Chem.

Soc. 1966, 88, 370-371; (b) Cookson, R. C.; Halton, B.; Stevens, I. D. R., J.

Chem. Soc. B. 1968, 767-774; (c) Staffilani, M.; Bonvicini, G.; Steed, J. W.;

Holman, K. T.; Atwood, J. L.; Elsegood, M. R. J., Organometallics 1998, 17,

1732-1740; (d) Lindsay, A. S., J. Chem. Soc. 1965, 1685.

7. Handbook of chemistry and physics online. 89th ed.; CRC Press: Boca Raton,

Fla., 2009.

8. Friebolin, H., Basic one- and two-dimensional NMR spectroscopy. 4th ed.;

WILEY-VCH: Weinheim ;, 2005; p 1-406.

Page 154: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

131

9. Garcia, C.; Aubry, A.; Collet, A., Bull. Chim. Soc. Fr. 1996, 133, 853-867

10. Huber, J. G.; Dubois, L.; Desvaux, H.; Dutasta, J.-P.; Brotin, T.; Berthault, P.,

J. Phys. Chem. A. 2004, 108, 9608-9615.

11. Lente, G.; Fabian, I.; Poe, A. J., New J. Chem. 2005, 29, 759-760.

12. Sato, T.; Uno, K., J. Chem. Soc., Perkin Trans. 1 1973, 895-900.

13. Lozach, B. D. T., Univerite’ Claude Bernard, Lyon, 1991. Garel, L. Doctoral

Thesis, Universite’ Claude Bernard, Lyon, 1995.

14. Mough, S. T.; Goeltz, J. C.; Holman, K. T. Angew. Chem. Int. Ed. 2004, 43,

5631-5635.

15. Guy, A.; Doussot, J.; Falguieres, A.; Prieur, B.; Baclet, B., Bull. Chim. Soc. Fr.

1996, 133, 1009.

16. Van Der Sluis, P.; Spek, A. L., Acta Crystallogr. Sect. A: Found. Crystallogr.

1990, 46, 194-201.

Page 155: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

132

CHAPTER 4: SYNTHESIS AND CHARACTERIZATION OF

CRYPTOPHANE-BASED METAL-ORGANIC POLYMER (CBMOP):

SINGLE CRYSTAL TO SINGLE CRYSTAL PARTIAL DESOLVATION

4.1. Introduction: Single Crystal to Single Crystal Processes

In many ways, the crystalline solid state has been considered a static system.

The classic model of a molecular crystal has been that its molecular components are

held rigidly in space through various intermolecular and intramolecular forces.

Furthermore, it was believed that the close-packed nature of the crystalline solid

prevents significant molecular movement within the crystal. Consequently,

significant molecular egress from a crystal (e.g. desolvation), ingress into the crystal

(e.g. sorption), or movement within the crystal commonly results in the degradation

and collapse of the single crystal into a powder. This behavior has been observed

frequently; for example, desolvation of all of the cryptophane inclusion compounds

reported in Chapter 2 result in fracture of the crystal. The breakdown of the single

crystal was often believed to be required for exogenous molecular species to react

with single crystals, as the reactions were thought to occur at or very near the

crystal‟s surface. The resulting cracks increased the accessible surface area for the

reaction and decrease the need for diffusion through the crystal bulk.

There has been an increased focus on research that looks at molecular crystals

that are capable of sustaining movement within the crystal and/or diffusion of small

exogenous molecules through the bulk of the crystal. In such instances, the diffusion

Page 156: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

133

exogenous molecules

through the bulk of the

crystal.1 In such

instances, the diffusion

does not necessarily

result in the destruction

of the single crystal, as

the changes may occur

in a single crystal to

single crystal fashion.

This phenomenon was

recently highlighted in

an issue of Australian

Journal of Chemistry.2

There are four major types of single crystal to single crystal processes: absorption,3

desolvation,4 guest exchange,

5 and polymerization

6 (or other reactions

7), as shown in

Figure 4.1. Adsorption involves the addition of small atomic or molecular guests to a

single-crystal, while desolvation involves the removal of volatile guests from the

crystal bulk.

Single-crystal to single-crystal desolvation has been known to produce dramatic

changes in the structural features of the crystal. For example, Suh and coworkers

Figure 4.1. Single-crystal to single-crystal processes.

Page 157: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

134

were able to drive off six pyridine molecules and 26 H2O molecules per asymmetric

unit in a single-crystal to single-crystal fashion, resulting in an incredible 35%

reduction in unit cell volume as well as a dramatic decrease of 5.1Å in bilayer

thickness8 (Figure 4.2)! Exposure to water-pyridine solvent vapor or immersion

within water-pyridine resulted in a complete reversion to the initial structure,

although the guest readsorption was not in a single-crystal to single-crystal fashion.

This is merely one exceptional example of many for single-crystal to single crystal

desolvation recently reported in the literature recently.5,9

Guest exchange is another single-crystal to single-crystal process, where

crystallized molecules are replaced by other molecules. To affect such an exchange,

Figure 4.2. a) [Ni2(C26H52N10)]3[BTC]46C5H5N36H2O bilayer viewed in stick form (left) and

spacefill (right). b) [Ni2(C26H52N10)]3[BTC]44H2O bilayer viewed in stick form (left) and

spacefill (right).

Page 158: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

135

a single crystal is submersed in a solution. The new guest molecules diffuse through

the existing crystal lattice, and replace the initially crystallized guests. Frequently,

the new guest molecules move through available channels. Saied and Wuest

described guest exchange in

“Deformation of Porous

Molecular Networks Induced by

the Exchange of Guests in Single

Crystals”9 (Figure 4.3). This

study examined the effect of

changing one molecular

component (carboxylic acids) in a

two-component molecular crystal

(tetrapyridone and carboxylic

acid). Submersion of a crystal in

a liquid does not result in dissolution and recrystallization, but rather in guest

exchange if the new carboxylic acid is smaller than the cocrystallized carboxylic acid.

In general, networks are capable of shrinkage and expansion; however network

shrinkage is more facile.10

While most single-crystal to single-crystal processes involve the removal,

addition or exchange of molecules from the molecular crystal, one class involves the

reactivity of the components of the crystal. Schmidt and coworkers first reported the

Figure 4.3. a) Structure of tetrapyridone crystallized

with isovaleric acid. b) Structure of tetrapyridone after

guest exchange with propionic acid. Note the decrease

in the c-axis with the smaller propionic acid.

Page 159: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

136

Figure 4.4. Single-crystal to single-crystal syntheses. Top left: Crystal structure of 1,6 triene.

Top Right: Crystal structure of photo-polymerized 1,6 triene. Bottom left: Hydrogen bond

mediated assembly of 4,4’ bipyridyl ethylene. Bottom right: [2+2] Photochemical synthesis of

[2.2]paracyclophane.

[2+2] photodimerization of cinnamic acids,11

and defined the experimental conditions

that are necessary for a successful reaction. Schmidt‟s so-called „Topochemical

Postulate‟ states that a “reaction in the solid state occurs with a minimum amount of

atomic or molecular movement.”12

Observations of photoreactions in crystallized

olefins have been shown to occur if the olefins are in parallel alignment and lie within

4.2Å of one another.13

Since Schmidt‟s initial finding, chemists have used the

topochemical postulate and principles of crystal engineering to design materials that

Page 160: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

137

Figure 4.5. One dimensional chain of CBMOP-solvent as viewed down the b-axis. Chain

propagates along the [1 0 1] direction. Hydrogen atoms, encapsulated DMF molecules, and

solvent molecules omitted for clarity. CTB arenes have been filled in.

will successfully photochemically react in a predictable fashion.14

In some cases,

these reactions may occur in a single-crystal to single-crystal fashion: two examples

include the polymerization of a 1,6 triene15

and the synthesis of a [2.2]

paracyclophane3d

(Figure 4.4).

The field of single-crystal to single-crystal transformations has only recently

gained significant attention in the literature. The increase in reported single-crystal to

single-crystal processes reflects several recent advances. The advent of the area

(CCD) detector in X-ray crystallography has increased the number of crystals

analyzed in a given period of time, allowing for more detailed analysis of molecular

crystal dynamics.16

Also, the increase in reported single-crystal to single-crystal phe-

Page 161: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

138

nomena has encouraged further research in the field, as the previous paradigm of the

“chemical cemetery”17

continues to give way to a more dynamic view of crystalline

solids. While it remains impossible to predict if a crystal will undergo a single-

crystal to single-crystal transformation, it is evident that dynamic behavior in the

crystalline solid will continue to be studied.

4.2. A 1-D Cryptophane-Derived Coordination Polymer

The solvothermal reaction of ligand anti-(±)-H39 with Cu(NO3)2 2.5 H2O and

pyridine in 2:1 DMF/MeOH yielded dark blue prismatic crystals. The low

temperature X-ray single crystal structure of these crystals (immediately transferred

to the low temperature stream after removal from the mother liquor) revealed a one-

dimensional polymeric material comprised of anti-(±)-93-

ligands that are linked by

coordination to copper(II) ions; the single crystals have a composition of [Cu1.5((±)-

9DMF)(C6H5N)3(MeOH)]∙xDMF∙yMe OH (ca. x = 1, y = 2, vide infra) hereafter

CBMOP∙solvent (Figure 4.5). This structure‟s asymmetric unit consists of one

cryptophane and 1.5 crystallographically distinct copper ions, and the coordination

about the copper ions is given in Figure 4.6. The coordination about Cu(1) has a

Jahn-Teller distorted octahedral coordination consisted of two crystallographically

equivalent trans carboxylates (Cu1-O3 = 1.925(4)Å; Cu1-O2 = 1.960(4), two trans

pyridine ligands (Cu1-N13S = 2.040(6)Å; Cu1-N17S = 2.010(6)Å) and two weakly

coordinating, crystallographically equivalent carbonyl oxygen atoms (Cu1-O1,4 =

Page 162: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

139

Figure 4.6. Coordination environment about metal centers.

2.519 Å). Two Cu(1) atoms link cryptophane molecules; one cryptophane molecule

is found within the asymmetric unit and one at (1/2-x, 5/2-y, -z).

Cu(2) has two trans carboxylates (Cu2-O6,6‟ = 1.963(3)Å), two pyridines (Cu2-

N13S = 2.040(6)Å; Cu2-N17S = 2.010(6)Å), and two axially coordinated MeOH

molecules (Cu2-O7, Cu2-O7‟; = 2.518(5)Å). The two carboxylate ligands are not

coplanar; the dihedral angle of O(5)-O(6)-O(6‟)-C(5‟) is 142. Cu(2) falls on a

Wykoff crystallographic position (0, y. ¼), which is a glide plane, and the two

cryptophane molecules that link through Cu(2) are in the related symmetrically

through the glide plane. One cryptophane linked through Cu(2) has atoms found in

the asymmetric unit and one at (-x, y, ½-z).

The one-dimensional polymer can be described as cryptophane dimers that are

linked to one another through the third m-xylyl bridge, similar to (±)-anti-

H39·(2C3H6O). While (±)-anti-H39 links cryptophanes to form a linear polymer,

CBMOP forms a one-dimensional bent polymer (Figure 4.5). Cu(1) linked two

cryptophanes to form a dimeric structure, which is analogous to the dimeric structure

Page 163: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

140

of anti-H39. Cu(2) linked the dimers to one another, such that the polymer

propagated along the [101] direction. The polymer‟s “kink” or bend can be attributed

to the m-xylyl bridge linked to Cu(2), which is rotated approximately 60 from that of

(±)-anti-H39 linearity (Figure 4.7). The stereochemistry of the cryptophanes along

the polymer was [{(+)-9-(-)-9}-{(-)-9-(+)-9}], where (+) and (-) indicate the helicity

of the cryptophanes within the chain .

Determining the composition

of CBMOPsolvent was complicated

by significant disorder of lattice-

solvent. This is an unfortunate

consequence of using bulky, elliptical

ligands since they pack inefficiently

in the solid-state. Though the

polymeric region of the structure was

generally well ordered, the lattice

solvent was not well-ordered. One DMF and two MeOH were assigned to this

electron density. A SQUEEZE18

analysis of CBMOP-solvent was performed, and

determined the solvent accessible volume of 3880 Å3

per asymmetric unit (485Å3/unit

cell, 20.2% of the unit cell volume), with 573 unmodeled e-s per unit cell (72 per

ASU). The formulation defined earlier for the lattice solvent (1DMF, 2MeOH) has a

total of 76e-s per ASU.

Figure 4.7. Overlay of cryptophane in anti-(±)-H39

and CBMOPsolvent viewed from top of

cryptophane. Note the difference in the

cryptophane bridges on the left. While (±)-anti-

H39 continues in a nearly straight path, CBMOP-

solvent is rotated approximately 60°.

Page 164: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

141

Thermogravimetric analysis of CBMOP-solvent was employed to confirm

the stoichiometry of the material. Subjecting CBMOP-solvent to a 10C/min

heating ramp, and monitoring mass as a function of temperature revealed a

continuous mass loss upon heating. One mass loss occurs between 50 and 100C, a

second apparent mass loss between 130C and 220C and a third, more steep mass

loss is found above 250C (Figure 4.8 Bottom). The highest temperature mass loss

corresponds with the decomposition of the material; however, the two lower

temperature mass losses likely correspond to the desolvation of the CBMOP-solvent

material. The lowest temperature mass loss most likely corresponds to the removal of

included lattice solvent molecules, especially considering that those solvent

molecules are highly disordered even at low temperatures. Furthermore, the forces

that are containing these molecules within the lattice are weak (van der Waals) and a

path exists to allow their release. The “second” mass loss likely corresponds to the

removal of coordinated solvent molecules, and possibly the encapsulated DMF

solvent molecule, which are held by stronger intermolecular forces (coordination

bonds, constrictive binding). An isothermal TGA experiment (Figure 4.8 Top) was

performed to quantify and possibly separate the two early mass losses so as to

understand the desolvation process. A very slow heating ramp followed by an

isotherm (0.2C/min up to 160C, isothermed for ~800 minutes) revealed three

discernable but still convoluted mass losses: a low temperature mass loss occurring

very near 30C, a transition occurring between 40C and 80C, and a higher

Page 165: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

142

temperature transition occurring between 100C and 150C. The first transition

Figure 4.8. Thermogravimetric analysis of CBMOP-solvent. Top: CBMOP mass as a function

of heating at 0.2°C/min from room temperature to 160°C followed by an isothermal hold at

160°C for ~ 800 minutes. Bottom: CBMOP mass as a function of heating at 10°C/min from

room temperature to 380°C.

Page 166: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

143

likely corresponds to surface-bound solvent molecules, as the solvent was removed

immediately, while the second and third transitions more likely correspond to lattice

desolvation and metal coordinated solvent loss of CBMOP. The temperatures are

lower in the second experiment; however, the heating ramp is much slower as well

(0.2C/min vs. 10C/min). This phenomenon is commonly observed in

thermogravimetric (and calorimetric) studies at various rates of heating.

4 mg). The second and third mass losses were calculated to be 11.7% and

12.0% respectively; the theoretical mass losses were calculated to be 7.5% for the

loss of lattice solvent and 14.7% for the loss of coordinated solvent (18.7% for loss

of coordinated solvent and encapsulated solvent). The experimental mass loss total

of 23.7% is intermediate between the theoretical mass loss total of 22.2% for loss of

coordinated and lattice solvent and 26.2% for loss of coordinated, lattice and

encapsulated solvent. It is quite possible that some, but not all encapsulated DMF

molecules have been driven from the cryptophanes‟ pores.

4.3. Unit Cell Changes in CBMOP

It was observed that single crystals of this material retained their single crystal

features long after their removal from the mother liquor. Several crystals (5+

crystals) of CBMOP-solvent were removed from the mother liquor, immediately

mounted, and placed in the low temperature nitrogen stream (173±2K) of the X-ray

single crystal diffractometer, and an orientation matrix and unit cell was obtained for

Page 167: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

144

Table 4.1. Unit Cell data for CBMOP∙solvent and CBMOP∙desolvated at RT and 173K.

CBMOP∙solvent @ 173K

[a]

CBMOP∙desolvated @ 173K

[b]

CBMOP∙solvent @ RT

[c]

CBMOP∙desolvated @ RT

[c]

a (Å) 53.0(1) 53.3(2) 53.28(7) 53.40(3)

b (Å) 16.09(2) 15.89(4) 16.26(2) 16.15(1)

c (Å) 23.13(3) 22.43(8) 23.40(4) 22.70(1)

Volume

(Å3)

19249(26) 18534(87) 19784(57) 19162(24)

[a] Average of seven crystals. [b] Average of six crystals. [c] Result from one crystal.

each crystal. The averaged initial unit cell parameters from these experiments are

given in Table 4.1. Each crystal was removed from the cold temperature stream, and

exposed to ambient conditions for a period of at least two hours. The single crystals

were then remounted on the diffractometer, and a second orientation matrix and unit

cell was obtained for each crystal at 173K. Though, the crystals‟ quality had

diminished somewhat after being exposed to ambient conditions, each crystal yielded

a similar, but statistically smaller unit cell, as shown in Table 4.1 (Monoclinic C,

19249(26)Å3 vs. Monoclinic C, 18534(87)Å

3). The total unit cell volume change

measured 715Å3 per unit cell, which corresponds to a volume decrease of 3.7%; this

change is statistically significant. This behavior implied that a single-crystal to

single-crystal desolvation process was occurring in this system.

The unit cell volume shrinkage was also monitored in real time by tracking

the cell volume of two crystals as a function of time that the crystals were exposed to

ambient conditions (Figure 4.9). The volume change occurred rapidly at room

Page 168: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

145

temperature, but the solvated crystals were stable indefinitely at 173K. It is

interesting to note that the rate of volume change occurs rapidly (completed in ca. 1

hour); however, the rate is deceleratory in nature, which is commonly observed in

solid-state desolvation.19

Figure 4.9. Unit cell change of CBMOP as a function of time at 25C. Data points consist of

individual unit cell matrices collected over time.

4.3.1. X-Ray Single Crystal Structure of CBMOP-desolvated

Despite the decrease in quality of CBMOP crystals upon desolvation, a

dataset of desolvated CBMOP, hereafter CBMOP-desolvated, was collected and the

structure solved. Not surprisingly, the single crystal structure of CBMOP-

desolvated was very similar to that of CBMOP-solvent; however, several notable

differences were observed beyond the unit cell volume change. There were

significant changes about the metal centers: Cu(1) lost approximately 0.4 pyridine

18400

18500

18600

18700

18800

18900

19000

19100

19200

19300

0 20 40 60 80 100 120 140

time (min)

Vo

lum

e (

Å3)

Page 169: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

146

molecules and 0.5 MeOH molecules while Cu(2) appeared to lose 0.4 pyridines.

Furthermore, the carboxylate-copper O-Cu-O angles changed from 169.1(2)˚ to

160.9(4)˚ about Cu1 and 178.8(1)˚ to 174.0(5)˚ about Cu2 respectively. The

reduction of the angles had the net effect of contracting the b- and c-axes. The unit

cell changes in the single crystal of CBMOP-desolvated were similar to those

observed in the unit cells of CBMOP-solvent that had been exposed to ambient

conditions for 2+ hours (Table 4.1). Furthermore, the number of unit cell electrons

that were not modeled in the single crystal structure, as determined by SQUEEZE,

was reduced from 573 in CBMOP-solvent to 397 in CBMOP-desolvated, which

corresponds to a loss of approximately 1.5 MeOH per asymmetric unit.

4.3.2. Explanation of Single Crystal Desolvation

As observed in Figure 4.10, the unit cell dimensions in the b- and c- axes

statistically decrease upon partial desolvation of the crystals. Since the crystal shrunk

as a result of single-crystal to single-crystal desolvation, a structural analysis was

performed to determine how the crystal desolvated. Examination of the global

structure of CBMOP revealed that solvent-filled channels exist along the c-axis, and

the included solvent has access to the surface of the crystal via these channels. The

channels are two-dimensional with a ladder-like structure, and the infinite chains that

extend along the c-direction are also connected to one another by short channels that

Page 170: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

147

roughly run parallel to the a-axis.

These channels, shown in Figure 4.12,

were observed to be larger in the

CBMOP-solvent structure relative to

the CBMOP-desolvated structure.

Face indexing of a CBMOP-solvent

crystal was performed to determine

what faces were expressed. Figure

4.11 is a schematic representation of

the crystal habit of CBMOP; it

revealed that the (0 0 1) face was the

predominant crystal face. As shown in

Figure 4.12, the solvent channels run

roughly normal to this face. The solvent,

therefore, is likely to leave the crystal

through its largest face.

4.4. Desolvation/Resolvation of Bulk CBMOP Observed by Powder X-ray

Diffraction

Powder X-ray diffraction (PXRD) was employed to confirm that the desolvation

behavior occurred in the bulk material, and to determine if one could regenerate the

Figure 4.10. Changes in c-axis (top) and b-axis

(bottom) before and after desolvation. Green

points represent individual CBMOP∙solvent

crystals, except for labeled spot. The orange

points below the green points represent

CBMOP∙solvent crystals after desolvation

(CBMOP∙desolvated.

Page 171: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

148

original CBMOP-solvent material by

resolvating the CBMOP-desolvated

material (i.e. suspend the material in

mother liquor). It was necessary to

calculate powder X-ray patterns from the

room temperature unit cell data for CBMOP-solvent (Figure 4.13 a-b) and CBMOP-

desolvated, (Figure 4.13 c); LAZY-PULVERIX,20

an X-SEED program interface,

was used to calculate the powder X-ray diffraction patterns. LAZY-PULVERIX was

also employed to broaden the narrow peaks in the calculated powder X-ray

diffraction patterns, to better match the broad peaks observed in powder X-ray

diffraction.

CBMOP-solvent was packed into an open capillary tube and allowed to

desolvate over a period of weeks. The PXRD patterns were obtained over time and

the resulting patterns were compared to the calculated powder patterns for CBMOP-

solvent and CBMOP-desolvated. The broadened theoretical pattern for CBMOP-

solvent (Figure 4.13.b) matched well with the experimental pattern for bulk

CBMOP-solvent (Figure 4.13.d), confirming that the bulk crystalline material

possesses the same structure as the single crystals. Furthermore, three reflections, the

(110), (220) and (221), could be uniquely assigned in the experimental pattern, as

Figure 4.11. Faces of crystal CBMOP.

Page 172: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

149

Figure 4.12. Top: Spacefill representations of the single crystal structures of CBMOPsolvent

(top) and CBMOPdesolvated (bottom) as viewed down the [001] direction. The lattice solvent

molecules have been removed for clarity. Note that the channels collapse upon desolvation.

Bottom: Connolly surface plot of CBMOP-solvent as viewed normal to the [010] direction. The

dark blue highlighted areas are the channels observed above. The channels form a ladder-like

structure extending along the c-axis, and having rungs that run roughly along the a-axis.

Page 173: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

150

these reflections were sufficiently distant from neighboring peaks to prevent

significant overlapping.

The calculated powder patterns for CBMOP-solvent and CBMOP-

desolvated predicted that the (110) and (220) reflections, as well as the diffraction

peak near 6.8 2θ, would shift to slightly lower 2θ values upon desolvation.

Experimentally this was confirmed, as seen in Figure 4.13.d-f, which showed that the

bulk material converted from CBMOP-solvent to CBMOP-desolvated over about 1

week in the capillary tube. However, after 22 days (Figure 4.13.h), the powder

pattern of the CBMOP-based material had broadened significantly, while

maintaining the same general features. This suggests that the material changes

beyond the single-crystal to single-crystal structural changes observed, but that the

same polymeric structure likely remains. Also, broadening of crystalline peaks may

indicate a reduction in crystallite size as a result of desolvation. The addition of

mother liquor to the highly desolvated material led to a complete restoration of the

solvated material, as evidenced by the powder diffraction pattern (Figure 4.13.j).

Thus, it appears that the initial stages of desolvation occur in a single crystal to single

crystal fashion, and this has been successfully characterized; however, the powder

diffraction data imply that even further desolvation occurs over the time frame of

days to weeks, resulting in fracture of the single crystals (See Figure 4.14 for model).

Though we were unable to structurally characterize the highly desolvated material,

the reversion to CBMOP-solvent upon addition of mother liquor after three weeks

Page 174: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

151

suggests that the highly desolvated CBMOP structure is similar to CBMOP-

desolvated.

Figure 4.13. Calculated powder X-ray diffractograms were determined using the program

LAZY-PULVERIX and the single-crystal x-ray diffraction data for CBMOP-solvent and

CBMOP-desolvated. The room temperature unit cells for each material were used to more

accurately match the experimental data, which were collected at room temperature. The

powder x-ray diffraction (PXRD) experiment was performed on one open capillary tube of

CBMOP collected over time. Experimental partial powder X-ray diffractograms (PXRD) of

the desolvation process of CBMOP∙solvent. a) Calculated PXRD pattern using the unit cell

CBMOP∙solvent obtained at room temperature. b) Calculated PXRD pattern of

CBMOP∙solvent that has been arbitrarily broadened to more accurately reflect the peaks widths.

c) Calculated PXRD of CBMOP∙desolvated at room temperature. d) Experimental PXRD

pattern of CBMOP at t = 0. e) PXRD pattern of CBMOP at t = 1 day. f) PXRD pattern of

CBMOP at t = 6 days. g) PXRD pattern of CBMOP at t = 8 days. h) PXRD pattern of CBMOP

at t = 14 days. i) PXRD pattern of CBMOP at t = 22 days. j) Experimental PXRD pattern of the

material in (i) after moistening the material with mother liquor.

4.5. Gas Sorption Study on CBMOP-desolvated

Many coordination polymer-based materials have been shown to successfully

absorb gases21

due to their permanent porosity. We wanted to examine the

Page 175: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

152

structurally flexible CBMOP-desolvated material to determine if this material would

exhibit some porous behavior and thus

adsorb gases.

In a collaborative effort with Robert

Fairchild and Susumu Kitagawa, synthesized

CBMOP-desolvated was tested for porosity

at Kyoto University using the FMS-BG (Bel

Inc.) automatic gravimetric absorption

measurement system.22

The CBMOP

material was heated to 80C for 5 hours

under high vacuum. The „empty‟ material was cooled to 77K and nitrogen gas

sorption was monitored gravimetrically. The data reveal a so-called Type X sorption

isotherm, indicating that the material does not possess permanent pores.

4.6. Synthesis of a Cryptophane-Dimer

Having found a synthetic protocol that successfully produced a single crystal of

a cryptophane-based MOF, attempts were made to modify this synthesis to produce a

multi-dimensional material. One modification included the addition of HNO3 to the

solution mixture. The result of acidification of the reaction medium was the slow

growth of dark blue crystals. X-ray analysis of these crystals revealed a metal-

organic cryptophane dimer, which was the first example of a metal-organic container

Figure 4.14. Proposed model of

desolvation of CBMOF over time.

Page 176: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

153

molecule dimer with formula [Cu(H9py)(py)3]9DMF, hereafter CBD or container-

based dimer. The quality of this structure was relatively poor, but the data was

sufficient to locate the atoms of the cryptophane, as well as the metal cations and its

corresponding ligands. The cryptophane dimer can be described as an “incomplete

polymerization” in that one carboxylic acid remains protonated while two carboxylic

acids have been deprotonated and have coordinated with copper(II) cations. The

structure of this material explained the crystal‟s softness and instability in air (and in

heated mother liquor), as the material was molecular, not an infinite polymeric

material. Unlike the polymer CBMOF, CBD exhibits no single-crystal-to single-

crystal behavior.

While the dimeric CBD appears to be similar to that of polymer CBMOF, there

are significant differences in the coordination and in the connectivity about the

copper cations. The copper metal center in CBD was five-coordinate (T-Shaped),

with three ancillary pyridine ligands, rather than four and six-coordinate metal centers

of CBMOF (square planar and Jahn-Teller distorted octahedral), with two ancillary

pyridines (Figure 4.15). The carboxylates are also syn to one another (O4C-O3C-

O4A-O3A = 7.3), rather than anti or gauche as in CBMOF. The unreacted

protonated carboxylic acid groups hydrogen bond to a DMF lattice solvent molecule

(O3B-O12X = 2.56Å).

Page 177: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

154

4.7. Conclusions

The first container-based polymers were synthesized and characterized

structurally. (±)-anti-H39 formed a linear 1-D polymer comprised of cryptophane

molecules that hydrogen bond to one another, while CBMOF was a bent 1-D

polymer comprised of cryptophane molecules that link through coordination via their

carboxylate moieties. CBMOF was found to retain its crystallinity upon partial

desolvation, and this single-crystal to single-crystal process was monitored by X-ray

crystallography (single crystal and powder). The unit cell shrinkage was monitored

in real time, revealing a deceleratory rate of cell shrinkage. The unit cell was found

to decrease by nearly 4% in volume, with minimal structural rearrangement. The

faces of the crystal were indexed, and solvent channels that lie perpendicular to the (0

0 1) face were shown to shrink upon desolvation. Both single crystal and powder x-

ray crystallography revealed that the single crystal to single crystal desolvation

observed is not complete; however, the completely desolvated CBMOF material

could not be structurally defined. The powder x-ray diffraction pattern taken after

several weeks suggested that the basic structural features remained intact at greater

levels of desolvation. Furthermore, the introduction of the highly desolvated powder

yielded an immediate return to the completely solvated CBMOF material. The

resolvation of a single crystal of CBMOF-desolvated caused the single crystal to

occlude and crack, meaning that while the resolvation was a bulk property, it was not

a single crystal to single crystal property.

Page 178: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

155

4.8. Experimental

4.8.1. General Methods

All solvents and reagents were used without further purification. Uncorrected

melting points were performed on a Thomas Hoover capillary melting point

apparatus. Elemental analysis data were collected using a Perkin Elmer 2400 CHN

microanalyzer. Infrared spectroscopy was performed using a Perkin-Elmer Spectrum

One FTIR, scanning from 4500 cm-1

to 500 cm-1

. Powder X-ray diffraction was

obtained on a Rigaku RAPID powder diffractometer (CuKα) equipped with a curved

image plate area detector. Single crystal X-ray diffraction data was obtained on a

Bruker-AXIS SMART diffractometer (MoKα) at -100ºC unless explicitly stated.

Lattice parameters were determined from least-square analysis and the reflection data

was integrated using SAINT. Structures were solved using direct methods and

refined by full matrix least-squares based on F2 using X-SEED.

4.8.2. Synthesis of New Materials

CBMOP: (±)-Anti-H39 (103 mg, 0.082mmol) and Cu(NO3)22.5H2O (72 mg,

0.31mmol) were dissolved in 2:1DMF/MeOH (36mL) and pyridine (3.6mL). The

solution was heated to 110˚C for 13 hours in a Parr bomb. The solution was

subsequently cooled at 0.1˚C/min to 90˚C, and held at 90˚C for 1 hour. The solution

was further cooled at 0.1 ˚C/min to 60˚C, isothermed at 60˚C for 1 hour, then cooled

Page 179: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

156

at 0.1˚C/min to 40˚C. Dark blue crystals were filtered and weighed (129 mg, yield =

90% (yield based on [Cu1.5((±)-9DMF)(C6H5N)3(MeOH)]∙2MeOH∙DMF per ASU

as estimated by SQUEEZE). Melting point and elemental analysis unable to be

determined due to the highly variable nature of the crystalline material. Powder X-

Ray diffraction of the bulk material possesses the same structure as that confirmed in

the single crystal determination.

CBD: (±)-Anti-H39 (1.5 mg, 0.0012 mmol) and Cu(NO3)22.5H2O (30.7 mg,

0.13 mmol) were dissolved in 2:1DMF/MeOH (1 mL) and pyridine (1 mL). After 11

days, an additional 0.6 mL of MeOH and 3 drops (0.1 mL) conc. HNO3 were added.

The solution was evaporated and dark blue crystals were observed. Mass, melting

point and elemental analysis were not determined due to the highly variable nature of

the crystalline material.

4.8.3. Crystal Structures

CBMOP-solvent: C94H89N6O20Cu1.5, Mr = 1690.08, 0.44 x 0.24 x 0.18 mm,

monoclinic, space group C2/c (no. 15), a = 52.87(1)Å, b = 16.102(3)Å, c =

23.170(4)Å, = 102.582(3). V = 19250(6)Å3, T = 173(2)K, Z = 8, calcd = 1.17

gcm3, Mo-K radiation, = 0.71073Å, 2max = 46, scans, 173(2)K, 38123 total

reflections, 13381 unique reflections, 8119 reflections with I2(I) (Rint = 0.062);

absorption correction SADABS (Tmin = 0.8435, Tmax = 0.9314, = 0.40 mm-1

),

Page 180: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

157

structure solution using SHELXS, refinement (against F2) with SHELX-97-2,

1107 parameters, 0 restraints, H atoms placed in calculated positions and refined with

a riding model, R1 = 0.0692 (I2(I)) and wR2 = 0.2037 (all data), residual electron

density max/min = 0.69/-0.66 e-/Å

3, GOF = 0.971. Disordered solvent molecules

were modeled using the SQUEEZE subroutine of the program PLATON,[1]

which

estimates the solvent accessible volume is 3880 Å3

to be occupied by 573 electrons

per unit cell (calculated 2 MeOH molecules and 1 DMF molecule).

CBMOP-desolvated: C94H89N4O24Cu1.5, Mr = 1712.08, 0.36 x 0.18 x 0.14

mm, monoclinic, space group C2/c (no.15), a = 53.319(6)Å, b = 15.912(2)Å, c =

22.444(2)Å, = 102.719(2). V = 18574(3) Å3, T = 173(2)K, Z = 8, calcd = 1.22

gcm-3

, Mo-K radiation, = 0.71073Å, 2max = 44.2, scans, 173(2)K, 51526 total

reflections, 11482 unique reflections, 4551 reflections with I2(I) (Rint = 0.1090);

absorption correction SADABS (Tmin = 0.8647, Tmax = 0.9441, = 0.42 mm-1

),

structure solution using SHELXS, refinement (against F2) with SHELX-97-2, 977

parameters, 4 restraints, H atoms placed in calculated positions and refined with a

riding model, R1 = 0.1090 (I2(I)) and wR2 = 0.3301 (all data), residual electron

density max/min = 0.75/-0.42 e-/Å

3, GOF = 0.938. The high values of the merging

and final R factor values are attributed to the degradation of the crystal as a result of

the desolvation process, or as a result of greater molecular motion as the remaining

species attempt to fill space. It was impossible to model the remaining disordered

Page 181: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

158

solvent; SQUEEZE analysis reveals the solvent-accessible volume to be 2994 Å3

which is occupied by 397 electrons (calculated. 1DMF and 0.5 MeOH molecules).

The program X-SEED[2]

was used as a graphical interface to SHELX and for the

generation of figures.

CBD: C94H89N6O20Cu1.5, Mr = 1690.08, 0.44 x 0.24 x 0.18 mm, triclinic,

space group P-1 (no. 2), a = 14.03(1)Å, b = 16.42(1)Å, c = 26.75(2)Å, = 99.90(2),

= 92.40(2), = 108.54(2). V = 5726()Å3, T = 173(2)K, Z = 2, calcd = 1.28 gcm

-3,

Mo-K radiation, = 0.71073Å, 2max = 44, scans, 173(2)K, ___ total reflections,

13799 unique reflections, 4031 reflections with I2(I) (Rint = 0.115); absorption

correction SADABS (Tmin = ___, Tmax = ___, = 0.27 mm-1

), structure solution using

SHELXS, refinement (against F2) with SHELX-97-2, 512 parameters, 0 restraints,

H atoms placed in calculated positions and refined with a riding model, R1 = 0.118

(I2(I)) and wR2 = 0.300 (all data), residual electron density max/min = 1.36/-1.28e

-

/Å3, GOF = ___. Disordered solvent molecules were modeled using the SQUEEZE

subroutine of the program PLATON,[1]

which estimates the solvent accessible

volume is 2170 Å3

to be occupied by 708 electrons per unit cell (calculated 18 DMF

molecules per unit cell). The poor quality of this dataset was a consequence of the

large number of lattice solvent molecules, which comprise approximately 23% of the

crystal‟s mass. This crystal is relatively unstable, even to temperature increases in

the mother liquor, which cause the crystals to crack and blacken.

Page 182: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

159

4.8.4. Unit Cell Analysis over time

Crystals were removed from the mother liquor and quickly placed into the

cold temperature stream (-100C) and centered. The crystals were indexed at the low

temperature, and then removed from the cold temperature stream for a period of time.

The crystals were then reintroduced to the cold temperature stream, recentered and

reindexed. The total unit cell volumes are plotted with respect to the total amount of

time that the crystals were exposed to room temperature conditions. Note that the

rate of volume change is greatest initially; this deceleratory rate change is common in

solid-state desolvation processes.

Page 183: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

160

4.9 References

1. (a) Atwood, J. L.; Barbour, L. J.; Jerga, A.; Schottel, B. L. Science, 2002,

298, 1000-1002. (b) Dobrzanska, L.; Lloyd, G. O.; Raubenheimer, H. G.;

Barbour, L. J. J. Am. Chem. Soc. 2006, 128, 698-699.

2. (a) Barbour, L, J. Aust. J. Chem. 2006, 59, 595-596. (b) Halder, G. J.;

Kepert, C. J. Aust. J. Chem. 2006, 59, 597-604. (c) Suh, M. P.; Cheon, Y.

E. Aust. J. Chem. 2006, 59, 605-612. (d) Friscic, T.; MacGillivray, L. R.

Aust. J. Chem. 2006, 59, 613-616. Wang, Z.; Zhang, Y.; Kurmoo, M.; Liu,

T.; Vilminot, S.; Zhao, B.; Gao, S. Aust. J. Chem. 2006, 59, 617-628.

3. (a) Matsuda, R.; Kitaura, R.; Kitagawa, S.; Kubota, Y.; Belosludov, R. V.;

Kobayashi, T. C.; Sakamoto, H.; Chiba, T.; Takata, M.; Kawazoe, Y.; Mita,

Y. Nature, 2005, 436, 238-241.

4. (a) Te, R. L.; Griesser, U. J.; Morris, K. R.; Byrn, S. R.; Stowell, J. G. Cryst.

Growth. Des. 2003, 3, 997-1004. (b) Wei, Q.; Nieuwenhuyzen, M.; Meunier,

F.; Hardacre, C.; James, S. L. Dalton Trans. 2004, 1807-1811.

5. (a) Zhang, J.-P.; Lin, Y.-Y.; Zhang, W.-X.; Chen, X.-M. J. Am. Chem. Soc.

2005, 127, 14162-14163. (b) Maji, T. K.; Mostafa, G.; Matsuda, R.;

Kitagawa, S. J. Am. Chem. Soc. 2005, 127, 17152-17153.

6. Irngartinger, H.; Skipinski, M. Tetrahedron, 2000, 56, 6781-6794.

Page 184: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

161

7 . (a) Kim, J. H.; Lindeman, S. V.; Kochi, J. K. J. Am. Chem. Soc. 2001, 123,

4951-4959. (b) Toh, N. L.; Nagarathinam, M.; Vittal, J. J. Angew. Chem.,

Int. Ed. 2005, 44, 2237-2241. (c) Devic, T.; Batail, P.; Avarari, N. Chem.

Commun. 2004, 1538-1539. (d) Armentano, D.; De Munno, G.;

Mastropietro, T. F.; Julve, M.; Lloret, F. J. Am. Chem. Soc. 2005, 127,

10778-10779.

8. Suh, M. P.; Ko, J. W.; Choi, H. J. J. Am. Chem. Soc. 2002, 124, 10976-

10977.

9. Saied, O.; Maris, T.; Wuest, J. D. J. Am. Chem. Soc. 2003, 125, 14956-

14957.

10. (a) Biradha, K.; Fujita, M. Angew. Chem., Int. Ed. 2002, 41, 3392-3394. (b)

Biradha, K.; Hongo, Y.; Fujita, M. Angew. Chem., Int. Ed. 2002, 41, 3395-

3398. (c) Ananchenko, G. S.; Udachin, K. A.; Dubes, A.; Ripmeester, J. A.;

Perrier, T.; Coleman, A. W. Angew. Chem., Int. Ed. 2006, 45, 1585-1588.

(d) Ohmori, O.; Kawano, M.; Fujita, M. J. Am. Chem. Soc. 2004, 126,

16292-16293.

11. Schmidt, G. M. J. Pure Appl. Chem. 1971, 27, 647-678.

12. Schmidt, G. M. J.; Cohen, M. D. J. Chem. Soc. 1964, 1996.

13. (a) Ramamurthy, V.; Venkatesan, K. Chem. Rev. 1987, 87, 433-481. (b)

Tanaka, K.; Toda, F. Chem. Rev. 2000, 100, 1025-1074.

Page 185: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

162

14. (a) Suzuki, T.; Fukushima, T.; Yamashita, Y.; Miyashi, T. J. Am. Chem. Soc.

1994, 116, 2793-2803. (b) Varshney, D. B.; Papaefstathiou, G. S.;

MacGillivray, L. R. Chem. Commun. 2002, 1964-1965. (c) Takahashi, S.;

Miura, H.; Kasai, H.; Okada, S.; Oikawa, H.; Nakanishi, H. J. Am. Chem.

Soc. 2002, 124, 10944-10945.

15. Hoang, T.; Lauher, J. W.; Fowler, F. W. J. Am. Chem. Soc. 2002, 124,

10656-10657.

16. An ISI Web of Knowledge search for “single-crystal to single-crystal”

processes yielded 4 matches between 1980 and 1990, and 59 matches in 2010

alone.

17. Dunitz, J. D.; Schomaker, V.; Trueblood, K. N. J. Phys. Chem. 1988, 92,

856.

18. Van Der Sluis, P.; Spek, A. L. Acta Crystallogr. Sect. A. 1990, 46, 194-201.

19. Brown, M. E. Introduction to Thermal Analysis: Techniques and

Applications, Chapman and Hall, London, 1988.

20. LAZY PULVERIX is a computer program designed to calculate x-ray and

neutron powder diffraction patterns from single-crystal x-ray and neutron

data. Yvon, K.; Jeitschko, W.; Parthe‟, E. J. Appl. Cryst. 1977, 10, 73-74.

21. (a) Rowsell, J. L. C.; Spencer, E. C.; Eckert, J.; Howard, J. A. K.; Yaghi, O.

M. Science, 2005, 305, 1350-1354. (b) Kepert, C. J.; Rosseinsky, M. J.

Page 186: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

163

Chem. Commun. 1999, 375-376. (c) Li, H.; Eddaoudi, M.; O‟Keeffe, M.;

Yaghi, O. M. Nature, 1999, 402, 276-279.

22. Kitaura, R.; Seki, K.; Akiyama, G.; Kitagawa, S. Angew. Chem., Int. Ed.

2003, 42, 428-431.

Page 187: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

164

CHAPTER 5: SOLID-STATE KINETICS OF SUPRAMOLECULAR HOST-

GUEST CLATHRATES

5.1. Introduction: Solid-State Kinetics

The concept of solid-state kinetics seems to be a contradiction in terms. Most

chemists are accustomed to thinking about kinetics in the gas-phase or in solution;

this can be attributed to several factors. First, the Arrhenius equation (k = Aexp(-Ea/

RT) was derived empirically, and found to best fit reactions occurring in the gas

phase and in solution.1 Even today, the Arrhenius equation is often used to report

kinetic data in terms of activation energy (Ea) and frequency factor (log A). Henry

Eyring expounded upon Arrhenius’s work, and found a relation between Arrhenius’

empirical variables and more meaningful thermodynamic quantities (Ea = ΔH‡

+ RT;

ln A ΔS‡).2 The Eyring equation (k = (kBT/h)*exp(-∆H

‡/RT)*exp(-∆S

‡/R)) allows

for direct calculation of these useful parameters, which consequently allows for the

determination of ΔG‡ at any temperature.

3 Furthermore, Eyring related these kinetic

parameters to transition state theory, meaning that the Eyring equation is a theoretical

construct rather than and not an empirical model. The Eyring equation is the

accepted means to study gas, solution-based, and mixed-phase reactions. It is crucial

to note that neither of these theories describes kinetics specifically in the solid-state,

and that both models are geared toward systems with far more mobile reactants that

are not limited in their ability to achieve a specific transition state. The Arrhenius

and Eyring models are often taught to students in General Chemistry or Physical

Page 188: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

165

Chemistry; however, kinetic study of solid-state processes is rarely discussed, even in

graduate level study.

Nonetheless, solid-state reactions and thermal processes can and do occur; a

single-crystal to single-crystal desolvation was described in Chapter 4. Reactions

also occur in polycrystalline and amorphous solids. Several important nonacademic

examples of solid state reactions include the dehydration of important materials such

as borax4 and carbamazepine dehydrate,

5 determining the activation energy to the

reactions of high energy materials,6 solid state polymerization,

7 and even cooking.

8

These reactions occur in a fairly reproducible fashion which implies that they are

following a reaction coordinate that is similar to reactions in solution. It is this belief

that has led solid-state scientists to study solid-state kinetics; however, there is a great

deal of debate in the literature as to the applicability of activation energy (and by

association ΔH‡

and ΔS‡) to solid state reactions,

9 because conditions in the solid

state do not allow reaction collisions to occur as in solution or in the gas phase.

Unlike solution kinetic studies, which often rely on spectroscopic methods,

reactions in the solid state are often monitored using either calorimetric or

gravimetric instrumentation.10

Thermogravimetric analysis (TGA) is an attractive

instrument for monitoring all processes that involve a mass change (dehydration,

desolvation), while differential scanning calorimetry (DSC) can be used to evaluate

and quantitate the heat-flow associated with solid-state processes. These instruments

measure changes in mass (TGA) and heat flow (DSC) as a function of temperature

Page 189: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

166

and time.11

In these experiments, the fractional extent of a solid state reaction () can

be measured against both variables, which has consequences on the kinetic

experiments performed. Two main approaches have been developed to deal with

these three variable experiments: isothermal and nonisothermal (or dynamic)

methods.12

In isothermal kinetics, various techniques are used to measure the variable ,

which is defined as the extent of the reaction and varies from 0 to 1. In isothermal

TGA experiments, one initially calculates in gravimetric experiments by

determining the mass change at time n and dividing by the total mass change. Alpha,

in turn, is fitted to various models (Table 5.1) that have been derived from observed

crystallization processes.13

The experimentally determined alpha is inserted into the

integral form of the equations given in Table 5.1, and each equation is plotted as a

function of time. The best-fit line is calculated for each model, and the linearity of

Table 5.1. Solid-state rate expressions for several reaction models.

Model Integral Form

g(α) = kt

Differential Form

f(α) = 1/k dα/dt

Zero-Order (R1) α 1

Contracting Area (R2) [1-(1- α)1/2

] 2(1- α)1/2

Contracting Volume (R3) [1-(1- α)1/3

] 3(1- α)2/3

1-D Diffusion (D1) α 2

1/2α

2-D Diffusion (D2) [(1- α)ln(1- α)] + α [-ln(1- α)]-1

3-D Diffusion-Jander (D3) [1-[(1- α)1/3

]2 [3(1- α)

2/3]/

[2(1-(1- α)1/3

)]

3-D Diffusion-Ginstiling-Brounshtein (D4) 1-(2α/3)-(1-α)2/3

3/[2((1-α)-1/3

-1)]

Avarami-Erofeyev (A2) [-ln(1-α)]1/2

2(1-α)[-ln(1-α)]1/2

Avarami-Erofeyev (A3) [-ln(1-α)]1/3

3(1-α)[-ln(1-α)] 2/3

Avarami-Erofeyev (A4) [-ln(1-α)]1/4

4(1-α)[-ln(1-α)]3/4

First-order (F1) [-ln(1-α)] (1-α)]

Page 190: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

167

each plot is statistically evaluated. The model that gives the most linear curve is

detetrmined to be the most appropriate kinetic model, and the slope of its best-fit line

represents the rate constant k. The determination of the rate constant at several

temperatures allows for the calculation of Arrhenius or Eyring activation parameters,

by plotting ln k vs. 1/T or ln (k/T) vs. 1/T respectively. This type of experiment is

more familiar to scientists, as solution kinetics studies are usually performed as a

series of isothermal experiments. Isothermal kinetics studies are considered to be

more accurate, but are often time-consuming.14

Nonisothermal kinetics experiments measure as a function of time and

temperature in a single experiment by heating the sample. The most common

nonisothermal kinetic experiments are performed at a constant rate, and

nonisothermal experiments have the general advantage of short experiment time as

compared to isothermal experiments. Nonisothermal experiments may use the

differential form of the models given in Table 5.1 to calculate the rate constant k.

There are additional nonisothermal models (Ozawa (OFW)15

, Kissinger,16

and Coats

and Redfern17

) given in Table 5.2, which monitor the rate of change of alpha or peak

temperature as a function of heating rate.18

Table 5.2. Nonisothermal kinetic models

β = heating rate. α = extent of reaction. Tmax = peak temperature.

Model Plot Slope

Ozawa ln β vs. 1/Tmax Ea/R

Kissinger ln( β/T2

max) vs. 1/Tmax Ea/R

Coats and Redfern ln[1-(1-α)1-n

/T2(1-n)] vs. 1/T -Ea/R

Page 191: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

168

5.2. Desolvation of Host-Guest Systems

Supramolecular chemists, in general, have a keen interest in identifying and

quantifying intermolecular interactions. The crystalline solid state, due to its close

packed nature, can be viewed as a supramolecular entity.19

Molecular crystals offer

an interesting environment in which to study the interplay of intermolecular

interactions. One means to study this is through the examination of host-guest

structures, where the host is defined as a large, stable organic molecule which has

functionality that allows for intermolecular interaction. In these systems, the guest is

usually a small molecule, normally a solvent molecule or other volatile species,

which cocrystallizes with the host molecule. While stable at low temperatures, the

guest molecules can be driven off at higher temperatures.

Constrictive binding in container molecules has been established in solution

studies; however, constrictive binding kinetic studies in the solid-state have not been

reported for this class of host-guest crystalline molecular materials. However, there

have been studies on related clathrate inclusion materials. Luigi Nassembini has

studied the desolvation kinetics in a number of simple host-guest inclusion

compounds, as described in “Physicochemical Aspects of Host-Guest Compounds.”20

In one example, Nassembini described the removal of volatiles, including THF,

DMSO, and acetone, from binapthol-solvent clathrate materials.21

In another paper,

Nassembini compared and contrasted the structures and activation parameters of

xanthenol clathrates (Figure 5.1).22

Similarly, this chapter explores the desolvation

Page 192: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

169

kinetics of THF in a

simple host-guest

clathrate system, in

which the host

(cyclotriveratrylene, or

CTV) is structurally

similar to a

cryptophane CTB

subunit. The kinetics of CTVTHF0.5 is examined and compared to the desolvation

kinetics of a cryptophane inclusion material (±)-anti-8THF. This specific

cryptophane inclusion material was chosen because it contained no THF in the lattice,

which could convolute the data interpretation.

5.3. Isothermal TGA Analysis of CTVTHF0.5

Cyclotriveratrylene (CTV) is a C3v symmetric cavitand derived from the acid

condensation of veratrole (1,2 dimethoxybenzene) and formaldehyde.23

The CTV

molecular scaffold has been modified by chemists to form pyramidic liquid crystals,24

larger molecular polyhedra25

and CTV coordination polymers,26

neutral molecule and

anion inclusion complexes,27

hemicryptophanes28

and cryptophanes by metal

coordination,29

iron-sulfur clusters,30

and amino acid glycoconjugates.31

CTV is a

classic inclusion host that forms stoichiometric clathrates when crystallized with

Figure 5.1. Series of xanthenol clathrates studied to relate structure

and desolvation kinetics.

Page 193: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

170

many small molecules, including

water,32

benzene and water,33

THF,

ethyl acetate, benzene, and other

small molecules.34

This behavior has

been well characterized by single

crystal diffraction. It was found that

CTV packing was influenced by the

nature of the included solvent

molecules. Solvents that are hydrogen bond donor induced CTV to pack in

columnar, cup-in-cup stacks in which the donor hydrogen bonds to CTV oxygens.

This is the so-called α-phase. (Figure 5.2).35

The β-phase is induced by

crystallization of CTV with solvents that do not have hydrogen bond donor functional

groups. The most important structural feature that results from the differing α and β

forms is that the β form shows out-of-plane methoxy groups, while the α form has in-

plane methoxy groups.

CTV(THF)0.5 crystallizes in the β form, as THF has no hydrogen bond donors.

The structure of this clathrate is given in Figure 5.2. The oxygen atom of the THF

guest molecule appears to weakly hydrogen bond with a CTV methoxy substituent,

with a C-H…

O = 3.55 and 3.69Å. As a consequence of the structure, it was expected

that the THF molecules would be somewhat stable to desolvation.

Figure 5.2. Crystal structure of CTV0.5THF.

CTV molecules shown in stick, THF molecules

shown in spacefill form.

Page 194: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

171

Figure 5.3. Nonisothermal TGA experiments of CTVTHF0.5 as a function of alpha (extent of

desolvation) at 5C/min and 10/min.

Initially, nonisothermal TGA experiments on CTV(THF)0.5 were performed to

determine the temperature range over which desolvation typically occurs. Figure 5.3

shows the results of two experiments, performed at 5 and 10C/min respectively.

There appears to be a two stage diffusion mechanism: an early stage desolvation that

is first observed at approximately 50C and a late stage desolvation that begins at

approximately 115C. Nonisothermal TGA of the CTV(THF)0.5 sample shows that a

gradual mass loss corresponding to the early stage desolvation was observed over a

50-60C (320K-380K) temperature range. However, the nonisothermal TGA of the

late stage desolvation was completed over a range of approximately 20C (380K-

400K). The early stage was not investigated with regard to desolvation

Page 195: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

172

kinetics, as it only represented a small fraction of the entire thermal event, as

evidenced by the TGA and DSC data (α range <0.25). The later stage desolvation,

which was also monitored by DSC data, was thoroughly studied by isothermal TGA.

Freshly filtered and air dried

samples were crushed lightly

(particle size ~ 0.3-0.7 mm) and

placed on the TGA pan, and

isothermally heated at appropriate

temperatures. As the

CTV(THF)0.5 sample reached the

isothermal temperature, the

material progressed through the

early stage desolvation transition.

Once the isothermal temperature

had been achieved, the material

was held at the prescribed

temperature overnight or until the

mass loss was complete. Figure

5.4 (top) shows the isothermal

mass losses over time for a sample of CTVTHF0.5 held at various temperatures

Figure 5.4. Top: Plot of alpha vs. time for an isothermal

TGA experiment for the desolvation of CTVTHF0.5 at

96.2 C. Bottom: Isothermal TGA data after application

of the three-dimensional diffusion model for the

desolvation of CTVTHF0.5C.

Page 196: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

173

(86C-106C), in which the mass loss data is converted to alpha, and plotted vs. time.

Alpha is also fit to the major functions described in Table 5.1, and plotted in Figure

5.4 (bottom). Analysis of each of these functions revealed that the isothermal kinetic

data best fit with the D3 model, which is the three-dimensional diffusion model for

desolvation (Figure 5.4, bottom). The rate

constant derived from the isothermal

experiments was calculated at various

temperatures, and both the Arrhenius and

the Eyring equations were plotted, and the

Eyring plot is given in Figure 5.5. From

the Arrhenius plot, the experimentally

derived activation energy was found to be

62 kcal/mol and the experimentally

derived frequency factor was 2.3 x 1033

. The values for these data were rather large,

and seemed to indicate an unfathomably high barrier to activation. When these

numbers were converted to enthalpy, entropy and free energy of activation, the

corresponding activation enthalpy and entropy were still very large (∆H‡

= 61(2)

kcal/mol, ∆S‡ = 84(6) cal/molK); however, the Gibbs free energy of activation for the

desolvation was determined to be 36 kcal/mol at 298K and 28 kcal/mol at 398K. The

free energy of activation revealed that the large enthalpic contribution to the

activation barrier was tempered by the highly positive activation entropy.

Figure 5.5. Eyring plot of desolvation of

CTVTHF0.5 derived from isothermal TGA

data.

Page 197: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

174

5.4. Nonisothermal Kinetic Analysis of CTVTHF0.5

One goal of this study was to compare isothermal and nonisothermal

methodology to determine if they give similar values. Two major instruments are

used in nonisothermal solid-state kinetics: DSC and TGA. Since nonisothermal

TGA of CTVTHF0.5 appeared to have two distinct mass losses, this method was not

chosen. Nonisothermal calorimetric analyses of CTVTHF0.5 were simpler to

interpret, as only one endothermic peak was found. Varying the heating rate from

1C/min to 20C/min resulted in maxima in endothermic heat flow ranging from

96C to 112C, respectively. Figure 5.6 reveals an overlay of DSC traces for

CTVTHF0.5 for all temperature ramps (1, 2.5, 5, 10, and 20C/min). As

expected, the desolvation temperature increases with increasing heating rate

(hereafter ). An Ozawa plot was generated (ln vs. 1/T) to analyze the DSC data

(Figure 5.7). The derived slope is defined as -Ea/R; the corresponding activation

energy was calculated to be 58 kcal/mol. The intercept of the Ozawa plot

corresponds to lnA, such that the frequency factor A = 1.7 x 1034

. Since the

Arrhenius parameters are related mathematically to the Eyring parameters, we can

calculate ∆H‡ and ∆S

‡. (∆H

‡ = 57(2) kcal/mol; ∆S‡ = 96(6) cal/molK) Statistically,

the experimentally determined values were similar in the isothermal and

Page 198: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

175

Figure 5.6. DSC overlay of desolvation of CTVTHF0.5 at varying heating rates.

nonisothermal experimental determinations of ∆H‡

(57(2) kcal/mol vs. 61(2) kcal/mol

respectively)

and ∆S‡

(96(6) cal/molK vs. 84(6)cal/molK respectively). The

agreement between the two sets of data lends credibility to the experimentally

determined values.

5.5. Nonisothermal DSC Analysis of (±)-anti-8THF

One goal of this experiment was to compare the desolvation kinetics of a

container-based material (cryptophanes) wherein the guest in encapsulated with a

Page 199: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

176

simple clathrate (CTV).

However, study of most

cryptophane materials was

complicated by the fact that

most cryptophane-based

materials possess both lattice-

included and encapsulated

solvents, which severely

complicate the desolvation kinetics. In fact, this behavior was clearly observed and

reported in Chapter 3, Figure 3.3 for (±)-anti-4THF3THF, where it was observed

that the 1st three THF equivalents are lost from the material at a much lower

temperature (<120C) than the last equivalent (>120C). Fortunately, cryptophane

(±)-anti-8THF was found to form a 1:1 complex with THF (Figure 5.8). The

structure clearly shows

encapsulation of THF within the

cryptophane cavity.

Since isothermal and

nonisothermal results for the

CTVTHF0.5 clathrate were

statistically identical,

Figure 5.7. Ozawa plot derived from nonisothermal DSC

experiments for the desolvation of CTVTHF0.5 at various

heating rates (1-20C/min).

Figure 5.8. Stucture of (±)-anti-8THF. Left: Side

view of complex. Right: Top view of complex.

Carbon: Gray; Oxygen: Red; Hydrogen: White.

Page 200: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

177

nonisothermal desolvation kinetics for (±)-anti-8THF were performed by DSC.

Figure 5.9 shows the overlay of DSC endotherms at a wide range of heating ramps

(1, 2.5, 5, 10, 20C/min). A comparison of the DSC desolvation endotherms for

the CTVTHF0.5 (Figure 5.6) and (±)-anti-8THF (Figure 5.9) shows significant

differences upon examination. The shapes of the endothermic peaks are quite

symmetric and sharp in the CTV desolvation, while the peaks are broad and

asymmetric in the cryptophane desolvation. In general, the shape of the DSC

endothermic peak may be a consequence of the “order” of the reaction, according to

Kissinger.16

Certainly, the shape of the curve has consequences in data analysis, as it

can be difficult to identify the temperature of maximum endothermic heat flow.

Also, the range of temperatures associated with desolvation is much larger in the

cryptophane (84C-

135C) as compared to

the CTV (96C-112C).

An Ozawa plot

was generated from the

DSC data for

desolvation of (±)-anti-

8THF (Figure 5.10).

It became immediately Figure 5.9. DSC overlay of desolvation of (±)-anti-8THF at

varying heating rates.

Page 201: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

178

apparent that the slope and intercept of this Ozawa curve were much smaller than in

Ozawa plot for CTVTHF0.5 (Figure 5.5). The calculated activation energy, Ea, was

15(2) kcal/mol and the frequency factor log A is 9(1). This result was surprising, as it

indicates that the kinetic

barrier to desolvation was

larger in the CTV inclusion

compound of THF as

compared to the (±)-anti-

8THF encapsulation

complex. When the data for

(±)-anti-8 was converted to

∆H‡

and ∆S‡, we found that ∆S

‡ was negative (-17 cal/molK). This implies that the

rate limiting step is the squeezing of the guest from the cryptophane. Also, as the

temperature increases, the barrier to desolvation also increases. In fact, the seemingly

disparate activation parameters result in statistically identical ∆G‡

at 100C when

comparing the two DSC experiments (Table 5.3).

Table 5.3 Activation parameters for desolvation of CTV·THF0.5 and (±)-anti-8THF.

∆H

(kcal/mol)

∆S‡

(cal/molK)

∆G‡

298

(kcal/mol)

∆G‡

373

(kcal/mol)

CTV (TGA) 61(2) 84(6) 36(3) 30(3)

CTV (DSC) 57(2) 96(6) 28(3) 21(3)

(±)-anti-8 (DSC) 14(1) -17(4) 19(2) 20(2)

Figure 5.10. Ozawa plot derived from nonisothermal DSC

data for the desolvation of (±)-anti-8THF.

Page 202: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

179

5.6. Discussion of Activation Parameters

CTVTHF0.5 desolvation is characterized by a very large activation enthalpy,

as well as a largely positive activation entropy. This data suggests that the transition

state is highly disordered with respect to the ground state, while (±)-anti-8THF

desolvation has a relatively small activation enthalpy, but a negative activation

entropy. The transition state for (±)-anti-8THF is more ordered than the ground

state, which may be explained in this system as a disordered guest squeezing through

a highly oriented host molecule to exit. Despite these differences, the two materials

have similar activation free energy of desolvation (∆G‡). We were surprised that the

container effect did not seem to have a significant effect on the activation free energy

of the desolvation of the cryptophane material. One explanation might be that the

encapsulated THF molecules can exit the molecule easily, as seen from the side view

in Figure 5.8. Compare that with the structure of (±)-anti-4THF, in which the

cryptophane appears to more effectively encapsulate the guest (Figure 5.11). Another

explanation may be that molecular

implosion described in Chapter 3 is the

rate-limiting step, and that this perturbation

in the solid-state causes the crystal to

degrade and desolvate. A more

comprehensive study of CTV and Figure 5.11. Side view of (±)-anti-4THF

complex.

Page 203: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

180

cryptophane clathrates may be useful in determining if there is a single rate-limiting

step that defines the desolvation of these materials.

5.7. Conclusions

The solid-state kinetics of desolvation were studied in two systems:

CTVTHF0.5 and (±)-anti-8THF. CTVTHF0.5 was analyzed using both isothermal

TGA as well as nonisothermal DSC techniques. The results for ∆H‡

and ∆S‡ were

statistically the same for the two different techniques. (±)-anti-8THF was analyzed

using nonisothermal DSC as well. While the values of ∆H‡

and ∆S‡

were very

different for the two systems, their values for ∆G‡

373 calculated from the DSC

experiments were statistically similar. The so-called “container effect” did not

seemingly enhance the stability of (±)-anti-8THF in the solid-state; however, its

structure may be sufficiently open to allow egress of the THF moiety. Another

theory is that the conversion of CTB to saddle twist in the solid-state is the rate-

limiting step in desolvation of these materials. Further study may help elucidate if

there is a common mechanism to desolvation.

5.8. Experimental

5.8.1. General Methods

All solvents and reagents were used without further purification

Page 204: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

181

Thermogravimetric analyses were performed using a TA Instruments TGA 2050

under a constant stream of nitrogen gas. Differential Scanning Calorimetry was

performed using a TA Instruments DSC

5.8.2. Sample Preparation

Clathrate samples (CTVTHF0.5, 8THF) were prepared by dissolving

material in THF and placing in the freezer. The materials were removed from the

freezer and the crystals were vacuum filtered before thermal analysis. The crystals

were air dried for approximately 5 minutes before thermal analysis to remove surface

solvent molecules.

For DSC experiments, between 1 mg and 10 mg of material were weighed in

a DSC pan, which was immediately sealed and vented with a pinhole through the lid.

The material was immediately run on the DSC instrument at heating rates ranging

from 1C to 20C/min. The endothermic transition was plotted, and the temperature

corresponding to the maximum endothermic heat flow was recorded.

For isothermal TGA experiments, between 1 mg and 10 mg of material was

placed on a tared TGA pan. The material was ramped to the appropriate temperature

and held isothermally. The mass loss was plotted relative to time at a given

temperature, which was used to generate an alpha vs. time plot.

Page 205: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

182

5.9 References

1. Atkins, P. W. Physical Chemistry, 1994, W. H. Freeman and Co., New York.

2. Eyring, H. J. Chem. Phys. 1935, 3, 107-116.

3. ∆G‡ = ∆H

‡ - T∆S‡

4. Ekmekyapar, A.; Baysar, A.; Kunkul, A. Ind. Eng. Chem. Res. 1997, 36,

3487-3490.

5. Surana, R.; Pyne, A.; Suryanarayanan, R. PharmSciTech, 2003, 4, 1-10.

6. Vyazovkin, S.; Clawson, J. A.; Wight, C. A. Chem. Mater. 2001, 13, 960-

966.

7. Shi, C.; Gross, S. M.; DeSimone, J. M.; Kiserow, D. J.; Roberts, G. W.

Macromolecules, 2001, 34, 2060-2064.

8. Bayram, M.; Omer, M. D. J. Food Eng. 2004, 64, 43-51.

9. Galwey, A. K.; Brown, M. E. Thermochim. Acta 2002, 386, 91-98.

10. Brown, M. E. Introduction to Thermal Analysis: Techniques and

Applications, 1988, Chapman and Hall, New York.

11. Wunderlich, B. Thermal Analysis, 1990, Academic Press, Boston.

12. Johnson Jr., D. W.; Gallagher, P. K. J. Phys. Chem. 1972, 76, 1474-1479.

13. Gotor, F. J.; Criado, J. M.; Malek, J.; Koga, N. J. Phys. Chem. A, 2000, 104,

10777-10782.

14. Ozawa, T. J. Therm. Anal. Cal. 2001, 64, 109-126.

15. Ozawa, T. J. Thermal Anal. 1970, 2, 301-324.

Page 206: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

183

16. Kissinger, H. E. Anal. Chem. 1957, 23, 1702-1706.

17. Coats, A. W.; Redfern, J. P. Nature 1964, 201, 68-69.

18. Doyle, C.; J. Appl. Polym. Sci. 1962, 6, 639-642.

19. Desiraju, G. R. The Crystal as a Supramolecular Entity, 1996, Wiley, New

York.

20. Nassimbeni, L. R. Acc. Chem. Res. 2003, 36, 631-637.

21. Nassimbeni, L. R.; Su, H. New J. Chem. 2002, 26, 989-995.

22. Jacobs, A.; Nassimbeni, L. R.; Su, H.; Taljaard, B. Org. Biomol. Chem.

2005, 3, 1319-1322.

23. Robinson, G. M. J. Chem. Soc. 1915, 102, 267-276.

24. Zimmermann, H.; Bader, V.; Poupko, R.; Wachtel, E. J.; Luz, Z. J. Am.

Chem. Soc. 2002, 124, 15286-15301.

25. Sumby, C. J.; Hardie, M. J. Angew. Chem., Int. Ed. 2005, 44, 6395-6399.

26. (a) Ahmad, R.; Hardie, M. J. Cryst. Growth Des. 2003, 3, 493-499. (a)

Ahmad, R.; Franken, A.; Kennedy, J. D.; Hardie, M. J. Chem--Eur. J. 2004,

10, 2190-2198. (c) Sumby, C. J.; Hardie, M. J. Cryst. Growth Des. 2005, 5,

1321-1324.

27. (a) Fairchild, R. M.; Holman, K. T. J. Am. Chem. Soc. 2005, 127, 16364-

16365. (b) Staffilani, M.; Bonvicini, G.; Steed, J. W.; Holman, K. T.;

Atwood, J. L.; Elsegood, M. R. J. Organometallics, 1998, 17, 1732-1740. (c)

Holman, K. T.; Orr, G. W.; Steed, J. W.; Atwood, J. L. Chem. Commun.

Page 207: SYNTHESIS AND SOLID STATE STUDIES OF CRYPTOPHANE BASED

184

1998, 2109-2110. (d) Holman, K. T.; Halihan, M. M.; Steed, J. W.; Jurison,

S. S.; Atwood, J. L. J. Am. Chem. Soc. 1995, 117, 7848-7849. (e) Holman,

K. T.; Steed, J. W.; Atwood, J. L. Angew. Chem., Int. Ed. Engl. 1997, 36,

1736-1738. (f) Zhang, S.; Echegoyen, L. J. Am. Chem. Soc. 2005, 127,

2006-2011.

28. Gautier, A.; Mulatier, J.-C.; Crassous, J.; Dutasta, J.-P. Org. Lett. 2005, 7,

1207-1210.

29. Zhong, Z.; Ikeda, A.; Shinkai, S.; Sakamoto, S.; Yamaguchi, K. Org. Lett.

2001, 3, 1085-1087.

30. van Strijdonck, G. P. F.; van Haare, J. A. E. H.; van der Linden, J. G. M.;

Steggerda, J. J.; Nolte, R. J. M. Inorg. Chem. 1994, 33, 999-1000.

31. van Ameijide, J.; Liskamp, R. M. J. Org. Biomol. Chem. 2003, 1, 2661-

2669.

32. Birnbaum, G. I.; Klug, D. D.; Ripmeester, J. A.; Tse, J. S. Can. J. Chem.

1985, 63, 3258-3263.

33. Cerrini, S.; Giglio, E.; Mazza, F.; Pavel, N. V. Acta Cryst. 1979, B35, 2605-

2609.

34. Holman, K. T. unpublished results.

35. Holman, K. T. Ph.D. Thesis, University of Missouri, 1998.