Thermodynamic studies of halogen bonding in …...ii Thermodynamic studies of halogen bonding in...

252
Thermodynamic studies of halogen bonding in solution and application to anion recognition by Mohammed Golam Sarwar A thesis submitted in conformity with the requirements For the degree of Doctor of Philosophy Department of Chemistry University of Toronto ©Copyright by Mohammed Golam Sarwar 2012

Transcript of Thermodynamic studies of halogen bonding in …...ii Thermodynamic studies of halogen bonding in...

  • Thermodynamic studies of halogen bonding in solution and application to anion recognition

    by

    Mohammed Golam Sarwar

    A thesis submitted in conformity with the requirements For the degree of Doctor of Philosophy

    Department of Chemistry University of Toronto

    ©Copyright by Mohammed Golam Sarwar 2012

  • ii

    Thermodynamic studies of halogen bonding in solution and application to anion recognition

    Mohammed Golam Sarwar Doctor of Philosophy

    Department of Chemistry University of Toronto

    2012 Abstract Halogen bonding (XB), the interaction between electron deficient halogen compounds and

    electron donors, is an established non-covalent interaction in the solid and gaseous phases.

    Understanding of XB in the solution phase is limited. This thesis describes experimental

    studies of XB interactions in solution, and the application of XB interactions in anion

    recognition.

    Chapter 1 is a brief review of current understanding of XB interaction: theoretical

    models, studies of XB in solid and gaseous phases and examples in biological systems are

    discussed. At the end of this chapter, halogen bonding in the solution phase is discussed,

    along with applications of halogen bonding in organic syntheses.

    In chapter 2, linear free energy relationships involving the thermodynamics of

    halogen bonding of substituted iodoaromatics are studied. The utility of substituent constants

    and calculated molecular electrostatic potential values as metrics of halogen bond donor

    ability are discussed. Density Functional Theory (DFT) calculations are shown to have useful

    predictive values for trends in halogen bond strength for a range of donor-acceptor pairs.

    Chapter 3 describes the development of new multidentate anion receptors based on

    halogen bonding. Bidentate and tridentate receptors were found to exhibit significantly

    higher binding constants than simple monodentate donors. These receptors show selectivity

    for halide anions over oxyanions. Using 19F NMR spectra at different temperature, the

    enthalpies and entropies of anion bindings for monodentate and tridentate receptors were

  • iii

    determined. The results indicate a positive entropy contribution to anion binding for both

    mono and tridentate receptors in acetone solvent.

    Finally in chapter 4, some mesitylene based receptors with 3-halopyridinium and 2-

    iodobenzimidazolium donors are introduced. The receptors perform halide anion recognition

    in aqueous solvent system through charge-assisted XB interactions. These findings can allude

    to utility in organic synthesis, supramolecular chemistry and drug design.

  • iv

    ACKNOWLWDGEMENTS

    First I owe many thank to my supervisor, Professor Mark S. Taylor for giving me

    opportunity to work with him throughout my study in this department. His indispensable

    guidance, constant encouragement, excellent supervision and thoughtful suggestions

    immensely contributed to the successful completion of this work. I cannot imagine any better

    environment for encouraging discussion and discovery.

    My appreciation goes to group members in this group for their friendship and help. I would

    like to thank Sunny Lai, Ali Rostami, Doris Lee, Elena Dimitrijevic, Corey McClay, Mike

    Chudzinski, . Many thanks to Pragma Roy, Bojan Dragisic, and Simon Chan, undergraduate

    students for their help in syntheses and titration experiments. I wish to thank the Department

    of Chemistry and the University of Toronto for financial support.

    At last but not the least, I am grateful to my grandfather, parents and siblings for teir

    unconditional support for me over the years.

  • v

    Table of Contents

    ABSTRACT……………………………………………………………………………

    ii

    ACKNOWLEDGMENTS……………………………………………………………..

    iv

    LIST OF TABLES….…………………………………………………………………

    viii

    LIST OF SCHEMES……………………………………………………………………

    x

    LIST OF FIGURES ………………………………………………………………….

    xii

    LIST OF ABBREVIATIONS………………………………………………………….

    xvii

    1. CHAPTER 1……………………………………………………………………..

    1

    1. Halogen bonding: A noncovalent interaction………………………………………

    1

    1.1. Introduction………………………………………………………………………..

    1

    1.2. History of the halogen bonding interaction………………………………………..

    1

    1.3. Halogen bonding in the gas phase…………………………………………………

    6

    1.3.1. Complex formation with ammonia, water and hydrogen sulfide………………

    7

    1.3.2. Complex formation with non-aromatic π–electron donors………………………

    9

    1.4. Halogen bonding in the solid state…………………………………………………

    11

    1.4.1. Halogen bonding in crystal engineering………………………………………..

    12

    1.4.2. Liquid crystals…………………………………………………………………..

    14

    1.4.3. Metal-organic frameworks………………………………………………………

    17

    1.4.4. Nanomaterials…………………………………………………………………...

    19

    1.5. Halogen bonding in medicinal chemistry…………………………………………

    21

    1.6. Halogen bonding in the solution phase……………………………………………

    24

    1.6.1. Halogen bonding in supramolecular chemistry…………………………………

    29

    1.6.2. Syntheses induced by halogen bonding……………………………………….. 32

  • vi

    1.7. Conclusions………………………………………………………………………... 34

    1.8. Bibliography………………………………………………………………………

    36

    2. CHAPTER 2……………………………………………………………………….

    42

    2. Thermodynamics of halogen bonding in solution: substituent, structural, and solvent effects…………………………………………………………………………

    42

    2.1. Introduction……………………………………………………………………….

    42

    2.1.1. Early studies of XB in solution phase…………………………………………..

    42

    2.1.2. XB studies in solution using IR spectroscopy………………………………….

    44

    2.1.3. XB studies in solution using 19F NMR pectroscopy…………………………..

    46

    2.1.4. XB studies in solution using 13C NMR spectroscopy………………………….

    47

    2.2. Results and discussion……………………………………………………………

    50

    2.2.1. Experimental determinations of Ka values for halogen bonding interactions…..

    50

    2.2.2. Substituent effects on the halogen-bond donor ability of iodoperfluoro- benzenes

    51

    2.2.3. Structural Effects on halogen bonding: a combined experimental and computational approach………………………………………………………………

    58

    2.3 Conclusions………………………………………………………………………...

    62

    2.4 Experimental details………………………………………………………………

    64

    2.5 Bibliography………………………………………………………………………

    71

    3. CHAPTER 3………………………………………………………………………

    76

    3. Development of anion receptors using halogen bonding interactions………………

    76

    3.1. Introduction………………………………………………………………………..

    76

    3.2. Results and discussion……………………………………………………………

    80

  • vii

    3.2.1 Association constants and thermodynamic data for interactions of C8F17I and C6F5I with anions………………………………………………………………………

    80

    3.2.2. Development of multidentate halogen bond donors……………………………

    84

    3.2.3. Solvent effects on halogen bonding……………………………………………

    96

    3.2.4. Anion receptor based on hydrogen and halogen bonding interactions…………

    100

    3.2.5. Assembly of ternary complex through hydrogen and halogen bonding interactions…………………………………………………………………………….

    101

    3.3. Conclusions and outlook………………………………………………………….

    102

    3.4. Experimental details………………………………………………………………

    105

    3.5. Bibliography………………………………………………………………………

    125

    4. CHAPTER 4………………………………………………………………………..

    130

    4. Development of cationic halogen bonding receptors based on halopyridinium and haloimidazolium groups….……………………………………………………………

    130

    4.1. Introduction……………………………………………………………………….

    130

    4.2. Result and discussion…………………………………………………………….

    132

    4.3. Conclusions………………………………………………………………………

    136

    4.4. Experimental……………………………………………………………………..

    137

    4.5. Bibliography………………………………………………………………………

    145

    5. CHAPTER 5………………………………………………………………………..

    130

    5. Lesson, hurdles and future directions.……………………………………………….

    146

  • viii

    List of Tables

    Table 1.01. Inhibition of hCatL influence by halogen bonding (XB)

    22

    Table 1.02. Binding constants (M-1) of complexes between pyridine derivatives and THF with halogen bonds donors (CCl4, CBr4, and I2).

    25

    Table 2.01. Equilibrium constant (Ka) for the complex formation of inorganic, perfluorinated haloalkenes (e.g., C6F13I) and the non fluorinated R−C≡C−I halogen bond donors with base pyridine in CCl4

    42

    Table 2.02. Equilibrium constants (Ka) for the complex formation of inorganic halogen bond donors with pyridine in CCl4 at 30 °C.

    43

    Table 2.03. Equilibrium constant (Kc) for halogen bonding of XB with Et2NCOMe in CCl4.

    45

    Table 2.04. Equilibrium constant (Kapp) for the complex of I−C≡CCN with Et2NCOMe in C6H6-CCl4 solvent mixtures.

    46

    Table 2.05. Halogen bonding association constants of para-substituted iodotetrafluorobenzenes (tri-n-butylphosphine oxide acceptor, cyclohexane solvent), with substituent constants (σ para and σ meta) and computed electrostatic potential at the iodine atom.

    53

    Table 2.06. Experimental and predicted binding energies of C6F5I and C8F17I with representative halogen bond acceptors.

    59

    Table 2.07. Computed Halogen Bonding Energies for Comparison to Experimental Data From Table 2.06.

    61

    Table 3.01. Binding constant of C6F5I and C8F17I with n-Bu4N+X− in acetone.

    84

    Table 3.02. Calculated geometries of 3.11·X− and 3.12·X−.

    86

    Table 3.03. Binding constant of 3.11 and 3.12 with n-Bu4N+X− in acetone.

    89

    Table 3.04. Binding constant of C6F5I, C6F4ICOOH, C6F4ICOOBn, and C6F4ICO(NH)Bn with n-Bu4N+X− in acetone

    90

    Table 3.05. Binding constant of receptors 3.11, 3.12, 3.13, and 3.14 with n-Bu4N+X− in acetone

    92

    Table 3.06. Binding constant of 3.21, 3.22 and 3.23 with n-Bu4N+X− in acetone

    95

    Table 3.07. Binding constant of C6F5I, C8F17I and 3.21 with n-Bu4N+Cl− in different solvent system.

    97

  • ix

    Table 3.08. Binding constant of C6F5I with n-Bu4N+Cl− in acetone with the presence of trace amount of water.

    98

    Table 3.09. Association constant 3.13, 3.14, 3.26, 3.27, and Cl− in acetone at 295 K.

    103

    Table 3.10. Calculated halogen-bonding complex (3.21–X– ) mass for comparison to actual mass (ESI)

    123

    Table 4.01. Binding constants (M-1) of receptor 4.07 (X = Cl), 4.08 (X = Br), and 4.09 (X = H) with n-Bu4N+X− in CD3CN:D2O (9:1) and DMSO.

    136

    Table 4.02. Binding constants of 4.13 and 4.15 with n-Bu4N+X− in CD3CN:D2O (8:2).

    137

    Table 4.03. Binding constants of 4.14 and 4.16 with n-Bu4N+X− in CD3CN:D2O (8:2).

    138

  • x

    List of Schemes

    Scheme 1.01. Mechanisms for the reaction of halogen molecules with alkenes (outer and inner complexes) in polar solvents in the dark.

    2

    Scheme 1.02. Halogen bonding and hydrogen bonding interactions: A) halogen bond donor (D−X) with acceptor A, B) hydrogen bond donor (D−H) with acceptor A, and C) hydrogen bond donor (D−H) with acceptor (D−X)

    3

    Scheme1.03. The formation of assembly consisting of functionalized AuNP with organic cross-linkers (BPEB) in solution.

    20

    Scheme 1.04. Stepwise generation of assemblies consisting of functionalized gold nanoparticles (AuNP-1) and organic cross-linkers (BPEB) on silicon organic monolayers (M1, M2).

    20

    Scheme 1.05. Halogen bond catalyzed reduction of 2-phenylquinoline (1.43).

    33

    Scheme 1.06. Reaction of benzhydryl bromide (1.45) with wet CD3CN catalyzed by 1.48 or 1.49.

    33

    Scheme 1.07. Baylis-Hillman reaction between aromatic aldehydes and Michael acceptors.

    34

    Scheme 1.08. Synthesis of N-arylaziridine (1.53) using halogen bonding interaction.

    34

    Scheme 3.01. formation of polyiodide( m and n integers > 0, n = 1 to 4)

    78

    Scheme 3.02. Complex formation of tetraiodide with 1,6-bis(trimethylammonium)-hexadiiodide (3.06).

    79

    Scheme 3.03. Synthesis of 4,4’-bis(2,3,4,5-tetrafluoro-6-iodophenyl)buta-1,3-diyne (3.11) and 3.12

    87

    Scheme 3.04. Synthesis of 4,4’-bis(2,3,4,5-tetrafluoro-6-iodophenyl)buta-1,3-diyne (3.11) and 3.12

    89

    Scheme 3.05. Synthesis of 2,3,4,5-tetrafluoro-6-iodobenzaldehyde (3.16)

    91

    Scheme 3.06. Synthesis of hexadentate receptor 3.25.

    96

    Scheme 3.07. Synthesis of calix[4]arene based intramolecular hydrogen and halogen bonding receptor (3.26).

    100

  • xi

    Scheme 3.08. Synthesis of ester 3.27 from 6-methylpyridin-2-amine.

    101

    Scheme 4.01. Syntheses of receptors 4.07, 4.08 and 4.09 when R = Cl, Br, and H respectively.

    133

    Scheme 4.02. Syntheses of benzimidazolium derived receptors 4.13 and 4.14 when R = CH3 and CH2Ph respectively

    134

    Scheme 4.03. Syntheses of benzimidazolium derived receptors 4.15 and 4.16 when R = CH3 and CH2Ph respectively

    135

  • xii

    List of Figures Figure 1.01. Halogen bonding in Br2 −1,4-dioxane, Br2 –hexamethylenetetramine and Br2 –acetone complexes

    2

    Figure 1.02. The computed B3LYP/6-31+G(d,p) electrostatic potential, on the 0.001 electrons/bohr surface, of C6F5I .

    5

    Figure 1.03. Molecular electrostatic potential surfaces (B3LYP/6-31+G**- LANLdp, Gaussian 03) of trifluoro-halomethane species: CF4 (left), CF3Cl (middle left), CF3Br (middle right), and CF3I (left).

    6

    Figure 1.04. Geometry of both XB and HB complexes a) NH3 contains lone pairs of electrons and the observed C3v geometries in b) H3N···Br2, c) H3N···BrCl, and d) H3N···HBr complexes.

    7

    Figure 1.05. Nonbonding pair electron density model of H2O and the observed geometries of a) H2O, b) H2O···HF, and c) H2O···ClF. The equilibrium geometry for H2O···HF is shown, for which the angle φ = 46(5)° and For H2O···ClF, the estimated value is φ ≈ 20°, (effective value for the zero-point state).

    8

    Figure 1.06. Geometries of a) Nonbonding pair electron density model of H2S and the observed geometries of b) H2S···HCl (φ = 93.8°), and d) H2S···ClF (φ = 95.8°).

    8

    Figure 1.07. The geometry of H2O···F2 , which is effectively planar in the zero-point state and geometry of H2S···F2 . The angle φ = 113°, and the internal rotation of H2S about its local C2 axis (angle y) gave a low-energy vibrational satellite in the rotational spectrum.

    9

    Figure 1.08. a) π-Bonding electron density model of ethene and ethyne observed geometries (perpendicular or planar, C2v) of a) ethene···HCl b) ethene···ClF c) ethyne···HCl and d) ethyne···ClF.

    9

    Figure 1.09. a) pseudo-π-electron density model of cyclopropane and observed geometries b) cyclopropane···ClF c) cyclopropane···HCl.

    10

    Figure 1.10 Observed geometries of a) benzene···ClF and b) benzene···HCl.

    10

    Figure 1.11: 1D supramolecular network of cocrystal of two layers top: 1,6-diiodoper-fluorohexane with trans-4,4’-azobis (pyridine) and bottom:1,8-diiodoperfluorooctane with trans-4,4’-azobis (pyridine).

    12

    Figure 1.12. 1D chain produced through combination of hydrogen and halogen bond.

    12

    Figure 1.13. Matching and mismatching various I–(CF2)m–I with bis-

  • xiii

    (trimethylammonium) alkane to separate α,ω-diiodoperfluoroalkane from their mixtures.

    13

    Figure 1.14. a) 2D sheet generated through a combination of self-complementary hydrogen bonds and I···I interactions.

    14

    Figure 1.15. The molecular structure of the XB complexes between alkoxystilbazole and 4-substituted iodotetrafluorobenzene.

    15

    Figure 1.16. The molecular structure of the XB complexes between 4-substituted pyridine and 4-substituted iodotetrafluorobenzene.

    16

    Figure 1.17. The bent-core halogen bonding complexs between stilbezole and 1,3-disubstituted perfluorobenzene.

    17

    Figure 1.18: Halogen bonded polymer.

    17

    Figure 1.19. Geometry of type I and II halogen-halogen interactions. Type I: θ1 = θ2; and type II: θ 1 = 90° involving the halogen bond acceptor role, θ 2 = 180° involving the halogen donor role.

    18

    Figure 1.20. Structure of the (2-haloimidazolium)2(TeX2)salt (1.14) and (DIETSe)2FeCl4 (1.15).

    19

    Figure 1.21. Binding mode of covalent inhibitors 1.16 and 1.17 at the active site of hCat L with its three pockets. Pocket S1 binds with covalent bond, ArSO2- group accommodates in pocket pocket S2 contains and the substituent at position 4 of the phenyl ring approaches the C=O group of Gly61 in the S3 pocket.

    22

    Figure 1.22. X-ray structure bound to Inhibitors KH-CB19 (1.18) and CLK3 reveals the halogen interaction with the hinge CO.

    23

    Figure 1.23. The chemical structures of adenosine triphosphate and 5,6-dichlorobenzimi- dazone-1-β-D-ribofuranoside (DRB, 1.19) bound to a Cdk9 hinge region

    24

    Figure 1.24. L-thyroxine, 3,5,3',5'-tetraiodothyronine (1.20) is the precursor of thyroid hormone.

    24

    Figure 1.25. . Equilibrium geometries of CF3X···N(Me)3 complexes (X = Cl, Br,

    or I).

    26

    Figure 1.26. Equilibrium geometries of CF3X···O(Me)2 complexes where ((X = Cl, Br, or I) and (CF3Br)2···O(Me)2

    27

    Figure 1.27. The computed B3LYP/LANLdp electrostatic potential, on the 0.001 electrons/bohr surface of I2.

    28

  • xiv

    Figure 1.28. Hydrogen bond and halogen assisted interaction between heterocyclic amines (2-aminopyridine 1.21, N,N-dimethylaminopyridine 1.22, N,-phenylamino pyridine 1.23, 1H-pyrrolo[2,3-b]pyridine 1.24) and diiodine

    28

    Figure 1.29. The symmetric (a) and asymmetric (b) geometries of the systems bis(pyridine) bromonium and iodonium triflates (1.25, X = Br, for 1.26, X = I and for 1.27, X = H).

    29

    Figure 1.30. Perylene diimide based C-shaped receptors (1.28), S-shaped receptors (1.29) and phenazine (1.30).

    30

    Figure 1.31. Interactions between the nitropyrrole containing bisaniline derivatives (1.31, 1.32, 1.33 and 1.34) with isothaloyl derivative (1.35). (φ = 30 ≈ 37°)

    31

    Figure 1.32. Nickel fluoride complex trans-[NiF(C5F4N)(PEt3)2]·C6F5I (1.36), trans-[NiF(C5F4N) (PEt3)2]·2C6F5I (1.37) and nickel chloride complexes trans-[NiCl(C4F2N)(PEt3)2] (1.38).

    32

    Figure 1.33. Ni, Pd, and Pt fluoride complex

    32

    Figure 2.01. Halogen bonding interactions between iodoalkynes and DMSO.

    48

    Figure 2.02. Chemical shift of C-1 of n-C4H9−C≡C−I (2.03) versus donor number (DN) of the solvent.

    48

    Figure 2.03. Chemical shift of C-1 of n-C4H9−C≡C−I (2.03) with respect to β.

    49

    Figure 2.04. Chemical shift of C-1 of n-C4H9−C≡C−I (2.03) verses solvent soft basicity Bsoft.

    49

    Figure 2.05. The computed B3LYP/6-31+G(d,p) electrostatic potential, on the 0.001 electrons/ bohr surface, of C6F5I. The iodine atom is at the right.

    51

    Figure 2.06. Structures of the substituted iodoperfluoroarenes employed in the substituent effect study.

    52

    Figure 2.07. Molecular electrostatic potential surfaces (B3LYP/6-31+G**-LANLdp; Gaussian ’03) of (left to right) iodopentafluorobenzene, 2.03a and 2.03b.

    54

    Figure 2.08. Correlation of 4-X-C6F4I···Bu3PO halogen bond strength log (Ka) with (a) σ para and (b) σmeta substituent constants of X.

    55

    Figure 2.09. Correlation of 4-X-C6F4I···Bu3PO halogen bond strength log(Ka) with the electrostatic potential density at the iodine atom, calculated with (a) AM1 and

    56

  • xv

    (b) DFT (B3LYP/6-31+G**/LANLdp) computational methods. Figure 2.10. Molecular electrostatic potential surfaces (B3LYP/6-31+G**-LANLdp; Gaussian ’03) of (iodoethynyl)benzene (left); 5-iodouracil.

    57

    Figure 2.11. Experimental binding data (–∆Gbinding) against values according to the model of Hunter and coworkers.

    60

    Figure 2.12. Experimental binding data (–∆Gbinding, Table 2.05) against DFT-calculated (–∆ECPB3LYP, Table 2.06) gas phase binding energy.

    62

    Figure 3.01. Bidentate coordination of nitrates by 2-bromo-1,3-diisopropyl-4,5-dimethyl imidazolium nitrate (3.01), infinite chains formed by: 3,5,3´-tri-iodo-L-thyronine (3.02), chain formation on sulfonate coordination in 4-chloropyridine-3-sulfonic acid (3.03)

    77

    Figure 3.02. Bidentate coordination of Br− with 2,5 dibromoaniline (3.04), Cl− with 2,5 dichloroaniline (3.05).

    78

    Figure 3.03. Heteroditopic receptor (3.09) and (3.09a).

    79

    Figure 3.04. Top: 19F NMR titration of C8F17I (acetone, 295 K; change inchemical shift |∆δ| (in ppm) vs [n-Bu4N+Cl−] (in mm)).

    81

    Figure 3.05. Van’t Hoff plot (ln(Ka) vs 1000/T) for the interaction of C8F17I with Cl− anion in acetone solvent.

    82

    Figure 3.06. Van’t Hoff plot (ln(Ka) vs 1000/T) for the interaction of C6F5I with Cl− anion in acetone solvent.

    83

    Figure 3.07. Hydrogen bonded molecular aggregation among 2-pyridones and dipyridones (3.10a, 3.10b, 3.10c).

    85

    Figure 3.08. Complex formation by propose receptors with halide anions. Complexes are 3.11·X− and 3.12·X−.

    86

    Figure 3.09. Solid-state structures of the halogen-bonded complex of a) 3.11···I− and n-Bu4N+ b) 2 × (3.11···I−).

    88

    Figure 3.10. Bidentate receptors.

    91

    Figure 3.11. Tridentate receptors of mesityle (3.21), amine (3.22) and benzene scaffold (3.23); spatial arrangement of ethyl groups in hexaethylbenzene (3.24).

    93

    Figure 3.12. Left: Calculated structure of the 3.21–Cl− complex (HF/6-31+G**-LANL2DZdp: H white; C gray; O red; F blue; I purple; Cl green). Right: Electrostatic potential surface of 3.21 (chloride-bound conformation, B3LYP/6-

    94

  • xvi

    31G*-LANL2DZdp: blue indicates sites of partial positive charge and red sites of partial negative charge). Figure 3.13. Job plot ([3.21+n-Bu4N+Cl−] = 11 mm, |χ∆δ| vs mole fraction χ).

    94

    Figure 3.14. Cu+2 coordinated hexadentate receptor 3.25.

    95

    Figure 3.15. Van’t Hoff plot (ln(Ka) vs 1000/T) for the interaction of receptor 3.21 with Cl− anion in acetone solvent.

    99

    Figure 3.16. Propose ternary complex employed by both hydrogen and halogen bonding. The donors are the molecules 3.26 and 3.13.

    100

    Figure 3.17. Tridentate receptors (3.21a, 3.21b, and 3.21c) and bidentate receptors syn bromoimidazoliophane (3.29.2PF6−).

    103

    Figure 3.18. Van’t Hoff plot (ln(Ka) vs 1000/T) for the interaction of C6F5I with Cl− anion in acetone solvent.

    119

    Figure 3.19. Van’t Hoff plot (ln(Ka) vs 1000/T) for the interaction of C8F17I with Cl− anion in acetone solvent.

    119

    Figure 3.20. Van’t Hoff plot (ln(Ka) vs 1000/T) for the interaction of receptor 3.21 with Cl− anion in acetone solvent.

    120

    Figure 3.21. Calibration curve of probe temperature (°C) with methanol solvent.

    Figure 4.01. Macrocyclic receptor 4.01·2PF6− having the interlocked nature of catenane.

    130

    Figure 4.02. Imidazolium based receptors 4.03 and 4.04 using anthracene as spacer or scaffold.

    131

    Figure 4.03. Imidazolium based receptors 4.05 and 4.06 using 1,3,5-trimethylbenzene-based scaffold.

    132

  • xvii

    List of Abbreviations

    CDCl3 deuterated chloroform

    B3LYP Becke, three-parameter, Lee-Yang-Parr

    CH2Cl2 methylene chloride

    CH3CN acetonitrile

    DCE 1,2-dichloroethane

    DCM dichloromethane

    DFT density functional theory

    DIC N,N’-diisopropylcarbodiimide

    DMAP 4-N,N-(dimethylamino)pyridine

    DMF N,N-dimethylformamide

    DMSO dimethyl sulfoxide

    EI electron impact

    equiv. equivalents

    ESI electrospray ionization

    Et ethyl

    Et3N triethylamine

    EtOAc ethyl acetate

    FID free induction delay

    g gram

    GC gas chromatography

    HRMS high-resolution mass spectrometry

    Hz Hertz

    IR infrared

    J coupling constant

    LDA lithium diisopropylamide

    LANL2DZ Los Alamos National Laboratory 2-double-z (density functional theory) M molarity

    Me methyl

    MgSO4 magnesium sulfate

    MHz megahertz

  • xviii

    MP2 Møller-Plesset Perturbation theory

    δ chemical shift in parts per million

    m/z mass-to-charge ratio

    NaHCO3 sodium bicarbonate

    Na2SO4 sodium sulfate

    NaHSO4 sodium bisulfate

    NaS2O3 sodium thiosulfate

    NBS N-bromosuccinimide

    NH4Cl ammonium chloride

    NMR nuclear magnetic resonance

    mmol millimole

    mL milliter

    µL microliter

    MS mass spectrometry

    n-Bu n-butyl

    OAc acetate

    Piv pivaloyl

    ppm parts per million

    py pyridine

    Rf perfluoroalkane/ perfluoroarene

    TBAF tetrabutylammonium fluoride

    TBACl tetrabutylammonium chloride

    TBABr tetrabutylammonium bromide

    TBAI tetrabutylammonium iodide

    t-Bu tert-butyl

    THF tetrahydrofuran

    TMS tetramethylsilane

    TMSOTf trimethylsilyl trifluoromethanesulfonate

    TOF time of flight

    UV ultraviolet

    Vis visible

  • 1

    Chapter 1

    Halogen Bonding: A Noncovalent interaction

    1.1. Introduction Halogenated compounds play key roles in polymer chemistry, crystal engineering, medicinal

    chemistry, and environmental science. The interaction between electron deficient halogen

    compounds and electron donors is known as halogen bonding. Halogen bonding has seen

    applications in areas such as crystal design, molecular recognition, topochemical reactions,

    molecular conductors, and liquid crystals; its role in structural biology has also been

    examined. The aims of this review is not to summarize all recent work related to halogen

    bonding, but rather to provide an overview of studies of the interaction in different phases.

    1.2. History of the halogen bonding interaction Although halogen bonding interactions were first observed by Guthrie in 1863,1 the term

    “halogen bond” was not coined until late 1970’s by Dumas.2 In Guthrie’s experiment, iodine

    powder was added to an ammonia solution to obtain an iodoammonium complex. Half a

    century later, the group of Odd Hassel used X-ray crystallography to probe the structure of

    the bromine−dioxane complex3 and the bromine−hexamethylene tetramine complex.4 The

    Br−Br bond length had elongated upon complex formation, changing from 2.28 Å in the free

    molecule (Br2) to 2.31 Å whereas Br···O distance in the complex was 2.71 Å, shorter than the

    sum of van der Waals (VDW) radii (3.35 Å). Similarly, in the Br2-hexamethylenetetramine

    complex, the Br−Br bond length was 2.3 Å and the Br···N distance 2.4 Å. The noncovalent

    O···Br−Br and N···Br−Br angles were found to be linear in both crystals. Hassel showed that

    “halogen molecule bridges” of the dioxane−halogen addition compounds were not only

    confined to halogen compounds formed with ethers and amines but had also been found in

    the 1:1 complex formed by acetone and bromine; the 2:1 ratio in bromine−

    hexamethylenetetramine complex as well. In the case of bromine−acetone complexes, the 1:1

    complex was formed by alternating acetone and bromine end chains, indicating that both

    electron pairs were capable of forming such bonds simultaneously (Figure 1.01).

  • 2

    O

    O

    O

    OBr Br

    N N

    NN

    BrBrO

    H3C CH3

    BrBrBrBr

    Br Br Figure 1.01. Halogen bonding in Br2 −1,4-dioxane, Br2 –hexamethylenetetramine and Br2 –acetone complexes.

    According to Hassel, charge transfer from electron donors (Lewis bases) to electron

    acceptors (iodine or bromine) was responsible for these interactions.5 In 1949 Benesi and

    Hildebrand investigated solutions of iodine in benzene and mesitylene, and reported the

    presence of absorption bands in the ultraviolet region. Their quantitative studies were able to

    determine the equilibrium constants 1.72 M-1 and 1.15 M-1 for the iodine−benzene complex

    in CCl4 and n-heptane solvent respectively, whereas equilibrium constants of the iodine-

    mesitylene complex in CCl4 and n-heptane solvent were 7.2 M-1 and 5.3 M-1, respectively.6

    At the same time, Mulliken worked out a quantum mechanical theory of complex resonance

    that explained halogen bonding as a charge-transfer interaction. The direction of a bond

    linking between a lone pair donor atom to a halogen atom was determined by the orientation

    of the orbitals of non-bonding electron pairs in the donor atom.7 He also proposed a

    mechanism for halogenation of alkenes involving an outer complex where there was little

    charge transfer between components and an inner complex where there was extensive charge

    distribution (Scheme 1.01).8

    Scheme 1.01. Mechanism for the reaction of halogen molecules with alkenes (outer and inner complexes) in polar solvents in the dark.

    C C XY C C

    XY

    C C

    X

    Youter complex inner complex

  • 3

    According to their Nobel lectures in 1966 and 1970, both Mulliken and Hassel respectively

    described interactions involving negative charge transfer from electron donors to electron

    acceptor halogens along with the properties of these systems elicited by X-ray

    crystallography.9 In his work, Dumas defined the “halogen bond” (XB), analogous with the

    hydrogen bond, by determining the association constants of pyridine derivatives with CX4 (X

    = Cl, and Br) and comparing them with those between pyridine derivatives and pyrrole. The

    attractive interaction between a halogen atom X from R–X (e,g. X = I, Cl, and Br), in which

    R is an electron withdrawing group and an atom or a group of atoms A, and the acceptor

    species is termed halogen bonding. Similarities between halogen bonding and hydrogen

    bonding interaction are depicted in Scheme 1.02, where A is a halogen bonding while B and

    C are hydrogen bonding interactions. The details of these interactions will be discussed later.

    D X A D H A D X

    DH

    ··· ··· ···

    A B C Scheme 1.02. Halogen bonding and hydrogen bonding interactions: A) halogen bond donor (D−X) with acceptor A, B) hydrogen bond donor (D−H) with acceptor A, and C) hydrogen bond donor (D−H) with acceptor (D−X).

    In 1979, Bertran and Rodriguez first reported the interaction of halogen bond donors (e.g.,

    chloroform, bromoform and iodoform) with different amines, ketones, ethers and esters by

    measuring the chemical shift (δ) change of the haloformic proton in different solvents

    relative to that in hexane.10 Analyzing the environment around halogen centers in hundreds

    of crystal structures, Parthasarathy et al. in 1986 studied the angular preferences of

    intermolecular forces around halogen centers; nucleophiles approach C−X along the backside

    of the bond whereas electrophiles approach nearly perpendicular to C−X.11 From the late

    1990’s there was a rapid increase in the number of publications involving halogen bonding,

    including applications in the solid state, studies in the gas phase, and computational work.

    The Milan group successfully used halogen bonding in crystal engineering, and to solve

    problems in supramolecular chemistry.12 The Legon group contributed to the research

    advancement by determining the symmetries, radial and angular geometries, as well as

  • 4

    intermolecular force constants Kσ of weakly bound halogen bonded complexes in the gas

    phase.13

    Trends in halogen bond donor ability have been investigated through both theoretical and

    experimental approaches, and the order I > Br > Cl >> F has emerged from many different

    studies in solid, liquid and gas phases, correlated positively with the polarizabilities of the

    halogen atoms and correlated negatively with their electronegativities. The donor abilities of

    these halogen atoms are also influenced by the hybridization of adjacent carbon atoms as the

    order of donor strength increases in the order X−X > X−C (sp) > X−C (sp2) > X−C (sp3).14

    Paradoxically, halogen atoms having partial negative charge in R−X compounds can act as

    electrophiles and accept electron density from Lewis bases.

    The group of Politzer has advanced an explanation for this phenomenon that is distinct from

    the charge-transfer hypothesis discussed above: it relies on a concept named ‘sigma-hole’.

    The σ-hole represents a region of positive electrostatic potential that develops along the

    covalent carbon−halogen bond as a result of anisotropic distribution of electron density

    around the halogen nucleus. As a result, the atomic radius of halogen atom along the covalent

    bond axis is smaller than that along the perpendicular direction.15

  • 5

    δ− δ+

    Figure 1.02. The computed B3LYP/6-31+G(d,p) electrostatic potential, on the 0.001 electrons/bohr surface, of C6F5I . The iodine atom is at the right. Red indicates negative charge density and blue positive charge density The molecular electrostatic potential surface of any ground state may be defined as the three-

    dimensional contour of the electronic density envelope.16,17 The electrostatic potential of any

    ground-state spherically symmetric atom is positive everywhere. On the contrary, when

    atoms are combined to form a molecule, the accompanying rearrangements of electronic

    charge normally produce one or more regions of negative potential. Figure 1.02 depicts the

    molecular electrostatic potential for C6F5I where a positive potential is developed on the

    outer most portion of the iodine surface along the C−I axis. This portion of electrostatic

    potential is represented by blue color and is called the σ-hole. Previously it was also observed

    by Allen group that the effective atomic radius of atom I along the extended C−I bond axis is

    smaller than the radius perpendicular to this axis, a phenomenon known as “polar

    flattening”.18 The σ-hole of a halogen bond donor can create the possibility of an attractive

    electrostatic interaction with a nucleophile to form a halogen bond which is along the C−I

    axis. The size of the σ-hole depends on the hybridization state and electronegativity of the

    halogen atom X (e.g. X = I, Cl, and Br) along with the electron withdrawing ability of the

    group attached with it.

  • 6

    Figure 1.03. Molecular electrostatic potential surfaces (B3LYP/6-31+G**- LANLdp, Gaussian 03) of trifluoro-halomethane species: CF4 (left), CF3Cl (middle left), CF3Br (middle right), and CF3I (left). Red indicates negative charge density and blue positive charge density: the plots have been set to the same color scale so that a visual comparison can be made. One view of bonding in CF3X places two unshared electron pairs for X in two p-orbitals and

    one unshared electron pairs in one s-orbital. The amount of s orbital contribution to the

    hybridization of the halogen atom influences the size of the σ-hole; the s-contribution to the

    bonding orbital of X decreases according to the following sequence F, Cl, Br, and I (25%,

    12%, 9% and 9%, respectively). Using natural bond order (NBO) population analysis,

    Politzer et al. showed that the share of bonding electrons by X in CF3X was 71%, 53%, 49%

    and 46% for F, Cl, Br, and I, respectively.15 The combined effects of hybridization and

    electronegativity make the fluorine atom the poorest halogen bond donor and make the

    iodine atom the best halogen bond donor. Moreover, incorporation of strongly electron

    withdrawing functionalities on the halogen donor increases the size of the σ-hole and thus

    increases the halogen bond donor ability. The electrostatic potential maps of CF4 (left),

    CF3Cl, CF3Br, CF3I and CF4 (right) illustrate trends in the magnitude of the σ-hole as a result

    of hybridization, electron withdrawing effects, and electronegativity (Figure 1.03).

    1.3. Halogen bonding in the gas phase The nature of the halogen bond in the gas phase was studied by Legon et al. using microwave

    rotational spectra of complexes B···XY, where B is a simple Lewis base and XY a dihalogen

    molecule. The radial and angular geometries as well as the strengths of the intermolecular

    binding of complexes B···XY were found closely parallel to those of the corresponding

    hydrogen-bonded systems B···HX. To carry out these measurements, a supersonic jet or

    beam of gas mixture was formed by premixing the components and then immediately

    diluting them with argon followed by expanding from a relatively high pressure vessel

  • 7

    through a small circular nozzle into a vacuum. The weakly bound complexes emerged as

    collisionless complexes that were frozen in their lowest rotational and vibrational energy

    states and were then interrogated by microwave radiation in a high electric field to record

    their rotational spectra. By analyzing the rotational spectra, different spectroscopic constants

    (e.g. moment of inertia Ia, intermolecular stretching force constant k and bond angles θ, φ)

    that contain much information about the complexes, were determined.13 The Legon group

    found that B···XY complexes were best described as Mulliken outer type complexes with

    very small charge redistribution on complex formation.8 On the other hand, almost negligible

    bond strength for B···F2 complex indicated the effective absence of halogen bonding to

    fluorine atom.17

    1.3.1. Complex formation with ammonia, water and hydrogen

    sulfide

    The angular geometries of halogen bonded or hydrogen bonded complexes of the type

    B···XY and B···HY were determined based on the molecular axis of Base B. The molecular

    NH

    HH N

    H

    HH Br Br N

    H

    HH Br Cl N

    H

    HH H Br

    Figure 1.04. Geometries of both XB and HB complexes ( H3N···Br2, H3N···BrCl, and H3N···HBr). axes of bases lie along the axis of a nonbonding (n) electron pair of acceptor B, along the

    local symmetry axis of a π (pseudo-π orbital) or along the axis of one of the n pairs when B

    carries both n- and π-pairs. The complexes H3N···Br2 and H3N···BrCl both have C3v geometry

    as shown in Figure 1.04 and are consistent with the above rules, where the electrophilic

    region of the halogen interacted with the n-pair on N so that XY lay along the C3 axis.19

  • 8

    H FOHH

    Cl FOHH

    OHH

    a b c

    ϕϕ

    Figure 1.05. Nonbonding pair electron density model of H2O and the observed geometries of a) H2O, b) H2O···HF, and c) H2O···ClF. The equilibrium geometry for H2O···HF is shown, for which the angle φ = 46(5)° and For H2O···ClF, the estimated value is φ ≈ 20°, (effective value for the zero-point state). The complexes of H2O with HF and ClF were expected to be pyramidal, but experimental

    evidence was in favor of effectively planar structures, indicating that the energy barrier

    between pyramidal and the planar conformation is low enough that the vibrational wave

    functions can be classified as the C2v point group; even in the zero-point state, the molecule

    tunnels rapidly between the pyramidal conformers. In Figure 1.05, the angle φ for both HB

    and XB complexes is shown.

    S

    H H

    S

    H H

    HCl S

    H H

    ClF

    a b c Figure 1.06. Geometries of a) Nonbonding pair electron density model of H2S and the observed geometries of b) H2S···HCl (φ = 93.8°), and d) H2S···ClF (φ = 95.8°).

    To determine the molecular angular geometry of halogen bonding complexes different

    isotopomers, H232S···35ClF, H232S···37ClF, H234S···35ClF, HD32S···35ClF, D232S···35ClF,

    DH32S···37ClF, and D232S···37ClF, of a complex formed by H2S with ClF was observed by

    pulsed-nozzle, Fourier-transform microwave spectroscopy. From these rotational spectra,

    vibrational satellites and Cl-nuclear quadrupole hyperfine components were obtained. By

    using Hamiltonian H operator, these ground state rotational spectra were analyzed and

    centrifugal distortion constants Djk, along with centrifugal distortion term χk, χD were

    obtained. From the value of Djk, χk, χD.as well as moment of inertia Ia, Ib, and Ic, the angle φ

    and ө were extracted.20 One of the rules for determining the equilibrium angular geometry of

    halogen bonded complex is the axis of the complex lies along the axis of non bonding

    electron paired carried by halogen bond acceptor. In terms of halogen bonding complex

    formation the donor-acceptor interactions are between halogen bond donor and non bonding

  • 9

    (n) lone pair of electron such as H2O···ClF. In case of H2S···ClF complex, the two non

    bonding electron pairs occupy two sp hybrids orbitals. The axes of these two orbitals are

    180° to each other and 90° to the H2S molecular plane. As a result, the axis of complex

    formed with halogen bond donor acceptor interaction is different from that of with H2S···ClF

    complex in Figure 1.06. Paradoxically, the F2 molecule showed halogen bond donor ability

    in the gas phase. The rotational spectrum of H2O···F2 complex indicated weak binding at zero

    point energy level with a planar geometry, whereas a pyramidal geometry was observed for

    the H2S···F2 complex with φ = 113° along with a vibrational satellite, suggesting that a low

    potential energy barrier separates two equivalent pyramidal conformations (Figure 1.07).13, 21 φ

    OH

    HF F S

    H H

    F F

    Figure 1.07. The geometries of H2O···F2, which is effectively planar in the zero-point state, and of H2S···F2. The angle φ = 113°, and the internal rotation of H2S about its local C2 axis (angle Y) gave a low-energy vibrational satellite in the rotational spectrum.

    1.3.2. Complex formation with non-aromatic π–electron donors

    Complexes between hydrogen bond donors and π-pair Lewis bases B showed that the HX

    stayed along the local symmetry axis of a π-pair orbital in the absence of a n-pair orbital.

    Both ethylene and ethyne bind with HX or XY perpendicular to their C2 axis. Angular

    geometries of the complexes of ethylene and ethyne with XY (Cl2, BrCl, ClF) in Figure 1.08

    showed a T shape where Xδ+ of XY interacts with the π-electrons as shown. Two π-orbitals in

    ethyne are perpendicular to each other, giving it a cylindrical shape; the interaction with HX

    or XY (HCl, HF, HBr, Cl2, Br2, ICl) resulted in a planar and T-shaped geometry.

    Figure 1.08. Observed geometries (perpendicular or planar, C2v) of ethene···HCl, ethene···ClF, ethyne···HCl, and ethyne···ClF.

  • 10

    Due to similar behaviors of cyclopropanes and alkenes, Coulson and Moffitt introduced the

    idea of a pseudo-π carbon−carbon bond generated by overlapping a pair of sp3 hybrid orbitals

    on adjacent carbons.22 The symmetry axis of the pseudo-π-orbital coincides with a median of

    the cyclopropane equilateral triangle. Hence the angular geometry of cyclopropane···ClF and

    cyclopropane···HCl lie along the extension of the median.

    H

    HH

    H

    H

    H

    Cl

    F

    H

    HH

    H

    H

    H

    H

    Cl

    H

    H

    H

    H

    H

    H

    a b c

    Figure 1.09. a) Pseudo-π-electron density model of cyclopropane and observed geometries of b) cyclopropane···ClF c) cyclopropane···HCl. In case of the interaction between benzene with ClF or HCl, the aromatic π-orbitals act as the

    halogen bond or hydrogen bond acceptor. The geometry of the benzene···ClF complex is

    shown in Figure 1.10. The Cl end of the ClF molecule approximately circulated the cyclic

    path in the potential energy minimum and sampled the π-electron density on one face of the

    aromatic molecule. This motion is therefore governed by a “Mexican-hat” type of potential

    function. Due to the lighter weight of the hydrogen atom and the absence of vibrational

    satellites associated with the intermolecular modes, the vibrational wave function for

    benzene···HCl complex was either strictly C6v equilibrium symmetry or would have preferred

    a specific carbon−carbon bond. However, in either case, the vibrational wavefunctions will

    have C6v symmetry.23

    H

    Cl

    a b

    φ = 23°

    H H

    HH

    H H

    φ = 14.4°Cl

    F

    H H

    HH

    H H

    Figure 1.10. Observed geometries of a) benzene···ClF and b) benzene···HCl.

  • 11

    In contrast with general perception, fluorine that is bound to a strong electron withdrawing

    group such as N or O may act as a halogen bond donor. By using DFT calculations at the

    B3PW91/6-31G(d,p) level of theory, Metrangolo et al. proposed the presence of a positive σ-

    hole on fluorine atom in CF4, CF3COF, FOCF2OF, HC=CF, O=CF2, FOC(O)F, F3CC(O)OF,

    F3CSO2)2NF, and FSO2CF3.24 This proposal was supported by gas phase studies on the

    H2S···F2 complex, H2O···F2 complex as well as the measurement of gas phase F2 elimination

    from F3− (98.4±10.6 KJmol-1).25 The experimental XB bond strength of F3− is comparable

    with that of the other trihalide anions (99 ± 5, 127 ± 7, and 126 ± 6 kJmol-1 for Cl3−, Br3−, and

    I3−, respectively).

    1.4. Halogen bonding in the solid state Due to advances in electronics and optics, single crystals are becoming increasingly

    important tools in industrial areas and play crucial roles in semiconductors, optical crystals,

    laser crystals, jewelry and watch industries. Crystal engineering is relevant to several

    important applications, including gas storage in metal-organic frameworks and the

    understanding of polymorphism in pharmaceuticals and explosives. Although hydrogen

    bonds and coordination bonds remain the most preferable options for crystal engineering

    strategies, halogen bonds have received increasing attention over the last decade. By

    analyzing crystal structures from the Cambridge Structural Database (CSD), Allen et al.

    found that the intermolecular contacts between carbon-bonded halogen (C−X, X = Cl, Br, or

    I), and electronegative atoms (O or N) were highly directional. They also proposed that the

    attractive nature of the interaction was due to electrostatic interactions although polarization,

    charge transfer, and dispersion contributions may also be important.18 In further studies on

    X-ray crystal structures, scatter plots illustrated that the angles between the covalent and non-

    covalent bonds around the halogen in C−I···N interaction were close to 175°.12 Similar

    behavior was observed for iodide ion interactions C−I···I−, but the C−I···O interaction angle

    showed a larger spread, the mean being ca. 165°. The interaction angle for hypervalent iodine

    showed a bimodal interaction geometry where two major interaction angles (ca. 175° and

    80°) were observed with the same interaction distance.26

  • 12

    1.4.1. Halogen bonding in crystal Engineering

    Halogen bonding was applied for synthesis of 1D, 2D and 3D networks in the crystalline

    state. Linear and zig zag chains were assembled with the use of bidentate 4,4’-bipyridine and

    different (1,2- or 1,3- or 1,4-) dibromotetrafluorobenzenes.27 Similarly, 1,6-diiodoper-

    fluorohexane and 1,8-diiodoperfluorooctane (electron-acceptor tectons) were assembled with

    the XB acceptor trans-4,4’-azobis(pyridine) to form halogen bonded 1D cocrystals as shown

    in Figure 1.11.28

    I

    I

    FF

    F

    FF

    FF

    FF

    FFF

    I

    I

    FF

    F

    FF

    FF

    FF

    FFF

    NN

    NN

    I

    FF

    F

    FF

    FF

    FF

    FFF

    I

    FF

    F

    FF

    FF

    FF

    FFF

    NN

    NN

    I

    F F

    F F

    I

    FF

    F F

    Figure 1.11. 1D supramolecular crystalline networks Top: 1,6-diiodoperfluorohexane with trans-4,4’-azobis(pyridine) and bottom:1,8-diiodoperfluorooctane with trans-4,4’-azobis(pyridine).

    Akeröy et al. have demonstrated the simultaneous action of both hydrogen bonds and

    halogen bonds for construction of cocrystal networks. Mixing hydrogen bond donor 2-

    aminopyrazine and 4-iodo-2,3,5,6-tetrafluorobenzoic acid gave a cocrystal in which the Py···I

    bond directed the assembly of the cocrystal and the amine−acid dimeric synthon was

    responsible for propagating the inherent geometry in the resulting supramolecules into

    infinite 1D chains.29

    O

    O

    F F

    I

    FF

    NN

    NH

    H

    HO

    OF F

    I

    FF

    H

    NN

    NH

    H

    Figure 1.12. 1D chain produced through a combination of hydrogen and halogen bonding.

  • 13

    Recently, halogen bonded cocrystal 1D networks have been used for resolving mixtures of

    α,ω-diiodoperfluoalkanes, and for isolation and characterization of both stable and unstable

    polyanions. Mixtures of α,ω-diiodoperfluoalkanes were synthesized by telomerization

    reaction of tetrafluoroethylene with iodine and purified by fractional distillation which left a

    mixture of heavier inseparable diiodoperfluoalkanes. Treating these inseparable

    diiodoperfluoalkanes I–(CF2)m–I, where m = 2, 4, 6, 8, 10 or 12 with a bis-

    (trimethylammonium) alkane diiodide [I−(CH3)3N+–(CH2)m+6–N+(CH3)3I−] formed

    supramolecular anions I−···I(CF2)mI···I−. The dications formed a crystal framework with the

    suitable superanions. The CH2 groups of the dications adopted the all-trans conformation,

    and the carbon chains were parallel and interdigitated, thereby forming the required crystal

    framework. When there are (m +6) CH2 and m CF2 groups in the bis-(trimethylammonium)

    alkane diiodide and the α,ω-diiodoperfluoroalkane respectively, there is a match between the

    length of the anion I−···I(CF2)mI···I− formed and the size of the cavities in the dication

    framework. When this rule is not maintained, there is a mismatch, the crystal packing is

    different and there is no cavity to trap the α,ω-diiodoperfluoroalkane. Based on matching and

    mismatching, various I–(CF2)m–I were separated from their mixtures. This was achieved by

    mixing solutions of I–(CF2)m–I (m = 2, 4, 6 and 8) with a solution of a given bis-

    (trimethylammonium) alkane diiodide (say, m = 4). Only the size-matched (m = 4) adduct

    crystallized from the solution and pure (m = 4) α,ω-diiodoperfluoroalkane was then readily

    recovered from the solid by raising the temperature (Figure 1.13).30

    N Nm

    I- I-(CF2)I I

    N Nm

    -I I-(CF2)I Im m

    m = 2, 4, 6, 8, 10, and 12 Figure 1.13. Matching and mismatching various I–(CF2)m–I with bis-(trimethylammonium) alkane to separate α,ω-diiodoperfluoroalkane from their mixtures.

    Bis(trimethylammonium)hexane diiodide was also used to encapsulate I2 from solution

    whereas bipyridinium-derived cations were used to encapsule I3− and I4−2. 31 In case of

    bipyridinium-derived cations, the synergistic effect of size and space matching were used as

    an effective strategy to develop a general and reliable protocol where the crystal

    stoichiometry is independent of the stoichiometry in the crystallization solution.32

  • 14

    Generally, construction of two dimensional 2D networks involves the use of tridentate

    halogen bond donors or XB acceptor tectons. The ability of halogen atoms to act as either

    donors or acceptors can also be used to generate 2D networks. For example using the

    bifurcated bonding ability of XB donor atom, a 2D sheet was created from 2,3,5,6-

    tetrafluoro-4-iodobenzaldehyde oxime (Figure 1.14).33

    N

    H

    HO

    FF

    F F

    IN

    H

    OH

    F F

    FF

    IN

    H

    HO

    F F

    FF

    I

    N

    H

    HO

    F F

    FF

    I

    Figure 1.14. A 2D sheet generated through a combination of self-complementary hydrogen bonds and I···I interactions.

    1.4.2. Liquid crystals

    Although most liquid crystals are covalent materials, non-covalent interactions such as

    hydrogen bonding, quadrupolar and charge transfer interactions, have been used to generate

    mesogenic species.34 Due to similarities with hydrogen bonding, halogen bonding has also

    been explored for the synthesis of liquid crystal materials. The liquid crystal state can be

    reached from the solid state either by the action of temperature (thermotropic liquid crystals)

    or solvent (lyotropic liquid crystals). The halogen bonding contribution to thermotropic

    liquid crystals can be divided into two portions: a) halogen bonded low molar mass liquid

    crystals b) supramolecular liquid-crystalline polymers. The first examples of low molar mass

    halogen bonded liquid crystals were obtained by self-assembly; the evaporation of equimolar

    solutions of iodopentafluorobenzene and 4-alkoxystilbazoles ((E)-4-(4-alkoxystyryl)

    pyridines) afforded the halogen bonded dimers 1.01 (Figure 1.15). This complex, formed by

    I of the iodopentafluorobenzene with the N of the alkoxystilbazole, was liquid crystalline in

    nature even though each of the components was non-mesomorphic. X-Ray diffraction of the

    complex of iodopentafluorobenzene and n-octyloxystilbazole confirmed the absence of

    quadrupolar phenyl and perfluorophenyl interactions, and the presence of halogen bond was

    clearly evident with an observed N···I distance of 2.811 Å and N···I−C angle of 168.4°.

  • 15

    Comparison with the transition temperatures of hydrogen-bonded systems (phenol or benzoic

    acid derivatives) suggested that halogen bond and hydrogen bond have effects of similar

    magnitudes in the fluid, liquid crystal phase.35

    .

    H2m+1Cm I F

    FF

    F F

    I

    FF

    F FOCnH2n+1

    1.01

    I

    FF

    F F

    OCnH2n+1

    1.02

    1.03

    NO

    H2m+1Cm NO

    H2m+1Cm NO

    H2m+1Cm NO

    m = 4, 6, 8, 10, 12n = 4, 6, 8, 10, 12

    FF

    F F

    OI

    OOCnH2n+1

    1.04

    H2m+1Cm NO

    FF

    F F

    IO OCnH2n+1

    1.05

    H2m+1Cm NO

    FF

    F F

    OI

    OOCnH2n+1

    F F

    FF

    1.06

    O

    NO = NO

    Figure 1.15. The molecular structure of the XB complexes between alkoxystilbazole and 4-substituted iodotetrafluorobenzene. Extensions of this work were reported by Bruce et al. who prepared new complexes from

    alkoxystilbazole derivatives, 4-substituted pyridine derivatives and 4-substituted

    iodotetrafluorobenzene derivatives. In supramolecular dimer 1.02 (m = 10, n = 6), all four

    phenyl rings are coplanar and the alkyl chains also adopt a linear trans arrangement coplanar

    with the phenyl rings. In the crystal structure, two different dimers of 1.02 (m = 10, n = 6)

    were held together in an antiparallel fashion in the plane of the aromatic rings by

    intermolecular H···F contacts (2.49 Ǻ). In the case of tetramer 1.03, the mean square plane of

  • 16

    the stilbazole rings rotates with respect to the XB donor ring by 31.1° and 13.3°, but in case

    of 1.02 the angle was 8.6°. The alkyl chains in 1.03 were not coplanar with the aromatic

    rings. However, transition temperatures of these complexes primarily depended on the

    stilbazole chain length, and on the alkyl chain length of the iodobenzene to a lesser extent.

    For complexes between octylpyridine and esters 1.07-1.10 (Figure 1.16), the transition

    temperatures were significantly lower than expected in spite of the presence of three aromatic

    rings.36

    n = 4, 6, 8, 10, 12

    FF

    F F

    OI

    OOCnH2n+1

    1.07

    NC8H17

    FF

    F F

    IOCnH2n+1

    NC8H17

    FF

    F F

    OI

    OOCnH2n+1

    F F

    FF

    NC8H17

    1.09

    1.10

    1.08

    FF

    F F

    IO OCnH2n+1

    NC8H17O

    Figure 1.16. The molecular structure of the XB complexes between 4-substituted pyridine and 4-substituted iodotetrafluorobenzene.

    The telechelic donors 1,3-diiodoperfluorobenzene or para-iodoperfluorophenol self-

    assembled with the stilbazole to give rise to bent core systems with 2:1 stoichiometry. Once

    again, both starting materials were nonmesogenic, but 1.11 and 1.12 complexes showed

    monotropic liquid crystalline phases on cooling (Figure 1.17).37

    .

  • 17

    CmH2m+1

    O

    FF

    I

    FF H

    NO

    N

    O

    H2m+1Cm

    F

    I

    FF

    F

    1.121.11

    NO = NO

    H2m+1Cm

    N

    O

    I

    CmH2m+1

    N

    O

    Figure 1.17. The bent-core halogen bonding complexs between stilbazole and 1,3-disubstituted perfluorobenzene.

    A layer-by-layer polymer 1.13 was assembled on the basis of halogen bonding between

    poly(4-(4-iodo-2,3,5,6-tetrafluorophenoxy)-butyl acrylate) and poly(4-vinylpyridine) shown

    in Figure 1.18. This multilayer water insoluble polymer was synthesized in methanol.38

    OO

    O

    In

    FF

    FF

    Nn

    1.13 Figure 1.18. Halogen bonded polymer.

    1.4.3. Metal-organic frameworks

    Due to their redox, magnetic, and optical properties, metal ions can play an interesting role in

    the area of supramolecular chemistry. The semi-occupied d and f orbitals may give rise to

    such properties as strong absorption, high quantum yields, luminescence, tunable redox

    states, and suitable excited state lifetimes. The splitting of d orbitals, in response to ligand

    field strength, can influence spin crossover phenomena along with physical properties.39 In

    organometallic networks, M−C bonds are longer than C−C bonds, leading to increases in the

    sizes of cavities and channel sizes of the crystal networks. Metal-organic compounds that

    contain halogen−halogen interactions are divided into two groups: type I, where 140° < θ1 =

  • 18

    θ2 < 160° and type II where θ1 is close to 180° and θ2 is close to 90°. The later type is halogen

    bonding.40

    C X

    X CC X X

    C

    Type I Type II

    θ1

    θ1

    θ2

    θ2

    Figure 1.19. Geometries of type I and II halogen−halogen interactions. Type I: θ1 = θ2; and type II: θ 1 = 90° involving the halogen bond acceptor role, θ 2 = 180° involving the halogen donor role.

    Halogenated and positive charged heterocyclic systems such as halopyridinium derivatives

    have been commonly used for maximizing halogen bond donor ability in complex

    formation.41 In M−X···X−C interactions involving halopyridinium derivatives, the halogen

    bonding order is I > Br > Cl and the bond angles are close to 90° and 180° for θ1 and θ2

    respectively. Complexes trans-[MCl2(3-X-py)2] were synthesized by the reaction of ML4-2

    (M = Pd, Pt) with 3-X-pyridine (X = Cl, Br, I) in acidic aqueous solution. After

    crystallization, M−X···X−C interactions were observed in the crystal. Similarly,

    halopyridinium derivatives also formed complexes with Cu, Au, Co, Pd, Ni and Ru salts by

    halogen bonding interactions. Halogen substituted imidazolium, imidazolinium,

    benzodithiazolium, and anilium derivatives also showed halogen bond donor properties.41

    The design of conducting materials was achieved relying on interactions of

    diiodo(ethylenedithio)tetraselenafulvalene (DIETSe) (donor molecules) with anionic

    acceptors. Here stronger π-d exchange was proposed to occur when intermolecular

    halogen···halogen contacts were present, promoting exchange interactions between the

    conducting π-electrons of the donors and the localized d-electrons of the anions.

    (DIETSe)2FeCl4 showed a metal–semiconductor transition at 11 K (Figure 1.20).42

  • 19

    N N

    X

    NN

    X

    TeX

    XX

    X

    X

    X

    S

    S

    Se

    Se

    Se

    Se I

    I

    FeCl Cl

    Cl Cl

    S

    S

    Se

    Se

    Se

    Se I

    I

    S

    S

    Se

    Se

    Se

    SeI

    I

    S

    S

    Se

    Se

    Se

    SeI

    I

    1.14 1.15 Figure 1.20. Structure of the (2-haloimidazolium)2(TeX2)salt (1.14) and (DIETSe)2FeCl4 (1.15).

    1.4.4. Nanomaterials

    Shirman et al. have employed halogen bonding as a tool for stepwise assembly of gold

    nanoparticles onto a planar surface.43 Previously they demonstrated the formation of highly

    crystalline halogen bonded thin films by vapor deposition on silicon substrates, where the

    solid state structure was independent of the substrate surface chemical functionalization.44

    They also demonstrated the supramolecular assembly of gold nanoparticles (AuNPs) by

    activating with an XB donor ligand. To form the AuNP-1, they substituted

    tetraoctylammonium bromide (TOAB) from AuNP−TOAB with the polar N-oxide moiety of

    (E)-4 (2,4,6-trifluoro-3,5-diiodostyryl)-pyridine-1-oxide (FIPO). The AuNP- FIPO particle

    was formed by coordination of the polar N-oxide moiety of FIPO to the gold surface and

    addition of a bifunctional halogen bond linker (BPEB) to the solvent mixture. The final

    aggregates AuNP−FIPO /BPEB which were formed could be controlled by time during the

    first assembly steps and by varying the concentration of the linker (BPEB). The aggregation

    process (scheme 1.03) was monitored by UV/Vis spectroscopy and TEM.45

  • 20

    Scheme1.03. The formation of assembly consisting of functionalized AuNP with organic cross-linkers (BPEB) in solution. Reprinted with permission from reference 42. Copyright 2011, American Chemical Society.

    The gold nanoparticles (AuNPs) were assembled on silicon and quartz substrates which were

    functionalized by p-chloro benzyl derivatives and reacted with (E)-1,2-di(pyridin-4-yl)ethane

    or (E)-4-(2-(pyridin-4-yl)vinyl)phenol to construct M1 or M2. The reactions of AuNP-1 with

    M1/ M2 in presence of BPEB assembled AuPN-1 on M1/ M2 planar surfaces. The progress

    of the assembly of M1-AuNP-1- BPEBB could be monitored by color change, from red to

    deep blue after 10 AuNP-1 deposition steps (scheme 1.04).

    N

    N

    =

    Scheme 1.04. Stepwise generation of assemblies consisting of functionalized gold nanoparticles (AuNP-1) and organic cross-linkers (BPEB) on silicon organic monolayers (M1, M2). Adapted and reprinted with permission from reference 42. Copyright 2011, American Chemical Society.

  • 21

    1.5. Halogen bonding in Medicinal chemistry Although halogens are not generally prevalent atoms in biological molecules, it is clear that

    they can play important roles in some biological systems. In medicinal chemistry, one of the

    common strategies to develop new drugs is insertion of halogen atom /atoms on hit or lead

    compounds to exploit their steric effect in the binding sites of molecular targets having

    deeper pockets. Other than steric effects, halogen atom incorporation may increase

    membrane permeability and/or blood-brain barrier (BBB) permeability or block sites of

    metabolism.46 Roughly one fourth of total number of papers and patents related to medicinal

    chemistry involve halogen atom insertion to the final compounds as per Scifinder, Web of

    Science, Science Direct and ACS. The majority of the halogenated drugs are fluorine

    containing drugs followed by chlorine, bromine and iodine. In structure-activity relationship

    (SAR) discussions the effect of halogenation is largely confined to steric effects or changes

    in lipophilicity. The role of halogen bonds in protein-ligand complexes was pointed out by

    Auffinger et al. (2004) who studied halogen bonding interactions between halogenated

    ligands and cyclin-dependent kinase 2 (CDK 2).47

    Recently, Hadegger et al. studied halogen bonding in human Cathepsin L (hCat L)

    inhibitors.48 The nitrile group was responsible for binding with the hCat L in S1 pocket,

    forming a covalent thioimidate adduct while ArSO2- interacts with S2 by lipophilic

    interaction and the other aryl group interacts with the oxyanion hole pocket S3 through a

    halogen bonding interaction involving p-X−Ar (X = Cl, Br, I, and CF3) (Figure 1.21).

    Experimental results were consistent with a halogen bonding interaction between Gly61 in

    pocket S3 with the inhibitors 1.16 (X = Cl, Br, I, and CF3). Methyl substitution at the 4 X-

    aryl position, which occupied the S3 receptor site, also showed moderate activity probably

    due to the flexibility and reorganization of S3. X- ray crystal structures showed the O···Cl−C

    angle was 174° (close to 180°) and the O···Cl distance was 3.10Å consistent with a halogen

    bonding definition. The free energy gain with this halogen bond was as much as ∆∆G = −2.6

    kcal mol-1 from X = H to X = Cl (van der Waals radii of Cl and CH3 are1.75 and 2.00 Å

    respectively).

  • 22

    N

    S

    Cl

    O

    X

    ONHN

    OO

    N

    S

    Cl

    O

    X

    ONHN

    OOO

    F3C

    SH

    HN

    O

    HN

    OCys25

    gly61

    S3

    S1

    S2

    1.16

    1.17

    Figure 1.21. Binding mode of covalent inhibitors 1.16 and 1.17 at the active site of hCat L with its three pockets.

    Table 1.01. Activity data for hCatL inhibitors.

    Inhibitor

    IC50 Log D X Inhibitor IC50 Log D

    1.16a 0.29 2.11 H 1.17a 0.59 n.d

    1.16b 0.13 2.57 Me 1.17b 0.81 n.d

    1.16c 0.34 2.36 F 1.17c 0.93 n.d

    1.16d 0.022 2.73 Cl 1.17d 0.036 2.19

    1.16e 0.012 2.96 Br 1.17e 0.0081 >3.0

    1.16f 0.0065 3.23 I 1.17f 0.020 n.d

    1.16g 0.095 3.12 CF3 1.17g 0.22 n.d

    Fedorov et al. (2011) introduced dichloroindoyl enaminonitrile KH-CB19, a potent and

    highly specific inhibitor of the CDC2-like kinase isoforms 1 and 4 (CLK1/CLK4). The ATP

    binding site in the protein kinase contains a hinge connection between the two protein lobes

    and forms H-bonds to bound ATP through NH and CO, leaving a hinge CO as an available

    accepting site. The X-ray crystal structure of KH-CB19 with CLK1 and CLK3 revealed a

    non-ATP-mimetic binding mode involving conformational changes in helix αC and the

  • 23

    phosphate binding loop and halogen bonding to the kinase hinge region. KHCB19 effectively

    suppressed phosphorylation of SR (serine/arginine) proteins in cells (Figure 22). 49, 50

    NH3C

    Cl

    Cl

    O

    O

    CN

    H2N

    18

    Figure 1.22. X-ray cocrystal structure bound to inhibitors KH-CB19 (1.18) and CLK3. Reprinted with permission from reference 49. Copyright 2011, Elsevier.

    As protein kinase Cdk9 inhibition contributes to anticancer activity, inhibitors have been

    under clinical investigation, including 5,6-dichlorobenzimidazone-1-β-D-ribofuranoside

    (DRB), studied by Baumli et al. (2010). DRB, an adenosine analog, binds to the ATP binding

    site via halogen bonds from the chlorine to the CO in the hinge region that connects the N-

    and C-terminal kinase lobes. It also induced conformational changes in the glycine-rich loop

    and the β3-αC loop that enclose the inhibitor as shown in Figure 1.23.51

  • 24

    NN

    N

    N

    O

    HOHO

    O

    N

    PO

    O

    P OO O

    O

    NH

    OHN

    OHN

    OHN

    HHPhe105

    Asp104

    Cys106

    NH

    OHN

    ONH

    OHN

    Phe105

    Asp104

    Cys106

    N

    N

    OHO

    HOOH

    Cl

    Cl

    ATPDRB (1.19)

    Figure 1.23. The chemical structures of adenosine triphosphate and 5,6-dichlorobenzimi- dazone-1-β-D-ribofuranoside (DRB, 1.19) bound to a Cdk9 Hinge Region An important iodinated compound in human physiology is L-thyroxine, 3,5,3',5'-tetra-

    iodothyronine (1.20), present in thyroid hormones. Halogen bonding is thought to play a role

    in thyroxine binding to its transport protein transthyretin, and this is also supported by short

    I···O contacts that result when bound. Thyroxine binding to RNA sequences has also been

    attributed to halogen bonding, and the T3 (thyroxine having three iodine atoms) and T4

    (thyroxine having four iodine atoms) forms have been identified as the active forms of the

    protein.27

    O

    OH

    O

    NH2

    I

    II

    IHO

    1.20 Figure 1.24. L-thyroxine, 3,5,3',5'-tetraiodothyronine (1.20) is the precursor of thyroid

    hormone.

    1.6. Halogen bonding in the solution phase In spite of enormous success in explorations of XB in computational studies, the gas phase,

    metal organic frameworks, and polymer chemistry, there were limited studies of halogen

    bonding in solution before 2010. The first such work was done by Hildebrand, Glascock

    (1909), and Benesi, Hildebrand (1949) to determine the binding constant of the iodine-

  • 25

    benzene complex in CCl4 and n-heptane solvent as well as the iodine-mesitylene complex in

    CCl4 and n-heptane solvent.6 Mulliken et al. (1959) studied the formation of iodine

    complexes with benzene and pyridine in heptane followed by binding constant determination

    using both UV and IR spectroscopy. The binding constant of Py·I2 at 26 °C was 200 M-1

    which roughly two orders of magnitude stronger than that of the benzene·I2 complex. The

    observed frequencies for Py·I2 and Bn·I2, 180 cm-1 and 205 cm-1 respectively, were slightly

    lower than that of I2 vapor (213 cm-1) because of partial loosening of intrahalogen bond as

    well as an increase in effective mass due to partial formation of a donor-halogen bond.52 The

    binding constants of pyridine and pyridine derivatives (4-MePy, 2,6-Me2Py, and 2,4,6-

    Me3Py) were determined in CCl4 and CBr4 and gave significantly lower values. Similar

    studies with I2 and pyridine derivatives were helpful for understanding steric and substituent

    effects on halogen bonding. In the same study, THF showed much weaker binding strength

    as shown in Table 1.02.53

    Table 1.02. Binding constants (M-1) of complexes between pyridine derivatives and THF with halogen bond donors (CCl4, CBr4, and I2).

    CCl4 CBr4 I2

    Py 0.14 ± 0.04 0.83 ± 0.16 117 ± 10

    4-MePy, , and 0.32 ± 0.07 1.2± 0.22 260 ± 10

    2,6-Me2Py 0.10 ± 0.02 0.41 ± 0.08 74 ± 7

    2,4,6-Me3Py 0.09 ± 0.02 0.64 ± 0.18 125 ± 10

    THF 0.08 ± 0.05 0.46 ± 0.10 1.9 ± 0.2

    To understand the halogen bonding interactions of haloforms with different amines, ketones,

    and esters, NMR spectroscopy was used based on the chemical shift (δ) change with respect

    to the same haloform proton chemical shift in hexane. On the other hand, chemical shift (∆δ)

    changes were poorly correlated for anisotropic solvents e.g benzene, pyridine derivatives,

    and amides.10 The 1:1 complex formed between trifluoroiodomethane and 2,4,6-trimethyl-

    pyridine in cyclopentane solution was studied, and the enthalpy and entropy of formation

    were −20.9 × 0.4 kJ mol-1 and −52.7 × 1.6 JK-1mol-1, respectively.54 Laurence and co-

    workers reported that the association constant of 1-cyano-2-iodoacetylene with Ph3PO was

    higher in CCl4 than in benzene. Kochi and co-worker also studied solvent effects on halogen

  • 26

    bonding in the context of the quinuclidine···CBr4 and CBr4···DABCO complexes. The

    association constants for CBr4···DABCO decreased as a function of solvent, in the order

    CH3CN > CHCl3 > CH3OH. Experimental investigation of n-C3F7I···toluene and 2-

    C3F7I···toluene complexes determined the enthalpies and entropies −2.9 kJmol-1 and −19.1

    JK-1mol-1 for 2-C3F7I···toluene-d8 complex and −2.7 kJmol-1 and −16.0 JK-1mol-1 1-

    C3F7I···toluene-d8 complex. These were correlated with the predicted results.55

    Cl

    FFF

    173.9ο Br/I

    FFF

    180ο

    NH

    H

    H

    H HHH

    H

    HN

    HH

    H

    H HHH

    H

    H

    Figure 1.25. Equilibrium geometries of CF3X···N(Me)3 complexes (X = Cl, Br, or I)

    The halogen bonding complexes between trimethylamine and trifluorohalomethane (CF3I,

    CF3Br, and CF3Cl) were investigated using FT-IR spectroscopy. Using MP2/6-311++G(d,p)

    +LanL2DZ basis set, the complexes of CF3X (X = Cl, Br, or I) with NMe3 showed C−X···N

    (180°) with C3v symmetry for CF3I···NMe3 and CF3Br···NMe3 complexes but Cs symmetry

    was observed for CF3Cl···NMe3 complex (angle C−Cl···N = 173.9°). The tilt of the

    CF3Cl···NMe3 complex was caused by additional stabilization of van dar Waals contact

    between one of the fluorine atoms and a hydrogen atom of one of the methyl groups in

    trimethylamine. The CF3 stretching frequencies of CF3I···NMe3, CF3Br···NMe3, and

    CF3Cl···NMe3 complexes in xenon, kryton and argon respectively were compared with the

    spectra of respective monomers and assigned to 1:1 complexes. Van’t Hoff plots of these

    complexes gave complexation enthalpies of −28.7(1) kJ mol-1for CF3I···NMe3, −18.3(1) kJ

    mol-1 for CF3Br···NMe3, and −8.9(2) kJ mol-1for CF3Cl···NMe3.56 On the other hand, the

    weakly bound molecular complexes between Me2O and CF3I, CF3Br, and CF3Cl in liquid

    argon and krypton were investigated, using both IR and Raman spectroscopy. Ab initio

    calculated geometries of CF3I···OMe2, CF3Br···OMe2, and CF3Cl···OMe2 complexes showed

    that the bond angles for O···X−C were 180 ± 0.5° in all cases. The experimental standard

    complexation enthalpies for 1:1 complexes were derived using the Van’t Hoff equation and

    found to be −6.8(3) kJ mol-1 for Me2O···CF3Cl, −10.2(1) kJ mol-1 for Me2O···BrCF3, and

  • 27

    −15.5(1) kJ mol-1for Me2O···ICF3. At higher concentrations and low temperatures for the

    Me2O / CF3Br mixture, the presence of an additional band near 916.8 cm-1 indicated the 1:2

    Me2O:(CF3Br)2 complex, whereas the band at 923.8 cm-1 corresponded to the 1:1 complex

    between Me2O / CF3Br at low concentration. The complexation enthalpy of the 1:2 complex

    is −17.8(5) kJ mol-1, which is slightly less than twice the experimental enthalpy −10.2(1) kJ

    mol-1 for the corresponding 1:1 complex (Figure 1.26).

    X

    FFF

    H

    HH

    O

    H

    HHO

    HHH

    H HH

    Br F

    FF

    Br

    F

    FF

    Figure 1.26. Equilibrium geometries of CF3X···O(Me)2 complexes where ((X = Cl, Br, or I) and (CF3Br)2···O(Me)2

    One of the most fascinating uses of halogen bonding in solution phase is Lewis acid-base

    scales (pKBI2) introduced by Laurence group.57 Unlike Bronsted acid and bases, there is no

    single reference Lewis acidity or basicity scale. Existing Lewis acid-base scales were

    constructed against a particular acid such as BF3, 4-FC6H4OH, I2, Li+, Na+, K+, Al+, Mn+,

    CpNi+, and CH3NH3+. These scales are based on dative bond, hydrogen bond, halogen bond,

    and cation-molecule interactions.58 The new diiodine basicity scale pKBI2 is quasi-orthogonal

    to the hydrogen bond, dative bond, and cation bonds scale. This scale covers an extended free

    energy range of 9 kcalmol-1 and the order of diiodine basicity of bases followed the sequence

    R2NH > R2PH ≈ R2Se > R2S > RI ≈ R2O > RBr > RCl > RF. According to the HSAB (Hard-

    Soft Acid-Base) theory, I2 is a soft acid. The diiodine basicity of halogen, chalcogen and

    pnictogen bases follow the sequences, I > Br > Cl > F, and Se > S > O, and N > P

    respectively. Contradictory to predictions based on HSAB, diiodine preferred to bind to

    amines than phosphanes. The diiodine basicities of oxygen functionalities occur in the

    following order NO ≈ AsO > SeO > PO > SO > C=O > -O- > SO2. In these series, the

    chalcogen and pnictogen oxides basicity appeared to be correlated with the electrostatic

    potential on oxygen atom. The diiodine basicity order also found in sulfur and nitrogen bases

    and these are PS >> -S- ~ C=S >> N=C=S and amine > pyridine > nitrile. The high accepts

    ability of the PS moiety can be explained in terms of the zwitterionic contribution of P+−S-,

  • 28

    and the order of nitrogen bases can be explained by the decreasing p character of the lone

    pair.

    Figure 1.27. The computed B3LYP/LANLdp electrostatic potential, on the 0.001 electrons/bohr surface of I2.

    The electrostatic potential surface of I2 reveals the σ-hole as well as a belt of negative

    electrostatic potential surrounding the σ-hole as shown in Figure 1.27. Exploiting the σ-hole

    of diiodine as a halogen bond donor and the surrounding belt of negative charge as a

    hydrogen bond acceptor, pKBI2 was determined for heterocyclic amines. The value of pKBI2

    of 2-aminopyridine (2.45) is higher than the pKBI2 of N,N-dimethylaminopyridine (0.95). The

    difference of 1.50 pKBI2 units is due partially to the steric effect of the NMe2 group compared

    to NH2 group and partially by the existence of an intramolecular NH···I hydrogen bond,

    assisting to the formation of N···I halogen bond. Though the diiodine scale is orthogonal to

    the manganese cation, lithium cation, aluminium cation, trimethylboron, and 4-fluorophenol

    basicity scales, it correlated significantly with cyclopentadienylnickel cation and copper

    cation basicity scales which can be explained due to the soft Lewis acidic character of I2,

    CpNi+, and Cu+.57

    N

    NH H

    II

    N

    NH3C CH3

    II

    N

    NH

    N

    NPh H

    II

    II

    1.21 1.22 1.23 1.24pKBI2 = 2.45 pKBI2 = 2.08pKBI2 = 0.95 pKBI2 = 2.10

    Figure 1.28. Hydrogen bond and halogen assisted interaction between heterocyclic amines (2-aminopyridine 1.21, N,N-dimethylaminopyridine 1.22, N,-phenylamino pyridine 1.23, 1H-pyrrolo[2,3-b]pyridine 1.24) and diiodine.

  • 29

    Recently, Erdélyi et al. (2012) successfully applied an NMR spectroscopic method called

    isotopic perturbation of equilibrium (IPE) to probe the symmetry of bis(pyridine) halonium

    halogen bonded systems (Figure 1.29). Perrin had previously shown that N···H···N-type HBs

    was asymmetric in solution, but symmetric in crystals. Analogous iodous 1.26 and bromous

    1.25 halogen bonded (XB) systems preferred symmetric arrangements in solution whereas

    1.26 preferred symmetric but 1.25 asymmetric arrangements in the solid state.59

    N

    X

    N

    N

    XN

    a b

    NX

    N

    b

    (D)H (D)H (D)H

    Figure 1.29. The symmetric (a) and asymmetric (b) geometries of the systems bis(pyridine) bromonium and iodonium triflates (1.25, X = Br, for 1.26, X = I and for 1.27, X = H).

    1.6.1. Halogen bonding in supramolecular chemistry

    The polyaromatic perylene diimide core of C- and S-shaped receptor clefts was designed to

    probe interactions between phenazine (1.30) nitrogens and halogens. Using UV/vis

    spectroscopy, the determined binding constants of halogen-containing clefts receptor 1.28a,

    1.28b and 1.28c with phenazine (1.30) were 480 ± 50, 1100 ± 100 and 2200 ± 200 M-1

    respectively in benzene solution whereas the binding constants of receptors 1.29a, 1.29b and

    1.29c were 780 ± 80, 900 ± 90 and 760 ± 80 M-1. The cleft-like shapes were well suited for

    the investigation of weak intermolecular forces because they provide convergent functional

    groups: the acid for hydrogen bonding, π-stacking, and halogen bonding interaction with the

    substrate held within. The energy differences were 0.5 and 0.9 kcal/ mol for nitrogen-

    chlorine (1.28b and 1.28a) and nitrogen−bromine interactions (1.28c and 1.28a). The energy

    differences were 0.3, 0.1, 0.6 kcal/ mol for among the C- and S-shaped (1.28a and 1.29a),

    (1.28b and 1.29b) (1.28c and 1.29c) receptors, respectively.60

  • 30

    N

    N

    N N

    O

    OO

    O

    N N

    O

    OO

    O

    OO

    O

    O

    XO

    HO

    X

    O

    HO

    a = H, b = Cl, C = Br

    C - shape

    S - shape

    1.28

    1.29

    1.30

    Figure 1.30. Perylene diimide based C-shaped receptors (1.28), S-shaped receptors (1.29) and phenazine (1.30). Adams et al. determined the free energy of arene-halogen interactions in CHCl3 solvent. The

    interactions between nitropyrrole-containing bisaniline derivatives (1.31, 1.32, 1.33, and

    1.34) with isothaloyl derivative (1.35) were studied with double-mutant-cycles. As the

    geometry of the zipper complexes is locked, the architecture of the system from one side

    allows them to determine the thermodynamic contribution of relatively weak interactions.

    The measured halogen−aromatic interaction (repulsive) free energies (∆∆G) were 3.0, 3.2

    and 1.2 kJmol-1 for 1.32, 1.33 and 1.34 where X= F, Cl and Br respectively, compared to the

    free energy with 1.31.61

  • 31

    N NH

    OX

    X

    X

    O

    NO2N

    O

    O

    NH

    N

    O2NNO2

    H

    H

    H

    1.35

    1.31 = H1.32 = F1.33 = Cl1.34 = Br

    φ

    Figure 1.31. Interactions between the nitropyrrole containing bisaniline derivatives (1.31, 1.32, 1.33 and 1.34) with isothaloyl derivative (1.35). (φ = 30 ≈ 37°)

    Brammer et al. (2008) reported experimental studies of the interaction enthalpies and

    entropies of C−F···I−M halogen bonds, determined by titration of a nickel fluoride complex

    trans-[NiF(C5F4N)(PEt3)2] with the halogen bond donor C6F5I. Thermodynamic data in

    heptane solvent for trans-[NiF(C5F4N)(PEt3)2]·C6F5I (∆H° = −26.1 kJmol−1, ∆S° = 63.4

    Jmol−1K−1) (1.36) and for trans-[NiF(C5F4N)(PEt3)2]·2C6F5I (∆H° = −21.1 kJmol−1, ∆S° =

    83.5 Jmol−1K−1) (1.37) indicated a higher entropic penalty for the trimolecular complex, as

    expected.62 For the synthesis of trans-[NiCl(C5F4N)(PEt3)2] (1.38), a halogen bonding

    interaction (Ni−F···Cl−C) might be involved in the substitution of the metal-bound fluorine

    by Me3SiCl in THF solvent (Figure 1.32).63 Although the nickel fluoride complex trans-

    [NiF(C5F4N)(PEt3)2] interacted with pentafluoroiodobenzene (C6F5I) in 1:1 and 1:2 ratios,

    the complex formation ratio in toluene was 1:1. In an extension of this work, Perutz et al.

    (2011) studied Ni, Pd, and Pt fluoride complexes in toluene (d8). Using 19F NMR titrations at

    various temperatures, the enthalpies and entropies of interaction were determined. The trend

    for halogen bonding interactions based on −∆H° was Pt > Pd > Ni. In terms of ligands,