<<

University of Mary Washington Eagle Scholar

Student Research Submissions

Spring 5-7-2018

Synthetic Routes to Bromo-Terminated Phosphonate Films and Alkynyl Compounds for Click Coupling

Poornima Sunder

Follow this and additional works at: https://scholar.umw.edu/student_research

Part of the Biochemistry Commons

Recommended Citation Sunder, Poornima, "Synthetic Routes to Bromo-Terminated Phosphonate Films and Alkynyl Pyridine Compounds for Click Coupling" (2018). Student Research Submissions. 226. https://scholar.umw.edu/student_research/226

This Honors Project is brought to you for free and open access by Eagle Scholar. It has been accepted for inclusion in Student Research Submissions by an authorized administrator of Eagle Scholar. For more information, please contact [email protected].

Synthetic Routes to Bromo-Terminated Phosphonate Films and Alkynyl Pyridine Compounds

for Click Coupling

Poornima Rachel Sunder

Thesis submitted to the faculty of University of Mary Washington in partial fulfillment of the requirements for graduation with Honors in Chemistry (2018)

ABSTRACT

Click reactions are a highly versatile class of reactions that produce a diverse range of products. Copper-catalyzed azide- cycloaddition (CuAAC) click reactions require an azide and a terminal alkyne and produce a coupled product that is “clicked” through a triazole ring that can have a variety of substituents. In this work, bromo-terminated phosphonate films on copper oxide surfaces were explored as the platform for click coupling, as the terminal azide needed for the reaction can be generated through an in situ SN2 reaction with a terminal bromo group. The reactions were characterized using model reactions in solution before being conducted on modified copper oxide surfaces. Copper oxide surfaces were modified with 11- hydroxyundecylphosphonic acid (1) through the tethering by aggregation and growth (TBAG) method. Since modification of these surfaces with 1 installs a hydroxyl-terminated film (2), synthetic routes to bromination of the species using tribromide, tosyl chloride and bromide, 2,4,6-trichloro-(1,3,5)-triazine, and were explored. with halo-pyridine and halo-terpyridine compounds was conducted to synthesize a desired alkynyl species. CuAAC was conducted using the synthesized bromo-terminated phosphonate film and a model terminal alkyne, and Surface Reflectance

Infrared analysis was used to characterize the films. Future work will focus on the use of the

CuAAC to couple a synthesized pyridyl alkyne species with the bromo-terminated films as a route towards catalytic applications.

i ACKNOWLEDGEMENTS

I would like to thank everyone in the University of Mary Washington’s Chemistry department for constantly inspiring me through their knowledge, creativity, and curiosity. I am thankful that I had the chance to look up to the faculty as my role models and am grateful that I had the chance to build strong relationships with many of them. I would also like to thank the students in the Chemistry department, and especially the senior class of 2018, that walked along side me in the journey through all the organic chemistry mechanisms, physical chemistry derivations, and biochemistry pathways. I am also thankful for my family and friends that never ceased to encourage me and push me beyond the milestones that, at times, seemed unreachable. I especially want to thank Ave Keefer for being such a kind and caring research partner and friend, and Katelyn Torres for providing much needed encouragement and company countless times. I am also grateful to Brook Andrews, who synthesized the phosphonic acid that I used as part of my research. A very special thanks goes to my research advisor, Dr. Nicole Crowder, not only for her endless amount of patience she bestowed upon me but also for the personal encouragement and empowerment that she provided on a regular basis.

I am grateful to have received generous financial support for the research conducted, and specifically financial support provided by the UMW Summer Science Institute and the UMW

Undergraduate Research Grant. This monetary support made it possible to purchase chemicals and equipment for research and to present my research in professional academic settings.

Thank you to the entire UMW community that adopted me as a transfer student and made me feel so incredibly welcomed and loved. My time here has been truly wonderful because of the strength of this community.

ii TABLE OF CONTENTS

Abstract…………………………………………………………………..…………...... ………....i

Acknowledgements……………………………………………………………..……...... ii

List of Figures………………………………………………………………...……....…...... iv

Introduction……………………………………………………..…………………...... ……….…1

Experimental…………………………………………………………………………...... 18

Results and Discussion………………………………………………………..…..……………..27

Conclusions and Future Work ……………..……………………………………………………52

References……………..………………………………………………………….……...... 53

Appendix...... 57

iii LIST OF FIGURES

Figure 1: General copper-catalyzed azide-alkyne cycloaddition coupling reaction

Figure 2: Mechanism for copper-catalyzed cycloaddition of azides and terminal , with [Cu] denoting the copper complex catalyzing the reaction

Figure 3: Copper-catalyzed azide-alkyne cycloaddition (CuAAC) onto copper surface, with bromo-terminated alkylphosphonates and phenylacetylene

Figure 4: Carboxylates and phosphonates can be covalently bound onto the surface of metal oxide

Figure 5: Tethering by Aggregation and Growth reaction scheme (a) and depiction of method (b)

Figure 6: Mechanism of chemisorption and covalent bond formation between phosphonic acid and metal oxide surface

Figure 7: TBAG deposition of 1 onto a copper oxide surface

Figure 8: Possible Synthetic Routes on terminal hydroxyl of 2 to yield 3

Figure 9: Bromination routes in solution using 4 as the starting material

Figure 10: Mechanism of alcohol bromination using phosphorus tribromide

Figure 11: Mechanism for tosylation of an alcohol using tosyl chloride

Figure 12: Bromination of a tosylate species through SN2 substitution using

Figure 13: Mechanism of bromination of an alcohol using carbon tetrabromide and triphenylphosphine

Figure 14: A proposed mechanism for bromination of an alcohol using trichlorotriazine and sodium bromide

Figure 15: Mechanism for Sonogashira palladium-catalyzed cross coupling of pyridyl halides with (trialkylsilyl)alkynes

Figure 16: Sonogashira coupling and desilylation reactions yield alkynyl

iv

Figure 17: CuAAC reactions of surface-bound bromo-terminated alkylphosphonates and synthesized alkynyl pyridines

Figure 18: Copper-catalyzed azide-alkyne cycloaddition reactions with 5 or 3 and phenylacetylene

Figure 19: Ligation of -based complexes to a copper surface with surface-bound pyridyl species as

Figure 20: Surface Reflectance IR (SRIR) of 2

Figure 21: NMR spectrum of commercially obtained 4

Figure 22: NMR spectrum of commercially obtained 5

Figure 23: IR spectrum of commercially obtained 5

Figure 24: NMR spectrum of crude 5 synthesized from 4 using PBr3

Figure 25: SRIR spectrum of 2 after PBr3 reaction

Figure 26: NMR spectrum of 4 after TsCl reaction

Figure 27: NMR spectrum of crude 5 after NaBr reaction

Figure 28: SRIR spectrum of 2 after reaction with TsCl

Figure 29: SRIR spectrum of 2 after NaBr reaction

Figure 30: NMR spectrum of 4 after TCT reaction

Figure 31: SRIR spectrum of 2 after TCT reaction

Figure 32: NMR spectrum of 5 after reaction with CBr4 /PPh3 and Kugelrohr distillation

Figure 33: SRIR spectrum of 2 after reaction with CBr4 /PPh3

Figure 34: NMR spectrum of 6

v Figure 35: NMR spectrum of 6 after Sonogashira reaction

Figure 36: NMR spectrum of first fraction of the Sonogashira product after flash chromatography

Figure 37: NMR spectrum of second fraction of the Sonogashira product after flash chromatography

Figure 38: Structures for possible side products of the Sonogashira reaction due to homocoupling of TMS-acetylene (top) and TIPS- acetylene (bottom)

Figure 39: NMR spectrum of 9 after Sonogashira reaction

Figure 40: NMR spectrum of 9 after Songashira reaction and desilylation, taken in CD2Cl2

Figure 41: IR spectrum of 5 after CuAAC reaction, dissolved in acetone

Figure 42: SRIR spectrum of 3 (from TCT synthesis) after CuAAC reaction

Figure 43: SRIR spectrum of 3 (from CBr4 synthesis) after CuAAC reaction

vi INTRODUCTION

Click Coupling: Copper-Catalyzed Azide-Alkyne Cycloaddition

Click chemistry has emerged as a key category of reactions that can easily build complex molecules from more simple ones, as they “click” together into highly stable coupled products.

These products contain functional groups such as triazole rings that are formed through highly irreversible reactions. These reactions are additionally advantageous in that they generally give high product yields and create byproducts that can be removed without follow-up methods involving purification by chromatography.1 Due to these properties, click reactions are attractive and have potential for creating a wide range of coupled products.

Copper (I)-catalyzed azide-alkyne cycloadditions (CuAAC) are an important class of click reactions in which compounds containing a terminal azide can be reacted with compounds with a terminal alkyne. The resulting coupled product is a five-membered triazole ring, as seen in

Figure 1.1-2 CuAAC holds an additional advantage in that the reaction can take place over broad temperature and pH ranges and is compatible with a diverse range of functional groups.3

N N CuSO4 R N N N + R 1 2 N NaC6H7O6 R R1 2

Figure 1: General copper-catalyzed azide-alkyne cycloaddition coupling reaction

CuAAC is a regioselective reaction and yields the 1,4-disubstituted 1,2,3-triazole.4 Figure

2 illustrates a proposed mechanism for the formation of the triazole ring through coupling of an azide and alkyne.4 In this mechanism, the copper (I) catalytic complex is generated when sodium ascorbate reduces copper (II) sulfate. The copper (I) complex coordinates the terminal alkyne and forms a copper acetylide species when the acetylenic proton is abstracted by a base, most

1 likely by a water or tert-butanol molecule in the reaction mixture. The azide is then coordinated with the copper complex, and the π electrons of the alkyne perform a nucleophilic attack on the azide, forming a six-membered metallacycle ring intermediate, which is stabilized by another copper complex. The nitrogen coordinated by both copper complexes then attacks the alpha carbon which creates the five-membered triazole ring coordinated by one of the copper complexes and simultaneously collapses the unstable metallacycle intermediate.4 Protonolysis results in the release of the CuAAC coupled product containing the triazole, while the copper catalyst is released to begin another catalytic cycle.5

N B O R1 N N O P R1 = O 10 R H + 2 [Cu]+ [Cu] B

R2 H R2 H B-H [Cu]+

R2 = B-H N R1 N N [Cu]+

R2 Cu R2 [Cu] N N N R [Cu]+ 1

R1 N R1 N N N [Cu] N N [Cu]+ R2 [Cu] Cu R2

Figure 2: Mechanism for copper-catalyzed cycloaddition of azides and terminal alkynes, with [Cu] denoting the copper complex catalyzing the reaction4,5

2 The terminal azide that is necessary to react in the click reaction can be directly synthesized, or it can be generated through an in situ SN2 reaction of an azide with a terminal bromo-group.1 The latter is known as a one-pot synthesis, because it simplifies the synthetic process by removing the need to isolate an azide-terminated compound prior to conducting a click reaction.1 This research project will focus on the synthesis of a bromo-terminated phosphonate film on a copper oxide surface (R1) and alkyne-terminated pyridine species (R2) to act as reactants in the in situ CuAAC reactions.

The reaction scheme in Figure 3 shows a route to triazole formation through CuAAC, with phenylacetylene acting as a model alkyne. The distinct bands due to the aromatic allow for ease of characterization of the coupled product by (IR).

O O NaN3, CuSO , 4 O N O NaC6H7O6 N P P Br 10 N O 10 O

Figure 3: Copper-catalyzed azide-alkyne cycloaddition (CuAAC) onto copper surface, with bromo-terminated alkylphosphonates and phenylacetylene

Surface Modification: Phosphonic Acid Attachment through Tethering by Aggregation and Growth (TBAG) Method

This research utilizes copper oxide surfaces as a platform for the bromo-functionalized species that can participate in a click reaction. and phosphonic acid species have been observed to self-assemble onto metal oxide surfaces as a monolayer, and these monolayers act as the anchor for further surface-based modifications.6-13 Several studies have suggested that phosphonic acids have the potential to bind to the metal oxide surface in a tridentate manner and

3 that coordination of the phosphonate to the surface may often be in this form.14-15 This tridentate bond formation is possible because there is an additional acid moiety on a phosphonic acid when compared with a carboxylic acid. Figure 4 demonstrates carboxylic acid and phosphonic acid binding on metal oxide surfaces.8

R 1. O R O

O O O R OH

2. Bake at 140 °C

OH OH O O

Native Oxide Layer

R R O P P

O O O O O

1. O OH P R OH

2. Bake at 140 °C

Figure 4: Carboxylates and phosphonates can be covalently bound onto the surface of metal oxides

Deposition of acid species onto metal surfaces is procedurally accomplished through a method known as tethering by aggregation and growth (TBAG). The TBAG method is a technique in which the molecule that is to be adhered onto the metal oxide surface is dissolved in a dilute solution. The metal oxide surfaces are then suspended in the solution in order for the molecules to chemisorb onto the surface as the solvent evaporates, as seen in Figure 5b.10,13,16 A self-assembled monolayer is formed when the phosphonic acids organize themselves in order to

4 orient the polar head groups towards the copper surface, with hydrogen bonding interactions occurring between the head group and the native oxide layer of the surface.

a) b) Phosphonic acids organized at the air-solvent interface sample holder O O 1. TBAG deposition 2. 130 °C, 48 hours HO O P P R slow solvent phosphonic 10 R O 10 evaporation acid film HO substrate dissolved phosphonic acid

Figure 5: Tethering by Aggregation and Growth reaction scheme (a) and depiction of method (b)

Figure 6 shows the phosphonic acid functional group as it interacts with the hydroxyl and

-oxo species that are present on the native oxide layer of the copper surface.15,17A series of hydrogen bonding interactions facilitate the interfacial proton transfers from the phosphonic acid to the metal surface. The -oxo groups are often a predominating species on copper oxide surfaces15,17 and can have hydrogen bonding interactions with the phosphonic acid species. This facilitates the opening of these groups through a second proton transfer to the surface, with the end result being a greater degree of surface modification by the phosphonic acid species and a more stable film.13,15

After chemisorption of the phosphonic acid monolayer onto the surface, baking the surfaces causes dehydration and formation of covalent bonds between the phosphonic acid head group and the metal oxide surface. The water generated in this process is removed from the surface through exposure to heat, and the acid becomes covalently bound to the metal surface as the phosphonate.10

5 R R R R R R

P O P O O P HO OH O P HO O O P P O µ-oxo OH HO OH HO OH HO OH HO O species OH OH2 O O O O M Phosphonic Acid Proton Transfer Heat, O Deposition O to Surface O loss of H2O O

R R R R R R

O P O P P O O P P O HO OH P O HO HO OH O O O O HO OH O H OH O 1. Opening of O M M µ-oxo groups M M Coordination Proton Transfer to Surface O O 2. Coordination O

Figure 6: Mechanism of chemisorption and covalent bond formation between phosphonic acid and metal oxide surface

Surface Modification: Synthesis

11-Hydroxyundecylphosphonic acid (1) was used as the phosphonic acid species to form the surface monolayer through TBAG deposition. This research will use previously synthesized

1 to form surface-bound 11-hydroxyundecylphosphonate (2), as seen in Figure 7. The terminal functional group on the chain that is not directly bound to the metal is free to participate in further reactions on the surface.

O O 1. TBAG deposition 2. 130 °C, 48 hours HO O P P OH 10 OH HO 10 O

1 2

Figure 7: TBAG deposition of 1 onto a copper oxide surface

6 After the phosphonate film is confirmed by surface reflectance infrared spectroscopy

(SRIR) to be covalently attached to the surface, further reactions can be conducted on the terminal hydroxyl group of the phosphonate in order to convert it to a bromo-terminated phosphonate film.

Synthetic Routes for Bromination

In this research, several reactions will be explored as synthetic routes to achieve conversion of the terminal hydroxyl on the phosphonate film to the bromo- functional group necessary to participate in the click reaction (Figure 8). One route utilizes phosphorus tribromide

18 (PBr3) as a bromination reagent through activation of the alcohol, followed by an SN2 reaction.

Another route allows for conversion of the alcohol group to a tosylate, which functions as a

19 leaving group when the bromide of NaBr performs a nucleophilic attack in this SN2 reaction.

Another synthetic route involves the conversion of a hydroxyl group to a bromo group through an Appel halogenation that uses carbon tetrabromide and triphenyl .20,21 A final route uses sodium bromide in the presence of a trichlorotriazine promoter for bromination.22

7 PBr3 , CH2Cl2

1) TsCl, pyr

2) NaBr, DMF O O O O P P OH O 10 O 10 Br 3

CBr4, PPh3

CH2Cl2

TCT, DMF NaBr

Figure 8: Possible Synthetic Routes on terminal hydroxyl of 2 to yield 3

Confirmation of reactions on surfaces can be difficult due to reliance on SRIR as the only characterization technique. In addition, the presence of a monolayer and further reactions from that monolayer may be difficult to detect due to possibly attenuated signals. Therefore, each bromination route proposed for the modified copper surfaces in Figure 8 was modeled using 1- heptanol (4) as the alcohol for solution-based reactions, as shown in Figure 9. This molecule was chosen because it contains a primary alcohol and a seven-carbon alkyl chain and is a good representative of the reaction that would occur on the surface. It is additionally advantageous to conduct these reactions in solution because multiple characterization techniques are now possible.

8 PBr3 , CH2Cl2

1) TsCl, pyr

2) NaBr, DMF

OH Br

5 5 4 5

CBr4, PPh3

CH2Cl2

TCT, DMF

NaBr

Figure 9: Bromination routes in solution using 4 as the starting material

Bromination Routes: Phosphorus Tribromide

As seen in Figure 10, the mechanism of bromination using phosphorus tribromide involves nucleophilic attack of the phosphorus by the oxygen of the hydroxyl group.19 This results in one of the bromides being liberated as a bromide anion. The remaining phosphorus species becomes a good leaving group that is ejected upon nucleophilic substitution by the bromide anion. This results in the production of the alkyl bromide and dibromophosphinous acid.

Br Br P Br Br Br P R Br O + O R H R H Br Br OH P

Br

Figure 10: Mechanism of alcohol bromination using phosphorus tribromide

9 Bromination Routes: Tosyl Chloride and Substitution

In the bromination route using tosyl chloride and sodium bromide, the formation of a tosylate intermediate replaces the alcohol with a much better leaving group that can be freed when the bromide anion performs a nucleophilic substitution on the tosylate species. The mechanism, seen in Figure 11, involves the oxygen of the hydroxyl performing a nucleophilic attack on the atom of the tosylate, directing the electrons onto one of the oxygens and forming a trigonal bipyramidal intermediate. This intermediate collapses and ejects a chloro- group, resulting in the formation of a chloride anion. The pyridine, in addition to acting as the solvent in the reaction mixture, also acts a base that abstracts a proton from the oxygen on the sulfonate. This results in the formation of the tosylate species and a protonated pyridine species.

O H R H O O R O S Cl Cl S O

O

N N H O H O O O S R S R

O O Cl

Figure 11: Mechanism for tosylation of an alcohol using tosyl chloride

When the tosylate is isolated, it can be reacted with sodium bromide, which allows the bromide anion to attack the electropositive carbon species that is alpha to the tosylate (Figure

12). This results in the release of the sodium tosylate salt and formation of the desired alkyl bromide species.

10 O

Br O S R R

O O Br- + Na+ Na O S

O

Figure 12: Bromination of a tosylate species through SN2 substitution using sodium bromide

Bromination Routes: Carbon Tetrabromide

In Appel halogenation, carbon tetrahalides have been used along with triphenylphosphine as another synthetic route to generation of halo-substituted species.20,21 The reaction mechanism, seen in Figure 13, for bromination using carbon tetrabromide is initiated when the lone pair of electrons on the phosphorus of the triphenylphosphine abstracts one of the bromo- groups on the carbon tetrabromide. This forms an “Appel salt” species that consists of a bromotriphenylphosphonium cation and a carbon tribromide anion.23 The species then abstracts the proton from the hydroxyl, generating an alkoxide anion that performs a nucleophilic substitution on the bromophosphonium and allows for the bromine to be released as a bromide anion. The resulting alkoxytriphenylphosphonium species is resonance-stabilized, with the phosphorus-oxygen bond having partial double-bond character. The bromide anion performs a nucleophilic attack on the resonance structure in which the oxygen is positively charged, generating the alkyl bromide and releasing a species as a side product.23

11 P Br P

Br H Br

Br C Br R O C Br Br Br

CBr3H

R O

P R O P

Br

R Br Br +

R O O P P

Figure 13: Mechanism of bromination of an alcohol using carbon tetrabromide and triphenylphosphine

Bromination Routes: Trichlorotriazine

The chlorination of has been performed recently using 2,4,6-trichloro-(1,3,5)- triazine (TCT), which has the advantage of being a relatively inexpensive reagent that can react with the alcohol species under relatively mild reaction conditions.22 Additionally, the formation of the corresponding bromoalkyl species has been observed with the addition of sodium bromide to the reaction mixture.22 Seen in Figure 14, the mechanism of this reaction involves the formation of a TCT-dimethylformamide (DMF) complex.22 One possible mechanism involves the movement of the lone pair of electrons on the nitrogen of DMF into the carbon-nitrogen bond, and the electrons in the carbonyl moving onto one of carbons in the aromatic triazine ring.

A tetrahedral intermediate forms as the electrons move onto an adjacent nitrogen atom, and the intermediate collapses to reform the carbon-nitrogen double bond in the triazine ring. The

12 oxygen of the hydroxyl performs a nucleophilic attack on the intermediate, generating a positively charged hemiacetal-like species. An intramolecular proton transfer allows for the electrons on the nitrogen to move into the carbon-nitrogen bond, expelling the triazine species as a leaving group. When sodium bromide is added into the mixture, the bromide anion performs a nucleophilic attack on the oxygen, which results in the regeneration of DMF and the release of the alkyl bromide product. The bromide anion is able to compete with the chloride anion as a nucleophile not only because is it present in a greater molar equivalent than the chloride but also because NaBr is more soluble than NaCl in DMF.24

R OH

Cl H Cl N Cl O N R Cl N O O N N N N O N N N H N Cl Cl N Cl

NaBr O Cl R O N H N R N H O O Br OH N N N N N R O N Cl

Cl N Cl R Br

Figure 14: A proposed mechanism for bromination of an alcohol using trichlorotriazine and sodium bromide

13 Sonogashira Coupling: Synthesis of a Pyridyl for Catalytic Complexation

For future applications of this work, the alkyne-terminated compound required for

CuAAC must have the ability to bind a transition metal complex. This research will utilize either a pyridine or a terpyridine compound, which functions as a monodentate or tridentate ligand, respectively, in transition metal complexes. The alkyne that will be used to attach the pyridyl group will be synthesized through a Sonogashira cross-coupling reaction.2 A proposed mechanism for Sonogashira coupling involving a palladium catalyst and copper (I) iodide is illustrated in Figure 15. The copper (I) co-catalyst coordinates with the ethynyl group, and trimethylamine acts as the base to deprotonate the ethynyl group to generate a copper acetylide species. The pyridine and the halide group coordinate to the palladium through an oxidative addition,25 and the halide is displaced when the palladium coordinates the ethynyl group in a transmetalation from the copper acetylide. The coupled alkyne is generated after trans/cis isomerization and reductive elimination.25

14 R1 R2

A

R = N 1 R 1 0 Pd (PPh3)2 A Ph3P Pd R2 R1 Br N PPh A= H or 3

PPh PPh3 3 X R Pd Br R1 Pd R2 1 R2 = Si X PPh X PPh3 3

X = CH3 or CH(CH3)2

[Cu] I

Cu R2

Et3NH I

H R2 H R2

Et3N [Cu] I

Figure 15: Mechanism for Sonogashira palladium-catalyzed cross coupling of pyridyl halides with (trialkylsilyl)alkynes

For the purpose of this research, a Sonogashira reaction will be used to couple a terpyridyl halide or pyridyl halide with triisopropylsilylacetylene (TIPS-acetylene) or trimethylsilylacetylene (TMS-acetylene), respectively, to form a terpyridyl triisopropylsilylalkyne (7) or pyridyl trimethylsilylalkyne (10), as seen in Figure 16.2

15 N Pd(PPh)3Cl2, CuI N N dry THF, i-Pr2NH, AgF, HClO4, NaClO4 TIPS-acetylene, N2 CH3CN, N2 Cl N TIPS N N reflux @ 80 °C N N N

6 7 8

Pd(PPh)3Cl2, CuI neat i-Pr2NH, KOH, MeOH TMS-acetylene, N2 CH2Cl2 Br N TMS N N 40 °C

9 10 11

Figure 16: Sonogashira coupling and desilylation reactions yield alkynyl pyridines

Desilylation of 7 or 10 results in the formation of the terminal alkyne 8 or 11 (Figure 16), which can then be “clicked” onto the bromo-functionalized surface via the in situ CuAAC reaction to yield 12, as demonstrated in Figure 17.

NaN , 3 O O CuSO4, A NaC6H7O6 N O O N P P A N 10 Br 10 O O N

N

A 12 A

N A= H or

Figure 17: CuAAC reactions of surface-bound bromo-terminated alkylphosphonates and synthesized alkynyl pyridines

CuAAC Reactions with Bromo-Terminated Species

In order to model the CuAAC reaction on bromo-terminated films, CuAAC will be conducted in solution using 5 as the model for a bromo-terminated species. Click reactions using

16 these bromo-terminated compounds and phenylacetylene yield 1-heptyl-4-phenyl-(1H)-1,2,3- triazole (13) and 1-(11-phosphonoundecyl)-4-phenyl-(1H)-1,2,3-triazole (14), as outlined in

Figure 18.

NaN3, CuSO , 4 N NaC H O 6 7 6 N 5 Br 5 N

5 13

NaN , O 3 O CuSO4, O N NaC6H7O6 O P P N O Br 10 O 10 N

3 14

Figure 18: Copper-catalyzed azide-alkyne cycloaddition reactions with 5 or 3 and phenylacetylene

Applications: Surface Catalysis

Ruthenium functions as a strong back-bonding metal center, and complexes containing ruthenium have been shown to reduce carbon dioxide as homogeneous catalysts in solution.26-28

Tethering the ruthenium catalytic complexes to a surface would not only allow for facile reuse of the catalyst, but it would also lower the cost of the carbon dioxide reduction process. Instead of chemically recovering the catalyst, the removal of the surface would manually recover the catalyst from the reaction vessel. The modified surface containing pyridyl or terpyridyl ligands

(12) would coordinate with ruthenium through the lone pair on the nitrogen, as shown with the terpyridyl ligand in Figure 19.2

17 O N O N O N N N P Ru(Cl) (DMSO) O Cl 2 4 P N 10 N O N 10 N O N Ru Cl

DMSO

N N

Figure 19: Ligation of ruthenium-based complexes to a copper surface with surface-bound pyridyl species as ligands

EXPERIMENTAL

Chemicals were obtained from Fisher Scientific, Acros Organics, and Alfa Aesar; reagent suppliers and purity are available in the Appendix. Chemicals were used as received unless otherwise noted. Sonication was performed using a 2510 Branson Sonicator. An MBraun Solvent

Purification System was used in order to obtain dry (THF) and dichloromethane.

Distillation procedures utilized a Kugelrohr Air Bath Oven.

Synthetic products were characterized using a 300 MHz Nuclear Magnetic Resonance

(NMR) Spectrometer and a Perkin-Elmer Spectrum RXI Fourier Transform Spectrometer for

Infrared Spectroscopy (IR). Copper surfaces were obtained from Fisher Scientific, and surface characterizations were conducted using the Perkin-Elmer Fixed Angle Specular Reflectance apparatus in conjugation with the IR. Background scans were taken of clean copper plates and subtracted from spectra for samples being evaluated by IR; all surface IRs were averaged over 64 scans at 4 cm-1resolution. IR spectra of products from reactions that were conducted in solution were taken using salt plates. NMR spectra were taken with samples dissolved in deuterated solvents containing 0.03% TMS as an internal standard. All NMR and IR spectrum peaks, unless otherwise noted, were assigned based on spectral characterization data.29-31

18 Preparation of Copper Surfaces

Copper plates (approximately 1 cm x 1 cm x 1mm) were labelled for identification using a Dremel Engraver and were suspended for all reactions using copper alligator clips obtained from Mouser Electronics. In order to clean the surfaces, the copper plates were suspended by copper clips and arranged around the edge of a 250 mL beaker. Plates were submerged in hexanes and sonicated for 30 minutes. Plates were then transferred to a clean beaker and sonicated in dichloromethane for 30 minutes. This process was repeated with three rounds of sonication in methanol for 15 minutes each. Plates were dried under nitrogen and stored in a gravity convection oven at 130°C until use.

Synthesis of 2 using the Tethering by Aggregation and Growth (TBAG) Method

A 0.60 mM solution of previously synthesized 16,10,32 (30.3 mg, 0.12 mmol) in dry tetrahydrofuran (THF) was made in a 200 mL volumetric flask. The solution was sonicated for 5 minutes in order for the phosphonic acid to be fully dissolved in THF. The phosphonic acid solution was added to 50 mL beakers, and clean copper plates were suspended in solution using copper alligator clips. The THF was allowed to evaporate below the level of each plate, and plates were heated at 140°C for 48 hours in a gravity convection oven. Plates were cooled and

-1 8 stored in a desiccator. SRIR (cm ): 2922.22- 2851.00 (CH2 stretch), 1092.54 (Cu-O-P), 1001.07

(P-O stretch).29

Synthesis of 5 from 4 using PBr3

4 (2.5 mL, 17.5 mmol) was combined with 45 mL of dry dichloromethane in a 100 mL round bottom flask and placed under nitrogen and allowed to cool to 0°C. Phosphorus tribromide

(1.14 mL, 12.0 mmol) was added over approximately 10 minutes to the flask via syringe. The

19 reaction was stirred and monitored by TLC using 4:1 hexanes:ethyl acetate and visualized using cerium ammonium molybdate (CAM) stain. The reaction was quenched with 60 mL water after

1.5 hours, and the mixture was extracted with dichloromethane (2 x 75 mL). The dichloromethane layers were combined, washed with saturated NaCl solution, dried over MgSO4, and filtered. The solvent was removed via rotary evaporation. The translucent, colorless liquid was filtered through a 1-cm plug of silica with 20% hexanes in dichloromethane, and the solvent

18 1 was removed by rotary evaporation. Yield: 1.5051 g, 52% crude. H NMR (300 MHz, CDCl3):

δ/ppm = 3.40 (2H), 1.85 (2H), 1.57-1.28 (8H), 0.87 (3H).

Synthesis of 3 from 2 using PBr3

Copper plates containing 2 were suspended in a three-neck 100 mL round bottom flask, and 50 mL dry CH2Cl2 was added in order to submerge plates in solvent. The reaction flask was cooled over ice and placed under nitrogen. PBr3 (0.10 mL, 0.12 mmol) was slowly added to the mixture via syringe, and the reaction was stirred for 3.5 hours. Water (40 mL) was added in order to quench the reaction, and plates were sonicated in dichloromethane for 20 minutes on each side. Plates were dried under N2 and stored in a desiccator prior to SRIR analysis.

Synthesis of 5 from 4 using TsCl/NaBr

4 (4.175 grams, 37.50 mmol) was dissolved in 20 mL pyridine and stirred in a 100 mL round bottom flask at 0°C. Tosyl chloride (TsCl, 10.78 grams, 56.50 mmol) was dissolved separately in 50 mL pyridine. The greenish TsCl/pyridine mixture was then added to the flask containing 4, and the reaction was allowed to proceed for 24 hours. Thin layer chromatography was conducted on the reaction mixture in 4:1 hexanes:ethyl acetate until the plate, developed

20 using a KMnO4 stain, indicated the absence of alcohol starting material. The reaction mixture was quenched with 100 mL 10% HCl and extracted with (3 x 75 mL). The ether layers were combined and washed with saturated NaCl solution, dried over magnesium sulfate, filtered, and evaporated via rotary evaporation. The crude product was clear with a yellow tint.33

1 Yield: 2.37 g, 34.8% crude. H NMR (300 MHz, CDCl3): δ/ppm = 2.45, 1.27 (8H), 0.87 (3H).

Sodium bromide (18.36 grams, 17.85 mmol) was partially dissolved in dimethylformamide (DMF). Since the sodium bromide did not completely dissolve, the solution was decanted from the undissolved solid and added to 0.8 g of the total product mixture previously obtained. The resulting reaction mixture was heated in a 250 mL round bottom flask at 60°C and allowed to proceed for 36 hours. The mixture was then diluted with 100 mL of water and extracted with 3 x 140 mL of pentane. The reaction mixture was then dried over MgSO4, filtered, and evaporated using a rotary evaporator. The product was clear with a light yellow tint

1 and had a negligible percent yield. H NMR (300 MHz, CDCl3): δ/ppm = 3.33 (2H), 1.79 (2H),

1.23 (8H), 0.82 (3H).

Synthesis of 3 from 2 using TsCl/NaBr

Plates modified with 2 were suspended in a 100 mL three-neck round bottom flask using copper alligator clips, and 40 mL of pyridine was added to submerge plates in solvent. TsCl

(0.030 g, 0.160 mmol) was dissolved in 20 mL of pyridine. The resulting greenish-yellow solution was added to the reaction vessel and allowed to stir at 0°C for 18 hours. Plates were dipped in ice-cold 10% HCl and dried under nitrogen. Plates were a burgundy color. After being dried in vacuum chamber for 1.5 hours, some light green splotches appeared on some plates.

-1 SRIR (cm ): 2997.84 (CH3 stretch), 1444.36 (S-O stretch), 747.59-688.40 (monosubstituted aromatic).

21 Plates were then suspended in DMF and allowed to stir at 60°C in a 100 mL round bottom flask. NaBr (0.06 g, 0.60 mmol) was dissolved in DMF, and the solution was added to the reaction vessel. The reaction was allowed to proceed for 20 hours, and the final reaction mixture was a light orange color. Plates were sonicated in pentane for 20 minutes and dried

-1 under nitrogen. SRIR (cm ): 2927.05-2855.30 (CH2 stretch).

Synthesis of 5 from 4 using TCT

(2,4,6)-Trichloro-(1,3,5)-triazine (TCT, 1.8361 grams, 10.0 mmol) was added to 20 mL

DMF and swirled in a 50 mL round bottom flask. The solution turned a golden yellow with the addition of DMF. NaBr (2.26 grams, 22.0 mmol) and 20 mL dichloromethane were added to the reaction vessel. The mixture was allowed to stir at room temperature for 19 hours. 1.3 mL (9.1 mmol) of 4 was added, and the mixture was allowed to react for 30 minutes. The solution was a grainy yellow mixture. The reaction mixture was washed with 20 mL water, 15 mL saturated sodium bicarbonate solution, 15 mL 1 M HCl, and 15 mL saturated NaCl. The mixture was dried over Na2SO4, filtered, and evaporated via rotary evaporation. The final reaction mixture was a clear and translucent liquid.22

Synthesis of 3 from 2 using TCT

TCT (0.036 g, 0.190 mmol) was added to 20 mL DMF in a 100 mL round bottom flask.

After all the solid had dissolved, 10 mL of dichloromethane and NaBr (0.076 g, 0.730 mmol) were added to the reaction vessel. The solution was allowed to stir at room temperature for 24 hours. Surfaces modified with 2 were suspended in the reaction vessel, and the reaction was allowed to proceed for 3 hours. Plates were removed from the reaction vessel and dipped in

22 saturated sodium bicarbonate solution, 1 M HCl, and saturated NaCl. Plates were dried in a vacuum chamber for approximately 20 minutes.

Synthesis of 5 from 4 using Carbon Tetrabromide

Carbon tetrabromide (13.2 g, 38.3 mmol) and 4 (4.1 g, 35.0 mmol) were combined in a

100 mL round bottom flask with 25 mL dichloromethane and cooled over ice.

Triphenylphosphine (PPh3, 11.3 g, 43.1 mmol) was added to the reaction vessel over 45 minutes.

The solution went from having a light yellow tint to having an orange tint. The reaction mixture was stirred for 18 hours and subsequently evaporated via rotary evaporation. A thick light yellow solid was observed. 40 mL of hexanes was added, and the mixture was evaporated via rotary evaporation after filtration of the solid. The resulting solution was translucent with a golden tint.

Yield: 8.33 g, 38.9% crude. Kugelrohr distillation was performed in order to remove

20 1 triphenylphosphine oxide byproduct. H NMR (300 MHz, CDCl3): δ/ppm = 3.41 (2H), 1.87 (2

H), 1.41-1.30 (8 H), 0.89 (3H).

Synthesis of 3 from 2 using Carbon Tetrabromide

Carbon tetrabromide (0.054 g, 0.16 mmol) was dissolved in 20 mL dry dichloromethane in a three-neck 100 mL round bottom flask, and the solution was cooled to 0°C. Copper surfaces modified with 2 were suspended in the solution, and the mixture was allowed to stir for 5 minutes. PPh3 (0.044 g, 0.17 mmol) was added to the reaction mixture over 10 minutes. The reaction was allowed to stir for 48 hours at room temperature. Plates were sonicated in hexanes

-1 and dried in a vacuum chamber. SRIR (cm ): 2927.47-2855.23 (CH2 stretch), 1436.71 (P-CH2 stretch).31

23 Synthesis of 7

Bis(triphenylphosphine)palladium (II) dichloride (Pd(PPh3)2Cl2, 27.0 mg, 0.0380 mmol), copper (I) iodide (7.6 mg, 0.040 mmol), and 6 (0.18 g, 0.67 mmol) were added to a 50 mL round bottom flask. Triisopropylacetylene (TIPS-acetylene, 0.375 ml, 1.70 mmol) was added to the flask, and the resulting mixture was placed under N2. 40 mL dry THF was added to the reaction vessel via syringe. The mixture was allowed to reflux at 80°C for 2 hours. The black solution was filtered through a 1-cm plug of silica, eluted with 10% EtOAc in CH2Cl2. The solvent was removed via rotary evaporation, and the crude product was a yellowish brown solid with a musty smell. Flash column chromatography revealed the presence of two products in the mixture, and the fractions containing each product were collected and evaporated via rotary evaporation.2

Modified Synthesis of 7

Pd(PPh3)2Cl2 (27.3 mg, 0.0390 mmol), copper (I) iodide (10.1 mg, 0.0530 mmol), and 6

(0.173 g, 6.50 mmol) were combined in a 100 mL round bottom flask. TIPS-acetylene (0.37 mL,

1.6 mmol) and 11 mL of diisopropylamine were added sequentially via syringe to the vessel. The solution was degassed through alternating exposure to high vacuum and nitrogen for 10 minutes.

The reaction mixture was then placed under N2 and refluxed at 80°C. The reaction was removed from heat after 2 hours. The resulting black mixture was filtered through silica and washed with

80 mL of 10% EtOH in CH2Cl2. A yellowish musty solid was isolated from the orange filtrate via rotary evaporation of the solvent.

Synthesis of 10

Pd(PPh3)Cl2 (0.0902 g, 0.130 mmol), copper (I) iodide (0.0592 g, 0.310 mmol), and 9

(1.023 g, 5.300 mmol) were added into a 100 mL round bottom flask, and the flask was flushed

24 with nitrogen. Dry diisopropylamine (15 mL) and trimethylsilylacetylene (TMS-acetylene, 0.80 mL, 5.7 mmol) were added to the reaction vessel via syringe. The reaction mixture progressed over 30 seconds from a grainy yellow color to an orange, brown, and finally a black colored solution. The reaction was allowed to stir at 40°C under nitrogen.

The reaction was quenched after 48 hours with 20 mL deionized water, extracted with 3 x

20 mL CH2Cl2, dried over MgSO4, filtered, and evaporated using a rotary evaporator. Kugelrohr distillation yielded a clear, translucent solid.34

Synthesis of 11

Pd(PPh3)Cl2 (0.1834 , 0.2600 mmol), copper (I) iodide (0.0507, 0.270 mmol), and 9

(1.00 g, 5.10 mmol) were added into a 100 mL round bottom flask, and the flask was flushed with nitrogen. Dry diisopropylamine (10 mL) and TMS-acteylene (0.80 mL, 5.7 mmol) were added to the reaction vessel through a pressure-equalizing funnel. The reaction mixture progressed over 30 seconds from a grainy yellow color to an orange, brown, and finally a black colored solution. The reaction was allowed to stir at 40°C under nitrogen. The reaction was quenched after 20 hours with 20 mL deionized water. Charcoal was added to the reaction mixture, and the mixture was filtered through Celite. The product was extracted twice with 40 mL CH2Cl2, dried over Na2SO4, filtered, and evaporated using a rotary evaporator.

For desilylation of the Sonogashira product, methanol (10 mL) and CH2Cl2 (5 mL) were added to the reaction vessel. KOH was dissolved in 5 mL methanol and added to the reaction vessel, and the reaction was allowed to proceed for 24 hours at room temperature. The reaction was quenched with 20 mL deionized water, and extracted with 3 x 30 mL CH2Cl2. The dark orange reaction mixture was dried over MgSO4, filtered, and then filtered through a 1-cm plug of

25 silica gel. The product was a yellow liquid when eluted with CH2Cl2. Solvent was removed via rotary evaporation to yield a brown liquid product.34

Synthesis of 13 from 5 and phenylacetylene

Phenylacetylene (0.20 mL, 2.0 mmol), 5 (0.23 mL, 2.0 mmol), (0.1376 g,

2.100 mmol), and 6 mL of 1:1 tert-butanol:deionized water were combined in a 20 mL scintillation vial. Sodium ascorbate (0.0424 g, 0.210 mmol) and copper (II) sulfate (0.0706 g,

0.440 mmol) were added to the reaction vessel, and the reaction mixture was allowed to stir at

60°C for 4 hours. The reaction was quenched with 15 mL ice water and was allowed to stir for an additional 5 minutes after the addition of 6 mL 10 % aqueous ammonium hydroxide. The solid precipitate was collected with a Hirsch funnel and allowed to dry overnight.1

Synthesis of 14 from 3 (resulting from CBr4 synthesis)

Phenylacetylene (0.30 mL, 2.7 mmol), sodium azide (0.1770 g, 2.700 mmol), 15 mL of

1:1 tert-butanol:deionized water, sodium ascorbate (0.0105 g, 0.0530 mmol), and copper (II) sulfate (0.0311 g, 0.190 mmol) were combined in 100 mL round bottom flask, and 3 was added to the reaction vessel. The reaction was stirred at 60°C for 24 hours. Plates were dipped in ice water, then 10% aqueous ammonium hydroxide, and dried under nitrogen.

Synthesis of 14 from 3 (resulting from TCT synthesis)

Phenylacetylene (0.30 mL, 2.7 mmol), sodium azide (0.177 g, 2.70 mmol), 15 mL of 1:1 mixture of tert-butanol and deionized water, sodium ascorbate (0.0105 g, 0.0530 mmol), and copper (II) sulfate (0.0311 g, 0.190 mmol) were combined in 100 mL round bottom flask, and 3 was added to the reaction vessel. The reaction was stirred at 60°C for 24 hours. Plates were then dipped in ice water, then 10% aqueous ammonium hydroxide, and dried under nitrogen.

26 RESULTS AND DISCUSSION

Synthesis of 2

The first step in the synthesis of the bromo-terminated films involved covalent attachment of a hydroxyl-terminated phosphonic acid onto copper oxide surfaces. Previously synthesized 1 was used to prepare a TBAG solution, and clean plates were immersed in the solution to allow adsorption of the species onto the copper surfaces. Baking the surfaces caused the loss of water and formation of covalent metal-oxygen bonds between the surface and the newly formed phosphonate. Surface Reflectance Infrared Spectroscopy (SRIR) was used in order to confirm the presence of the bound species, as seen in Figure 20. The signal at 1092.54 cm-1 indicated the presence of a metal-oxygen-phosphorus bond between the phosphonic acid species and the copper surface. The moderate stretches at 2851.00 cm-1 and 2922.22 cm-1 indicated the presence of methylene stretches from the alkyl chain of the surface-bound phosphonate species, and the signal at 1001.07 cm-1 indicates a phosphorus-oxygen bond.35 These peaks confirm that the surface is covalently modified with the 11-hydroxyundecylphosphonate. With this confirmation that the phosphonate containing a terminal hydroxyl was covalently bound to the surface, synthetic routes towards bromination of the hydroxyl group were explored.

27

Figure 20: Surface Reflectance IR (SRIR) of 2

Characterization of 4 and 5

NMR spectra were taken for commercially available 1-heptanol and 1-bromoheptane in

CDCl3 in order to compare the effectiveness of PBr3, TsCl/NaBr, TCT, and CBr4 as bromination reagents in solution. Figure 21 shows a triplet at a chemical shift of 3.63 ppm, which represents the two hydrogens on the α carbon to the hydroxyl group in 1-heptanol. In the NMR of 1- bromoheptane (Figure 22), a triplet is present at 3.40 ppm, which is the typical chemical shift for a protons adjacent to a Br.29 Additionally, the protons on the β carbon to the hydroxyl group appear at 1.54 ppm, while the protons on the β carbon to the bromo- group appear at a chemical shift of 1.84 ppm. The presence or absence of these signals was used to verify that the bromination reaction had taken place. In both 1-heptanol and 1-bromoheptane, signals appear for the methyl hydrogens at 0.87-0.89 ppm, and at 1.29 ppm for the remaining aliphatic protons.

28

Figure 21: NMR spectrum of commercially obtained 4

Figure 22: NMR spectrum of commercially obtained 5

29 An IR spectrum of commercial sample of 1-bromoheptane was taken as a reference in assigning the peaks in the SRIRs for each surface bromination reaction, particularly as a reference for the C-Br stretch. In the spectrum of 1-bromoheptane in Figure 23, the peaks at

-1 -1 2929.85 -2856.91cm indicate the CH2 stretch, and the peak at 2958.24 cm indicates the CH3 stretch. Peaks in the 690-515 cm-1 range and 1300-1190 cm-1 range indicate the presence of the

C-Br bond, which appear in this spectrum at 646.62 cm-1 and 1298.80-1200.25 cm-1.29

Figure 23: IR spectrum of commercially obtained 5

Bromination of 4 from 5 using PBr3

The NMR spectrum (Figure 24) of the crude product of the bromination using phosphorus tribromide revealed a triplet present at chemical shift of 3.40 ppm and a multiplet at

30 1.85 ppm, which indicated the presence of 1-bromoheptane. However, the peaks at 3.65 ppm and

1.57 ppm indicated that 1-heptanol starting material was also present in the reaction mixture. The peak at 5.31 ppm represented residual dichloromethane solvent, and water was also present in the mixture, as indicated by the peak at 1.69 ppm. Additionally, the peak at 4.05 ppm indicated the presence of isopropanol,36 which could have been introduced into the reaction mixture through an NMR tube that contained residual isopropanol. Despite the presence of several impurities in the reaction mixture, the NMR revealed that the bromination using phosphorus tribromide, although incomplete, worked in solution. Therefore, it was subsequently utilized as a reagent for bromination on copper surfaces.

Figure 24: NMR spectrum of crude 5 synthesized from 4 using PBr3

31 Synthesis of 3 from 2 using PBr3

Copper plates with 2 were submerged and reacted with PBr3. The SRIR revealed that the expected peaks around 2850 cm-1 and 2910 cm-1, indicating the presence of phosphonate methylene stretches, were not detectable. In addition, the peak near 1090 cm-1 indicating the phosphorus-oxygen-metal bond was absent (Figure 25). PBr3 reacts violently with water; therefore, the addition of water into the reaction vessel most likely lead to the formation of which had the potential to interact with the phosphonate linkages that were present on the copper surface.37 The peaks at 3447.63 cm-1 and 3330.24 cm-1 were attributed to water adhering to the surface, and the peaks present at 1034.97 - 985.40 cm-1 indicated the presence of a phosphorus-oxygen bond from a byproduct.

Figure 25: SRIR spectrum of 2 after PBr3 reaction

32 Synthesis of 5 from 4 using TsCl/NaBr

Another bromination route involved converting 4 to the corresponding tosylate, which acts as a better leaving group for the nucleophilic attack by the bromide ion. Aromatic signals were present at 7.28 ppm and 7.68 ppm in the NMR spectrum (Figure 26), confirming the presence of aromatic hydrogens, but these large peaks most likely represented residual pyridine from the reaction mixture. Tosylate aromatic peaks typically appear at 7.79 ppm and 7.34 ppm, and are difficult to discern in Figure 26 due to the overwhelming presence of pyridine signals in the aromatic region of the spectrum.38 The signal at 0.87 ppm for the hydrogens of the on the alkyl chain was present, but the signal for the methyl hydrogens on the toluene substituent that should have been present at 2.45 ppm as a singlet was seen only as a small peak.

Methylene hydrogens adjacent to the oxygen of the tosylate are typically present near 4.02 ppm,38 but were conditionally assigned as the peak at 3.53 ppm in Figure 26. Bromination using

NaBr was conducted as a follow-up procedure, even though evidence indicating substitution of the alcohol with a tosyl group was inconclusive based on this NMR spectrum.38

33

Figure 26: NMR spectrum of 4 after TsCl reaction

The percent yield for the NaBr reaction was negligible, as might be expected based on the discussion of Figure 26, but the NMR spectrum in Figure 27 revealed the presence of the bromo- group installed on the molecule, with the triplet at 3.33 ppm corresponding to the hydrogens on the α carbon to the bromo- group. However, the peaks at 3.56 and 3.46 ppm indicated the hydrogens on the carbons alpha to the hydroxyl and the tosylate, respectively. This indicated that neither the tosylation nor the bromination reaction went to completion. The NMR of the crude product also showed residual DMF as an impurity, as indicated by the peaks at 7.99-7.95 ppm,

2.89, and 2.80.29 Another impurity that was present was isopropanol, as indicated by the signal at

4.08 ppm.

34

Figure 27: NMR spectrum of crude 5 after NaBr reaction

Synthesis of 3 from 2 using TsCl/NaBr.

Copper plates with 2 were suspended in the tosylation reaction mixture. The presence of the surface-bound tosylate species was confirmed by SRIR (Figure 28), as the peak at

2997.84 cm-1 indicate the presence of an sp3 hybridized C-H bond that is present in the methyl group on the tosylate. The peak at 1444.36 cm-1 also confirmed the presence of an S-O bond, which is also present on the tosylate. Additionally, the peaks at 688.40 cm-1 and 747.59 cm-1 represent the peaks for the para-disubstituted aromatic ring structure of the tosylate. These signals confirmed the presence of the tosylate species on the surface, but the peak near 1090 cm-1 indicating the metal-oxygen-phosphorus bond was not distinct in the SRIR. Although, the tosylate species could not be confirmed to be bound to the surface, the bromination using sodium bromide was still conducted on the surface.

35

Figure 28: SRIR spectrum of 2 after reaction with TsCl

Following the reaction of the surfaces with NaBr, peaks are present for the CH2 of the alkyl chain of the phosphonate at 2927.15 cm-1 and 2855.30 cm-1. However, the characteristic peaks for a bromo-group, including the stretches in the 690-515 cm-1 region and 1300-1190 cm-1, were absent (Figure 29). Additionally, the large peak at 1414.79 cm-1 may represent unreacted tosylate.29 More importantly, the peak around 1090 cm-1 cannot be identified in either Figure 28 or 29, indicating that there is no conclusive evidence that the phosphonate species remains bound to the copper surface after either the tosylation or the bromination reaction.

36

Figure 29: SRIR spectrum of 2 after NaBr reaction

Synthesis of 5 from 4 using TCT

Another bromination route involved adding sodium bromide to trichlorotriazine (TCT), allowing for the nucleophilic substitution of the hydroxyl by a bromide that is facilitated by the triazine intermediate. The NMR of the product, seen in Figure 30, showed a triplet at 3.62 ppm, which indicated that the alcohol starting material had not been converted into 5, which would have had a peak that appeared around 3.4 ppm. A signal near 3.5 ppm, which would indicate the occurrence of a chlorination reaction of the alcohol by the TCT, was absent. This indicated that no halogenation reaction had occurred and that the amount of TCT in the reaction mixture may need to be higher in order to promote halogenation of the alcohol. Residual DMF was present

37 from the solvent in the reaction mixture, seen at 2.97, 2.89, and 8.01 ppm. Other impurities present in the crude product are residual dichloromethane, acetone, and water, present at 5.31,

2.18, and 1.66, respectively.

Figure 30: NMR spectrum of 4 after TCT reaction

Synthesis of 3 from 2 using TCT

The SRIR spectrum (Figure 31) for the TCT-reacted plates showed no indication of the surface-bound phosphonate film, as the peak indicating the oxygen-metal bond expected around

1090 cm-1 was not discernible on this SRIR. The only species that appears to be present on this copper surface is the DMF from the reaction mixture that is adsorbed to the surface, as indicated by the methyl stretches at 2930.47 cm-1 and 2865.29 cm-1.

38

Figure 31: SRIR spectrum of 2 after TCT reaction

Synthesis of 5 from 4 using CBr4

A final route that was explored for alcohol bromination was using CBr4 in the presence of triphenylphosphine. The NMR analysis of the product after reacting 4 with CBr4 revealed a spectrum that was nearly identical to that of the commercial sample of 5, with the signal at 3.41 ppm indicating the protons on the alpha carbon to the bromo- group. Additionally, the peak at

1.85 ppm represents the aliphatic hydrogens on the beta carbon to the bromo- group, and signals at 0.88-1.41 ppm indicate the remaining aliphatic hydrogens on the alkyl chain. Peaks at 7.51-

7.68 ppm revealed the presence of aromatic rings present in the triphenylphosphine remaining in the product sample as the only major impurity. After Kugelrohr distillation of the crude product, these peaks for triphenylphosphine oxide disappeared in the NMR spectrum (Figure 32). This route appeared to be the most efficient route to bromination of the alcohol in solution based on

39 relatively high purity of the sample. This reaction was conducted on copper surfaces in order to confirm the hypothesis that this reaction would work on the surface to the same degree.

Figure 32: NMR spectrum of 5 after reaction with CBr4 /PPh3 and Kugelrohr distillation

Synthesis of 3 from 2 using CBr4

Copper plates with 2 were then reacted with CBr4 and PPh3 in order to attempt the surface bromination using these reagents. The typical range for C-Br stretches are 650-510 cm-1 and

1300-1190 cm-1. There were no stretches in these ranges; instead, the SRIR spectrum (Figure 33) of the surface bound species revealed the presence of bands at 720.00 cm-1, 691.50 cm-1, and

669.58 cm-1, which fall into the range of peaks that are characteristic for aromatic compounds.

-1 These peaks, along with the peak at 1436.71 cm indicating a P-CH2 bond, suggest the presence of a phosphorus species on the surface. The stretch around 2850 cm-1 and 2927 cm-1 were still present from the alkyl chain of the phosphonate film, and the peak at 1114.52 cm-1 indicates the

40 presence of a phosphorus-oxygen double bond, most likely indicating residual triphenylphosphine oxide on the surface. The peak for the phosphorus-oxygen-metal bond around 1090 cm-1 is once again indistinguishable on the IR. This leads to inconclusive evidence of preserved phosphonate linkage on the copper surface.

Figure 33: SRIR spectrum of 2 after reaction with CBr4 /PPh3

None of the SRIR spectra taken after each bromination reaction could provide confirmation of the surface-bound species, as the phosphorus-oxygen-metal bond expected near

1090 cm-1 was not distinct in any of the spectra. It was concluded that the general noise on the spectra may have been overwhelming this peak, and click reactions would be conducted on the surfaces (synthesized from CBr4 and TCT) in order to explore this possibility.

Sonogashira Coupling Reactions

The Sonogashira cross-coupling reaction was expected to produce the pyridyl alkyne substrate for click. This molecule could then be involved in the in situ CuAAC with the bromo- terminated surface, yielding a surface-tethered pyridyl complex.

41 Synthesis of 7

An NMR spectrum (Figure 34) of starting material 6 was taken in order to compare with the spectrum of the expected product (7).

Figure 34: NMR spectrum of 6

The first Sonogashira reaction involved coupling a chloro-terpyridine species with a

(triisopropylsilyl)alkyne to yield an alkynyl terpyridine species. The NMR spectrum (Figure 35) of the product prior to conducting flash column chromatography showed peaks in the aromatic region (8.71 ppm, 8.61 ppm, 8.50 ppm, 7.89 ppm, and 7.38 ppm), possibly suggesting that the terpyridine coupled with the TIPS-acetylene that was present in the solution, as indicated by the peak at 1.62 ppm and 1.09 ppm representing the silylalkyl group.

42

Figure 35: NMR spectrum of 6 after Sonogashira reaction

The replacement of a chloro- group on the terpyridine with a TIPS group would be expected to make the molecule more nonpolar, and, therefore, the Rf value of the product would expected to be lower when carrying out TLC in a polar solvent. TLC carried out in the follow-up for the reaction showed two spots, with a lower spot in the product mixture. This suggested that the reaction generated a product that was possibly the expected Sonogashira product.

Flash column chromatography that was performed allowed for separation of the two products that were present in the final reaction mixture. The NMR spectra that were taken revealed that the product with the lower Rf (Figure 36) did not contain any peaks in the aromatic region, indicating that it did not contain pyridyl groups, while the product with the higher Rf

(Figure 37) showed peaks in the aromatic region and indicated the presence of the pyridyl groups.

43

Figure 36: NMR spectrum of first fraction of the Sonogashira product after flash chromatography

Figure 37: NMR spectrum of second fraction of the Sonogashira product after flash chromatography

44 It is possible that the new spot on the TLC that is more nonpolar was the product of oxygen-promoted homocoupling of the TIPS-acetylene. The products from this homocoupling of the silyl-acteylene groups can be seen in Figure 38. In Figure 36, there are signals at 1.09 ppm due to the hydrogens of the methyl groups of the TIPS and at 1.26 ppm due to the hydrogen on the central carbon of the isopropyl groups.

Si Si

Si Si

Figure 38: Structures for possible side products of the Sonogashira reaction due to homocoupling of TMS-acetylene (top) and TIPS- acetylene (bottom)

The spectrum of the compound with the higher Rf (Figure 37) had several peaks that were identical to peaks present in the starting material (Figure 34); it was therefore concluded that the major component of the mixture was unreacted starting material.

Modified Synthesis of 7

The Sonogashira reaction set up was repeated, but alternating high vacuum and nitrogen exposure of the vessel prior to the transmetallation reaction as a modification was performed in order to minimize the chances of exposure of the reactants to oxygen in the air. TLC revealed that the starting material and product had the same Rf values, thus the product was unreacted starting material. The reaction was unsuccessful, most likely due to the presence of oxygen in the reaction vessel, which lead to the homocoupling of the acetylene reagent. Although efforts were

45 made to sparge the vessel with nitrogen to remove dissolved oxygen, this may not have been completely effective.

Synthesis of 10

The Sonogashira reaction conditions were again modified, as the substrates in this reaction were 9, instead of the chloro-terpyridine, and trimethylsilylacetylene instead of triisopropylsilylacetylene. Since these reactants were not as sterically bulky as in the previous iterations and the bromo group is a better leaving group than the chloro group for the transmetallation, it was hypothesized that the Sonogashira reaction utilizing these compounds as substrates would be more effective. Additionally, literature suggested that the product could be isolated by extraction using dichloromethane.34 The NMR of the Sonogashira product (Figure

39) revealed a peak at 0.19 ppm that represented nine hydrogens of the trimethylsilyl group.

However, the absence of aromatic peaks at 7.31 ppm and 8.55 ppm that would be expected from the pyridine ring in the product indicated that the coupling most likely did not occur. Once again, homocoupling of the TMS-acetylene may have created this peak.39 The impurities at 1.24 and

1.56 ppm represented hexanes and water, respectively.

46

Figure 39: NMR spectrum of 9 after Sonogashira reaction

Synthesis of 11

In the final modification of the Sonogashira reaction conditions, the amount of palladium catalyst and copper (I) iodide cocatalyst that was traditionally used34 was doubled. Additionally, a pressure-equalizing funnel was utilized in order to deposit the dry diisopropylamine and TMS- acetylene in order to minimize the risk of exposure of the reaction vessel to oxygen present in the air. The final change in the procedure involved utilizing the crude product from the Sonogashira reaction directly in the desilylation procedure as a recommendation from one of the authors of the paper from the Holmes research group.34 The expected peaks were absent from the NMR spectrum (Figure 40),39 and that this modification did not increase the percent yield of the product of Sonogashira coupling. The peak at 5.25 ppm represents residual dichloromethane, and the large peak in the aromatic region is unassignable. A strategy to utilize in future research

47 would involve very slow addition of the (trialkylsilyl)alkyne in order to reduce the likelihood of the oxygen-sensitive homocoupling.

Figure 40: NMR spectrum of 9 after Songashira reaction and disilylation, taken in CD2Cl2

Synthesis of 13 from 5 and phenylacetylene

In order to characterize the click reaction that was to be performed on the bromo- terminated films, the reaction was performed in solution using commercially available 5 as the bromo-terminated molecule and phenylacetylene as the reactant containing the terminal alkyne.

Performing a CuAAC with these molecules would form a triazole ring, with one substituent being an alkyl chain and the other substituent being a phenyl ring (13), as seen in Figure 18.

Peaks were present in 900-690 cm-1 range (901.84 cm-1, 785.24 cm-1), at 1996.07 cm-1, and

1092 cm-1 that indicated the presence of the monosubstituted aromatic ring substituent. However, the peak in the 2200-2020 cm-1 range (2144.22 cm-1) indicates an azide species in the crude

48 product, indicating success of the azidation but not of the formation of the triazole through the click reaction.40 Additionally, a peak that is characteristic for a triazole ring is not present between 3300-3010 cm-1, leading to inconclusive evidence of formation of the CuAAC product.

Figure 41: IR spectrum of 5 after CuAAC reaction, dissolved in acetone

Synthesis of 14 from 3

The SRIR of 3 (resulting from reaction with TCT) after the CuAAC reaction showed the lack of distinct methylene peaks for the surface bound alkyl chain and for the P-O-Cu bond that ensures covalent linkage of the species to the copper surface. The new peak at 2065.93 cm-1 after the click reaction (Figure 42) once again represents free azide, not the successful formation of the triazole ring. The lack of any other distinct peaks on this IR spectrum indicates that the phosphonate film is no longer present on the surface. This is supported by the fact that the IR indicated the lack of film after the TCT reaction in Figure 31.

49

Figure 42: SRIR spectrum of 3 (from TCT synthesis) after CuAAC reaction

The SRIR of 3 (resulting from reaction with CBr4) after the CuAAC reaction showed retention of the methylene peaks at 2852.74 cm-1 and 2923 cm-1, as well as the peak for the phosphorus-oxygen-metal bond between the surface and the compound at 1088.21 cm-1 (Figure

43). Additionally, peaks were present in the 900-690 cm-1 range that confirmed the presence of a monosubstituted aromatic ring. The new peak at 2057 cm-1 after the click reaction, however, indicated the presence of an azide species. This evidence, along with the absence of a peak for the triazole ring in the 3300-3010 cm-1 region, indicated that triazole formation could not be confirmed. This could be due to the water peak at 3353.84 cm-1 overwhelming any other possible peaks in the region of the spectrum that would contain the triazole peak.

50

Figure 43: SRIR spectrum of 3 (from CBr4 synthesis) after CuAAC reaction

Bromination of 2 towards the assembly of a surface-bound catalyst

Of all the bromination routes characterized in solution, the route using carbon tetrabromide completely converted the alcohol into a bromo- group. However, triphenylphosphine oxide was observed as a side product on the surface and may be difficult to remove. Carbon tetrabromide and triphenylphosphine will be utilized for bromination of hydroxyl-terminated phosphonate films in future research, but optimization of the PBr3 and TCT reactions will also be performed.

51 CONCLUSIONS AND FUTURE WORK

Synthetic routes for both components that are involved as reactants in click reactions were explored. With the bromo-terminated R1 group being a phosphonate film covalently bound to copper oxide, and the R2 being the alkyne containing one or several pyridyl substituents, coupling of these components creates the potential for pyridyl-based ligand modification of copper oxide surfaces. Synthesizing a bromo-terminated film from 2 is possible through bromination of the alcohol using carbon tetrabromide and triphenylphosphine as reagents. Future work on the bromination routes will involve optimization of the reaction parameters for the PBr3 and TCT bromination reactions. Future work related to the Sonogashira reaction should focus on minimizing oxygen exposure to the reaction flask or reaction mixture, possibly through very slow addition of the silylacetylene reagent into the reaction flask. An additional technique that could be explored is freeze-pump-thaw as a way of removing dissolved oxygen from the solvent.

With the success of the Sonogashira coupling of a silylalkyne with a halo-pyridyl species, along with the desilylation of the coupled product, copper-catalyzed click coupling can occur on the bromo-terminated film through an in situ azidation and triazole formation. The resulting compound will be reacted with a species containing a ruthenium metal center in order to form a surface tethered transition metal complex.

52 REFERENCES

1. Sharpless, W.D.; Wu, P.; Hansen, T. V.; Lindberg, J. G. Just Click It: Undergraduate Procedures for the Copper (I) Catalyzed Formation of 1,2,3-Triazoles from Azides and Terminal Acetylenes. J. Chem. Ed. 2005, 82, 1833-1836

2. Gerken, J. B.; Rigsby, M. L.; Ruther, R. E.; Perez-Rodriguez, R. J.; Guzei, I. A.; Hamers, R. J.; Stahl, S. S. Modular Synthesis of Alkyne-Substituted Ruthenium Polypyridyl Complexes Suitable for “Click” Coupling. Inorg. Chem. 2013, 52, 2796-2798.

3. Himo, F.; Lovell, T.; Hilgraf, R.; Rostovtsev, V. V.; Noodleman, L.; Sharpless, K. B.; Fokin, V. V. Copper (I)- Catalyzed Synthesis of Azoles. DFT Study Predicts Unprecedented Reactivity and Intermediates. J. Am. Chem. Soc. 2005, 127, 210-216.

4. Worrell, B. T.; Malik, J. A.; Fokin, V. V. Direct Evidence of a Dinuclear Copper Intermediate in Cu(I)- Catalyzed Azide-Alkyne Cycloadditions. Science. 2013, 340, 457- 460.

5. Singh, M. S.; Chowdhury, S.; Koley, S. Advances of azide-alkyne cycloaddition-click chemistry over the recent decade. Tetrahedron. 2016, 72, 5257-5283.

6. Putvinski, T.M.; Schilling, L. M.; Katz, H. E.; Chidsey, C. E. D.; Mujsce, A. M.; Emerson, A. B. Self-Assembly of Organic Multilayers with Polar Order Using Zirconium Bonding between Layers. J. Am. Chem. Soc. 1990, 6, 1567-1571.

7. Pawsey, S.; McCormick, M.; De Paul, S.; Graf, R.; Lee, Y. S.; Reven, L.; Spiess, H. W. 1H Fast MAS NMR Studies of Hydrogen-Bonding Interactions in Self-Assembling Monolayers. J. Am. Chem. Soc. 2003, 125, 4174-4184.

8. Pawsey, S.; Yach, K.; Reven, L. Self-Assembly of Carboxyalkylphosphonic Acids on Metal Oxide Powders. Langmuir. 2002, 18, 5205-5212.

9. Queffelec, C.; Petit, M.; Pascal, J.; Knight, D. A.; Bujoli, B. Surface Modification Using Phosphonic Acids and . Chem. Rev. 2012, 112, 3777-3807.

10. Andrews, B.; Almahdali, S.; James, K.; Ly, S.; Crowder, K. N. Copper oxide surfaces modified by alkylphosphonic acids with terminal pyridyl-based ligands as a platform for supported catalysis. Polyhedron. 2016, 114, 360-369.

53 11. Vega, A.; Thissen, P.; Chabal, Y. J. Environment-Controlled Tethering by Aggregation and Growth of Phosphonic Acid Monolayers on Silicon Oxide. Langmuir. 2012, 28, 8046-8051.

12. Raman, A. and Gawalt, E. S. Self-Assembled Monolayers of Alkanoic Acids on the Native Oxide Surface of SS316L by Solution Deposition. Langmuir. 2007, 23, 2284- 2288.

13. Hanson, E.L.; Schwartz, J.; Nickel, B.; Koch, N.; Danisman, M. F. Bonding Self- Assembled, Compact Organophosphonate Monolayers to Native Oxide Surface of Silicon. J. Am. Chem. Soc. 2003, 125, 16074-16080.

14. Kim, P.; Jones, S. C.; Hotchkiss, P. J.; Haddock, J. N.; Kippelen, B.; Marder, S. R.; Perry, J. W. Phosphonic Acid-Modified Barium Titanate Polymer Nanocomposites with High Permittivity and Dielectric Strength. Adv. Mater. 2007, 19, 1001-1005.

15. Fonder, G.; Minet, I.; Volcke, C.; Devillers, S.; Delhalle, J.; Mekhalif, Z. Anchoring of alkylphosphonic acid derivatives molecules on copper oxide surfaces. App. Surf. Sci. 2011, 257, 6300-6307.

16. Al Sadat, W. I. and Archer, L. A. The O2-assisted Al/CO2 electrochemical cell: A system for CO2 capture/conversion and electric power generation. Sci. Adv. 2016, 2, 1-10.

17. Keszthelyi, T.; Paszti, Z.; Rigo, T.; Hakkel, O.; Telegdi, J.; Guzi, L. Investigation of Solid Surfaces Modified by Langmuir - Blodgett Monolayers Using Sum-Frequency Vibrational Spectroscopy and X-ray Photoelectron Spectroscopy. J. Phys. Chem. B. 2006, 110, 8701-8714.

18. Weintraub, J. G. and Parquette, J. R. Synthesis of Unsymmetrically Branched Dendrimeric Wedges up to the Fourth Generation Based on 2,3-Dihydroxybenzyl Alcohol. J. Org. Chem. 1999, 64, 3796-3797.

19. Klein, D. Organic Chemistry, 3rd edition; John Wiley and Sons, Inc.: Hoboken, 2017; 530.

20. Baughman, T. W.; Sworen, J. C.; Wagener, K. B. The facile preparation of alkenyl metathesis synthons. Tetrahedron, 2004, 60, 10943-10948.

54 21. Denton, R. M.; An, J.; Adeniran, B.; Blake, A. J.; Lewis, W.; Poulton, A. M. Catalytic Phosphorus (V)-Mediated Nucleophilic Substitution Reactions: Development of a Catalytic . J. Org. Chem. 2011, 76, 6749-6767.

22. De Luca, L.; Giacomelli, G.; Porcheddu, A. An Efficient Route to Alkyl Chlorides from Alcohols using the Complex TCT/DMF. Org. Lett. 2002, 4, 553-555.

23. Abdugadar, A. Activation of Alcohols Towards Neocleophilic Substitution: Conversion of Alcohols to Alkyl Halides. M. S. Dissertation, University of Northern Colorado, 2012.

24. Scrosati, B.; Vincent, C. A. Data Series; Pergamon Press: Elmsford, 1980.

25. Chinchilla, R. and Najera, C. The Sonogashira Reaction: A Booming Methodology in Synthetic Organic Chemistry. Chem. Rev. 2007, 107, 874-922.

26. Morris, A.; Meyer, G.; Fujita, E. Molecular Approaches to the Photocatalytic Reduction of Carbon Dioxide for Solar Fuels. Acc. Chem. Res. 2009, 42, 1983-1994.

27. Kumar, B.; Llorente, M.; Froehlich, J.; Dang, T.; Sathrum, A.; Kubiak, C.P.

Photochemical and Photoelectrochemical Reduction of CO2 Annu. Rev. Phys. Chem. 2012, 63, 541-569.

28. Keith, J. A. and Carter, E. A. Theoretical Insights into Pyridinium-Based

Photoelectrocatalytic Reduction of CO2. J. Am. Chem. Soc. 2012, 134, 7580-7583.

29. Silverstein, R. M. and Webster F. X. Spectrometric Identification of Organic Compounds, 3rd ed.; John Wiley & Sons, Inc, 1998.

30. Pavia, D. L.; Lampman, G. M.; Kriz, G. S.; Vyvyan, J. R. Introduction to Spectroscopy. 4th ed.; Brooks/Cole, Cengage Learning, 2009.

31. Bellamy, L. J. The Infrared Spectra of Complex Molecules, Vol. 1, 3rd ed.; Methuen, 1954.

32. Adden, N.; Gamble, L. J.; Castner, D. G.; Hoffmann, A.; Gross, G.; Menzel, H. Phosphonic Acid Monolayers for Binding of Bioactive Molecules to Titanium Surfaces. Langmuir. 2006, 22, 8197-8204.

55 33. Ceskis, B. A. and Maiisenkov, A. M. Effective Synthesis of the chiral components of the pheromone of leaf roller moths Adoxophyes. Zhurnal organicheskoi khimii. 1988, 24, p103-eoa.

34. Holmes, B. T.; Pennington, W. T.; Hanks, T. W. Efficient Synthesis of a Complete Donor/Acceptor bis(Aryl)diyne Family. Synth. Comm. 2003, 33, 2447-2461.

35. MacPhall, R. A.; Strauss, H. L.; Snyder, R. G. C-H stretching modes and the structure of n-alkyl chain. 2. Long, all-trans chains. J. Phys. Chem. 1984, 88, 334-341.

36. Gottlieb, H. E.; Kotlyar, V.; Nudelman, A. NMR Chemical Shifts of Common Laboratory Solvents as Trace Impurities. J. Org. Chem. 1997, 62, 7512-7515.

37. PubChem Compound Database; CID=24614, https://pubchem.ncbi.nlm.nih.gov/compound/24614 (accessed Apr. 8, 2018).

38. Asano, K.; Matsubara, S. Amphiphilic Organocatalyst for Schotten-Baumann-Type Tosylation of Alcohols under Organic Solvent Free Condition. Org. Lett. 2009, 11, 1757- 1759.

39. Ziessel, R.; Suffert, J.; Youinou, M. General Method for the Preparation of Alkyne- Functionalized Oligopyridine Building Blocks. J. Org. Chem. 1996, 61, 6536-6546.

40. Sun, S.; Wu, P. Mechanistic Insights into Cu(I)-Catalyzed Azide-Alkyne “Click” Cycloaddition Monitored by Real Time Infrared Spectroscopy. J. Phys. Chem. A. 2010, 114, 8331-8336.

56 APPENDIX

Chemicals

Reagent Supplier Purity 1-bromoheptane Sigma Aldrich 99% 4-bromopyridine hydrochloride Oakwood Chemical Reagent Grade 4’-chloro-2, 2’:6’, 2’’-terpyridine Acros Organics 99% 1-heptanol Acros Organics 98% (2,4,6)-trichloro-(1,3,5)-triazine Acros Organics 99% (cyanuric chloride) Bis(triphenylphosphine)palladium(II) Acros Organics 98% dichloride tert-Butanol Fisher Scientific Certified ACS Carbon tetrabromide Acros Organics 98% Chloroform-d Acros Organics 98% Copper (I) iodide Acros Organics 98% Copper (II) Sulfate pentahydrate Acros Organics 98% Diisopropylamine Acros Organics 99+% Ethyl acetate Fisher Chemical HPLC grade Ethyl Ether Fisher Chemical Certified ACS Hydrochloric Acid Fisher Chemical Certified ACS plus Hexanes Fisher Chemical HPLC Grade L-ascorbic acid sodium salt Acros Organics 99% Magnesium sulfate, anhydrous Fisher Scientific Certified ACS Methanol Fisher Scientific HPLC grade Methylene Chloride Fisher Chemical GC/MS grade Methyl -d6 Acros Organics 99.9 atom % Sodium azide Acros Organics 99% Sodium bicarbonate Acros Organics 99.5% Sodium bromide Allied Chemical and Dye Reagent Grade Corporation Sodium Chloride Fisher Chemical Certified ACS Sodium Sulfate, anhydrous Fisher Scientific Certified ACS Diisopropylamine Acros Organics 99.5%, extra dry N,N-dimethylformamide Fisher Scientific Certified ACS Phenylacetylene Acros Organics 98% Phosphorus tribromide Acros Organics 99% Pyridine Fisher Scientific Reagent Grade Tetrahydrofuran Fisher Scientific HPLC grade p-Toluenesulfonyl chloride Acros Organics 99+% (triisopropylsilyl)acetylene Acros Organics Reagent Grade Trimethylsilylacetylene TCI >98.0% Triphenylphosphine Alfa Aesar 99+%

57

POORNIMA RACHEL SUNDER

2471 PORT POTOMAC AVENUE [email protected] WOODBRIDGE, VA 22191 (571) 659-1213

EDUCATION

University of Mary Washington, Fredericksburg, Virginia Fall 2016- May 2018 B.S., Biochemistry

RESEARCH EXPERIENCE

University of Mary Washington, Department of Chemistry, Fredericksburg, VA Spring 2017-Spring 2018 Advisor: K. Nicole Crowder

Presentations:

Sunder, P. R. Crowder, K. N. Assembling a Surface-Bound Ruthenium Electrocatalyst Through the Use of Click Chemistry. Presented at Summer Science Institute Symposium, University of Mary Washington, Fredericksburg, VA, July 27, 2017.

Heisey, S. Andrews, B. Sunder, P. R. Keefer, A. Crowder, K. N. ω-Functionalized Self-Assembled Monolayers of Phosphonates as a Pathway to Tethered Electrocatalysis. Poster Session, ACS National Meeting, August 2017.

Sunder, P. R. Crowder, K. N. Synthesis of Phosphonic Acids to use in Click Chemistry to Assemble a Surface- Bound Ruthenium Electrocatalyst. Presented at South East Regional Meeting of the ACS, Charlotte, NC, November 10, 2017.

AWARDS

UMW Dean’s List Fall 2016-Spring 2017

American Chemical Society Undergraduate Award in Organic Chemistry 2017

John Cope ’83 Memorial Scholarship 2017

Atlantis Project Fellowship, Athens, Greece 2018

American Institute of Chemists Student Award in Biochemistry 2018

58