ANION SENSING IN POLAR MEDIA BY FLUORECENT SMALL MOLECULE RECEPTORS VIA HALOGEN BONDING by HANNAH JOSEPHINE BATES A DISSERTATION Presented to the Department of Chemistry and Biochemistry and the Division of Graduate Studies of the University of Oregon in partial fulfillment of the requirements for the degree of Doctor of Philosophy September 2022 DISSERTATION APPROVAL PAGE Student: Hannah Josephine Bates Title: Anion Sensing in Polar Media by Fluorescent Small Molecule Receptors via Halogen Bonding This dissertation has been accepted and approved in partial fulfillment of the requirements for the Doctor of Philosophy degree in the Department of Chemistry and Biochemistry by: Dr. Ramesh Jasti Chairperson Dr. Darren Johnson Co-Advisor Dr. Michael Haley Co-Advisor Dr. Amanda Cook Core Member Dr. Keat Ghee Ong Institutional Representative and Krista Chronister Vice Provost for Graduate Studies Original approval signatures are on file with the University of Oregon Division of Graduate Studies. Degree awarded September 2022. ii © 2022 Hannah Josephine Bates iii DISSERTATION ABSTRACT Hannah Josephine Bates Doctor of Philosophy Department of Chemistry and Biochemistry September 2022 Title: Anion Sensing in Polar Media by Fluorescent Small Molecule Receptors via Halogen Bonding Anions are a ubiquitous class of molecule which play critical roles in many of the environments with which humans concern themselves. From our physiology to nuclear waste to our water supply, anions have been the focus of much of the host- guest chemistry to have emerged within the last five decades. Host-guest chemists have used supramolecular tools, including halogen and hydrogen bonding, along with many others, to reversibly binding these and many other analytes. The purposes of binding are as wide ranging as real-time molecular recognition, catalysis, mechanically interlocked molecules, purification, and many others. Due to the many factors that can contribute to the efficacy of host-guest systems, including solvent effects, pocket size, molecular geometry, pH and others, most systems must be fined tuned for their particular application. One frequently encountered challenge is the competitive nature of any polar solvents, especially water, in solution state binding. Halogen bonding has recently come onto the scene as a potential answer to this problem, often demonstrating higher binding affinities than comparable hydrogen bonding molecules in polar environments. Despite its promise, much remains to be learned about how to best deploy halogen bonding motifs. iv This lacuna motivates the work covered in this thesis, which explores the design principles behind incorporating halogen bond donors effectively in a variety of arylethynyl; systems which were inspired by similar hydrogen bonding scaffolds published by the Johnson and Haley groups. After examining the successes and pitfalls of these halogen bond receptors in Chapter 3, Chapters 4 and 5 go on to report on the use of halogen bond and hydrogen bond systems in tandem, which are shown to bind the chloride anion notably well in appreciably polar organic solvents. These studies, suggest that the accepted definition of a halogen bond, as presented by IUPAC and discussed in Chapter 1, biases chemists to place undue importance on ensuring that the location of a polarizing group is as close as possible to the halogen-bond donor. Finally, future studies that can help flush out the ideas established in this thesis are reviewed in Chapter 6. This dissertation includes published and unpublished co-authored material. v CURRICULUM VITAE NAME OF AUTHOR: Hannah Josephine Bates GRADUATE AND UNDERGRADUATE SCHOOLS ATTENDED: University of Oregon, Eugene Sarah Lawrence College, Bronxville DEGREES AWARDED: Doctor of Philosophy, Chemistry, 2022, University of Oregon Master of Science, Chemistry, 2022, University of Oregon Bachelor of Art, Chemistry, 2018, Sarah Lawrence College AREAS OF SPECIAL INTEREST: Physical Organic Chemistry Supramolecular Chemistry Anion Recognition Halogen Bonding PROFESSIONAL EXPERIENCE: Associate Scientist, Thermo Fisher Scientific, May – December 2021 Graduate Research Fellow, University of Oregon, 2019 – 2022 Graduate Teaching Fellow, University of Oregon, 2018 – 2019 GRANTS, AWARDS, AND HONORS: ARCS Scholar Award, ARCS Foundation Oregon, 2018-2021 PUBLICATIONS: H. J. Bates, T. P. de Faria, A. S. Trom, F. I. D. Longnight, M. P. Miller, L. N. Zakharov, M. M. Haley & D. W. Johnson, manuscript in prep. T. P. de Faria, H. J. Bates, A. S. Trom, F. I. D. Longnight, M. P. Miller, L. N. Zakharov, M. M. Haley & D. W. Johnson, manuscript in prep. vi J. P. Bard, H. J. Bates, C.-L. Deng, L. N. Zakharov, M. M. Haley, J. Org. Chem. 2019, acs.joc.9b02132. vii ACKNOWLEDGMENTS I would like to take this space to thank the countless people who offered me support through my time in grad school as well as before it. Let me begin by thanking my parents, who kept a home secure enough that my curiosity could grow uninhibited and were always a source of unconditional love. I also am so grateful to my younger siblings, Mason, Austin and Riley who, despite their youth, never fail to inspire me with their amazing strengths. Another thank you is deserved by the many STEM students at Sarah Lawrence College including: Brooke, Isabella, Ellie, Ella, Sarah, and so many others; there was such a beautiful and tight-knit, field-spanning scientific community at SLC and it wouldn’t have existed without you guys. Another shout out to Emma, Ollie, and Jules for being friends I could talk to when I needed a break from science. Finally, a thank you to my many amazing professors from SLC, who taught me so much; Prof. Mike Siff and Prof. Colin Abernethy deserve a special shout out. As an advisor, Mike ensured that all of his mentees knew exactly how seriously we should take college (reminding all of us that school was not “real life”, no matter how important it felt) and helping me keep a cool head until the calmer winds prevailed once again. As a lecturer, Colin transferred his enthusiasm for the field of chemistry to every student who walked through his doors; beyond being just engaging though, Colin was a safe harbor for any question, chemistry related or not, he was already ready to lend a hand or an ear as the situation required. Without the support of all of the aforementioned people I would have never made it all the way to grad school. This brings me to the myriad of amazing people that I have met in grad school. First, let me thank my advisors, Professors Darren Johnson and Mike Haley who have viii done an admirable job wrangling much of my chaotic energy into a productive four-year stint here at UO. Thanks to my other committee members, Ramesh Jasti, Amanda Cook and Keat Ghee Ong, who all gave me both the opportunity to graduate and some amazing advice my 3rd year that I still think about frequently. Another thank you to my advisors from Thermo Fisher, Adam York, Dan Cash and Eric Welch who made sure that my 6 months there were filled with lots of learning and laughs alike. Also, a huge shout out to the front office staff at UO, Janet, Christi, Kathy, Jim and Rob who keep this program running and the people in it as happy as possible. Next let me give a hearty thank all of my labmates, past and present, in the DWJ, Haley labs and beyond – Jess, Sean, Merideth, Jordan, Hazel, Josh, Jeremy, Chun-lin, Trevor, Ngoc-Minh, Jacob, Henry, Thaís, Scout, Faith, Luca, Willow, Alex, Grace, Ben, Jack, Kiana, Efrain, Nolan, Gabby, Bella, Allison, Nick, Olivia, Mark, Russell and Haley – the amazing people that I met in grad school were easily my favorite part of the experience. Jeremy, Chun-lin, Jess, Hazel, and Thaís you guys have taught me so much and never fail to pick up the phone when I need advice or to bounce ideas off of someone. Jordan, you made every new member to the lab (myself included) feel so welcome when they joined and were always the first to person make the fun-loving and balanced culture of our lab a talked-about priority. Sean, you introduced me to motorcycles and consequently ensured every day to and from work was something to smile about. Thaís, Scout and Faith whether board games or titrations you guys saturated every moment with fun, love and support and I am unspeakably grateful for our palio- ship. Willow, Luca and Alex I am so happy my time here has overlapped with your own and that I can leave knowing that the lab is in good hands that will continue to guide it ix steadily. Ben, you have been such an amazing friend, housemate and confidant for the past four years and know that Massachusetts is a great spot to consider for your post grad school aspirations. Finally thank you to my fiancé, David. The past seven years have been filled with things that I never could have expected but together we handled them with so much more grace than I ever could have alone, and I am so grateful for the support, sage advice, good food and comfort that you never seem to run out of. No matter what the lab held for me I always look forward to coming home to you and Ezra and Kali (our two desperately snuggly kittens that never fail to put a smile on my face). x To David and Thaís – The past several years have often been trying, but that hardship was always salved by your generous love and friendship. Both of you, in your own inimitable ways, have shown me new and beautiful meanings to the word “partner” and I could not be more grateful to have you in my life. xi TABLE OF CONTENTS Chapter Page I. INTRODUCTION ................................................................................................... 1 II. AMPLIFICATION OF THE QUANTUM YIELDS OF 2‐λ5‐PHOSPHA- QUINOLIN-2-ONES THROUGH PHOSPHORUS CENTER MODIFICATION .... 11 2.1 Introduction ...................................................................................................... 11 2.2 Results and Discussion .................................................................................... 13 2.3 Conclusion ....................................................................................................... 19 2.4 Experimental Section ....................................................................................... 20 III. TOWARDS THE MOLECULAR RECOGNITION OF THE HYDROSULFIDE AND CHLROIDE ANIONS WITH AN ARYL ETHYNYL RECEPTOR ................ 25 3.1 Introduction to DWJ-Haley Lab Receptors and Halogen Binding .................. 25 3.2 Methods............................................................................................................ 28 3.3 Results and Discussion .................................................................................... 30 3.4 Conclusion ....................................................................................................... 36 IV. SUBSTITUENT EFFECTS IN A SERIES OF COMPACT ARYLETHNYL “MONODENTATE” HALOGEN BONDING HOSTS FOR ANION ....................... 38 4.1 Introduction ...................................................................................................... 38 4.2 Results and Discussion .................................................................................... 40 4.3 Conclusion ....................................................................................................... 48 V. FUTURE RECEPTORS AND FUTURE DIRECTIONS: AN UPDATE ON CURRENT WORK AND WHERE IT IS GOING...................................................... 49 5.1 Introduction ...................................................................................................... 49 xii Chapter Page 5.2 Synthesis of Tridentate Hybrid Receptor 1 ...................................................... 50 5.3 Results and Discussion of Tridentate Hybrid Receptor 1 ................................ 51 5.4 Synthesis of Quinoline-Based Receptor 2 ....................................................... 53 5.5 Computation Work with Library of Receptors ................................................ 55 5.5 Conclusion ....................................................................................................... 57 VI. CONCLUSION...................................................................................................... 58 APPENDICES ............................................................................................................. 61 A. Supplementary Information For Chapter 2 ....................................................... 61 B. Supplementary Information For Chapter 3 ....................................................... 113 C. Supplementary Information For Chapter 4 ....................................................... 122 D. Supplementary Information For Chapter 4 ....................................................... 181 REFERENCES CITED ................................................................................................ 203 xiii LIST OF FIGURES Figure Page 1. Figure 1.1 – Examples from the literature of phosphate receptors containing a) bis-urea motifs, b) hydrogen-bond donors and ferrocene motifs, and c) indoles. ............................................................................................ 03 2. Figure 1.2 – Receptor scaffolds for a) Gen 1, hydrogen-bonding where X=N, N+–R, CH; R= H, Me, F, OMe; R'= H, tBu, CF3, SO2Me, SO2CF3 ; R''= H, NO2, CF3, OMe and b) Gen 2, halogen bonding where X=N, CH; R= H, Me, R'= H, tBu, CF3, SO2Me, SO2CF3; Z= H, I, Br. The central aromatic ring containing X is referred to as the core of the receptor, and the ligands attached to the core referred to as arm(s). ............................................................................ 05 3. Figure 1.3 – A pictogram that represents halogen bonding where blue is associated with low electron density and red is associated with high electron density. ............................................................................................. 06 4. Figure 1.4 –The first XB receptor (a) informing modifications for a hydrosulfide receptor (b). ............................................................................................................ 07 5. Figure 1.5 – Family of One-Armed Receptors studied in Chapter Four............... 08 6. Figure 2.1 – Well-studied coumarin and carbostyril scaffolds (top) compared to phosphaquinolinone analogues (bottom). .............................................................. 12 7. Figure 2.2 – Absorption (solid lines) and fluorescence (dotted lines) spectra of 2 in CHCl3 at 298 K. ..................................................................................................... 15 8. Figure 2.3 – Selected bond length and dihedral angle in the optimized S0 and S1 structures of 2f calculated by DFT and TD-DFT methods at the PCM(CHCl3)- PBE0/TZVP level of theory, respectively. ............................................................ 17 9. Figure 2.4 – a) Characteristic PN-heterocycle dimer for 2f with the O···N distance (Å) shown as well as b) bond angles and c) torsional angles formed within monomers upon dimerization. Ellipsoids drawn at 30% probability. ........ 19 10. Figure 3.1 – Original DWJ-Haley pyridinium halogen bonding receptor and the second iteration derived from it, designed to sense hydrosulfide anions .............. 26 11. Figure 3.2 – Crystallographic data of receptor 1 binding DCM ........................... 31 12. Figure 3.3 – Absorbance data and binding isotherm for receptor 1 with TBACl in CHCl3 ..................................................................................................................... 33 xiv 13. Figure 4.1 – Previously studied bis-arylethynyl XB receptor, a. Novel “one-armed” neutral, b, and charged, c, arylethynyl receptors. .................................................. 40 14. Figure 4.2 – ESP maps of the entire family of receptors with the σ-hole values underneath each structure. Neutral structures are depicted on the left and methylated structures on the right. ......................................................................... 41 15. Figure 4.3 – Hammett plot between 5a-d and I–. The slope of the line indicated that substituents in this receptor has little role in affecting the sensitivity of the halogen-bond donor. .............................................................................................. 45 16. Figure 4.4 – Crystals structures of: a, 5a binding to chloride; b, 5a binding to iodide; c, 5b binding to iodide; and d, 5d binding to iodide. ............................................. 46 17. Figure 5.1 – The two receptors that will be the focus of this chapter. .................. 49 18. Figure 5.2 – (a) Representative 1H NMR titrations of the tridentate receptor 1 with TBACl, and (b) the binding isotherm associated with that titration ............. 52 19. Figure 5.3 – Family of one arm receptors discussed in chapter 4 and Table 5.3 ................................................................................................................ 56 xv LIST OF TABLES Table Page 1. Table 2.1 – Photophysical Properties and HOMO–LUMO Energy Gaps of Heterocycles 2a ...................................................................................................... 15 2. Table 2.2 – Dimerization Constants and Energies for 2 ....................................... 18 3. Table 4.1 – Binding constants, in M–1, of charged receptors 5a-d with guest chloride and their calculated sigma-hole values in kJ/mol. ................................... 44 4. Table 4.2 – XB or HB distances, in Å, of receptors 5a, 5b, and 5d binding to chloride or iodide as calculated from their crystal structures in Figure 4. ............. 46 5. Table 5.1 – Titrations done with receptor 1 & the halides at r.t. in ACN 10% DMSO via 1H NMR .............................................................................................. 53 6. Table 5.2 – Summarized the calculated energy value of sigma holes for XB donors in all of the receptors discussed in Chapters 4 and 5 (first eight entries refer to the one arm family..................................................................................... 56 xvi LIST OF SCHEMES Scheme Page 1. Scheme 2.1 – Synthesis of Phosphaquinolinones 2............................................... 14 2. Scheme 3.1 – Synthetic route to access second generation DWJ-Haley XB anion receptor .................................................................................................................. 29 3. Scheme 4.1 – Synthetic pathway and yields for the formation of the neutral, 4a-d, and charged, 5a-d, receptors. ................................................................................. 42 4. Scheme 5.1 – The synthetic route to host receptor 1. ............................................ 51 5. Scheme 5.2 – Synthetic scheme for the R = Bn of the derivatives of receptor 2 ............................................................................................................... 55 xvii CHAPTER ONE INTRODUCTION This chapter contains work in the process of being published and was written by Thaís P. de Faria and Hannah Bates, with editorial assistance from Douglas H. Banning, Dr. Trevor A. Shear, and Professors Michael M. Haley and Darren W. Johnson. This review article accounts the science and history behind anion binding, host-guest chemistry, some of the first papers on the project and foreshadows future work. However, it is by no means a comprehensive review of the field, rather a few pertinent examples have been selected for discussion to give a general sketch of the state of the field. Anions are prevalent in the human body and vital to life. They play a variety of different roles from producing electrical signals to maintaining cell volume.1 While there are many anions present in living systems, chloride (Cl–) and the protic forms of phosphate (i.e. HPO42-, H2PO4–) play a critical role in maintaining basic metabolic pathways within the human body. Two types of phosphorous species (organic and inorganic) are prevalent in our body's biological processes. An example of organic phosphates are the nucleotide phosphates found in adenosine monophosphate, diphosphate, and triphosphate (AMP, ADP, and ATP, respectively), molecules known for supplying energy for most biological reactions, participating in metabolism, and transferring genetic information.2 Inorganic phosphates, on the other hand, play a part in bone formation and help maintain bone density.3 They are normally seen in adult human serum at concentration levels of 0.8-1.45 1 mM;4 however, when present in excess, inorganic phosphates can ultimately lead to cardiovascular disease due to vascular calcification and renal failure.5 Deficient concentration of inorganic phosphate can lead to negative physical effects, like osteomalacia, central nervous system dysfunction, and muscle weakness, amongst many other disorders.6 Chloride anions are found within human cells in concentration levels between 5 and 15 mM1 and are vital to cellular processes, including balancing the charges of transfer proteins and building membrane potential. The natural role and regulation of Cl– is so critical that there are even chloride transport channels in intracellular organelles and the plasma membrane.7,8 Mutations of these channels lead to disruption of chloride movement in and out of the cell and have been linked to many genetic diseases, such as cystic fibrosis and the degeneration of muscles.9 Because these anions play such important roles within the human body, there is significant interest for their detection and monitoring in physiologically-relevant media. Previous approaches to quantify the concentration of these anions has largely involved dialysis and sample dilution,10,11 microelectrodes (for measurement within individual cells),12,13 and imaging using yellow fluorescent proteins.14,15 These detection methods all have major drawbacks, including difficulty in performing the measurements, pH sensitivity, and low selectivity. Because of these challenges, the use of supramolecular chemistry in this area of research has been of growing interest, as supramolecular host- guest chemistry is well studied and often allows for the ability to fine-tune the electronic and steric properties of the hosts used.16–19 Recently, attention has shifted to the use of 2 these small molecule sensors that can reversibly bind and display either a colorimetric or fluorescent response upon anion binding and, finally, can be performed in vivo.2,20–25 Many research groups have been attempting to conquer the feat of designing and developing a small molecule receptor that displays a selective preference for reversibly binding one targeted anion over others, that is sensitive enough to detect minute concentrations, and that exhibits some form of optoelectronic response to binding.19,26,27 One example of a small molecule receptor that has shown reversible binding to dihydrogenphosphate (H2PO4–) in dimethyl sulfoxide (DMSO) involves a scaffold with a hydrogen-bond donating (bis)thiourea motif (Figure 1.1a); however, the receptor has a modest limit of detection, requiring at least ~1 mM concentration of analyte for detection.28 Another example of promising binding to H2PO4– in acetonitrile includes a redox-based sensor that contains a ferrocene unit in addition to hydrogen-bond (HB) donors (Figure 1.1b).29 The HB units provide initial binding, but after oxidation of ferrocene, the host-guest interactions get further stabilized with the electrostatic interactions of the ferrocenium moieties.30 In addition to binding strongly with the phosphate anion, it also binds well with fluoride, and, therefore, lacks selectivity. A more recent example of phosphate anion detection comes from Sessler et al., who synthesized a. b. O c. NO2 O ONH O S NH HN S N NNH HN O NH HN HN Fe N N H H Figure 1.1. Examples from the literature of phosphate receptors containing a) bis- urea motifs, b) hydrogen-bond donors and ferrocene motifs, and c) indoles. 3 an indole-based receptor that showed promising colorimetric response upon binding (Figure 1.1c); however, this receptor lacks water solubility.31 Some examples of successful receptors and sensors that are able to detect chloride through non-covalent interactions include: 1) nucleic acid-based hosts that can detect anions over a wide pH range, but they either lack anion specificity because of their ion- pairing recognition mechanism or lack functionality for targeting some cellular compartments;32–35 2) chloride-templated rotaxanes that use redox-active interlocked systems to act as electrochemical sensors for anions, although binding only occurred in minimally polar solvents and did not show fully reversible oxidation;36 3) foldamers that have the ability to selectively coil around chloride and bind it tightly (albeit, the energy of organization of the helix was only optimal in certain solvents);37–42 and numerous others.43–46 The original, or Gen 1, host-guest receptors to emerge from the Johnson and Haley labs have shown to bind a range of anions through non-covalent, reversible interactions.47–52 The tunable scaffold includes urea (HB) motifs (Figure 1.2a) as well as three functional groups (R, R', and R'') that can be synthetically modified to optimize the optoelectronics, selectivity, and binding properties of the receptors. In addition, the conjugated backbone of the arylethynyl scaffold has allowed for some receptors to show a fluorometric response upon binding.47–49 Finally, all but one of the recent Johnson/Haley receptors have only demonstrated anion binding in organic media or wet organic solvents, which does not necessarily reflect a physiologically-relevant environment. 4 Figure 1.2. Receptor scaffolds for a) Gen 1, hydrogen-bonding where X=N, N+–R, CH; R= H, Me, F, OMe; R'= H, tBu, CF3, SO2Me, SO2CF3 ; R''= H, NO2, CF3, OMe and b) Gen 2, halogen bonding where X=N, CH; R= H, Me, R'= H, tBu, CF3, SO2Me, SO2CF3; Z= H, I, Br. The central aromatic ring containing X is referred to as the core of the receptor, and the ligands attached to the core referred to as arm(s). The most promising arylethynyl bis-urea scaffold to show chloride binding in water was synthesized and studied by Michelle Watt et al. in 2015.21 The way that solubility was achieved in this case was by protonating the core pyridine nitrogen using 1% TFA in water. The most notable result in this paper was that a “turn-on” fluorescence was reported in water/DMSO mixtures and increased with a higher percentage of water added, peaking at about 80% water.21 The observed fluorescence was attributed to a mechanism called “aggregation-induced emission” (AIE), that occurs through J- aggregation and is widely used in fluorophores to sense analytes.53 Although titration studies demonstrated a strong selectivity for chloride, the receptor exhibited weak binding (Ka ≤ 100 M–1) in DMSO/H2O mixtures and the fluorescence was not strong enough for in vivo cell studies. Alas, the search for a water-soluble arylethynyl receptor that can selectivly bind to chloride in aqueous media and displays strong fluorescent properties has continued in the Johnson and Haley laboratories. Firstly, to accomplish the goal of using our supramolecular receptors in physiological environments, we have to optimize the reversible interactions in polar, 5 protic solvents, like water. Sensing anions in water is a difficult mission for chemists due to the inherently low charge density which is a consequence of the large anion size. This makes the non-covalent electrostatic interactions between anions and receptors less efficient. In addition to this, anions have high free energies of solvation in water;54 consequently, the receptors must energetically compete with the medium.55 According to Hunter et al., a promising strategy to achieve aqueous anion detection employs halogen bonding (XB) interactions, which exhibit stronger anion binding in polar, protic solvents than their HB analogs.56 Halogen bonding is a type of reversible non-covalent binding interaction that has been gaining interest in the scientific community since it was first reported.57 The nature of this interaction was defined by IUPAC in 201358 as: R—X···Y, where X is a polarizable halogen covalently bound to an R group and Y is an electronically-dense halogen bond acceptor (Figure 1.3). This attractive interaction is facilitated by the δ– X = I, Br X δ+ Figure 1.3. A pictogram that represents halogen bonding where blue is associated with low electron density and red is associated with high electron density. halogen atom's ability to shift electron density with the help of the covalently bonded R group and, thus, reveal an area of positive electrostatic potential, named the σ-hole for its electron deficiency, directly opposite to the R–X σ-bond.59 Since the σ-hole is located on the opposite side of halogen, X, to the R—X bond, the covalently attached “R” substituent influences the size of the partially positive hole with its electron-withdrawing 6 character. Conventinal wisdom holds that more withdrawing the substituent, R, the larger the σ-hole and, therefore, the stronger the binding to the XB acceptor. Surrounding the σ- hole is an area of high electronegativity and this feature gives the σ-hole its strong directionality, demonstrating a strong preference for non-covalent bonding at a 180˚ angle from the XB acceptor.60,61 New anion receptors, or Gen 2 receptors (Figure 1.2b), were developed in the Johnson/Haley lab to utilize halogen bonding. In place of the bis-urea hydrogen binding motif found in Gen 1 receptors, iodine was installed as a halogen bond donor in the binding pocket. To enhance the XB interactions, strong electron-withdrawing functional groups at the R’ positions can be used to increase the polarization of XB donors, and consequently, the strength of binding. Dr. Jessica Lohrman synthesized and characterized the first XB Johnson and Haley receptor (Figure 1.4a).62 It was shown to bind to the halide anion family, as well as produce a “turn-on” fluorescence response; however, the receptor was not sufficiently soluble in water for biological studies. Since the desired applications of this receptor Figure 1.4. a) The first XB receptor informing modifications for b) a hydrosulfide receptor. require water solubility, modifications to achieve solvation in aqueous media must be made. This receptor inspired another XB receptor, Figure 1.4b, for use with the hydrochacogenide guests, mainly hydrosulfide, HS-, which will be discussed in more detail in Chapter 3. 7 We then set out to explore what is herein described as Gen 3 receptors (Figure 1.5). These receptors were highly influenced by the results of Dr. Lohrman’s XB receptor.62 With the target analyte, chloride, proving to be too small for the binding pocket available in the “two-arm” receptor, more simplified “one-armed” analogues were designed with: the purpose of simplifying/creating a smaller pocket for the receptors, converting their preference of binding chloride over iodide and studying their geometry of binding. Figure 1.5. Family of One-Armed Receptors studied in Chapter Four We investigated the meta isomer of the mono-dentate receptor and discovered that, in addition to halogen bonding, this receptor also utilizes hydrogen-bonding to non- covalently bind to anionic guests. The small hybrid–part XB, part HB–pocket was able to reverse the preference of binding seen in Gen. 2 receptors, now binding the smaller halides better than the larger (i.e. the target analyte, chloride, exhibited strongest binding). This study also revealed that the strongest contributor to the XB donor’s σ-hole is the methyl pyridinium ion located on the core aryl ring and that the tunable functional groups on the XB donor’s aryl ring play little part in increasing strength of binding through thorough theoretical, solution- and solid-state experimental structure/property relationship studies. 8 Finally, in Chapter Five, the fourth generation of Johnson and Haley receptors will be discussed. This chapter features two hybrid receptors (which use both HB and XB to bind anions). One receptor contains a quinoline backbone that enables the molecule to preorganize into what we anticipate being its preferred binding geometry. Its binding pocket consists of a triazole-polarized XB donor as well as multiple HB donors. We anticipate this hybrid receptor to demonstrate an optospectroscopic shift upon binding, allowing anion affinity to be measured by UV-Vis as well as 1H NMR. Notably, depending upon the identity of the HB donors, this molecule could have a preoganized state in either its cationic or neutral form. This allows for the molecule to reduce the entropic energy penalty associated with maintaining a single rigid conformation, like the one associated with binding. The second receptor is a larger version of the one-arm receptors discussed in Chapter Four, bearing a urea HB donor and pyridnium polarized XB donor. This receptor was designed to have a smaller pocket that seen in the bisarylethynyl receptors discussed in Chapter Three, with the aim of flipping the prefernce of the host receptor so that it demonstrates higher bind affinities for the smaller halides over the larger ones. Finally, yet to be mentioned is the contents of Chapter Two, which features work done on P, N containing heterocyclic fluorophores. This relatively new class of molecule is actually a spin off project, derived from the unexpected product produced when a previous graduate student used a phosphorous containing catalyst on a precursor to one of the DWJ/Haley lab HB receptors. The synthesis as well as the photophysical properties of this molecule are discussed and analyzed in detail in this chapter. 9 Considering the significance of anions and the important roles they play in biological and environmental systems, the need for monitoring their concentrations in aqueous media cannot be understated. Hydrogen bonding has long been used by nature and synthetic chemists alike for this purpose. And, while halogen bonding had been used in solid state crystallographic engineering for some time, it is only within the last two decades that solution-state XB has come onto the scene as an alternative molecular recognition strategy with particular promise in more competitive, polar solvents compared to that of their HB counterparts. In this short time, the field of halogen bonding has already demonstrated successful applications not just in molecular recognition but also in transmembrane transport, catalysis, template-directed synthesis, molecular motion, and many others niches as well. Despite this rapid expansion of research into halogen bonding, the number of reported receptors is still dwarfed by that of their hydrogen bonding analogues and much research remains to be done. This research is sure to be concomitant to further investigation of other sigma-hole driven interactions, like those seen in chalcogen and pnictogen bonding. There also is much to be learned about how to best use these complementary tools in tandem for even greater selectivity and strength of binding. With so much promise and still much nuance left to elucidate it is certain that the use of halogen bonding for anion binding, as well as a myriad of other applications, will continue to garner increasing attention for years to come. 10 CHAPTER TWO AMPLIFICATION OF THE QUANTUM YIELDS OF 2‐λ5‐PHOSPHAQUINOLIN-2-ONES THROUGH PHOSPHORUS CENTER MODIFICATION This chapter includes previously published and co-authored material from Bard, J.P., Bates, H.J., Deng, C.-L., Zakharov, L.N., Johnson, D.W., Haley, M.M. “Amplification of the Quantum Yields of 2-λ5-Phosphaquinolin-2-ones Through Phosphorus Center Modification.” J. Org. Chem. 2020, 85, 85–91. This work highlights our work on understanding the effects of placing a phenyl group in place of the traditional phenoxy group on the P center of the phosphaquinolinone scaffold. It draws structure-property relationships with both the sumpramolecular and photophysical properties. 2.1 Introduction. Small molecule fluorophores are used ubiquitously throughout many different fields, including chemical biology, molecular probe development, and materials for industrial and environmental sensing.1–4 In many of these applications, a fluorophore must exhibit a few characteristics to be considered optimal: large Stokes shift, high brightness, and red-shifted emission. One such example of a molecule that meets these specifications is coumarin (Figure 2.1).5–14 An impressive number of coumarin- containing compounds have been reported throughout the literature that are collected either from in-lab syntheses or by isolation from natural sources.15,16 This scaffold has 11 been the subject of a variety of synthetic modifications, diversification by adding groups onto the backbone, or incorporation of the coumarin system into larger ring networks.17–19 Through these modifications, a tremendous breadth of understanding upon the structure– activity relations has been developed20–24 that have guided the design of many useful derivatives, which have emerged for applications in chemosensing and many other areas.25–31 Figure 2.1 Well-studied coumarin and carbostyril scaffolds (top) compared to phosphaquinolinone analogues (bottom). Alongside the many derivatives of the parent coumarin scaffold, there is the nitrogen-containing structural analogue known as carbostyril (Figure 2.1).32–37 Though not as widely utilized as coumarin, carbostyril is the subject of many structure–property relationship studies, and it shows promise for use in both pharmaceutical discovery and fluorescence imaging applications.38,39 These carbostyril analogues expand on the applications of the coumarin family through modifying the lactone core to a lactam. With further alteration of this core, new applications, functionality, and fluorescent properties are expected from this widely used fluorophore. Recently, we reported a series of phosphorus- and nitrogen-containing (PN) phosphaquinolinone 1 derivatives (Figure 2.1).40,41 This scaffold, which is one of only a handful of similar heterocycles,42–52 is also an isostere to carbostyril and coumarin, with 12 the only difference being the replacement of the lactam carbonyl with an isolobal, chiral phosphorus center. We have performed a variety of structure–property studies that looked at the effects of both acene core modification and substitution at various points on the scaffold.40,41,53–56 In these studies, it was found that the emission wavelength can be moderately red-shifted through careful substitution of various groups on the backbone, affording significant Stokes shifts and modest quantum yields. On the basis of these design principles, this moiety has recently been implemented in a fluorescent receptor for HSO – 4 in acidic media, showing promise for future applications of this scaffold that take advantage of both its exceptional hydrogen bonding capabilities and its inherent fluorescence.54 As clearly shown for the coumarin and carbostyril motifs, systematic modification of various structural aspects can lead to very useful derivatives. In our continuing efforts to study the 2-λ5-phosphaquinolin-2-one skeleton, the next facet that we wanted to explore was the variation of the group attached to the phosphorus center. Disclosed herein is the substitution of a phenyl ring in place of the standard phenoxy upon the phosphorus center, generating a racemic mixture of heterocycle 2 (Figure 2.1). We also hypothesize that this would increase the quantum yield through rigidifying the scaffold. With these modifications, the reported compounds could have greater potential for applications in phosphorus-containing chemosensors and fluorophores, expanding on this pre-existing group of molecules.57–65 2.2 Results and Discussion. The synthesis of 2 starts from key arylethynylaniline intermediate 3, prepared following previously reported methods.40 Aniline 3 is then reacted with diphenyl 13 phenylphosphonite (PhP(OPh)2) in pyridine at 100 °C. Subsequent hydrolysis in THF at 60 °C furnishes phenyl-appended heterocycles 2 in modest to good yields (Scheme 2.1). Scheme 2.1 Synthesis of Phosphaquinolinones 2. The photophysical properties of 2a–2f in CHCl3 are shown in Figure 2.2 and are compiled in Table 2.1. All derivatives share a common λmax at ca. 300 nm, and the lowest energy absorption peaks range from 343 to 381 nm. The absorption coefficients for this scaffold stay within the range of 1.5 × 104 (for 2a) to 2.2 × 104 M–1 cm–1 (for 2e and 2f). The λem values range from 447 (for 2a) to 515 nm (for 2f) with Stokes shifts on the order of 6350–7000 cm–1. Interestingly, the emission spectra of 2 show a ca. 20 nm bathochromic shift from those of the analogous congeners of 1,40 and the quantum yields of this scaffold show a dramatic improvement, on the order of a 4–5-fold increase in most cases. Brightness values range from 6.84 × 103 (for 2b) to 1.14 × 104 M–1 cm–1 (for 2a), which are now on par with several optimized coumarin derivatives.66 Fluorescence lifetime measurements were also performed (Figure C.6), and the radiative (kr) and nonradiative (knr) decay rate constants were determined. The kr values range from 0.06 to 0.19 ns–1, and the knr values vary from 0.06 to 0.18 ns–1, showing either equal rates or a slightly larger knr in most cases. 14 Figure 2.2 Absorption (solid lines) and fluorescne (dotted lines) specta 2 in CHCl3 at 298 K. Table 2.1 Photophysical properties and HOMO-LUMO Energy Gaps of Heterocycles 2a These values elucidate a potential explanation for the increased quantum yields when compared to similar values of phosphaquinolinones 1.40 For 1, the kr values range from 0.10 to 0.30 ns–1, which show similar values, whereas the knr values vary from 0.30 to 3.0 ns–1, which are substantially faster in most cases. The diminished ratio of knr to kr seen for 2 may suggest that the reduced degrees of freedom may indeed be the cause of the increased quantum yields. 15 To gain further understanding of these experimental results, the frontier orbitals for heterocycles 2 were calculated (Tables 4.1 and C.1–C.7). The narrowest HOMO– LUMO gap is seen for 2f (3.90 eV), and the largest is for 2a (4.40 eV). This trend arises because of a higher magnitude of HOMO destabilization than that of the stabilization of the LUMO with more donating substituents (Table C.1). These values also follow a similar trend to the optical gaps (Tables 4.1 or C.1). TD-DFT was then used to examine the S0 to S1 transition. It was found that the S0 to S1 transition is dominated by the HOMO–LUMO transition (Table C.1). Additionally, the distributions of the HOMOs and LUMOs show a slightly more pronounced separation due to a larger HOMO localization at the phosphorus center (Figure C.4). These observations suggest that π to π* transitions are dominant, but there may be some intramolecular charge transfer (ICT) occurring in the excited state. The emission of 2 was then examined in solvents of varying polarities (Figures C.7–C.13 and Tables C.11–C.16). In these studies, bathochromatic shifting is observed in every case with more polar solvents. Compound 2f showed the greatest shifting, ranging from emission wavelengths of 504 to 531 nm and Stokes shifts of 6070 and 7070 cm–1 in cyclohexane and acetonitrile, respectively. The TD–DFT optimized S1 state near the Franck–Condon geometry of 2f shows that the dihedral angle between the parent core and appended 4-cyanophenyl substituent becomes smaller compared to the ground state (Figure 4.3). Additionally, the C–C bond connecting them is shortened by ca. 0.035 Å. The considerable geometric changes lead to a more conjugated system at the S1 state, in which the HOMO–LUMO energy gap decreases by 0.9 eV (Figure C.5); thus, the computed emission wavelength of 541 nm is 16 within the observed emission maxima (Table 4.1). These computational results could further explain the large Stokes shift for this type of PN-heterocycle. Figure 2.3 Selected bond length and dihedral angle in the optimized S0 and S1 structures of 2f calculated by DFT and TD-DFT methods at the PCM(CHCl3)-PBE0/TZVP level of theory, respectively. In addition to improved fluorescence properties, we were curious to see how phenyl substitution would affect the strength of hydrogen bond dimerization we typically observe for phosphaquinolinones. Variable concentration (VC) NMR experiments were performed in water-saturated CDCl3 to assess the strength of dimerization (Tables 2.2 and C.17–C.21, Figures C.14–C.23w). Heterocycles 2 exhibit dimerization strengths of 22 (for 2e and 2f) to 82 M–1 (for 2b). While these values are roughly 70–80% smaller than those measured for the analogous congeners of 1, the strengths of dimerization for 2 again exceed those of many typical head-to-tail hydrogen bonded dimers.67 This result suggests that this new entry into the phosphaquinolinone family can still be implemented in supramolecular systems, as found in 1.54 17 Table 2.2 Dimerization Constants and Energies for 2 Single crystals suitable for X-ray diffraction were grown by slowly diffusing pentane into a CHCl3 solution of 2f, and the resultant data are shown in Figures 4.4 and C.1–C.3. The structure of 2f still features the typical meso-dimer between racemates (Figure 2.4a); however, the N···O distance in the dimer of 2f (2.874 Å) is longer than those of heterocycles 1 (2.768–2.821 Å),40 which supports the observation that the molecule should form a weaker dimer in the solution-state as well. This weakened hydrogen bonding interaction can potentially be explained by examining the pseudo six-membered ring formed between the monomers. The N···O–P (115.53°), O–P–N (116.66°), and P– N···O (104.57°) angles formed between the participating atoms in the dimer formation show significant deviation from the ideal 120° orientation (Figure 4.4b), likely caused by the large O–P–N–H torsional angle of 55.35° (Figure 2.4c). With an angle so much larger than the analogous angle found in the crystal structures of several derivatives of 1 (ca. 30– 40°),40 there is a less ideal orientation for the two monomers to associate, slightly weakening the interaction overall. By comparing optimized geometries of the meso-dimer of 2f and its −OPh analogue, there are some additional steric clashes in 2f among the C– 18 H atoms in the phenyl ring and N–H moieties, according to the noncovalent interactions (NCI) plot (Figure C.24). Moreover, the natural bond orbital (NBO) analyses predict a total contribution of the nO → σNH* interactions of 23.1 kcal mol–1 for 2f and 25.0 kcal mol–1 for the respective −OPh analogue (Figure C.25). Therefore, the strength of dimerization may decrease to some degree due to the weaker primary hydrogen bonding and extra steric hindrance, in agreement with the observed diminished Kdim for 2. Figure 2.4 a) Characterization PN-heterocycle dimer for 2f with the O···N distance (Å) shown as well as b) bond angles and c) torsional angles formed within monomers upon dimerization. Ellipsoids drawn at 30% probability. 2.3. Conclusion. In summary, we have shown the effects of the attachment of a phenyl group on the phosphorus center of the phosphaquinolinone scaffold. This new class of PN- heterocycles not only has large Stokes shift values (up to 7000 cm–1) but also shows a marked 4–5-fold increase in the quantum yield when compared to previously reported phenoxy-substituted compounds. Additionally, this modification retains the strong dimerization strengths of the scaffold in both the solid and solution states. This new 19 modification deepens the fundamental understanding of the phosphaquinolinone scaffold and allows for further possibilities in the applications of this scaffold as a biologically or industrially relevant fluorophore, like the coumarin and carbostyril scaffolds. 2.4. Experimental Section General All air- or water-free reactions were performed under a N2 atmosphere using Schlenk techniques. Column chromatography was performed using silica gel (240–300 mesh), with solvent systems being referenced to the most abundant solvent. NMR spectra were acquired at room temperature on a Varian Inova 500 instrument (1H: 500 MHz, 13C: 126 MHz, 19F: 471 MHz, 31P: 202 MHz) or a Bruker Avance III HD 500 apparatus equipped with a Prodigy multinuclear cryoprobe (1H: 500 MHz, 13C: 126 MHz). 1H and 13C chemical shifts (δ) are expressed in parts per million (ppm) relative to residual CHCl3 shifts (1H: 7.26 ppm, 13C: 77.16 ppm) or residual DMSO shifts (1H: 2.50 ppm, 13C: 39.52 ppm). 31P and 19F NMR spectra are referenced to 85% H3PO4 (δ 0 ppm) and to CFCl3 (δ 0 ppm), respectively, as the external standards. UV–vis spectra were recorded using an Agilent Technologies Cary 60 UV–vis spectrophotometer in HPLC-grade CHCl3. Fluorescence emission spectra were recorded using a Horiba Jobin Yvon FluoroMax-4 fluorimeter exciting at 365 nm. Quantum yields (φ) were determined through a comparison of the emission and absorption intensities of the analyte to that of a 0.1 M H2SO4/quinine sulfate solution.68 Fluorescence lifetime measurements were recorded using a Horiba FluoroHub Single Photon Counting Controller with a TemPro Fluorescence Lifetime System attachment. High-resolution mass spectra (HRMS) were 20 recorded on a Waters XEVO G2-XS mass spectrometer. 2-Ethynylanilines 3a–3f40 and phenyl diphenylphosphonite (PhP(OPh)2)69 were prepared as previously described. General Synthetic Procedure for Phosphaquinolinone 2 2-Ethynylaniline3 (1.0 equiv.) and PhP(OPh)2 (2.0 equiv.) were dissolved in pyridine (ca. 0.35 M). The vessel was sealed and heated to 100 °C for 24 h in an oil bath. The mixture was then diluted with toluene, and the solvent was removed in vacuo. This was repeated three times to remove all residual pyridine. The crude material was dissolved in THF, and ca. five drops of water were added. The solution was stirred at 60 °C for 1 h before being dried (Na2SO4), filtered, and concentrated in vacuo. The crude mixture was then purified by column chromatography on silica gel. Reported yields are given for >95% pure material (by 1H NMR spectroscopy), though subsequent recrystallization from hexanes and CH2Cl2 was used to achieve analytically pure material. Phosphaquinolinone 2a Compound 2a was synthesized from 3a (462 mg, 1.9 mmol, 1 equiv.) and PhP(OPh)2 (1.11 g, 3.8 mmol, 2 equiv.). Column chromatography (1:1 EtOAc:CH2Cl2, Rf = 0.20) gave 2a (149 mg, 21%) as a pale brown solid: mp > 250 °C; 1H NMR (500 MHz, DMSO-d6) δ 10.28 (d, J = 3.8 Hz, 1H), 8.14 (d, J = 2.0 Hz, 1H), 8.05 (d, J = 30.1 Hz, 1H), 7.82 (ABm, J = 8.4 Hz, 4H), 7.81–7.77 (m, 1H), 7.69–7.64 (m, 2H), 7.58–7.51 (m, 1H), 7.49–7.43 (m, 2H), 7.21 (d, J = 8.5 Hz, 1H); 13C{1H} NMR (126 MHz, DMSO-d6) δ 143.1 (d, J = 3.8 Hz), 140.9 (d, J = 11.7 Hz), 139.3, 135.8, 133.8, 132.6, 132.5 (d, J = 2.7 Hz), 132.4 (d, J = 137.0 Hz), 132.2 (d, J = 10.8 Hz), 128.6 (d, J = 13.2 Hz), 128.1 (d, J = 6.1 Hz), 126.9 (d, J = 115.9 Hz), 119.2, 119.0 (d, J = 9.9 Hz), 118.5, 117.7 (d, J = 8.1 21 Hz), 110.6, 102.1; 31P{1H} NMR (202 MHz, DMSO-d6) δ 7.77; HRMS (ASAP) [M + H]+ calcd for C22H15N3OP 368.0953, found 368.0977. Phosphaquinolinone 2b Compound 2b was synthesized from 3b (700 mg, 2.4 mmol, 1 equiv.) and PhP(OPh)2 (1.40 g, 4.8 mmol, 2 equiv.). Recrystallization from CH2Cl2 and hexanes gave 2b (580 mg, 58%) as a yellow solid: mp > 250 °C; 1H NMR (500 MHz, DMSO-d6) δ 10.14 (d, J = 4.1 Hz, 1H), 8.16 (d, J = 30.1 Hz, 1H), 8.06 (d, J = 2.2 Hz, 1H), 7.85 (d, J = 8.3 Hz, 2H), 7.80 (d, J = 8.4 Hz, 2H), 7.73–7.63 (m, 3H), 7.53 (td, J = 7.4, 1.5 Hz, 1H), 7.49– 7.43 (m, 2H), 7.25 (d, J = 8.5 Hz, 1H); 13C{1H} NMR (126 MHz, DMSO-d6) δ 142.5 (d, J = 3.7 Hz), 141.1 (d, J = 12.0 Hz), 139.9, 132.7 (d, J = 136.9 Hz), 132.6, 132.4 (d, J = 2.7 Hz), 132.2 (d, J = 10.7 Hz), 130.5, 128.5 (d, J = 13.2 Hz), 128.1 (d, J = 6.1 Hz), 127.4 (d, J = 3.7 Hz), 126.5 (d, J = 116.3 Hz), 124.5 (q, J = 271.2 Hz), 120.6 (q, J = 32.3 Hz), 118.6 (d, J = 13.0 Hz), 118.5, 117.4 (d, J = 7.9 Hz), 110.5; 31P{1H} NMR (202 MHz, DMSO-d6) δ 7.86; 19F NMR (471 MHz, DMSO-d6) δ −59.92; HRMS (ASAP) [M + H]+ calcd for C22H15N2OF3P 411.0874, found 411.0909. Phosphaquinolinone 2c Compound 2c was synthesized from 3c (645 mg, 2.6 mmol, 1 equiv.) and PhP(OPh)2 (1.5 g, 5.1 mmol, 2 equiv.). Column chromatography (1:1:1 hexanes:EtOAc:CH2Cl2, Rf = 0.10) followed by two rounds of recrystallization from CH2Cl2 and hexanes gave 2c (300 mg, 31%) as a yellow solid: mp > 250 °C; 1H NMR (500 MHz, DMSO-d6) δ 9.80 (d, J = 4.2 Hz, 1H), 7.99 (d, J = 29.8 Hz, 1H), 7.83 (d, J = 8.4 Hz, 2H), 7.78 (d, J = 8.5 Hz, 2H), 7.73 (d, J = 2.5 Hz, 1H), 7.69–7.57 (m, 2H), 7.56–7.49 (m, 1H), 7.48–7.38 (m, 3H), 7.11 22 (d, J = 8.7 Hz, 1H); 13C{1H} NMR (126 MHz, DMSO-d6) δ 141.3 (d, J = 12.0 Hz), 139.4, 138.3 (d, J = 3.7 Hz), 132.8 (d, J = 136.6 Hz), 132.5, 132.3, 132.2, 130.7, 129.9, 128.5 (d, J = 13.1 Hz), 128.1 (d, J = 6.2 Hz), 126.5 (d, J = 116.8 Hz), 123.6, 120.2 (d, J = 12.6 Hz), 118.6, 118.4 (d, J = 8.1 Hz), 110.3; 31P{1H} NMR (202 MHz, DMSO-d6) δ 7.69; HRMS (ASAP) [M + H]+ calcd for C21H15N2OPCl 377.0611, found 377.0641. Phosphaquinolinone 2d Compound 2d was synthesized from 3d (151 mg, 0.69 mmol, 1 equiv.) and PhP(OPh)2 (463 mg, 1.4 mmol, 2 equiv.). Column chromatography (1:1:1 hexanes:EtOAc:CH2Cl2, Rf = 0.20) gave 2d (144 mg 61%) as a yellow solid: mp > 250 °C; 1H NMR (500 MHz, CDCl3) δ 7.73 (d, J = 8.1 Hz, 2H), 7.66 (d, J = 30.4 Hz, 1H), 7.71–7.66 (m, 2H), 7.54 (d, J = 8.2 Hz, 2H), 7.47–7.42 (m, 2H), 7.38–7.32 (m, 3H), 7.06 (t, J = 7.5 Hz, 1H), 6.97 (d, J = 8.0 Hz, 1H), 6.83 (br s, 1H); 13C{1H} NMR (126 MHz, CDCl3) δ 141.5 (d, J = 11.8 Hz), 141.0 (d, J = 3.0 Hz), 138.8 (d, J = 4.3 Hz), 132.7 (d, J = 10.8 Hz), 132.6 (d, J = 2.8 Hz), 132.5, 132.2 (d, J = 139.3 Hz), 131.5, 131.2, 128.6 (d, J = 13.8 Hz), 128.4 (d, J = 6.2 Hz), 126.0 (d, J = 119.5 Hz), 121.4, 119.4 (d, J = 12.1 Hz), 118.8, 117.3 (d, J = 7.7 Hz), 111.6; 31P{1H} NMR (202 MHz, CDCl3) δ 10.52; HRMS (ASAP) [M + H]+ calcd for C21H16N2OP 343.1000, found 343.1030. Phosphaquinolinone 2e Compound 2e was synthesized from 3e (549 mg, 2.0 mmol, 1 equiv.) and PhP(OPh)2 (1.3 g, 4.0 mmol, 2 equiv.). Column chromatography (1:1:1 EtOAc:CH2Cl2, Rf = 0.25) gave 2e (520 mg, 65%) as a yellow solid: mp > 250 °C; 1H NMR (500 MHz, CDCl3) δ 7.73 (d, J = 8.0 Hz, 2H), 7.72–7.67 (m, 2H), 7.68 (d, J = 30.3 Hz, 1H), 7.53 (d, J = 8.1 Hz, 2H), 23 7.48– 7.39 (m, 3H), 7.37–7.32 (m, 2H), 6.90 (d, J = 8.4 Hz, 1H), 6.48 (br s, 1H), 1.35 (s, 9H); 13C{1H} NMR (126 MHz, CDCl3) δ 144.4, 141.7 (d, J = 12.1 Hz), 141.5, 136.4, 132.7 (d, J = 10.8 Hz), 132.5, 132.5, 132.4 (d, J = 139.2 Hz), 129.1, 128.5 (d, J = 13.6 Hz), 128.4 (d, J = 6.4 Hz), 127.6, 125.8 (d, J = 119.9 Hz), 119.0, 118.9 (d, J = 8.4 Hz), 116.9 (d, J = 7.3 Hz), 111.4, 34.4, 31.5; 31P{1H} NMR (202 MHz, CDCl3) δ 10.44; HRMS (ASAP) [M + H]+ calcd for C25H24N2OP 399.1628, found 399.1629. Phosphaquinolinone 2f Compound 2f was synthesized from 3f (430 mg, 1.7 mmol, 1 equiv.) and PhP(OPh)2 (969 g, 3.3 mmol, 2 equiv.). Column chromatography (1:1:1 hexanes:EtOAc:CH2Cl2, Rf = 0.25) gave 2f (200 mg, 31%) as a pale yellow solid: mp > 250 °C; 1H NMR (500 MHz, CDCl3) δ 7.73–7.65 (m, 2H), 7.71 (d, J = 8.6 Hz, 2H), 7.59 (d, J = 30.4 Hz, 1H), 7.53 (d, J = 8.2 Hz, 2H), 7.47–7.42 (m, 1H), 7.37–7.31 (m, 2H), 6.98–6.93 (m, 2H), 6.88 (d, J = 8.5 Hz, 1H), 6.60 (br s, 1H), 4.04 (q, J = 7.0 Hz, 2H), 1.43 (t, J = 7.0 Hz, 3H); 13C{1H} NMR (126 MHz, CDCl3) δ 153.5, 141.6 (d, J = 12.0 Hz), 140.7, 132.8 (d, J = 10.7 Hz), 132.7, 132.6 (d, J = 3.0 Hz), 132.5, 132.1 (d, J = 140.4 Hz), 128.5 (d, J = 13.8 Hz), 128.4 (d, J = 6.4 Hz), 126.7 (d, J = 119.9 Hz), 119.9, 119.8 (d, J = 11.8 Hz), 118.8, 118.2 (d, J = 7.7 Hz), 114.8, 111.5, 64.3, 15.0; 31P{1H} NMR (202 MHz, CDCl3) δ 10.51; HRMS (ASAP) [M + H]+ calcd for C23H20N2O2P 387.1261, found 387.1283. 24 CHAPTER THREE TOWARDS THE MOLECULAR RECOGNITION OF THE HYDROSULFIDE AND CHLROIDE ANIONS WITH AN ARYL ETHYNYL RECEPTOR 3.1. Introduction to DWJ-Haley Lab Receptors and Halogen Binding. Research groups at the University of Oregon have been collaborating to synthesize supramolecular receptors for ion recognition for over a decade now.1–7 Like many in the supramolecular field of host guest chemistry much of our efforts have been focused on anions, which are conventionally thought of as more difficult to bind that their cationic counterparts, due to their relatively high pH sensitivity, powerful solvation energies and lower charge density.8–10 Most of the receptors out of the DWJ and Haley laboratories were rigid and highly conjugated diethynyl system that not only can bind anions but also have an inherent fluorescence that allows for spectroscopic recognition upon binding. Originally the receptors featured HB donors, with amine and urea functional groups, and were used to bind the halide anions: chloride, bromide and iodide. Eventually, through careful design and several structure-property relationship studies, the binding capabilities of these hosts were expanded to include oxoanions - nitrate, phosphate and perchlorate - as well as hydrochalcogenides - hydrosulfide and hydroselenide.2,4,11 While these guests were bound by HB donors with appreciable affinities, competition from similar guest molecules became apparent when attempting to bind them in highly competitive polar solvents like DMSO and water.6,11,12 As stated earlier in this dissertation (see Chapter I), XB donors have typically demonstrated stronger binding in 25 more competitive environments when compared to analogous HB donors.13 With this in mind, Dr. Jess Lohrman synthesized and characterized a XB receptor from the traditional arylethynyl core seen in DWJ-Haley lab receptors (see structure 0 in Figure 3.1).3 This scaffold maintained the highly extended conjugation that gave previous hydrogen bonding analogues of these scaffolds their inherent fluorescence but, in place of the traditional HB donors, two Figure 3.1 – Comparison between (top) the previously published “first generation” DWJ-Haley XB receptor 0 & iodine atoms were installed to act (bottom) the current generation XB receptor 1 that is the focus of this chapter as the XB donors. These XB donors were polarized by sulfonyl functional groups, bridging acetylene units and a methylated pyridinium core to yield a sigma hole. When titrated with the halides in DMSO, the receptor 0 demonstrated Kas ranging from 690 to 3600 M-1 with the strongest binding affinity being demonstrated for the largest halide, iodide. Having proven the utility of anion binding with this class of XB receptors, the next step was to investigate whether XB interactions would also be an effective strategy for binding the hydrochalcogenides, initially hydrosulfide (HS-), in highly polar media, 26 like those encounters in cells. With this purpose in mind, XB receptor 1 was designed, building upon the design principles mentioned above. The new structure maintained the highly conjugated arylethynyl scaffold, enabling spectroscopic measurement of binding interactions by UV-Vis. Additionally, the central pyridine ring was inverted, giving a 1,2,6- orientation between the two ethynyl arms and the central pyridine nitrogen, as opposed to the original 1,3,5 geometry. This orients the pyridine’s lone pair towards the intended internal binding pocket of 1. Previous work with similar urea-based systems has shown that an internally-directed pyridine core has weakened overall binding to Cl- and HS- while concomitantly demonstrating a higher selectivity for HS- over Cl-, compared to its benzene ring analog.2 This is possibly explained as a destabilizing electron-electron repulsion between the pyridine’s lone pair and the anionic guest, Cl-. However, if the guest anion has a polarized hydrogen atom, then this internal lone pair could also serve as a HB acceptor to the anion’s HB donor.14 In a system like the one seen in receptor 1 it is postulated that, with a protic guest, the host could coordinate the guest with bidentate halogen bond donation and a single point of hydrogen bond acceptance at the pyridine core. However, further studies will have to be done to confirm this as the rigid 180 ° directionality of XB donors might hold the guest molecule too far away to experience either attractive or repulsive forces from the lone pair of the pyridine core. To mitigate this electron repulsion between the pyridine lone pair and to add a hydrogen bond donation site, a charged pyridinium core created by either protonation or methylation (1+H and 1+Me in Figure 3.1) would serve as a very strong receptor for the halides. Additionally, the charge of the pyridinium moiety would both serve to further polarize the XB donating iodine atoms through induction, thus creating stronger XB 27 donors, as well as add a site of HB donor to the central pocket! It should be noted, however, that this would not be an effective strategy for making an HS- host, as that hydrochalcogenide is a particularly nucleophilic guest. A previous group member, Dr. Hazel Fargher, had documented similar problems with a neutral HB host and hydroselenide (HSe-) in which the nucleophile attacked one of the acetylene carbons, causing decomposition to begin within minutes of guest addition.12 The assumption while designing the next generation of XB receptors was that any positively charged host (like 1+H or 1+Me) would open itself up to decomposition by the reactive hydrochalcogenides, which would likely result in breaking the aromaticity of the molecule in order to accommodate the addition of HS-. This vulnerability drove many of the changes seen between receptor 0 and the iterations of receptor 1. The methane sulfonyl group was exchanged for a trifluoro methyl group, which is less challenging synthetically and won’t promote inviting unwanted reactions with the guest (but is otherwise a relatively similar EWG, having a p of 0.54, compared to the 0.72 of SO2Me – although the lion’s share of the polarization in receptor 0 came from the methylate pyridinium).15 The pyridine ring was inverted, as discussed above, and the protonated or methylated pyridinium functionalities (1+H and 1+Me) were set to be exclusively used with the less reactive halide guests. 3.2 Methods. The modular nature of the synthesis of receptor 1 makes it easily tunable and synthetically accessible. In the initial attempt to synthesize receptor 1 (Appendix A for Chapter 3 – Scheme A.2), a triazene functional group was incorporated on the 28 trifluoromethyl sidearms via a Sandmeyer reaction;16 however, an alternative synthetic path was required because the triazene moiety produced an impurity which made isolation of the final product impractical. To avoid this impurity, Scheme 3.1 was developed, wherein the aniline side arms were left unprotected, completely side stepping the reaction required to append the triazene species, and still furnishing the final product. To start, 4-trifluormethylaniline is iodinated under acidic conditions to create 2-iodo-1- methyl-4-(trifluoromethyl) benzene, 2, in very good yield. Afterwards, a Sonogashira cross-coupling with trimethylsilyl acetylene (TMSA) provides the trimethylsilyl- protected ethynyl aniline 3.3,17,18 This was subsequently deprotected with a basic workup to afford the respective terminal acetylene that is immediately subjected to a second Sonogashira cross coupling with 2, 6-dibromopyridine, at a 2.5:1 molar ratio, to obtain the dianiline 4.19 After being purified via silica gel flash column chromatography followed by recrystallization, the final step is a variant on the Sandmeyer reaction which KI, KIO TMSA, NHNH2 3 NH 2 HCl, MeOH 2 Pd(PPh3)2Cl2, CuI F C H2O F C I THF, DIPA F3C 3 3 TMS 80% 39% 2 3 1. K2CO3 THF, MeOH 2. 2,6 Dibromopyridine N NaNO2 Pd(PPh3)4, CuI F3C CF3 HI, KI NF3C CF3 50 °C -10 °C THF, DIPA NH2 H2N ACN 21% I 30% I 4 1 Scheme 3.1 – Synthetic route to access second generation DWJ-Haley XB anion receptor 1 cleaves the amine and installs two iodines inside the binding pocket. The final receptor is purified through recrystallizations and a series of plugs to give 1 in 30% yields.3 This reaction pathway not only avoids the aforementioned impurities, but it also removes the low-yielding protection step altogether. 29 3.3 Results and Discussion. To corroborate the identity of the scaffold and give valuable information about solid-state host-guest interactions, crystals of 1 suitable for x-ray crystallography were obtained via slow evaporation of a concentrated solution of 1 and TBACl in dichloromethane (DCM) (Figure 3.2). The C(I) bond lengths fell around a normal range of 2.100 or 2.0092 Å and all aryl protons demonstrated a C(H) bond length ranging from 0.949 to 0.951 Å. The receptor can be seen in its W-conformation, with the XB donors directed outside of the intended binding pocket. Interestingly, in place of the desired Cl- guest, a molecule of DCM was bound in a 2:1 host:DCM ratio. The W-conformation adopted by the host is likely due to the size of the DCM molecule. This geometry allows for C-H groups meta to the XB donor to coordinate with the Cl atoms of the DCM (at C(H)-Cl distances of 3.779 Å and 3.769 Å). Notably, one of the C-H groups of the DCM can be seen interacting with the internal pyridine nitrogen (C(H) ··· N distance = 3.248 Å) of the host molecule, perhaps supporting some of the earlier speculation on the benefits of having a lone pair in the pocket of the host. Finally, a third C-H HB, from a second host receptor, also coordinates to one of the guest’s Cls (C(H)-Cl = 3.678 Å) to complete multidentate binding of DCM in a 2:1 host-guest complex. Contrary to what 30 was expected or intended, the coordination of the guest with 1 occurs via HBing rather than XBing. This result served to emphasize the challenge of predicting binding geometry and although it is hard to say for certain whether a smaller guest, capable of more energetically favorable interactions, would persuade the receptor to adopt the intended U-shaped conformation, it is Figure 3.2: Crystallographic data of receptor 1 binding DCM abundantly clear that the DCM molecule is too large for such a geometry. To build upon the initial host-guest interactions seen with receptor 1 in the solid- state as well as previous success demonstrated by previous DWJ-Haley HB receptors, initial investigations into the binding nature of this host in solution were undertaken. Cl- was chosen as the first guest to be tested because these anions are relatively similar in size (Cl- having an ionic radius of 1.81 Å relatively the 1.84 Å of S-2) but Cl- is far less volatile, dangerous, and difficult to work with than HS-. These were carried out with tetrabutylammonium chloride (TBACl) as a Cl- source in chloroform, a non-competitive solvent (λ cutoff < 245 nm). CHCl3 has a relatively low dielectric constant, ε =4.8, in comparison to water, where ε = 80.4, which solvates charges, but not large organic host receptors, especially well.20 As mentioned before, this is one of the greatest challenges associated with anion binding in aqueous environments. Consequently, the non- 31 competitive nature of CHCl3 mitigated the likelihood that weak binding interactions will be out competed by solvent interactions and therefore can be observed. There are a number of ways in which binding data can be obtained so that binding strength can be calculated; these include isothermal titration calorimetry (ITC), UV-Vis, Fluorescence, potentiometric or NMR spectroscopy titration data. 1H, and 19F NMR titrations were tested as possible routes for determining the association constant of this receptor (1); however, none of these NMR active atoms played a direct enough role in coordination of the guest during the solution state binding event to track anion binding. As a result, titrations were performed using UV-Vis spectroscopy instead (Figure 3.3a). These allowed for the generation of an isotherm which plots the relative absorbance units at the λ max (275 nm & 325 nm)of the host as a function of the concentration of guest relative to the concentration of host (Figure 3.3b).21 With nonlinear regression analysis, utilizing a 1:1 host-guest complexation fitting model, titration isotherms can be used to quantify the binding affinity of the H-G system. Here, BindFit is used to process the raw UV-Vis data collected.22 In a typical isotherm, the absorbance of the host is expected to increase for a period before binding saturation is seen, indicating saturation of the host receptor. In the isotherm of receptor 1, no leveling off is seen, indicating weak to negligible binding. This was supported by a highly variable association constant measuring anywhere in the range of 0.3 - 6.7  103 M-1. 32 A 0.35 0.33 B 0.3 0.25 0.28 325 nm 0.2 0.23 274 nm 0.15 0.1 0.18 0.05 0 0.13 250 300 350 0 100 200 300 Wavelength (nm) [G]/[H] Figure 3.3 – (a) Absorbance data & (b) binding isotherm for host 1 with TBACl in CHCl3. From both the work of Dr. Jess Lohrman with receptor 0 and similar receptors from Dr. Orion Berryman, it has been demonstrated that the bisarylethynyl receptor’s pocket size is preferential to the iodide anion due to its larger size, which is likely a contributing factor to the limited interaction between chloride and the neutral host, and might explain the weak binding discussed above.3,23 Restrictive 180° XB angles do not allow binding in the internal pocket close. For chloride to be coordinated to both XB donors, it would have to fit further inside the pocket which would pull it off course of the sigma hole and hit the electron belt instead. Another possible explanation for the apparent weak binding of receptor 1 is the neutral charge of the molecule. The neutral analog of receptor 1 was a relatively ineffective host compared to that of the hypothetical cationic versions of the receptor, 1+H and 1+Me.3 A final consideration for this point is the internal facing pyridine core – which, when neutral, selects for HS- over Cl-, but has been shown to decrease binding strength for both.2 When designing the receptor at the center of this chapter, 1, the main guest of interest was HS-. But the other guest of interest hadn’t been forgotten; it was hypothesized that the pyridinium core (seen in 1+H and 1+Me) would enhance Cl- 33 Absorbance Units (AU) Absorbance Units (AU) binding. The cationic core holds much promise in the form of an additional HB donor as well as in the generation of a charged species, with attractive electrostatics for the anionic guests. To this end, several protonation experiments were carried out in DCM. Initial protonation attempts were carried out by introducing the host to a variety of acids. It wasn’t until using triflic acid (CF3SO3H) that precipitate of the supposed pyridinium was able to be collected. Many of the initial acids likely failed because they are less effective in organic solvents, having higher pKas than those reported in water because of their relatively unstable conjugate bases. Triflic acid, having a sufficiently charge disperse conjugate base, was finally able to protonate the core pyridine (that might look misleadingly basic in organic conditions, considering it also has a decent ability to delocalize charge). When an 1H NMR spectrum was taken of the protonated compound in acetone-d6 it did demonstrate an expected shift in all of the resonance environments. However, it was notable that there was not an easily identifiable peak associated with the proton assumed to be on the pyridine. This is likely due to exchange with the solvent under the highly acidic conditions in which the NMR spectrum was taken. This could potentially be resolved by taking the NMR spectrum in another solvent, at a lower temperature, in an acidic solvent system (like 1% TFA acetone-d6 or similar) or with a shorter time delay between collection. But regardless of how we choose to trouble shoot characterizing that receptor, it quickly became clear that the protonated 1+H was too labile to be applied in practical settings for anion recognition (1+Me presents a more promising target, with regards to stability of the host as it would not be directly influenced as easily by changing pH). 34 Due to its high acidity, the receptor 1+H proved unstable in water, deprotonating upon any aqueous treatment. This means that, while this receptor would likely show stronger binding when protonated, the pyridinium is too acidic to be used successfully in the more competitive polar environments it was designed to compete in. Though it should be noted that despite its water instability, it is still likely possible to do titrations with the protonated receptor in organic solvents. Thus, knowing that this wouldn’t be a newly developed small molecule fluorescent probe able to compete with water for anions, as we had wanted, we decided to forgo further characterization (like fluorescent studies) and instead move on to a new hybrid host design. If the next generation of receptors are to be successful in the aim of getting into water and cells, then they should abide (as closely as possible) to Lipinski’s rule of five, which is a set of parameters intended to assess how likely a small molecule is to be cell permeable.24 These parameters include having a molecular mass of less than 500 amu, as well as limiting the amount of HB donors and acceptors to less than 5 and 10, respectively. Future chapters in this thesis feature receptors which do a much better job following these general rules of thumb, with the hope of find more success not only in aqueous environment, but in vivo applications specifically. This was accomplished by abandoning the two-arm structure for a more compact, lighter “one arm” motif and leverage the best parts of both HB and XB systems by using both in tandem! While initial experiments with receptors 1 and 1+H seemed less successful than we might have hoped there is still much promise to be found in the 1+Me scaffold. The methylated pyridinium could serve to further polarize the XB donors, add an electrostatic attraction to all anion guests, as well as append 3 additional C-H HB donors to the inside of the pocket. This 35 final change might also serve to shrink the effective size of pocket, potentially switching the preference for larger halides seen in receptor 0. 3.4. Conclusion. In conclusion, a new fluorescent aryl ethynyl receptor with XB capabilities was synthesized and characterized. Characterizations were carried out via 1H NMR, 19F NMR, UV-vis spectroscopy, and X-Ray crystallography. Preliminary data suggests that the neutral host does not strongly bind chloride, even in a non-competitive solvent, which bodes ill for its ability to bind the more reactive HS- guest. This supports data previously seen by XB hosts out of the DWJ and Haley lab, which showed a preference for larger halides.3 While there have been many efficient and clever Cl- receptors published since Park and Simmons’ first host in 1968, few of them are both fluorescent and sufficiently strong enough to bind chloride in water.19,25 Continued efforts shall be made in optimizing this receptor to close the gap between anion and cation binding in aqueous media. This remains incredibly important for applications in vivo, where the high solvation energy of anions make them significantly more difficult to track and quantify than their positively charged counterparts. Bridge to Chapter IV. The next generation of XB receptors, alluded to in the above paragraphs, are the focus of the next two Chapters, IV and V, which explore a derivatized family of one- armed arylethynyl hybrid, or XB and HB, receptors and then another new tridentate hybrid receptor, respectively. The family of one-armed receptors were intended to 36 elucidate some of the more fundamental preferences for mixed HB and XB systems with a simpler host and, hopefully, so more promise for in vivo studies. This is in part motivated by the difficulty our colleague Dr. Jess Lohrman experienced trying to solubilize the receptor into polar environments, making it nonviable for work in cells, as well as the weak bind that the large flexible bis-arylethynyl XB host scaffold, 0, displayed towards Cl-, given its huge pocket.3 Additionally, in her receptor, chloride was bound out of pocket, something thought to be due to the scaffold’s large pocket size. Thus, new receptors (in Chapters 4 and 5) were tailor-made with smaller pockets, to increase the receptors’ affinity towards Cl- anion recognition. 37 CHAPTER FOUR SUBSTITUENT EFFECTS IN A SERIES OF COMPACT ARYLETHNYL “MONODENTATE” HALOGEN BONDING HOSTS FOR ANION This chapter includes previously published and co-authored material from T. P. de Faria, H. J. Bates, A. S. Trom, F. I. D. Longnight, M. P. Miller, L. N. Zakharov, M. M. Haley & D. W. Johnson, manuscript in prep. This work investigates the structural changes, made by modification of tunable groups with various EWGs and EDGs, to assess the influence of substituent effects in neutral versus charged receptors in this series. The results of the comparison of experimental binding trends and theoretical calculations will heavily influence future designs of small-molecule, halogen-bonding receptors from this scaffold family to bind to smaller anions more successfully and selectively in polar solvents. 4.1. Introduction. Anions are ubiquitous in nature—they play especially critical roles within our cells as well as in our external environment. Some of the most fundamental anions, the halides (I—, Br—, Cl—, F—), are crucial to human health and well-being. Examples include the misregulation of chloride in cells associated with disease states like cystic fibrosis1 and iodine deficiency causing thyroid disease and goiter formation.2 Despite their significance, non-covalent binding of these anions still proves to be difficult for several reasons, but most notably because the entire halide family shares the same spherical shape, similar basicities for the three heaviest halides, and only slightly increasing sizes, leaving little margin for error when designing a suitable selective binding pocket. 38 Additionally, anions are well-solvated, particularly by the polar, protic media that make up many of the more interesting applications of these systems, in vivo sensing, for example.3-5 In fact, the Hofmeister series illustrates how, especially as halides get smaller, the energy required to free a halide anion from water becomes increasingly costly in energetic terms.6,7 In spite of these challenges, there have been many supramolecular systems that successfully bind anions, like chloride, in competitive polar protic solvents.8-12 Many of these hydrogen bonding hosts feature multi-dentate coordination that seems necessary for these hosts to compete in aqueous media. Further preorganization through macrocyclization often further enhances binding, although can be synthetically challenging and may make cellular uptake and compatibility less likely, per Lipinski’s Rules, for in vivo cell studies.13,14 Recent work in the field of host-guest chemistry has expanded the supramolecular toolset beyond that of traditional hydrogen-bonding (HB) to now also include sigma-hole interactions, like those seen in halogen-15, chalcogen-16 and pnictogen-17 bonding.18 Previous studies by other groups have shown that halogen-bonding (XB), while often comparable in strength to hydrogen-bonding, tends to be favored in polar protic solvents.9,19,20 For this reason, our group, among many others, has recently begun investigating halogen bonding as a tool for binding halides and other anions.21–24 The work we present here was largely influenced by our previously studied bis-aryl-ethynyl or “two-armed” receptors (Figure 4.1a), whose pocket size seemed to be too large to accommodate our target anion, chloride.22 A crystallographic study showed the “two-arm” receptor changing configurations and binding chloride out of the intended pocket, unlike larger anions like iodide which fit 39 well. The new studies reported herein focus in on a compact “one-armed” structure (Figure 4.1b and 4.1c) that (1) eliminates the pre-determined pocket in hopes to shift binding preference to chloride over iodide and (2) enables the study of small, discreet changes on binding geometry.25,26 This paper will also describe the modification of the R substituents and how various electron-withdrawing and -donating groups affect the binding affinity in neutral vs charged receptors in this family. These results will allow for a number of observations to be made about the fundamental nature of the halogen bond—something often Figure 4.1. Previously studied bis-arylethynyl XB receptor, a. Novel “one-armed” neutral, b, and charged, c, arylethynyl receptors. reductively characterized as an attractive R–X -- :A interaction (where R is a sufficiently electron withdrawing atom, X is a halide and A is an anion or other Lewis base)— through a structure-property relationship study.27 4.2. Results and Discussion. Electrostatic potential surface (ESP) maps were generated for all potential neutral, 4a-d, and charged, 5a-d, host receptors (Figure 4.2) to see which synthetic alteration would have the greatest effect on the XB donor’s ability to noncovalently bind guests: the 40 positively charged N-methylpyridinium core, or the tunable R groups attached to the “arm” of the receptors. The results of the maps corroborate trends our group has seen in previous computational studies.22 When looking at the area of partial positivity— quantified and described as Vs,max—created from the σ-hole of the XB donor, it shows a Figure 4.2. ESP maps of the entire family of receptors with the σ-hole values underneath each structure. Neutral structures are in panel a, and methylated structures in panel b. This figure clearly illustrates the outsized effect the methylated pyridinium unit plays in polarizing our halogen bond donor (compared to the changing ID of the FG para to it). decrease in magnitude of about 11 kcal/mol as the R-groups para to the iodo XB donor become more electron-donating in the neutral pyridyl receptors 4a-d. In the pyridine-N-Me cationic receptors 5a-d, the change is similar with a 9 kcal/mol decrease across the series of electron withdrawing/donating R-groups. However, when comparing the neutral receptors with similar σp values as their charged 41 analogues, the Vs,max values significantly increase with a ~54-58 kcal/mol difference providing evidence that what contributes more in the XB σ-hole enhancement of these arylethynyl scaffolds is the presence of the cationic N- methylpyridinium core, despite its distance from the XB donor. To further verify this theoretical trend, all the scaffolds were synthesized for later solution-state binding studies, according to the modular synthetic path shown in Scheme 4.1. The starting aniline with the desired R-group was iodinated using KI Scheme 4.1. Synthetic pathway and yields for the formation of the neutral, 4a-d, and charged, 5a-d, receptors. and KIO3 to furnish 1a-d. Diazotization of the aniline group followed by trapping with Et2NH afforded triazenes 2a-d. Next, Sonogashira cross-coupling with 3- ethynylpyridine gave compounds 3a-d. The triazene group was then transformed 42 into an iodine, the XB donor chosen for each of our scaffolds, by heating with HI/I2 to produce neutral hosts 4a-d. The neutral receptors were transformed into the charged receptors through alkylation using MeI with I– as the counter ion, 5a-d • I–. To enhance the apparent association constants for these charged receptors, the iodide counterion was then exchanged for the less competitive PF –6 (5a-d • PF6–). The solution-state binding trend seen between our group’s previously studied two-arm analogue and the series of halides showed that the preferred binding pocket fit best with the largest anion, iodide;22 bromide demonstrated the second-strongest affinity and chloride last. To test the hypothesis that the two-arm pocket was too large to preferentially bind chloride over larger anions, 1H NMR titrations were performed with 1 mM solutions of single arm host receptor, 5a • PF6–, and 20-30 mM guest solutions of the tetrabutylammonium salts of chloride, bromide, and iodide in CD3CN. Association constants (Ka) were determined by tracking proton peaks and fitting the change in chemical shift to a 1:1 binding model using the Bindfit software.28,29 Receptor 5a • PF6– bound to Cl–, Br–, and I– with Ka values of 137, 104, and 103 M–1, respectively, demonstrating a modest yet inverse trend seen from our previous two-arm host analogue. This provides evidence that by changing the size of the pocket and reducing the entropic penalty associated with the previous host’s pre-organization, it is possible to shift the preference of anionic guest to slightly favor chloride despite the fact that iodide typically features stronger halogen bonding interactions in polar solvents, presumably due to lower solvation energies.30 The proton peaks that exhibited the largest change in chemical environment helped to elucidate the solution-state binding geometry of the receptors, which matched that of the calculated lowest energy state of the charged receptor (Figure 43 4.2) with the iodine XB donor facing the methylated pyridinium core. In all cases, the two hydrogens active in binding the halide guests are the pyridyl C–H in between the alkyne and the N atom, and one of the C–Hs of the pyridinium methyl group (Figure C1). A similar geometry was observed in the solid-state crystal data of our analogous bidentate XB host receptor that showed binding with the XB donor twisted out of the too- large pocket and HB participation by the polarized C–Hs on the core and the methyl of the pyridinium. Neutral receptors 4a-d did not show measurable binding to chloride in acetonitrile, demonstrating that XB donors with σ-holes with Vs,max of ~17-29 kcal/mol may not be significantly electron-withdrawn to participate in halogen binding. Charged receptors 5a-d•PF6– , and most notably those with lower σp values, were still able to bind to chloride in a trend that follows well with their calculated Vs,max values (Table 4.1). Table 4.1. Binding constants, in M–1, of charged receptors 5a-d with guest chloride and their calculated sigma-hole values in kJ/mol. 5a • PF – 5b 6 • PF –6 5c • PF –6 5d • PF –6 Ka with Cl– (M–1) 137 ± 5 120 ± 6 98 ± 4 85 ± 2 Vs,max (kJ/mol) 82.8 80.0 77.2 73.8 With the binding constants determined, we sought out to better comprehend the receptors’ binding sensitivity to chloride through the use of the Hammett relationship (log(K R Ha /Ka ) vs σp) by fitting the Kas to eq 1. R log 𝐾aH = 𝜌σp + ε (1)𝐾a Plotting these values against the σp of the R substituents revealed that there is a weak linear response between the Ka and the electron-donating/-withdrawing nature of the substituent (Figure 4.3). The equation inset in Figure 4.3 describes the parameters of 44 Figure 4.3. Hammett plot between 5a-d and I–. The slope of the line indicated that substituents in this receptor has little role in affecting the sensitivity of the halogen-bond donor. The dotted lines represent the 95% confidence interval. the linear fit, resolved through linear regression. The slope of the line’s positive magnitude tells us that chloride binding is favored with more electron-withdrawing para R substituents. The value of ρ, 0.173, reveals that the halogen bond in this system is less sensitive to substituent effects compared to the standard, benzoic acid; therefore, the σp of the tunable R group is a statistically insignificant predictor of binding sensitivity in the one-armed arylethynyl receptor, further corroborating the hypothesis that the methylpyridinium dominates this family of receptors’ binding affinity compared to the para R–X’s σp values. This result is contrary to what we found for CH H-bond donors in related neutral bis-arylethynyl urea scaffolds. The CH hydrogen-bonding motifs in those systems were more susceptible to para substituent effects than that of our novel XB receptor with a ρ value of 0.71.31 It remains to be determined if substituent effects in charged CH receptors would be mitigated by the overall charge of the host, a topic we will address in the future. After getting a handle on the one-armed XB receptors’ binding geometries and properties in solution-state, we set out to study solid-state binding geometries. Crystals suitable for x-ray diffraction were grown of 5a,d • – I (Figures 4.4b and 4.4d) 45 from a 2:1 mixture of CH2Cl2:acetone via slow evaporation. These structures revealed that the solid-state geometry of binding is similar to what we see in solution-state. Atomic distances calculated in the crystal structures were less than the sum of the van der Waals radii of the guest and the iodine XB donor, the aryl H, and the methyl H. These results indicate supramolecular binding occurs with not only our intentional XB donor, but also with the polarized CHs as H-bond donors. Table 4.2. XB or HB distances, in Å, of receptors 5a, 5b, and 5d binding to chloride or iodide as calculated from their crystal structures in Figure 4.4. 5a • Cl– 5a • I– 5b • I– 5d • I– C–H – aryl -- X distance (Å) 2.904 3.173 2.928 3.187 C–Hme -- X– distance (Å) 2.833 3.522 3.626 3.558 C–I -- X– distance (Å) 3.403 3.705 N/A 3.824 When comparing the binding distances in 5a and 5d, both binding I–, in the solid-state (Table 4.2), there is little difference in either of the C–H--I– distances of the HB donors – about 0.01-0.03 Å. The distance between the XB donor and a c b d Figure 4.4. Crystals structures of: a, 5a binding to chloride; b, 5a binding to iodide; c, 5b binding to iodide; and d, 5d binding to iodide. 46 iodide, C–I--X–, however, shows an order of magnitude greater difference with 0.12 Å. Receptor 5a, with the most electron-withdrawing substituent, showed a closer distance of guest to XB donor than 5d, the receptor with the most electron- donating substituent, suggesting that even in solid-state binding, there is a slight preference for substituents with electron-withdrawing characteristics. Crystals of 5a • Cl– were grown from a 1:2 acetone:CH3CN mixture via evaporation. When comparing the SO2CF3 receptor, 5a, binding to the largest guest, I–, and smallest guest, Cl–, an even larger shift is observed. The aryl C–H--I– distance changes by 0.27 Å, the methyl C–H--I– by 0.69 Å, and the C–H--X– by 0.30 Å with the guest halide becoming more tightly bound to the receptor with decreasing size of the anion. This trend is also shown subtly in the solution-state binding studies with a stronger Ka observed for chloride over iodide by 27 M–1. Finally, crystals of 5b – • I (Figure 4.4c) were grown from a mixture of 1:1:1 CH2Cl2:acetone:CH3CN. The crystals in this sample showed solid state binding of iodide with only the HB donors, C–H(aryl and methyl)--I–, which appeared to hold no trend with the previous bonding distances. The aryl C–H to iodide distance is 2.93 Å, the second shortest aryl C–H to iodide distance of the four crystal structures; however, the methyl C–H to iodide distance is 3.63 Å, the largest methyl C–H to iodide distance. Notably, this structure is showing a binding geometry with the XB donor facing away from the iodide guest and not participating in binding at all. This result begs further investigation to help understand the significance of solvent effects in supramolecular binding through crystal packing of host-guest molecules. 47 Conclusion In conclusion, the data reported for this compact series of receptors suggest that when designing XB receptors with an arylethynyl foundation, a charged pyridinium core will provide a greater enhancement to the size and strength of the σ-hole than any adjacent functional groups. This knowledge allows for the purpose of the R group to shift away from that of only a XB enhancer to, instead, a number of other purposes including being a modular handle for other desired applications, such as polymerization, increased water-solubility, various cell compartment targeting functional groups, fluorophores, etc. without disrupting the most important portion of the anion recognition (the charge- polarized XB donor). Additionally, recognizing that the R-group does not contribute significantly to the polarization of the XB donor in these charged hosts allows for future host-guest chemists to dodge the synthetic pitfalls that come with appending extremely electron withdrawing functional groups, like SO2CF3. These design principles, in addition to the binding demonstrated by the halogen/hydrogen bonding receptor in a polar solvent, will inform future host receptors for the halide anions. 48 CHAPTER FIVE FUTURE RECEPTORS AND FUTURE DIRECTIONS: AN UPDATE ON CURRENT WORK AND WHERE IT IS GOING 5.1. Introduction. Throughout the previous chapters of this thesis, several host receptors were discussed in detail. Their strengths and weaknesses have been scrutinized to further the understanding of effective host-guest design principles, motivated by the prospect of more selective future hosts. In this chapter, two final hybrid XB-HB scaffolds will be explored for their potential as hosts for the halide anions (Figure 5.1). Specifically, both hybrid receptors were designed to have a smaller binding pocket than the receptors discussed in Chapter 3, and thus should show a preference for the smaller Cl– anion. The deliberate use of XB and HB donors in tandem is also expected to show some of the HB enhanced XB that Berryman and coworkers have discussed in past publications (see red dashed line in Figure 5.1).1 These interactions have been shown to both further strengthen and stabilize the XB donor as well as preorganize the binding pocket into the geometry hypothesized to be best for binding. These hosts (1 and 2) Figure 5.1. The two also feature functional moieties and strategies developed receptors that will be the focus of this chapter. in previous studies: this includes i) reliable anion 49 recognition motifs used in the early generation of XB and HB hosts developed in the DWJ/Haley laboratories; ii) urea (or related guanidine) HB donors, iii) acetylene bridging units, and iv) XB polarization via a pyridinium group (discussed in Chapters 3 and 4) or the well-known iodotriazoles popularized by Beer et al..2–11 In addition to these motifs, there are design strategies derived from the success of the family of one arm receptors, discussed in Chapter 4. Both receptors 1 and 2 feature XB and HB donors working in tandem to bind anion guests, which produced appreciable affinity for the halide anions in the more polar organic solvent, ACN. Additionally, much like the family of one-armed receptors, the size of the binding pocket of receptors 1 and 2 is smaller compared to the large cavity seen in the earliest generations of the DWJ-Haley XB receptors (reviewed in Chapter 3). Our hypothesis is that the smaller binding pocket of 1 and 2 will invert the preference of the host to preferentially bind Cl– over the iodine anion, I–.6 The beginning of this chapter will focus on the synthesis and results from receptor 1. The latter half will focus on the synthesis of receptor 2, further design principles that inspired the scaffold, and some computational work which compares 2 to the receptors discussed in Chapter 4. 5.2. Synthesis of Tridentate Hybrid Receptor 1. Tridentate hybrid receptor 1 was synthesized in seven steps (Scheme 5.1). The first five synthetic steps were performed by Thaís de Faria. First, benzocaine is iodinated under acidic conditions to yield intermediate 1.1. A TMS protected alkyne is then deprotected and installed through Sonogashira cross-coupling of 1.1 with the acetylene. The ethynylaniline precursor 1.2 is then subjected to a second Sonogashira cross- 50 I Br N NH2 KI, KIO NH2 1. TMSA, K2CO3 THF, MeOH NH3 2 HCl, MeOH 2. 1.1, Pd(PPh3)2Cl2, CuI Pd(PPh3)4, CuIO O O H2O, 50 °C I N2, THF, TEA N2, 50 °C O O O TMS THF, TEA 1.1 1.2 H N 10 mol% CF3 NH2 NH NH2 O Br OCN5 mol% CuI, NaI O I O Dioxane, 110 °C O toluene, 55 °C N overnight N overnight 1.3 1.4 CF3 CF3 CF3 O NH MeI O NH AgPF O NH6 NH NH 110 °C Acetone, Water NH O I O I O I PF6 70% 95% O O I O N N N 1.5 1.6 1 Scheme 5.1. The synthetic route to host receptor 1. coupling reaction with 4-bromo-3-iodopyridine to yield 1.3. The bromine on the pyridine moiety is exchanged for iodine because iodine is known to be a stronger XB donor.12 Next, a urea HB donor is installed via addition of trifluoromethylphenyl isocyanate to the aniline moiety on 1.4. Neutral receptor 1.5 was refluxed with MeI to yield a pyridinium moiety, which was expected to polarize the XB donor. Finally, 1.6 undergoes a counter ion exchange with AgPF6, to give the final host receptor in good yield (95%). 5.3. Results and Discussion of Tridentate Hybrid Receptor 1. To test the anion binding ability of 1, titrations with the halides Cl–, Br–, and I– were performed in 10% DMSO in ACN. Unlike previous generations of XB hosts (like those discussed in Chapter 3) the incorporation of HB donors enabled the binding constant to be measured by 1H NMR spectroscopy, which concomitantly gives 51 information about binding geometry and stoichiometry (example shown in Figure 5.1a). The downfield chemical shift of the urea protons (starred in Figure 5.2) indicates their participation in binding and can be plotted against the concentration of guest (16-42 mM) present, relative to that of host (0.73-1.90 mM), and fitted to 1:1 binding isotherms through non-linear regression to give the association constants reported in Table 5.1.13,14 2.5 2 Predicted N(H) 1Predicted N(H) 2 Predicted C(H) 1.5 Experimental N(H)1 Experimental N(H)2 Experimental C(H) 1 0.5 0 0 10 [G]/[H] 20 30 A B Figure 5.2. (a) Representative 1H NMR spectroscopy titrations of the tridentate receptor 1 with TBACl, where the urea protons are starred and CH HB donor is marked with a triangle, and (b) the binding isotherm associated with that titration. Interestingly a third peak can also be seen shifting over the course of the titration (marked with a triangle in Figure 5.2) this is suspected to correspond to the hydrogen in the meta position, relative to our XB donor, acting as a C(H) HB donor. This suggests that, with the smaller anions, the iodine XB donor flips out of pocket and the anionic guest experiences tridentate coordination from exclusively HB donors, rather than the intended hybrid binding. These binding affinities indicates that the new receptor design successfully inverted the preference of the XB hosts to the smaller halide, Cl– (the reverse trend demonstrated by the first DWJ-Haley XB receptor reviewed in Chapter 3) as Cl– shows the strongest affinity for the receptor 1. This is an important finding because it highlights 52 𝜹 - 𝜹 0 that it is possible to tailor binding pocket size in future generations of these hosts to improve guest selectivity. Interestingly, Receptor 1 shows a stronger affinity for Cl– than for Br–, indicating that the size of the binding pocket is an important factor in anion binding and this small pocket favored smaller anions. If the size trend was to continue, it would be expected that then the affinity for I– would be even weaker. Instead, I– binding demonstrated a complexity not seen in the titrations with either Cl– or Br–. This is likely due to I– binding with this host differently from the other two halides, as indicated by the appearance of new peaks which growing in up field of the original receptor, suggesting a second species forming in solution. Additionally, when titrated with TBAI, the host peaks shifting only weakly downfield and decreasing in the intensity of the host peaks as more guest is added. Further experiments will be done to discern whether this is due to slow exchange, aggregation, some other complex binding pattern, or decomposition. These include performing diffusion-ordered spectroscopy (DOSY) NMR experiments (to assess whether aggregation is occurring), increasing the delay between scans (to better resolve any slow exchange), and performing VT NMR if both methods fail to elucidate clearer results. Additionally, post titration samples could be condensed and analyzed to verify whether decomposition has occurred. Solid state data from x-ray quality crystals could also help elucidate whether the larger anion is potentially binding out of pocket and/or interacting with more than one host molecule at a time. Finally, it should be noted that this titration of receptor 1 is conducted a fairly polar solvent system, which is relatively competitive, and demonstrates a marked improvement in anion recognition over that observed from data obtained in CHCl3, in the titrations discussed in Chapter 3. 53 Table 5.1. Titrations done with receptor 1 & the halides at r.t. in ACN 10% DMSO via 1H NMR - *the titration sets obtained for the I– anion cannot be fitted to a 1:1 binding model with 1 (consistent with the other anions) because there is a complex binding event occurring (See appendix D for more details) it is possible that slow exchange or higher order aggregation is occurring with the I– anion that was not seen with the Cl– and Br– anions. Halide Anion Guest Ka (M-1) Ionic Radii (pm)15 Cl– 440 181 Br– 90 196 I– * 220 Synthesis of Quinoline-Based Receptor 2. The quinoline-based receptor 2 highlights all the design features discussed over the course of this thesis. Its extended π system gives the host (and many of its intermediates) an inherent bright fluorescence opening the door to equal parts beautiful and useful applications, like bio-imaging.16–18 It is also likely that the binding of this receptor could be tracked by 1H NMR, fluorimetry or UV-Vis spectroscopy. A new motif is seen in the triazole ring, bearing the pendant XB donor. Triazoles are seen throughout host-guest chemistry literature as an effective means of polarizing XB donors while maintaining a charge neutral molecule.10,19–21 Finally, the nitrogen atom in the quinoline backbone enables the preorganization of the molecule into a hydrogen bonded 6- membered ring (shown in blue in Figure 5.1) with the guanidino HB donor group, which would reduce the entropic penalty associated with maintaining the conformation predicted to be most favorable for binding.22 The synthesis for receptor 2 (Scheme 5.2) began with the protection of the amine functional group on the 2-aminoquinoline by acetyl chloride under basic conditions to afford 2.1. After this, the lone pair on the nitrogen atom in the amino group coordinated n-BuLi for a regioselective iodination in the 3-position of the quinoline intermediate,23 before then undergoing deprotection by refluxing in acid overnight. 2-Amino 3- 54 iodoquinoline (intermediate 2.3) then is subjected to a Sonogashira cross coupling with the selectively deprotected 1,4-bis(trimethylsilyl)butadiyne (2.3a). The outermost diyne relative to the quinoline is then iodinated with n-iodosuccinimide, which undergoes a click reaction with benzyl azide to furnish the XB triazole ring in 2.6. Finally, the HB donor is added via either a cyanamide or isocyanate addition. Precursors 2.1-2.4 were synthesized and characterized completely and, while the latter intermediates and receptor 2.5-2 were synthesized, they were not fully isolated and characterized due to time constraints. Synthesis of 2 was confirmed by high resolution MS; more details on all of these molecules’ properties can be found in Appendix D. O Cl 1. nBuLi TEA O 2. I I HCl I2 O N NH2 DCM .t. N N 1. Et2O, N2 Dioxane 0 °C to r -70 °C to 0 °C N N r.t. then reflux N NH2 H 2. -70 °C to -50 °C H 98% 2.1 73% 2.2 96% 2.3 TMS 1. MeLi•LiBr H2. NH4Cl TMS Et2O, RT TMS 2.3a Quant. TMS I I 1. AgFaq, 20 minPd(PPh3)4, CuI 2. NIS, 1h N NH2 THF, TEA N NH H2O, ACN, r.t. N NH 89% 2 31% 2 2.3 2.4 2.5 I N N N N 4 mol% Lutidine, CuI(PPh3)3, N N10 mol % Sc(OTf) N 33Bn H2NCN I I DMF H2O, 100° C N NH N NH2 66% N NH2 H N NH2 2.5 2.6 2 Scheme 5. 2 Synthetic scheme for derivatives of receptor 2 (R = benzyl). 5 .5. Computation Work with Library of Receptors. Although empirical tests were not performed with receptor 2, computational work by Michael Miller was used to determine the energy levels Figure 5.3 Family of one arm receptors discussed in associated with σ-hole formation in nearly all the receptors chapter 4 and Table 5.3 55 discussed in this thesis for direct comparisons between the systems. All calculations are done with Gaussian 09, using B97D3 as the level of theory and def2tzvp as the basis set. These are listed in Table 5.3 and a variety of observations can be made from them. First, as established in Chapter 4, it is clear the primary driver behind XB donor polarization in the family of one-armed receptors (easily referenced in Figure 5.3) is the methylation of the pyridine moiety; the cationic hosts are always more than 50 kcal/mol stronger than their neutral analogs. This is particularly notable because it suggests that the IUPAC definition of the XB interaction (discussed in detail in Chapter 1) might be biasing host guest chemists to ascribe an undue significance to place EWG groups as near as possible to the XB donor. In the family of one arm hosts the nearest EWG (the pendant R-groups listed in Table 5.3) are much less effective at forming the σ-hole needed for XB donation when compared to the methylation of the nitrogen atom on the other side of the molecule. The complexity of effectively polarizing the XB donors is further emphasized by the results of the calculations done on receptor 1, in which after methylation of the much closer nitrogen atom, the σ-hole value actually decreases by 4 kcal/mol. This result is surprising to us given the previously noted influence of the positive charge on the σ-hole. To verify and study this result, we would perform 1H NMR spectroscopy titrations to be done with the neutral version of receptor 1 (labelled 1.5 in Scheme 5.1) to determine binding constants and compare with the methylated receptor. Finally, the calculated σ- hole value for the quinoline receptor shows that the triazole moiety will elicit similar energy levels comparable to an aryl ring with an appended EWG. Although further 56 empirical and computation data must be collected to reinforce this idea, the suggestion that electron-deficient aryl rings can be used as comparable drivers of XB donors, relative to the much more regularly used triazole units, could open the door to novel XB host designs in future receptors. Table 5.3. Summary of the calculated energy value of sigma holes for XB donors in all of the receptors discussed in Chapters 4 and 5 (first 8 entries refer to the one arm family in Chapter 4) Chapter Number Receptor Name Sigma Hole Energy (kcal/mol) 4 4a Neutral SO2CF3 28.5 4 4b Neutral CF3 24.3 4 4c Neutral H 19.0 4 4d Neutral t-Bu 17.5 4 5a Methylated SO2CF3 82.8 4 5b Methylated CF3 80.0 4 5c Methylated H 77.2 4 5d Methylated t-Bu 73.8 5 2 Quinoline Receptor 26.8 5 1 Neutral 11.8 5 1.5 Methylated 7.1 5.6. Conclusion. The receptors presented in this chapter highlight the complexity of designing an effective XB donor. Despite the nuance involved with XB host design, the latest generation of DWJ-Haley XB-HB hybrid receptors, like 1 (and those discussed in Chapter 4), have demonstrated a noteworthy ability to bind Cl– in polar solvents. These will undoubtedly pave the way towards future DWJ-Haley receptors and their use in aqueous media, fulfilling a long sought-after goal of binding these important anion species in critical environments, such as in vivo and in aqueous ecological environments. 57 CHAPTER SIX CONCLUSIONS AND OUTLOOKS The research in this thesis discussed many of the most recent XB and HB receptors to come out of the DWJ and Haley laboratories for anion recognition. Chapter Two begins by exploring modifications made to improve the quantum yield, and other photophysical properties of a relatively newly discovered P-N containing heterocyclic fluorophore. This family of molecules beautifully highlights the nuanced understanding that can be derived from thorough structure-property relationships. It also sets the groundwork for thinking about many of the other fluorescent molecules discussed throughout this thesis. In Chapter Three the second XB receptor to come out of the DWJ and Haley group is discussed. This receptor was designed with two goals in mind: first, binding the hydrochacogenide, HS- and second, competing in a significant way with polar media. While falling short of these goals the failed hypothesis did reinforce fundamental design principles suggested by the first DWJ and Haley XB receptor (also discussed in Chapter Three). These design principles were then used to conceptualize and synthesize the much more successful receptors that are the subject of Chapters Four and Five. In Chapter Four a family of one-armed receptors capable of binding the halide anions in the polar solvent acetonitrile are discussed. In this family the size of the pocket is sufficiently shrunk to flip the preference of the XB guest to bind the smaller halide (Cl- ) over that of the larger halides (Br- or I-). This family demonstrated some HB 58 coordination of these anions, making it the first hybrid receptor to be published by the DWJ and Haley labs. Finally, In Chapter Five, two more hybrid receptors are explored. The first is a slightly larger one-arm receptor that demonstrate a similar preference for the smaller halide (Cl-) with an even larger association constant in more competitive solvent conditions (10% DMSO in ACN). It also was the subject of relatively surprising computation work, which suggested that methylating the pyridine ring (on which the XB donor is appended) had a minimal effect on the size of the σ-hole, bucking the outsized effect that methylation had with the earlier XB receptors. This result begs much further investigation. The small pocket of this receptor could also be an interesting case study in the effect of changing the identity of the XB donor (namely, from I to Br) and how that effects the halide guest preference, given a slightly less crowded pocket. Also mentioned in Chapter Five is the synthesis of a quinoline based scaffold that combines many of the design principles explored through the rest of the thesis. It has fluorescent properties, can preorganize into the state hypothesized to be best for binding and could be synthesized in either a neutral or cationic motif, depending upon the pH the receptor is intended to be utilized at. Further synthesis and characterization remain to be done with the quinoline receptor and should be finished shortly. The receptors discussed in Chapters Four and Five, in particular, present a library of receptors which seem to buck the IUPAC definition of a halogen bond as a simple interaction between R—X···Y, where X is a polarizable halogen covalently bound to an R group and Y is an electronically-dense halogen bond acceptor. This definition biases supramolecular chemists to give undue important to the electron withdrawing nature of 59 groups closest to the XB donor. In Chapter 4 it is shown that the much closer functional groups (which ranged from highly electron withdrawing to slightly electron donating) had a meager effect on XB polarization, relative to the methylation of the pyridine ring on the otherside of the molecule. But, as evidence by the receptors explored in Chapter 5, the answer is not as simple as adding a cation charge anywhere on the XB donor molecule. On the contrary, despite its signifigantly closer position to the XB donor in this receptor the methylation of the pyridine ring served to only marginally increase the size of the σ-hole. In conclusion, anions are an incredibly important class of molecule that are ubiquitious in human physiology and ecological niches around the world. Unfortunately, given their signifigance, the environments in which they are most pertinent (polar, aqueous ones) are also the mediums in which anions are the hardest to pin down. Their low charge density, high solvation energy and pH sensitivy makes them more difficult to bind than their cationic counterparts in water. Water’s ability to act as both an HB donor and acceptor makes it a highly competitive solvent for any synthetic host looking to bind anions. This makes XB an incredibly attractive tool, as study after study has shown that these σ-hole interactions outcompete the HB counterparts in competitive polar media in otherwise analogous host receptors. Despite this promise, and the growing interest of the past two decades, much still remains to be discovered about the best way to polarize and utilize XB donors for anion coordination. This, along with puzzling out how to best use this tool in tandem with other recognition techniques, including the other σ-hole interactions (chalcogen and pnictogen bonding) or anion···π interactions, is sure to drive research interests for decades to come. 60 APPENDIX A SUPPLEMENTARY CONTENT FOR CHAPTER TWO Experimental Details 1. Crystallographic Data for 2f General. Diffraction intensities for 2f were collected at 173 K on a Bruker Apex2 CCD diffractometer using CuK radiation, = 1.54178 Å. Space group was determined based on systematic absences. Absorption correction was applied by SADABS.1 Structure was solved by direct methods and Fourier techniques and refined on F2 using full matrix least-squares procedures. All non-H atoms were refined with anisotropic thermal parameters. H atoms were refined in calculated positions in a rigid group model except the H atom at the N atom involved in H-bond which was found on the residual density map and refined with isotropic thermal parameter. In the crystal structure the main molecules form a dimer unit via N-H•••O H-bonds which are packed in columns. Solvent molecules CHCl3 fill out empty space between such the columns and are highly disordered. These disordered solvent molecules were treated by SQUEEZE.2 The corrections of the X-ray data by SQUEEZE are 220 electron/cell; the required values are 232 electron/cell for four CHCl3 molecules in the full unit cell. All calculations were performed by the Bruker SHELXL-2014 package.3 Crystallographic Data for 2f: C24H20Cl3O2P, M = 505.74, 0.15 x 0.08 x 0.06 mm, T = 173(2) K, Monoclinic, space group P21/c, a = 17.7865(8) Å, b = 8.1543(4) Å, c = 18.4012(8) Å,  = 117.199(3), V = 2373.7(2) Å3, Z = 4, Dc = 1.415 Mg/m3, μ(Cu) = 4.335 mm–1, F(000) = 1040, 2θmax = 133.15°, 13893 reflections, 4116 independent reflections [Rint = 0.0465], R1 = 0.0570, wR2 = 0.1562 and GOF = 1.021 for 4116 reflections (257 parameters) with I>2(I), R1 = 0.0699, wR2 = 0.1630 and GOF = 1.021 for all reflections, max/min residual electron density +0.588/–0.275 eÅ–3. CCDC 1944053. 61 Figure A.1. ORTEP drawing of 2f with selected bond lengths listed; thermal ellipsoids drawn at 30% probability. Crystals grown with slow diffusion of pentane into a solution of 2 in CHCl3. Figure A.2. ORTEP drawing of the racemic dimer of 2f; thermal ellipsoids drawn at 30% probability. 62 Figure A.3. ORTEP drawing of the crystal packing of 2f looking down the b-axis. 63 2. Frontier Orbital Values and TD-DFT Excitations Ground state calculations Table A.1. Calculated Frontier Orbitals and First Excitation Values for 2a S0 → S1 cmpd EHOMO ELUMO EDFT computed (nm), osc. Strength 2a –6.800 –2.405 4.395 341, 0.505 2b –6.750 –2.319 4.431 337, 0.535 2c –6.531 –2.296 4.235 354, 0.392 2d –6.510 –2.294 4.303 348, 0.486 2e –6.481 –2.178 4.207 358, 0.361 2f –6.353 –2.146 3.896 389, 0.195 a Calculated at the PBE0/TZVP level of theory; values reported in eV. Figure A.4. Frontier orbital shapes (isovalue = 0.3) of heterocycles 2; calculated at the PBE0/TZVP level of theory. Excited state calculations These initial structures were optimized using the functional PBE04,5 (25% full-range HF exchange) and TZVP basis set6 as implemented in Gaussian 09.7 In addition, all the optimized structures were confirmed by frequency analysis and the number of imaginary frequencies was zero. TD-DFT vertical excitation calculations and geometry optimization of the first excited state (S1) were performed at the same level of theory. The PCM solvation model8,9 was used to account for the solvent effects of the chloroform. 64 Figure A.5. HOMO-LUMO pictorial representations (isovalue = 0.3), energies and their differences (ΔE) in the optimized S0 and S1 structures of 2f calculated by the DFT and TD-DFT methods at the PCM(CHCl3)-PBE0/TZVP level of theory, respectively. 65 3. Geometry Optimizations and Coordinates Ground state geometries Table A.2. Cartesian coordinates of the optimized geometry of 2a determined at the PBE0/TZVP level of theory Atom x y z C -4.85224 -0.99453 0.24314 C -5.06467 0.36876 -0.01362 C -3.99796 1.19772 -0.26557 C -2.68944 0.69768 -0.26948 C -2.46321 -0.67090 -0.00881 C -3.55816 -1.49676 0.24538 N -1.62915 1.54068 -0.48129 P -0.05268 1.09981 -0.96544 C 0.01373 -0.57214 -0.28708 C -1.13106 -1.21587 0.02816 C 1.32888 -1.22391 -0.17733 C 2.32937 -1.00360 -1.13093 C 3.55719 -1.62907 -1.02594 C 3.81064 -2.48996 0.04330 C 2.82419 -2.71342 1.00652 C 1.60209 -2.08011 0.89538 O 0.24116 1.20873 -2.43673 H -6.07224 0.76605 -0.01630 H -4.16680 2.24964 -0.46799 H -3.38832 -2.54835 0.44662 H -1.86453 2.51486 -0.62006 H -1.06114 -2.25919 0.32868 H 2.13135 -0.35505 -1.97675 H 4.32091 -1.45850 -1.77493 H 3.02505 -3.37115 1.84342 H 0.85418 -2.23445 1.66438 C 5.07615 -3.13557 0.15614 N 6.10029 -3.65817 0.24771 C 1.00594 2.19506 0.01314 C 1.97814 2.94304 -0.64517 66 C 0.86604 2.29013 1.39720 C 2.80595 3.78889 0.08033 H 2.07444 2.85806 -1.72158 C 1.69724 3.13388 2.11728 H 0.10786 1.70840 1.91080 C 2.66562 3.88319 1.45833 H 3.56138 4.37440 -0.43113 H 1.59006 3.20904 3.19339 H 3.31411 4.54305 2.02388 C -5.95519 -1.85468 0.50419 N -6.85123 -2.54976 0.71725 Zero-point correction = 0.306046 (Hartree/Particle); Thermal correction to energy = 0.328311; Thermal correction to enthalpy = 0.329256; Thermal correction to Gibbs free energy = 0.252414; Sum of electronic and zero-point energies = -1426.281789; Sum of electronic and thermal energies = -1426.259523; Sum of electronic and thermal enthalpies = -1426.258579; Sum of electronic and thermal free energies = -1426.335421. Table A.3. Cartesian coordinates of the optimized geometry of 2b determined at the PBE0/TZVP level of theory Atom x y z C 4.33256 0.25239 0.07266 C 4.36740 -1.12574 -0.14433 C 3.19680 -1.82533 -0.36097 C 1.96424 -1.16425 -0.36534 C 1.91814 0.22872 -0.13593 C 3.11747 0.91433 0.07864 N 0.79435 -1.87023 -0.54497 P -0.71412 -1.24659 -1.00223 C -0.55995 0.42532 -0.35685 C 0.66485 0.93655 -0.08900 C -1.78476 1.23382 -0.22255 C -2.83620 1.11740 -1.14210 C -3.97972 1.88534 -1.01112 C -4.09599 2.78768 0.05013 67 C -3.05815 2.90828 0.97984 C -1.92102 2.13323 0.84339 O -1.06879 -1.33987 -2.44972 H 5.31560 -1.64935 -0.14765 H 3.22816 -2.89645 -0.53066 H 3.08031 1.98498 0.25349 H 0.90195 -2.87056 -0.65853 H 0.72898 1.98857 0.18318 H -2.74143 0.43697 -1.98120 H -4.78302 1.79523 -1.73288 H -3.15443 3.59763 1.81044 H -1.13530 2.20738 1.58703 C -5.27351 3.57973 0.18954 N -6.22685 4.22181 0.30316 C -1.87332 -2.16682 0.02964 C -2.99233 -2.74612 -0.56830 C -1.67521 -2.28375 1.40740 C -3.90785 -3.44425 0.21021 H -3.13095 -2.64787 -1.63956 C -2.59376 -2.97976 2.18044 H -0.80363 -1.83295 1.87151 C -3.70888 -3.55955 1.58133 H -4.77702 -3.89756 -0.25393 H -2.44117 -3.07162 3.25034 H -4.42561 -4.10321 2.18781 C 5.58540 1.02927 0.31687 F 5.71872 2.05264 -0.54614 F 6.68573 0.27257 0.20460 F 5.60776 1.57440 1.54852 Zero-point correction = 0.312616 (Hartree/Particle); Thermal correction to energy = 0.335712; Thermal correction to enthalpy = 0.336656; Thermal correction to Gibbs free energy = 0.257793; Sum of electronic and zero-point energies = -1670.857410; Sum of electronic and thermal energies = -1670.834315; Sum of electronic and thermal enthalpies = -1670.833370; Sum of electronic and thermal free energies = -1670.912234. 68 Table A.4. Cartesian coordinates of the optimized geometry of 2c determined at the PBE0/TZVP level of theory Atom x y z C 4.76530 -0.84753 -0.21352 C 4.95235 0.51406 0.01777 C 3.85707 1.32309 0.25213 C 2.56393 0.79136 0.25768 C 2.37737 -0.58478 0.01677 C 3.49994 -1.39313 -0.21621 N 1.46788 1.61180 0.44837 P -0.06873 1.13040 0.96633 C -0.10263 -0.53610 0.28870 C 1.05876 -1.16265 -0.01869 C -1.40505 -1.21599 0.17783 C -2.41834 -1.00066 1.12223 C -3.63596 -1.64912 1.01591 C -3.86748 -2.52873 -0.04558 C -2.86886 -2.74659 -1.00019 C -1.65681 -2.09032 -0.88811 O -0.35888 1.22487 2.42868 H 5.95108 0.93373 0.01736 H 4.00140 2.38257 0.43708 H 3.36185 -2.45261 -0.39989 H 1.67299 2.59782 0.54848 H 1.01106 -2.21043 -0.31023 H -2.23659 -0.33867 1.96189 H -4.40883 -1.48285 1.75717 H -3.05318 -3.41765 -1.83095 H -0.90062 -2.23953 -1.65088 C -5.12144 -3.19771 -0.15958 N -6.13694 -3.74019 -0.25236 C -1.17198 2.18283 0.00032 C -2.19830 2.86079 0.65770 C -1.02047 2.31216 -1.38225 C -3.06654 3.67011 -0.06545 H -2.30167 2.75103 1.73180 69 C -1.89224 3.11874 -2.10022 H -0.22075 1.78506 -1.89284 C -2.91380 3.79763 -1.44141 H -3.86269 4.20063 0.44554 H -1.77547 3.22002 -3.17380 H -3.59348 4.42828 -2.00484 Cl 6.14889 -1.86580 -0.50514 Zero-point correction = 0.298482 (Hartree/Particle); Thermal correction to energy = 0.320018; Thermal correction to enthalpy = 0.320963; Thermal correction to Gibbs free energy = 0.245463; Sum of electronic and zero-point energies = -1793.507492; Sum of electronic and thermal energies = -1793.485956; Sum of electronic and thermal enthalpies = -1793.485012; Sum of electronic and thermal free energies = -1793.560511. Table A.5. Cartesian coordinates of the optimized geometry of 2d determined at the PBE0/TZVP level of theory Atom x y z C 4.87957 -2.33820 -0.55510 C 5.40340 -1.07460 -0.27234 C 4.56839 -0.01982 0.04974 C 3.18339 -0.20570 0.09586 C 2.63866 -1.47362 -0.19415 C 3.51235 -2.52485 -0.51570 N 2.34180 0.85710 0.37539 P 0.75066 0.75975 0.93800 C 0.26395 -0.80418 0.19669 C 1.21640 -1.68909 -0.18898 C -1.17160 -1.12470 0.11863 C -2.06875 -0.69107 1.10523 C -3.41442 -1.00319 1.02777 C -3.89511 -1.75842 -0.04594 C -3.01376 -2.19077 -1.04236 C -1.67175 -1.86848 -0.95930 O 0.53974 0.85819 2.41437 H 6.47569 -0.91258 -0.30076 H 4.98133 0.95935 0.27178 70 H 3.08676 -3.49915 -0.73586 H 2.80016 1.74837 0.51685 H 0.89018 -2.67426 -0.51917 H -1.69940 -0.12656 1.95449 H -4.09666 -0.67053 1.80128 H -3.38796 -2.76407 -1.88250 H -1.00217 -2.17782 -1.75409 C -5.28099 -2.08196 -0.12992 N -6.40367 -2.34505 -0.19896 C -0.07738 2.10620 0.06374 C -0.88916 2.97892 0.78798 C 0.06664 2.26802 -1.31618 C -1.55170 4.01136 0.13447 H -0.98883 2.84211 1.85939 C -0.59871 3.29914 -1.96466 H 0.70009 1.58949 -1.87873 C -1.40739 4.16985 -1.23917 H -2.18113 4.69171 0.69792 H -0.48711 3.42552 -3.03617 H -1.92673 4.97494 -1.74843 H 5.53887 -3.16167 -0.80375 Zero-point correction = 0.308146 (Hartree/Particle); Thermal correction to energy = 0.328403; Thermal correction to enthalpy = 0.329347; Thermal correction to Gibbs free energy = 0.257198; Sum of electronic and zero-point energies = -1334.037270; Sum of electronic and thermal energies = -1334.017013; Sum of electronic and thermal enthalpies = -1334.016068; Sum of electronic and thermal free energies = -1334.088218. Table A.6. Cartesian coordinates of the optimized geometry of 2e determined at the PBE0/TZVP level of theory Atom x y z C 4.35248 0.21049 0.07235 C 4.33995 -1.17064 -0.16139 C 3.16166 -1.86961 -0.36931 C 1.93366 -1.21055 -0.35366 C 1.90847 0.17770 -0.11411 71 C 3.12016 0.85067 0.09372 N 0.74735 -1.90753 -0.51892 P -0.74060 -1.26142 -0.99010 C -0.56897 0.40515 -0.33619 C 0.66375 0.89745 -0.06005 C -1.78056 1.23340 -0.20894 C -2.83472 1.12515 -1.12718 C -3.96518 1.91354 -1.00623 C -4.06710 2.83118 0.04353 C -3.02744 2.94420 0.97248 C -1.90413 2.14776 0.84649 O -1.09672 -1.34365 -2.43972 H 5.27024 -1.72536 -0.18728 H 3.19132 -2.94032 -0.54750 H 3.06805 1.91990 0.27852 H 0.84423 -2.90826 -0.63491 H 0.73924 1.94775 0.21772 H -2.75058 0.43374 -1.95842 H -4.76930 1.82839 -1.72779 H -3.11227 3.64495 1.79488 H -1.11817 2.21716 1.59035 C -5.22999 3.64586 0.17141 N -6.17152 4.30702 0.27540 C -1.92693 -2.16603 0.02901 C -3.05613 -2.71550 -0.57772 C -1.73891 -2.29931 1.40660 C -3.99141 -3.39891 0.19059 H -3.18671 -2.60599 -1.64895 C -2.67669 -2.98045 2.17009 H -0.85902 -1.87219 1.87731 C -3.80214 -3.52974 1.56168 H -4.86839 -3.82863 -0.28143 H -2.53109 -3.08441 3.23994 H -4.53410 -4.06162 2.16049 C 5.63766 1.00664 0.30357 C 5.74841 2.11833 -0.75049 H 5.77949 1.69730 -1.75979 72 H 4.90443 2.81180 -0.70109 H 6.66472 2.69643 -0.59303 C 6.88485 0.12698 0.20409 H 6.88632 -0.66501 0.95924 H 6.97790 -0.33603 -0.78307 H 7.77572 0.74011 0.36767 C 5.60079 1.63565 1.70446 H 5.53400 0.86334 2.47670 H 6.51095 2.21793 1.88161 H 4.74564 2.30666 1.82388 Zero-point correction = 0.420029 (Hartree/Particle); Thermal correction to energy = 0.445957; Thermal correction to enthalpy = 0.446901; Thermal correction to Gibbs free energy = 0.362564; Sum of electronic and zero-point energies = -1491.024374; Sum of electronic and thermal energies = -1490.998446; Sum of electronic and thermal enthalpies = -1490.997502; Sum of electronic and thermal free energies = -1491.081839. Table A.7. Cartesian coordinates of the optimized geometry of 2f determined at the PBE0/TZVP level of theory Atom x y z C 4.47588 -0.49543 -0.11516 C 4.55087 0.88383 0.10641 C 3.39489 1.62204 0.31345 C 2.14145 1.01628 0.30167 C 2.05466 -0.37183 0.06988 C 3.22937 -1.10406 -0.13277 N 0.98169 1.76709 0.45969 P -0.51584 1.18392 0.96994 C -0.42817 -0.48686 0.30598 C 0.77737 -1.03553 0.01956 C -1.67841 -1.25610 0.18321 C -2.71804 -1.10476 1.11169 C -3.88606 -1.83701 0.99485 C -4.04100 -2.73965 -0.06140 C -3.01604 -2.89482 -1.00048 C -1.85444 -2.15464 -0.87810 73 O -0.83990 1.26621 2.42738 H 5.50541 1.39434 0.12376 H 3.46976 2.69084 0.48664 H 3.17401 -2.17336 -0.30815 H 1.11904 2.76509 0.55595 H 0.80395 -2.08652 -0.26323 H -2.59360 -0.42479 1.94740 H -4.67904 -1.71914 1.72410 H -3.14145 -3.58368 -1.82771 H -1.07896 -2.25603 -1.62926 C -5.24310 -3.49585 -0.18623 N -6.21661 -4.10928 -0.28810 C -1.68565 2.14206 -0.01896 C -2.78685 2.71979 0.61264 C -1.51424 2.28933 -1.39736 C -3.71017 3.44529 -0.13109 H -2.90542 2.59902 1.68402 C -2.44008 3.01266 -2.13633 H -0.65624 1.84030 -1.88745 C -3.53719 3.59020 -1.50293 H -4.56501 3.89692 0.36070 H -2.30706 3.12759 -3.20671 H -4.25970 4.15509 -2.08268 O 5.54569 -1.30193 -0.32098 C 6.84268 -0.72057 -0.29182 C 7.84268 -1.82535 -0.52364 H 7.01082 -0.23888 0.68006 H 6.92205 0.04825 -1.07137 H 8.85712 -1.41944 -0.50995 H 7.76385 -2.58608 0.25638 H 7.67376 -2.30144 -1.49231 Zero-point correction = 0.368617 (Hartree/Particle); Thermal correction to energy = 0.393025; Thermal correction to enthalpy = 0.393970; Thermal correction to Gibbs free energy = 0.312175; Sum of electronic and zero-point energies = -1487.682390; Sum of electronic and thermal energies = -1487.657981; Sum of electronic and thermal enthalpies = -1487.657037; Sum of electronic and thermal free energies = -1487.738831. Representative excited state geometry 74 Table A.8. Cartesian coordinates for compound 2f (S1) Atom x y z C 4.19754 0.83011 0.04258 C 4.43859 -0.30536 -0.75296 C 3.34960 -0.99099 -1.24675 C 2.02893 -0.57501 -0.98158 C 1.76910 0.59205 -0.18219 C 2.87518 1.25433 0.31293 N 0.98402 -1.27177 -1.48236 P -0.70551 -1.07234 -1.17069 C -0.75456 0.54008 -0.40889 C 0.43970 1.06313 0.08325 C -2.03546 1.16742 -0.22284 C -3.26152 0.49519 -0.47023 C -4.47749 1.09997 -0.26071 C -4.54758 2.42475 0.21555 C -3.34562 3.11750 0.45118 C -2.13156 2.50968 0.23727 O -1.47016 -1.33310 -2.42262 H 5.44102 -0.64467 -0.97311 H 3.50091 -1.87627 -1.85654 H 2.75354 2.14216 0.92323 H 1.20261 -2.08660 -2.04452 H 0.38969 1.94048 0.71954 H -3.25594 -0.52378 -0.83955 H -5.39269 0.55342 -0.45885 H -3.38120 4.14509 0.79654 H -1.23161 3.09014 0.40068 C -5.79638 3.04886 0.44241 N -6.82004 3.56105 0.62818 C -1.02320 -2.31567 0.10097 C -1.88954 -3.36767 -0.20702 C -0.43713 -2.24196 1.36922 C -2.16463 -4.33891 0.74848 H -2.34218 -3.41212 -1.19192 75 C -0.71523 -3.21626 2.31544 H 0.22229 -1.41627 1.61968 C -1.57745 -4.26674 2.00635 H -2.84004 -5.15319 0.50835 H -0.26328 -3.15571 3.29972 H -1.79397 -5.02592 2.75041 O 5.14428 1.57982 0.58991 C 6.52855 1.25640 0.38195 C 7.34720 2.28195 1.12018 H 6.73510 1.27453 -0.69282 H 6.71378 0.24638 0.76081 H 8.40933 2.06646 0.98366 H 7.14692 3.28556 0.73949 H 7.12518 2.25966 2.18909 Geometries of 2 meso-dimer and analogous 1 meso-dimer Table A.9. Cartesian coordinates for compound 2f·2f (meso-dimer) Atom x y z P -1.95724 0.60235 -0.15810 O -1.13426 0.91794 1.07453 O -4.86819 -5.45211 -0.32378 N -1.72792 -0.90711 -0.85743 N -6.92721 6.54145 1.19270 C -3.73766 0.57706 0.11224 C -4.39950 -0.59867 0.11116 H -5.45545 -0.58653 0.37173 C -3.84837 -1.89655 -0.20763 C -2.54204 -2.01925 -0.69556 C -2.07237 -3.28570 -1.07125 H -1.07359 -3.37896 -1.48229 C -2.86947 -4.39809 -0.93597 H -2.50799 -5.37646 -1.22553 76 C -4.17029 -4.28817 -0.42388 C -4.65455 -3.04354 -0.07520 H -5.65812 -2.91755 0.30781 C -4.41821 1.86371 0.36187 C -5.70140 2.08442 -0.14977 H -6.17936 1.31735 -0.74569 C -6.35626 3.28249 0.06722 H -7.34684 3.44695 -0.33573 C -5.72281 4.29106 0.79408 C -4.43772 4.09287 1.29858 H -3.95194 4.87940 1.86064 C -3.79259 2.88822 1.08154 H -2.80197 2.72981 1.49055 C -6.39194 5.54025 1.01540 C -1.60300 1.78258 -1.48582 C -0.61820 2.74784 -1.31084 H -0.06610 2.78759 -0.38140 C -0.34559 3.65095 -2.33102 H 0.42602 4.39841 -2.19596 C -1.06490 3.59304 -3.51740 H -0.85633 4.29959 -4.31093 C -2.05424 2.62939 -3.69172 H -2.61274 2.58634 -4.61790 C -2.32146 1.72118 -2.67885 H -3.08473 0.96234 -2.81326 C -6.18694 -5.38880 0.20600 H -6.80431 -4.74129 -0.42516 H -6.15489 -4.95833 1.21243 77 C -6.73651 -6.79556 0.23574 H -7.74918 -6.78877 0.63901 H -6.76584 -7.21498 -0.76966 H -6.11603 -7.43412 0.86410 H -0.75384 -1.07946 -1.13391 P 1.95156 -0.64866 0.19274 O 1.08688 -0.96110 -1.01172 O 4.69506 5.48289 0.53051 N 1.68726 0.83535 0.93440 N 7.06869 -6.35013 -1.57736 C 3.71838 -0.55160 -0.13926 C 4.34473 0.64290 -0.10761 H 5.39135 0.67402 -0.40224 C 3.76786 1.90828 0.28504 C 2.47046 1.97330 0.80554 C 1.97625 3.20980 1.24574 H 0.98434 3.25755 1.68084 C 2.74001 4.34850 1.13996 H 2.35923 5.30349 1.47902 C 4.03159 4.29657 0.59628 C 4.54060 3.08158 0.18481 H 5.53862 2.99783 -0.22352 C 4.43069 -1.80350 -0.46702 C 5.73087 -2.00838 0.00685 H 6.19822 -1.25657 0.63004 C 6.41634 -3.17373 -0.28147 H 7.41988 -3.32672 0.09277 C 5.79805 -4.16492 -1.04428 78 C 4.49709 -3.98263 -1.51265 H 4.02348 -4.75542 -2.10340 C 3.82065 -2.81055 -1.22348 H 2.81680 -2.66345 -1.60302 C 6.50213 -5.37893 -1.34047 C 1.69613 -1.88524 1.49285 C 0.64469 -2.78840 1.38896 H -0.01285 -2.74941 0.53080 C 0.44143 -3.73291 2.38765 H -0.38232 -4.43095 2.30814 C 1.29730 -3.78013 3.48005 H 1.14343 -4.52010 4.25539 C 2.35300 -2.87906 3.58311 H 3.01798 -2.91732 4.43630 C 2.55094 -1.92835 2.59314 H 3.36803 -1.21946 2.67405 C 6.00897 5.47632 -0.01461 H 6.65036 4.82286 0.58592 H 5.97823 5.08342 -1.03618 C 6.51872 6.89823 0.00350 H 7.52896 6.93518 -0.40423 H 6.54128 7.28223 1.02310 H 5.87721 7.54089 -0.59909 H 0.70937 0.98002 1.21499 Table A.10. Cartesian coordinates for compound 2f‘·2f‘ (meso-dimer for -OPh analogue) Atom x y z P 1.92141 0.79131 -0.26147 O 0.78543 1.51236 -0.92423 79 O 6.53956 -3.99151 -1.19988 N 2.02262 -0.82536 -0.62475 N 4.97709 8.03803 0.16496 C 3.59102 1.36451 -0.52551 C 4.57382 0.46252 -0.73510 H 5.58336 0.84492 -0.86408 C 4.43734 -0.97249 -0.81836 C 3.18131 -1.58317 -0.74275 C 3.09026 -2.97766 -0.82535 H 2.11643 -3.44875 -0.76712 C 4.22370 -3.74153 -0.97276 H 4.15972 -4.82045 -1.03252 C 5.48901 -3.14105 -1.05234 C 5.58959 -1.76691 -0.97746 H 6.54735 -1.26819 -1.03595 C 3.87197 2.80945 -0.37778 C 5.03179 3.22763 0.28458 H 5.69956 2.49146 0.71451 C 5.32555 4.57143 0.42127 H 6.22247 4.88754 0.93716 C 4.44618 5.52323 -0.09631 C 3.27888 5.12301 -0.74537 H 2.60112 5.86640 -1.14344 C 2.99450 3.77540 -0.88166 H 2.09196 3.46511 -1.39128 C 4.74073 6.91971 0.04861 C 7.83932 -3.42380 -1.31227 H 8.06387 -2.84264 -0.41180 H 7.86840 -2.74857 -2.17381 C 8.82206 -4.55834 -1.48203 H 9.83191 -4.16050 -1.58264 80 H 8.79387 -5.22113 -0.61745 H 8.58521 -5.13637 -2.37497 H 1.13240 -1.32490 -0.52414 P -1.83870 -0.83501 -0.24159 O -0.58006 -1.54000 0.16537 O -6.03526 4.24836 0.94534 N -1.79480 0.81008 -0.03092 N -4.96926 -7.90305 1.34215 C -3.39028 -1.28890 0.51650 C -4.28646 -0.31809 0.79650 H -5.24232 -0.62508 1.21435 C -4.11342 1.10510 0.61877 C -2.88448 1.63685 0.21450 C -2.74464 3.02402 0.08792 H -1.78912 3.43301 -0.21776 C -3.81060 3.85580 0.33656 H -3.71118 4.92879 0.23273 C -5.05144 3.33451 0.73214 C -5.19474 1.96973 0.87834 H -6.13228 1.53222 1.19349 C -3.71128 -2.72292 0.68277 C -4.99895 -3.18586 0.38912 H -5.74136 -2.49828 0.00330 C -5.33121 -4.51660 0.56144 H -6.32798 -4.86844 0.33091 C -4.36480 -5.41187 1.02114 C -3.07251 -4.96965 1.30158 H -2.32861 -5.67016 1.65720 C -2.74907 -3.63508 1.12870 H -1.74774 -3.29074 1.35074 C -4.70057 -6.79521 1.19888 81 C -7.30470 3.76745 1.37118 H -7.69766 3.06517 0.62875 H -7.19240 3.23607 2.32210 C -8.21923 4.96014 1.52189 H -9.20286 4.63183 1.85789 H -8.33183 5.47687 0.56903 H -7.81697 5.65942 2.25452 H -0.90390 1.25228 -0.29356 O 1.65012 0.93158 1.34869 O -2.03326 -1.14197 -1.84182 C 2.42198 0.19621 2.24001 C 3.51657 0.80032 2.83895 C 2.06499 -1.11320 2.52760 C 4.27398 0.07137 3.74738 H 3.75515 1.82865 2.59475 C 2.83279 -1.83172 3.43459 H 1.19713 -1.54270 2.04046 C 3.93610 -1.24381 4.04341 H 5.12751 0.53465 4.22565 H 2.56580 -2.85461 3.66811 H 4.52892 -1.80929 4.75059 C -3.06789 -0.51671 -2.52823 C -4.25761 -1.20300 -2.71875 C -2.88728 0.76802 -3.01906 C -5.29228 -0.58260 -3.40703 H -4.35624 -2.20940 -2.33039 C -3.93080 1.38009 -3.70142 H -1.94065 1.27030 -2.86233 C -5.13323 0.70995 -3.89317 H -6.22390 -1.11125 -3.56359 H -3.80062 2.38351 -4.08590 82 H -5.94271 1.19250 -4.42543 83 4. Photophysical Properties in Different Solvents Figure A.6. Fluorescence decay curve of 2d vs Ludox Prompt sample. Solvent effects on fluorescence Figure A.7. Emission spectra of 2a in various solvents. 84 Table A.11. Photophysical properties of 2a in various solvents Cyclohexane Toluene Chloroform DCM THF Acetonitrile λabs 353 351 348 349 358 350 λem 445 446 452 448 456 456 Stokes Shift (nm) 92 95 104 99 98 106 Figure A.8. Emission spectra of 2b in various solvents. Table A.12. Photophysical properties of 2b in various solvents Cyclohexane Toluene Chloroform DCM THF Acetonitrile λabs 346 346 343 344 358 345 λem 442 442 447 445 453 458 Stokes Shift (nm) 96 96 104 101 95 113 85 Figure A.9. Emission spectra of 2c in various solvents. Table A.13. Photophysical properties of 2c in various solvents Cyclohexan Toluen Chlorofor Acetonitril e e m DCM THF e λabs 366 365 360 365 367 363 λem 467 466 474 472 480 486 Stokes Shift (nm) 101 101 114 107 113 123 86 Figure A.10. Emission spectra of 2d in various solvents. Table A.14. Photophysical properties of 2d in various solvents Cyclohexan Toluen Chlorofor DCM THF Acetonitrile e m e λabs 355 353 352 365 361 357 λem 458 461 467 465 473 475 Stokes Shift (nm) 103 108 115 100 112 118 87 Figure A.11. Emission spectra of 2e in various solvents. Table A.15. Photophysical properties of 2e in various solvents Cyclohexan Toluen Chlorofor DCM THF Acetonitrile e m e λabs 365 361 360 362 368 365 λem 467 469 473 478 488 493 Stokes Shift (nm) 102 108 113 116 120 128 88 Figure A.12. Emission spectra of 2f in various solvents. Table A.16. Photophysical properties of 2f in various solvents Cyclohexan Toluen Chlorofor e e m DCM THF Acetonitril e λabs 386 384 381 386 387 386 λem 504 508 515 518 524 531 Stokes Shift (nm) 118 124 134 132 137 145 Figure A.13. Fluorescent image of 2f in Acetonitrile, THF, DCM, Chloroform, Toluene, Cyclohexane (left to right) 89 5. Experimental and Computational Examination of the Dimer Solution-state studies of dimer strength General procedure for variable concentration (VC) NMR experiments. CDCl3 was added in a 1:1 ratio to H2O, shaken vigorously, then allowed to separate. The organic phase was then separated and used to prepare ca. 10 mM solutions of heterocycle 2. These solutions were then diluted through addition of known amounts of the H2O-saturated CDCl3 solvent, with 31P NMR spectra being collected after each addition. The 31P NMR spectra were chosen over those of 1H NMR for these experiments since they give comparable results to 1H NMR signals and the 1H NMR signals become impossible to track at lower concentrations due to complex splitting of multiple signals by the phosphorus center. The chemical shift of the phosphorus signal was tracked and fitted to generate the dimerization values.10 Figure A.14. Binding isotherm (red), fit (blue), and residuals (inset) of VC NMR of 2b. 90 Table A.17. Experimental and fitting data for VC NMR experiment of 2b Concentration Experimental (mM) δ (ppm) Fit δ (ppm) 8.60 11.3755 11.37636588 8.32 11.3583 11.35553759 8.06 11.3315 11.33535491 7.48 11.2871 11.28750165 6.97 11.2465 11.24300097 6.14 11.1654 11.16255996 5.49 11.09 11.09160143 4.96 11.0247 11.02836421 4.16 10.918 10.92008376 3.58 10.8299 10.83033275 2.96 10.7164 10.72063772 2.53 10.6342 10.6324564 2.11 10.5421 10.53825939 1.76 10.4816 10.44649514 1.46 10.3465 10.36343269 1.25 10.2875 10.29857999 1.10 10.2345 10.24644727 0.877 10.1592 10.16767483 0.731 10.1191 10.11087501 0.585 10.0577 10.05013847 Figure A.15. Representative 31P NMR spectra for VC NMR of 2b. 91 Figure A.16. Binding isotherm (red), fit (blue), and residuals (inset) of VC NMR of 2c. Table A.18. Experimental and fitting data for VC NMR experiment of 2c Concentration Experimental (mM) δ (ppm) Fit δ (ppm) 14.0 11.2039 11.2064793 13.6 11.1831 11.1885475 13.2 11.1672 11.1711442 12.2 11.1341 11.1297717 11.4 11.0951 11.0911563 10.0 11.0212 11.0209934 8.96 10.9617 10.9586984 8.09 10.9091 10.9028468 6.79 10.8023 10.8064342 5.84 10.7176 10.7257245 4.84 10.6209 10.6260187 4.12 10.5419 10.5449533 3.45 10.4777 10.4573812 2.86 10.4277 10.3710209 2.39 10.2748 10.2918817 2.04 10.1932 10.2294034 1.79 10.1464 10.1787126 1.43 10.0982 10.1012771 1.19 10.0425 10.0447712 0.954 10.0123 9.98368372 92 Figure A.17. Representative 31P NMR spectra for VC NMR of 2c. 93 Figure A.18. Binding isotherm (red), fit (blue), and residuals (inset) of VC NMR of 2d. Table A.19. Experimental and fitting data for VC NMR of 2d Concentration Experimental Fit δ (mM) δ (ppm) (ppm) 12.6 10.7014 10.694546 12.2 10.6799 10.679281 11.8 10.6658 10.664594 10.9 10.6234 10.630185 10.2 10.5926 10.598709 9.00 10.5362 10.543093 8.03 10.5069 10.495405 7.25 10.4546 10.453992 6.09 10.3830 10.385454 5.24 10.3280 10.330913 4.34 10.2696 10.267033 3.70 10.2173 10.217883 3.09 10.1697 10.167531 2.57 10.1452 10.120592 2.14 10.0711 10.079881 1.83 10.0379 10.049254 1.60 10.0143 10.025362 1.28 9.9839 9.9904767 1.07 9.9667 9.9662182 0.856 9.9540 9.9410965 94 Figure A.19. Representative 31P NMR spectra for VC NMR of 2d. 95 Figure A.20. Binding isotherm (red), fit (blue), and residuals (inset) of VC NMR of 2e. Table A.20. Experimental and fitting data for VC NMR of 2e Concentration Experimental (mM) δ (ppm) Fit δ (ppm) 11.6 10.7552 10.75070613 11.2 10.7379 10.73691784 10.9 10.7255 10.72367435 10.1 10.6966 10.69273486 9.39 10.6656 10.66454316 8.27 10.609 10.61499453 7.39 10.5668 10.57278564 6.68 10.5217 10.53634355 5.60 10.4786 10.47648406 4.83 10.4291 10.42926445 3.99 10.3749 10.37444554 3.41 10.3314 10.33263873 2.85 10.3066 10.29015441 2.36 10.2735 10.25087549 1.97 10.2087 10.21706939 1.69 10.1835 10.19180097 1.48 10.1671 10.17218912 1.18 10.1399 10.14371351 0.985 10.1185 10.12402601 0.788 10.109 10.1037383 96 Figure A.21. Representative 31P NMR spectra for VC NMR of 2e. 97 Figure A.22. Binding isotherm (red), fit (blue), and residuals (inset) of VC NMR of 2f. Table A.21. Experimental and fitting data for VC NMR of 2f Concentration Experimental (mM) δ (ppm) Fit δ (ppm) 11.8 10.7356 10.7285531 11.4 10.7052 10.712116 11.0 10.689 10.6963319 10.3 10.6654 10.6594712 9.59 10.6183 10.6259015 8.45 10.5703 10.5669423 7.55 10.5243 10.51676 6.83 10.4787 10.4734671 5.72 10.4056 10.4024241 4.93 10.3464 10.3464462 4.08 10.2772 10.2815335 3.48 10.2254 10.2320847 2.91 10.1666 10.1818864 2.41 10.1348 10.1355237 2.01 10.1038 10.0956594 1.72 10.0768 10.0658868 1.51 10.0456 10.0427937 1.21 10.01 10.0092865 1.01 9.9864 9.98613666 0.805 9.9561 9.96229545 98 Figure A.23. Representative 31P NMR spectra for VC NMR experiment of 2f. Computational studies on dimer interactions The geometries for dimeric complexes of 2f and its -OPh counterpart were fully optimized without symmetry constraint by using the functional M06-2X11 (accounting for the contributions of H-bond/dispersion forces) and TZVP basis set as implemented in Gaussian 09. The PCM solvation model was used to account for the effects of the chloroform environment. All of the optimized structures were confirmed by frequency calculations to be minima using the same level of theory. The solution-phase optimized geometries were used in the following calculations. The natural bond order (NBO) analyses were carried out using NBO 3.1 embedded in Gaussian 09 package. The NCI plot were performed by using the Multiwfn program.12 99 Figure A.24. The NCI plots for (a) 2f and (b) its -OPh analogue meso-dimeric motifs at the PCM(CHCl3)-M06-2X/TZVP equilibrium geometries. NCI regions are represented as solid surfaces and blue−green−red scaling from −0.02 < sign(2)(r) < 0 (in a.u.), where red surface indicates strong repulsion, blue surface strong attraction and green surface relatively weak interactions. Isosurface cutoff for NCI = 0.5. 100 Figure A.25. The NBO charge transfer stabilization energy [second-order perturbation energies E(2), in kcal/mol] for the indicated H-bond interactions in (a) 2f and (b) its -OPh analogue at PCM(CHCl3)-M06-2X/TZVP equilibrium geometries of the meso-dimeric motifs. The indicated H-bond interactions shown by the broken line, and H-bond distances in Å. 6. References: (1) Sheldrick, G. M. Bruker/Siemens Area Detector Absorption Correction Program, Bruker AXS, Madison, WI, 1998. (2) Van der Sluis, P.; Spek, A. L. BYPASS: an effective method for the refinement of crystal structures containing disordered solvent regions. Acta Cryst., Sect. A 1990, A46, 194−201. (3) Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3−8. 101 (4) Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model J. Chem. Phys. 1999, 110, 6158-6169. (5) Bremond, E.; Savarese, M.; Su, N. Q.; Perez-Jimenez, A. J.; Xu, X.; Sancho-Garcia, J. C.; Adamo, C. Benchmarking density functionals on structural parameters of small- /medium-sized organic molecules J. Chem. Theory Comput. 2016, 12, 459–465. (6) Schuafer, A.; Huber, C.; Ahlrichs, R. Fully optimized contracted Gaussian basis sets of triple zeta valence quality for atoms Li to Kr J. Chem. Phys. 1994, 100, 5829–5835. (7) Gaussian 09, Revision E.01, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian, Inc., Wallingford CT, 2013. (8) Miertus, S.; Scrocco, E.; Tomasi, Electrostatic interaction of a solute with a continuum. A direct utilizaion of AB initio molecular potentials for the prevision of solvent effects J. Chem. Phys. 1981, 55, 117–129. (9) Barone, V.; Cossi, M.; Tomasi, J. Geometry optimization of molecular structures in solution by the polarizable continuum model J. Comput. Chem. 1998, 19, 404–417. (10) Thordarson, P. Determining association constants from titration experiments in supramolecular chemistry. Chem. Soc. Rev. 2011, 40, 1305−1323. (11) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals Theor. Chem. Acc. 2008, 120, 215–241. (12) Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer J. Comput. Chem. 2012, 33, 580–592. 102 7. Copies of NMR Spectra Figure A.26. 1H NMR spectrum of 2a in DMSO-d6. Figure A.27. 13C NMR spectrum of 2a in DMSO-d6. 103 Figure A.28. 31P NMR spectrum of 2a in DMSO-d6. Figure A.29. 1H NMR spectrum of 2b in DMSO-d6. 104 Figure A.30. 13C NMR spectrum of 2b in DMSO-d6. Figure A.31. 31P NMR spectrum of 2b in DMSO-d6. 105 Figure A.32. 19F NMR spectrum of 2b in DMSO-d6. Figure A.33. 1H NMR spectrum of 2c in DMSO-d6. 106 Figure A.34. 13C NMR spectrum of 2c in DMSO-d6. Figure A.35. 31P NMR spectrum of 2c in DMSO-d6. 107 Figure A.36. 1H NMR spectrum of 2d in CDCl3. Figure A.37. 13C NMR spectrum of 2d in CDCl3. 108 Figure A.38. 31P NMR spectrum of 2d in CDCl3. Figure A.39. 1H NMR spectrum of 2e in CDCl3. 109 Figure A.40. 13C NMR spectrum of 2e in CDCl3. Figure A.41. 31P NMR spectrum of 2e in CDCl3. 110 Figure A.42. 1H NMR spectrum of 2f in CDCl3. Figure A.43. 13C NMR spectrum of 2f in CDCl3. 111 Figure A.44. 31P NMR spectrum of 2f in CDCl3. 112 APPENDIX B SUPPLEMENTARY CONTENT FOR CHAPTER THREE Synthesis General Methods: All reagents were purchased from conventional commercial sources and used as received, unless otherwise noted. All NMR spectra were acquired at room temperature on a 1H NMR, 13C NMR, and 19F spectra were obtained using a Bruker 500 MHz spectrometer (1H 500 MHz, 13C 126 MHz, 19F 471 MHz). All 1H and 13C chemical shifts () are reported in ppm relative to the residual CHCl3 (1H: 7.26 ppm, 13C: 77.16 ppm) or (CH3)2CO (1H: 2.05, 13C: 206.26, 29.84). Scheme B1: The second, more successful, synthetic path for reaching host receptor 1 Scheme B2: The original, low yielding, synthetic path for accessing receptor 1 113 2-Iodo-4-Trifluoromethyl Aniline (1.4) KI (2.1638g, 13 mmol) and KIO3 (1.5762, 7.4 mmol) were added to a RBF containing DI water (60mL) and set stirring. 2-iodo-4-trifluoromethyl aniline (3.0216g, 18.8 mmol) was then dissolved in MeOH (10 mL) and added to the aqueous solution dropwise. Once everything was dissolved in solution 1M HCl was added dropwise (20.5 mL, 20.5 mmol). The reaction was then left to stir overnight. The next day the reaction was worked up via extraction with DCM. The organic phase was washed twice with saturated solution of NaHCO3, then water and finally brine. The organic phase was then dried with NaSO4, filtered and the solvent removed in vacuo to yield 1.4 in quantitative yields (3.73 g). 1H NMR (500 MHz, Acetone-d6) δ = 7.87 (s, 1H), 7.42 (d, J=9.8, 1H), 6.94 (d, J=8.5, 1H), 5.53 (s, 2H). 4-(Trifluoromethyl)-2-((trimethylsilyl)ethynyl)aniline (1.1) The iodo-aniline (8.899g, 31.0 mmol) was dissolved in 200 mL of a 50/50 mixture of THF and DIPA then set outgassing under N2 for ~3.3 hours. After this time Pd(PPh3)2Cl2. (1.088g, 1.55 mmol) and CuI (0.118g, 0.62 mmol) was weighed out and added to the air free flask. Finally, TMSA (12.95 mL, 186 mmol) was added to the solution dropwise via an addition funnel. The solution was run overnight and worked up the next day, assuming TLC control showed all of the starting material gone. The crude product was purified by 114 column chromatography on silica with EtOAc/Hexanes. The product was produced in 79% yield (6.30 g). 2,2′-(2,6-pyridinediyldi-2,1-ethynediyl)bis[4-(trifluoromethyl)-benzeneamine (1.2) The TMS protected aniline was (10.3084g, 40 mmol) was dissolved in a minimal amount of a 50/50 mixture of MeOH and THF. K2CO3 (31.295g, 240 mmol) was then added (not all of it will dissolve). The reaction was monitored by TLC control and, after completion, the reaction was filtered and solvent removed by rotary evaporation. The crude product was then redissolved in EtOAc and washed thrice with a saturated aqueous solution of ammonium chloride, then water and finally brine. The organic phase was then dried with NaSO4 and filtered before the EtOAc was removed by rotary evaporator to give the deprotected product in quantitative yields. This was then dissolved in 200 mL of 50/50 THF and DIPA and set outgassing under N2. In a second flask 2,6-dibromopyridine (5.26 g, 17.2 mmol) was set sparging in 50 mL of THF and DIPA. After sparging was complete Pd(PPh3)4 (1.37g, 1.72 mmol) and CuI (0.09047g, 0.65 mmol) was added to the first RBF. Then the solution of “elbow” was then cannulated into the first RBF dropwise. The solution was set stirring overnight, and the next day solvent was removed in vacuo and then purified over silica by column chromatography to furnish the bis-aniline in 21% yield (3.74 g). 115 Host Receptor (1) The bis-aniline (5.53 g, 12.3 mmol) was dissolved in 300 mL of ACN and set stirring in a brine and ice bath until it reaches -10 °C. In a separate RBF concentrated HI was diluted with 100 mL ACN and then cannulated in dropwise, the solution went from a light yellow to a bright orange. Then a 10 mL solution of NaNO2 was added by syringe, dropwise, after which the solution was a dark red. The solution was monitored by TLC and after ~90 minutes the reaction had run to completion. At this point the reaction was cannulated over to a quench solution of KI (5.615g, 30.9 mmol) in water. This solution was washed twice with sodium bicarbonate, water and brine before the organic phase was then dried with NaSO4. The ACN was then removed by rotary evaporation and the product was redissolved in DCM and run through a silica plug. The filtrate was reduced, in vacuo to furnish the final receptor in a 30% yield (2.46 g). 116 NMR Spectra HJB-Exp4-PC-Acetone.1.fid NH2 F I F F A (s) B (d) C (d) D (s) 7.87 7.42 6.94 5.53 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 f1 (ppm) Figure C.1. 1H NMR spectrum of 2-Iodo-4-Trifluoromethyl Aniline (1.4) in acetone-d6. HJB-Exp12-PF-Acetone.1.fid 2100 2000 1900 NH2 1800 F 1700 CH3 1600 F F Si 1500 CH3 H3C 1400 1300 1200 1100 C (s) 7.48 1000 B (d) A (d) D (s) 7.37 6.90 0.26 900 800 700 600 500 400 300 200 100 0 -100 -200 -300 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 f1 (ppm) Figure C.2. 1H NMR spectrum of 4-(Trifluoromethyl)-2-((trimethylsilyl)ethynyl)aniline (1.1) in acetone-d6. 117 1 00 7.4778. 0.41 7.3757 1.00 1.32 7.86821.05 0.47 7.3575 7.4274 6.9110 1.02 0.99 1.03 1.00 7.4079 6.8966 6.9518 1.04 1.02 6.9349 1.92 1.96 2.13 5.5338 4.0527 2.8488 2.8051 Water 2.0500 CDCl3 1.9635 2.0500 Acetone 2.0500 CDCl3 1.1966 0.2587 8.84 8.17 0.2561 0.1920 HJB-Exp14-NPDTspot-Acetone.1.fid 6000 5500 5000 4500 4000 F F F N F 3500 F F 3000 C (dd) NH2 H2N 7.75 B (dd) A (d) 7.44 6.97 2500 D (m) 7.63 2000 1500 1000 500 0 -500 9.0 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 f1 (ppm) Figure C.3. 1H NMR spectrum of 2,2′-(2,6-pyridinediyldi-2,1-ethynediyl)bis[4- (trifluoromethyl)-benzeneamine (1.2) in acetone-d6. Figure C.4. 1H NMR spectrum of host receptor (1.3) in acetone-d6. 118 7.7814 7.7657 7.7317 7.7169 7.6592 1.92 1.43 7.6548 1.88 1.48 7.6155 1.05 0.96 7.6001 7.4295 6.9798 1.00 1.00 6.9627 6.9445 1.85 2.8383 Water 2.0500 Acetone UV-Vis titrations: General Titration Procedure: A stock solution of host receptor 1 (0.1 mM) was prepared in CHCl3 and sonicated until dissolved. The stock solution was then used to prep experimental host solutions at a concentration of XXX +/- uM of which 1.75 mL was transferred to a septum sealed quartz cuvette. To maintain a constant host concentration, the remaining experimental host solution was then used to dissolve NBu4Cl salt which was prepared in vol. flasks and transferred to a septum sealed vial. Aliquots of NBu4Cl were added to the host cuvette with a gas-tight Hamilton syringe, and a spectrum was recorded after each addition. Binding constants were determined by non-linear regression fit to a 1:1 model using the Bindfit software. Figure C.5. Representative UV-Vis spectroscopic titration of 1 with TBACl in CHCl3 out to 32 equivalents. 0.18 0.16 Predicted Abs 274 0.14 Predicted Abs 375 0.12 Experimental Abs 274 0.1 Experimental Abs 375 0.08 0.06 0.04 0.02 0 0 5 10 15 20 25 30 35 [G]/[H] 119 ∆ in Abs 0.002 0 0 5 10 15 20 25 30 35 -0.002 274 -0.004 [G]/[H] 375 Figure C.6. The absorption binding isotherm and residuals of TBACl and host 1 at 274 and 325 nm. Table C.1. Calculated association constants (Ka) for 1 in CHCl3 obtained by fitting to a 1:1 host-guest model in Bindfit. Addition # VGuest Equiv. (𝜇L) Guest M Host M Anion 𝛌1: 274.0 𝛌2: 325.0 1 5 4.8023E-06 2.5216E-05 0.190449 0.13496858 0.13089709 2 5 9.5834E-06 2.5216E-05 0.38005531 0.13717921 0.13191755 3 10 1.9082E-05 2.5216E-05 0.75676211 0.13897592 0.13253617 4 20 3.7831E-05 2.5216E-05 1.50030567 0.14299439 0.13376707 5 20 5.6256E-05 2.5216E-05 2.23097401 0.1472857 0.1352873 6 20 7.4364E-05 2.5216E-05 2.94909869 0.15005948 0.13586836 7 20 9.2164E-05 2.5216E-05 3.65499997 0.1531345 0.13683006 8 25 0.00011399 2.5216E-05 4.52065786 0.15732685 0.13772386 9 25 0.00013537 2.5216E-05 5.36828121 0.16341333 0.1397128 10 50 0.0001768 2.5216E-05 7.0116326 0.17213741 0.14254235 11 50 0.00021658 2.5216E-05 8.58924994 0.1794094 0.14429528 12 50 0.00025481 2.5216E-05 10.1049999 0.18651174 0.14666025 13 50 0.00029156 2.5216E-05 11.5624518 0.19426057 0.14910294 14 50 0.00032692 2.5216E-05 12.9649056 0.20099705 0.15107825 15 50 0.00036097 2.5216E-05 14.3154166 0.20796648 0.15375777 120 Residuals 16 100 0.00042543 2.5216E-05 16.871741 0.22334722 0.15857396 17 100 0.00048545 2.5216E-05 19.2517671 0.23541191 0.16212441 18 100 0.00054146 2.5216E-05 21.4731248 0.24931185 0.168189 19 200 0.00064299 2.5216E-05 25.4993358 0.26696813 0.1726637 20 400 0.00081219 2.5216E-05 32.2096873 0.30114302 0.18446825 Table C.2. Titration of 1 with TBACl 121 APPENDIX C SUPPLEMENTARY CONTENT FOR CHAPTER FOUR Experimental Details General Methods. Unless otherwise noted all reactions were run open to the atmosphere and consequently were exposed to air and water. All reagents were used as purchased from commercial sources. The 1H NMR, 13C NMR, and 19F spectra were obtained using a Bruker 500 MHz spectrometer (1H 500 MHz, 13C 126 MHz, 19F 471 MHz), and the 1H NMR titrations were performed on a Varian 500 MHz spectrometer (1H 500.10 MHz) using d3- ACN. All 1H NMR and 13C NMR chemical shifts (δ) are reported in parts per million and referenced to residual solvent peaks (CHCl3: 1H 7.26 ppm, 13C 77.16 ppm; (CH3)2CO: 1H 2.05ppm, 13C 29.84 and 206.26 ppm; CH3CN 1H 1.94 ppm, 13C 118.26 and 1.32 ppm). Masses for novel compounds were determined with a Waters Xevo G2-XS ToF spectrometer. 122 Full Receptor Synthesis. 4-(trifluoromethylthio)acetamide. 4-(trifluoromethylthio)aniline (7.4 mL, 71.8 mmol) was added under N2 to an oven-dried, vacuum-cooled round bottom flask. Acetic anhydride (10.2 mL, 108 mmol) was added to the flask dropwise and a white precipitate crashed out upon addition. The mixture was stirred at room temperature for one hour. Afterwards, the solid was dissolved using ethyl acetate and washed with water and brine. The organic layer was dried (Na2SO4), filtered, and concentrated to give the crude protected aniline. The white powdered product was purified by running the crude product through a plug with dichloromethane. (12 g, 71% yield). 1H NMR (500 MHz, Acetone-d6) δ 9.44 (s, 1H), 7.80 (d, J = 8.7 Hz, 2H), 7.63 (d, J = 8.7 Hz, 2H), 2.11 (s, 3H). N-[4-[(trifluoromethyl)sulfonyl]phenyl]acetamide. In an oven-dried, vacuum-cooled round bottom flask, the 4-(trifluoromethyl)thio phenyl acetamide (12 g, 51.0 mmol) was dissolved in 200 mL of dry dichloromethane and cooled to 0 ºC. mCPBA (26 g, 153 mmol) was added in 4 portions under a heavy N2 (g) flow. The reaction was allowed to warm up to room temperature while stirring overnight. The 123 reaction mixture was diluted with more dichloromethane and Na2S2O3 was used to quench any excess mCPBA. The mixture was extracted with dichloromethane and the combined organic layers were washed with Na2S2O3, NaHCO3, water and brine, then dried (Na2SO4) and concentrated to afford a white precipitate (11.4 g, 84% yield). 1H NMR (500 MHz, Acetone-d6) δ 9.84 (s, 1H), 8.08 – 8.00 (m, 4H), 2.18 (s, 3H). N-[4-[(trifluoromethyl)sulfonyl]phenyl]aniline. N-[4-[(trifluoromethyl)sulfonyl]phenyl]acetamide (6.2 g, 23.15 mmol) was dissolved in 200 mL of EtOH; KOH (1.43 g, 25.46 mmol) was added, and the reaction mixture was set to 60 ºC and allowed to stir. After 3 hours, the mixture was concentrated to afford a crude solid which was then dissolved in water. White precipitate formed and collected via vacuum filtration to afford pure product (4.7 g, 90% yield). 1H NMR (500 MHz, Chloroform-d) δ 7.75 (d, J = 8.8 Hz, 2H), 6.75 (d, J = 8.8 Hz, 2H), 4.54 (s, 2H). Iodoaniline derivatives, 1a-d. The aniline derivative (20.76 mmol, 1.0 equiv.) was dissolved in 40 mL methanol. In another flask, KIO3 (7.27 mmol, 0.35 equiv.) and KI (13.7 mmol, 0.66 equiv.) were dissolved in 100 mL water. The aqueous solution was added to the solution containing the aniline derivative then 1 M HCl (23 mL, 1.1 equiv.) was added dropwise. After the 124 reaction stirred at room temperature overnight, the mixture was extracted with ethyl acetate. The combined organic layers were then washed with Na2S2O3, NaHCO3, water, and brine, then dried (Na2SO4) and concentrated to afford crude product. R= SO2CF3 (red/brown solid, 92% yield). 1H NMR (500 MHz, Chloroform-d) δ 8.22 (d, J = 2.1 Hz, 1H), 7.74 (dd, J = 8.6, 2.1 Hz, 1H), 6.81 (d, J = 8.6 Hz, 1H), 4.99 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 153.83, 142.11, 132.45, 123.86, 121.27, 118.88, 118.68, 116.09, 113.44, 81.85, 77.16. 19F NMR (471 MHz, CDCl3) δ -78.79. MASS SPEC NEEDED. R= CF3 (brown solid, 79% yield). 1H NMR (500 MHz, Chloroform-d) δ 7.87 (s, 1H), 7.38 (dd, J = 8.5, 2.2 Hz, 1H), 6.74 (d, J = 8.4 Hz, 1H), 4.41 (s, 2H). R= H commercially available R= tBu (brown solid, 66% yield). 1H NMR (500 MHz, Chloroform-d) δ 7.65 (d, J = 1.7 Hz, 1H), 7.19 (d, J = 2.0 Hz, 1H), 6.70 (d, J = 8.4 Hz, 1H), 3.99 (s, 2H), 1.28 (d, J = 1.7 Hz, 9H). Iodotriazene derivatives, 2a-d. The iodoaniline derivative (8.55 mmol, 1.0 equiv.) was dissolved in 30 mL acetonitrile and set to cool at -10 ºC. Concentrated HCl (5.7 mL, 8.0 equiv.) was added to the flask. In another flask, a solution of NaNO2 (20.51 mmol, 2.4 equiv.) in 30 mL water was prepared. The aqueous NaNO2 solution was added dropwise to the solution of the iodoaniline derivative and left to stir at -10 ºC. This reaction mixture was monitored by TLC for consumption of starting material. Once consumed, the reaction mixture was 125 added to a quench solution of: HNEt2 (85.45 mmol, 10 equiv.) and K2CO3 (49.56 mmol, 5.8 equiv.) in 50 mL of acetonitrile and 100 mL of water at -10 ºC. This quenched reaction mixture was allowed to warm up to room temperature while continuously stirring overnight. The reaction was extracted with ether and the combined organic layers were washed with Na2S2O3, NaHCO3, water, and brine. The organic layer was collected and dried (Na2SO4) and concentrated to afford the crude iodotriazene that was purified by column chromatography (gradient from pure hexanes to 1:9, dichloromethane:hexanes) to afford pure iodotriazene. R= SO2CF3 (orange solid, 25% yield). 1H NMR (500 MHz, Chloroform-d) δ 8.11 (d, J = 2.2 Hz, 1H), 7.56 (dd, J = 8.6, 1.9 Hz, 1H), 7.27 (d, J = 8.7 Hz, 1H), 3.60 (qd, J = 7.2, 5.1 Hz, 4H), 1.07 (dt, J = 30.0, 7.2 Hz, 6H). 13C NMR (126 MHz, CDCl3) δ 157.06, 141.51, 131.04, 126.15, 123.75, 121.16, 118.56, 117.34, 115.97, 96.24, 50.42, 43.64, 14.29, 10.69. 19F NMR (471 MHz, CDCl3) δ -77.75. MASS SPEC NEEDED. R= CF3 (orange solid, quant. yield). 1H NMR (500 MHz, Chloroform-d) δ 8.07 (s, 1H), 7.51 (dd, J = 8.6, 2.0 Hz, 1H), 7.42 (d, 1H), 3.84 (q, J = 7.2 Hz, 4H), 1.34 (dt, J = 24.0, 7.3 Hz, 6H). R= H (orange solid, quant. yield). 1H NMR (500 MHz, Chloroform-d) δ 7.86 (dd, J = 7.9, 1.4 Hz, 1H), 7.38 (dd, J = 8.1, 1.6 Hz, 1H), 7.29 (ddd, J = 8.3, 7.2, 1.4 Hz, 1H), 6.85 (td, J = 7.5, 1.6 Hz, 1H), 3.81 (q, J = 7.2 Hz, 4H), 1.34 (t, J = 7.2 Hz, 6H). R= tBu (orange solid, 89% yield). 1H NMR (500 MHz, Chloroform-d) δ 7.88 (d, J = 2.0 Hz, 1H), 7.60 – 7.01 (m, 2H), 3.80 (q, J = 7.2 Hz, 4H), 1.74 – 0.92 (m, 15H). 126 TMS-protected 3-ethynylpyridine. After a mixture of 10 mL tetrahydrofuran and 10 mL acetonitrile, 3-bromopyridine (0.31 mL, 3.16 mmol), Pd(PhCN)2Cl2 (0.16 mmol, 5 mol%) and CuI (0.16 mmol, 5 mol%) was outgassed with N2 (g) for at least 20 minutes, the ligand P(tBu)3 (0.316 mmol, 10 mol%) was added. The alkyne, TMSA (4.74 mmol, 1.5 equiv.) was then added dropwise. The reaction was allowed to stir at room temperature under a steady flow of N2 (g) overnight. The reaction mixture was run through a silica/celite plug and washed with DCM to afford the crude product that was then purified by column chromatography (gradient from pure hexanes to 2:8, ethylacetate:hexanes) to afford a pure brown oil. (0.549 g, quant. yield). 1H NMR (500 MHz, Chloroform-d) δ 8.59 (s, 1H), 8.41 (dd, J = 4.9, 1.9 Hz, 1H), 7.62 (dt, J = 7.8, 1.8 Hz, 1H), 7.10 (dd, J = 7.9, 4.8 Hz, 1H), 0.16 (s, 6H). Triazene receptor intermediate, 3a-d. The 3-ethynylpyridine (1.71 mmols, 1.0 equiv.) was dissolved in Et2O and set to stir in a flask. Tetrabutylammonium fluoride (TBAF, 2.40 mmols) was then added to the solution and left to stir until the starting material had disappeared by TLC (~ 20 min.). The reaction mixture was quenched with NH4Cl and extracted with water and brine x2. The organic layer was dried (NaSO4), and the solvent was removed in vacuo to furnish 3- ethynylpyridine in quantitative yields. While this was going on a new flask of 50% THF 127 and 50% DIPA (20 mL total) was outgassed with N2(g) for at least twenty minutes. Once isolated 3-ethynylpyridine (2.28 mmol, 1.3 equiv.) was dissolved in a minimal amount of THF and outgassed for at least 1 min/mL of solvent. After outgassing, the iodotriazene compound (1.76 mmol, 1.0 equiv.), Pd(PPh3)4 (0.18 mmol, 0.10 equiv.), and CuI (0.11 mmol, 0.06 equiv.) were added to the RBF with the THF/DIPA mixture and set to stir. The separate solution of 3-ethynylpyridine in THF was then cannulated dropwise to the iodotriazene/catalyst mixture and the reaction was left stirring at 45 °C overnight. The next day, the reaction was passed through a silica/celite plug, washed with DCM, and concentrated to afford the crude product that was purified by column chromatography (gradient from 1:9, DCM:hexanes to 100% DCM) R= SO2CF3 (yellow solid, 69% yield). 1H NMR (500 MHz, Chloroform-d) δ 8.77 (s, 1H), 8.58 (s, 1H), 8.17 (d, J = 2.2 Hz, 1H), 7.85 (dd, J = 8.8, 2.2 Hz, 1H), 7.81 (d, J = 7.9 Hz, 1H), 7.71 (d, J = 8.8 Hz, 1H), 7.31 (dd, J = 7.9, 4.8 Hz, 1H), 3.91 (dq, J = 11.1, 7.2 Hz, 4H), 1.38 (dt, J = 30.4, 7.2 Hz, 6H). CHECKING WITH MIKE IN SUBGROUP 13C NMR (126 MHz, CDCl3) δ 159.95, 153.74, 150.39, 139.86, 137.42, 132.52, 126.79, 125.30, 124.71, 122.71, 120.43, 120.12, 119.17, 117.53, 93.63, 90.41, 51.86, 44.69, 15.84, 12.24. 19F NMR (471 MHz, CDCl3) δ -78.49. HRMS (TOF-MS-ES+) for C18H17N4SO2F3 [M+H] +: calcd 410.10, found 410.1024. R= CF3 (76% yield). NEED NMRS. HRMS (TOF-MS-ES+) for C18H17N4F3 [M+H] +: calcd 346.14 , found 346.1405. R= H (yellow solid, 97% yield). 1H NMR (500 MHz, Chloroform-d) δ 8.77 (s, 1H), 8.53 (s, 1H), 7.79 (dd, J = 7.8, 1.8 Hz, 1H), 7.54 (dd, J = 7.6, 1.5 Hz, 1H), 7.45 (d, J = 8.4 Hz, 1H), 7.34 – 7.27 (m, 2H), 7.10 (t, J = 7.5 Hz, 1H), 3.82 (q, J = 7.2 Hz, 4H), 1.33 (t, J = 128 7.2 Hz, 6H). NEED DRY HNMR. NEED CARBON. HRMS (TOF-MS-ES+) for C17H18N4 [M+H] +: calcd 278.15, found 278.1531. R= tBu (light brown oil, quant. yield). 1H NMR (500 MHz, Chloroform-d) δ 7.88 (d, J = 2.0 Hz, 1H), 7.60 – 7.01 (m, 2H), 3.80 (q, J = 7.2 Hz, 4H), 1.74 – 0.92 (m, 15H). 13C NMR (126 MHz, CDCl3) δ 207.02, 152.31, 150.42, 148.14, 147.85, 147.66, 138.25, 129.58, 127.08, 123.10, 116.83, 116.79, 92.40, 89.24, 34.52, 31.41, 31.04. HRMS (TOF- MS-ES+) for C21H26N4 [M+H] +: calcd 334.22, found 334.2157. Neutral ethynyl pyridine receptor, 4a-d. HI (0.54 mmol, 3.0 equiv.) and I2 (0.54 mmol, 3.0 equiv.) were dissolved in ACN (5 mL) then set to stir at 60 °C. While the first solution was coming up to temperature the triazene (0.18 mmol, 1.0 equiv.) was dissolved in ACN (5 mL), which was then added dropwise into the first flask and stirred at 60 °C while monitoring via TLC. After the starting material was consumed (~40 min), the mixture was extracted with EtOAc and washed with NaHCO3, Na2S2O3, and water x2 each. The organic layer was then dried with Na2SO4 and concentrated to form an oil which was purified on silica via column chromatography (gradient from 100% hexanes to 3:7 ethylacetate:hexanes). R= SO2CF3 (tan solid, 40% yield). 1H NMR (500 MHz, Acetone-d6) δ 8.88 (s, 1H), 8.66 (d, J = 3.9 Hz, 1H), 8.50 (d, J = 8.4 Hz, 1H), 8.23 (d, J = 2.3 Hz, 1H), 8.06 (dt, J = 7.8, 1.9 Hz, 1H), 7.83 (dd, J = 8.4, 2.3 Hz, 1H), 7.50 (dd, J = 8.0, 4.8 Hz, 1H). NEED TO 129 CHECK CARBON WITH MIKE IN SUBGROUP.13C NMR (126 MHz, Acetone) δ 152.99, 150.93, 142.45, 139.48, 134.12, 132.75, 132.13, 131.48, 124.46, 121.88, 119.78, 119.29, 116.70, 113.65, 93.60, 92.97. 19F NMR (471 MHz, CDCl3) δ -79.33. HRMS (TOF-MS-ES+) for C14H7INSO2F3 [M+H]+: calcd 436.92, found 436.9194. R= CF3 (yellow solid, 31% yield). 1H NMR (500 MHz, Chloroform-d) δ 8.62 (d, J = 2.5 Hz, 1H), 8.42 (dd, J = 4.9, 1.7 Hz, 1H), 8.01 (d, J = 8.3 Hz, 1H), 7.89 – 7.75 (m, 1H), 7.70 (d, J = 2.2 Hz, 1H), 7.35 – 7.13 (m, 2H). NEED TO CHECK CARBON WITH MIKE IN SUBGROUP. 19F NMR (471 MHz, CDCl3) δ -58.44. HRMS (TOF-MS-ES+) for C14H7INF3 [M+H] +: calcd 372.96, found 372.9575. HRMS (TOF-MS-ES+) for C13H8IN [M+H] +: calcd 304.97, found 304.970. R= H (tan solid, 20% yield). 1H NMR (500 MHz, Chloroform-d) δ 8.83 (d, J = 2.1 Hz, 1H), 8.58 (dd, J = 5.0, 1.7 Hz, 1H), 7.92 – 7.85 (m, 2H), 7.54 (dd, J = 7.7, 1.7 Hz, 1H), 7.35 (td, J = 7.6, 1.2 Hz, 1H), 7.31 (ddd, J = 7.8, 4.9, 0.8 Hz, 1H), 7.05 (td, J = 7.8, 1.7 Hz, 1H). R= tBu (yellow solid, 87% yield). 1H NMR (500 MHz, Chloroform-d) δ 8.84 (s, 1H), 8.57 (d, J = 3.3 Hz, 1H), 7.89 (dt, J = 7.9, 2.0 Hz, 1H), 7.79 (d, J = 8.4 Hz, 1H), 7.57 (d, J = 2.4 Hz, 1H), 7.31 (dd, J = 7.9, 4.9 Hz, 1H), 7.09 (dd, J = 8.4, 2.4 Hz, 1H), 1.32 (s, 9H). 13C NMR (126 MHz, CDCl3) δ 152.49, 151.68, 149.09, 138.75, 138.72, 130.16, 128.84, 127.96, 123.36, 120.54, 97.64, 95.49, 89.14, 34.89, 31.34. HRMS (TOF-MS-ES+) for C17H16IN [M+H] +: calcd 361.03, found 361.0327. 130 Methylated ethynyl pyridine. The neutral receptor (0.067 mmol) was added to a bomb flask followed by an excess of MeI (5 mL). The reaction vessel was then sealed and set to stir at 110 °C overnight. Over the next 24 hours, the product would precipitate out of solution. The yellow precipitate was collected and washed with minimal DCM to afford pure product. R= SO2CF3 (yellow solid, quant. yield). 1H NMR (500 MHz, Acetone-d6) δ 9.65 (s, 1H), 9.30 (d, J = 6.1 Hz, 1H), 8.94 (d, J = 8.1 Hz, 1H), 8.53 (d, J = 8.4 Hz, 1H), 8.42 – 8.32 (m, 1H), 8.26 (d, J = 2.4 Hz, 1H), 7.91 (dd, J = 8.4, 2.3 Hz, 1H). NEED CARBON. 19F NMR (471 MHz, Acetone) δ -79.27. HRMS (TOF-MS-ES+) for C14H7I2NSO2F3 [M+H] +: calcd 451.94, found 451.9429. R= CF3 (yellow solid, quant. yield). 1H NMR (500 MHz, Acetone-d6) δ 9.62 (s, 1H), 9.30 (d, J = 6.2 Hz, 1H), 8.89 (d, J = 8.1 Hz, 1H), 8.38 (t, J = 7.2 Hz, 1H), 8.29 (d, J = 8.3 Hz, 1H), 8.00 (d, J = 2.2 Hz, 1H), 7.60 (d, J = 8.2 Hz, 1H), 4.76 (s, 3H). NEED CARBON. 19F NMR (471 MHz, Acetone-d6) δ -63.70. HRMS (TOF-MS-ES+) for C14H7I2NF3 [M+H] +: calcd 387.98, found 387.9810. R= H (yellow solid, quant. yield). 1H NMR (500 MHz, Acetone-d6) δ 9.40 (s, 1H), 9.16 (d, J = 6.1 Hz, 1H), 8.84 (d, J = 8.1 Hz, 1H), 8.38 – 8.31 (m, 1H), 8.05 (d, J = 8.0 Hz, 1H), 7.71 (dd, J = 7.9, 1.7 Hz, 1H), 7.59 – 7.52 (m, 1H), 7.30 (td, J = 7.7, 1.5 Hz, 1H), 4.71 (s, 3H). NEED CARBON. NEED FLUORINE. HRMS (TOF-MS-ES+) for C13H8I2N [M+H] +: calcd 319.99, found 319.9936. R= tBu (yellow solid, quant. yield). 1H NMR (500 MHz, Acetone-d6) δ 9.48 (s, 1H), 9.23 (d, J = 6.1 Hz, 1H), 8.84 (d, J = 8.0 Hz, 1H), 8.39 – 8.32 (m, 1H), 7.94 (d, J = 8.5 Hz, 1H), 7.73 (d, J = 2.4 Hz, 1H), 7.37 (dd, J = 8.5, 2.5 Hz, 1H), 4.74 (s, 3H), 1.34 (s, 9H). 131 NEED CARBON. No peaks shown on 19F NMR. HRMS (TOF-MS-ES+) for C17H16I2N [M+H] +: calcd 376.06, found 376.0562. Methylated PF6 pyridine receptor, 5a-d. The methylated receptor (0.423 mmol, 1.0 equiv.) was dissolved in 10 mL of acetone and 40 mL of water and set to stir. To this flask, AgPF6 (1.068 mmol, 2.5 equiv.) was added and stirred for about 40 minutes during which time gray precipitate (AgI) formed. This precipitate was filtered off and the filtrate solvent mixture was removed in vacuo until only water remained. At this point white precipitate (final product) began to crash out of the water and was collected via filtration and washed with minimal water to give the final pure product. R= SO2CF3 (white solid, quant. yield). 1H NMR (500 MHz, Acetone-d6) δ 9.52 (s, 1H), 9.24 (d, J = 6.2 Hz, 1H), 8.95 (d, J = 8.2 Hz, 1H), 8.57 (d, J = 8.4 Hz, 1H), 8.41 (t, 1H), 8.27 (d, J = 2.3 Hz, 1H), 7.94 (dd, J = 8.4, 2.3 Hz, 1H), 4.74 (s, 3H). NEED CARBON. 19F NMR (471 MHz, Acetone) δ -71.74, -73.24, -79.32. R= CF3 (white solid, quant. yield). NEED PROTON AND CARBON IN ACETONE. 19F NMR (471 MHz, Acetone-d6) δ -63.79, -71.90, -73.41. R= H (white solid, quant. yield). 1H NMR (500 MHz, Acetone-d6) δ 9.32 (s, 1H), 9.07 (d, J = 6.1 Hz, 1H), 8.80 (d, J = 8.1 Hz, 1H), 8.29 (dd, J = 8.2, 6.2 Hz, 1H), 8.02 (d, J = 8.0 Hz, 1H), 7.71 (dd, J = 7.8, 1.6 Hz, 1H), 7.54 (t, J = 7.6 Hz, 1H), 7.28 (td, J = 7.7, 1.6 132 Hz, 1H), 4.65 (s, 3H). 13C NMR (126 MHz, Acetone) δ 148.34, 147.55, 145.97, 140.09, 134.29, 132.54, 129.45, 129.12, 128.23, 124.80, 101.12, 99.26, 85.80, 49.52. 19F NMR (471 MHz, Acetone) δ -71.64, -73.14. R= tBu (white solid, quant. yield). 1H NMR (500 MHz, Acetone-d6) δ 9.34 (s, 1H), 9.09 (d, J = 6.1 Hz, 1H), 8.81 (d, J = 8.1 Hz, 1H), 8.33 – 8.25 (m, 1H), 7.93 (d, J = 8.4 Hz, 1H), 7.73 (d, J = 2.5 Hz, 1H), 7.36 (dd, J = 8.5, 2.5 Hz, 1H), 4.67 (s, 3H), 1.33 (s, 9H) ). CHECKING CARBON WITH MIKE IN SUBGROUP. 13C NMR (126 MHz, Chloroform-d) δ 152.88, 148.35, 147.55 (d, J = 14.2 Hz), 145.94 (d, J = 18.2 Hz), 139.80 (d, J = 22.4 Hz), 131.42 (d, J = 24.9 Hz), 130.25 (d, J = 8.7 Hz), 129.15 (d, J = 21.9 Hz), 127.90, 124.96, 99.75, 97.42, 85.31, 49.54, 35.28, 31.13. 19F NMR (471 MHz, Acetone) δ -71.91, -73.41. Titrations General Methods. 1H NMR spectra were acquired at room temperature on a Varian Inova 500 MHz spectrometer (500.11 MHz). 1H chemical shifts (δ) are expressed in ppm relative to residual CH3CN (calibrated to 1.94 ppm) shifts. General Procedure. A 4.0 mL host solution (0.9-1.5 mM in CD3CN) was prepared. Of this, 500 μL was added to a septum-sealed NMR tube. A 19-30 mM host/guest (TBAX, where X = I, Br, or Cl) stock solution was prepared using the remaining 3.5 mL of host solution. Aliquots of the host/guest solution were added to the NMR tube using Hamilton gas-tight syringes, and 1H NMR spectra were recorded at 25 °C after each addition of 133 guest. The Δδ of the CHaryl and CHmethyl protons were used to follow the progress of the titration, and association constants were determined using the Thordarson method. Figure C.1. Representative 1H NMR titration done in CD3CN with 5a where participating protons can be seen shifting with increasing chloride concentration. Table C.1. Representative 1H NMR Titration of 5a • PF6– with I– in CD3CN at 25 ºC. VGuest [Host] Entry [I– ] (M) Equiv. δ CHaryl δ CHmethyl (μL) (M) 0 0 1.32E-03 0.00E+00 0.00 8.888 4.322 1 5 1.32E-03 2.08E-04 0.16 8.892 4.324 2 5 1.32E-03 4.11E-04 0.31 8.897 4.326 3 10 1.32E-03 8.06E-04 0.61 8.905 4.329 4 10 1.32E-03 1.19E-03 0.90 8.912 4.332 5 10 1.32E-03 1.55E-03 1.18 8.918 4.334 6 15 1.32E-03 2.08E-03 1.58 8.928 4.339 7 15 1.32E-03 2.57E-03 1.95 8.934 4.340 8 15 1.32E-03 3.05E-03 2.31 8.940 4.343 9 20 1.32E-03 3.64E-03 2.76 8.947 4.346 10 20 1.32E-03 4.19E-03 3.18 8.954 4.349 134 11 20 1.32E-03 4.71E-03 3.57 8.960 4.351 12 30 1.32E-03 5.44E-03 4.12 8.967 4.354 13 50 1.32E-03 6.51E-03 4.93 8.978 4.359 14 100 1.32E-03 8.26E-03 6.26 8.992 4.364 15 200 1.32E-03 1.07E-02 8.14 9.008 4.371 16 500 1.32E-03 1.41E-02 10.68 9.016 4.375 17 500 1.32E-03 1.58E-02 11.97 9.028 4.379 18 500 1.32E-03 1.68E-02 12.75 9.035 4.381 Binding Isotherm of I– with 5a • PF – 0.0066 CH (aryl) Residuals0.2 CH (methyl) Residuals CH (aryl) Predicted CH (methyl) Predicted 0.15 CH (aryl) ExperimentalCH (methyl) Experimental 0.1 0 0 2 4 6 8 10 12 14 0.05 0 0 2 4 6 8 10 12 14 [G]/[H] -0.006 [G]/[H] Figure C.2. Representative binding isotherm and residuals for I– titration of receptor 5a • PF6– in CD3CN at 25 °C determined by 1H NMR spectroscopy. Table C.2. Representative 1H NMR Titration of 5a • PF6– with Br– in CD3CN at 25 ºC. VGuest [Host] Entry [Br– ] (M) Equiv. δ CHaryl δ CHmethyl (μL) (M) 0 0 1.25E-03 0.00E+00 0.00 8.888 4.323 1 5 1.25E-03 2.34E-04 0.19 8.896 4.326 2 5 1.25E-03 4.64E-04 0.37 8.905 4.329 3 10 1.25E-03 9.10E-04 0.73 8.922 4.335 4 10 1.25E-03 1.34E-03 1.07 8.938 4.341 5 10 1.25E-03 1.75E-03 1.40 8.951 4.345 6 15 1.25E-03 2.35E-03 1.87 8.967 4.351 7 15 1.25E-03 2.91E-03 2.32 8.982 4.357 135 δ (ppm) Residuals 8 15 1.25E-03 3.44E-03 2.75 8.994 4.361 9 20 1.25E-03 4.11E-03 3.28 9.008 4.366 10 20 1.25E-03 4.73E-03 3.78 9.022 4.371 11 20 1.25E-03 5.32E-03 4.25 9.033 4.375 12 30 1.25E-03 6.14E-03 4.90 9.047 4.381 13 50 1.25E-03 7.35E-03 5.87 9.065 4.387 14 100 1.25E-03 9.32E-03 7.45 9.093 4.398 15 200 1.25E-03 1.21E-02 9.68 9.124 4.408 16 500 1.25E-03 1.59E-02 12.71 9.157 4.420 17 500 1.25E-03 1.78E-02 14.24 9.160 4.422 Binding Isotherm of I– with 5a • PF – 0.0066 CH (aryl) Residuals 0.2 CH (methyl) Residuals CH (aryl) Predicted CH (methyl) Predicted 0.15 CH (aryl) Experimental CH (methyl) Experimental 0.1 0 0 2 4 6 8 10 12 14 0.05 0 0 2 4 6 8 10 12 14 [G]/[H] -0.006 [G]/[H] Figure C.3. Representative binding isotherm and residuals for Br– titration of receptor 5a • PF6– in CD3CN at 25 °C determined by 1H NMR spectroscopy. Table C.3. Representative 1H NMR Titration of 5a • PF6– with Cl– in CD3CN at 25 ºC. VGuest [Host] δ Entry [Cl– ] (M) Equiv. δ CHaryl (μL) (M) CHmethyl 0 0 9.58E-04 0.00E+00 0.00 8.887 4.322 1 5 9.58E-04 1.99E-04 0.21 8.905 4.328 2 5 9.58E-04 3.93E-04 0.41 8.911 4.330 3 10 9.58E-04 7.72E-04 0.81 8.941 4.340 136 δ (ppm) Residuals 4 10 9.58E-04 1.14E-03 1.19 8.957 4.346 5 10 9.58E-04 1.49E-03 1.55 8.972 4.351 6 15 9.58E-04 1.99E-03 2.08 9.006 4.362 7 15 9.58E-04 2.46E-03 2.57 9.019 4.367 8 15 9.58E-04 2.91E-03 3.04 9.037 4.372 9 20 9.58E-04 3.48E-03 3.63 9.057 4.380 10 20 9.58E-04 4.01E-03 4.19 9.079 4.386 11 20 9.58E-04 4.51E-03 4.71 9.090 4.390 12 30 9.58E-04 5.20E-03 5.43 9.108 4.396 13 50 9.58E-04 6.23E-03 6.50 9.134 4.405 14 100 9.58E-04 7.90E-03 8.25 9.170 4.417 15 200 9.58E-04 1.03E-02 10.73 9.205 4.429 16 500 9.58E-04 1.35E-02 14.08 9.262 4.449 17 500 9.58E-04 1.51E-02 15.77 9.263 4.448 Binding Isotherm of Cl– with 5a • PF – 0.0126 CH(aryl) Residuals 0.4 CH(methyl) Residuals 0.35 0.3 0.25 CH(aryl) Predicted CH(methyl) Predicted 0.2 CH(aryl) Experimental 0 0 5 10 15 CH(methyl) Experimental 0.15 0.1 0.05 0 0 2 4 6 8 10 12 14 16 [G]/[H] -0.012 [G]/[H] Figure C.4. Representative binding isotherm and residuals for Cl– titration of receptor 5a • PF6– in CD3CN at 25 °C determined by 1H NMR spectroscopy. Table C.4. Representative 1H NMR Titration of 5b • PF6– with Cl– in CD3CN at 25 ºC. VGuest [Host] Entry [Cl– ] (M) Equiv. δ CHaryl δ CHmethyl (μL) (M) 137 δ (ppm) Residuals 0 0 1.35E-03 0.00E+00 0.00 8.864 4.318 1 5 1.35E-03 2.46E-04 0.18 8.874 4.322 2 5 1.35E-03 4.88E-04 0.36 8.888 4.327 3 10 1.35E-03 9.57E-04 0.71 8.911 4.335 4 10 1.35E-03 1.41E-03 1.04 8.932 4.342 5 10 1.35E-03 1.84E-03 1.37 8.951 4.349 6 15 1.35E-03 2.47E-03 1.83 8.974 4.357 7 15 1.35E-03 3.06E-03 2.26 8.993 4.364 8 15 1.35E-03 3.62E-03 2.68 9.010 4.370 9 20 1.35E-03 4.32E-03 3.20 9.029 4.376 10 20 1.35E-03 4.98E-03 3.69 9.045 4.382 11 20 1.35E-03 5.59E-03 4.14 9.060 4.387 12 30 1.35E-03 6.45E-03 4.78 9.079 4.394 13 50 1.35E-03 7.72E-03 5.72 9.106 4.403 14 100 1.35E-03 9.80E-03 7.26 9.140 4.415 15 200 1.35E-03 1.27E-02 9.44 9.176 4.428 16 500 1.35E-03 1.67E-02 12.39 9.192 4.434 17 500 1.35E-03 1.87E-02 13.88 9.217 4.442 Binding Isotherm of Cl– with 5b • PF – 0.0126 CH(aryl) Residuals 0.4 CH(methyl) Residuals CH(aryl) Predicted CH(methyl) Predicted CH(aryl) Experimental 0.3 CH(methyl) Experimental 0.2 0 0 2 4 6 8 10 12 14 0.1 0 0 2 4 6 8 10 12 14 [G]/[H] -0.012 [G]/[H] Figure C.5. Representative binding isotherm and residuals for Cl– titration of receptor 5b • PF6– in CD3CN at 25 °C determined by 1H NMR spectroscopy. 138 δ (ppm) Residuals Table C.5. Representative 1H NMR Titration of 5c • PF6– with Cl– in CD3CN at 25 ºC. VGuest [Host] Entry [Cl– ] (M) Equiv. δ CHaryl δ CHmethyl (μL) (M) 0 0 1.36E-03 0.00E+00 0.00 8.838 4.309 1 5 1.36E-03 2.45E-04 0.18 8.847 4.312 2 5 1.36E-03 4.85E-04 0.36 8.857 4.316 3 10 1.36E-03 9.52E-04 0.70 8.877 4.324 4 10 1.36E-03 1.40E-03 1.03 8.891 4.330 5 10 1.36E-03 1.83E-03 1.35 8.907 4.336 6 15 1.36E-03 2.45E-03 1.80 8.918 4.340 7 15 1.36E-03 3.04E-03 2.23 8.941 4.348 8 15 1.36E-03 3.60E-03 2.64 8.954 4.354 9 20 1.36E-03 4.29E-03 3.16 8.969 4.359 10 20 1.36E-03 4.95E-03 3.64 8.982 4.363 11 20 1.36E-03 5.56E-03 4.09 8.997 4.371 12 30 1.36E-03 6.41E-03 4.72 9.013 4.377 13 50 1.36E-03 7.68E-03 5.65 9.034 4.384 14 100 1.36E-03 9.75E-03 7.17 9.064 4.397 15 200 1.36E-03 1.27E-02 9.32 9.094 4.405 16 500 1.36E-03 1.66E-02 12.23 9.138 4.424 17 500 1.36E-03 1.86E-02 13.70 9.142 4.425 139 Binding Isotherm of Cl– with 5c • PF – 0.0126 CH(aryl) Residuals0.4 CH(aryl) Predicted CH(methyl) Residuals 0.35 CH(methyl) Predicted CH(aryl) Experimental 0.3 CH(methyl) Experimental 0.25 0.2 0 0 2 4 6 8 10 12 14 0.15 0.1 0.05 0 0 2 4 6 8 10 12 14 [G]/[H] -0.012 [G]/[H] Figure C.6. Representative binding isotherm and residuals for Cl– titration of receptor 5c • PF6– in CD3CN at 25 °C determined by 1H NMR spectroscopy. Table C.6. Representative 1H NMR Titration of 5d • PF6– with Cl– in CD3CN at 25 ºC. VGuest [Host] Entry [Cl– ] (M) Equiv. δ CHaryl δ CHmethyl (μL) (M) 0 0 1.73E-03 0.00E+00 0.00 8.828 4.310 1 5 1.73E-03 2.33E-04 0.14 8.834 4.312 2 5 1.73E-03 4.62E-04 0.27 8.843 4.316 3 10 1.73E-03 9.07E-04 0.53 8.859 4.322 4 10 1.73E-03 1.33E-03 0.77 8.873 4.328 5 10 1.73E-03 1.75E-03 1.01 8.885 4.333 6 15 1.73E-03 2.34E-03 1.35 8.903 4.340 7 15 1.73E-03 2.90E-03 1.68 8.917 4.345 8 15 1.73E-03 3.43E-03 1.98 8.929 4.350 9 20 1.73E-03 4.09E-03 2.37 8.944 4.356 10 20 1.73E-03 4.72E-03 2.73 8.958 4.361 11 20 1.73E-03 5.30E-03 3.07 8.970 4.366 12 30 1.73E-03 6.11E-03 3.54 8.985 4.372 13 50 1.73E-03 7.32E-03 4.24 9.006 4.380 14 100 1.73E-03 9.29E-03 5.38 9.037 4.392 15 200 1.73E-03 1.21E-02 6.99 9.078 4.407 16 500 1.73E-03 1.58E-02 9.18 9.106 4.418 140 δ (ppm) Residuals 17 500 1.73E-03 1.78E-02 10.28 9.110 4.419 Binding Isotherm of Cl– with 5d • PF – 0.0086 CH(aryl) Residuals 0.35 CH(methyl) Residuals CH(aryl) Predicted 0.3 CH(methyl) Predicted CH(aryl) Experimental 0.25 CH(methyl) Experimental 0.2 0 0.15 0 2 4 6 8 10 12 14 16 0.1 0.05 0 0 2 4 6 8 10 12 14 [G]/[H] -0.008 [G]/[H] Figure C.7. Representative binding isotherm and residuals for Cl– titration of receptor 5d • PF6– in CD3CN at 25 °C determined by 1H NMR spectroscopy. Linear Regression Fitting and Statistics of log(KaR/KaH) and σp Figure C.8. Linear regression residuals of log(KaR/KaH) of Cl– and σp. 141 δ (ppm) Residuals X-Ray Crystallography Waiting on Lev for crystal data Computational Details General. Gas phase optimization and frequency analyses were performed for structures 4a-d and 5a-d via Gaussian 09.1 All computations were ran using B97-D3 density functional theory2 with a Def2-TZVP basis set.3 This work utilized the resources of the University of Oregon’s high-performance computer, Talapas. Electrostatic potential surface (ESP) maps and figures were generated using Gaussian 09. The −hole values (VS,max) were assessed using Multiwfn, an open-source software.4 Table C.7. Cartesian coordinates for 5a. X Y Z H 0.56677 3.82829 -1.70301 C 0.47593 2.87918 -1.18175 C 0.35303 0.46906 0.15123 C -0.729 2.18606 -1.13094 C -0.7917 0.97061 -0.45739 C 1.5382 1.19649 0.07591 142 H -1.61602 2.58904 -1.61598 H -1.73203 0.42135 -0.415 C 0.3235 -0.77869 0.84392 C 0.32554 -1.82742 1.43 C 0.34676 -3.08209 2.12455 C 0.31521 -5.58703 3.40306 C -0.70468 -3.97748 1.89278 C 1.37881 -3.44364 3.00595 C 1.36104 -4.69326 3.64051 C -0.71983 -5.21743 2.54135 H -1.50099 -3.71063 1.19888 H 2.16216 -4.9904 4.31555 H 0.32672 -6.56422 3.88476 I 3.00298 -2.17607 3.42256 S -2.03813 -6.37912 2.20256 O -2.68494 -5.90191 0.98831 O -1.41507 -7.69345 2.21649 C -3.12485 -6.14675 3.64653 F -4.24422 -6.91894 3.59412 F -2.49414 -6.44563 4.81836 F -3.54626 -4.85335 3.76174 N 1.57981 2.38196 -0.57785 C 2.85121 3.122 -0.65002 H 3.38779 2.99487 0.29388 H 2.64531 4.18346 -0.80889 H 3.42333 2.70879 -1.4843 H 2.46056 0.84427 0.53021 143 Table C.8. Cartesian coordinates for 5b. X Y Z H -1.53275 3.76493 -1.78037 C -1.38771 2.79981 -1.30259 C -1.01487 0.40151 0.00457 C -1.15997 1.64474 -2.04358 C -0.97322 0.43414 -1.38441 C -1.24573 1.58185 0.70669 H -1.12618 1.68659 -3.13058 H -0.79381 -0.47476 -1.9582 C -0.82222 -0.8228 0.71243 C -0.66365 -1.83578 1.33849 C -0.47692 -3.03151 2.10796 C -0.10428 -5.43182 3.51025 C -0.25728 -4.23164 1.41783 C -0.50752 -3.04179 3.50985 C -0.32245 -4.24067 4.2072 C -0.06163 -5.43967 2.10936 H -0.23499 -4.23489 0.32754 H -0.34636 -4.26692 5.29513 H 0.03456 -6.35798 4.06876 I -0.82882 -1.30413 4.64765 N -1.43308 2.75067 0.04866 C -1.66487 3.9834 0.82046 H -2.27816 3.74318 1.69296 H -2.19203 4.70953 0.19661 144 H -0.688 4.36885 1.12262 C 0.15158 -6.72706 1.34718 F 0.87956 -6.54656 0.21043 F -1.02302 -7.28983 0.95554 F 0.80825 -7.67594 2.06716 H -1.2821 1.60883 1.7926 Table C.9. Cartesian coordinates for 5c. X Y Z H -0.50817 4.43527 -1.02704 C -0.48071 3.41393 -0.65688 C -0.29988 0.84185 0.31428 C -1.55026 2.54177 -0.82891 C -1.45968 1.24356 -0.33803 C 0.74749 1.74798 0.46379 H -2.45066 2.87027 -1.3443 H -2.29234 0.553 -0.46944 C -0.17051 -0.48405 0.82552 C -0.01911 -1.59256 1.26293 C 0.19677 -2.90886 1.78928 C 0.50705 -5.50189 2.8093 C -0.91261 -3.75041 1.94049 C 1.46802 -3.37775 2.155 C 1.62015 -4.67304 2.66367 C -0.75745 -5.0417 2.44871 H -1.90818 -3.40942 1.66417 145 H 2.59794 -5.05314 2.9514 H 0.62686 -6.50848 3.20466 I 3.20272 -2.2032 1.9755 N 0.64043 3.00945 -0.01598 C 1.75685 3.95646 0.14343 H 2.49346 3.55124 0.84114 H 1.35543 4.89386 0.53772 H 2.20639 4.10522 -0.84168 H -1.62362 -5.68993 2.56328 H 1.67384 1.47574 0.96261 Table C.10. Cartesian coordinates for 5d. X Y Z H -0.54887 4.16441 -0.7044 C -0.51057 3.13011 -0.3736 C -0.31151 0.52189 0.49514 C -1.654 2.34004 -0.31213 C -1.55373 1.02329 0.12489 C 0.80703 1.34771 0.41795 H -2.62042 2.74803 -0.6018 H -2.44459 0.39789 0.17638 C -0.17184 -0.8223 0.9528 C -0.01105 -1.94687 1.34322 C 0.21552 -3.28282 1.81312 C 0.52861 -5.91198 2.71504 146 C -0.88599 -4.14825 1.90275 C 1.48227 -3.74708 2.18082 C 1.63549 -5.05929 2.63044 C -0.75653 -5.47874 2.35418 H -1.86683 -3.77401 1.61326 H 2.61127 -5.44248 2.92192 H 0.7066 -6.9238 3.07203 I 3.20389 -2.5473 2.09002 N 0.69275 2.6251 -0.01688 C 1.89375 3.47511 -0.07706 H 2.06597 3.87004 0.9272 H 1.73022 4.28734 -0.78969 H 2.73915 2.86676 -0.40925 C -1.9942 -6.39185 2.43575 C -2.62474 -6.53738 1.03391 H -1.90222 -6.94502 0.31692 H -2.97777 -5.57963 0.63624 H -3.48855 -7.21207 1.05718 C -3.03161 -5.77468 3.39845 H -2.6044 -5.62872 4.39768 H -3.90929 -6.42335 3.50238 H -3.39314 -4.801 3.05015 C -1.6871 -7.81715 2.94834 H -1.26652 -7.80011 3.96056 H -0.97955 -8.33808 2.29277 H -2.59754 -8.42723 2.98847 H 1.79828 1.0032 0.70098 147 Table C.11. Cartesian coordinates for 4a. X Y Z H 0.12833 4.06952 -1.50739 C 0.09491 3.1812 -0.8827 C 0.03654 0.8788 0.59272 C -0.17191 3.27212 0.47482 C -0.20114 2.0976 1.22376 C 0.29564 0.89081 -0.77145 H -0.3525 4.23526 0.93901 H -0.40727 2.13987 2.28997 N 0.32942 2.01632 -1.52557 C 0.01373 -0.34359 1.33028 C -0.00662 -1.36417 1.96397 C -0.03199 -2.58061 2.72196 C -0.07289 -5.02761 4.1081 C -0.0645 -3.79005 2.01574 C -0.0238 -2.60209 4.12583 C -0.04077 -3.82364 4.81352 C -0.09988 -5.0029 2.71216 H -0.05583 -3.77896 0.92616 H -0.02193 -3.8516 5.90152 H -0.06927 -5.96817 4.65676 I 0.02064 -0.85142 5.2863 S -0.11451 -6.54016 1.8013 O 0.52307 -6.25747 0.52367 O 0.46727 -7.53929 2.68596 C -1.8991 -6.82928 1.58524 148 F -2.16671 -7.97659 0.90112 F -2.55334 -6.92652 2.7788 F -2.50587 -5.81711 0.90043 H 0.48998 -0.02813 -1.31777 Table C.12. Cartesian coordinates for 4b. X Y Z H -0.61441 3.63972 -1.96172 C -0.53895 2.79105 -1.28756 C -0.36464 0.58185 0.31563 C -0.33555 2.97446 0.07163 C -0.24718 1.84722 0.88575 C -0.56718 0.50125 -1.05576 H -0.24789 3.97198 0.48754 H -0.08809 1.96166 1.95474 N -0.65718 1.57898 -1.87258 C -0.27887 -0.59429 1.12059 C -0.20507 -1.57517 1.81055 C -0.11666 -2.74377 2.63628 C 0.05041 -5.10381 4.14315 C -0.18469 -3.99616 2.00876 C 0.03487 -2.68174 4.02822 C 0.11909 -3.86099 4.7776 C -0.11009 -5.18559 2.7531 H -0.29901 -4.04976 0.92543 H 0.24008 -3.82845 5.85847 H 0.12133 -6.01176 4.74169 149 I 0.14907 -0.86677 5.07889 C -0.16172 -6.5257 2.05983 F -0.94596 -6.51885 0.94655 F 1.06761 -6.94205 1.65112 F -0.65058 -7.51908 2.853 H -0.66647 -0.45794 -1.55648 Table C.13. Cartesian coordinates for 4c. X Y Z H -0.25865 3.86017 -1.8606 C -0.23302 2.99815 -1.19981 C -0.15813 0.75799 0.3676 C -0.31725 3.15014 0.17583 C -0.27884 2.00716 0.97164 C -0.08068 0.70888 -1.01796 H -0.41053 4.1356 0.61829 H -0.34277 2.0964 2.05291 N -0.11545 1.80251 -1.81792 C -0.11516 -0.43375 1.15264 C -0.07923 -1.42898 1.82488 C -0.03577 -2.6154 2.6287 C 0.05698 -5.00698 4.09191 C 0.0479 -3.8504 1.97172 C -0.07296 -2.58755 4.03035 C -0.02645 -3.78348 4.75821 C 0.09408 -5.04084 2.70004 150 H 0.07779 -3.89261 0.88475 H -0.0543 -3.77786 5.84527 H 0.093 -5.93357 4.65986 I -0.19811 -0.80048 5.12778 H 0.15904 -5.9935 2.17987 H 0.01453 -0.23682 -1.54442 Table C.14. Cartesian coordinates for 4d. X Y Z H -0.03841 3.62563 -1.89212 C -0.03046 2.78586 -1.20288 C -0.04513 0.60119 0.44255 C 0.31885 2.95765 0.12793 C 0.31043 1.84289 0.96365 C -0.37971 0.53078 -0.90328 H 0.59119 3.93671 0.50593 H 0.57942 1.94808 2.01141 N -0.3803 1.59676 -1.74051 C -0.06423 -0.56177 1.27045 C -0.07811 -1.53237 1.97848 C -0.09364 -2.6899 2.82471 C -0.13067 -5.03315 4.35472 C -0.15686 -3.95149 2.21089 C -0.0484 -2.6109 4.21911 C -0.06724 -3.78194 4.97962 C -0.17676 -5.1487 2.95721 151 H -0.19087 -3.99424 1.12298 H -0.033 -3.74219 6.06609 H -0.1425 -5.90975 4.9973 I 0.04936 -0.78885 5.25586 C -0.24619 -6.5068 2.23602 C -1.534 -6.57894 1.38733 H -1.5567 -5.8155 0.60201 H -2.4252 -6.43685 2.0101 H -1.62591 -7.55187 0.89045 C 0.98089 -6.66774 1.31281 H 1.01089 -5.90612 0.52599 H 0.97446 -7.64367 0.81337 H 1.91516 -6.59015 1.88152 C -0.26077 -7.7213 3.19174 H 0.64495 -7.7607 3.80811 H -0.3107 -8.66344 2.63263 H -1.12885 -7.69792 3.86082 H -0.66609 -0.41026 -1.36486 References (1) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, R. J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al. Gaussian 09, Revision E.01. Wallingford, CT 2013. (2) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J. Comput. Chem. 2011, 32, 1456–1465. (3) Weigend, F.; Ahlrichs, R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. (4) Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580–592. 152 NMR Spectra N TMS Figure C.9. 1H NMR spectrum of TMS-protected 3-ethynylpyridine. S F3C O N H Figure C.10. 1H NMR spectrum of 4-(trifluoromethylthio)acetamide. Wet with water and ethyl acetate. 153 O O S F3C O N H Figure C.11. 1H NMR spectrum of N-[4-[(trifluoromethyl)sulfonyl]phenyl]acetamide. O O S F3C NH2 Figure C.12. 1H NMR spectrum of N-[4-[(Trifluoromethyl)sulfonyl]phenyl]aniline. 154 O O F C S I 3 NH2 Figure C.13. 1H NMR spectrum of 1a. O O S IF3C NH2 Figure C.14. 13C NMR spectrum of 1a. 155 Figure C.15. 19F NMR spectrum of 1a. O O S I F3C N3Et2 Figure C.16. 1H NMR spectrum of 2a. 156 O O S I F3C N3Et2 Figure C.17. 13C NMR spectrum of 2a. Figure C.18. 19F NMR spectrum of 2a. 157 N O O S CF3 Et2N3 Figure C.19. 1H NMR spectrum of 3a. Wet with water and grease. N O O S CF3 Et2N3 questionable carbon Figure C.20. 13C NMR spectrum of 3a. Contains grease. 158 Figure C.21. 19F NMR spectrum of 3a. N O O S CF3 I Figure C.22. 1H NMR spectrum of 4a. Wet with water and grease. 159 N O O S CF3 I questionable carbon Figure C.23. 13C NMR spectrum of 4a. Figure C.24. 19F NMR spectrum of 4a. 160 N I O O S CF3 I Figure C.25. 1H NMR spectrum of 5a • I–. Wet with water. Figure C.26. 13C NMR spectrum of 5a • I–. 161 Figure C.27. 19F NMR spectrum of 5a • I–. N PF6 O O S CF3 I Figure C.28. 1H NMR spectrum of 5a • PF6 –. 162 Figure C29. 13C NMR spectrum of 5a • PF6 –. Figure C.30. 19F NMR spectrum of 5a • PF6 –. 163 F3C I NH2 Figure C.31. 1H NMR spectrum of 1b. F3C I N3Et2 Figure C.32. 1H NMR spectrum of 2b. 164 Figure C.33. 1H NMR spectrum of 3b. N CF3 I Figure C.34. 1H NMR spectrum of 4b. 165 N CF3 I questionable carbon here need another in chloroform or if good, need to retake proton and fluorine in acetone Figure C.35. 13C NMR spectrum of 4b. Figure C.36. 19F NMR spectrum of 4b. 166 N I CF3 I Figure C.37. 1H NMR spectrum of 5b • I–. Figure C.38. 13C NMR spectrum of 5b • I–. 167 Figure C.39. 19F NMR spectrum of 5b • I–. N PF6 CF3 I need in acetone Figure C.40. 1H NMR spectrum of 5b • PF6 –. 168 Figure C.41. 13C NMR spectrum of 5b • PF6 –. Figure C.42. 19F NMR spectrum of 5b • PF6 –. 169 I N3Et2 Figure C.43. 1H NMR spectrum of 2c. N Et2N3 need dried proton Figure C.44. 1H NMR spectrum of 3c. 170 Figure C.45. 13C NMR spectrum of 3c. N I Figure C.46. 1H NMR spectrum of 4c. 171 N I I Figure C.47. 1H NMR spectrum of 5c • I–. Figure C.48. 13C NMR spectrum of 5c • I–. 172 N PF6 I Figure C.49. 1H NMR spectrum of 5c • PF6 –. N PF6 I Figure C.50. 13C NMR spectrum of 5c • PF6 –. 173 Figure C.51. 19F NMR spectrum of 5c • PF6 –. tBu I NH2 Figure C.52. 1H NMR spectrum of 1d. Wet with DCM and water. 174 tBu I N3Et2 Figure C.53. 1H NMR spectrum of 2d. N tBu Et2N3 Figure C.54. 1H NMR spectrum of 3d. Wet with water and ethyl acetate. 175 N tBu Et2N3 Figure C.55. 13C NMR spectrum of 3d. N tBu I Figure C.56. 1H NMR spectrum of 4d. Wet with water and ethyl acetate. 176 N tBu I Figure C.57. 13C NMR spectrum of 4d. N I tBu I Figure C.58. 1H NMR spectrum of 5d • I–. Wet with water. 177 Figure C.59. 13C NMR spectrum of 5d • I–. Figure C.60. 19F NMR spectrum of 5d • I–. 178 N PF6 tBu I Figure C.61. 1H NMR spectrum of 5d • PF6 –. N PF6 tBu I checking this with Mike Figure C.62. 13C NMR spectrum of 5d • PF6 –. 179 Figure C.63. 19F NMR spectrum of 5d • PF6 –. 180 APPENDIX D SUPPLEMENTARY CONTENT FOR CHAPTER FIVE Synthesis General Methods: All reagents were purchased from conventional commercial sources and used as received, unless otherwise noted. All NMR spectra were acquired at room temperature on a Bruker Avance-III-HD 600 MHz (1H 600 MHz, 13C 151 MHz, 19F 565 MHz) spectrometer with a Prodigy multinuclear broadband BBO CryoProbe. 1H and 13C chemical shifts () are reported in ppm relative to the residual CHCl3 (1H: 7.26 ppm, 13C: 77.16 ppm) or CH3CN (1H: 1.94, 13C: 118.26). Methylated Urea and Iodo-Pyridine (1.6) The receptor (0.030g, 0.0518 mmol) was weighed out into a narrow bomb flask and a small stir bar was added. Then ~1 to 2 mL of MeI was carefully pipetted over the scaffold until it was fully submerged. The bomb flask was then sealed and placed in a metal bead bath at room temperature. This bath was then set to heat to 110 °C and watched carefully. As soon as there was visible precipitation of the methylated receptor, as an orange-brown solid, the flask was removed from heat. This typically happened as soon as the flask had reached temperature, if not slightly before. The solution was let to cool down to room temperature and, upon chilling, was 181 carefully filtered. The precipitate was then washed with cold DCM before being scraped off of the filter paper into a massed scintillation vial to yield 0.026g of product (70%). 1H NMR (500 MHz, Acetone-d6) δ = 9.25 (s, 1H), 9.05 (s, 1H), 8.62 (d, J=4.9, 1H), 8.50 – 8.47 (m, 2H), 8.22 (d, J=2.1, 1H), 8.09 (dd, J=8.9, 2.1, 1H), 7.77 (d, J=8.5, 3H), 7.73 (d, J=4.9, 1H), 7.65 (d, J=8.5, 3H), 4.37 (q, J=7.1, 3H), 2.84 (s, 3H), 1.38 (t, J=7.1, 3H). 13C NMR (126 MHz, Acetone-d6) δ = 165.53, 157.96, 152.30, 149.66, 145.41, 143.81, 137.04, 134.49, 132.66, 127.59, 126.97 (q, J=3.9), 125.31, 124.49, 124.43, 119.55, 119.38, 111.02, 99.94, 96.49, 92.09, 61.62, 54.96, 14.58. 19F NMR (471 MHz, Acetone-d6) δ = -62.32. Tridentate Host Receptor (1) The methylated scaffold (0.026g, 0.03605 mmol) was dissolved in ~6 mL of acetone (sonicating as needed to help solubilize it). Then ~3 mL of water was added to the scintillation vial which was set stirring. AgPF6 was then added and the reaction was left to stir for 1-2 hours. Upon visible precipitation the solution was filtered until translucent (ppt is not product) the filtrate is then subjected to reduced pressure to remove the acetone, at which point the receptor will crash out of the water. This aqueous solution was then filtered a second time to yield the receptor as a white solid (0.025g, 94%). 1H NMR (500 MHz, Acetone-d6) δ = 9.14 (s, 1H), 9.08 (s, 1H), 8.66 182 (d, J=4.9, 1H), 8.49 (d, J=8.9, 1H), 8.47 (s, 1H), 8.23 (d, J=2.1, 1H), 8.13 – 8.09 (m, 1H), 7.75 (t, J=7.5, 3H), 7.66 (d, J=8.5, 2H), 4.37 (q, J=7.1, 2H), 1.38 (t, J=7.1, 3H). 19F NMR (471 MHz, Chloroform-d) δ = -57.14, -66.70, -68.20. 31P NMR (202 MHz, Acetone-d6) δ = -134.68 – - 153.89 (m). Protected Quinoline (2.1) Aminoquinoline (1.00g, 6.9 mmols) was dissolved in DCM and set stirring in an ice bath. Once the reaction mixture reached 0 °C TEA (.91 mL, 7.6 mmols) was added to the flask followed by pivaloyl chloride (1.2 mL, 8.7mmols). After stirring for 15 minutes the ice bath was removed and the reaction was monitored by TLC. After completion the reaction mixture was washed with three times with NaHCO3 and dried over Na2SO4. Finally, solvent was removed in vacuo to afford a pale-yellow oil (1.56 g, 6.83 mmols, 99% yield). If needed the product was purified by column chromatography over ethyl acetate and hexanes, though this could typically be avoided by running the reaction overnight. 1H NMR (500 MHz, Chloroform-d) δ 8.46 (d, J = 9.0 Hz, 1H), 8.27 (s, 1H), 8.17 (d, J = 9.0 Hz, 1H), 7.82 (dq, J = 8.4, 0.9 Hz, 1H), 7.78 (dd, J = 8.0, 1.5 Hz, 1H), 7.66 (ddd, J = 8.4, 6.9, 1.5 Hz, 1H), 7.45 (ddd, J = 8.1, 6.9, 1.2 Hz, 1H), 1.38 (s, 7H). 13C NMR (126 MHz, Chloroform-d) δ = 177.67, 151.38, 146.70, 138.69, 130.12, 127.72, 127.32, 126.50, 125.24, 114.42, 40.14, 27.65. 183 Protected Iodo-Quinoline (2.2) Three round bottoms and two stir bars were left in a 150 °C oven to dry overnight. Dry Et2O was procured from a solvent still, transferred into one of the dry round bottom flasks with a dry stir bar in it and set under N2. Some of this Dry Et2O was cannulated over to a second RBF, that has a stir bar and I2 (1.24 equiv.) already in it. This second flask was set stirring, so the I2 can fully dissolve before it is needed. The protected quinoline was then added to the first flask and set this stirring in a salted ice bath until it reached -20 °C. After that 2.2M nBuLi in hexanes was added dropwise via syringe at which point the pale yellow solution grew brighter before changing to a lime green color. After 3 hours of stirring at -20 °C the reaction mixture was a solid brown and is cooled to -80 °C in an acetone and dry ice bath. Once the desired temperature had been reached the solution of I2 was cannulated in dropwise at which point the mixture changed back to a warmer yellow. The reaction mixture was left to stir overnight, slowly coming back up to room temperature. The next day the mixture was washed with saturated aqueous solution of Na2S2O3, then dried and solvent removed. 1H NMR (500 MHz, Chloroform-d) δ = 8.61 (s, 1H), 8.27 (s, 1H), 8.03 (d, J=8.5, 1.0, 1H), 7.69 (ddd, J=8.4, 6.9, 1.4, 1H), 7.65 (d, J=8.2, 1.4, 1H), 7.48 (ddd, J=8.1, 6.9, 1.2, 1H), 1.42 (s, 9H). 13C NMR (126 MHz, Chloroform-d) δ = 176.29, 148.40, 148.01, 146.52, 130.66, 128.87, 127.62, 126.56, 126.32, 85.86, 40.52, 27.75. 184 3-Iodo-2-Amino-Quinoline (2.3) Iodoaminoquinoline (1.5 mmols, 0.55g, 1 equiv) was dissolved in a minimal amount of dioxane and set stirring in a round bottom flask. Then a solution of 4 N HCl (9.3 mmols, 2.39 mL, 6 equiv.) was added and let stir at room temperature for about an hour and a half. Then the solution was heated to reflux and left stirring overnight. The next day the reaction mixture has transitioned from a translucent pale yellow to a dark blue. The reaction mixture was then quenched with saturated aqueous solution of NaHCO3 until a basic pH was reached and washed in ether with brine. Finally, the organic phase was dried with Na2SO4, filtered and the solvent was removed in vacuo to give a crude product. This was purified by an ethyl acetate/hexanes plug to give pure product in an 83% yield. 1H NMR (500 MHz, Chloroform-d) δ = 8.40 (s, 1H), 7.63 (dd, J=8.5, 1.1, 1H), 7.57 (ddd, J=8.4, 6.8, 1.5, 1H), 7.52 (dd, J=8.0, 1.4, 1H), 7.25 (ddd, J=8.1, 6.8, 1.2, 1H), 5.40 (s, 2H). 13C NMR (126 MHz, CDCl3) δ 155.01, 147.37, 147.25, 130.41, 126.42, 125.92, 125.21, 123.15, 80.51, 77.04, 0.03. TMS Protected Diyne Quinoline (2.4) TMS-protected buta-1,3-diyne was selectively deprotected according to known literature procedures with MeLiLiBr. Iodoaminoquinoline was dissolved in a 75 mL of a 50/50 mixture of THF and TEA and set stirring under N2 sparging until completely outgassed. The CuI (0.02 185 equiv.) and Pd(PPh3)2Cl2 (0.05 mol) were added and the solution stirred at room temperature overnight. The next day, after verifying that the reaction ran to completion by TLC, the solvent was removed in vacuo and the product purified by column chromatography on silica gel with EtOAc and hexanes. 1H NMR (500 MHz, Chloroform-d) δ = 8.07 (s, 1H), 7.60 (dt, J=8.8, 1.0, 1H), 7.56 (dtd, J=8.8, 3.7, 3.2, 1.5, 2H), 7.26 – 7.22 (m, 1H), 5.39 (s, 2H), 0.26 (s, 10H). 13C NMR (126 MHz, Chloroform-d) δ = 156.86, 147.40, 142.69, 1313C NMR (126 MHz, Chloroform-d) δ = 156.86, 147.40, 142.69, 131.35, 127.58, 126.05, 123.36, 122.97, 104.74, 93.49, 87.32, 80.30, 72.02, -0.33.1.35, 127.58, 126.05, 123.36, 122.97, 104.74, 93.49, 87.32, 80.30, 72.02, -0.33. Iodo Diyne Quinoline (2.5) The diyne quinoline (1.00 equiv) was dissolved in ACN and set stirring under N2 to outgas. The round bottom flask was then wrapped in aluminum foil, to ensure dark conditions, and AgF (1.00 equiv.) was added and let stir for 20 minutes. After 20 minutes NIS (1.20 equiv.) was added, and the reaction was monitored by TLC until done. Upon the disappearance of all starting material the reaction was rotovapped down and then washed in water and EtOAc three times before drying over Na2SO4 and removing the solvent to yield the product as a brown solid. This product was collected and used as is for the next step without further purification. 186 Cyclized Quinoline (2.6Bn) The iodo-diyne (1 equiv.) was dissolved in a minimal amount of DMF and set stirring under N2. Following outgassing lutidine (4 mol%), the azide (1 equiv.), and CuI(PPh3)3 (10 mol%) are added. These are left stirring at room temperature overnight. After completion the reaction mixture was diluted with EtOAc and washed with a saturated aqueous solution of NH4Cl three times. The organic layer was then dried with Na2SO4 and concentrated under reduced pressure to yield the crude product as an oil. This was purified by column chromatography over silica gel using a EtOAc and Hexanes mixture (3:1) to yield the product. Quinoline Host Receptor (2Bn) The newly cyclized compound (0.0429 g, 0.0951 mmol) was dissolved in a minimal amount of a 50/50 mixture of water and 1,4 dioxane and set stirring. Cyanamide (0.0048 g, 0.1141 mmol) and Sc(OTf)3 catalyst (0.0047g, 0.0095 mmol) were added to the reaction. This was then heated to 100 °C and left stirring for 2 days. The crude mixture was then taken up in EtOAc and washed with a saturated solution of NaCO3H, to crash the product out as its carbonate salt analog. While 187 the final product was never able to be isolated alone its existence in the crude reaction mixture was confirmed by mass spec. HRMS (TOF-MS-ES+) for C21H16IN7 [M+H] +: calcd 493.05, found 493.0512. 188 NMRs Figure D.1. 1H NMR spectrum of protected quinoline Figure D.2. 13C NMR spectrum of protected quinoline 189 Exp37-Product-CDCl3.1.fid I O CH3 N N CH3 H CH3 8.7 8.6 8.5 8.4 8.3 8.2 8.1 8.0 7.9 7.8 7.7 7.6 7.5 7.4 7.3 7.2 7.1 f1 (ppm) 9.5 9.0 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 -0.5 -1.0 f1 (ppm) Figure D.3. 1H NMR spectrum of protected iodo-quinoline Exp37-Carbon-CDCl3.1.fid I O CH3 N N CH3 H CH3 220 210 200 190 180 170 160 150 140 130 120 110 100 90 80 70 60 50 40 30 20 10 0 -10 -20 f1 (ppm) Figure D.4. 13C NMR spectrum of protected iodo-quinoline 190 0.97 8.6134 0.97 0.90 0.96 0.96 1.03 176.2920 1.00 0.90 8.2677 7.2600 CDCl3 148.4029 148.0058 8.03600.96 146.5197 8.0191 130.6637 128.8700 127.6196 7.7090 126.5621 7.7062 126.3186 7.6923 7.6782 0.96 7.6754 1.03 7.6590 7.6428 7.4936 1.00 7.4776 85.8587 7.4635 77.1600 CDCl3 7.2600 CDCl3 40.5184 8.66 27.7478 Exp6-HJB-Acetone.1.fid I N NH2 11.0 10.5 10.0 9.5 9.0 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 f1 (ppm) Figure D.5. 1H NMR spectrum of 3-Iodo-2-Amino-Quinoline Figure D.6. 13C NMR spectrum of 3-Iodo-2-Amino-Quinoline 191 1.00 8.5682 7.6476 1.08 7.6334 2.01 7.5426 1.07 7.2325 1.68 6.0163 2.8435 Water 2.0500 Acetone Exp33.2-F5-HJB-CDCl3.1.fid H3C CH3 Si CH3 N NH2 8.1 8.0 7.9 7.8 7.7 7.6 7.5 7.4 7.3 7.2 5.5 5.4 5.3 f1 (ppm) 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 -0.5 f1 (ppm) Figure D.7. 1H NMR spectrum of TMS protected diyne quinoline Exp33-F5-Carbon-CDCl3.1.fid H3C CH3 Si CH3 N NH2 160 150 140 130 120 110 100 90 80 70 60 50 40 30 20 10 0 f1 (ppm) Figure D.8. 13C NMR spectrum of TMS protected diyne quinoline 192 156.8645 1.00 1.00 8.0749 147.3993 142.6889 1.02 2.09 0.87 7.2600 CDCl3 131.3534 127.5785 126.0489 123.3629 7.6129 122.9708 7.5956 1.02 7.5718 7.5627 2.09 7.5582 104.7448 7.5501 1.81 7.5471 93.4887 7.2600 CDCl3 7.2470 87.3176 0.87 7.2443 7.2310 80.3007 7.2285 4.1110 EtOAc 77.1600 CDCl3 72.0243 1.81 2.0416 Acetone 1.2547 EtOAc 0.8801 EtOAc 8.80 0.2551 -0.3333 Exp13-HJB-PT-RC-CDCl3.1.fid N N N I N NH2 8.3 8.2 8.1 8.0 7.9 7.8 7.7 7.6 7.5 7.4 7.3 7.2 7.1 f1 (ppm) 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 f1 (ppm) Figure D.9. 1H NMR spectrum of cyclized quinoline TPD_3_12_ppt_acetone.1.fid F F F O NH NH H3C O I O N 9.3 9.2 9.1 9.0 8.9 8.8 8.7 8.6 8.5 8.4 8.3 8.2 8.1 8.0 7.9 7.8 7.7 7.6 7.5 f1 (ppm) 10.5 10.0 9.5 9.0 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 f1 (ppm) Figure D.10. 1H NMR spectrum of precursor 1.5 193 1.00 9.1289 1.00 8.1622 1.02 9.0629 2.04 1.14 7.6369 3 00 7.5807. 1.00 2 23 7.3690. 1.02 0 81 7.2600 CDCl3. 0.93 1.03 8.63950.93 0.88 8.6305 0.92 8.49951.03 1.05 8.48180.88 2.16 8.4578 0.99 1.95 0.92 8.2278 2.01 5.6366 8.2238 2.04 5.4135 5.2994 DCM 1.05 8.1101 8.0884 7.7706 2 16 7.7535. 0 99 7.7141 1.00 . 1.95 7.7044 7.6660 7.6488 3.4850 Et2O 2.35 4.3758 4.3616 2.04 1.14 2.8108 3.00 1.2078 Et2O 2.23 0.81 7.2600 CDCl3 2.0500 Acetone 3.14 1.3811 Figure D.11. 19F NMR spectrum of precursor 1.5 Exp56-Methylated-Acetone.1.fid F F F O H NH3C O NH O I 9.3 9.2 9.1 9.0 8.9 8.8 8.7 8.6 8.5 8.4 8.3 8.2 8.1 8.0 7.9 7.8 7.7 7.6 7.5 I- f1 (ppm) N+ CH3 10.0 9.5 9.0 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 f1 (ppm) Figure D.12. 1H NMR spectrum of precursor 1.6 194 1.05 9.2515 0.96 9.0520 1.05 0.96 8.6281 8.6182 1.00 8.4949 1.74 1.00 8.4842 0.93 8.4773 1.02 1.74 8.2259 1.87 8.2218 1.01 8.1051 1.97 0.93 8.1011 1.02 8.0874 8.0833 7.7750 7.7578 1.87 7.7331 1.01 7.7232 1.97 7.6628 7.6454 5.6212 DCM 4.3869 4.3726 2.15 4.3584 4.3442 2.8766 H2O 3.38 2.8434 2.0500 Acetone 2.99 1.3786 Figure D.13. 13C NMR spectrum of precursor 1.6 Figure D.14. 19F NMR spectrum of precursor 1.6 195 Exp59-moredilute-acetone.2.fid 9.4 9.3 9.2 9.1 9.0 8.9 8.8 8.7 8.6 8.5 8.4 8.3 8.2 8.1 8.0 7.9 7.8 7.7 7.6 7.5 7.4 f1 (ppm) 10.0 9.5 9.0 8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 f1 (ppm) Figure D.15. 1H NMR spectrum of Host Receptor 1PF6 Figure D.16. 19F NMR spectrum of Host Receptor 1PF6 196 0.75 0.93 0.75 9.1446 0.93 9.0836 1.00 0.60 0.96 0.94 1.00 3.10 2.00 8.66211.00 8.6523 8.4990 0.60 8.4813 0.96 8.4681 0.94 8.2307 8.2270 1.00 8.1128 8.0987 5.6231 DCM 7.7705 3.10 7.7517 7.7406 2.14 2.00 7.6688 7.6516 2.8365 Water 2.0500 Acetone 3.27 Figure D.17. 31P NMR spectrum of Host Receptor 1PF6 197 1H NMR Titrations. General Methods. 1H NMR spectra were acquired at room temperature on a Varian Inova 500 MHz spectrometer (1H: 500.11 MHz). 1H chemical shifts () are expressed in ppm relative to residual CH3CN (1H: 1.94 ppm) shifts. A solution of host in 10% DMSO-d6/CD3CN (0.73-1.9 mM) was prep in 3 or 4 mL volumetric flasks. 500 L of this host solution was added to a septum-sealed NMR tube. The remaining host solution was used to prepare a host/guest (NBu4Cl, NBu4Br, or NBu4I) stock solution in 2 or 3 mL volumetric flasks, respectively. Aliquots of the host/guest solution were incrementally added to the NMR tube using gas-tight Hamilton syringes, and 1H NMR spectra were recorded, at 25 °C, after each addition of guest. The ∆ of the NH protons were tracked to follow the progress of the titration and association constants were determined using Thordarson’s Bindfit software. Table D.1. Representative titration of receptor 1 with Cl- in 10% DMSO-d6/CD3CN. Entry VGuest (𝜇L) [Guest] (M) [Host] (M) Equiv. Anion 𝛅1 (ppm) 𝛅2 (ppm) 0 0 0 0.000946816 0.00 9.66250 8.515 1 5 0.000275384 0.000946816 0.29 9.70870 8.531 2 5 0.000545368 0.000946816 0.58 9.87850 8.5806 3 10 0.00106976 0.000946816 1.13 10.28150 8.8036 4 10 0.001574364 0.000946816 1.66 10.56070 8.933 5 10 0.002060278 0.000946816 2.18 10.77210 9.0358 6 15 0.002756318 0.000946816 2.91 10.99810 9.1398 7 15 0.003415725 0.000946816 3.61 11.15700 9.2183 8 15 0.004041315 0.000946816 4.27 11.27180 9.2709 9 20 0.004827181 0.000946816 5.10 11.38460 9.3212 10 20 0.005562752 0.000946816 5.88 11.47290 9.3645 11 20 0.006252706 0.000946816 6.60 11.53340 9.3942 12 30 0.007210975 0.000946816 7.62 11.60970 9.4305 13 50 0.008631856 0.000946816 9.12 11.68670 9.4651 14 100 0.010956936 0.000946816 11.57 11.78640 9.5129 15 200 0.014246072 0.000946816 15.05 11.87130 9.5539 16 500 0.018694494 0.000946816 19.74 11.92860 9.5814 17 500 0.020946164 0.000946816 22.12 11.9434 9.5876 198 2.5 Predicted N(H)1 2 Predicted N(H)2 Experimental 1.5 N(H)1 Experimental N(H)2 1 0.5 0 0 5 10 15 20 25 [G]/[H] 0.2 N(H)1 N(H)2 0 0 5 10 15 20 -0.2 [G]/[H] Figure D.19. Representative binding isotherm for Cl- titration of receptor 1 in 10% DMSO- d6/CD3CN as determined by 1H NMR spectroscopy. Table D.2. Representative titration of receptor 1 with Br- in 10% DMSO-d6/CD3CN. Entry Vguest (𝜇L) [Guest] (M) [Host] (M) Equiv. Anion 𝛅1 (ppm_ 𝛅2 (ppm) Host Alone 0 0 0.00132554 0 9.66550 8.51620 1 2.5 0.00024569 0.00132554 0.19 9.66700 8.51720 2 2.5 0.00048896 0.00132554 0.37 9.68070 8.52180 3 5 0.00096832 0.00132554 0.73 9.70790 8.53520 4 5 0.00143839 0.00132554 1.09 9.76420 8.56010 5 10 0.00235165 0.00132554 1.77 9.85890 8.60300 6 10 0.00323077 0.00132554 2.44 9.93740 8.63740 199 Residuals 𝜹 - 𝜹 0 7 10 0.00407762 0.00132554 3.08 10.00420 8.66780 8 10 0.00489396 0.00132554 3.69 10.06420 8.69550 9 15 0.00606477 0.00132554 4.58 File Corrupted 10 15 0.00717553 0.00132554 5.41 10.20370 8.75910 11 15 0.00823076 0.00132554 6.21 10.25520 8.78090 12 20 0.0095583 0.00132554 7.21 10.31050 8.80530 13 30 0.01139644 0.00132554 8.60 10.38580 8.83800 14 50 0.01410987 0.00132554 10.64 10.47160 8.87870 15 100 0.01851921 0.00132554 13.97 10.58820 8.92280 16 200 0.02469228 0.00132554 18.63 10.70940 8.98500 17 300 0.03039049 0.00132554 22.93 10.79140 9.02320 18 500 0.03566662 0.00132554 26.91 10.84980 9.04790 19 500 0.03864878 0.00132554 29.16 10.86270 9.05340 1.2 1 0.8 0.6 0.4 Predicted N(H)1 0.2 Predicted N(H)2 Experimental 0 N(H)1 0 2 4 6 8 10 12 14 16 [G]/[H] 0.2 N(H)1 0.1 N(H)2 0 0 5 10 15 20 -0.1 [G]/[H] Figure D.20. Representative binding isotherm for Br- titration of receptor 1 in 10% DMSO- d6/CD3CN as determined by 1H NMR spectroscopy. 200 Residuals 𝜹 - 𝜹 0 Table D.3. Representative titration of receptor 1 with I- in 10% DMSO-d6/CD3CN. Vguest Equiv. Entry (𝜇L) [Guest] (M) [Host] (M) Anion 𝛅1 (ppm_ 𝛅2 (ppm) Host Alone 0 0 0.0010415 0 9.6622 8.5159 1 5 0.00025483 0.0010415 0.24467354 9.6641 8.517 2 5 0.00050466 0.0010415 0.48454956 9.6884 8.5273 3 10 0.0009899 0.0010415 0.95046259 9.6947 8.5326 4 10 0.00145684 0.0010415 1.398794 9.7042 8.5358 5 10 0.00190648 0.0010415 1.83052055 9.7095 8.5385 6 15 0.00255056 0.0010415 2.44893965 9.7131 8.5405 7 15 0.00316075 0.0010415 3.03481038 9.7168 8.5418 8 15 0.00373964 0.0010415 3.59063646 9.7211 8.544 9 20 0.00446684 0.0010415 4.28886426 9.7292 8.547 10 20 0.0051475 0.0010415 4.94240548 9.7317 8.5472 11 20 0.00578595 0.0010415 5.55541701 9.7327 8.5487 12 30 0.00667269 0.0010415 6.40682192 9.738 8.5506 13 50 0.00798751 0.0010415 7.66924988 9.7437 8.5537 14 100 0.01013902 0.0010415 9.73504109 9.755 8.5587 15 200 0.01318263 0.0010415 12.6573799 9.7683 8.5667 16 500 0.01729899 0.0010415 16.6097233 9.7885 8.5841 17 500 0.01938257 0.0010415 18.6102922 9.7969 8.5769 18 500 0.02064098 0.0010415 19.8185566 9.8006 8.5802 201 0.16 0.14 0.12 0.1 0.08 0.06 Predicted N(H) 1 0.04 Predicted N(H) 2 0.02 Experimental N(H) 1 Experimental N(H) 2 0 0 5 10 15 20 25 [G]/[H] 0.02 N(H) 1 N(H) 2 0 0 5 10 15 20 25 -0.02 [G]/[H} Figure D.21. Representative binding isotherm for I- titration of receptor 1 in 10% DMSO- d6/CD3CN as determined by 1H NMR spectroscopy. 202 Residuals 𝜹 - 𝜹 0 REFERENCES Chapter 1: (1) Sessler, J. L.; Gale, P. A.; Cho, W.-S. Anion Receptor Chemistry; Royal Society of Chemistry: Cambridge, 2006. (2) Zyryanov, G. V; Palacios, M. A.; Anzenbacher Jr., P. Rational Design of a Fluorescence- Turn-On Sensor Array for Phosphates in Blood Serum Zuschriften. Angew. Chemie 2007, 119, 7995–7998. (3) Hruska, K. A.; Mathew, S.; Lund, R.; Qiu, P.; Pratt, R. Hyperphosphatemia of Chronic Kidney Disease. Kidney Int. 2008, 74 (2), 148–157. (4) Schmidt, R. F.; Thews, G. Human Physiology, 2nd ed.; Springer: Berlin, 1989. (5) Kanbay, M.; Goldsmith, D.; Akcay, A.; Covic, A. Phosphate - The Silent Stealthy Cardiorenal Culprit in All Stages of Chronic Kidney Disease: A Systematic Review. Blood Purif. 2009, 27 (2), 220–230. (6) Feng, H.; Wang, Y.; Jia, H.; Zhang, R.; Han, Q.; Meng, Q.; Zhiqiang, Z. Selective Detection of Inorganic Phosphates in Live Cells Based on a Responsive Fluorescence Probe. New J. Chem. 2017, 41, 9623–9630. (7) Scott, C. C.; Gruenberg, J. Ion Flux and the Function of Endosomes and Lysosomes: PH Is Just the Start: The Flux of Ions across Endosomal Membranes Influences Endosome Function Not Only through Regulation of the Luminal PH. BioEssays 2011, 33 (2), 103– 110. (8) Jentsch, T. J.; Stein, V.; Weinreich, F.; Zdebik, A. A. Molecular Structure and Physiological Function of Chloride Channels. Physiol. Rev. 2002, 82 (2), 503–568. (9) Berend, K.; Van Hulsteijn, L. H.; Gans, R. O. B. Chloride: The Queen of Electrolytes? Eur. J. Intern. Med. 2012, 23 (3), 203–211. 203 (10) Beck, A.; Ernst, M. Kinetic Modeling and Selectivity of Anion Exchange in Donnan Dialysis. J. Memb. Sci. 2015, 479, 132–140. (11) Matsuyama, K.; Sen, A. C.; Perrin, J. H. The Effects of PH, Calcium and Chloride Ions on the Binding of Tolmetin to Human Serum Albumin: Circular Dichroic, Dialysis and Fluorometric Measurements. J. Pharm. Pharmacol. 1987, 39 (3), 190–195. (12) Scanziani, M.; Häusser, M. Electrophysiology in the Age of Light. Nature 2009, 461 (7266), 930–939. (13) Saunders, J. H. Liquid and Solid-State Cl- -Sensitive Microelectrodes. Characteristics and Application to Intracellular Cl- Activity in Balanus Photoreceptor. J. Gen. Physiol. 1977, 70 (4), 507–530. (14) Nagai, T.; Ibata, K.; Park, E. S.; Kubota, M.; Mikoshiba, K.; Miyawaki, A. A Variant of Yellow Fluorescent Protein with Fast and Efficient Maturation for Cell-Biological Applications. Nat. Biotechnol. 2002, 20, 87–90. (15) Tutol, J. N.; Peng, W.; Dodani, S. C. Discovery and Characterization of a Naturally Occurring, Turn-On Yellow Fluorescent Protein Sensor for Chloride. Biochemistry 2019, 58 (1), 31–35. (16) Yoon, D. W.; Gross, D. E.; Lynch, V. M.; Lee, C. H.; Bennett, P. C.; Sessler, J. L. Real- Time Determination of Chloride Anion Concentration in Aqueous-DMSO Using a Pyrrole-Strapped Calixpyrrole Anion Receptor. Chem. Commun. 2009, No. 9, 1109–1111. (17) Langton, M. J.; Serpell, C. J.; Beer, P. D. Anion Recognition in Water: Recent Advances from a Supramolecular and Macromolecular Perspective. Angew. Chemie - Int. Ed. 2016, 55 (6), 1974–1987. (18) Yu, X. H.; Hong, X. Q.; Mao, Q. C.; Chen, W. H. Biological Effects and Activity Optimization of Small-Molecule, Drug-like Synthetic Anion Transporters. Eur. J. Med. Chem. 2019, 184, 111782. 204 (19) Liu, Y.; Zhao, W.; Chen, C.-H.; Flood, A. H. Chloride Capture Using a C–H Hydrogen- Bonding Cage. Science (80-. ). 2019, 365, 159–161. (20) Tutol, J. N.; Lee, J.; Chi, H.; Faizuddin, F. N.; Abeyrathna, S. S.; Zhou, Q.; Morcos, F.; Meloni, G.; Dodani, S. C. A Single Point Mutation Converts a Proton-Pumping Rhodopsin into a Red-Shifted, Turn-on Fluorescent Sensor for Chloride. Chem. Sci. 2021, 12, 5655–5663. (21) Watt, M. M.; Engle, J. M.; Fairley, K. C.; Robitshek, T. E.; Haley, M. M.; Johnson, D. W. “Off-on” Aggregation-Based Fluorescent Sensor for the Detection of Chloride in Water. Org. Biomol. Chem. 2015, 13 (14), 4266–4270. (22) Khatua, S.; Choi, S. H.; Lee, J.; Kim, K.; Do, Y.; Churchill, D. G. Aqueous Fluorometric and Colorimetric Sensing of Phosphate Ions by a Fluorescent Dinuclear Zinc Complex. Inorg. Chem. 2009, 48 (7), 2993–2999. (23) Zhang, Y. M.; Lin, Q.; Wei, T. B.; Wang, D. D.; Yao, H.; Wang, Y. L. Simple Colorimetric Sensors with High Selectivity for Acetate and Chloride in Aqueous Solution. Sensors Actuators, B Chem. 2009, 137 (2), 447–455. (24) Bazany-Rodríguez, I. J.; Martínez-Otero, D.; Barroso-Flores, J.; Yatsimirsky, A. K.; Dorazco-González, A. Sensitive Water-Soluble Fluorescent Chemosensor for Chloride Based on a Bisquinolinium Pyridine-Dicarboxamide Compound. Sensors Actuators, B Chem. 2015, 221, 1348–1355. (25) Huang, X. H.; Lu, Y.; He, Y. B.; Chen, Z. H. A Metal-Macrocycle Complex as a Fluorescent Sensor for Biological Phosphate Ions in Aqueous Solution. European J. Org. Chem. 2010, No. 10, 1921–1927. (26) Martí, I.; Bolte, M.; Burguete, M. I.; Vicent, C.; Alfonso, I.; Luis, S. V. Tight and Selective Caging of Chloride Ions by a Pseudopeptidic Host. Chem. - A Eur. J. 2014, 20 (24), 7458–7464. 205 (27) Kim, D. S.; Lynch, V. M.; Nielsen, K. A.; Johnsen, C.; Jeppesen, J. O.; Sessler, J. L. A Chloride-Anion Insensitive Colorimetric Chemosensor for Trinitrobenzene and Picric Acid. Anal. Bioanal. Chem. 2009, 395 (2), 393–400. (28) Buhlmann, P.; Nishizawa, S.; Xiao, K. P.; Umezawa, Y. Strong Hydrogen Bond-Mediated Complexation of H2PO4- by Neutral Bis-Thiourea Hosts. Tetrahedron 1997, 53 (5), 1647–1654. (29) Scherer, M.; Sessler, J. L.; Gebauer, A.; Lynch, V. A Bridged Pyrrolic Ansa-Ferrocene. A New Type of Anion Receptor. Chem. Commun. 1998, 85–86. (30) Snowden, T. S.; Anslyn, E. V. Anion Recognition: Synthetic Receptors for Anions and Their Application in Sensors. Curr. Opin. Chem. Biol. 1999, 3 (6), 740–746. (31) Sessler, J. L.; Cho, D. G.; Lynch, V. Diindolylquinoxalines: Effective Indole-Based Receptors for Phosphate Anion. J. Am. Chem. Soc. 2006, 128 (51), 16518–16519. (32) Prakash, V.; Saha, S.; Chakraborty, K.; Krishnan, Y. Rational Design of a Quantitative, PH-Insensitive, Nucleic Acid Based Fluorescent Chloride Reporter. Chem. Sci. 2016, 7 (3), 1946–1953. (33) Chakraborty, K.; Veetil, A. T.; Jaffrey, S. R.; Krishnan, Y. Nucleic Acid-Based Nanodevices in Biological Imaging. Annu. Rev. Biochem. 2016, 85, 349–373. (34) Labuda, J.; Oliveira Brett, A. M.; Evtugyn, G.; Fojta, M.; Mascini, M.; Ozsoz, M.; Palchetti, I.; Paleček, E.; Wang, J. Electrochemical Nucleic Acid-Based Biosensors: Concepts, Terms, and Methodology (IUPAC Technical Report). Pure Appl. Chem. 2010, 82 (5), 1161–1187. (35) Jin, J.; Ouyang, X.; Li, J.; Jiang, J.; Wang, H.; Wang, Y.; Yang, R. Nucleic Acid- Modulated Silver Nanoparticles: A New Electrochemical Platform for Sensing Chloride Ion. Analyst 2011, 136 (18), 3629–3634. (36) Evans, N. H.; Beer, P. D. A Ferrocene Functionalized Rotaxane Host System Capable of the Electrochemical Recognition of Chloride. Org. Biomol. Chem. 2011, 9 (1), 92–100. 206 (37) Parks, F. C.; Sheetz, E. G.; Stutsman, S. R.; Lutolli, A.; Debnath, S.; Raghavachari, K.; Flood, A. H. Revealing the Hidden Costs of Organization in Host-Guest Chemistry Using Chloride-Binding Foldamers and Their Solvent Dependence. J. Am. Chem. Soc. 2022, 144 (3), 1274–1287. (38) Liu, Y.; Parks, F. C.; Sheetz, E. G.; Chen, C. H.; Flood, A. H. Polarity-Tolerant Chloride Binding in Foldamer Capsules by Programmed Solvent-Exclusion. J. Am. Chem. Soc. 2021, 143 (8), 3191–3204. (39) Hua, Y.; Liu, Y.; Chen, C. H.; Flood, A. H. Hydrophobic Collapse of Foldamer Capsules Drives Picomolar-Level Chloride Binding in Aqueous Acetonitrile Solutions. J. Am. Chem. Soc. 2013, 135 (38), 14401–14412. (40) John, E. A.; Massena, C. J.; Berryman, O. B. Helical Anion Foldamers in Solution. Chem. Rev. 2020, 120 (5), 2759–2782. (41) Suk, J.; Jeong, K.-S. Indolocarbazole-Based Foldamers Capable of Binding Halides in Water. J. Am. Chem. Soc. Commun. 2008, 130 (36), 11868–11869. (42) Hua, Y.; Flood, A. H. Flipping the Switch on Chloride Concentrations with a Light-Active Foldamer. J. Am. Chem. Soc. 2010, 132 (37), 12838–12840. (43) Gale, P. A.; Caltagirone, C. Anion Sensing by Small Molecules and Molecular Ensembles. Chem. Soc. Rev. 2015, 44 (13), 4212–4227. (44) Tutol, J. N.; Kam, H. C.; Dodani, S. C. Identification of MNeonGreen as a PH-Dependent, Turn-On Fluorescent Protein Sensor for Chloride. ChemBioChem 2019, 20 (14), 1759– 1765. (45) Zhang, P. F.; Yang, G. P.; Li, G. P.; Yang, F.; Liu, W. N.; Li, J. Y.; Wang, Y. Y. Series of Water-Stable Lanthanide Metal-Organic Frameworks Based on Carboxylic Acid Imidazolium Chloride: Tunable Luminescent Emission and Sensing. Inorg. Chem. 2019, 58 (20), 13969–13978. 207 (46) Miyaji, H.; Sessler, J. L. Off-the-Shelf Colorimetric Anion Sensors. Angew. Chemie - Int. Ed. 2001, 40 (1), 154–157. (47) Carroll, C. N.; Berryman, O. B.; Johnson, C. A.; Zakharov, L. N.; Haley, M. M.; Johnson, D. W. Protonation Activates Anion Binding and Alters Binding Selectivity in New Inherently Fluorescent 2,6-Bis(2-Anilinoethynyl)Pyridine Bisureas. Chem. Commun. 2009, No. 18, 2520–2522. (48) Carroll, C. N.; Coombs, B. A.; McClintock, S. P.; Johnson, C. A.; Berryman, O. B.; Johnson, D. W.; Haley, M. M. Anion-Dependent Fluorescence in Bis(Anilinoethynyl)Pyridine Derivatives: Switchable ON-OFF and OFF-ON Responses. Chem. Commun. 2011, 47 (19), 5539–5541. (49) Tresca, B. W.; Zakharov, L. N.; Carroll, C. N.; Johnson, D. W.; Haley, M. M. Aryl C- H⋯Cl-Hydrogen Bonding in a Fluorescent Anion Sensor. Chem. Commun. 2013, 49 (65), 7240–7242. (50) Gavette, J. V.; Mills, N. S.; Zakharov, L. N.; Johnson, C. A.; Johnson, D. W.; Haley, M. M. An Anion-Modulated Three-Way Supramolecular Switch That Selectively Binds Dihydrogen Phosphate, H2PO4-. Angew. Chemie - Int. Ed. 2013, 52 (39), 10270–10274. (51) Hartle, M. D.; Hansen, R. J.; Tresca, B. W.; Prakel, S. S.; Zakharov, L. N.; Haley, M. M.; Pluth, M. D.; Johnson, D. W. A Synthetic Supramolecular Receptor for the Hydrosulfide Anion. Angew. Chemie - Int. Ed. 2016, 55 (38), 11480–11484. (52) Eytel, L. M.; Gilbert, A. K.; Görner, P.; Zakharov, L. N.; Johnson, D. W.; Haley, M. M. Do CH–Anion and Anion–π Interactions Alter the Mechanism of 2:1 Host–Guest Complexation in Arylethynyl Monourea Anion Receptors? Chem. - A Eur. J. 2017, 23 (17), 4051–4054. (53) Würthner, F.; Kaiser, T. E.; Saha-Möller, C. R. J-Aggregates: From Serendipitous Discovery to Supramolecular Engineering of Functional Dye Materials. Angew. Chemie - Int. Ed. 2011, 50 (15), 3376–3410. 208 (54) Schmidtchen, F. P.; Berger, M. Artificial Organic Host Molecules for Anions. Chem. Rev. 1997, 97 (5), 1609–1646. (55) Palacios, M. A.; Nishiyabu, R.; Marquez, M.; Anzenbacher, P. Supramolecular Chemistry Approach to the Design of a High-Resolution Sensor Array for Multianion Detection in Water. J. Am. Chem. Soc. 2007, 129 (24), 7538–7544. (56) Robertson, C. C.; Perutz, R. N.; Brammer, L.; Hunter, C. A. A Solvent-Resistant Halogen Bond. Chem. Sci. 2014, 5 (11), 4179–4183. (57) Metrangolo, P.; Neukirch, H.; Pilati, T.; Resnati, G. Halogen Bonding Based Recognition Processes: A World Parallel to Hydrogen Bonding. Acc. Chem. Res. 2005, 38 (5), 386– 395. (58) Desiraju, G. R.; Ho, P. S.; Kloo, L.; Legon, A. C.; Marquardt, R.; Metrangolo, P.; Politzer, P.; Resnati, G.; Rissanen, K. Definition of the Halogen Bond ( IUPAC Recommendations 2013 ). Pure Appl. Chem. 2013, 85 (8), 1711–1713. (59) Clark, T.; Hennemann, M.; Murray, J. S.; Politzer, P. Halogen Bonding: The σ-Hole. J. Mol. Model. 2007, 13 (2), 291–296. (60) Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimagi, A.; Resnati, G.; Terraneo, G. The Halogen Bond. Chem. Rev. 2016, 116, 2478–2601. (61) Beale, T. M.; Chudzinski, M. G.; Sarwar, M. G.; Taylor, M. S. Halogen Bonding in Solution: Thermodynamics and Applications. Chem. Soc. Rev. 2013, 42 (4), 1667–1680. (62) Lohrman, J. A.; Deng, C.-L.; Shear, T. A.; Zakharov, L. N.; Haley, M. M.; Johnson, D. W. Methanesulfonyl-Polarized Halogen Bonding Enables Strong Halide Recognition in an Arylethynyl Anion Receptor. Chem. Commun. 2019, 55, 1919–1922. 209 Chapter 2: (1) L. Basabe-Desmonts, D. N. Reinhoudt and M. Crego-Calama, Chem. Soc. Rev., 2007, 36, 993. (2) L. D. Lavis and R. T. Raines, ACS Chem. Biol., 2008, 3, 142–155. (3) M. Vendrell, D. Zhai, J. C. Er and Y.-T. Chang, Chem. Rev., 2012, 112, 4391–4420. (4) J. Du, M. Hu, J. Fan and X. Peng, Chem. Soc. Rev., 2012, 41, 4511. (5) I. Tabakovic, K. Tabakovic, M. Trkovnik and Z. Stunic, Organic Preparations and Procedures International, 1985, 17, 133–137. (6) Y. Song, Z. Chen and H. Li, COC, 2012, 16, 2690–2707. (7) A. Stefanachi, F. Leonetti, L. Pisani, M. Catto and A. Carotti, Molecules, 2018, 23, 250. (8) R. S. Becker, S. Chakravorti, C. A. Gartner and M. de Graca Miguel, Faraday Trans., 1993, 89, 1007. (9) S. M. Sethna and N. M. Shah, Chem. Rev., 1945, 36, 1–62. (10) K. Tabaković and I. Tabaković, Journal of Heterocyclic Chemistry, 1992, 29, 383– 385. (11) T. M. Pereira, D. P. Franco, F. Vitorio and A. E. Kummerle, CTMC, 2018, 18, 124– 148. (12) M. Riveiro, N. De Kimpe, A. Moglioni, R. Vazquez, F. Monczor, C. Shayo and C. Davio, CMC, 2010, 17, 1325–1338. (13) D. Cao, Z. Liu, P. Verwilst, S. Koo, P. Jangjili, J. S. Kim and W. Lin, Chem. Rev., 2019, 119, 10403–10519. (14) J. L. Abernethy, J. Chem. Educ., 1969, 46, 561. 210 (15) R. D. H. Murray, J. Méndez and S. A. Brown, The natural coumarins: occurrence, chemistry, and biochemistry, Wiley, Chichester ; New York, 1982. (16) Anamika, D. Utreja, Ekta, N. Jain and S. Sharma, COC, 2019, 22, 2509–2536. (17) S. Govori, Synthetic Communications, 2016, 46, 569–580. (18) F. G. Medina, J. G. Marrero, M. Macías-Alonso, M. C. González, I. Córdova- Guerrero, A. G. Teissier García and S. Osegueda-Robles, Nat. Prod. Rep., 2015, 32, 1472–1507. (19) M. Tasior, D. Kim, S. Singha, M. Krzeszewski, K. H. Ahn and D. T. Gryko, J. Mater. Chem. C, 2015, 3, 1421–1446. (20) X. Liu, Z. Xu and J. M. Cole, J. Phys. Chem. C, 2013, 117, 16584–16595. (21) S. Nad, M. Kumbhakar and H. Pal, J. Phys. Chem. A, 2003, 107, 4808–4816. (22) G. Jones and M. A. Rahman, J. Phys. Chem., 1994, 98, 13028–13037. (23) L. Cisse, A. Djande, M. Capo-Chichi, F. Delatre, A. Saba, A. Tine and J.-J. Aaron, Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, 2011, 79, 428–436. (24) M. Cigáň, J. Donovalová, V. Szöcs, J. Gašpar, K. Jakusová and A. Gáplovský, J. Phys. Chem. A, 2013, 117, 4870–4883. (25) A. Gandioso, R. Bresolí-Obach, A. Nin-Hill, M. Bosch, M. Palau, A. Galindo, S. Contreras, A. Rovira, C. Rovira, S. Nonell and V. Marchán, J. Org. Chem., 2018, 83, 1185–1195. (26) Z.-S. Wang, Y. Cui, K. Hara, Y. Dan-oh, C. Kasada and A. Shinpo, Adv. Mater., 2007, 19, 1138–1141. (27) F. Bu, R. Duan, Y. Xie, Y. Yi, Q. Peng, R. Hu, A. Qin, Z. Zhao and B. Z. Tang, Angew. Chem. Int. Ed., 2015, 54, 14492–14497. 211 (28) J.-X. Chen, W. Liu, C.-J. Zheng, K. Wang, K. Liang, Y.-Z. Shi, X.-M. Ou and X.-H. Zhang, ACS Appl. Mater. Interfaces, 2017, 9, 8848–8854. (29) K. Hara, T. Sato, R. Katoh, A. Furube, Y. Ohga, A. Shinpo, S. Suga, K. Sayama, H. Sugihara and H. Arakawa, J. Phys. Chem. B, 2003, 107, 597–606. (30) S. A. Swanson, G. M. Wallraff, J. P. Chen, W. Zhang, L. D. Bozano, K. R. Carter, J. R. Salem, R. Villa and J. C. Scott, Chem. Mater., 2003, 15, 2305–2312. (31) Y. Jung, J. Jung, Y. Huh and D. Kim, Journal of Analytical Methods in Chemistry, 2018, 2018, 1–11. (32) A. B. Ahvale, H. Prokopcová, J. Šefčovičová, W. Steinschifter, A. E. Täubl, G. Uray and W. Stadlbauer, Eur. J. Org. Chem., 2008, 2008, 563–571. (33) M. B. de Macedo, R. Kimmel, D. Urankar, M. Gazvoda, A. Peixoto, F. Cools, E. Torfs, L. Verschaeve, E. S. Lima, A. Lyčka, D. Milićević, A. Klásek, P. Cos, S. Kafka, J. Košmrlj and D. Cappoen, European Journal of Medicinal Chemistry, 2017, 138, 491–500. (34) G. C. Enoua, G. Lahm, G. Uray and W. Stadlbauer, J. Heterocyclic Chem., 2014, 51, E263–E275. (35) G. A. Strohmeier, W. M. F. Fabian and G. Uray, HCA, 2004, 87, 215–226. (36) G. Uray, K. S. Niederreiter, F. Belaj and W. M. F. Fabian, HCA, 1999, 82, 1408– 1417. (37) W. M. F. Fabian, K. S. Niederreiter, G. Uray and W. Stadlbauer, Journal of Molecular Structure, 1999, 477, 209–220. (38) T. Tashima, Bioorganic & Medicinal Chemistry Letters, 2015, 25, 3415–3419. (39) J. R. Casley-Smith, R. G. Morgan and N. B. Piller, N Engl J Med, 1993, 329, 1158– 1163. 212 (40) J. P. Bard, C.-L. Deng, H. C. Richardson, J. M. Odulio, J. E. Barker, L. N. Zakharov, P. H.-Y. Cheong, D. W. Johnson and M. M. Haley, Org. Chem. Front., 2019, 6, 1257–1265. (41) C. L. Vonnegut, A. M. Shonkwiler, M. M. Khalifa, L. N. Zakharov, D. W. Johnson and M. M. Haley, Angew. Chem. Int. Ed., 2015, 54, 13318–13322. (42) M. J. S. Dewar and V. P. Kubba, J. Am. Chem. Soc., 1960, 82, 5685–5688. (43) I. G. M. Campbell and J. K. Way, J. Chem. Soc., 1961, 2133. (44) Y. Sun and N. Cramer, Angew. Chem. Int. Ed., 2017, 56, 364–367. (45) W. Tang and Y.-X. Ding, J. Org. Chem., 2006, 71, 8489–8492. (46) J. Yan, Q. Li, J. A. Boutin, M. P. Renard, Y. Ding, X. Hao, W. Zhao and M. Wang, Acta Pharmacologica Sinica, 2008, 29, 752–758. (47) D. Zhao, C. Nimphius, M. Lindale and F. Glorius, Org. Lett., 2013, 15, 4504–4507. (48) S. Park, B. Seo, S. Shin, J.-Y. Son and P. H. Lee, Chem. Commun., 2013, 49, 8671. (49) L. Liu, A.-A. Zhang, Y. Wang, F. Zhang, Z. Zuo, W.-X. Zhao, C.-L. Feng and W. Ma, Org. Lett., 2015, 17, 2046–2049. (50) Z.-Q. Lin, W.-Z. Wang, S.-B. Yan and W.-L. Duan, Angew. Chem. Int. Ed., 2015, 54, 6265–6269. (51) Z. Wang, B. S. Gelfand and T. Baumgartner, Angew. Chem. Int. Ed., 2016, 55, 3481– 3485. (52) Y.-N. Ma, X. Zhang and S. Yang, Chem. Eur. J., 2017, 23, 3007–3011. (53) N. A. Takaesu, E. Ohta, L. N. Zakharov, D. W. Johnson and M. M. Haley, Organometallics, 2017, 36, 2491–2493. 213 (54) C. Deng, J. P. Bard, J. A. Lohrman, J. E. Barker, L. N. Zakharov, D. W. Johnson and M. M. Haley, Angew. Chem. Int. Ed., 2019, 58, 3934–3938. (55) C.-L. Deng, J. P. Bard, L. N. Zakharov, D. W. Johnson and M. M. Haley, J. Org. Chem., 2019, 84, 8131–8139. (56) C.-L. Deng, J. P. Bard, L. N. Zakharov, D. W. Johnson and M. M. Haley, Org. Lett., 2019, 21, 6427–6431. (57) T. A. Schaub, S. M. Brülls, P. O. Dral, F. Hampel, H. Maid and M. Kivala, Chem. Eur. J., 2017, 23, 6988–6992. (58) C. Romero‐Nieto, A. López‐Andarias, C. Egler‐Lucas, F. Gebert, J. Neus and O. Pilgram, Angew. Chem. Int. Ed., 2015, 54, 15872–15875. (59) A. Fukazawa, H. Osaki and S. Yamaguchi, Asian Journal of Organic Chemistry, 2014, 3, 122–127. (60) E. Regulska and C. Romero-Nieto, Dalton Trans., 2018, 47, 10344–10359. (61) N. Grenon and T. Baumgartner, Inorg. Chem., 2018, 57, 1630–1644. (62) P. Gong, K. Ye, J. Sun, P. Chen, P. Xue, H. Yang and R. Lu, RSC Adv., 2015, 5, 94990–94996. (63) X.-D. Jiang, J. Zhao, D. Xi, H. Yu, J. Guan, S. Li, C.-L. Sun and L.-J. Xiao, Chem. Eur. J., 2015, 21, 6079–6082. (64) F. Mathey, Chem. Rev., 1988, 88, 429–453. (65) N. Avarvari, P. Le Floch and F. Mathey, J. Am. Chem. Soc., 1996, 118, 11978– 11979. (66) W.-C. Sun, K. R. Gee and R. P. Haugland, Bioorganic & Medicinal Chemistry Letters, 1998, 8, 3107–3110. 214 (67) J. R. Quinn, S. C. Zimmerman, J. E. Del Bene and I. Shavitt, J. Am. Chem. Soc., 2007, 129, 934–941. (68) A. M. Brouwer, Pure and Applied Chemistry, 2011, 83, 2213–2228. (69) K. Kuroda, Y. Maruyama, Y. Hayashi and T. Mukaiyama, BCSJ, 2009, 82, 381–392. Chapter 3: (1) C. L. Vonnegut, B. W. Tresca, D. W. Johnson and M. M. Haley, Chemistry - An Asian Journal, 2015, 10, 522–535. (2) M. D. Hartle, R. J. Hansen, B. W. Tresca, S. S. Prakel, L. N. Zakharov, M. M. Haley, M. D. Pluth and D. W. Johnson, Angew. Chem. Int. Ed., 2016, 55, 11480–11484. (3) J. A. Lohrman, C.-L. Deng, T. A. Shear, L. N. Zakharov, M. M. Haley and D. W. Johnson, Chem. Commun., 2019, 55, 1919–1922. (4) H. A. Fargher, N. Lau, L. N. Zakharov, M. M. Haley, D. W. Johnson and M. D. Pluth, Chemical Science, 2019, 10, 67–72. (5) N. Lau, L. N. Zakharov and M. D. Pluth, Chem. Commun., 2018, 54, 2337–2340. (6) L. M. Eytel, H. A. Fargher, M. M. Haley and D. W. Johnson, Chem. Commun., 2019, 55, 5195–5206. (7) O. B. Berryman, C. A. Johnson, C. L. Vonnegut, K. A. Fajardo, L. N. Zakharov, D. W. Johnson and M. M. Haley, Crystal Growth & Design, 2015, 15, 1502–1511. (8) P. D. Beer and P. A. Gale, Angew. Chem. Int. Ed., 2001, 31. (9) M. J. Langton, C. J. Serpell and P. D. Beer, Angew. Chem. Int. Ed., 2016, 55, 1974– 1987. 215 (10) S. Kubik, Chem. Soc. Rev., 2010, 39, 3648. (11) L. M. Eytel, A. C. Brueckner, J. A. Lohrman, M. M. Haley, P. H.-Y. Cheong and D. W. Johnson, Chem. Commun., 2018, 54, 13208–13211. (12) H. A. Fargher, N. Lau, H. C. Richardson, P. H.-Y. Cheong, M. M. Haley, M. D. Pluth and D. W. Johnson, J. Am. Chem. Soc., 2020, 142, 8243–8251. (13) A. Borissov, I. Marques, J. Y. C. Lim, V. Félix, M. D. Smith and P. D. Beer, J. Am. Chem. Soc., 2019, 141, 4119–4129. (14) K. Kavallieratos, C. M. Bertao and R. H. Crabtree, J. Org. Chem., 1999, 64, 1675– 1683. (15) Corwin. Hansch, A. Leo and R. W. Taft, Chem. Rev., 1991, 91, 165–195. (16) L. Xu, Z. Liu, W. Dong, J. Song, M. Miao, J. Xu and H. Ren, Org. Biomol. Chem., 2015, 13, 6333–6337. (17) Y. Wang, Y. Zhou, X. Ma and Q. Song, Org. Lett., 2021, 23, 5599–5604. (18) M. Jash, B. Das and C. Chowdhury, J. Org. Chem., 2016, 81, 10987–10999. (19) M. M. Watt, J. M. Engle, K. C. Fairley, T. E. Robitshek, M. M. Haley and D. W. Johnson, Org. Biomol. Chem., 2015, 13, 4266–4270. (20) Y. Liu, A. Sengupta, K. Raghavachari and A. H. Flood, Chem, 2017, 3, 411–427. (21) P. Thordarson, Chem. Soc. Rev., 2011, 40, 1305–1323. (22) D. Brynn Hibbert and P. Thordarson, Chem. Commun., 2016, 52, 12792–12805. (23) A. M. S. Riel, D. A. Decato, J. Sun, C. J. Massena, M. J. Jessop and O. B. Berryman, Chem. Sci., 2018, 9, 5828–5836. (24) C. A. Lipinski, Drug Discovery Today: Technologies, 2004, 1, 337–341. 216 (25) I. J. Bazany-Rodríguez, D. Martínez-Otero, J. Barroso-Flores, A. K. Yatsimirsky and A. Dorazco-González, Sensors and Actuators B: Chemical, 2015, 221, 1348–1355. Chapter 4: (1) K. Berend, L. H. van Hulsteijn and R. O. B. Gans, Eur. J. Intern. Med., 2012, 23, 203. (2) C. D. Meletis, J Evid. Based Complementary Altern. Med., 2011, 16, 190. (3) K. J. Bowman, Science, 2019, 365, 124. (4) J. Pancholi and P. D. Beer, Coord. Chem. Rev., 2020, 213281. (5) R. Hein and P. D. Beer, Chem. Sci., 2022, 13, 7098. (6) W. Kunz, J. Henle and B. W. Ninham, Curr. Opin. Colloid Interface Sci., 2004, 9, 19. (7) K. P. Gregory, G. R. Elliott, H. Robertson, A. Kumar, E. J. Wanless, G. B. Webber, V. S. J. Craig, G. G. Andersson and A. J. Page, Phys. Chem. Chem. Phys., 2022, 24, 12682. (8) Y. Liu, W. Zhao, C. H. Chen and A. H. Flood, Science, 2019, 365, 159. (9) Craig. C. Robertson, J. S. Wright, E. J. Carrington, R. N. Perutz, C. A. Hunter and L. Brammer, Chem. Sci., 2017, 8, 5392. (10) B. W. Tresca, L. N. Zakharov, C. N. Carroll, D. W. Johnson and M. M. Haley, Chem. Commun., 2013, 49, 7240. (11) M. M. Watt, J. M. Engle, K. C. Fairley, T. E. Robitshek, M. M. Haley and D. W. Johnson, Org. Biomol. Chem., 2015, 13, 4266. (12) C. N. Carroll, B. A. Coombs, S. P. McClintock, C. A. Johnson II, O. B. Berryman, D. W. Johnson and M. M. Haley, Chem. Commun., 2011, 47, 5539. 217 (13) C. A. Lipinski, F. Lombardo, B. W. Dominy and P. J. Feeney, Adv. Drug Deliv. Rev., 1997, 23, 3. (14) C. A. Lipinski, Drug Discov. Today Technol., 2004, 1, 337. (15) E. Corradi, S. V. Meille, M. T. Messina, P. Metrangolo and G. Resnati, Angew. Chem. Int. Ed., 2000, 39, 1782. (16) H. A. Fargher, N. Lau, L. N. Zakharov, M. M. Haley, D. W. Johnson and M. D. Pluth, Chem. Sci., 2019, 10, 67. (17) L. M. Lee, M. Tsemperouli, A. I. Poblador-Bahamonde, S. Benz, N. Sakai, K. Sugihara and S. Matile, J. Am. Chem. Soc., 2019, 141, 810. (18) M. S. Taylor, Coord. Chem. Rev. 2020, 413, 213270. (19) A. Borissov, I. Marques, J. Y. C. Lim, V. Félix, M. D. Smith and P. D. Beer, J. Am. Chem. Soc., 2019, 141, 4119. (20) C. C. Robertson, R. N. Perutz, L. Brammer and C. A. Hunter, Chem. Sci., 2014, 5, 4179. (21) T. Bunchuay, A. Docker, A. J. Martinez‐Martinez and P. D. Beer, Angew. Chem. Int. Ed., 2019, 58, 13823. (22) J. A. Lohrman, C.-L. Deng, T. A. Shear, L. N. Zakharov, M. M. Haley and D. W. Johnson, Chem. Commun., 2019, 55, 1919. (23) D. A. Decato, A. M. S. Riel and O. B. Berryman, Crystals, 2019, 9, 522. (24) A. M. S. Riel, D. A. Decato, J. Sun, C. J. Massena, M. J. Jessop and O. B. Berryman, Chem. Sci., 2018, 9, 5828. (25) M. Giese, M. Albrecht, T. Krappitz, M. Peters, V. Gossen, G. Raabe, A. Valkonen and K. Rissanen, Chem. Commun., 2012, 48, 9983. (26) M. Giese, M. Albrecht and K. Rissanen, Chem. Commun., 2016, 52, 1778. 218 (27) G. R. Desiraju, P. S. Ho, L. Kloo, A. C. Legon, R. Marquardt, P. Metrangolo, P. Politzer, G. Resnati and K. Rissanen, Pure Appl. Chem., 2013, 85, 1711. (28) D. Brynn Hibbert and P. Thordarson, Chem. Commun., 2016, 52, 12792. (29) P. Thordarson, Chem. Soc. Rev., 2011, 40, 1305. (30) S. W. Robinson, C. L. Mustoe, N. G. White, A. Brown, A. L. Thompson, P. Kennepohl and P. D. Beer, J. Am. Chem. Soc., 2015, 137, 499. (31) H. A. Fargher, N. Lau, H. C. Richardson, P. H.-Y. Cheong, M. M. Haley, M. D. Pluth and D. W. Johnson, J. Am. Chem. Soc., 2020, 142, 8243. Chapter 5: (1) A. M. S. Riel, R. K. Rowe, E. N. Ho, A.-C. C. Carlsson, A. K. Rappé, O. B. Berryman and P. S. Ho, Acc. Chem. Res., 2019, 52, 2870–2880. (2) O. B. Berryman, C. A. Johnson, L. N. Zakharov, M. M. Haley and D. W. Johnson, Angew. Chem. Int. Ed., 2008, 47, 117–120. (3) B. W. Tresca, L. N. Zakharov, C. N. Carroll, D. W. Johnson and M. M. Haley, Chem. Commun., 2013, 49, 7240. (4) C. L. Vonnegut, B. W. Tresca, D. W. Johnson and M. M. Haley, Chemistry - An Asian Journal, 2015, 10, 522–535. (5) L. M. Eytel, A. C. Brueckner, J. A. Lohrman, M. M. Haley, P. H.-Y. Cheong and D. W. Johnson, Chem. Commun., 2018, 54, 13208–13211. (6) J. A. Lohrman, C.-L. Deng, T. A. Shear, L. N. Zakharov, M. M. Haley and D. W. Johnson, Chem. Commun., 2019, 55, 1919–1922. (7) H. A. Fargher, N. Lau, L. N. Zakharov, M. M. Haley, D. W. Johnson and M. D. Pluth, Chemical Science, 2019, 10, 67–72. 219 (8) H. A. Fargher, N. Lau, H. C. Richardson, P. H.-Y. Cheong, M. M. Haley, M. D. Pluth and D. W. Johnson, J. Am. Chem. Soc., 2020, 142, 8243–8251. (9) G. Turner, A. Docker and P. D. Beer, Dalton Trans., 2021, 50, 12800–12805. (10) R. Tepper, B. Schulze, H. Görls, P. Bellstedt, M. Jäger and U. S. Schubert, Org. Lett., 2015, 17, 5740–5743. (11) N. L. Kilah, M. D. Wise, C. J. Serpell, A. L. Thompson, N. G. White, K. E. Christensen and P. D. Beer, J. Am. Chem. Soc., 2010, 132, 11893–11895. (12) L. Fotović, N. Bedeković and V. Stilinović, Crystal Growth & Design, 2021, 21, 6889–6901. (13) P. Thordarson, Chem. Soc. Rev., 2011, 40, 1305–1323. (14) D. Brynn Hibbert and P. Thordarson, Chem. Commun., 2016, 52, 12792–12805. (15) R. D. Shannon, Acta Cryst A, 1976, 32, 751–767. (16) Y. Jung, J. Jung, Y. Huh and D. Kim, Journal of Analytical Methods in Chemistry, 2018, 2018, 1–11. (17) J. Du, M. Hu, J. Fan and X. Peng, Chem. Soc. Rev., 2012, 41, 4511. (18) F. Zapata, A. Caballero, N. G. White, T. D. W. Claridge, P. J. Costa, V. Félix and P. D. Beer, J. Am. Chem. Soc., 2012, 134, 11533–11541. (19) D. Mungalpara, S. Stegmüller and S. Kubik, Chem. Commun., 2017, 53, 5095–5098. (20) J. Shang, W. Si, W. Zhao, Y. Che, J.-L. Hou and H. Jiang, Org. Lett., 2014, 16, 4008– 4011. (21) A. Borissov, I. Marques, J. Y. C. Lim, V. Félix, M. D. Smith and P. D. Beer, J. Am. Chem. Soc., 2019, 141, 4119–4129. 220 (22) F. C. Parks, E. G. Sheetz, S. R. Stutsman, A. Lutolli, S. Debnath, K. Raghavachari and A. H. Flood, J. Am. Chem. Soc., 2022, 144, 1274–1287. (23) J. M. Fourquez, A. Godard, F. Marsais and G. Quéguiner, Journal of Heterocyclic Chemistry, 1995, 32, 1165–1170. 221