<<

UNDERSTANDING BASED BIOLOGY THROUGH

ADVANCEMENTS IN CHEMICAL BIOLOGY

BY

THOMAS POOLE

A Dissertation Submitted to the Graduate Faculty of

WAKE FOREST UNIVERSITY GRADUATE SCHOOL OF ARTS AND SCIENCES

in Partial Fulfillment of the Requirements

for the Degree of

DOCTOR OF PHILOSOPHY

Chemistry

May 2017

Winston-Salem, North Carolina

Approved By:

S. Bruce King, Ph.D., Advisor

Leslie B. Poole, Ph.D., Chair

Patricia C. Dos Santos, Ph.D.

Paul B. Jones, Ph.D.

Mark E. Welker, Ph.D.

ACKNOWLEDGEMENTS First and foremost I would like to thank professor S. Bruce King. His approach to science makes him an excellent role model that I strive to emulate. He is uniquely masterful at parsing science into important pieces and identifying the gaps and opportunities that others have missed. I thank him for the continued patience and invaluable insight as I’ve progressed through my graduate studies. This work would not have been possible without his insight and guiding suggestions.

I would like to thank our collaborators who have proven invaluable in their support. Cristina Furdui has provided the knowledge, context, and biochemical support that allowed our work to be published in esteemed journals. Leslie Poole has provided the insight into redox biology and guidance towards appropriate experiments.

I thank my committee members, Mark Welker and Patricia Dos Santos who have guided my graduate work from the beginning. Professor Dos Santos provided the biological perspective to evaluate my redox biology work. In addition, she helped guide the biochemical understanding required to legitimize my independent proposal.

Professor Welker provided the scrutinizing eye for my early computational work and suggested validating the computational work with experimental data. Professor Paul

Jones was kind enough to join my committee on late notice, and I look forward to his questions about radicals. Professor Leslie Poole will act as my committee chair and I thank her for taking the time out of her busy schedule to lead the assessment. I thank you all for your valuable time in reviewing my work.

I thank all of my professors at Wake Forest. Professor Paul Jones helped hone my understanding of reactive intermediates and synthesis from a rudimentary understanding ii

to a complex tool box capable of being applied throughout my work. Professor Amanda

Jones bolstered my understanding of physical organic chemistry and provided the basis of

knowledge (strain concepts) that lead to my first publication here. Professor Marcus

Wright expanded my understanding of spectroscopy to the point where I feel confident in

accurately assigning structure to complex molecules with a variety of techniques. In

addition, Marcus has provided continued support and advice that has proven valuable both in science and in life. Professor Mark Welker opened my eyes to the possibilities of transition metal organic chemistry, and provided the background to understanding and successfully executing the ring closing metathesis shown in this work.

I would like to thank all members of Team King, past and present. My lab mates throughout the years have provided invaluable insight, advice, and company when the going got tough. Zhengrui Miao and I started at the same time and he has been an excellent sounding board, and source of insightful alternative perspectives. Julie Reisz

contributed heavily to my first publication here, and provided strong advice for how to navigate graduate school.

Most importantly, I’d like to thank my family and friends. Mom and Dad, you have given me everything I need to succeed. My accomplishments would not be possible without you. You have raised me to be the scientist and person I am today. A sincere thanks to my girlfriend Kelsey Shore, I don't know how you’ve put up with me as I’ve written this dissertation. Thank you for the continued support and understanding. I’d also like to thank my bike people, you all keep my wheels rollin.

iii

TABLE OF CONTENTS

ACKNOWLEDGEMENTS ...... ii

TABLE OF CONTENTS ...... iv

LIST OF FIGURES ...... vi

LIST OF SCHEMES...... ix

ABBREVIATIONS ...... xii

LIST OF STRUCTURES ...... xiv

ABSTRACT ...... xv

CHAPTER 1 SULFENIC ACIDS AND REDOX BIOLOGY ...... 1 1.1 An Introduction to Sulfenic Acids in Redox Biology ...... 2 1.2 Disease ...... 7 1.3 Redox Balance and Signaling ...... 11 1.4 and Sulfenic acids ...... 19 1.5 Synopsis ...... 32

CHAPTER 2 STRAINED CYCLOALKYNES AS NEW TRAPS ...... 34 2.1 Introduction ...... 36 2.2 Evaluating the reactivity and selectivity of strained for sulfenic acids ...... 37 2.3 Evaluating BCN and BCN-Bio in ...... 45 2.4 Conclusion ...... 62 2.5 Materials and Methods ...... 63

CHAPTER 3 OPTIMIZING BASED SULFENIC ACID TRAPS ...... 79 3.1 Reactions with ...... 79 3.2 Addition to Alkynes ...... 84 3.3 Optimizing Strain and Electronics of Alkynes...... 86 3.4 Results ...... 92

iv

3.5 Discussion and Future Aims ...... 107 3.6 Methods Section ...... 110

CONCLUSIONS AND FUTURE OUTLOOK ...... 132

REFERENCES ...... 134

SCHOLASTIC VITA ...... 153

v

LIST OF FIGURES

Figure 1. Radical initiation, propagation, and termination...... 3

Figure 2. Lipid oxidation...... 4

Figure 3. Oxidations of nitrogenous bases, ribose backbone, and amino acids ...... 5

Figure 4. Oxidation of guanine to 8-oxoguanine...... 9

Figure 5. Oxidation of arachidonic acid by ROS or lipoxygenase...... 10

Figure 6. Hypothetical transition state of AhpC ...... 13

Figure 7. are recycled by thioredoxins ...... 14

Figure 8. hyperoxidation and resurrection ...... 16

Figure 9. Biological reactions and fates of sulfenic acids...... 20

Figure 10. Oxidation states & nucleophilic/electrophilic reactions ...... 22

Figure 11. Stable organic sulfenic acids ...... 22

Figure 12. AhpC with resolving and peroxidatic cysteine thiols ...... 24

Figure 13. Early dicarbonyl probes for sulfenic acids...... 26

Figure 14. Commercially available 1,3 dicarbonyl probes...... 27

Figure 15. Properties of nucleophilic probes...... 29

Figure 16. The electrophilic probe NBD-Cl ...... 31

Figure 17. Examples of sulfenic acid reacting with pi bonds...... 32

Figure 18. UV-Vis Kinetics; Fries Acid + BCN in acetonitrile...... 40

Figure 19. UV-Vis Kinetics; Fries Acid + BCN in 50:50 MeCN:H2O ...... 41

Figure 20. UV-Vis Kinetics; Fries Acid + trans-cyclooctene (tCOT) in acetonitrile...... 42

Figure 21. Fries Acid and dimedone in 15% MeCN:85% H2O...... 44

Figure 22. Fries Acid + Dimedone or AcOH have nearly identical UV-Vis spectrum. . 45

Figure 23. The reactions of C165A AhpC-SOH with BCN or BCN-Bio ...... 47 vi

Figure 24. Treatment of C165A AhpC-SO2H with BCN-Bio...... 48

Figure 25. Treatment of C165A AhpC-SNO and C165A AhpC-S-S-Cys with BCN-Bio 49

Figure 26. Treatment of C165A AhpC-SH with BCN-Bio...... 50

Figure 27. Western blot analysis in non-reducing conditions of C165A AhpC cysteine oxoforms...... 51

Figure 28. Polymeric addition to AhpC-SOH with high concentration of BCN-Bio...... 52

Figure 29. Positive ion MS2 (CID) spectrum of the Cys-containing peptide of C165A AhpC and BCN...... 53

Figure 30. Positive ion MS2 (CID) spectrum of the Cys-containing peptide of C165A AhpC and BCN-Bio ...... 54

Figure 31. Stability of the C165A AhpC-S(O)-BCN-Bio adduct (40 ・ M) toward

biochemical reductants (1 mM) as judged by ESI-TOF MS...... 56

Figure 32. Western blot analysis of AhpC-S(O)-BCN-Bio adduct stability in the presence of reductants: ...... 57

Figure 33. Reactivity of BCN-Bio with oxidized cell lysates...... 58

Figure 34. Labeling of endogenous protein SOH using BCN-Bio and TCEP...... 59

Figure 35. Viability of SCC-61 cells in the presence of BCN-Bio and dimedone as measured by the MTT assay...... 60

Figure 36. Labeling of RSOH in live cells and purified AhpC, vs DCP-Bio1...... 62

Figure 37. radical cycle...... 81

Figure 38. Methods of moderating strain ...... 87

Figure 39. Combining strain and electronic activation...... 89

Figure 40. Computational evaluation of methane sulfenic acid reacting with BCN ...... 94

Figure 41. Five alkynes were reacted with Fries acid under pseudo first order conditions...... 96

vii

Figure 42. The benchmark alkynes were evaluated at the aforementioned 6 levels of theory and results were compared...... 97

Figure 43. Graphing computational energy barrier vs empirical kinetics ...... 98

Figure 44. The effect of electronics on energy barriers ...... 99

Figure 45. Exocyclic heteroatoms align gauche with the alkyne, while endocyclic atoms align anti and have greater effect on electronics...... 100

Figure 46. Endocyclic and exocyclic heteroatoms were evaluated for their effect on energy barriers at the “B” level of theory (B3LYP 6-31G(2d))...... 101

Figure 47. Evaluating strain energy as a function of heteroatom location and ring size. 102

Figure 48. UV-Vis Kinetics; Fries Acid + DIBO in acetonitrile...... 127

Figure 49. UV-Vis Kinetics; Fries Acid + BCN in acetonitrile...... 128

Figure 50. UV-Vis Kinetics; Fries Acid + COT in acetonitrile...... 129

Figure 51. UV-Vis Kinetics; Fries Acid + DBCO in acetonitrile...... 130

Figure 52. UV-Vis Kinetics; Fries Acid + DIBOTM in acetonitrile...... 131

viii

LIST OF SCHEMES

Scheme 1. Superoxide dismutation ...... 7

Scheme 2. Synthesis of Fries acid ...... 21

Scheme 3. The 1,3 diketone known as dimedone reacts with sulfenic acids via a nucleophilic mechanism to form covalent bonds...... 25

Scheme 4. The reaction of AhpC with BCN-Bio ...... 35

Scheme 5. The trapping of sulfenic acids by dimedone and 9-hydroxymethyl- bicyclo[6.1.0]nonyne (BCN)...... 37

Scheme 6. Reactions of BCN with a cysteine-derived sulfenic acid and Fries acid yield alkenyl ...... 38

Scheme 7. Reaction of thermally generated organic-soluble cysteine-derived sulfenic acid with BCN ...... 64

Scheme 8. Reaction of thermally generated organic-soluble cysteine-derived sulfenic acid with tCOT(trans-Cyclooctene) ...... 65

Scheme 9. The Reaction of Fries Acid with BCN...... 66

Scheme 10. The reaction of Fries Acid with tCOT...... 67

Scheme 11. Synthesis of BCN-Bio ...... 68

Scheme 12. Reactivity of BCN with various small molecule sulfur oxoforms ...... 70

Scheme 13. Stability of BCN-Alkenylsulfoxide to common reducing agents ...... 71

Scheme 14. Thiols can add to alkynes twice...... 80

Scheme 15. Addition of cysteine to BCN ...... 83

Scheme 16. The reaction of BCN with ...... 84

Scheme 17. MSBTA is capable of blocking thiols and persulfides while sparing sulfenic acids...... 86

Scheme 18. Alkyne electronics effect regiochemistry of RSOH addition...... 91

ix

Scheme 19. Retrosynthetic analysis of MOCO with alcohol handle...... 103

Scheme 20. Attempted synthesis of MOCO-OH...... 105

Scheme 21. Synthesis of BMOCO-OH...... 106

Scheme 22. Synthesis of MOCO with linker...... 107

Scheme 23. Alternative alkyne synthesis ...... 109

Scheme 24. Grignard addition to , 1-(allyloxy)hex-5-en-2-ol ...... 112

Scheme 25. Alcohol protection to ((1-(allyloxy)hex-5-en-2-yl)oxy)(tert- butyl)diphenylsilane ...... 113

Scheme 26. Ring closing metathesis to (Z)-tert-butyldiphenyl((3,4,5,8-tetrahydro-2H- oxocin-3-yl)oxy)silane ...... 114

Scheme 27. Attempted bromination of (Z)-tert-butyldiphenyl((3,4,5,8-tetrahydro-2H- oxocin-3-yl)oxy)silane ...... 115

Scheme 28. Grignard addition to , forming 1-(2-(allyloxy)phenyl)but-3-en-1-ol ...... 116

Scheme 29. Protection of secondary alcohol to form ((1-(2-(allyloxy)phenyl)but-3-en-1- yl)oxy)(tert-butyl)dimethylsilane ...... 117

Scheme 30. Ring closing metathesis to form (Z)-tert-butyl((5,6-dihydro-2H- benzo[b]oxocin-6-yl)oxy)dimethylsilane ...... 118

Scheme 31. Attempted bromination of (Z)-tert-butyl((5,6-dihydro-2H-benzo[b]oxocin-6- yl)oxy)dimethylsilane ...... 119

Scheme 32. Parikh Doering oxidation to form 4-((tert-butyldimethylsilyl)oxy)butanal 120

Scheme 33. Grignard addition to aldehyde, forming 1-((tert-butyldimethylsilyl)oxy)non- 8-en-4-ol ...... 121

Scheme 34. O- allylation to form ((4-(allyloxy)non-8-en-1-yl)oxy)(tert- butyl)dimethylsilane ...... 122

x

Scheme 35. Ring closing metathesis to form (Z)-tert-butyldimethyl(3-(3,4,5,8-tetrahydro- 2H-oxocin-2-yl)propoxy)silane ...... 123

Scheme 36. Bromination to form tert-butyl(3-(6,7-dibromooxocan-2-yl)propoxy)- dimethylsilane ...... 124

Scheme 37. Bromine elimination to form (E)-(3-(7-bromo-3,4,5,8-tetrahydro-2H-oxocin- 2-yl)propoxy)(tert-butyl)dimethylsilane ...... 125

xi

ABBREVIATIONS ROS – reactive species

RNS – reactive specieses

GPx – glutathione peroxidase

Prx – peroxiredoxin

AhpC – alkyl reductase subunit C (a peroxiredoxin)

Trx - thioredoxin

GSH – glutathione

RIPA Buffer – radioimmunoprecipitation buffer

LA - linoleic acid

AA – arachidonic acid

EPA - eicosapentaenoate

DHA – docosahexaenoate

SOD - superoxide dismutase

NADPH - nicotinamide adenine dinucleotide phosphate

AD - Alzheimer’s disease

PD - Parkinson’s disease

ALS - amyotrophic lateral sclerosis

xii

PTP - protein tyrosine phosphatase

PTEN - tumor suppressor phosphatase and tensin homolog

GAPDH - glyceraldehyde-3-phosphate dehydrogenase

BCN - bicyclononyne

COT - cyclooctyne

MOCO – mono oxygen cyclooctyne

DFCO – difluorocyclooctyne

MOFO - monofluorinated cyclooctyne

DIBO – dibenzocyclooctyne

DIBAC – dibenzoazacyclooctyne

TMDIBO – tetramethoxydibenzocyclooctyne

DBCO – dibenzocyclooctyne

xiii

LIST OF STRUCTURES

OH OS

OOHO H

O H Fries Acid dimedone bicyclononyne (BCN)

O O H H HN O NH H S H BCN-Biotin (BCN-Bio) cyclooctyne COT

O O

O O N H2N OH O OH dibenzocyclooctynol tetramethoxydibenzocyclooctyne dibenzocyclooctyne (DIBO) (TMDIBO) (DBCO)

O OO O monooxygencyclooctyne - dioxygencyclooctyne (MOCO) monooxygencyclooctyne (DOCO) (BMOCO)

xiv

ABSTRACT Sulfenic acids are fleeting intermediates formed from the oxidation of cysteine

thiols by a number of biologically relevant oxidants. Sulfenic acids are quickly converted to other species including , , sulfinic acids, or sulfonic acids, leading to changes in protein function. The resulting structural and functional changes

can modulate enzyme activity, activate transcription factors, and lead to signaling

cascades. Understanding the intricacies of sulfenic acid signaling is important for

understanding healthy cellular function and disease states. Identifying the site, timing, and conditions of protein sulfenic acid formation is crucial to understanding cellular redox regulation.

The transient nature of sulfenic acids necessitates study with bioorthogonal traps.

Current methods for trapping and analyzing sulfenic acids involve the use of dimedone and other nucleophilic 1, 3-dicarbonyl probes that form covalent adducts with cysteine- derived protein sulfenic acids. As a mechanistic alternative, we have evaluated highly strained bicyclo[6.1.0]nonyne (BCN) derivatives as concerted traps of sulfenic acids.

These strained cycloalkynes react efficiently with sulfenic acids in proteins and small molecules yielding stable alkenyl products at rates more than 100x greater than

1, 3-dicarbonyl reagents enabling kinetic competition with physiological sulfur chemistry.

We have demonstrated the selectivity of BCN based probes for sulfenic acids by protein lysate and small molecule experiments. Our experiments show no reactivity with thiols, S-nitrosothiols, sulfinic acids, and sulfonic acids. In contrast, recent literature indicates that BCN and strained alkynes are prone to reacting with thiols under radical conditions via “thiol-yne” reactions, and with persulfides via a mechanism similar to the

xv

reaction of alkynes with sulfenic acids. However, it is not completely clear to what

extent thiol-yne and persulfide trapping reactions occur under biological conditions. In

support of the selectivity of strained alkynes, they are widely used as bioorthogonal

reagents in copper free click chemistry. Indeed, the enhanced rates and predictable

reactivity of strained alkyne probes make them valuable tools for evaluating transient

sulfenic acids in complex cellular environments

We have employed computations to identify alternative strained alkynes which

should react quickly and more selectively than BCN. Density functionals and hybrid

density functionals have been employed with a range of basis sets. We calibrated the

computational results with experimental results from three synthesized and two

commercially available strained alkynes. The B3LYP-6-31G(2d) theory correlated best

with experimental data and was used to assess barrier heights between sulfenic acids and

hypothetical alkynes. Computational experiments identified cyclooctynes with

endocyclic propargylic heteroatoms as the most viable sulfenic acid traps. Synthesis of a

number of choice heterocyclic alkyne candidates was attempted. The enhanced rates

demonstrated by current alkynes identify them as new bioorthogonal probes that should

facilitate the discovery of previously unknown sulfenic acid sites and their parent proteins.

The synthesis and computations described in this work should prove valuable to continued research in this area.

xvi

CHAPTER 1

SULFENIC ACIDS AND REDOX BIOLOGY

Thomas Poole

Department of Chemistry, Wake Forest University, Winston-Salem, NC 27109

1

1.1 An Introduction to Sulfenic Acids in Redox Biology

Interest in the field of redox biology is rapidly expanding. Redox biology studies both

the consequences of aberrant oxidation and the complexities of signal transduction

through the carefully orchestrated networks of cellular communication.1,2 Reactive

oxygen species (ROS) are the primary currency of redox biology. ROS are both the

expected product and deleterious consequence of cellular respiration. ROS primarily

originate in mitochondria, where they begin as highly reactive species prone to

indiscriminate reactivity, but still capable of participating in controlled signal

transduction.3 The mitochondria quickly convert these stronger oxidants to more docile

species which can function locally, move outward into the cell as controlled messengers,

or act as destructive emissaries in times of stress.4 Sulfenic acids are a vital tool used by

the conductors of this symphony.5 The fleeting existence of sulfenic acids marks the

point where an oxidant meets it fate. Sulfenic acids are the switch which triggers a cell to

either perform its daily duties or mount a defense against oxidative assailants. The

research herein focuses on chemical approaches to delineate the role sulfenic acids play

in redox biology. Only through understanding the integral role of sulfenic acids in redox

biology, can we paint an accurate likeness of cellular function.6

Oxidative Balance and Damage

Reactive oxygen species are formed as both a consequence and desired product of

natural cellular processes. They include superoxide, peroxide, ,

hypohalorous acid, and peroxynitrite.7 Under normal conditions their location and concentration is tightly controlled. However, under conditions of stress where their concentrations and lifetimes may be higher, their highly reactive nature lends them to

2

aberrant and often self-propagating reactivity. For instance, superoxide and hydrogen

peroxide are capable of forming hydroxyl radical which can damage proteins, lipids,

polysaccharides, and DNA.1 After initiation, the free radicals can remove a hydrogen

atom [H•] from a number of species, the resulting alkyl radical can then react with O2 and

form an alkyl peroxyl radical which can then continue to remove H• from other species

(Figure 1). Some of the most highly oxidizable molecules include lipids and poly

unsaturated fatty acids such as linoleic acid (LA), arachidonic acid (AA),

eicosapentaenoate (EPA), docosahexaenoate (DHA), and cholesterol (Figure 2).8 These oxidations readily form adjacent to sites of unsaturation, due to greater stability of the resulting radicals. Lipid peroxidation has been implicated in many diseases, including atherosclerosis, heart disease, cancer, diabetes, and neurodegenerative disorders.1,9,10

Figure 1. Radical initiation, propagation, and termination.

3

Figure 2. Lipid oxidation. Oxidation occurs primarily at allylic sites which form more stable radicals.

Further damage can be incurred when macro molecules such as proteins and DNA are the target of damage. Similar to lipids, the nitrogenous bases of DNA can have an H• abstracted leading to a cascade of reactions. Under aerobic conditions, the radical usually reacts with oxygen, forming a peroxy species which is then capable of propagating the radical chain by abstracting a H• from another source.11–13 Additionally, amino acids and

proteins can be oxidized in similar ways to DNA and lipids. Starting with an abstraction

step on side chain (1) or backbone (2), then reacting with oxygen, and possibly

propagating, leading to further damage.14 Interestingly, some amino acids are particularly

good or bad at propagating radical chain reactions, for instance the branched chain amino

acids valine, leucine, and isoleucine are as much as 50x more efficient at propagating free

radical reactions when compared to other amino acids.15 A primary source of natural

ROS capable of performing these reactions is from within the mitochondria.14,16,17

4

Figure 3. Oxidations of nitrogenous bases, ribose backbone, and amino acids

The mitochondrial respiratory chain is a major source of cellular ROS.3,18 The

- initial reactive oxygen product of these reactions is commonly superoxide (O2• ).

Superoxide is produced by one electron reduction of O2, primarily as a respiratory chain

product in the inner mitochondrial membrane.4 The factors governing superoxide

generation rate are highly complex and dependent on everything from O2 concentration,

5

energy levels, organism, tissue, hormonal status, to age.19 After production, the primary

destination of superoxide is the mitochondrial matrix, and intermembrane space. While

superoxide is highly reactive and capable of oxidizing many targets, it rarely reacts with

other species due to dedicated enzymatic processes for converting it to more manageable species.

Superoxide dismutases convert superoxide to oxygen and as shown in Scheme 1.20 Superoxide in the intermembrane space is converted to hydrogen

peroxide and water by the copper-zinc superoxide dismutase one (SOD1), while

superoxide in the matrix is dismutated by the manganese dependent dismutase (SOD2). 4

While superoxide conversion to hydrogen peroxide can occur spontaneously, SODs prevent the buildup of superoxide and aberrant oxidation by accelerating the reaction over 10,000 fold to near diffusion controlled rates.4,21 While elevated superoxide in the mitochondria leads to excess oxidation, the consequences are much more severe in the matrix rather than the intermembrane space. For example, SOD2 (matrix) has been found to be essential in murine models, full MnSOD knockouts were found to die within a few days,22 and heterozygous mice had severe deficiencies and were highly sensitive to

aerobic environments.23,24 While complete SOD1 (intermembrane) deficiency is less

severe, it results in reduced glucose tolerance, muscular atrophy, and ROS sensitivity.25,26

Once SODs have converted most of the superoxide to hydrogen peroxide, it can be handled by cysteine and sulfenic acid redox pathways e.g. peroxiredoxins.5,27,28

6

Scheme 1. Superoxide dismutation

The mitochondrial respiratory chain, NADPH oxidases, and the dismutation of

2,29,30 superoxide are major sources of cellular H2O2. Hydrogen peroxide is reduced by the

oxidation of cysteine thiols to sulfenic acids in redox sensitive proteins. Specifically,

H2O2 is reduced by catalases, peroxiredoxins, glutathione peroxidases, and ascorbate

peroxidases, depending on mitochondrial location. Catalases, peroxiredoxins and

glutathione peroxidases reduce peroxides around 107 M-1s-1 and thus scavenge most of

the peroxide.30–32 A kinetic competition analysis estimates that Prx 3 will be the target for up to 90% of hydrogen peroxide generated in the mitochondrial matrix.32 Oxidants

within the mitochondrial intermembrane space can freely flow into the cytosol via

porins33 which allow unrestricted flow of molecules up to 5 kDa. However, flow between

the matrix and intermembrane space is tightly controlled by substrate-specific carriers.4,33

Once in the cytosol, most hydrogen peroxide reacts with the cysteine thiols of recycling enzymes to form sulfenic acids and begin the recycling process.31,32

1.2 Disease

Sulfenic acids are one of the core signaling molecules within redox biology,

likewise they help carry the overall tone of adaptive or maladaptive processes. Sulfenic

acids and their parent cysteine thiols participate in the control and movement of ROS

from their origin in the mitochondria and through their efflux to the rest of the cell. Their

signaling is carefully balanced in an adaptive cellular setting. However, some diseases

7

such as cancer can take advantage of these signaling pathways to promote their growth,

proliferation, and metastasis.1,34 Other diseases associated with redox balance occur when

cellular regulation systems are overwhelmed, tipping the scales towards oxidation. When

left unchecked, excess ROS and RNS begin to damage proteins,15,35,36 lipids,8,37 and even

DNA.11–13,38 The damage of these vital molecules can lead to consequences ranging from decreased cellular efficiency, inflammation, apoptosis, and chronic disease states.

Associated diseases include but are not limited to osteoarthritis,39 cardiovascular disease,

diabetes mellitus, and increased cancer risk involve abnormal ROS control.1,40 Peroxide

reducing enzymes known as peroxiredoxins act as an excellent barometer for the redox

state of cells, and as a result they make excellent starting points when trying to

understand the redox implications of diseases.1,27

Cancer

Damage to DNA as shown in Figure 3 is widely accepted as one of the major

causes of cancer.11,41,42 The majority of ROS induced mutations leading to cancer are

initiated by damage to guanine, leading to guanine to thymine transversions (Figure 4).43–

46 If these changes affect critical oncogenes or suppressor genes, then carcinogenesis can

occur.47 The resulting cancer cells often have robust respiratory systems that enable

expedient growth and proliferation. As a result they often have abnormal mitochondria,48 and knowing that ROS are produced in the mitochondria, it logically follows that cancer cells produce large amounts of ROS.49 In addition, mitochondria control cell death

pathways and upregulation of glycolysis.50 For instance, the Prx systems of the cells are

often upregulated to prevent ROS induced apoptosis, and buffer the extra ROS produced during enhanced respiration.51 In fact, upregulation of almost all Prx isoforms leads to a

8

tumor promoting effect.1 For instance upregulation of Prx1 is seen in almost all cancers,

and is associated with increased tumor grade52 and increased recurrence after radiative

treatment.53 However, there are a few cases where it suppresses tumors.54 In human

breast cancer cells deficient of Prx1, there was a downregulation of the c-Myc oncogene,

leading to decreased tumorigenesis.55

Figure 4. Oxidation of guanine to 8-oxoguanine. The oxidative modification of guanine to 8-oxoguanine is one of the most common oncogenic modifications. It pairs preferentially with cytosine, resulting in GC to TA conversion after replication.56

Atherosclerosis

Cardiovascular diseases (coronary artery disease, stroke, and peripheral vascular

disease) are the leading cause of death in the western world, with 614,348 deaths in

2014.57 Cardiovascular diseases are primarily caused by atherosclerosis and

inflammation.58,59 Atherosclerosis is the narrowing and hardening of arteries due to lipid

build-up on the arterial walls. The oxidative modification hypothesis of atherosclerosis

states that one of the causative factors is the aberrant oxidation of low-density lipoprotein

(LDL).60,61 In support of this hypothesis is the presence of higher concentrations of

oxidized molecules within plaque, including proteins and lipids. Aberrant oxidation of

lipids or even enzyme mediated oxidation of the proinflammatory arachidonic acid in

9

phospholipids, can be recognized by the immune system and lead to inflammation

60,62,63 (Figure 5). ROS that can perform these oxidations include H2O2, , superoxide, hydroxyl radical, and nitric oxide. Oxidized lipoproteins activate receptors on

endothelial cells and upregulate adhesion and inflammation molecules.62,65–67 Foam cells particularly promote inflammation and growth of atherosclerotic plaque.68 Pro- inflammatory cytokines and growth factors secreted by the foam cells increase ROS in the plaque-lesion.69 Elevated ROS leads to increased production of extracellular matrix

proteins by vascular muscle cells causing hardening of the arteries, buildup of plaque and

associated pathogenesis.64 As a protective effect Prxs have been shown to be upregulated

in response to increased endothelial ROS, and protect against inflammation and

associated atherosclerosis.70–72

Figure 5. Oxidation of arachidonic acid by ROS or lipoxygenase. Aberrant oxidation of

lipids or even enzyme mediated oxygenation, activates low density lipoprotein receptors

leading to inflammation and disease progression.

Neurodegenerative diseases

The brain is a region with high energy requirements and high mitochondrial

density. The greatest risk factor for neurodegeneration including Alzheimer’s disease

(AD), Parkinson’s disease (PD), amyotrophic lateral sclerosis (ALS), and Huntington’s

disease is aging. Mitochondria are thought to be major contributors to aging through their accumulation of mutations and production of ROS.73 While oxidative damage isn’t

10

always indicated as a causal factor of neurodegenerative diseases, it has a high rate of co-

morbidity. In Alzheimer’s disease, oxidative damage precedes pathogenic plaque build-

up74 and apoptotic cells co-localize with neurons under oxidative stress.75 In Parkinson’s

disease inhibition of respiratory complex 1 (which is known to increase ROS) and

glutathione depletion were found in patients with early Parkinson’s.76,77 It was found that

neurons treated with a peptide (amyloid beta peptide) associated with AD showed

overexpression of Prx1 which was shown to be antiapoptotic by counteracting injury.78 In human brains with PD there is increased S-nitrosylation of Prx2 which prevents its peroxidatic function, making dopaminergic neurons particularly susceptible to oxidative stress.79

1.3 Redox Balance and Signaling

The initial generation of superoxide in the mitochondria begins the flow and

signaling of oxidants throughout the cell. First the superoxide is quickly converted by

superoxide dismutases to H2O2, primarily in the mitochondria, also throughout the cell.

The quick conversion of superoxide to hydrogen peroxide results in a short lifetime of the

superoxide and limited direct use in cellular signaling. However, the hydrogen peroxide

28 that is produced is “implicated as the main transmitter of redox signals.” H2O2 in

2 aerobic cells typically exists at steady state of about 10 nM. The direct reaction of

hydrogen peroxide with most protein thiols and low molecular weight thiols like

glutathione is quite slow. The oxidation of thiols by hydrogen peroxide is commonly

thought to proceed through formation of the thiolate, then nucleophilic attack on oxygen

with –OH leaving forming a sulfenic acid. Under simple conditions –OH is a very poor

leaving group, resulting in a high reaction barrier of ca. 46.6 kcal/mol80 with nucleophiles

11

and a slow reaction rate of around 20 M-1 s-1 with thiolates.31,81. However, dedicated

detoxifying enzymes such as glutathione peroxidase (GPx), and peroxiredoxin (Prx) react

quite quickly owing to transition state stabilization by their active sites (Figure 6).82 The near diffusion limited rate allows peroxiredoxins to scavenge most H2O2, making them

the “dominant peroxidases in most organisms.”83

Peroxiredoxins83

Peroxiredoxins were first thought to be secondary to other peroxide catalysts such as Gpx

and catalase.84 However it is now known that they react at rates as high as 107 M-1 s-1, and have a high cellular abundance.85,86 Their concentration and high reaction rates allow

them to function as primary redox machinery, in fact it has been estimated that 99% of

peroxide in the cytosol and 90% of peroxide in the mitochondria will react with

peroxiredoxins rather than small molecules such as, proteins, lipids, enzymes, or other

thiols.31,32 This efficient reactivity is enabled by a robust catalytic cycle that centers

around sulfur-based redox reactivity.

The catalytic cycle centers around a cysteine thiol within Prxs, referred to as the

peroxidatic cysteine thiol or “CP”. Prior to beginning the cycle, the protein must fold into

a fully folded or “FF” form which facilitates the transition state. The enzymatic pocket

helps deprotonate the cysteine thiol into a thiolate which is able to react with peroxide at

20 M-1 s-1.31,81 However, the interaction isn’t limited to deprotonation. In addition, the

pocket facilitates –OH acting as a leaving group in an SN2- type fashion. Hydroxide

which is normally a poor leaving group, is helped through the transition state by a local

arginine, as well as backbone N-H bonds, allowing for rate enhancement up to the

12

observed 107 M-1 s-1 (Figure 6).82 This reaction results in generation of sulfenic acid or

Cys-SOH.

Figure 6. Hypothetical transition state of AhpC. The peroxidatic reaction above adapted

from a crystal structure with bound hydrogen peroxide. The S to OA bond is partially

formed, and the OA to OB bond is partially broken. The local hydrogen bonding

facilitates the buildup of partial negative charges on OB and the Cp sulfur. This figure

was adapted from Hall et al.82

In the locally folded form of Prx, the sulfenic acid isn’t accessible for attack from

the resolving cysteine. However, the protein is typically in fast equilibrium between the

FF form and the locally unfolded form (UF), allowing for access to the CP and formation.87 When locally unfolded, the sulfenic acid can form a disulfide bond with a variety of partners including local resolving (CR), a CR of another adjacent Prx,

or with a intermolecular partner such as glutathione. The disulfide formation locks the

enzyme into the open form, allowing for recycling of the enzyme.

13

The recycling of Prx disulfides into thiols, is commonly performed by thioredoxin

(Figure 7) but can also be performed by a variety of molecules ranging from glutathione,

dithiothreitol (not biological), or specific Prx reductases similar to thioredoxin. With the

enzyme being locked into the open form, it is believed that recycling proceeds via sulfur

attack on the resolving cysteine sulfur, followed by displacement of the peroxidatic

cysteine thiolate.88–90 Once reduction is complete, the enzyme is able to refold into the

fully folded form and accept peroxide for reduction.

H O 2 2 HS S Prx Trx 2H2O HS S

HS Trx Prx reductase HOS S HS Prx Trx S HS

Figure 7. Peroxiredoxins are recycled by thioredoxins.

Hyperoxidation

If a secondary peroxide is present in the before resolution or recycling occurs, then hyperoxidation can occur. The sulfenate or RSO- form of the sulfenic acid

can react with another peroxide, in the FF active site, leading to formation of a sulfinic

91 acid or RSO2H. This sulfinic acid is no longer able to participate in the catalytic cycle

and effectively deactivates the enzyme. Interestingly, kinetic analysis of the

hyperoxidation rate of sulfenic acids in some PrxII reveals that this event occurs about

1000 times slower than the initial oxidation to sulfenic acid.92 Interestingly, this second

14

oxidation is still several orders of magnitude faster than oxidation of most (non catalytic)

reduced protein thiols to sulfenic acids.92 However, some Prxs can be oxidized even

further to the .91

Hyperoxidation relies on the protein being in the fully folded conformation and

presence of peroxide in the active site. If the equilibrium between FF and UF lies further

toward the unfolded side, or if there are low peroxide concentrations, then hyperoxidation

is slower. However, it is most strongly dependent on the concentration of peroxide due

to the rapid equilibrium between FF and UF states.93 Kinetic assessments which

manipulate the concentration of hydrogen peroxide can compare the relative propensity

for Prxs to hyperoxidize. The unit used to compare relative oxidizability is Chyp1% which

refers to the concentration at which 1% of the enzyme becomes deactivated per catalytic

cycle.94 The propensity for hyperoxidation and inactivation may be of importance in

cellular signaling. Of particular interest is the Prx1 subfamily which is highly susceptible to hyperoxidation and can undergo an additional recycling or “resurrection” step.93 This

“resurrection” step involves ATP-dependent conversion of the inactive sulfinic acid, or sulfinate, back to the sulfenic acid that can then progress through a normal recycling step

(Figure 8). The reaction involves transfer of a terminal phosphoryl group from ATP to the sulfinate oxygen, followed by attack on the sulfur by sulfiredoxin thiol, to displace the phosphoryl group and create the mixed Prx-S(O)-S-Srx .95 This mixed thiosulfinate can be reduced by displacement of the Prx-SOH with glutathione or thioredoxin, and the resulting sulfenic acid can react with a resolving sulfur and be reduced by normal means.96

15

S SR Srx

RSH H O 2 2 HS normal HS Prx Trx Prx 2H2O recycling S HS HS Prx OH O S S HS Prx Srx HOS HS HS Srx H2O2 Prx Prx O O O HO S S SH P 2 ADP O HO 2H O ATP O OH 2 P Srx HO OH

Figure 8. Peroxiredoxin hyperoxidation and resurrection

Signaling

Hydrogen peroxide signaling can proceed through three mechanisms which likely all occur to varying extents, depending on concentrations, redox states, and location within cells. First, H2O2 reacts directly with signaling proteins, and the usual redox

regulators of the cell such as Prx, can mediate this by adjusting peroxide

concentrations.2,97 Secondly, Prx can be oxidized and then relay this signal to specific

signaling proteins.2,98 Third, the oxidation of Prx can be transferred to a thioredoxin which then relays oxidation to a signaling protein.2,99 As more kinetic data emerges it

appears that the limited cellular lifetime and concentration of H2O2 makes it unlikely to

react with most signaling proteins under normal conditions.31,97 However, all signaling

methods are dependent on the nature of the substrate protein, H2O2 concentration, pH, and a number of other factors.

16

The direct oxidation of reactive thiols in signaling proteins such as phosphatases

Cdc25B and protein tyrosine phosphatase 1b have been shown to be 5-6 orders of magnitude slower than the Cp reactivity in Prxs, making them very unlikely to be

31 oxidized by H2O2. In addition, Prx has concentrations much higher than signaling

proteins like phosphatases and transcription factors. The combination of concentrations,

diffusion coefficients, and reactivity give diffusion distances of 2000 and 4 µm for Prx

and protein tyrosine phosphatase (PTP) respectively, to illustrate that direct oxidation is a

rare event.31 Nevertheless, a number of signaling proteins with slow reacting cysteines,

including PTPs, and the tumor suppressor phosphatase and tensin homolog (PTEN), were

found to be oxidized in biological tissue.100,101

These observed limitations of H2O2 signaling have resulted in the development of

the “floodgate hypothesis,” where short term increases in H2O2 concentration allow for

oxidation of less reactive thiols and signaling events to occur due to Prx inactivation.

93,102 The first evidence in support of this theory was the discovery of varying sensitivity

to hyperoxidation from sulfenic to sulfinic acids in different Prxs,93 which would lead to

inactivation because sulfinic acids aren’t able to be reduced by common Prx reductants

such as Trx and Grx. Additionally, the inactivation of Prx1 can be achieved by

103,104 31 phosphorylation. However, with levels of GSH being around 2 mM H2O2 would still be reduced before reacting with signaling proteins.2 A thorough assessment of the

relative kinetics with signaling proteins and GSH indicates that GSH competition for

H2O2 is negligible and the H2O2 concentrations required to oxidize signaling proteins (14

to 120 µM) are much higher than the usual sub-micromolar concentrations.30 However it

17

may be possible for such high concentrations105,106 to exist in local compartments during

signaling events.107

The transfer of oxidizing equivalents from highly reactive thiols, such as Prx, to

lower reactive signaling proteins requires complex physical interaction between the two

species and thiol-disulfide exchange. The inherent complexity of such reactions make

them hard to identify and prove. Interestingly, the transcription factor Yap1 is a signaling

protein that appears to be controlled via disulfide redox exchange.30 The reduced form of

Yap1 is capable of forming a mixed disulfide with Gpx3 which then exchanges for an

intramolecular disulfide which can no longer leave the nucleus and increases the

transcription of Prx and catalase.108,109 Interestingly, deletion of all 5 Prx genes and all 3

Gpx genes in yeast yields a viable strain that is no longer able to adjust transcription in

110 response to H2O2. This strongly supports the idea that transfer of oxidative equivalents from professional peroxidases (Gpx, Prx) is dominant, at least in yeast. By this mechanism Prx is oxidized to the sulfenic acid which can then form inter- or intramolecular disulfides and transfer these oxidative equivalents via protein-protein interactions and disulfide exchange to down-stream regulators.99,111

As a continuation of the oxidative equivalent transfer concept, thioredoxins (Trx)

transfer oxidation to signaling proteins (SP) in addition to performing their role as

reductants of Gpx, Prx, and other proteins. Trx itself reacts slowly with H2O2 resulting in

oxidative regulation of Trx being primarily controlled by Prx.31,112 An early example of

Trx reacting with SP was the dimerization of reduced Trx1 or Trx2 with Ask-1 kinase

which is then released upon oxidation of Trx.113 As an example of Trx-Prx interaction pathway in fission yeast, the SP known as Pap1 (transcription factor) is activated by

18

oxidized (disulfide) form of Prx, but only when Trx is depleted.114,115 A benefit of this

Prx-Trx interaction and subsequent interaction with SP is that it requires protein to protein interactions which are highly specific, making them ideal for controlled signaling interactions.

1.4 Cysteine and Sulfenic acids

Sulfenic Acids are characterized by sulfur singly bonded to carbon and to an OH group (RSOH). Sulfenic acids are highly reactive and as a result their existence is either short lived or facilitated by their local environment. In biological systems they form when cysteine is oxidized by reactive oxygen species (ROS) including hydrogen peroxide, hydroperoxides, hypochlorous acids, and peroxynitrite.7,28 Once formed, sulfenic acids can be irreversibly oxidized to sulfinic or sulfonic acid, form a thiosulfinate with another

RSOH, or react with a thiol to form an inter or intramolecular disulfide (Figure 1,

Reactions 1-6).116,117 These reactions trigger innumerable changes in the redox biology of

cells including regulating enzyme catalysis and transcription factors.7 A full

characterization of sulfenic acid chemistry is crucial to understanding the redox

regulation of cells. Identifying their parent protein and under what conditions they form

will lead to a better understanding of how cells respond to oxidative stress.

19

Figure 9. Biological reactions and fates of sulfenic acids. All steps are easily reversible except 2 and 3.

History/Significance

The first organic sulfenic acid, known as Fries Acid, was synthesized in 1912

(Scheme 2),118 but it wasn't until 1955119 that evidence began to emerge supporting the existence of RSOH as biologically significant species. Absent distinguishable spectroscopic features, early evidence for RSOH in proteins was indirect. First, precedence existed for cysteine thiol derivatives in similar oxidation states, specifically sulfenyl iodides (RS-I). 119 Secondly, in 1962 oxidation of cysteine thiols in papain (a cysteine protease) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) was shown to introduce an electrophilic sulfur species believed to be RSOH.120 Not only did this evidence support their existence, it indicated they played significant roles in regulating protein function. These early studies showcased the importance of RSOH in redox

20

biology and provided incentive for study leading toward the current state of scientific understanding.

Scheme 2. Synthesis of Fries acid

Chemical Character

Most reactivity of this sulfur containing can be attributed to sulfur’s 0

oxidation state which allows for both nucleophilic and electrophilic character (Figure 10,

A).116 The nucleophilic character is seen as it reacts with other sulfenic acids to form thiosulfinates (Figure 10, B) and its electrophilic character can be seen in its reaction with thiols to form disulfides (Figure 10, C). Additionally, the sulfur can be further oxidized to its +2 and +4 oxidation states to form sulfinic and sulfonic acids. All of these reactions

are dependent on the microenvironment around the sulfenic acid and proximity to required reactants.28,97,121 While few organic sulfenic acids exist, the ones that do offer

valuable insight into the effects of microenvironment on RSOH reactivity. For example,

Fries Acid is stabilized by hydrogen bonding with an adjacent (Figure 11). 122

Other sulfenic acids such as triptycene SOH use steric hindrance to prevent reactivity at 21 the sulfur via nucleophilic attack (Figure 11).123 While these sulfenic acids give insight into their structure and spectral properties, they are not perfect models of protein-based reactivity. Applying small molecule stabilizing concepts to protein sulfenic acids suggests that RSOH can be stabilized by proximity to groups capable of acting as hydrogen bond acceptors. Additionally, if access to the RSOH is hindered, then it is unlikely that these sulfur species can react to form disulfides or thiosulfinates, or react with potential traps.

Figure 10. Oxidation states & nucleophilic/electrophilic reactions

Figure 11. Stable organic sulfenic acids

22

AhpC – alkyl hydroperoxide reductase C (a peroxiredoxin)

Protein-based sulfenic acid would present the ideal model for understanding SOH

biological reactivity. Upon oxidation to the sulfenic acid most protein-based cysteines

quickly react to form disulfide bonds.124–126 The cysteine that is oxidized by hydrogen

peroxide is called the peroxidatic cysteine (CP) and the cysteine that attacks to form the

disulfide is called the resolving cysteine (CR) (Figure 12). Interestingly, there is a protein-

based sulfenic acid that has been stabilized by mutation of its resolving cysteine. In AhpC,

it was found that Cys46 and Cys165 quickly formed disulfide bonds. Mutation of each of

these residues separately, to generate three separate mutants with serine replacing

cysteine 46 (AhpC46S), cysteine 165 (AhpC165S), or both cysteine 46 and 165

(AhpC46S, 165S) showed no gross changes to tertiary structure. However, it was shown

that the two AhpC46S and AhpC46S, 165S mutants showed no peroxidatic activity.

While the 165 (AhpC165S) mutant still reacted with peroxide, indicating the remaining

124 C46 was the CP which is capable of reacting with hydrogen peroxide. These 165

mutants form stable sulfenic acids as verified and quantified by mass spectrum analysis,

reaction with traps, or UV-vis after chemical modification.127 AhpC stable sulfenic acid

acts as ideal model systems for assessing the reactivity of SOH traps used to probe the workings of redox biology.

23

Figure 12. AhpC with resolving and peroxidatic cysteine thiols

Chemical Methods for Protein Sulfenic Acid Detection

Due to the inherently complex environments in which most sulfenic acids exist,

and their propensity for fast modification to other functional groups, analysis of sulfenic

acids by normal means (x-ray, NMR, IR, UV-Vis) proves quite challenging.6 As a result, much of their reactivity has been defined by indirect chemical means. A number of traps have been developed that exploit the nucleophilic or electrophilic character of sulfenic acids. Of particular interest is 5,5-dimethyl-1,3-cyclohexanedione or dimedone, which reacts selectively through nucleophilic attack on the sulfur (Scheme 3) forming a covalent bond as originally identified by Allison in 1974.128–131 In aqueous buffers at neutral pH,

dimedone is selective for sulfenic acids in that it doesn't react with other electrophilic or protein-based functional groups such as , s-nitrosothiols, methionine sulfoxide, or .7,94,129,132 Due to the particular usefulness of dimedone and dimedone-like 1,3

diketone based molecules, it has spawned a number of fluorescent and biotin modified

probes.

24

Scheme 3. The 1,3 diketone known as dimedone reacts with sulfenic acids via a

nucleophilic mechanism to form covalent bonds.

The first 1,3 diketone based probes were reported by Poole et al. in 2005.132 As a proof of concept, these probes demonstrated that attaching fluorophores to 1,3 diketone functionality could enable selective and covalent trapping and quantification of sulfenic acids in complex media (Figure 13, A). A few years later, the 1,3 diketone functionality was elaborated to fluorescein, rhodamine, biotin (Figure 13, B), and clickable azide probes.133 The new fluorescent tags enabled imaging at much lower concentrations, and

had readily identifiable MS-MS fragments for positive identification of labeling. These early probes demonstrated the utility of 1,3 dicarbonyl molecules for covalently modifying sulfenic acids and their parent proteins.

25

Figure 13. Early dicarbonyl probes for sulfenic acids. Conjugation to fluorophores for fluorescent analysis, and to biotin for affinity based isolation.

Now, a number of modified 1,3-dicarbonyl molecules are commercially available.

The highly useful DCP-Bio1 contains a 1,3-dicarbonyl warhead that attaches to protein sulfenic acids, and is linked by an to biotin which can facilitate affinity isolations from complex media.7,94,133 The cyclopentanedione and biotin based probe known as BP1 acts similarly to dimedone and allows for streptavidin affinity isolations, but has increasing reactivity at lower pH (Figure 14).134 The fluorescent DCP-Rho1 is cell permeable and selectively modifies sulfenic acids for identification by fluorescence

(Figure 14).133,135 The alkyne β-ketoester (Alk-β-KE) from the Furdui lab allows for attachment to the β-ketoester to a variety of groups via click chemistry, making for a highly customizable probe.136 The 1,3-diketone attached to azide functionality known as

DCP-N3 is highly customizable and easily attached to alkynes via click chemistry (Figure

14).134 The alkyne counterpart is known as DCP-Alkyne which can also be attached to

26

106 useful azides via click chemistry. Or the heavy version, known as Dyn-2-d6 allows for

quantitative analysis via MS when paired with the light version.137

Commercialy Available Probes

Et2N

DCP-Rho1 O O + S H O NEt2 H O N NH OOH Cl- S N H HN O O O N BP1 O

OOH OOH O OH O O D D N O 3 D D D D Alk- -KE DCP-N3 DCP-Alkyne DYn-2-d6

Figure 14. Commercially available 1,3 dicarbonyl probes. A wide range of 1,3

dicarbonyl probes are now commercially available at KeraFAST138 for identifying,

isolating, and quantifying sulfenic acids in complex media.

In an effort to further exploit the electrophilic nature of sulfenic acids, Carroll et

al. performed a survey of existing C-reactive nucleophiles and their reactivity towards

sulfenic acids.139,140 They discovered that the linear forms of di-carbonyl probes were

more reactive than the cyclic forms (Figure 15). Interestingly, this trend followed the

same parallel to the keto-enol equilibrium, with cyclic species existing more as the enol.

This observation appeared to suggest that enol stabilization of the C-centered anion resulted in decreased reactivity towards sulfenic acids. Also, when the pKa of the reactive center was increased, the reactivity decreased. To prevent enol stabilization but still maintain a low pKa, they removed the carbonyl and replaced it with electron

27

withdrawing groups such as , nitro, or cyano, resulting in an increase of

reactivity.139,140 This eventually resulted in complete removal of carbonyls to yield C- nucleophiles that are more than 150x faster than dimedone.139 Selectivity studies

performed with Fmoc and CBz protected amino acids, cysteine (thiol), serine (alcohol), lysine (), cysteine (disulfide), and sulfinic acid (BnSO2Na) in pH 7.4 buffer,

indicated no cross-reactivity. However, these nucleophiles have not been as well vetted

as the traditional dicarbonyl probes and may present unforeseen problems.

28

Figure 15. Properties of nucleophilic probes. The nucleophilic 1,3 dicarbonyls exist in a keto to enol equilibrium. More linear structures have more keto structure and are more nucleophilic. Removing carbonyls decreases the keto to enol resonance even further and destabilizes the anion site. Probes with destabilized anions have increased reactivity with sulfenic acids.

As shown in Figure 10, sulfenic acids have both nucleophilic and electrophilic character. While their electrophilic character has been well explored, much less work has

29

been done with their nucleophilic character. Probes exploiting the nucleophilic character

of sulfenic acids would likely have cross-reactivity with other nucleophiles. Nevertheless,

the small molecule 4-chloro-7-nitrobenzofurazan (NBD-Cl) has been used to exploit the

nucleophilic character of RSOH and probe its function.127 This reagent is traditionally

used to modify thiol, amino, and hydroxyl groups. However, at neutral pH in the

oxidized C165S mutant of AhpC (an alkyl hydroperoxide reductase), the probe only

modified cysteinyl residues.127 The reaction product of NBD-Cl and RSOH is a sulfoxide

which can be detected via its absorbance maxima at 347 nm and distinguished from the

thiol product which absorbs at 420 nm. Additionally, ESIMS was used to verify the

addition of NBD to the Cys46-SOH of AhpC. 127 This supported the idea of RSOH

existing as an intermediate in the peroxidatic mechanism of AhpC.127,141,142 However, the mechanism of nucleophilic attack on NBD-Cl was unknown at this point and could have been via the sulfur to form a sulfoxide or via the sulfenate to form an ester. 127 As expected this probe has limited applications due to its tendency to react with other biological nucleophiles.

30

BH H O 347 nm S Ahp-C O S Ahp-C Cl N N O O N N NO NO2 2 sulfoxide (NBD-Cl)

Ahp-C 420 nm Cl S S N N Ahp-C H O O N N

NO2 NO2 (NBD-Cl) thioether

Figure 16. The electrophilic probe NBD-Cl. NBD-Cl reacts with both sulfenic acids and thiols. The difference in UV-Vis absorbance can be used to quantify the relative

SOH/SH ratios.

While the nucleophilic and electrophilic character of sulfenic acids encompass the dominant reactivity in biological settings, their propensity to perform pericyclic reactions is also relevant in probe development. Interestingly, one method for generating sulfenic acids involves the elimination of sulfoxides to form olefins (Figure 17, A) which proceed through a pericyclic transition state.143 Additionally, alkynes have been demonstrated to react with sulfenic acids such as triptycene sulfenic acid (Figure 17, B).144 However, the

addition of sulfenic acids to most alkynes requires heating and is too slow to be of use in

biological trapping (Figure 17, C & D).145 While Carroll et al. in 2014 reported that the

reaction of alkynes with sulfenic acids “are not facile in biological systems,” 117 it was

31 soon to be demonstrated otherwise. Reactions with alkynes are known to be accelerated by the introduction of strain, as is the case with copper free click reactions.146 Applying these concepts to the strain promoted trapping of sulfenic acids by alkynes yielded a trap that is more than 100x faster than traditional 1,3-diketone based probes and the intricacies of alkyne based traps will be thoroughly explored in chapters 2 and 3.147

Figure 17. Examples of sulfenic acid reacting with pi bonds. A. Thermal sulfoxide elimination to form sulfenic acid and . B. Triptycene sulfenic acid addition to electron poor alkyne with heating. C & D. Cysteine sulfenic acid addition to alkyne with heating.

1.5 Synopsis

This chapter has summarized mechanisms and importance of sulfur based redox biology in cellular signaling and the development of disease states. The control of ROS 32 begins with the generation of superoxide in the mitochondria and its dismutation to more manageable species such as hydrogen peroxide. The flow and signaling of hydrogen peroxide is managed by peroxiredoxins and associated sulfur species as it moves outwards from the mitochondria. Movement can be either natural diffusion or controlled efflux. Along the way, sulfur species chaperone the signaling and generation of ROS.

Cysteine thiols are oxidized to sulfenic acids, which then can form disulfides, or be oxidized further to sulfinic or sulfonic acids. The parent proteins of these sulfur species can act as recycling species such as peroxiredoxins or thioredoxins, or function as signaling molecules controlling enzymes and modulating the transcription of genes. The outcome is largely controlled by kinetic factors associated with the parent proteins, and the specificity of protein-protein interactions allowing targeted transfer of oxidative equivalents and signaling. Unfortunately the study of such processes is limited by the fleeting intermediacy of some sulfur species such as sulfenic acids, and the complexities of their cellular enviornments. The aim of this dissertation is to help delineate the importance of sulfenic acids and their parent proteins through the development of fast and biorthogonal traps.

33

CHAPTER 2

STRAINED CYCLOALKYNES AS NEW PROTEIN SULFENIC ACID TRAPS

Thomas H. Poole,=,† Julie A. Reisz,=,‡ Weiling Zhao,‡ Leslie B. Poole,§,∥ Cristina M.

Furdui,‡ ,∥ and S. Bruce King*,†,∥

†Department of Chemistry, Wake Forest University, Winston-Salem, North Carolina,

‡Department of Internal Medicine, Wake Forest School of Medicine, Winston-Salem,

North Carolina, §Department of Biochemistry, Wake Forest School of Medicine,

Winston-Salem, North Carolina, ∥Center for Molecular Communication and Signaling,

Wake Forest University, Winston-Salem, North Carolina, USA

The work of chapter 2 was published in the Journal of The American Chemical Society in

2014 and entails the development of novel sulfenic acid traps.147 The manuscript, including figures and schemes, was drafted by Thomas Poole and Julie A. Reisz. Editing was performed by Leslie B. Poole, Cristina, M. Furdui, and S. Bruce King. Since publication, changes in both format and contents were made to adapt this work for the dissertation format. The organic chemistry research including kinetics and synthesis were performed by Thomas Poole, while the biological & enzymatic work including protein kinetics, cell viability, cell permeability, cell labeling, and relative lysate reactivity (BCN-Bio vs dimedone) were performed by Julie A. Reisz.

34

ABSTRACT: Protein sulfenic acids are formed by the reaction of biologically relevant

reactive oxygen species with protein thiols. Sulfenic acid formation modulates the

function of enzymes and transcription factors either directly or through the subsequent

formation of protein disulfide bonds. Identifying the site, timing, and conditions of

protein sulfenic acid formation remains crucial to understanding cellular redox regulation.

Current methods for trapping and analyzing sulfenic acids involve the use of dimedone

and other nucleophilic 1, 3-dicarbonyl probes that form covalent adducts with cysteine-

derived protein sulfenic acids. As a mechanistic alternative, the present study describes

highly strained bicyclo[6.1.0]nonyne (BCN) derivatives as concerted traps of sulfenic

acids. These strained cycloalkynes react efficiently with sulfenic acids in proteins and

small molecules yielding stable alkenyl sulfoxide products at rates more than 100x

greater than 1, 3-dicarbonyl reagents enabling kinetic competition with physiological

sulfur chemistry (Scheme 4). Similar to the 1, 3-dicarbonyl reagents, the BCN

compounds distinguish the sulfenic acid oxoform from the thiol, disulfide, sulfinic acid and S-nitrosated forms of cysteine while displaying an acceptable cell toxicity profile.

The enhanced rates demonstrated by these strained alkynes identify them as new

bioorthogonal probes that should facilitate the discovery of previously unknown sulfenic

acid sites and their parent proteins.

Scheme 4. The reaction of AhpC with BCN-Bio

35

2.1 Introduction

Protein sulfenic acids (RSOH) arise from the reaction of reactive oxygen species, such as

hydrogen peroxide, alkyl peroxides and peroxynitrite, with cysteine thiols.148,149 Sulfenic acids may further oxidize to sulfinic or sulfonic acids, condense with other sulfenic acids to form thiosulfinates, or react with thiols to yield disulfides.116 This wide variety of sulfur-based chemistry marks protein sulfenic acids as the initial product toward

potentially irreversible oxidative damage. This versatile chemistry also allows their participation in the reversible control and modulation of important cellular processes, such as transcription and enzymatic cascade pathways, which directly influence biological outcomes.150,151 Most information regarding the biological roles of sulfenic

acids comes from studies using nucleophilic 1, 3-dicarbonyl-based probes, such as

dimedone (Scheme 5). The high reactivity of sulfenic acids limits their cellular lifetime,

permitting these probes to access and label only a fraction of existing sulfenic acid sites,

which ultimately constrains the understanding of the biological roles of these species.117

Sulfenic acids also react with and alkynes via a concerted mechanism to give alkyl and alkenyl sulfoxides.129,143 The introduction of strain energy, as shown with

alkyne-azide “click” chemistry, should increase the reactivity and rate of this mechanistic

alternative to current sulfenic acid trapping methods (Scheme 5).152,153 Specifically, we report the bicyclo[6.1.0]nonyne (BCN) and its biotinylated derivative (BCN-Bio) efficiently and selectively trap protein sulfenic acids at superior rates to 1, 3-dicarbonyl based probes, identifying a new group of bioorthogonal protein sulfenic acid probes

(Scheme 5).154

36

Scheme 5. The trapping of sulfenic acids by dimedone and 9-hydroxymethyl-

bicyclo[6.1.0]nonyne (BCN).

2.2 Evaluating the reactivity and selectivity of strained alkynes for sulfenic acids

The known bicyclo[6.1.0]nonyne (BCN) rapidly reacts with small molecule

sulfenic acids. Treatment of a thermally generated organic-soluble cysteine-derived sulfenic acid with BCN gives a mixture of diastereomeric alkenyl sulfoxides as determined by NMR spectroscopy and liquid chromatography-mass spectrometry in 84% isolated yield (Scheme 6).145 Similar reaction of BCN with Fries acid, a stable

anthraquinone-derived sulfenic acid, also produces the diastereomeric alkenyl sulfoxide

in 99% yield (Scheme 6).122 The results clearly show for the first time that this strained

cycloalkyne reacts with small molecule model sulfenic acids in organic systems. These

reactions likely proceed via a concerted cycloaddition-like mechanism quite distinct from

the accepted nucleophilic addition of the 1, 3-dicarbonyl probes to the sulfenic

acids.129,143,155The alkenyl sulfoxide product did not react with the nucleophilic reducing

agents dithiothreitol (DTT) and tris(2-carboxyethyl)phosphine (TCEP) demonstrating its stability.

37

Scheme 6. Reactions of BCN with a cysteine-derived sulfenic acid and Fries acid yield

alkenyl sulfoxides.

Monitoring the decrease in absorbance at 453 nm by ultraviolet-visible (UV/vis)

spectroscopy as a function of time provides kinetic information for the reaction of BCN

and Fries acid (Figure 18). A kinetic analysis of these reactions, performed under pseudo-

first-order conditions in organic solvent, gives a second-order rate constant of ca. 25 M-

1s-1 (Figure 18) and similar experiments in a 50:50 mixture of acetonitrile:ammonium

bicarbonate buffer give a second order rate constant of ca. 12 M-1s-1 (Figure 19). These rate constants are significantly greater than the reported value of 0.05 M-1s-1 for

dimedone-based probes with model protein sulfenic acids and suggest strained alkynes

may act as useful protein sulfenic acid traps.7 Similar experiments with trans-cyclooctene

38

(tCOT), a strained cyclic alkene,156,157 give a second-order rate constant of ca. 0.01 M-1s-1 and produce the corresponding alkyl sulfoxides with both the cysteine-derived sulfenic acid and Fries acid (Figure 20). Given the kinetic differences between BCN and tCOT, further biological trapping studies focused on BCN and its derivatives. Experimental limitations including the limited aqueous solubility of the cysteine-derived sulfenic acid and the low reactivity of Fries acid, prevent a direct kinetic comparison of BCN with dimedone in an aqueous environment (Figure 21).The enhanced rate of reaction provides evidence that the ring strain of BCN accelerates the reaction. This rate increase likely arises from bending of the normally 180° oriented bonds formed from the overlap of the sp and sp3 hybridized orbitals into a conformation that approaches the geometry of the transition state.

39

Figure 18. UV-Vis Kinetics; Fries Acid + BCCN in acetonitrile. The kinetic tests were performed at 3mM or 1.5mM of BCN, and 0.3mM Fries acid. The disappearance of

Fries Acid was monitored at 453 nm under pseudo first order kinetic conditions.

SigmaPlot was utilized to plot data using exponential decay with the equation: f =

(-bx) y0+a(e ).

40

Figure 19. UV-Vis Kinetics; Fries Acid + BCN in 50:50 MeCN:H2O (pH 7.5). The kinetics tests were performed at 3mM or 1.5mM of BCN, and 0.3mM Fries acid. The disappearance of Fries Acid was monitored at 453 nm under pseudo first order kinetic conditions. SigmaPlot was utilized to plot data using exponential decay with the

(-bx) equation: f = y0+a(e ).

41

Figure 20. UV-Vis Kinetics; Fries Acid + trans-cyclooctene (tCOT) in acetonitrile. An excess of tCOT was used to allow for pseudo 1sst order conditions. The disappearance of

Fries Acid was monitored at 453 nm under pseudo first order kinetic conditions.

SigmaPlot was utilized to plot data using exponential decay with the equation: f =

(-bx) y0+a(e ).

Reactivity of Fries Acid with Dimedone

Monitoring the reaction of dimedone with Fries Acid under pseudo first order kinetic conditions would be an ideal method for directly comparing the rate of the BCN probes to that of dimedone. However, it appears that dimedone and Fries acid react

42

through acid-base chemistry. In a paper by Benkovic et al., the authors combine

dimedone with Fries Acid anion and claim sulfenic acid trapping as monitored by UV-vis

spectroscopy.159 However, dimedone has a pKa of 5.15 and Fries Acid has a pKa of ca.

7.1.132 These constants suggest that mixture of dimedone with Fries Acid anion should

produce the enolate of dimedone and protonated Fries Acid through an acid base reaction.

Indeed, UV-vis spectroscopic experiments support this hypothesis as combination of

Fries Acid and NaOH in equimolar concentrations gives a UV-Vis spectrum identical to that reported by Benkovic and consistent with the reported spectrum of the Fries Acid anion (Figure 21, red line).159 Addition of one equivalent of dimedone gives an

absorbance spectrum identical to that of protonated Fries Acid (~460 nm peak, blue line).

Subsequent addition of one equivalent of NaOH regenerates the Fries Acid anion

spectrum (675 nm peak, green line) characteristic of an acid base reaction between

dimedone and the anion of Fries Acid and not adduct formation (Figure 21).

1.2 Fries Acid (0.3 mM) + Dimedone Reversibility 1

0.8 0.3 mM NaOH

0.6 0.3 mM NaOH + 0.3 mM Dimedone

Absorbance 0.4 0.3 mM NaOH + 0.3 mM 0.2 Dimedone, + 0.3 mM NaOH 0 350 400 450 500 550 600 650 700 Wavelength (nm)

43

Figure 21. Fries Acid and dimedone in 15% MeCN:85% H2O. Fries acid (3 mM, 0.1 mL,

MeCN) and NaOH (3 mM, 0.1 mL, H2O) were added to a mixture of MeCN and water

(0.05 mL MeCN, 0.75 mL H2O), to give equimolar (0.3 mM) concentrations of Fries

Acid and NaOH (15% MeCN:85% H2O) resulting in a characteristic peak at 675 nm

corresponding to the Fries acid anion.159 Addition of 1 equivalent of dimedone (3mM,

0.1mL, MeCN) results in an immediate decrease of the anion peak and a simultaneous

increase of an absorbance peak at ~460 nm, which is identical to the absorbance of Fries

Acid (protonated). Addition of another equivalent of NaOH (3mM, 0.1mL, H2O) results in the immediate reformation of the Fries acid anion peak at 675 nm.

Reactivity of Fries Acid with Dimedone and AcOH

For comparison, solutions of Fries acid with 10 eq. of AcOH, and Fries acid with

10 eq. dimedone were prepared (15% MeCN:85% H2O, Figure 22). The combination of

Fries Acid with AcOH shows the absorbance of protonated Fries Acid (purple line), Fries

Acid with dimedone gives a similar absorbance (light blue line, Figure 22). Additionally,

the incubation of Fries Acid anion with dimedone (10 mM each) for 2 h provided a bright

orange-red solution with small amounts of an orange precipitate. Analysis of this solution

by ESI-MS and NMR showed no sign of adduct formation. The precipitate is believed to

be the thiosulfinate formed by the condensation of two sulfenic acids. Accurate identification of this precipitate was prevented by its insolubility in common solvents.

Taken together, the high similarity between the UV-vis spectra, facile reversibility, and

the lack of NMR and MS evidence of adduct formation strongly suggest that the

44 dominant reaction between Fries acid and dimedone is an acid/base reaction, not adduct formation as reported by Benkovic et al.159

1.4 Fries Acid (0.3 mM) + Dimedone or AcOH 1.2

1

0.8

0.6 3mM AcOH

Absorbance 0.4 3mM Dimedone

0.2

0 350 400 450 500 550 600 650 700 Wavelength (nm)

Figure 22. Fries Acid + Dimedone or AcOH have nearly identical UV-Vis spectrum.

Fries acid (0.3 mM) and AcOH (3.0 mM, 15% MeCN:85% H2O) produced the characteristic absorbance of protonated Fries Acid (purple line). Fries acid (0.3 mM) and dimedone (3.0 mM, 15% MeCN:85% H2O) produced an absorbance spectrum (light blue line) similar to that of protonated Fries Acid.

2.3 Evaluating BCN and BCN-Bio in proteins

The stabilized sulfenic acid of the C165A mutant of the alkyl hydroperoxidase

AhpC protein (AhpC-SOH) provides an opportunity to measure the reaction efficiency of strained cyclooctynes with a protein sulfenic acid. This cysteine-based bacterial peroxidase forms an inter-subunit disulfide bond through oxidation of C46 to the sulfenic

45

acid followed by condensation with C165 of an adjacent AhpC monomer. Mutation of

C165 to alanine or serine stabilizes the reactive SOH at C46, allowing for the evaluation

of chemical probe reactions with the otherwise transient SOH. Conjugation of BCN to

biotin by a DCC coupling forms the biotin ester (BCN-Bio). Monitoring the reaction of

BCN-Bio and C165A AhpC-SOH (20,600 amu) by electrospray ionization time-of-flight mass spectrometry (ESI-TOF MS) yields a peak at 20,976 amu corresponding to the formation of the expected alkenyl sulfoxide adduct (Figure 23A). Unreacted AhpC-SOH is observed in the gas phase as a mixture of sulfenic acid and (S-N condensation product, 20,582 amu) as previously described.134 Incubation of BCN-Bio

with various AhpC oxoforms including AhpC-SO2H, AhpC-SNO and AhpC-S-S-Cys

fails to yield adducts as judged by ESI-TOF MS revealing the protein sulfenic acid

selectivity of BCN-Bio (Figure 24 and Figure 25). The lack of an adduct of BCN-Bio

with C165A AhpC-SH (thiol) indicates that thiol-yne, reactions, potentially complicating side reactions, do not occur in this system (Figure 26).160 Lack of significant cross- reactivity of BCN-Bio with these oxoforms, particularly AhpC-SNO, that could possibly undergo radical addition or cycloaddition with BCN-Bio,161,162 was further confirmed by

Western blotting under non-reducing conditions (Figure 27).

46

Figure 23. The reactions of C165A AhpC-SOH (40 M) with BCN or BCN-Bio (100

M) at r.t. in 50 mM NH4HCO3, pH 7.5. A. ESII-TOF mass spectrum of the AhpC-SOH reaction with BCN-Bio at 30 min. B. Time course of adduct fformation. The concentration of AhpC-SOH adduct with BCN or BCN-Bio along the reaction time course was determined based on relative abundance of adducct among thhe total ion abundances of the prominent species in the ESI-TOF mass spectra.

47

Figure 24. Treatment of C165A AhpC-SO2H with BCN-Bio. A: Untreated sample of

C165A AhpC-SO2H (40 M, 20616 amu) at pH 7.5. B-C: Incubation of AhpC-SO2H (40

M) with BCN-Bio (100 M) at r.t. and pH 7.5. B: 10 min, C: 60 min. A small amount of AhpC-SOH/alkyne adduct is observed at 20976.5 amu, likely resulting from the minor

SN and SOH species observed in Panel C prior to treatment.

48

Figure 25. Treatment of C165A AhpC-SNO and C165A AhpC-S-S-Cys with BCN-Bio.

A: Formation of C165A AhpC-SNO (20612.7 amu) and C165A AhpC-S-S-Cys (20702.9 amu). B: Incubation of AhpC-SNO and AhpC-S--S-Cys (40 M total protein) with BCN-

Bio (100 M) for 80 min at r.t. (in dark) and pH 7.5. C: Unttreated sample of AhpC-SNO and AhpC-S-S-Cys (40 M total protein) at 80 min at r.t. (in dark) and pH 7.5.

49

Figure 26. Treatment of C165A AhpC-SH with BCN-Bio. A: Untreated sample of

C165A AhpC-SH (40 M, 20584.6 amu) at pH 77.5. B: Incubation of AhpC-SH (40 M) with BCN-Bio (100 M) and TCEP (5 mM) for 60 min at r.t. and pH 7.5.

50

Figure 27. Western blot analysis in non-reducing conditions of C165A AhpC cysteine oxoforms. AhpC cysteine oxoforms (40 M) incubated with BCN-Bio (100 M) for 1 h at r.t. and pH 7.5.

Efficient trapping of AhpC-SOH occurs at a much lower concentration of BCN-

Bio (100 M) than reported concentrations of various 1, 3-dicarbonyl-based probes

(often 1-5 mM).7,117,163 A MS-based kinetic analysis of protein sulfoxide formation from the reactions of C165A AhpC-SOH (40 M) with both BCN and BCN-Bio (100 M) reveals the time-dependent increase in adduct formation and gives second order rate constants of 13.3 and 16.7 M-1s-1, respectively, values several hundred-fold greater than reactions with 1, 3-diketone-based SOH probes (Figure 23B).7 The increased reaction rates potentially allow BCN and BCN-Bio to trap more reactive, transient protein sulfenic acids. Experiments with higher concentrations of BCN-Bio (1-5 mM) result in protein labeling accompanied by cycloalkyne polymerization as judged by ESI-TOF MS (Figure

28). Trypsin digestion followed by tandem MS analysis verifies the addition of BCN to

C46 with an XCorr of 5.9 for a 4+ charged peptide providing a high confidence site assignment and adduct mass (Figure 29). Analogous MS2 results were obtained by labeling AhpC-SOH with BCN-Bio followed by trypsin digestion (Figure 30).

Comparatively, adducts of these cyclooctynes and AhpC-SOH produce much higher quality MS2 spectra compared to dimedone.136

51

Figure 28. Polymeric addition to AhpC-SOH with high concentration of BCN-Bio. ESI-

TOF MS analysis of the reaction of C165A AhpC-SOH (40 M) and BCN-Bio (5 mM) at

15 min (r.t. in 50 mM NH4HCO3, pH 7.5).

52

Figure 29. Positive ion MS2 (CID) spectrum of the Cys-containing peeptide of C165A

AhpC and BCN. The spectrum results from the reaction of AhpC-SOH (40 M) with

BCN (100 M) at pH 7.5. Protein species were digested with trypsin and the resulting peptides separated by nanoLC.

53

Figure 30. Positive ion MS2 (CID) spectrum of the Cys-containing peeptide of C165A

AhpC and BCN-Bio. The spectrum results from the reaction of AhpC-SOH (40 M) with

BCN-Bio (100 M) at pH 7.5. Protein species were digested with trypsin and the resulting peptides separated by nanoLC.

The chemical stability of the AhpC alkenyl sulfoxide adducts of BCN-Bio was investigated using H2O2 and several biochemical reductants. Of particular interest were

DTT and TCEP due to their widespread use in MS-based proteomics workflows and common endogenous Michael donors like glutathione and cysteine. Since initial experiments did not suggest H2O2-mediated oxidation of the alkenyl sulfoxide to the sulfone, preparation of the AhpC-S(O)- BCN-Bio adduct (20,977 amu) in this experiment included a final step of H2O2 treatment to consume protein species (e.g., SH, SN, SOH) that would interfere with reductant chemistry. A small amount of SN remained.

54

Following removal of unreacted BCN-Bio and H2O2, AhpC tagged with BCN-Bio (40

M total protein) was treated with 1 mM DTT, TCEP,-mercaptoethanol, GSH, or N- acetyl cysteine for 1 h at r.t. (Figure 31). Formation of low amounts of thiol (DTT, TCEP) and the corresponding mixed disulfides of beta-mercaptoethanol (-ME), glutathione

(GSH), and N-acetylcysteine (NAC) were observed. Since there was no significant decrease in the AhpC-SOH adduct with BCN-Bio (Figure 31) and Western blot analysis

(Figure 32), we conclude these products form from the unreacted SN species.

Additionally, Michael adducts of AhpC labeled with BCN-Bio were not observed under these conditions indicating the overall stability of the protein adducts of BCN-Bio.

55

Figure 31. Stability of the C165A AhpC-S(O)-BCN-Bio adduct (40 M) toward biochemical reductants (1 mM) as judged by ESI-TOF MS. Reactions were performed in

50 mM NH4HCO3 (pH 7.5) for 1 h at r.t. A: Untreated control of C165A AhpC-S(O)-

BCN-Bio generated via the reaction of C165A AhpC-SOH (40 M) with 4 (250 M) (4

o h, pH 7.5, r.t.) followed by treatment with H2O2 (750 M) overnight (4 C). H2O2 and

BCN-Bio were removed, and the protein species were treatted with B: DTT (1 mM). C:

TCEP (1 mM). D: -mercaptoethanol (1 mM). E: reduced glutathione (1 mM). F: 1 mM

N-acetyl cysteine. The species at 20,616 amu is C165A AhpC-SO2H, and the species marked with an asterisk (*) at 20,632.9 amu is C165A AhpC-SO3H. Percent abundance values for the C165A AhpC-S(O)-BCN-Bio adduct were determined using the ion abundance relative to the sum of ion abundance values for all relevant protein species observed in each mass spectrum.

56

Figure 32. Western blot analysis of AhpC-S(O)-BCN-Bio adduct stability in the pressence of reductants: DTT, TCEP, -mercaptoethanol, glutathione, aand N-acetyl cysteine (1 mM each, 1 h at r.t.). Reductants were not removed before SDS-PAGE and Laemmli buffer without additional reductant was used to prepare samples.

Given the efficient reactions of BCN annd BCN-Bio with AhpC-SOH, we tested the ability of BCN-Bio to trap protein sulfenic acids in cell lysates. Oxidation with increasing amounts of H2O2 in the presence of BCN-Bio shows concentration-dependent

SOH trapping (Figure 33A). Reaction efficiency increases as the concentration of H2O2 increases from 2.5 M to 5 M and then decreases as the concentration of H2O2 is raised further, likely due to the over-oxidation of cysteine residues to sulfinic and sulfonic acids.

Comparative analysis with iodoacetamido-biotin (biotin-IAM)-treated lysate sets the level of SOH and labeling with BCN-Bio at an upper limit of 25% of total thiol content

(Figure 33B).

57

Figure 33. Reactivity of BCN-Bio with oxidized cell lysates. A: Reduced protein lysates were spiked with C165A AhpC for loading control and aliquotted into 60 g fractions. Fractions were treated with BCN-Bio (100 M) and increasing concentrations of H2O2. A control sample of lysate (lane 2) was supplemented with TCEP then treated with BCN-Bio (100 M). B: Reduced protein lysate (60 g fraction) was treated with iodoacetomido-biotin (100 M) to label all thiol content and compared to the sample in

Panel A treated with 5 M H2O2 and 100 M of BCN-Bio.

The utility of BCN probes in labeling endogenous protein SOH was tested by lysing squamous cell carcinoma cells in a modified radioimmunoprecipitation (RIPA) buuffer containing BCN-Bio (100 M) for 30 min, providing a robust level of biotinylated proteins (Figure 34). In comparison, cells lysed in buffer containing TCEP (10 mM) for

30 min then incubated with BCN-Bio (100 M) for the same duration instead demonstrated a sharply decreased amount of protein biotinylation, demonstrating the selectivity of BCN for sulfenic acids. The small amount of BCN labeling in TCEP-treated 58 lysates may result from incomplete reduction of SOH, particularly in buried protein microenvironments. Our studies involving the incubation of BCN-Bio with cell lysates under highly stringent reducing conditions followed by Western blot analysis show no evidence of cross-reactivity with available protein thiols or other amino acids, including their post-translational modifications (Figure 34).

Figure 34. Labeling of endogenous protein SOH using BCN-Bio and TCEP. Where indicated, SCC-61 cells were lysed in mRIPA buffer containing TCEP (10 mM) fofor 30 min on ice, then treated with BCN-Bio (0.1 mM) for an additional 30 min on ice (n = 3).

For the TCEP-untreated samples, cells were lysed in mRIPA buffer conntaining BCN-Bio

(0.1 mM) for 30 min on ice (n = 3). DMSO was present at 0.2% (v/v) in all samples.

Lysates were clarified by centrifugation at 4oC, then reactions quenched by the removal of unreacted TCEP and BCN-Bio.

59

Figure 35. Viability of SCC-61 cells in the presence of BCN-Bio and dimedone as measured by the MTT assay.

With the intent to use BCN or BCN-Bio to label protein sulfenic acids in live cells, the cytotoxicity profile of BCN was assessed in comparison to dimedone using a human squamous cell carcinoma cell line and the 3-(4, 5-dimethylthiazolyl-2)-2,5- diphenyltetrazolium bromide) (MTT) cell viabillity assay. The addition of BCN-Bio to these cells results in death with an IC50 of 199.3 ± 27.3 µM at 48 h (Figure 35). In comparison, dimedone induces cell death with an IC50 of 1.46 ± 0.12 mM at 72 h (Figure

35). While more toxic than dimedone, the high reactivity of BCN-Bio allows for use at concentrations below 100 M and shorter incubation times that limit its toxicity.

Cell membrane permeability of the bicyclononynes allong with the efficiency of in vivo SOH labeling was tested using BCN-Bio in comparison to the biotinylated dimedone-based DCP-Bio1. SCC-61 cells were incubated with media containing DCP-

Bio1 (0.025 - 1 mM) or BCN-Bio (0.025 - 0.1 mM) for 30 min, washed twice with cold

PBS, then lysed. Reducing SDS-PAGE and Western blotting demonstrate the membrane

60

permeability of both probes and increased labeling with BCN-Bio relative to DCP-Bio1

(Figure 36A). Similarly, the efficiency of SOH labeling in a recombinant protein was

explored using C165A AhpC-SOH (20 M) in the presence of BCN (20 M) and

dimedone (5 mM). Quenching of the competition reaction at 15 min using size exclusion resin reveals a much higher abundance of the AhpC-SOH adduct with BCN (adduct observed at 20,750.6 amu) compared to dimedone (20,722.6 amu) (Figure 36B).

61

Figure 36. Labeling of RSOH in live cells and purified AhpC, vs DCP-Bio1. A: Labeling of protein sulfenic acids in live cells using DCP-Bio1 (1 mM; 0.1 mM; 0.025 mM) and

BCN-Bio (0.1 mM; 0.05 mM; 0.025 mM). SCC-61 cells were incubated for 30 min

o (37 C, 5% CO2) in the presence of each probe (final concentration of DMSO was 0.2%), washed twice with cold PBS, then lysed with mRIPA buffer in the absence of probe. B:

ESI-TOF mass spectra for the labeling of C165A AhpC-SOH (20 M, top panel) with 1

(20 M) and dimedone (5 mM) at 15 min in 25 mM Tris pH 7.0 (r.t., bottom panel).

2.4 Conclusion

In summary, strained cyclic alkynes (BCN and BCN-Bio) rapidly react with sulfenic acids to yield alkenyl sulfoxide adducts. Using a model protein sulfenic acid, these

62

compounds yield stable products with clear ionization states that facilitate identification and MS analysis. By reacting through concerted pathways, these traps provide a distinct mechanistic alternative to the nucleophilic 1, 3-dicarbonyl compounds. Kinetic analysis with purified protein, lysates and live cells reveals reaction rates exceeding those of dimedone and dimedone based-probes by more than two orders of magnitude (16.7 M-1s-1 vs 0.05 M-1s-1, 334x)7 making these one of the fastest characterized sulfenic acid traps

described to date.147 Carroll’s recent work describes new C-nucleophiles that react with a

model sulfenamide 215x faster than dimedone.139 Cycloalkynes also demonstrate

bioorthogonal reactivity with the sulfenic acid by not reacting with the thiol, disulfide,

sulfinic acid, or S- oxoforms of cysteine. The kinetic profile of these reagents with

proteins allows their use at low concentrations that minimize cell toxicity. While the rate

will vary with the individual protein,7 the excellent bioorthogonal selectivity combined

with an enhanced rate and MS/MS compatible profile make strained cycloalkynes

valuable new tools for detecting protein sulfenic acids in vitro or in vivo.

2.5 Materials and Methods

General Considerations

All chemicals were purchased from Sigma Aldrich Chemical Company and used as

received. TLC was performed on Sorbent polyester-backed Silica G plates with UV254

indicator. Visualization was accomplished with UV light unless otherwise indicated.

Solvents for extraction and purification were of technical grade and used as received.

Liquid chromatography–mass spectrometry (LC-MS) solvents were HPLC grade. For

small molecule experiments, ESI-MS was performed on an Agilent 100 Series LC/MSD

ion trap. 1H and 13C NMR spectra were recorded using a Bruker Avance 300 MHz NMR

63

spectrometer or Bruker 500 MHz NMR spectrometer. Chemical shifts are given in ppm

(δ); multiplicities are indicated by s (singlet), d (doublet), t (triplet), q (quartet), m

(multiplet). UV–vis spectroscopy was performed on a Cary 50 UV-vis spectrophotometer.

Scheme 7. Reaction of thermally generated organic-soluble cysteine-derived sulfenic

acid with BCN

A solution of diethyl 2-(2-(((R)-3-(tert-butoxy)-2-((tert-butoxycarbonyl)amino)-

3-oxopropyl)sulfinyl)propan-2-yl)malonate (150 mg, 0.30 mmol) and bicyclo[6.1.0]non-

4-yn-9-ylmethanol (BCN, 12.5 mg, 0.083 mmol) in DCM (2 mL) was heated to 35 °C and stirred for 12 hours.11,12The resulting solution was concentrated in vacuo and purified

by column chromatography on silica gel (EtOAc:MeOH, 99:1) to afford diastereomers of

tert-butyl (tert-butoxycarbonyl)((9-(hydroxymethyl)bicyclo[6.1.0]non-4-en-4-

yl)sulfinyl)-D-alaninate (31 mg, 84%) as a colorless oil. Rf 0.5 (EtOAc:MeOH, 19:1,

+ 1 KMnO4). ESI-MS calcd for C22H37NO6S [M+H] : 444.2; found 444.5. H NMR (300

MHz, CDCl3): δ 6.44 (q, J = 6.8 Hz, 1H), 5.75 (br s, 0.5H), 5.50 (br s, 0.5H), 4.59 – 4.28

(m, 1H), 3.69 (d, J = 7.5 Hz, 2H), 3.20 (br s, 1H), 3.04 – 2.82 (m, 1H), 2.67 – 1.98 (m,

13 7H), 1.96 – 0.80 (m, 27H), C NMR (75 MHz, CDCl3): δ 169.71, 169.43, 169.37, 155.60,

155.37, 143.94, 143.80, 143.19, 136.32, 135.47, 135.07, 134.10, 83.32, 83.27, 83.08,

64

83.03, 80.42, 80.28, 77.56, 59.92, 59.88, 54.52, 53.87, 51.43, 50.88, 28.57, 28.23, 28.21,

28.02, 27.82, 27.54, 27.51, 24.35, 24.29, 24.22, 24.11, 24.07, 23.55, 23.51, 23.35, 23.30,

22.85, 21.65, 21.62, 21.37, 19.52, 19.25, 19.18, 19.13, 19.03, 18.84, 18.80.

Scheme 8. Reaction of thermally generated organic-soluble cysteine-derived sulfenic

acid with tCOT(trans-Cyclooctene)

A sealed tube was charged with DCM (2 mL), diethyl 2-(2-(((R)-3-(tert-butoxy)-

2-((tert-butoxycarbonyl)amino)-3-oxopropyl)sulfinyl)propan-2-yl)malonate145 (175 mg,

0.35 mmol) and (E)-cyclooctene (78 mg, 0.71 mmol) and heated to 40 °C with stirring

for 12 hours. The resulting solution was concentrated in vacuo and purified by column

chromatography on silica gel (EtOAc:hexanes, 1:2) to afford diasteromers of tert-butyl

(tert-butoxycarbonyl)(cyclooctylsulfinyl)-D-alaninate (100 mg, 70%) as a colorless oil.

+ Rf 0.4 (EtOAc:hexanes, 1:1, I2). ESI-MS calcd for C20H37NO5S [M +Na] : 426.2; found

1 426.5. H NMR (CDCl3, 500 MHz): δ 5.80 (d, J = 8.4 Hz, 1H), 5.61 (d, J = 6.3 Hz, 1H),

4.55 (td, J = 7.9, 3.5 Hz, 1H), 4.47 (q, J = 5.7 Hz, 1H), 3.26 (dd, J = 13.2, 5.8 Hz, 1H),

3.16 (dd, J = 13.0, 7.6 Hz, 1H), 2.96 (dd, J = 13.2, 4.9 Hz, 1H), 2.90 (dd, J = 13.2, 3.7 Hz,

1H), 2.84 – 2.74 (m, 1H), 2.06 – 1.94 (m, 2H), 1.94 – 1.70 (m, 4H), 1.70 – 1.57 (m, 5H),

13 1.49 – 1.34 (m, 29H). C NMR (126 MHz, CDCl3) δ 169.74, 169.35, 155.67, 155.45,

65

83.38, 83.04, 80.41, 80.24, 60.54, 59.80, 51.48, 50.99, 50.83, 49.75, 28.58, 28.24, 28.22,

26.68, 26.66, 26.63, 26.23, 26.21, 26.17, 26.05, 26.02, 26.00, 25.87, 25.72.

H OH H O O O S O S H + H H r.t. MeCN OH O O (99%) BCN

Scheme 9. The Reaction of Fries Acid with BCN.

A solution of Fries Acid (15 mg, 0.058 mmol), and bicyclo[6.1.0]non-4-yn-9- ylmethanol154 (BCN, 8.7 mg, 0.058 mmol) in MeCN (10 mL) was stirred at r.t. for 20

minutes.122 The resulting solution was concentrated in vacuo to yield diastereomers of 1-

((9-(hydroxymethyl)bicyclo[6.1.0]non-4-en-4-yl)sulfinyl)anthracene-9,10-dione which

required no further purification (23.6 mg, 99%) as an orange solid. Rf 0.4 (EtOAc:MeOH,

+ 1 19:1). ESI-MS calcd for C24H22O4S [M +H] : 407.1; found 407.4. H NMR (500 MHz,

CDCl3): δ 8.72 (ddd, J = 25.9, 7.8, 1.3 Hz, 2H), 8.48 (ddd, J = 10.4, 7.7, 1.3 Hz, 2H),

8.33 – 8.26 (m, 2H), 8.26 – 8.19 (m, 2H), 8.03 (dt, J = 10.0, 7.8 Hz, 2H), 7.88 – 7.74 (m,

4H), 7.13 – 7.03 (m, 2H), 3.58 (d, J = 7.4 Hz, 0.5H), 3.55 (d, J = 7.4 Hz, 0.5H) 3.52 –

3.40 (m, 3H), 2.60 (q, J = 5.5 Hz, 0.5H), 2.57 (q, J = 5.5 Hz, 0.5H), 2.49 – 2.34 (m, 4H),

2.34 – 2.22 (m, 2H), 2.15 – 2.02 (m, 2H), 1.93 (dq, J = 14.6, 5.4 Hz, 1H), 1.82 (dq, J =

14.4, 5.5 Hz, 1H), 1.59 – 1.39 (m, 5H), 1.39 – 1.28 (m, 2H), 1.23 – 0.93 (m, 5H), 0.77 (p,

13 J = 16.6 Hz, 1H), 0.67 – 0.56 (m, 1H), 0.05 – -0.01 (m, 1H). C NMR (126 MHz, CDCl3)

δ 183.15, 182.87, 182.67, 182.60 148.58, 148.24, 144.94, 144.06, 141.30, 140.41, 135.08,

66

135.04, 134.91, 134.88, 134.86, 134.44, 134.38, 133.32, 132.78, 131.72, 131.67, 131.22,

131.14, 129.99, 129.84, 127.70, 127.68, 127.64, 127.53, 60.05, 59.87, 28.87, 27.47, 24.44,

23.74, 23.48, 23.43, 23.20, 21.91, 21.65, 21.08, 19.80, 19.59, 18.62, 17.56.

Scheme 10. The reaction of Fries Acid with tCOT.

A solution of Fries Acid (10 mg, 0.038 mmol), and (E)-cyclooctene (13 mg, 0.118 mmol) in MeCN (10 mL) in a sealed vial wrapped with foil was stirred for 12 hours. The resulting solution was concentrated in vacuo and purified by column chromatography on silica gel (EtOAc:hexanes, 1:2) to afford 1-(cyclooctylsulfinyl)anthracene-9,10-dione

(12.5 mg, 90%) as an orange solid. Rf 0.4 (EtOAc:hexanes, 1:1). ESI-MS calcd for

+ 1 C22H22O3S [M +H] : 367.1; found 367.4. H NMR (CDCl3, 300 MHz): δ 8.53 (dd, J =

7.9, 1.3 Hz, 1H), 8.46 (dd, J = 7.7, 1.3 Hz, 1H), 8.36 – 8.24 (m, 2H), 8.00 (t, J = 7.8 Hz,

1H), 7.90 – 7.79 (m, 2H), 3.09 (tt, J = 9.9, 3.8 Hz, 1H), 2.66 – 2.48 (m, 1H), 2.28 – 2.11

13 (m, 1H), 2.02 – 1.02 (m, 16H) C NMR (75 MHz, CDCl3) δ 184.17, 182.68, 148.93,

135.22, 135.04, 134.86, 134.30, 133.40, 133.05, 131.84, 130.38, 129.61, 127.86, 127.71,

61.98, 31.09, 27.12, 26.66, 26.58, 26.45, 25.57, 22.80.

67

Scheme 11. Synthesis of BCN-Bio

D-(+)-Biotin (72 mg, 0.29 mmol), hydroxybenzotriazole (7.9 mg, 0.058 mmol),

and activated molecular sieves were added to a flask and subjected to three vacuum/argon

purge cycles. Anhydrous DMF (2.5 mL) was added and the solution was heated to 60 °C

for 30 minutes. The solution was cooled to r.t. and a solution of

dicyclohexylcarbodiimide in DCM (0.32 mL, 0.32 mmol) was added dropwise and stirred

for 3 h (solution goes cloudy). Bicyclo[6.1.0]non-4-yn-9-ylmethanol (53 mg, 0.35 mmol)

and N,N-dimethylaminopyridine (0.4 mg, 0.003 mmol) were added and the flask was

wrapped in aluminum foil, heated to 60 °C for 4 h, then stirred at r.t. for 24 h.154 The mixture was filtered and washed with DCM/MeOH (1:1) and the filtrate was concentrated in vacuo. Column chromatography on silica gel (EtOAc:MeOH, 19:1,

KMnO4) affords BCN-Bio (70 mg, 63%) as a colorless oil. Rf 0.25 (EtOAc:MeOH, 19:1).

+ 1 ESI-MS calcd for C20H28N2O3S [M +H] : 377.18; found 377.2. H NMR (300 MHz,

MeOD) δ 4.35 (ddd, J = 7.9, 5.0, 1.0 Hz, 1H), 4.16 (dd, J = 7.9, 4.5 Hz, 1H), 4.05 (d, J =

8.2 Hz, 2H), 3.06 (ddd, J = 8.9, 5.7, 4.5 Hz, 1H), 2.79 (dd, J = 12.7, 5.0 Hz, 1H), 2.57 (d,

J = 12.7 Hz, 1H), 2.21 (t, J = 7.2 Hz, 2H), 2.15 – 1.94 (m, 6H), 1.71 – 1.38 (m, 7H), 1.40

– 1.18 (m, 4H), 0.90 – 0.66 (m, 2H). 13C NMR (75 MHz, MeOD) δ 175.79, 166.11, 99.66,

63.62, 63.46, 61.76, 61.64, 57.05, 41.08, 35.04, 30.20, 29.77, 29.53, 26.08, 21.96, 21.44,

18.59.

68

UV-Vis Kinetics; Fries Acid + BCN

Stock solutions of Fries Acid (3 mM) in MeCN and BCN (30 mM) in MeCN

were prepared. The Fries Acid stock (0.1 mL) was added to MeCN (0.8 mL) in a 1 mL

UV-vis cuvette and lastly an aliquot of the stock solution of BCN (0.1 mL) was added,

followed by a quick shake of the cuvette. The cuvette was immediately loaded into a

Cary 50 UV-vis spectrophotometer and measurements began. This experiment was

repeated 1x at identical concentrations and 2x at half the BCN concentrations (0.05 mL

of BCN and 0.85 mL of MeCN). UV-vis data was recorded at 453 nm with 6 second intervals and averaged at each time point. SigmaPlot was utilized to plot data using

(-bx) exponential decay with the equation: f = y0+a(e ).

UV-Vis Kinetics; Fries Acid + BCN, Aqueous

Stock solutions of Fries Acid (3 mM) and BCN (30 mM) were prepared in MeCN.

The Fries Acid stock (0.1 mL) was added to MeCN (0.3 mL) and ammonium bicarbonate

buffer (0.5 mL, 50 mM, pH 7.5) in a 1 mL UV-vis cuvette, then an aliquot of the stock

solution of BCN (0.1 mL) was added, followed by a quick shake of the cuvette. The

cuvette was immediately loaded into a Cary 50 UV-vis spectrophotometer and

measurements began. This experiment was repeated 1x at identical concentrations and 2x

at half the concentration of BCN (0.05 mL of BCN, 0.35 mL MeCN, and 0.5 mL buffer).

UV-vis data was recorded at 453 nm with 6 second intervals and averaged at each time

point. SigmaPlot was utilized to plot data using exponential decay with the equation: f =

(-bx) y0+a(e ).

69

UV-Vis Kinetics; Fries Acid + tCOT

Stock solutions of Fries Acid (3 mM) and tCOT (60 mM) were prepared in MeCN.

The Fries Acid stock (0.1 mL) was added to MeCN (0.8 mL) in a 1 mL UV-vis cuvette, then tCOT (0.1 mL) was added, followed by a quick shake of the cuvette. The cuvette

was immediately loaded into a Cary 50 UV-vis spectrophotometer and measurements

began. UV-vis data was recorded at 453 nm with 60 second intervals. SigmaPlot was

(-bx) utilized to plot data using exponential decay with the equation: f = y0+a(e ).

Scheme 12. Reactivity of BCN with various small molecule sulfur oxoforms

A stock solution of BCN (133.3 mM) was prepared with HPLC grade MeCN.

Additional stocks (133.3 mM) of reduced glutathione, s-nitrosoglutathione, oxidized glutathione, phenylsulfinic acid sodium salt, and phenylsulfonic acid were prepared with

50:50 MeCN:ammonium bicarbonate buffer (50 mM, pH 7.5). The BCN stock solution

was combined with each oxoform stock separately to final concentrations of BCN (10

mM) and oxoform (10 mM) in 50:50 MeCN:ammonium bicarbonate buffer (50 mM, pH

7.5). The solutions were analyzed by ESI-MS (positive ion mode) for disappearance of

BCN and sulfur oxoform at both 1 h and 24 h. No change in ESI-MS was observed. UV-

Vis tracking of s-nitrosoglutathione at 335 nm164 showed no decrease over a 24h period.

70

Additionally, the PhSO2Na and PhSO3H solutions were concentrated in vacuo and

determined to be unchanged via 1H NMR.

Scheme 13. Stability of BCN-Alkenylsulfoxide to common reducing agents

The BCN-sulfoxide product was combined with DTT and TCEP in separate solutions to final concentrations of 10 mM each in 50:50 MeOH:ammonium bicarbonate buffer (50 mM, pH 7.5). The reactions were monitored by TLC for disappearance of sulfoxide at 1 h, 24 h and 72 h. No change in TLC was observed, and the solutions were concentrated in vacuo and analyzed by 1H NMR, revealing no change in sulfoxide

starting material. Additional analysis via ESI-MS showed no change in the MS spectrum, with strong signals (m/z) for BCN-sulfoxide at 444.3 [M+H]+ and 466.3 [M+Na]+.

Generation of AhpC-SOH and AhpC-SO2H

The C165A mutant of Salmonella typhimurium AhpC was overexpressed and

purified from E. coli as previously described.124,165 Mutant AhpC was pre-reduced with

DTT (10 mM) for 30 min at ambient temperature, and DTT was then removed by passing

the solution through a Bio-Gel P6 spin column equilibrated with ammonium bicarbonate

(50 mM). Protein concentration was determined using the solution absorbance at 280 nm

71

( = 24,300 M-1 cm-1).165 The sulfenic acid species was generated via treatment with 1

equivalent of hydrogen peroxide for 30-45 seconds at room temperature (pH 7-7.5 buffer)

with quenching by passage through a Bio-Gel spin column equilibrated with ammonium

bicarbonate (50 mM). The sulfinic acid was generated via treatment of AhpC-SH with 2

equivalents of hydrogen peroxide for 45-75 seconds at room temperature with similar

removal of unreacted substrate. Formation of each oxidized species was confirmed by

ESI-TOF MS.

Generation of AhpC-SNO and AhpC-S-S-Cys

C165A AhpC was pre-reduced as detailed above then transnitrosated with freshly

prepared S-nitrosocysteine in HEPES/DTPA buffer (25 mM HEPES, 1 mM DTPA, pH

7.7) for 1 h in the dark as previously described.166 Cys-SNO was removed using Bio-Gel

spin columns equilibrated with ammonium bicarbonate (50 mM) and formation of AhpC-

SNO and the mixed disulfide AhpC-S-S-Cys were confirmed by ESI-TOF MS.

Reactivity of AhpC oxoforms with strained cyclooctynes

Strained cyclooctyne probes were pre-treated with immobilized TCEP for 1 h at

room temperature, then spun to concentrate beads and retain supernatant. Freshly

prepared AhpC-SOH (40 M), AhpC-SO2H (40 M), or the AhpC-SNO/AhpC-S-S-Cys mixture (40 M total protein) was incubated at room temperature in the presence of the

strained alkyne probe (100 M) in ammonium bicarbonate buffer (50 mM, pH 7.5).

Reactions involving AhpC-SNO were protected from light. Aliquots (40-50 L) were quenched at various time points by passage through Bio-Gel spin columns pre-

72

equilibrated with either 0.1% formic acid in water or ammonium bicarbonate (50 mM,

pH 7.5) for analysis by ESI-TOF MS.

Reactivity of the C165A AhpC-S(O)-BCN-Bio adduct with H2O2 and reductants

C165A AhpC-SOH (40 M) was freshly prepared as detailed above, then treated

with BCN-Bio (250 M) at pH 7.5 for 4 h (r.t., 50 mM NH4HCO3) to ensure maximum

o adduct formation, then incubated overnight at 4 C in the presence of H2O2 (750 M) to convert unreacted protein species to the irreversible oxoforms sulfinic and sulfonic acid.

The resulting mixture was passed through a Bio-Gel spin column to remove BCN-Bio and H2O2 and analyzed by ESI-TOF MS for both the amount of AhpC-S(O)-BCN-Bio adduct and to demonstrate the stability of the adduct toward excess H2O2. AhpC labeled

with BCN-Bio (40 M protein) was then treated with reductants DTT, TCEP, -

mercaptoethanol, glutathione, and N-acetyl cysteine (1 mM each) at pH 7.5 (1 h, r.t.). An

aliquot from each reaction was analyzed by SDS-PAGE (using non-reducing 4x Laemmli

buffer) and Western blotting, while the remainder of each sample was quenched by

passage through Bio-Gel spin columns pre-equilibrated with 0.1% formic acid in water

for analysis by ESI-TOF MS.

Electrospray ionization time-of-flight mass spectrometry

ESI-TOF MS analyses were performed on an Agilent 6120 MSD-TOF system

operating in positive ion mode with the following settings: capillary voltage of 3500 V,

nebulizer gas pressure of 30 psig, drying gas flow of 5 L/min, fragmentor voltage of 175

V, skimmer voltage of 65 V, and gas temperature of 325oC. Samples were introduced via

direct infusion at a flow rate of 20 L/min using a syringe pump (KD Scientific). Mass

73

spectra were averaged and deconvoluted using the Agilent MassHunter Workstation

software v B.02.00.

nanoLC-MS/MS analysis

Digestions were performed at pH 7.5-8.0 using Trypsin Gold (Promega) overnight

at 37oC. The resulting peptides were analyzed on a Dionex UltiMate3000 splitless nanoLC system coupled to a Thermo Orbitrap Velos Pro high-resolution mass spectrometer. Peptides were separated using a gradient of buffer A (0.1% formic acid/2% acetonitrile/98% water) and buffer B (0.1% formic acid/20% water/80% acetonitrile) over 60 minutes (2 to 85% B) at a flow rate of 300 nL/min with the column held at 35oC.

Eluant was introduced to the mass spectrometer via positive nanospray ESI with the

following settings: capillary temperature 200oC, spray voltage 1.8 kV, spray current 100

mA. The mass spectrometer was operated in data-dependent acquisition mode using

Xcalibur v. 2.1 (Thermo). After a full scan (150-2000 m/z range) at high resolution

(60,000), the top 15 most intense precursor ions were isolated and fragmented using collision-induced dissociation (CID). Dynamic exclusion was enabled with a repeat duration of 30 seconds and an exclusion duration of 9.5 seconds. The normalized collision energy was set at 35%, activation Q at 0.25, and activation time at 10 ms.

Acquired raw data were processed using Proteome Discoverer v 1.4 (Thermo).

General SDS-PAGE and Western blot procedures

Proteins were separated by SDS-PAGE and transferred to a nitrocellulose membrane (0.45 µm, BioRad). Blocking was achieved with 5% (w/v) BSA (5% nonfat milk for AhpC detection) in TBS buffer containing 0.1% Tween 20 (TBS-T). Protein

74

biotinylation was detected by overnight incubation with anti-biotin HRP (Cell Signaling

Technologies) diluted 1:1000-1:2000 in 5% (w/v) BSA in TBS-T. AhpC content was

detected using anti-AhpC (1:2000) in 5% (w/v) non-fat milk in TBS-T. -actin levels

were detected using anti--actin (1:10,000) (Cell Signaling Technologies). The secondary

antibody used for detection of AhpC and -actin was anti-rabbit HRP (1:2000-1:5000 in

1% BSA, TBS-T) (Cell Signaling Technologies). Western blots were developed using

Western Lightning Plus ECL (PerkinElmer) or WesternBright ECL (Bioexpress) reagents

followed by exposure to autoradiography film (BioExpress).

Labeling of protein SOH in lysate using BCN-Bio

Human squamous cell carcinoma cells (SCC-61) were cultured in DMEM/F12

medium supplemented with 10% FBS (Invitrogen) at 37°C and 5% CO2. Cells were lysed

with modified RIPA (mRIPA) buffer (50 mM Tris pH 7.4, 1% NP-40, 0.25% sodium deoxycholate, 15 mM NaCl, 1 mM EDTA, 1 mM NaF, supplemented with protease and

phosphatase inhibitor tablets (Roche)) and incubated on ice for 1 h followed by

centrifugation at 13,000 rpm for 10 min. The supernatant was retained, and as needed,

passed through Bio-Gel spin columns pre-equilibrated with PBS. Immobilized TCEP

resin (Thermo; 0.4 mL of suspension) was placed in a clean tube and pre-washed twice

with 0.5 mL of 100 mM NH4HCO3 then once with 0.5 mL of 50 mM citrate bis-Tris

propane pH 8.5. Washes consisted of resuspending resin by vortexing, spinning to collect resin, and removing supernatant. Neutralization of TCEP resin was confirmed by monitoring pH of supernatant. Lysate flow-through in PBS was combined (460 L of

1.24 mg/mL) and added to the pre-washed TCEP resin. The mixture was rotated end-

75

over-end at r.t. for 45 min, and beads were then removed by filtration. The reduced

protein solution was spiked with C165A AhpC (9 g) for loading control and aliquotted in 60 g fractions. Lysates were treated with 100 M of BCN-Bio and the following

H2O2 concentrations: 2.5, 5, 10, 20, 50, and 500 M. A control sample containing 60 g

of lysate was treated with 5 mM TCEP to prevent protein oxidation then 100 M of 4 was immediately added. Two additional controls included untreated lysate and lysate incubated with 100 M of iodoacetyl-LC-biotin (Thermo). Final reaction volumes were

0.1 mL, and all samples were incubated for 30 min at r.t. in the dark. Reactions were quenched via the addition of 4x Laemmli buffer containing 10% (v/v) 2-mercaptoethanol.

Labeling of protein SOH at lysis with BCN-Bio

SCC-61 cells were cultured as detailed above. One hour before treatment, media was removed and replaced with media devoid of FBS. Media was removed, cells were washed twice with ice cold PBS, then lysed with mRIPA (0.2 mL) in the presence of

TCEP (10 mM) or BCN-Bio (0.1 mM) and incubated for 30 min on ice. BCN-Bio was then added to the TCEP-treated lysates to a final concentration of 0.1 mM. DMSO was present at 0.2% (v/v) and all reaction times were limited to 30 min on ice (all incubations with BCN-Bio were performed simultaneously). Each treatment was performed in triplicate. Lysates were clarified via centrifugation at 13,000 rpm for 10 min at 4oC and

excess probe was removed as described above. Samples were normalized and replicates

analyzed individually by reducing SDS-PAGE and Western blotting for protein

biotinylation and -actin levels (loading control).

76

Cell viability assay

Cell viability was determined using SCC-61 cells and a modified MTT

assay.167,168 Briefly, 50,000 cells per well were plated in 24-well plates and incubated

overnight. Cells were then treated with either dimedone (0.1 to 3.0 mM) or BCN-Bio

(0.01 to 1 mM) for 48-72 h. The SOH probes were added as solutions in DMSO to the

media such that the final concentration of DMSO was < 0.3%. At the end of the

incubation period, MTT was added and the cultures were incubated for 4 h at 37°C. The

dark crystals formed were dissolved by the addition of an equal volume of sodium

dodecyl sulfate/dimethylformamide (SDS/DMF) extraction buffer. Plates were then

incubated overnight at 37°C. A 100 L aliquot of each soluble fraction was transferred

into a 96-well microplate, and the absorbance at 570 nm was measured. The mean

absorbance of the biological replicates (3 replicates for BCN-Bio; 4 replicates for

dimedone) was calculated and survival data is reported relative to untreated control cells.

Labeling of protein SOH in live cells using BCN-Bio and DCP-Bio1

SCC-61 cells were cultured in 6-well plates as detailed above. One hour before treatment, media was removed and replaced with media devoid of FBS. For treatment, media was removed, and each well treated in triplicate with 1 mL media containing various concentrations of BCN-Bio (25 M, 50 M, 100 M) or DCP-Bio1 (25 M, 100

M, 1 mM). DMSO was present at a final concentration of 0.2% (v/v). The cells were then incubated for 30 min at 37°C and 5% CO2. The media was removed, the cells

washed twice with ice cold PBS, then lysed with mRIPA buffer 1 h on ice followed by

centrifugation at 13,000 rpm for 10 min at 4oC. Excess probe was removed using Bio-Gel

77 spin columns pre-equilibrated with PBS, samples were normalized, and replicates analyzed individually. Proteins were separated by reducing SDS-PAGE and analyzed by

Western blotting for protein biotinylation and -actin levels (loading control).

78

CHAPTER 3

OPTIMIZING ALKYNE BASED SULFENIC ACID TRAPS

The work of chapter 2 serves as a proof of concept for a novel approach to trapping sulfenic acids.147 The highly strained bicyclononyne reacts with sulfenic acids via a

pericyclic reaction mechanism to form vinyl sulfoxides in a bioorthogonal manner. The

application of strain to the alkyne bonds was inspired by Bertozzi et al.146 and accelerates

the reaction to a biologically useful rate. The enhanced rate is of particular importance

due to the rapid reaction of sulfenic acids, such as their resolution with resolving thiols in

peroxiredoxins.82,83,92,97 However, the enhanced rate comes at the expense of stability.

The probe is prone to decomposition and polymerization as shown in Figure 28. In

addition, reports exist of reactions with thiols and persulfides.140,160,169–172 Improving

upon the stability and selectivity of these probes is important for the accurate assignment

of proteomic data, and the understanding of redox signaling.

3.1 Reactions with thiols

With selectivity being highly important for labeling sulfenic acids, it is important

to understand the possible side reactions. The reaction between thiols and alkynes is

commonly referred to as the “thiol-yne” reaction.172,173 Thiol-yne reactions been

observed in both organic172 and aqueous169 media. For unstrained terminal alkynes, it

appears that two equivalents of thiol are able to add, first through a thiol-yne reaction,

and secondly through a “thiol-ene” reaction (Scheme 14).172 Indeed, these reactions

appear to be so robust that they have been called “click” reactions and are applied to the

79

reproducible synthesis of polymers.173,174 Interestingly in polymer synthesis it has been

shown that double addition of thiols to alkynes is the primary product when reacting

stoichiometric equivalents of alkynes and thiols suggests that the secondary thiol-ene

reaction occurs quickly after the rate limiting thiol-yne reaction.175

Scheme 14. Thiols can add to alkynes twice. First the rate limiting thiol-yne, then the faster thiol-ene reaction.

Thiol-yne reactions are believed to occur via a catalytic radical mechanism

(Figure 37).176 Starting with formation of a thiyl radical, followed by radical addition to an alkyne, forming a vinyl radical. The vinyl radical regenerates the thiyl radical which

adds to an alkene, forming an alkyl radical which then reforms the thiyl radical,

completing the cycle. In some cases, simple exposure to a small amount of sunlight is

enough to initiate the reaction,172,177 while in other cases photo lamps and radical

initiators are used to promote the reaction.176 However, in a recent case the use of base increases reaction progression,171 which suggests the reaction proceeds through

deprotonation and nucleophilic attack. However, the authors note the use of the radical

scavenger TEMPO significantly reduced their yields and suggests a radical

mechanism.171

80

Figure 37. Thiol radical cycle. This figure was adapted from Fairbanks et al.176

However, there still appears to be some debate about the full nature of the mechanism with some reports that it happens spontaneously.160,178 Analysis of the

“spontaneous” cases reveals the potential for alternative explanations. The addition of

octanethiol to cyclooctyne reported by Fairbanks et al. indicates addition can occur in the

absence of light or a photoinitator.178 However, this reaction only occurs in unpurged

flasks and an unknown mechanism involving reactive oxygen species may exist.178 This

report is highly suggestive of sulfur oxidation to the sulfenic acid and subsequent addition to the cyclooctyne. Another report of “spontaneous” addition to alkynes is from the Bertozzi lab where tandem mass spectrometry (TMS), eliminates the possibility of sulfenic acid addition to the alkyne.179 Likewise, the thiol addition only occurs when the

81 alkyne contains two strongly electron-withdrawing propargylic fluorines, potentially by a nucleophilic mechanism in this particular case.179

The most conclusive evidence supporting a radical based mechanism was found from the combination of cysteine and BCN (3:1) in deuterated solvent (1:2, CD3CN/D2O), followed by 1 week of stirring (Scheme 15). The resulting product existed as a 87:13 mixture of the deuterium:hydrogen containing product.160 If a vinyl anion intermediate was formed via nucleophilic attack, then the anion would be quickly deuterated from the solvent. However, if a vinyl radical was formed, it would be more selective for abstraction of a hydrogen radical from a thiol, thus producing non-deuterated products.

Interestingly this case strongly suggests radical reactivity; however there is no initiator present.

82

Scheme 15. Addition of cysteine to BCN. BCN and 3 eq. of cysteine were stirred for 1 week at r.t. in deuterated solvents. The thiol-yne reaction appears to have occurred via a radical mechanism. The vinyl radical produced by addition of thiyl radical to the alkyne, would remove a thiol hydrogen much more easily than a deuterium from D2O. However,

a nucleophilic addition would have a vinyl anion intermediate which would quickly and

quantitatively remove a deuterium from the solvent.

With a radical mechanism being the most likely, it raises the question of where

the radicals are coming from. In lysate or in vitro experiments, the radicals could come

from ambient light initiation. Fairbanks et al. suggest they could come from molecular

oxygen creating reactive oxygen species.178 In complicated cell and biological

environments the radical initiation could come from endogenous sources such as

mitochondrial generated superoxide. Indeed, a number of one electron oxidants of thiols

● ● - ● exist, including hydroxyl radical ( OH), carbonate ( CO3 ), nitrogen dioxide ( NO2),

● ●- 31,180–182 peroxy radicals (ROO ), and obviously superoxide (O2 ). With thiyl radicals,

clearly present under physiological conditions. It raises the question of how to deal with

them or how to design a trap that doesn’t react with them.

83

3.2 Persulfide Addition to Alkynes

In addition to reacting with radical thiols, strained alkynes have been reported to

react with persulfides.170 Specifically, they have been shown to react with alkyl

persulfides and BSA (bovine serum albumin) persulfide.170 The mechanism may proceed

through a thiosulfoxide that spontaneously converts to the vinyl sulfoxide or thioether.170

Conversion to the thioether produces elemental sulfur that can be detected via fluorescent spectroscopy (Scheme 16).170 While, the exact mechanism or product of the reaction is

not completely clear, the postulated mechanism is identical to the concerted reaction of

RSOH with strained alkynes. In support of this mechanism, the model persulfide

tritylSSH reacts with BCN at a very similar rate to that of Fries acid with BCN, at 18.8

and 28.7 M-1s-1 respectively.170 With the mechanism most likely being identical to the

sulfenic acid mechanism, it may be hard to avoid side reaction with persulfides.

R R H R R S CHCl , S S 0 S 3 S -S S MeCN, S HO H H H H H Buffer Fluoresence Spectroscopy R=trityl,RSSR R = t-butyl, RSR S R S R = penSSH, RSSR and RSR R = BSA-SSH, RSR H

Scheme 16. The reaction of BCN with persulfides. Persulfides are believed to react with

BCN at rates similar to sulfenic acids.170

Strained alkynes clearly react with persulfides, thiols, and sulfenic acids to form covalent bonds to the parent protein or small molecule.160,169–172 Currently, this lack of bio-

84

orthogonality limits the use of strained alkynes in complex systems to MS-based analysis

which can differentiate between the product species based on their mass.183 However,

when working in less complex systems where potential additives can be readily utilized,

radical trapping or elimination of undesired species can help increase selectivity.160,184

These tactics may be most useful in lysate experiments where the natural cellular processes have already been disrupted and the addition of reducing, blocking, or trapping agents will be of less consequence.160

A recent report by van Geel et al. suggests the use of thiol blocking reagents such as iodoacetamide (IAM) to cap the thiols and prevent thiol interference with alkyne azide

click reactions.160 In addition, this method should block any side reactivity with persulfides, owing to their reactivity with IAM.185,186 Indeed, they observe decreased thiol

background reactivity and more selective click reactions.160 Interestingly, IAM also

reacts with sulfenic acids and some of the increased selectivity they observe may be due

to blocked RSOH.187–189 As a better alternative to IAM, methylsulfonyl benzothiazole

acetate (MSBTA) will likely be more selective for RSSH and RSH over RSOH (Scheme

17).187,188 However, many blocking reagents would significantly change the redox state of

a cell and be less than ideal for in vivo labeling. Another approach would be to trap all

present radicals with a scavenger such as TEMPO171, however this would also be less

than ideal in living tissue. Lastly, the use of the reducing agent β-mercapto-ethanol

(BME) has been shown to reduce persulfides and increase the selectivity of BCN.184,190

85

Scheme 17. MSBTA is capable of blocking thiols and persulfides while sparing sulfenic acids.

With the use of such additives being less than ideal, the development of an intrinsically selective probe is desired. Preferably, a probe would be selective, without the use of blocking agents that might interfere with natural processes. Adjusting the reactivity for sulfenic acids over thiols or persulfides could potentially be achieved by adjusting strain and electronic properties of the alkyne. In support of this idea, reports show that cyclooctynes vary significantly in their background binding and reactivity with thiols.179 Significant precedent exists for changing the strain and electronics of alkynes to

optimize reactivity with azides.191–194

3.3 Optimizing Strain and Electronics of Alkynes.

The most straightforward way to increase the reaction rate of alkynes with azides

or sulfenic acids is by strain activation.195 Ideally, strain serves to bend the alkyne into a

shape that more closely resembles the transition state, thus leading to a lower energy

barrier and faster reaction. Activating traps with strain is a careful balancing process

86

where too much strain leads to instability. The easiest method of introducing strain is by

adjusting ring size. Larger alkyne containing rings are less strained than smaller, with

cyclooctynes being the smallest stable cyclic alkyne (Figure 38).196 Fine tuning of strain

can be achieved by introducing atoms with different bond lengths into the ring (Figure

38). For instance, introducing silicon into a cycloheptyne may increase the ring size enough to achieve stability.197,198 Similarly, heteroatoms such as oxygen and nitrogen

have shorter bond lengths199 and can be introduced into cyclooctynes to adjust strain and

electronics. Lastly, increasing the s-hybridization of ring carbons shortens bond

lengths,200–202 and fusing rings such as arenes or cyclopropane increases rigidity and

strain.147,154 Properly tuning an alkyne for reactivity and selectivity will likely require

using one or more of these methods to achieve optimal strain.

Figure 38. Methods of moderating strain; ring size, heteroatoms, multiple bonds, hybridization, and fused rings. *Bond lengths are generic bond lengths obtained primarily from the Computational Chemistry Comparison and Benchmark DataBase.199

87

Electronic Activation

Both computational and experimental studies have demonstrated that

manipulating the electronic properties of alkenes and alkynes can have significant effects

on their reactivity with SOH.143,155 Numerous early examples show that conjugation of

alkenes to carbonyls increases the reactivity with SOH,143 suggesting electron poor

alkenes and alkynes are more reactive. Other studies show V shaped Hammett plots for

the elimination of sulfoxides from ethyl to form styrenes, suggesting that both

electron donating and electron withdrawing can accelerate SOH addition to

alkenes.203 Another study indicates electron rich phenyl alkynes react more quickly with

phenyl sulfenic acid.155 Adjusting the electronics of alkynes via the introduction of

electron withdrawing groups at the propargylic position was most notably achieved by

the Bertozzi group in the synthesis of difluorocyclooctynes for Huisgen cycloadditions.204

However, these alkynes are relatively laborious and expensive to synthesize and their solubility is poor.205

Alternatively, attaching benzene rings at the propargylic position has the dual

benefit of increasing strain and provides an attachment point for electron donating groups

(EDG) and electron withdrawing groups (EWG, Figure 39). Introducing inductively

withdrawing heteroatoms into the rings would have the dual benefit of decreasing

electron density and increasing ring strain (Figure 39). The combination of both strain

and electronic activation should yield the most stable, fast and selective trap. Strain

activation will bend the alkyne, destabilizing the starting alkyne, and move it closer to the

transition state geometry (ΔEs). The electronic activation will increase HOMO/LUMO

overlap, leading to a more stabilized transition state and lower transition state energy

88

(ΔEe). Combination of these two activations will result in a much lower transition state

‡ barrier (ΔEs+e , Figure 39).

O RS H EWG

Ee ‡ EWG Es ‡ Es+e

E ‡ R O s Ea S H

Activation

Energya O R S

Figure 39. Combining strain and electronic activation. A theoretical energy diagram for the reaction of alkynes with sulfenic acids.

Computational Design- precedent-literature

The synthesis of cyclic alkynes is not particularly facile, and the number of potential combinations of ring size and electronic activation quickly balloon to unreasonable numbers. Computational analysis could potentially be used to narrow down possible synthetic targets for empirical analysis with sulfenic acids. However, computations have not been applied the reaction of sulfenic acid with strained alkynes. 89

In 2001, Jenks et al. performed the first computational and empirical analysis of the elimination of sulfoxides to form sulfenic acids and pi bonds.143 They found that the

MP2 level of theory with the 6–311+G(3df,2p) fit experimental values quite well.143 This

work shows that strain increased the reactivity of alkenes, electron poor alkenes reacted

more quickly, and simple alkynes had energy barriers of around 11.0 kcal/mol (ΔE‡).

Unfortunately, the computational expense of MP2 scales as a N4 function, making it too

computationally expensive for larger systems.155

More recently Aversa et al. compared a number of density functionals for their

ability to accurately model the addition of phenyl sulfenic acid to aryl alkynes.155 Aversa et al. found the MPW1B95 density functionals from the Truhlar206 to be the most

accurate for modeling energy and molecular geometries of sulfenic acid additions to

alkynes.155 In addition, they found the 6–31+G(d,p),S(3df ) to be a good compromise of size and accuracy for their system.155 In corroboration with other studies, this work

reveals the Markovnikov product is favored for electron donating groups, and the anti-

Markovnikov product was favored for electron withdrawing substituents (Scheme

18).155,207 Interestingly they did not find that electron withdrawing groups attached to the

alkyne accelerated the reaction.155 The regiochemistry may be due to their use of phenyl sulfenic acid, a relatively electron poor species. However, their computations almost perfectly predicted the relative barrier heights of their empirical results.155

90

Scheme 18. Alkyne electronics effect regiochemistry of RSOH addition.

Recently Freeman et. al. modeled sulfenic acids using a wide range of hybrid

density functionals, MP2, and the 6-311+G(d,p) basis set, they modeled complex

transition states for a number of sulfenic acid reactions. 130,208,209 The computational

models included the reaction of alkane sulfenic acids with hydrogen thioperoxide (HSOH)

to form thiosulfinates,208 the reaction of cysteine sulfenic acid with dimedone,209 and cysteine sulfenic acids reacting with dimedone to form O-sulfenate products.130 These computations show the fundamental importance of thoroughly analyzing conformations, solvent, and hydrogen bonding in the context of dimedone based reactions with sulfenic acids. Luckily the transition state of sulfenic acids with alkynes proceeds through a comparably simple concerted cyclic transition state.143

The choice of theory and basis set in computational chemistry is non-trivial.210

The aforementioned computations by Jenks,143 Aversa,155 and Freeman130,208,209 all use different levels of theory and different basis sets. While the MP2 theory used by Jenks143 would be highly suitable for small systems, it would be computationally too expensive for cyclooctyne transition states with 12-20+ heavy atoms. A good balance of computational expense and demonstrated accuracy has been achieved by density

91

functionals. The hybrid density functionals MPW1B95206 has been proven accurate in

very similar alkyne systems155 and a number of updated density functionals from the

Truhlar group exist.211–216 The M06-2X hybrid density functionals from the Truhlar group stand out owing to their exceeding accuracy and availability.214 The MO6-2x

functionals out-perform many other methods on a wide range of benchmarks214,216 including those assessing barrier heights, 214,216 and should work well for sulfenic acid barrier heights.

3.4 Results

The number of potential modifications to ring strain and alkyne electronics reveals a large number of permutations that would prove excessive to synthesize.

Computational chemistry provides the opportunity to identify targets with high potential

for successful trapping of sulfenic acids. We have employed the use of MPW1B95,206

M06-2X,214 and the popular B3LYP to assess barrier heights for the addition of sulfenic acid to strained alkynes. In an effort to directly mimic the work by Aversa, the

MPW1B95 functions combined with the 6-31+G(d,p),S(3df ) basis set were used to assess barrier heights. The M06-2X and B3LYP functionals will be combined with the triple-zeta 6-311+G(d,p)130 and 6-311+G(d,p),S(3df) basis set to evaluate the need for

extra d-functionals when modeling sulfur.217–219 Geometries of sulfenic acids, alkynes

and product sulfoxides were optimized at B3LYP-6-31G(2d) and refined at each higher

level of theory with tight convergence and frequency verification. Transition state

geometries were predicted with QST2 at B3LYP-6-31G(2d), and optimized at higher

levels with verification of one imaginary frequency corresponding to the bond

forming/breaking vibrations.

92

As an initial benchmark molecule, BCN was evaluated at these combinations of

theories and basis sets. The different levels of theory (Named A1, B, B1, B2, M1,and M4

in Figure 40) provided barrier heights that varied significantly. Interestingly, A1 returned

a negative energy barrier which is not a realistic possibility. However, negative energy barriers are not without literature precedent, as Jenks found near zero and negative energy barriers for gas phase eliminations of sulfoxides owing to lower zero-point energies in the transition state.143 Indeed, removing zero-point energies and converting ΔE to ΔG

reveals positive transition states. The variation between these energy barriers suggests

they should not be used for directly predicting absolute kinetics. However, the variation

does not eliminate their usefulness in the evaluation of relative energy barriers and

kinetics between similar reactions. Unfortunately, no literature precedent exists for

computationally evaluating the addition of sulfenic acids to strained alkynes.

93

Figure 40. Computational evaluation of methane sulfenic acid reacting with BCN at a variety of theories.

The lack of literature precedent for these type of computations necessitates grounding the computations with experimental results. A group of five readily synthesized or commercially available alkynes were purchased or synthesized (Figure 41) and served as benchmark compounds for the computations. Cyclooctyne (COT) was synthesized according to methods reported by Meier et al.,220 tetramethoxydibenzocyclooctyne (TMDIBO) waas synthesized according to Stöckmann et

94

al.,221 and BCN was either synthesized according to the procedure by Dommerholt et al. or purchased commercially.154 Dibenzocyclooctynol (DIBO) and dibenzocyclooctyne

(DBCO) were purchased commercially. These molecules have a variety of ring strain

owing to the inclusion of sp2 atoms in the ring, or addition of a fused cyclopropane in the

case of BCN. In addition, TMDIBO is both strained and electron rich, allowing it to

serve as a benchmark for both electronic and strain activation. The reaction of these

strained alkynes with the stable sulfenic acid, Fries Acid, was monitored by UV-Vis. The

alkynes were added to a solution of Fries Acid in MeCN under pseudo first order

conditions as previously reported and the disappearance of the characteristic Fries Acid

peak at 453 nm was tracked.147 These conditions provide a clear kinetic assessment that

can be related to the computations. As expected, the more highly strained alkynes react

quicker. In addition, the electron donating methoxy groups allowed TMDIBO to react

more quickly than DIBO supporting the idea that electronic activation plays an important

role in facilitating the transition state.

95

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

10 or 20x alkyne MeCN O r.t. O

O O

O O N H N H H COT OH OH 2 O (0.7 M-1 s-1) DIBO TMDIBO -1 -1 -1 -1 DBCO HO (1.7 M s ) (4.8 M s ) BCN (10.1 M-1 s-1) (20.0 M-1 s-1)

Figure 41. Five alkynes were reacted with Fries acid under pseudo first order conditions.

Computational evaluation of barrier heights at the aforementioned levels of theory produced interesting results. First and most notably, the barrier heights weren’t always positive. However, consideration of zero-point energies removes this computational artifact.143 While there was significant variation in the overall magnitude of computed

barrier heights, they still predicted the relative reaction rates with a high level of accuracy.

At almost all levels of theory, the computations were able to predict which alkynes would

react most quickly (Figure 42). Graphing these alkynes in order of reaction rate eases

visualization (Figure 42). Parallel lines indicate a high level of congruence between

levels of theory. However, positive slopes suggest inaccuracies when evaluating relative

kinetics. Of particular interest are the results from the lowest level of theory (B, B3LYP-

6-31G(2d)), which ranks the alkynes from fastest to slowest with complete accuracy.

Going a step further and graphing computational barrier heights vs empirically derived

96 rates reveals the simple B level of theory has the strongest linear relationship (Figure 43) which strongly suggests applying this level of theory to the evaluation of relative kinetics will produce useful results.

Figure 42. The benchmark alkynes (Figure 41) were evaluated at the aforementioned 6 levels of theory and results were compared. Parallel lines inndicate the alkynes had very similar relative kinetics, while their absolute values varied significantly.

97

Figure 43. Graphing computational energy barriier vs empirical kinetics and fitting with an exponential curve, shows a high level of correlation for all levels of theory. However,

B matched empirical benchmarks with the highest degree of ccorrelation.

Initial exploratory computations were performed to assess the effect of electron density on barrier height and reaction kinetics (Figure 44). Strain levels were kept static and electronics were varied from electron rich, to neutral, annd to electron poor. Adding four methoxy groups to DIBO lowered the energy barrier byy around 0.8 kcal (ΔΔE‡) per mole and plugging this value into the Arrhenius equation(k = Ae-Ea/(RT)) reveals that the donating groups should be around 3.8x faster which is quite close to the empirical result of 2.82x. On the opposite end of the electronic spectrum, fluorines should accelerate the reaction by around 0.7 kcal which translates to an increase in rate of about 3.3x. These

98

results are congruent with previous reports of “V” shaped Hammett plots, where both

increasing and decreasing electron density accelerated reactions of RSOH and

alkynes.143,203

Figure 44. The effect of electronics on energy barriers was evaluated at B3LYP 6-

31G(2d).

As an alternative method of tuning electron density, heteroatoms can be inserted at the propargylic position inside and outside the ring. While they cannot be inserted directly on the alkyne for stability reasons, they can be positioned on the alpha carbon for a substantial electronic effect. Similar to the findings of Alabugin et al.193 for alkyne

azide click reaction, heteroatom orientation was important for influencing transition state

energetics. When the heteroatoms are gauche, as is the case with exocyclic atoms, they

cannot effectively accept electron density from the alkyne. However when antiperiplanar

as is the case with endocyclic atoms, they can accept electron density, ease alkyne

bending, and delocalize the transition state (Figure 45). As expected, endocyclic

heteroatoms lowered the energy barrier significantly more than exocyclic heteroatoms

99

(Figure 46). A single exocyclic oxygen lowered the barrier of COT from 3.82 kcal to 2.0

kcal, while an endocyclic one lowers the barrier to 0.5 kcal. Going a step further and

adding two endocyclic lowers the energy barrier to -1.27 kcal, corresponding to

an increase in rate of around 5500x. As expected, have a similar effect, and

protonated nitrogen has an even greater effect.

H OH H H

H H H OH H H H H R O R O gauche anti

R O S H R O S H

‡ OH Ea ‡ Ea O

OH O

R O R O S H S H OH

O

Figure 45. Exocyclic heteroatoms align gauche with the alkyne, while endocyclic atoms align anti and have greater effect on electronics.

100

Figure 46. Endocyclic and exocyclic heteroatoms were evaluated for their effect on

energy barriers at the “B” level of theory (B3LYP 6-31G(2d)).

In addition to changing the electronics of the alkyne, heteroatoms significantly

effect the strain of ring systems (Figure 47). One would expect heterocyclic alkynes to

have increased strain, owing to the smaller C-heteratom bond length. However, the

electronics significantly decrease the energy required to bend the alkyne bond. The

endocyclic heteroatoms are able to form a donor-acceptor relationship with π to σ*C-X hyperconjugation leading to a stabilized structure. Stabilizing ring strain may make these structures easier to handle than highly strained structures like BCN which are prone to polymerization and instability.147,193,221 Owing to their ability to stabilize transition states,

and decrease strain, we propose endocyclic heteroatoms as ideal candidates for use in

sulfenic acid traps.

101

Figure 47. Evaluating strain energy as a function of heteroatom location and ring size.

*Strain energies were computed at the B3LYP 6-31G(d) level of theory by Alabugin et al., except for the 3rd row which was computed by us using identical methods.193

Target Molecule:

Electron poor cyclooctynes react more quickly with sulfenic acids than basic

cycloalkynes. 143,155 The highly electron poor di-fluoro cyclooctynes (DFCO) from the

Bertozzi lab should react quickly with sulfenic acids. However, its poor solubility may

exclude it from some applications and it has been shown to react with thiols more quickly

than other strained alkynes.179,205 Alternatively, cyclooctynes with endocyclic oxygen or

nitrogen at the propargylic position have enhanced rate over exocyclic heteroatoms. As

102

an ideal target the mono oxygen cyclooctynes (MOCO) would be uncharged, relatively

soluble, and fast reacting. In addition, there may be some relatively simple synthetic

routes.

Starting from the top (Scheme 19), MOCO would ideally have a handle, such as

an alcohol for easy conjugation to biotin or other useful species. Working backward, the

alkyne can be formed via a bromination/elimination reaction.154 The cyclic alkene can be

made by performing ring closing metathesis (RCM) on the corresponding diene.222,223

The diene can be formed from simple nucleophilic attack on the readily available epoxide, allyl glycidyl .224

Bromination/ Elimination RCM O O O O

MOCO HO OPG OPG MOCO-OH O + Cuprate/ Br O Grignard 250mL 100mL $37.70 $38.00

Scheme 19. Retrosynthetic analysis of MOCO with alcohol handle.

In the forward direction, the Grignard reagent was easily formed from allyl bromide or it can be purchased as the pre-formed Grignard, which we found to produce higher yields with greater ease. In the same flask, a catalytic amount of copper iodide facilitates regioselective addition to the epoxide with a 40-60% yield (Scheme 20). The resulting alcohol was protected with excess 4-dimethylaminopyridine (DMAP) and tert- butyldimethylsilyl chloride (TBDMSCl), yielding the protected diene in 95% yield. The

103 ring closing metathesis reaction proved a challenge. Catalysts including Grubbs I,

Hoveyda Grubbs, and Grubbs II were applied without success. These catalysts were combined with a variety of solvents and temperatures including DCM and Toluene ranging 0 °C to 100 °C, and concentrations from 0.001M to 0.1M, without success. A breakthrough came from a report that emphasized the early formation of oligomers followed by “backbiting” to form ring closed products.225 Observation of polymeric products and knowledge of the “backbiting” mechanism led to the theory that oligomers were forming quickly and consuming the catalyst when heating from lower temperatures.

In support of this theory, heating prior to adding the catalyst achieved excellent yields.

The final conditions were diene (2.5 mM) at 80 °C in toluene followed by addition of

Grubbs I (0.1 eq.), resulting in a 70% yield of the desired ring closed product. However, the subsequent bromination reaction proved problematic, with deprotection occurring simultaneous to bromination. Attempts were made to use alternative protecting groups, however these failed to yield product as well.

104

Scheme 20. Attempted synthesis of MOCO-OH.

In a parallel fashion, the synthesis of a benzene-fused mono oxygen cyclooctyne

(BMOCO) was attempted (Scheme 21). This molecule offers similar electronics, and increased strain. This target arises from allyl Grignard addition to a commercially available aldehyde to yield the secondary alcohol (73%). The resulting alcohol was efficiently protected in 80% yield, giving a diene that could be easily cyclized via ring closing metathesis (70%). However, the resulting alkyne could not be efficiently brominated (Scheme 21). Once again, bromination of these strained compounds with adjacent alcohols appears to be problematic.

105

Scheme 21. Synthesis of BMOCO-OH.

Opting for a target without an oxygen directly on the ring seemed like a logical approach. The Grignard reagent of 5-bromo-1-pentene formed easily and added to the desired aldehyde (70%, from crude aldehyde) smoothly (Scheme 22). The resulting alcohol was then deprotonated with sodium hydride and converted to the allyl ether (49%) with an excess of allyl bromide. The resulting diene could be converted to the cyclooctene in almost quantitative yield (93%). Satisfyingly, this compound was converted to the di-bromide in 69% yield over two steps. This dibromide was subjected to various elimination conditions. At 0 °C the starting material disappeared over about

20 minutes, but subsequent heating quickly led to formation of many undesired products.

Analysis of the same reaction without heating revealed that the single elimination product forms easily at 0 °C (90%). Attempting the elimination of dibromide and monobromide

106

with a variety of bases including lithium diisopropylamide (LDA), lithium

bis(trimethylsilyl) (LiHMDS), and 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) in a

variety of solvents at a range of temperatures produced no desired product. Potential

future directions for this target molecular functionality include alternative methods and

conditions for performing the final elimination, or alternative methods for synthesizing

the alkyne.

Scheme 22. Synthesis of MOCO with alkyl linker.

3.5 Discussion and Future Aims

For the purpose of evaluating relative kinetics, the use of higher levels of theory

does not appear necessary. A strong linear relationship exists between calculated activation barrier and empirical kinetics for the whole span of basis sets and density functionals (Figure 43). However, the linear relationship is stronger for simple “B”

density functionals. Endocyclic propargylic heteroatoms in cyclooctyne rings give a

107

significant rate enhancement to the trapping of sulfenic acids. In addition they are able to

decrease the strain associated with alkyne bending which may offer some selectivity

benefits and increase overall stability.193

A number of desirable synthetic targets still exist including MOCO, DOCO, and

nitrogen containing heterocycles. A number of synthetic routes can be envisioned for the

synthesis of these molecules. However, all of the synthesis will likely center around

alkyne formation, which should be formed relatively late in the synthesis owing to its

high reactivity. Strained alkynes in cyclooctyl system can be synthesized by bromine eliminations as emphasized in this work and others.154 Alkynes can also be synthesized

from by conversion to semicarbazones, oxidation with SeO2 to selenium diazoles

and pyrolysis or elimination with n-butyllithium to form the desired alkyne (Scheme

220,226,227 23A). However, SeO2 can yield undesired regiochemistry in asymmetric rings.

Alternatively the ketone can be converted to the enol triflate, which would likely proceed

with the desired regiochemsitry (Scheme 23B).221 In addition, there may be some metal

mediated reactions that may be useful (Scheme 23C).228

108

Scheme 23. Alternative alkyne synthesis

While this work points towards a number of alkynes with increased reactivity and stability, it does not address the subject of selectivity. Future studies would benefit from computations to evaluate the kinetics of reaction with persulfides to determine if there is a type of alkyne that may be resistant to persulfide trapping. Additionally, more information is needed about thiyl radical reactivity with alkynes. The cysteine thiol bond dissociation energy (BDE) is around 87 kcal/mol229,230 and will react quite quickly with alkynes of normal electronics.160,172,178 However, electron poor alkynes will form electron poor radicals upon reaction which are less stable.231

109

In summary, this chapter points towards a number of ways to accelerate the

trapping of sulfenic acids with alkynes. Computations match empirical benchmarks with

a high level of accuracy. The synthesis of target molecules was attempted, and progress

was made to desired traps. Alkynes clearly have some cross reactivity with persulfides170 and thiyl radicals.160,172,176 However, the extent to which this occurs is unclear and varies between alkynes.179 If needed, mass spectrometry can discern between products.183 As a testament to their selectivity strained alkynes have been used extensively as bioorthogonal reagents in copper free Huisgen cycloadditions.146,153,205,232,233 Indeed, a

current paper indicates no selectivity problems when trapping sulfenic acids in the

presence of thiols, even with highly strained alkynes such as BCN.234 While current

strained alkynes offer room for improvement, their fast rates and selectivity make them

valuable tools for evaluating the redox proteome.

3.6 Methods Section

General Considerations

All chemicals were purchased from Sigma Aldrich Chemical Company and used as

received. TLC was performed on Sorbent polyester-backed Silica G plates with UV254

indicator. Visualization was accomplished with UV light unless otherwise indicated.

Solvents for extraction and purification were of technical grade and used as received.

Liquid chromatography–mass spectrometry (LC-MS) solvents were HPLC grade. For

small molecule experiments, ESI-MS was performed on an Agilent 100 Series LC/MSD

ion trap. 1H and 13C NMR spectra were recorded using a Bruker Avance 300 MHz NMR

spectrometer or Bruker 500 MHz NMR spectrometer. Chemical shifts are given in ppm

110

(δ); multiplicities are indicated by s (singlet), d (doublet), t (triplet), q (quartet), m

(multiplet). UV–vis spectroscopy was performed on a Cary 50 UV-vis spectrophotometer.

Computational Considerations

All computations were performed using Gaussian 09, Revision C.01,

M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R.

Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M.

Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg,

M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y.

Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro,

M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R.

Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J.

Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken,

C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R.

Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G.

A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B.

Foresman, J. V. Ortiz, J. Cioslowski, and D. J. Fox, Gaussian, Inc., Wallingford CT,

2010.

The authors acknowledge the Distributed Environment for Academic Computing (DEAC) at Wake Forest University for providing HPC resources that have contributed to the research results reported within this paper. URL: http://www.deac.wfu.edu

Geometries of sulfenic acids, alkynes and product sulfoxides were optimized at B3LYP-

6-31G(2d) (referred to as “B”) and refined at each higher level of theory with tight

111

convergence and frequency verification. Transition state geometries were predicted with

QST2 at B3LYP-6-31G(2d), and optimized at higher levels with verification of one

imaginary frequency corresponding to the bond forming/breaking vibrations.

Synthetic procedures

Scheme 24. Grignard addition to epoxide, 1-(allyloxy)hex-5-en-2-ol

A heavy stir bar and 3.5g (144.5 mmol, 2.5 eq.) of magnesium turnings were added to a

250 mL RBF which was evacuated and filled with argon 3x. The solid was stirred

vigorously for 30 min to break up the magnesium. The flask was charged with dry THF

(50 mL), cooled to 0 °C, a small iodine crystal was added, and 7 g (5mL, 57.8 mmol, 1

eq.) of allylbromide in THF (50 mL) was added dropwise over 1 hour. The solution was

cooled to -78 °C and 723 mg (3.8 mmol, 0.066 eq.) of CuI was carefully added with

positive argon pressure. The allyl glycidyl ether (4.53 mL, 38.15 mmol, 0.66 eq.) was

added dropwise, and the dewar was covered for slow overnight warming. After stirring

for 10 hours the solution was filtered through celite, quenched with NH4Cl (50 mL, 1/2 saturated, ca. 200 g/L), THF was removed under vacuum, extracted DCM (50 mL, 2x), washed with deionized water, dried over MgSO4, and concentrated under vacuum. The

resulting oil was purified by Biotage LC yielding 1-(allyloxy)hex-5-en-2-ol (3.4g, 21.7

mmol, 57%, THP-K2-63). 1H NMR (300 MHz, Chloroform-d) δ 6.00 – 5.72 (m, 2H),

5.34 – 5.13 (m, 2H), 5.09 – 4.90 (m, 2H), 4.02 (dt, J = 5.6, 1.5 Hz, 2H), 3.87 – 3.74 (m,

112

1H), 3.47 (dd, J = 9.5, 3.1 Hz, 1H), 3.29 (dd, J = 9.5, 7.8 Hz, 1H), 2.32 (d, J = 3.5 Hz,

1H), 2.27 – 2.07 (m, 2H), 1.64 – 1.44 (m, 2H). 13C NMR (75 MHz, Chloroform-d) δ

138.42 , 138.34 , 134.63 , 117.27 , 114.93 , 74.57 , 72.30 , 69.89 , 32.44 , 29.82 .

Scheme 25. Alcohol protection to ((1-(allyloxy)hex-5-en-2-yl)oxy)(tert- butyl)diphenylsilane

A solution of diene (100mg, 0.64 mmol), TBDPSCl (176 mg, 1.32 mmol, 3 eq.), and

DMAP (242 mg, 1.98 mmol, 4 eq.) in dry DCM (20 mL) were stirred overnight (35 °C), washed with HCl (20 mL, 1 M, 2x), washed with brine, dried over MgSO4, and concentrated under vacuum. Chromatography via Biotage LC (0-10% EtOAc) yielded

((1-(allyloxy)hex-5-en-2-yl)oxy)(tert-butyl)diphenylsilane (235 mg, 0.6 mmol, 90%). 1H

NMR (300 MHz, Chloroform-d) δ 7.75 – 7.61 (m, 4H), 7.46 – 7.29 (m, 6H), 5.83 – 5.58

(m, 2H), 5.18 – 5.01 (m, 2H), 4.95 – 4.81 (m, 2H), 3.86 (q, J = 5.5 Hz, 1H), 3.82 – 3.72

(m, 2H), 3.32 (d, 2H), 2.14 – 2.00 (m, 2H), 1.69 – 1.55 (m, 2H), 1.05 (d, J = 3.0 Hz, 9H).

113

Scheme 26. Ring closing metathesis to (Z)-tert-butyldiphenyl((3,4,5,8-tetrahydro-2H-

oxocin-3-yl)oxy)silane

Grubbs I (33mg, 0.04 mmol, 0.1 eq.) was added in one portion under positive argon

pressure to a solution of ((1-(allyloxy)hex-5-en-2-yl)oxy)(tert-butyl)diphenylsilane (157

mg, 0.4 mmol) in dry toluene (160 mL, 2.5 mM) at 80 °C. The solution was stirred for 2

hours, filtered through celite, concentrated under vacuum, and purified via Biotage LC to

yield (Z)-tert-butyldiphenyl((3,4,5,8-tetrahydro-2H-oxocin-3-yl)oxy)silane (103 mg, 0.28

mmol, 70%, THP-K3-23). 1H NMR (300 MHz, Chloroform-d) δ 7.66 (dq, J = 6.3, 1.7 Hz,

4H), 7.47 – 7.30 (m, 6H), 5.61 (dt, J = 10.4, 2.1 Hz, 1H), 5.23 (d, J = 11.7 Hz, 1H), 4.16

(dd, J = 2.4, 1.2 Hz, 1H), 4.03 – 3.85 (m, 2H), 3.64 – 3.56 (m, 2H), 2.57 (d, J = 11.5 Hz,

1H), 2.13 (d, J = 10.0 Hz, 1H), 1.90 – 1.74 (m, 1H), 1.70 – 1.53 (m, 1H), 1.05 (s, 9H).

13C NMR (75 MHz, CDCl3) δ 135.94, 134.46, 134.32, 129.78, 129.73, 129.38, 128.22,

127.70, 127.66, 77.58, 77.52, 77.36, 77.16, 77.09, 76.98, 76.74, 76.68, 75.28, 70.82,

70.29, 34.96, 31.07, 29.86, 27.13, 26.05, 22.13, 19.37.

114

Scheme 27. Attempted bromination of (Z)-tert-butyldiphenyl((3,4,5,8-tetrahydro-2H- oxocin-3-yl)oxy)silane

A solution of Br2 (ca. 0.1 mmol, 2 eq) in dry DCM (1 mL) was added dropwise to (Z)- tert-butyldiphenyl((3,4,5,8-tetrahydro-2H-oxocin-3-yl)oxy)silane (15 mg, 0.052 mmol) in

DCM (1 mL) at 0 °C. Addition produced a reddish brown color that quickly disappeared upon stirring. Addition was continued until the reddish brown color persisted. The solution was quenched with 10% Na2S2O3 (10 mL) and stirred until the color dissipated. The solution was extracted with DCM (10 mL, 2x), washed with deionized water, dried over MgSO4, and concentrated under vacuum. The GC-MS showed two equally present diastereomeric peaks, splitting patterns were indicative of one bromine, and a parent mass of 206/208. 13C-NMR and 1H-NMR showed no protecting group, and there are no longer any alkene peaks. The presence of 14 separate carbon peaks suggest diasteromers. Taken together this information suggests a monobromination product, ether formation, and deprotection. Identical products were observed when a TIPS protecting group was used (THP-K3-51, and THP-K3-43). (m/z)

206/208 [M]+ 1H NMR (300 MHz, Chloroform-d) δ 4.65 – 4.54 (m, 1H), 4.54 – 4.42 (m,

2H), 4.41 – 4.29 (m, 2H), 4.24 – 4.16 (m, 2H), 4.05 – 3.86 (m, 2H), 3.83 – 3.75 (m, 1H),

3.74 – 3.63 (m, 2H), 3.64 – 3.52 (m, 1H), 3.09 – 2.85 (m, 1H), 2.68 – 2.48 (m, 1H), 2.31

115

– 2.14 (m, 2H), 2.12 – 1.97 (m, 2H), 1.94 – 1.79 (m, 2H). 13C NMR (75 MHz, CDCl3) δ

80.36, 80.02, 78.39, 77.02, 71.00, 70.76, 67.28, 65.75, 52.54, 50.32, 31.45, 31.41, 30.07,

28.78.

Scheme 28. Grignard addition to aldehyde, forming 1-(2-(allyloxy)phenyl)but-3-en-1-ol

Allyl magnesium bromide (6.7 mL, 1M in THF, 1.1 eq.) was added dropwise to a solution of 2-(allyloxy)benzaldehyde (1g, 6.16 mmol) in dry THF (60 mL) at -78 °C.

After stirring for 2 hours the reaction was quenched by dropwise addition of deionized water, concentrated in vacuum, diluted into DCM (100 mL), washed with DI H2O, dried over MgSO4, THF was removed in vacuum, purified via Biotage LC 3-12% EtOAC to yield 1-(2-(allyloxy)phenyl)but-3-en-1-ol (920 mg, 73%) as a yellow oil (THP-K3-27).

1H NMR (300 MHz, Chloroform-d) δ 7.36 (dd, J = 7.5, 1.8 Hz, 1H), 7.22 (t, 1H), 6.97 (td,

J = 7.5, 1.1 Hz, 1H), 6.87 (dd, J = 8.2, 1.1 Hz, 1H), 6.15 – 5.96 (m, 1H), 5.96 – 5.74 (m,

1H), 5.42 (dd, J = 17.3, 1.6 Hz, 1H), 5.30 (dd, J = 10.6, 1.5 Hz, 1H), 5.20 – 4.95 (m, 3H),

4.59 (dt, J = 5.1, 1.6 Hz, 2H), 2.56 (tdd, J = 13.4, 7.2, 5.7 Hz, 3H).

116

Si OH TBDMSCl, O DMAP

DCM O O 80%

Scheme 29. Protection of secondary alcohol to form ((1-(2-(allyloxy)phenyl)but-3-en-1-

yl)oxy)(tert-butyl)dimethylsilane

To a solution of 1-(2-(allyloxy)phenyl)but-3-en-1-ol (750 mg, 3.67 mmol) in DCM (36 mL) was added TBDMSCl (609 mg, 4 mmol, 1.1 eq.) and DMAP (896 mg, 7.33 mmol, 2 eq.), stirred overnight, washed with HCl (50 mL, 1 M, 2x), brine, dried over MgSO4 , concentrated under vacuum, and purified by Biotage LC to yield ((1-(2-

(allyloxy)phenyl)but-3-en-1-yl)oxy)(tert-butyl)dimethylsilane (940 mg, 80%). 1H NMR

(300 MHz, Chloroform-d) δ 7.59 (dd, J = 7.6, 1.8 Hz, 1H), 7.33 – 7.23 (m, 1H), 7.06 (t, J

= 1.1 Hz, 0H), 6.91 (dd, J = 8.2, 1.1 Hz, 1H), 6.14 (d, J = 10.6 Hz, 0H), 6.03 – 5.88 (m,

1H), 5.53 (dq, J = 17.2, 1.7 Hz, 1H), 5.38 (dq, J = 10.5, 1.6 Hz, 1H), 5.30 (dd, J = 7.0, 4.5

Hz, 1H), 5.16 – 5.09 (m, 1H), 5.09 – 5.05 (m, 1H), 4.70 – 4.61 (m, 2H), 2.59 – 2.42 (m,

2H), 1.00 (s, 9H), 0.14 (s, 3H), -0.00 (s, 3H).

117

Scheme 30. Ring closing metathesis to form (Z)-tert-butyl((5,6-dihydro-2H- benzo[b]oxocin-6-yl)oxy)dimethylsilane

Grubbs I (141mg, 0.17 mmol, 0.1 eq.) was added in one portion under positive argon pressure to a solution of ((1-(2-(allyloxy)phenyl)but-3-en-1-yl)oxy)(tert- butyl)dimethylsilane (550 mg, 1.72 mmol) in dry toluene (750 mL, 2.3 mM) at 80 °C.

The solution was stirred for 2 hours, filtered through celite, concentrated under vacuum, and purified via Biotage LC to yield (Z)-tert-butyl((5,6-dihydro-2H-benzo[b]oxocin-6- yl)oxy)dimethylsilane (360 mg, 1.24 mmol, 72%). 1H NMR (300 MHz, Chloroform-d) δ

7.41 (dd, J = 7.6, 1.8 Hz, 1H), 7.24 (td, 1H), 7.12 (td, J = 7.4, 1.4 Hz, 1H), 6.99 (dd, J =

7.9, 1.4 Hz, 1H), 5.79 – 5.63 (m, 1H), 5.38 – 5.28 (m, 1H), 5.22 (dd, J = 7.7, 5.3 Hz, 1H),

13C NMR (75 MHz, CDCl3) δ 155.24, 137.19, 129.00, 128.64, 128.43, 126.43, 124.49,

122.01, 73.36, 71.90, 39.20, 26.11, 26.05, 25.79, 18.36, -4.43, -4.46. 4.81 – 4.70 (m, 1H),

4.66 – 4.54 (m, 1H), 0.93 (s, 9H), 0.11 (s, 3H), 0.08 (s, 3H).

118

Scheme 31. Attempted bromination of (Z)-tert-butyl((5,6-dihydro-2H-benzo[b]oxocin-6-

yl)oxy)dimethylsilane

A solution of Br2 (ca. 0.2 mmol, 2 eq) in dry DCM (1 mL) was added dropwise to (Z)-

tert-butyl((5,6-dihydro-2H-benzo[b]oxocin-6-yl)oxy)dimethylsilane (30 mg, 0.1 mmol) in DCM (1 mL) at 0 °C. Addition produced a reddish brown color that quickly disappeared upon stirring. Addition was continued until the reddish brown color

persisted. The solution was quenched with 10% Na2S2O3 (10 mL) and stirred until the

color dissipated. The solution was extracted with DCM (10 mL, 2x), washed with

deionized water, dried over MgSO4, and concentrated under vacuum. The resulting oil was purified and analyzed thoroughly via GC-MS, 1H-NMR, 13C-NMR, and 2D NMR

and there appeared to be an abnormal mono-bromination product (THP-K3-41). 1H

NMR (500 MHz, Chloroform-d) δ 7.22 (td, J = 8.5, 7.1, 1.8 Hz, 1H), 7.08 (dd, J = 7.7,

1.8 Hz, 1H), 7.02 (dd, J = 8.2, 1.2 Hz, 1H), 6.97 (td, J = 7.4, 1.3 Hz, 1H), 6.47 (dd, J =

12.0, 1.7 Hz, 1H), 5.58 (dd, J = 11.9, 6.2 Hz, 1H), 4.98 (dd, J = 13.6, 2.6 Hz, 1H), 4.73

(ddd, J = 9.8, 6.2, 1.7 Hz, 1H), 4.46 (dd, J = 13.6, 1.9 Hz, 1H), 4.07 (ddd, J = 9.8, 2.6, 1.9

Hz, 1H), 0.89 (s, 9H), 0.09 (s, 3H), 0.03 (s, 3H). 13C NMR (75 MHz, CDCl3) δ 156.49,

133.04, 132.36, 130.13, 129.23, 123.54, 122.29, 121.46, 73.30, 72.70, 56.40, 25.93, 18.32,

-4.42, -4.67.

119

Scheme 32. Parikh Doering oxidation to form 4-((tert-butyldimethylsilyl)oxy)butanal

DMSO (4.34 mL, 5 eq.) and TEA (5.1 mL, 3 eq.) were added to a solution of 4-((tert- buutyldimethylsilyl)oxy)butan-1-ol (2.5g, 12.23 mmol) in dry DCM (60 mL, 0.2M). The solution was cooled to -5 °C in a brine bath and sulfur trioxiide pyridine complex (3.89 g,

24.46 mmol, 2 eq.) was added in one portion. After 2 hours the solution was diluted into

200 mL DCM, washed with HCl (100 mL, 2 M, 2x), brine, dried over Na2SO4, and concentrated under vacuum to yield 2.49 g of crude 4-((tert- buutyldimethylsilyl)oxy)butanal (2.49 g, 12.3 mmol, ~100%, THP-K3-77) that could be used without further purification. *Extended storage showed decomposition; this crude material should be used within a few hours. 1H NMR (300 MHz, Chloroform-d) δ 9.78 (s,

120

1H), 3.64 (t, J = 6.0 Hz, 2H), 2.49 (t, J = 7.1, 1.7 Hz, 3H), 1.85 (p, J = 7.1, 6.0 Hz, 2H),

0.88 (s, 9H), 0.03 (s, 6H).

Scheme 33. Grignard addition to aldehyde, forming 1-((tert-butyldimethylsilyl)oxy)non-

8-en-4-ol

A heavy stir bar and magnesium turnings (0.5 g, 20.8 mmol, excess) were added to a 100 mL RBF which was evacuated and filled with argon 3x. The solid was stirred vigorously for 30 min to break up the magnesium. The flask was charged with dry THF (35 mL), cooled to 0 °C, a small iodine crystal was added, and 5-bromopent-1-ene (607 mg, 5.03 mmol, 434 µl, 1.5 eq.) was added dropwise and stirred until the solution turned a metallic grey color and remained that color for about 30 minutes(ca. 1 hour total). The solution was cooled to -78 °C, crude 4-((tert-butyldimethylsilyl)oxy)butanal (610 mg, 3.01 mmol,

0.9 eq.) in THF (10 mL) was added dropwise. After a few hours and slight ambient warming, the solution was filtered through celite, quenched with NH4Cl (50 mL, 1/2 saturated or 200 g/L), THF was concentrated under vacuum, extracted with DCM (50 mL,

2x), washed with deionized water, dried over MgSO4, and concentrated under vacuum, yielding 1-((tert-butyldimethylsilyl)oxy)non-8-en-4-ol (630 mg, 70%, THP-K3-79)) as a crude oil that appeared pure via NMR. 1H NMR (300 MHz, Chloroform-d) δ 5.89 – 5.73

121

(m, 1H), 5.05 – 4.90 (m, 2H), 3.69 – 3.63 (m, 1H), 3.62 (s, 1H), 2.51 (d, J = 4.3 Hz, 1H),

2.12 – 2.03 (m, 2H), 1.71 – 1.53 (m, 4H), 1.53 – 1.40 (m, 3H), 0.90 (s, 9H), 0.07 (s, 6H).

Scheme 34. O- allylation to form ((4-(allyloxy)non-8-en-1-yl)oxy)(tert-

butyl)dimethylsilane

Purified 1-((tert-butyldimethylsilyl)oxy)non-8-en-4-ol (3.33 g, 12.2 mmol) was added

dropwise to a mixture of NaH (440 mg, 1.1 eq.) in dry DMF (100 mL) prepared by

rinsing and decanting NaH (734 mg, 70% in mineral oil) with dry pentanes (3x).

Hydrogen gas formation was continually flushed out with argon. After the solution

ceased to produce gas, allyl bromide (2.95, 24.4 mmol, 2 eq.) was added dropwise, and

heated to 50 °C for two hours. The solution was quenched with water (1 mL), diluted

with EtOAc (300 mL), washed with 5% LiCl solution (300 mL, 3x), dried over MgSO4, concentrated under vacuum , and subjected to two hours of high vacuum (to remove residual DMF). The resulting oil was dry loaded onto 20g of silica, and subjected to two hours of high vacuum. The silica was split into two portions, loaded on 100g Biotage LC columns, and purified via a toluene:pet ether gradient (40:60 – 80:20) to yield ((4-

(allyloxy)non-8-en-1-yl)oxy)(tert-butyl)dimethylsilane (1.8 g, 5.75 mmol, 47%, ) as a clear oil. 1H NMR (300 MHz, Chloroform-d) δ 6.03 – 5.69 (m, 2H), 5.32 – 4.87 (m, 4H),

3.97 (dt, J = 5.6, 1.4 Hz, 2H), 3.68 – 3.53 (m, 2H), 3.39 – 3.22 (m, 1H), 2.14 – 1.97 (m,

122

2H), 1.67 – 1.38 (m, 8H), 0.89 (s, 9H), 0.05 (s, 6H). 13C NMR (75 MHz, Chloroform-d) δ

138.99 , 135.72 , 116.51 , 114.61 , 78.81 , 69.97 , 63.43 , 34.03 , 33.50 , 30.16 , 28.76 ,

26.12 , 24.81 , 18.50 , -5.11 .

Scheme 35. Ring closing metathesis to form (Z)-tert-butyldimethyl(3-(3,4,5,8-tetrahydro-

2H-oxocin-2-yl)propoxy)silane

Grubbs I (26 mg, 0.032 mmol, 0.1 eq.) was added in one portion under positive argon pressure to a solution ((4-(allyloxy)non-8-en-1-yl)oxy)(tert-butyl)dimethylsilane (100 mg,

0.32 mmol) in dry toluene (150 mL, 2.15 mM) at 80 °C. The solution was stirred for 30 minutes, another 0.5 eq. of Grubbs I (13 mg, 0.016 mmol, 0.05 eq.) was added in one portion under positive argon pressure, stirred for 30 min, filtered through celite, concentrated under vacuum, and purified via Biotage LC to yield (Z)-tert- butyldimethyl(3-(3,4,5,8-tetrahydro-2H-oxocin-2-yl)propoxy)silane (68 mg, 75%) as a clear oil. Careful TLC (5% EtOAc, I2) is required to discern starting material from product and making custom capillaries is suggested. Owing to the challenge of separating starting material from product, it may be useful to proceed to bromination without purification. 1H NMR (300 MHz, Chloroform-d) δ 5.73 – 5.55 (m, 1H), 5.44 (dtd, J =

11.5, 2.9, 1.4 Hz, 1H), 4.47 – 4.32 (m, 1H), 3.98 (ddt, J = 17.4, 3.4, 1.7 Hz, 1H), 3.71 –

123

3.47 (m, 3H), 2.75 – 2.59 (m, 1H), 2.21 – 2.04 (m, 1H), 1.86 – 1.27 (m, 4H), 0.89 (s, 9H),

0.04 (s, 6H). 13C NMR (75 MHz, Chloroform-d) δ 130.00 , 128.22 , 80.29 , 68.59 , 63.42 ,

31.83 , 30.11 , 29.60 , 26.12 , 25.28 , 24.21 , 18.50 , -5.12 .

Scheme 36. Bromination to form tert-butyl(3-(6,7-dibromooxocan-2-yl)propoxy)- dimethylsilane

Bromine (147 mg, 1.84 mmol, 2 eq.) in DCM (1 mL) was added to crude (Z)-tert- butyldimethyl(3-(3,4,5,8-tetrahydro-2H-oxocin-2-yl)propoxy)silane (262 mg, 0.92 mmol) in DCM (10 mL) at 0 °C. Addition produced a reddish brown color that quickly disappeared upon stirring. Addition was continued until the reddish brown color persisted. The solution was extracted DCM (20 mL, 2x), washed with deionized water, dried over MgSO4, and concentrated under vacuum, yielding 422 mg of a crude oil.

Careful TLC with PMA stain was used to determine the primary product (5% EtOAc, 0.3

RF) and purification via Biotage LC yielded pure tert-butyl(3-(6,7-dibromooxocan-2-

yl)propoxy)dimethylsilane (278 mg, 68%) as a clear oil and mixture of diasteriomers. 1H

NMR (300 MHz, Chloroform-d) δ 4.60 (ddd, J = 7.8, 5.4, 2.6 Hz, 1H), 4.47 (td, J = 7.8,

3.6 Hz, 2H), 4.37 (td, J = 6.9, 3.2 Hz, 1H), 4.26 (dd, J = 13.0, 3.2 Hz, 1H), 4.11 – 3.89 (m,

2H), 3.77 – 3.39 (m, 6H), 2.79 – 2.58 (m, 2H), 2.39 – 2.10 (m, 2H), 2.00 – 1.35 (m, 16H),

0.89 (s, 18H), 0.04 (s, 12H). 13C NMR (75 MHz, Chloroform-d) δ 81.87 , 80.68 , 71.21 ,

124

67.51 , 63.30 , 63.09 , 59.84 , 58.78 , 58.56 , 56.07 , 33.97 , 33.72 , 33.52 , 33.22 , 31.46 ,

31.05 , 29.53 , 29.34 , 26.12 , 24.77 , 22.08 , 18.49 , -5.12 .

Scheme 37. Bromine elimination to form (E)-(3-(7-bromo-3,4,5,8-tetrahydro-2H-oxocin-

2-yl)propoxy)(tert-butyl)dimethylsilane

K-OtBu (50 mg, 0.45 mmol, 4 eq.) was added in one portion to a solution of tert-butyl(3-

(6,7-dibromooxocan-2-yl)propoxy)dimethylsilane (50 mg, 0.11 mmol), in dry THF (2 mL) at -5 °C in a brine bath. After 30 minutes stirring the solution was diluted with 50

mL DCM, washed 3x with deionized water, dried over MgSO4, and concentrated under

vacuum, purified via Biotage LC to yield (E)-(3-(7-bromo-3,4,5,8-tetrahydro-2H-oxocin-

2-yl)propoxy)(tert-butyl)dimethylsilane (36 mg, 90%, THP-K3-87, 2D NMR available)

as a clear oil. 1H NMR (300 MHz, Chloroform-d) δ 6.07 (t, J = 9.7, 8.0, 1.8 Hz, 1H),

4.43 (d, J = 17.3, 2.1, 0.9 Hz, 1H), 4.02 (dt, J = 17.4, 1.6 Hz, 1H), 3.72 – 3.52 (m, 4H),

2.72 (t, J = 11.3 Hz, 1H), 2.24 – 2.09 (m, 1H), 1.92 – 1.67 (m, 1H), 1.67 – 1.35 (m, 10H),

0.89 (s, 9H), 0.05 (s, 6H). 13C NMR (75 MHz, Chloroform-d) δ 129.79 , 124.47 , 81.50 ,

73.38 , 63.21 , 31.07 , 29.42 , 29.17 , 26.50 , 26.12 , 23.48 , 18.50 , -5.12 .

125

UV-Vis Kinetics; Fries Acid + Alkyne

Stock solutions of Fries Acid (1 mM) in MeCN and Alkyne (20 mM) in MeCN

were prepared. The Fries Acid stock (0.1 mL) was added to MeCN (0.8 mL) in a 1 mL

UV-vis cuvette and lastly an aliquot of the stock solution of Alkyne (0.1 mL) was added,

followed by a quick shake of the cuvette. The cuvette was immediately loaded into a

Cary 50 UV-vis spectrophotometer and measurements began. This experiment was

repeated at half alkyne concentrations (0.05 mL of 20 mM stock, 0.85 mL MeCN). UV- vis data was recorded at 453 nm with 6 second intervals and averaged at each time point.

SigmaPlot was utilized to plot data using exponential decay with the equation: f =

(-bx) y0+a(e ).

126

Figure 48. UV-Vis Kinetics; Fries Acid + DIBO in acetonitrrile.

127

Figure 49. UV-Vis Kinetics; Fries Acid + BCN in acetonitrile.

128

Figure 50. UV-Vis Kinetics; Fries Acid + COT in acetonitrile.

129

Figure 51. UV-Vis Kinetics; Fries Acid + DBCO in acetonittrrile.

130

Figure 52. UV-Vis Kinetics; Fries Acid + DIBOTTM in acetonitrile.

131

CONCLUSIONS AND FUTURE OUTLOOK

This work aims to develop fast and bioorthogonal traps to help delineate the importance of sulfenic acids and their parent proteins in complex cellular environments.

Chapter 1 serves as background and starts with the basics of oxidant generation and the deleterious consequences of cellular oxidation. The state of thiol literature, redox biology, and sulfenic acid detection are reviewed. The properties of current traps and probes are discussed as context for the development of new probes in Chapter 2.

As a proof of concept the strained alkyne known as bicyclo[6.1.0]nonyne (BCN) is shown to react with sulfenic acids at much higher rates than traditional 1, 3-dicarbonyl- based probes. BCN reacts with small molecule and protein based sulfenic acids with a high level of selectivity and limited cellular toxicity. BCN and BCN-Biotin conjugates are shown to be selective for sulfenic acids over thiols, s-nitrosothiols, sulfinic acids, and sulfonic acids. The enhanced rates and selectivity demonstrated by these strained alkynes identify them as new bioorthogonal probes that should facilitate the discovery of previously unknown sulfenic acid sites and their parent proteins.

The work of Chapter 3 entails efforts to improve upon strained alkyne based sulfenic acid traps. Selectivity considerations are discussed, including reactions with persulfides and thiols. The current computational literature on sulfenic acids and alkyl sulfoxide eliminations is reviewed. Computations are identified as a method to help narrow down the many combinations of strain and electronic activation. A number of computational methods are considered and compared to experimental results. The

132

B3LYP-6-31G(2d) theory correlated best with experimental data and was used to assess

barrier heights between sulfenic acids and hypothetical alkynes.

Computations were used to explore the relative effect of endocyclic and exocyclic

heteroatoms on stabilizing the transition state of sulfenic acids and alkynes.

Cyclooctynes with endocyclic propargylic oxygens were identified as ideal candidates for

accelerated sulfenic acid trapping. Significant progress was made towards the synthesis

of three unique cyclooctynes with endocyclic propargylic oxygens. The synthesis and

computations described in this work should prove valuable to continued research in this

area.

It is clear that strained alkynes offer a significant kinetic improvement over

traditional 1, 3-dicarbonyl- based probes. It is likely that tuning the reactivity of alkynes

by adjusting electronics and strain can yield fast and selective sulfenic acid probes.

Future research in this area would benefit from a full experimental and computational assessment of persulfide-alkyne and thiol-yne reactivity. However if necessary, the use of mass spectrum based analysis can differentiate between trapped persulfides, thiols, and sulfenic acids. The enhanced rates demonstrated by current and theoretical alkynes identify them as alternative probes that should facilitate the discovery of previously

unknown sulfenic acid sites and their parent proteins.

133

REFERENCES

(1) Park, M. H.; Jo, M.; Kim, Y. R.; Lee, C.-K.; Hong, J. T. Roles of Peroxiredoxins in Cancer, Neurodegenerative Diseases and Inflammatory Diseases. Pharmacol. Ther. 2016, 163, 1–23. (2) Netto, L. E. S.; Antunes, F. The Roles of Peroxiredoxin and Thioredoxin in Hydrogen Peroxide Sensing and in Signal Transduction. Mol. Cells 2016, 39 (1), 65–71. (3) Nickel, A.; Kohlhaas, M.; Maack, C. Mitochondrial Reactive Oxygen Species Production and Elimination. J. Mol. Cell. Cardiol. 2014, 73, 26–33. (4) Riemer, J.; Schwarzlander, M.; Conrad, M.; Herrmann, J. M. Thiol Switches in Mitochondria: Operation and Physiological Relevance. Biol. Chem. 2015, 396 (5), 465–482. (5) Poole, L. B. The Basics of Thiols and Cysteines in Redox Biology and Chemistry. Free Radic. Biol. Med. 2015, 0, 148–157. (6) Furdui, C. M.; Poole, L. B. Chemical Approaches to Detect and Analyze Protein Sulfenic Acids. Mass Spectrom Rev 2014, 33 (2), 126–146. (7) Klomsiri, C.; Nelson, K. J.; Bechtold, E.; Soito, L.; Johnson, L. C.; Lowther, W. T.; Ryu, S.-E.; King, S. B.; Furdui, C. M.; Poole, L. B. Use of Dimedone-Based Chemical Probes for Sulfenic Acid Detection: Evaluation of Conditions Affecting Probe Incorporation into Redox-Sensitive Proteins. Methods Enzymol. 2010, 473, 77–94. (8) Yin, H. Y.; Xu, L. B.; Porter, N. A. Free Radical Lipid Peroxidation: Mechanisms and Analysis. Chem. Rev. 2011, 111 (10), 5944–5972. (9) Pajares, M.; Jimenez-Moreno, N.; Dias, I. H. K.; Debelec, B.; Vucetic, M.; Fladmark, K. E.; Basaga, H.; Ribaric, S.; Milisav, I.; Cuadrado, A. Redox Control of Protein Degradation. Redox Biol. 2015, 6, 409–420. (10) Park, J.-G.; Oh, G. T. The Role of Peroxidases in the Pathogenesis of Atherosclerosis. BMB Rep. 2011, 44 (8), 497–505. (11) Valko, M.; Rhodes, C. J.; Moncol, J.; Izakovic, M.; Mazur, M. Free Radicals, Metals and in Oxidative Stress-Induced Cancer. Chem. Biol. Interact. 2006, 160 (1), 1–40. (12) Cadet, J.; Wagner, J. R.; Shafirovich, V.; Geacintov, N. E. One-Electron Oxidation Reactions of Purine and Pyrimidine Bases in Cellular DNA. Int. J. Radiat. Biol. 2014, 90 (6), 423–432.

134

(13) Cadet, J.; Wagner, J. R. DNA Base Damage by Reactive Oxygen Species, Oxidizing Agents, and UV Radiation. Cold Spring Harb Perspect Biol 2013, 5 (2). (14) Davies, M. J. Protein Oxidation and Peroxidation. Biochem. J. 2016, 473, 805–825. (15) Gebicki, S.; Gebicki, J. M. Formation of Peroxides in Amino-Acids and Proteins Exposed to Oxygen Free-Radicals. Biochem. J. 1993, 289, 743–749. (16) Chen, Y.-R.; Zweier, J. L. Cardiac Mitochondria and Reactive Oxygen Species Generation. Circ. Res. 2014, 114 (3), 524–537. (17) Kalogeris, T.; Bao, Y.; Korthuis, R. J. Mitochondrial Reactive Oxygen Species: A Double Edged Sword in Ischemia/reperfusion vs Preconditioning. Redox Biol. 2014, 2, 702–714. (18) Murphy, M. P. How Mitochondria Produce Reactive Oxygen Species. Biochem. J. 2009, 417, 1–13. (19) Barja, G. Mitochondrial Oxygen Radical Generation and Leak: Sites of Production in State 4 and 3, Organ Specificity, and Relation to Aging and Longevity. J. Bioenerg. Biomembr. 1999, 31 (4), 347–366. (20) Sheng, Y.; Abreu, I. A.; Cabelli, D. E.; Maroney, M. J.; Miller, A.-F.; Teixeira, M.; Valentine, J. S. Superoxide Dismutases and Superoxide Reductases. Chem. Rev. 2014, 114 (7), 3854–3918. (21) Turrens, J. F. Mitochondrial Formation of Reactive Oxygen Species. J Physiol 2003, 552 (Pt 2), 335–344. (22) Li, Y. B.; Huang, T. T.; Carlson, E. J.; Melov, S.; Ursell, P. C.; Olson, T. L.; Noble, L. J.; Yoshimura, M. P.; Berger, C.; Chan, P. H.; et al. Dilated Cardiomyopathy and Neonatal Lethality in Mutant Mice Lacking Manganese Superoxide-Dismutase. Nat. Genet. 1995, 11 (4), 376–381. (23) Lebovitz, R. M.; Zhang, H.; Vogel, H.; Cartwright, J., Jr.; Dionne, L.; Lu, N.; Huang, S.; Matzuk, M. M. Neurodegeneration, Myocardial Injury, and Perinatal Death in Mitochondrial Superoxide Dismutase-Deficient Mice. Proc Natl Acad Sci U A 1996, 93 (18), 9782–9787. (24) Flynn, J. M.; Melov, S. SOD2 in Mitochondrial Dysfunction and Neurodegeneration. Free Radic. Biol. Med. 2013, 62, 4–12. (25) Muller, F. L.; Song, W.; Liu, Y.; Chaudhuri, A.; Pieke-Dahl, S.; Strong, R.; Huang, T. T.; Epstein, C. J.; Roberts, L. J., 2nd; Csete, M.; et al. Absence of CuZn Superoxide Dismutase Leads to Elevated Oxidative Stress and Acceleration of Age- Dependent Skeletal Muscle Atrophy. Free Radic Biol Med 2006, 40 (11), 1993– 2004.

135

(26) Garratt, M.; Pichaud, N.; Glaros, E. N.; Kee, A. J.; Brooks, R. C. Superoxide Dismutase Deficiency Impairs Olfactory Sexual Signaling and Alters Bioenergetic Function in Mice. Proc Natl Acad Sci U A 2014, 111 (22), 8119–8124. (27) Rhee, S. G. Overview on Peroxiredoxin. Mol. Cells 2016, 39 (1), 1–5. (28) Winterbourn, C. C. Chapter One - The Biological Chemistry of Hydrogen Peroxide. In Methods in Enzymology; Packer, E. C. and L., Ed.; Hydrogen Peroxide and Cell Signaling, Part C; Academic Press, 2013; Vol. 528, pp 3–25. (29) Sies, H. Role of Metabolic H2O2 Generation REDOX SIGNALING AND OXIDATIVE STRESS. J. Biol. Chem. 2014, 289 (13), 8735–8741. (30) Marinho, H. S.; Real, C.; Cyrne, L.; Soares, H.; Antunes, F. Hydrogen Peroxide Sensing, Signaling and Regulation of Transcription Factors. Redox Biol. 2014, 2, 535–562. (31) Winterbourn, C. C.; Hampton, M. B. Thiol Chemistry and Specificity in Redox Signaling. Free Radic. Biol. Med. 2008, 45 (5), 549–561. (32) Cox, A. G.; Winterbourn, C. C.; Hampton, M. B. Mitochondrial Peroxiredoxin Involvement in Defence and Redox Signalling. Biochem. J. 2010, 425, 313–325. (33) Marchissio, M. J.; Francés, D. E. A.; Carnovale, C. E.; Marinelli, R. A. Mitochondrial Aquaporin-8 Knockdown in Human Hepatoma HepG2 Cells Causes ROS-Induced Mitochondrial Depolarization and Loss of Viability. Toxicol. Appl. Pharmacol. 2012, 264 (2), 246–254. (34) Wei, Q.; Jiang, H.; Xiao, Z.; Baker, A.; Young, M. R.; Veenstra, T. D.; Colburn, N. H. Sulfiredoxin–Peroxiredoxin IV Axis Promotes Human Lung Cancer Progression through Modulation of Specific Phosphokinase Signaling. Proc. Natl. Acad. Sci. U. S. A. 2011, 108 (17), 7004–7009. (35) Grimm, S.; Hoehn, A.; Davies, K. J.; Grune, T. Protein Oxidative Modifications in the Ageing Brain: Consequence for the Onset of Neurodegenerative Disease. Free Radic. Res. 2011, 45 (1), 73–88. (36) Hawkins, C. L.; Davies, M. J. Generation and Propagation of Radical Reactions on Proteins. Biochim. Biophys. Acta-Bioenerg. 2001, 1504 (2–3), 196–219. (37) Bartesaghi, S.; Wenzel, J.; Trujillo, M.; Lopez, M.; Joseph, J.; Kalyanaraman, B.; Radi, R. Lipid Peroxyl Radicals Mediate Tyrosine Dimerization and Nitration in Membranes. Chem. Res. Toxicol. 2010, 23 (4), 821–835. (38) Kujoth, G. C.; Hiona, A.; Pugh, T. D.; Someya, S.; Panzer, K.; Wohlgemuth, S. E.; Hofer, T.; Seo, A. Y.; Sullivan, R.; Jobling, W. A.; et al. Mitochondrial DNA Mutations, Oxidative Stress, and Apoptosis in Mammalian Aging. Science 2005, 309 (5733), 481–484.

136

(39) Wood, S. T.; Long, D. L.; Reisz, J. A.; Yammani, R. R.; Burke, E. A.; Klomsiri, C.; Poole, L. B.; Furdui, C. M.; Loeser, R. F. Cysteine-Mediated Redox Regulation of Cell Signaling in Chondrocytes Stimulated With Fibronectin Fragments. Arthritis Rheumatol. 2016, 68 (1), 117–126. (40) Rahman, K. Studies on Free Radicals, Antioxidants, and Co-Factors. Clin. Interv. Aging 2007, 2 (2), 219–236. (41) Ames, B. N. Dietary Carcinogens and Anticarcinogens. Science 1983, 221 (4617), 1256–1264. (42) Loft, S.; Poulsen, H. E. Cancer Risk and Oxidative DNA Damage in Man. J. Mol. Med. Berl. Ger. 1996, 74 (6), 297–312. (43) Waris, G.; Ahsan, H. Reactive Oxygen Species: Role in the Development of Cancer and Various Chronic Conditions. J. Carcinog. 2006, 5, 14. (44) Lunec, J.; Holloway, K. A.; Cooke, M. S.; Faux, S.; Griffiths, H. R.; Evans, M. D. Urinary 8-Oxo-2′-Deoxyguanosine: Redox Regulation of DNA Repair in vivo?1. Free Radic. Biol. Med. 2002, 33 (7), 875–885. (45) Du, M.-Q.; Carmichael, P. L.; Phillips, D. H. Induction of Activating Mutations in the Human c-Ha-Ras-1 Proto-Oncogene by Oxygen Free Radicals. Mol. Carcinog. 1994, 11 (3), 170–175. (46) Shen, Z.; Wu, W.; Hazen, S. L. Activated Leukocytes Oxidatively Damage DNA, RNA, and the Nucleotide Pool through Halide-Dependent Formation of Hydroxyl Radical. Biochemistry (Mosc.) 2000, 39 (18), 5474–5482. (47) Ames, B. N.; Shigenaga, M. K.; Gold, L. S. DNA Lesions, Inducible DNA Repair, and Cell Division: Three Key Factors in Mutagenesis and Carcinogenesis. Environ. Health Perspect. 1993, 101 (Suppl 5), 35–44. (48) Song, I.-S.; Kim, H.-K.; Jeong, S.-H.; Lee, S.-R.; Kim, N.; Rhee, B. D.; Ko, K. S.; Han, J. Mitochondrial Peroxiredoxin III Is a Potential Target for Cancer Therapy. Int. J. Mol. Sci. 2011, 12 (10), 7163–7185. (49) Cerutti, P. A. Oxy-Radicals and Cancer. The Lancet 1994, 344 (8926), 862–863. (50) Gogvadze, V.; Orrenius, S.; Zhivotovsky, B. Mitochondria in Cancer Cells: What Is so Special about Them? Trends Cell Biol. 2008, 18 (4), 165–173. (51) Chatterjee, S.; Feinstein, S. I.; Dodia, C.; Sorokina, E.; Lien, Y.-C.; Nguyen, S.; Debolt, K.; Speicher, D.; Fisher, A. B. Peroxiredoxin 6 Phosphorylation and Subsequent Phospholipase A2 Activity Are Required for Agonist-Mediated Activation of NADPH Oxidase in Mouse Pulmonary Microvascular Endothelium and Alveolar Macrophages. J. Biol. Chem. 2011, 286 (13), 11696–11706. (52) Cha, M.-K.; Suh, K.-H.; Kim, I.-H. Overexpression of Peroxiredoxin I and thioredoxin1 in Human Breast Carcinoma. J. Exp. Clin. Cancer Res. 2009, 28, 93.

137

(53) Woolston, C. M.; Storr, S. J.; Ellis, I. O.; Morgan, D. A. L.; Martin, S. G. Expression of Thioredoxin System and Related Peroxiredoxin Proteins Is Associated with Clinical Outcome in Radiotherapy Treated Early Stage Breast Cancer. Radiother. Oncol. 2011, 100 (2), 308–313. (54) Neumann, C. A.; Fang, Q. Are Peroxiredoxins Tumor Suppressors? Curr. Opin. Pharmacol. 2007, 7 (4), 375–380. (55) Egler, R. A.; Fernandes, E.; Rothermund, K.; Sereika, S.; de Souza-Pinto, N.; Jaruga, P.; Dizdaroglu, M.; Prochownik, E. V. Regulation of Reactive Oxygen Species, DNA Damage, and c-Myc Function by Peroxiredoxin 1. Oncogene 2005, 24 (54), 8038–8050. (56) Bont, R. D.; Larebeke, N. van. Endogenous DNA Damage in Humans: A Review of Quantitative Data. Mutagenesis 2004, 19 (3), 169–185. (57) FastStats http://www.cdc.gov/nchs/fastats/leading-causes-of-death.htm (accessed Oct 5, 2016). (58) Ross, R. Atherosclerosis--an Inflammatory Disease. N. Engl. J. Med. 1999, 340 (2), 115–126. (59) Libby, P. Inflammation in Atherosclerosis. Nature 2002, 420 (6917), 868–874. (60) Navab, M.; Ananthramaiah, G. M.; Reddy, S. T.; Lenten, B. J. V.; Ansell, B. J.; Fonarow, G. C.; Vahabzadeh, K.; Hama, S.; Hough, G.; Kamranpour, N.; et al. Thematic Review Series: The Pathogenesis of Atherosclerosis The Oxidation Hypothesis of Atherogenesis: The Role of Oxidized Phospholipids and HDL. J. Lipid Res. 2004, 45 (6), 993–1007. (61) Navab, M.; Berliner, J. A.; Watson, A. D.; Hama, S. Y.; Territo, M. C.; Lusis, A. J.; Shih, D. M.; Lenten, B. J. V.; Frank, J. S.; Demer, L. L.; et al. The Yin and Yang of Oxidation in the Development of the Fatty Streak. Arterioscler. Thromb. Vasc. Biol. 1996, 16 (7), 831–842. (62) Chen, M.; Masaki, T.; Sawamura, T. LOX-1, the Receptor for Oxidized Low- Density Lipoprotein Identified from Endothelial Cells: Implications in Endothelial Dysfunction and Atherosclerosis. Pharmacol. Ther. 2002, 95 (1), 89–100. (63) Mehta, J. L.; Chen, J.; Hermonat, P. L.; Romeo, F.; Novelli, G. Lectin-Like, Oxidized Low-Density Lipoprotein Receptor-1 (LOX-1): A Critical Player in the Development of Atherosclerosis and Related Disorders. Cardiovasc. Res. 2006, 69 (1), 36–45. (64) Glass, C. K.; Witztum, J. L. Atherosclerosis. Cell 2001, 104 (4), 503–516. (65) Takei, A.; Huang, Y.; Lopes-Virella, M. F. Expression of Adhesion Molecules by Human Endothelial Cells Exposed to Oxidized Low Density Lipoprotein. Influences of Degree of Oxidation and Location of Oxidized LDL. Atherosclerosis 2001, 154 (1), 79–86.

138

(66) Klouche, M.; May, A. E.; Hemmes, M.; Messner, M.; Kanse, S. M.; Preissner, K. T.; Bhakdi, S. Enzymatically Modified, Nonoxidized LDL Induces Selective Adhesion and Transmigration of Monocytes and T-Lymphocytes through Human Endothelial Cell Monolayers. Arterioscler. Thromb. Vasc. Biol. 1999, 19 (3), 784–793. (67) Khan, B. V.; Parthasarathy, S. S.; Alexander, R. W.; Medford, R. M. Modified Low Density Lipoprotein and Its Constituents Augment Cytokine-Activated Vascular Cell Adhesion Molecule-1 Gene Expression in Human Vascular Endothelial Cells. J. Clin. Invest. 1995, 95 (3), 1262–1270. (68) Shibata, N.; Glass, C. K. Regulation of Macrophage Function in Inflammation and Atherosclerosis. J. Lipid Res. 2009, 50 (Supplement), S277–S281. (69) Madamanchi, N. R.; Vendrov, A.; Runge, M. S. Oxidative Stress and Vascular Disease. Arterioscler. Thromb. Vasc. Biol. 2005, 25 (1), 29–38. (70) Kisucka, J.; Chauhan, A. K.; Patten, I. S.; Yesilaltay, A.; Neumann, C.; Van Etten, R. A.; Krieger, M.; Wagner, D. D. Peroxiredoxin1 Prevents Excessive Endothelial Activation and Early Atherosclerosis. Circ. Res. 2008, 103 (6), 598–605. (71) Mowbray, A. L.; Kang, D.-H.; Rhee, S. G.; Kang, S. W.; Jo, H. Laminar Shear Stress up-Regulates Peroxiredoxins (PRX) in Endothelial Cells: PRX 1 as a Mechanosensitive Antioxidant. J. Biol. Chem. 2008, 283 (3), 1622–1627. (72) Guo, X.; Yamada, S.; Tanimoto, A.; Ding, Y.; Wang, K.-Y.; Shimajiri, S.; Murata, Y.; Kimura, S.; Tasaki, T.; Nabeshima, A.; et al. Overexpression of Peroxiredoxin 4 Attenuates Atherosclerosis in Apolipoprotein E Knockout Mice. Antioxid. Redox Signal. 2012, 17 (10), 1362–1375. (73) Corral-Debrinski, M.; Horton, T.; Lott, M. T.; Shoffner, J. M.; Flint Beal, M.; Wallace, D. C. Mitochondrial DNA Deletions in Human Brain: Regional Variability and Increase with Advanced Age. Nat. Genet. 1992, 2 (4), 324–329. (74) Praticò, D.; Uryu, K.; Leight, S.; Trojanoswki, J. Q.; Lee, V. M. Increased Lipid Peroxidation Precedes Amyloid Plaque Formation in an Animal Model of Alzheimer Amyloidosis. J. Neurosci. Off. J. Soc. Neurosci. 2001, 21 (12), 4183– 4187. (75) Reddy, P. H.; McWeeney, S.; Park, B. S.; Manczak, M.; Gutala, R. V.; Partovi, D.; Jung, Y.; Yau, V.; Searles, R.; Mori, M.; et al. Gene Expression Profiles of Transcripts in Amyloid Precursor Protein Transgenic Mice: Up-Regulation of Mitochondrial Metabolism and Apoptotic Genes Is an Early Cellular Change in Alzheimer’s Disease. Hum. Mol. Genet. 2004, 13 (12), 1225–1240. (76) Sherer, T. B.; Betarbet, R.; Testa, C. M.; Seo, B. B.; Richardson, J. R.; Kim, J. H.; Miller, G. W.; Yagi, T.; Matsuno-Yagi, A.; Greenamyre, J. T. Mechanism of Toxicity in Rotenone Models of Parkinson’s Disease. J. Neurosci. Off. J. Soc. Neurosci. 2003, 23 (34), 10756–10764.

139

(77) Schapira, A. H.; Cooper, J. M.; Dexter, D.; Clark, J. B.; Jenner, P.; Marsden, C. D. Mitochondrial Complex I Deficiency in Parkinson’s Disease. J. Neurochem. 1990, 54 (3), 823–827. (78) Cimini, A.; Gentile, R.; Angelucci, F.; Benedetti, E.; Pitari, G.; Giordano, A.; Ippoliti, R. Neuroprotective Effects of PrxI over-Expression in an in Vitro Human Alzheimer’s Disease Model. J. Cell. Biochem. 2013, 114 (3), 708–715. (79) Fang, J.; Nakamura, T.; Cho, D.-H.; Gu, Z.; Lipton, S. A. S-Nitrosylation of Peroxiredoxin 2 Promotes Oxidative Stress-Induced Neuronal Cell Death in Parkinson’s Disease. Proc. Natl. Acad. Sci. 2007, 104 (47), 18742–18747. (80) Bach, R. D.; Su, M. D. Inductive Versus Coulombic Effects on the Barriers to Oxygen-Atom Transfer from Alkyl Hydroperoxides - Model Studies on 4-Alpha- Flavin Hydroperoxide. J. Am. Chem. Soc. 1994, 116 (12), 5392–5399. (81) Winterbourn, C. C.; Metodiewa, D. Reactivity of Biologically Important Thiol Compounds with Superoxide and Hydrogen Peroxide. Free Radic. Biol. Med. 1999, 27 (3–4), 322–328. (82) Hall, A.; Parsonage, D.; Poole, L. B.; Karplus, P. A. Structural Evidence That Peroxiredoxin Catalytic Power Is Based on Transition State Stabilization. J. Mol. Biol. 2010, 402 (1), 194–209. (83) Karplus, P. A. A Primer on Peroxiredoxin Biochemistry. Free Radic. Biol. Med. 2015, 80, 183–190. (84) Hofmann, B.; Hecht, H.-J.; Flohé, L. Peroxiredoxins. Biol. Chem. 2002, 383 (3–4), 347–364. (85) Parsonage, D.; Karplus, P. A.; Poole, L. B. Substrate Specificity and Redox Potential of AhpC, a Bacterial Peroxiredoxin. Proc. Natl. Acad. Sci. U. S. A. 2008, 105 (24), 8209–8214. (86) Peskin, A. V.; Low, F. M.; Paton, L. N.; Maghzal, G. J.; Hampton, M. B.; Winterbourn, C. C. The High Reactivity of Peroxiredoxin 2 with H2O2 Is Not Reflected in Its Reaction with Other Oxidants and Thiol Reagents. J. Biol. Chem. 2007, 282 (16), 11885–11892. (87) Adén, J.; Wallgren, M.; Storm, P.; Weise, C. F.; Christiansen, A.; Schröder, W. P.; Funk, C.; Wolf-Watz, M. Extraordinary Μs-Ms Backbone Dynamics in Arabidopsis Thaliana Peroxiredoxin Q. Biochim. Biophys. Acta 2011, 1814 (12), 1880–1890. (88) Lian, F.-M.; Yu, J.; Ma, X.-X.; Yu, X.-J.; Chen, Y.; Zhou, C.-Z. Structural Snapshots of Yeast Alkyl Hydroperoxide Reductase Ahp1 Peroxiredoxin Reveal a Novel Two-Cysteine Mechanism of Electron Transfer to Eliminate Reactive Oxygen Species. J. Biol. Chem. 2012, 287 (21), 17077–17087.

140

(89) Jönsson, T. J.; Ellis, H. R.; Poole, L. B. Cysteine Reactivity and Thiol-Disulfide Interchange Pathways in AhpF and AhpC of the Bacterial Alkyl Hydroperoxide Reductase System. Biochemistry (Mosc.) 2007, 46 (19), 5709–5721. (90) Poole, L. B. Bacterial Defenses against Oxidants: Mechanistic Features of Cysteine- Based Peroxidases and Their Flavoprotein Reductases. Arch. Biochem. Biophys. 2005, 433 (1), 240–254. (91) Sarma, G. N.; Nickel, C.; Rahlfs, S.; Fischer, M.; Becker, K.; Karplus, P. A. Crystal Structure of a Novel Plasmodium Falciparum 1-Cys Peroxiredoxin. J. Mol. Biol. 2005, 346 (4), 1021–1034. (92) Peskin, A. V.; Dickerhof, N.; Poynton, R. A.; Paton, L. N.; Pace, P. E.; Hampton, M. B.; Winterbourn, C. C. Hyperoxidation of Peroxiredoxins 2 and 3 RATE CONSTANTS FOR THE REACTIONS OF THE SULFENIC ACID OF THE PEROXIDATIC CYSTEINE. J. Biol. Chem. 2013, 288 (20), 14170–14177. (93) Wood, Z. A.; Poole, L. B.; Karplus, P. A. Peroxiredoxin Evolution and the Regulation of Hydrogen Peroxide Signaling. Science 2003, 300 (5619), 650–653. (94) Nelson, K. J.; Klomsiri, C.; Codreanu, S. G.; Soito, L.; Liebler, D. C.; Rogers, L. C.; Daniel, L. W.; Poole, L. B. Use of Dimedone-Based Chemical Probes for Sulfenic Acid Detection: Methods to Visualize and Identify Labeled Proteins. Methods Enzymol. 2010, 473. (95) Jönsson, T. J.; Johnson, L. C.; Lowther, W. T. Structure of the Sulphiredoxin– peroxiredoxin Complex Reveals an Essential Repair Embrace. Nature 2008, 451 (7174), 98–101. (96) Lowther, W. T.; Haynes, A. C. Reduction of Cysteine Sulfinic Acid in Eukaryotic, Typical 2-Cys Peroxiredoxins by Sulfiredoxin. Antioxid. Redox Signal. 2011, 15 (1), 99–109. (97) Winterbourn, C. C. Reconciling the Chemistry and Biology of Reactive Oxygen Species. Nat. Chem. Biol. 2008, 4 (5), 278–286. (98) Wadley, A. J.; Aldred, S.; Coles, S. J. An Unexplored Role for Peroxiredoxin in Exercise-Induced Redox Signalling? Redox Biol. 2015, 8, 51–58. (99) Netto, L. E. S.; Oliveira, M. A. de; Tairum, C. A.; Neto, J. F. da S. Conferring Specificity in Redox Pathways by Enzymatic Thiol/disulfide Exchange Reactions. Free Radic. Res. 2016, 50 (2), 206–245. (100) Meng, T.-C.; Fukada, T.; Tonks, N. K. Reversible Oxidation and Inactivation of Protein Tyrosine Phosphatases In Vivo. Mol. Cell 2002, 9 (2), 387–399. (101) Kwon, J.; Lee, S.-R.; Yang, K.-S.; Ahn, Y.; Kim, Y. J.; Stadtman, E. R.; Rhee, S. G. Reversible Oxidation and Inactivation of the Tumor Suppressor PTEN in Cells Stimulated with Peptide Growth Factors. Proc. Natl. Acad. Sci. U. S. A. 2004, 101 (47), 16419–16424.

141

(102) Rhee, S. G.; Woo, H. A. Multiple Functions of Peroxiredoxins: Peroxidases, Sensors and Regulators of the Intracellular Messenger H2O2, and Protein Chaperones. Antioxid. Redox Signal. 2010, 15 (3), 781–794. (103) Rawat, S. J.; Creasy, C. L.; Peterson, J. R.; Chernoff, J. The Tumor Suppressor Mst1 Promotes Changes in the Cellular Redox State by Phosphorylation and Inactivation of Peroxiredoxin-1 Protein. J. Biol. Chem. 2013, 288 (12), 8762–8771. (104) Woo, H. A.; Yim, S. H.; Shin, D. H.; Kang, D.; Yu, D.-Y.; Rhee, S. G. Inactivation of Peroxiredoxin I by Phosphorylation Allows Localized H2O2 Accumulation for Cell Signaling. Cell 2010, 140 (4), 517–528. (105) Chen, K.; Kirber, M. T.; Xiao, H.; Yang, Y.; Keaney, J. F. Regulation of ROS Signal Transduction by NADPH Oxidase 4 Localization. J. Cell Biol. 2008, 181 (7), 1129–1139. (106) Paulsen, C. E.; Truong, T. H.; Garcia, F. J.; Homann, A.; Gupta, V.; Leonard, S. E.; Carroll, K. S. Peroxide-Dependent Sulfenylation of the EGFR Catalytic Site Enhances Kinase Activity. Nat. Chem. Biol. 2012, 8 (1), 57–64. (107) Brito, P. M.; Antunes, F. Estimation of Kinetic Parameters Related to Biochemical Interactions between Hydrogen Peroxide and Signal Transduction Proteins. Front. Chem. 2014, 2. (108) Delaunay, A.; Pflieger, D.; Barrault, M.-B.; Vinh, J.; Toledano, M. B. A Thiol Peroxidase Is an H2O2 Receptor and Redox-Transducer in Gene Activation. Cell 2002, 111 (4), 471–481. (109) Boronat, S.; Domènech, A.; Paulo, E.; Calvo, I. A.; García-Santamarina, S.; García, P.; Encinar del Dedo, J.; Barcons, A.; Serrano, E.; Carmona, M.; et al. Thiol-Based H2O2 Signalling in Microbial Systems. Redox Biol. 2014, 2, 395–399. (110) Fomenko, D. E.; Koc, A.; Agisheva, N.; Jacobsen, M.; Kaya, A.; Malinouski, M.; Rutherford, J. C.; Siu, K.-L.; Jin, D.-Y.; Winge, D. R.; et al. Thiol Peroxidases Mediate Specific Genome-Wide Regulation of Gene Expression in Response to Hydrogen Peroxide. Proc. Natl. Acad. Sci. 2011, 108 (7), 2729–2734. (111) Gutscher, M.; Sobotta, M. C.; Wabnitz, G. H.; Ballikaya, S.; Meyer, A. J.; Samstag, Y.; Dick, T. P. Proximity-Based Protein Thiol Oxidation by H2O2- Scavenging Peroxidases. J. Biol. Chem. 2009, 284 (46), 31532–31540. (112) Chae, H. Z.; Chung, S. J.; Rhee, S. G. Thioredoxin-Dependent Peroxide Reductase from Yeast. J. Biol. Chem. 1994, 269 (44), 27670–27678. (113) Saitoh, M.; Nishitoh, H.; Fujii, M.; Takeda, K.; Tobiume, K.; Sawada, Y.; Kawabata, M.; Miyazono, K.; Ichijo, H. Mammalian Thioredoxin Is a Direct Inhibitor of Apoptosis Signal-Regulating Kinase (ASK) 1. EMBO J. 1998, 17 (9), 2596–2606.

142

(114) Calvo, I. A.; Boronat, S.; Domènech, A.; García-Santamarina, S.; Ayté, J.; Hidalgo, E. Dissection of a Redox Relay: H2O2-Dependent Activation of the Transcription Factor Pap1 through the Peroxidatic Tpx1-Thioredoxin Cycle. Cell Rep. 2013, 5 (5), 1413–1424. (115) Brown, J. D.; Day, A. M.; Taylor, S. R.; Tomalin, L. E.; Morgan, B. A.; Veal, E. A. A Peroxiredoxin Promotes H2O2 Signaling and Oxidative Stress Resistance by Oxidizing a Thioredoxin Family Protein. Cell Rep. 2013, 5 (5), 1425–1435. (116) Roos, G.; Messens, J. Protein Sulfenic Acid Formation: From Cellular Damage to Redox Regulation. Free Radic. Biol. Med. 2011, 51 (2), 314–326. (117) Gupta, V.; Carroll, K. S. Sulfenic Acid Chemistry, Detection and Cellular Lifetime. Biochim. Biophys. Acta BBA - Gen. Subj. 2014, 1840 (2), 847–875. (118) Fries, K. 1912, 45, 2965. (119) Fraenkel-Conrat, H. The Reaction of Tobacco Mosaic Virus with Iodine. J Biol Chem 1955, 217 (1), 373–381. (120) Pihl, A.; Lange, R. The Interaction of Oxidized Glutathione, Cystamine Monosulfoxide, and Tetrathionate with the-SH Groups of Rabbit Muscle D- Glyceraldehyde 3-Phosphate Dehydrogenase. J Biol Chem 1962, 237, 1356–1362. (121) Martínez-Acedo, P.; Gupta, V.; Carroll, K. S. Proteomic Analysis of Peptides Tagged with Dimedone and Related Probes. J. Mass Spectrom. 2014, 49 (4), 257– 265. (122) Bruice, T. C.; Sayigh, A. B. The Structure of Anthraquinone-1-Sulfenic Acid (Fries’ Acid) and Related Compounds. J. Am. Chem. Soc. 1959, 13 (81), 3416–3420. (123) McGrath, A. J.; Garrett, G. E.; Valgimigli, L.; Pratt, D. A. The Redox Chemistry of Sulfenic Acids. J. Am. Chem. Soc. 2010, 132 (47), 16759–16761. (124) Ellis, H. R.; Poole, L. B. Roles for the Two Cysteine Residues of AhpC in Catalysis of Peroxide Reduction by Alkyl Hydroperoxide Reductase from Salmonella Typhimurium. Biochemistry (Mosc.) 1997, 36 (43), 13349–13356. (125) Wood, Z. A.; Schroder, E.; Robin Harris, J.; Poole, L. B. Structure, Mechanism and Regulation of Peroxiredoxins. Trends Biochem Sci 2003, 28 (1), 32–40. (126) Poole, L. B.; Hall, A.; Nelson, K. J. Overview of Peroxiredoxins in Oxidant Defense and Redox Regulation. 2011. (127) Ellis, H. R.; Poole, L. B. Novel Application of 7-Chloro-4-Nitrobenzo-2-Oxa-1,3- Diazole to Identify Cysteine Sulfenic Acid in the AhpC Component of Alkyl Hydroperoxide Reductase. Biochemistry (Mosc.) 1997, 36 (48), 15013–15018.

143

(128) Benitez, L. V.; Allison, W. S. The Inactivation of the Acyl Phosphatase Activity Catalyzed by the Sulfenic Acid Form of Glyceraldehyde 3-Phosphate Dehydrogenase by Dimedone and Olefins. J Biol Chem 1974, 249 (19), 6234–6243. (129) Allison, W. S. Formation and Reactions of Sulfenic Acids in Proteins. Acc. Chem. Res. 1976, 9, 293. (130) Freeman, F. Mechanism of the Cysteine Sulfenic Acid O-Sulfenylation of 1,3- Cyclohexanedione. Chem. Commun. 2014, 50 (31), 4102. (131) Wagner, G. Mechanistic Aspects of the Reaction of Dimedone Derivatives with Sulfenic Acids and Other Sulfur Compounds—a Computational Study. Tetrahedron 2013, 69 (35), 7243–7252. (132) Poole, L. B.; Zeng, B. B.; Knaggs, S. A.; Yakubu, M.; King, S. B. Synthesis of Chemical Probes to Map Sulfenic Acid Modifications on Proteins. Bioconjug Chem 2005, 16 (6), 1624–1628. (133) Poole, L. B.; Klomsiri, C.; Knaggs, S. A.; Furdui, C. M.; Nelson, K. J.; Thomas, M. J.; Fetrow, J. S.; Daniel, L. W.; King, S. B. Fluorescent and Affinity-Based Tools To Detect Cysteine Sulfenic Acid Formation in Proteins. Bioconjug. Chem. 2007, 18 (6), 2004–2017. (134) Qian, J.; Klomsiri, C.; Wright, M. W.; King, S. B.; Tsang, A. W.; Poole, L. B.; Furdui, C. M. Simple Synthesis of 1,3-Cyclopentanedione Derived Probes for Labeling Sulfenic Acid Proteins. Chem. Commun. 2011, 47 (32), 9203–9205. (135) Klomsiri, C.; Rogers, L. C.; Soito, L.; McCauley, A. K.; King, S. B.; Nelson, K. J.; Poole, L. B.; Daniel, L. W. Endosomal H2O2 Production Leads to Localized Cysteine Sulfenic Acid Formation on Proteins during Lysophosphatidic Acid- Mediated Cell Signaling. Free Radic. Biol. Med. 2014, 71, 49–60. (136) Qian, J.; Wani, R.; Klomsiri, C.; Poole, L. B.; Tsang, A. W.; Furdui, C. M. A Simple and Effective Strategy for Labeling Cysteine Sulfenic Acid in Proteins by Utilization of Beta-Ketoesters as Cleavable Probes. Chem. Commun. 2012, 48 (34), 4091–4093. (137) Yang, J.; Gupta, V.; Carroll, K. S.; Liebler, D. C. Site-Specific Mapping and Quantification of Protein S-Sulphenylation in Cells. Nat. Commun. 2014, 5. (138) Cysteine Protein Oxidation - KeraFAST https://www.kerafast.com/c-301-protein- cysteine-oxidation.aspx (accessed Oct 10, 2016). (139) Gupta, V.; Carroll, K. S. Rational Design of Reversible and Irreversible Cysteine Sulfenic Acid-Targeted Linear C-Nucleophiles. Chem. Commun. 2016, 52 (16), 3414–3417. (140) Gupta, V.; Carroll, K. S. Profiling the Reactivity of Cyclic C-Nucleophiles towards Electrophilic Sulfur in Cysteine Sulfenic Acid. Chem. Sci. 2016, 7 (1), 400– 415.

144

(141) Poole, L. B. Measurement of Protein Sulfenic Acid Content. Curr Protoc Toxicol 2008, Chapter 17, Unit17 2. (142) Poole, L. B.; Ellis, H. R. Identification of Cysteine Sulfenic Acid in AhpC of Alkyl Hydroperoxide Reductase. Methods Enzymol. 2002, 348, 122–136. (143) Cubbage, J. W.; Guo, Y.; McCulla, R. D.; Jenks, W. S. Thermolysis of Alkyl Sulfoxides and Derivatives: A Comparison of Experiment and Theory. J Org Chem 2001, 66 (26), 8722–8736. (144) Nakamura, N. A Stable Sulfenic Acid, 9-Triptycenesulfenic Acid - Its Isolation and Characterization. J. Am. Chem. Soc. 1983, 105 (24), 7172–7173. (145) Aversa, M. C.; Barattucci, A.; Bonaccorsi, P.; Giannetto, P. L-Cysteine, a Versatile Source of Sulfenic Acids. Synthesis of Enantiopure Analogues. J. Org. Chem. 2005, 70 (6), 1986–1992. (146) Agard, N. J.; Prescher, J. A.; Bertozzi, C. R. A Strain-Promoted [3+2] Azide- Alkyne Cycloaddition for Covalent Modification of Blomolecules in Living Systems. J. Am. Chem. Soc. 2004, 126 (46), 15046–15047. (147) Poole, T. H.; Reisz, J. A.; Zhao, W.; Poole, L. B.; Furdui, C. M.; King, S. B. Strained Cycloalkynes as New Protein Sulfenic Acid Traps. J Am Chem Soc 2014, 136 (17), 6167–6170. (148) Poole, L. B.; Karplus, P. A.; Claiborne, A. Protein Sulfenic Acids in Redox Signaling. Annu. Rev. Pharmacol. Toxicol. 2004, 44 (1), 325–347. (149) Reddie, K. G.; Carroll, K. S. Expanding the Functional Diversity of Proteins through Cysteine Oxidation. Curr. Opin. Chem. Biol. 2008, 12 (6), 746–754. (150) Paulsen, C. E.; Carroll, K. S. Orchestrating Redox Signaling Networks through Regulatory Cysteine Switches. ACS Chem. Biol. 2010, 5 (1), 47–62. (151) Poole, L. B.; Nelson, K. J. Discovering Mechanisms of Signaling-Mediated Cysteine Oxidation. Curr. Opin. Chem. Biol. 2008, 12 (1), 18–24. (152) Wilson, M. R.; Taylor, R. E. Strained Alkenes in Natural Product Synthesis. Angew. Chem. Int. Ed. 2013, 52 (15), 4078–4087. (153) Jewett, J. C.; Bertozzi, C. R. Cu-Free Click Cycloaddition Reactions in Chemical Biology. Chem. Soc. Rev. 2010, 39 (4), 1272. (154) Dommerholt, J.; Schmidt, S.; Temming, R.; Hendriks, L. J. A.; Rutjes, F. P. J. T.; van Hest, J. C. M.; Lefeber, D. J.; Friedl, P.; van Delft, F. L. Readily Accessible Bicyclononynes for Bioorthogonal Labeling and Three-Dimensional Imaging of Living Cells. Angew. Chem. Int. Ed. 2010, 49 (49), 9422–9425.

145

(155) Aversa, M. C.; Barattucci, A.; Bonaccorsi, P.; Contini, A. Addition of Sulfenic Acids to Monosubstituted Acetylenes: A Theoretical and Experimental Study. J. Phys. Org. Chem. 2009, 22 (11), 1048–1057. (156) Royzen, M.; Yap, G. P. A.; Fox, J. M. A Photochemical Synthesis of Functionalized Trans-Cyclooctenes Driven by Metal Complexation. J. Am. Chem. Soc. 2008, 130 (12), 3760–3761. (157) Blackman, M. L.; Royzen, M.; Fox, J. M. Tetrazine Ligation: Fast Bioconjugation Based on Inverse-Electron-Demand Diels-Alder Reactivity. J. Am. Chem. Soc. 2008, 130, 13518–13519. (158) Royzen, M.; Yap, G. P. A.; Fox, J. M. A Photochemical Synthesis of Functionalized Trans -Cyclooctenes Driven by Metal Complexation. J. Am. Chem. Soc. 2008, 130 (12), 3760–3761. (159) Liu, C. T.; Benkovic, S. J. Capturing a Sulfenic Acid with Arylboronic Acids and Benzoxaborole. J. Am. Chem. Soc. 2013, 135 (39), 14544–14547. (160) van Geel, R.; Pruijn, G. J. M.; van Delft, F. L.; Boelens, W. C. Preventing Thiol- Yne Addition Improves the Specificity of Strain-Promoted Azide–Alkyne Cycloaddition. Bioconjug. Chem. 2012, 23 (3), 392–398. (161) Talipov, M. R.; Khomyakov, D. G.; Xian, M.; Timerghazin, Q. K. Computational Design of S-Nitrosothiol “click” Reactions. J. Comput. Chem. 2013, 34 (18), 1527– 1530. (162) Cavero, M.; Motherwell, W. B.; Potier, P. Studies on the Intermolecular Free Radical Addition of Thionitrites to Alkenes: A Convenient Method for the Preparation of α-Tritylthio and Related Derivatives. Tetrahedron Lett. 2001, 42 (26), 4377–4379. (163) Poole, L. B.; Klomsiri, C.; Knaggs, S. A.; Furdui, C. M.; Nelson, K. J.; Thomas, M. J.; Fetrow, J. S.; Daniel, L. W.; King, S. B. Fluorescent and Affinity-Based Tools To Detect Cysteine Sulfenic Acid Formation in Proteins. Bioconjug. Chem. 2007, 18 (6), 2004–2017. (164) Hart, T. W. Some Observations Concerning the S-Nitroso and S-Phenylsulphonyl Derivatives of L-Cysteine and Glutathione. Tetrahedron Lett. 1985, 26 (16), 2013– 2016. (165) Poole, L. B.; Ellis, H. R. Flavin-Dependent Alkyl Hydroperoxide Reductase from Salmonella Typhimurium. 1. Purification and Enzymatic Activities of Overexpressed AhpF and AhpC Proteins. Biochemistry (Mosc.) 1996, 35 (1), 56–64. (166) Bechtold, E.; Reisz, J. A.; Klomsiri, C.; Tsang, A. W.; Wright, M. W.; Poole, L. B.; Furdui, C. M.; King, S. B. Water-Soluble Triarylphosphines as Biomarkers for Protein S-Nitrosation. ACS Chem. Biol. 2010, 5 (4), 405–414.

146

(167) Griscelli, F.; Li, H.; Cheong, C.; Opolon, P.; Bennaceur-Griscelli, A.; Vassal, G.; Soria, J.; Soria, C.; Lu, H.; Perricaudet, M.; et al. Combined Effects of Radiotherapy and Angiostatin Gene Therapy in Glioma Tumor Model. Proc. Natl. Acad. Sci. U. S. A. 2000, 97 (12), 6698–6703. (168) Zhao, W.; Kridel, S.; Thorburn, A.; Kooshki, M.; Little, J.; Hebbar, S.; Robbins, M. Fatty Acid Synthase: A Novel Target for Antiglioma Therapy. Br. J. Cancer 2006, 95 (7), 869–878. (169) Ekkebus, R.; van Kasteren, S. I.; Kulathu, Y.; Scholten, A.; Berlin, I.; Geurink, P. P.; de Jong, A.; Goerdayal, S.; Neefjes, J.; Heck, A. J. R.; et al. On Terminal Alkynes That Can React with Active-Site Cysteine Nucleophiles in Proteases. J. Am. Chem. Soc. 2013, 135 (8), 2867–2870. (170) Galardon, E.; Padovani, D. Reactivity of Persulfides Toward Strained Bicyclo[6.1.0]nonyne Derivatives: Relevance to Chemical Tagging of Proteins. Bioconjug. Chem. 2015, 26 (6), 1013–1016. (171) Sun, Y.; Ma, X.; Cheng, K.; Wu, B.; Duan, J.; Chen, H.; Bu, L.; Zhang, R.; Hu, X.; Deng, Z.; et al. Strained Cyclooctyne as a Molecular Platform for Construction of Multimodal Imaging Probes. Angew. Chem. Int. Ed. 2015, 54 (20), 5981–5984. (172) Hoogenboom, R. Thiol–Yne Chemistry: A Powerful Tool for Creating Highly Functional Materials. Angew. Chem. Int. Ed. 2010, 49 (20), 3415–3417. (173) Lowe, A. B.; Hoyle, C. E.; Bowman, C. N. Thiol-Yne Click Chemistry: A Powerful and Versatile Methodology for Materials Synthesis. J. Mater. Chem. 2010, 20 (23), 4745–4750. (174) Hoyle, C. E.; Lowe, A. B.; Bowman, C. N. Thiol-Click Chemistry: A Multifaceted Toolbox for Small Molecule and Polymer Synthesis. Chem. Soc. Rev. 2010, 39 (4), 1355. (175) Konkolewicz, D.; Gray-Weale, A.; Perrier, S. Hyperbranched Polymers by Thiol−Yne Chemistry: From Small Molecules to Functional Polymers. J. Am. Chem. Soc. 2009, 131 (50), 18075–18077. (176) Fairbanks, B. D.; Scott, T. F.; Kloxin, C. J.; Anseth, K. S.; Bowman, C. N. Thiol−Yne Photopolymerizations: Novel Mechanism, Kinetics, and Step-Growth Formation of Highly Cross-Linked Networks. Macromolecules 2009, 42 (1), 211– 217. (177) Dondoni, A. The Emergence of Thiol–Ene Coupling as a Click Process for Materials and Bioorganic Chemistry. Angew. Chem. Int. Ed. 2008, 47 (47), 8995– 8997. (178) Fairbanks, B. D.; Sims, E. A.; Anseth, K. S.; Bowman, C. N. Reaction Rates and Mechanisms for Radical, Photoinitated Addition of Thiols to Alkynes, and

147

Implications for Thiol−Yne Photopolymerizations and Click Reactions. Macromolecules 2010, 43 (9), 4113–4119. (179) Beatty, K. E.; Fisk, J. D.; Smart, B. P.; Lu, Y. Y.; Szychowski, J.; Hangauer, M. J.; Baskin, J. M.; Bertozzi, C. R.; Tirrell, D. A. Live-Cell Imaging of Cellular Proteins by a Strain-Promoted Azide–Alkyne Cycloaddition. ChemBioChem 2010, 11 (15), 2092–2095. (180) Trujillo, M.; Alvarez, B.; Radi, R. One- and Two-Electron Oxidation of Thiols: Mechanisms, Kinetics and Biological Fates. Free Radic. Res. 2016, 50 (2), 150–171. (181) Schöneich, C. Thiyl Radicals and Induction of Protein Degradation. Free Radic. Res. 2016, 50 (2), 143–149. (182) Winterbourn, C. C. Are Free Radicals Involved in Thiol-Based Redox Signaling? Free Radic. Biol. Med. 2015, 80, 164–170. (183) Cuevasanta, E.; Möller, M. N.; Alvarez, B. Biological Chemistry of Hydrogen and Persulfides. Arch. Biochem. Biophys. (184) Tian, H.; Sakmar, T. P.; Huber, T. A Simple Method for Enhancing the Bioorthogonality of Cyclooctyne Reagent. Chem Commun 2016, 52 (31), 5451– 5454. (185) Pan, J.; Carroll, K. S. Persulfide Reactivity in the Detection of Protein S- Sulfhydration. ACS Chem. Biol. 2013, 8 (6), 1110–1116. (186) Dóka, É.; Pader, I.; Bíró, A.; Johansson, K.; Cheng, Q.; Ballagó, K.; Prigge, J. R.; Pastor-Flores, D.; Dick, T. P.; Schmidt, E. E.; et al. A Novel Persulfide Detection Method Reveals Protein Persulfide- and Polysulfide-Reducing Functions of Thioredoxin and Glutathione Systems. Sci. Adv. 2016, 2 (1), e1500968. (187) Zhang, D.; Devarie-Baez, N. O.; Li, Q.; Lancaster, J. R.; Xian, M. Methylsulfonyl Benzothiazole (MSBT): A Selective Protein Thiol Blocking Reagent. Org. Lett. 2012, 14 (13), 3396–3399. (188) Zhang, D.; Macinkovic, I.; Devarie-Baez, N. O.; Pan, J.; Park, C.-M.; Carroll, K. S.; Filipovic, M. R.; Xian, M. Detection of Protein S-Sulfhydration by a Tag-Switch Technique. Angew. Chem. Int. Ed. 2014, 53 (2), 575–581. (189) Reisz, J. A.; Bechtold, E.; King, S. B.; Poole, L. B.; Furdui, C. M. Thiol-Blocking Interfere with Labeling and Detection of Protein Sulfenic Acids. FEBS J. 2013, 280 (23), 6150–6161. (190) Behshad, E.; Parkin, S. E.; Bollinger, J. M. Mechanism of Cysteine Desulfurase Slr0387 from Synechocystis Sp. PCC 6803: Kinetic Analysis of Cleavage of the Persulfide Intermediate by Chemical Reductants †. Biochemistry (Mosc.) 2004, 43 (38), 12220–12226.

148

(191) Dommerholt, J.; van Rooijen, O.; Borrmann, A.; Guerra, C. F.; Bickelhaupt, F. M.; van Delft, F. L. Highly Accelerated Inverse Electron-Demand Cycloaddition of Electron-Deficient Azides with Aliphatic Cyclooctynes. Nat. Commun. 2014, 5, 5378. (192) Debets, M. F.; van Berkel, S. S.; Dommerholt, J.; Dirks, A. T.; Rutjes, F. P.; van Delft, F. L. Bioconjugation with Strained Alkenes and Alkynes. Acc Chem Res 2011, 44 (9), 805–815. (193) Gold, B.; Dudley, G. B.; Alabugin, I. V. Moderating Strain without Sacrificing Reactivity: Design of Fast and Tunable Noncatalyzed Alkyne–Azide Cycloadditions via Stereoelectronically Controlled Transition State Stabilization. J. Am. Chem. Soc. 2013, 135 (4), 1558–1569. (194) Tummatorn, J.; Batsomboon, P.; Clark, R. J.; Alabugin, I. V.; Dudley, G. B. Strain-Promoted Azide–Alkyne Cycloadditions of Benzocyclononynes. J. Org. Chem. 2012, 77 (5), 2093–2097. (195) Garcia-Hartjes, J.; Dommerholt, J.; Wennekes, T.; van Delft, F. L.; Zuilhof, H. Electronic Effects versus Distortion Energies During Strain-Promoted Alkyne- Azide Cycloadditions: A Theoretical Tool to Predict Reaction Kinetics: Strain- Promoted Alkyne-Azide Cycloadditions. Eur. J. Org. Chem. 2013, 2013 (18), 3712–3720. (196) Bach, R. D. Ring Strain Energy in the Cyclooctyl System. The Effect of Strain Energy on [3+2] Cycloaddition Reactions with Azides. J. Am. Chem. Soc. 2009, 131 (14), 5233–5243. (197) de Almeida, G.; Townsend, L. C.; Bertozzi, C. R. Synthesis and Reactivity of Dibenzoselenacycloheptynes. Org. Lett. 2013, 15 (12), 3038–3041. (198) de Almeida, G.; Sletten, E. M.; Nakamura, H.; Palaniappan, K. K.; Bertozzi, C. R. Thiacycloalkynes for Copper-Free Click Chemistry. Angew. Chem. Int. Ed. 2012, 51 (10), 2443–2447. (199) CCCBDB Experimental bond lengths http://cccbdb.nist.gov/expbondlengths1x.asp (accessed Nov 15, 2016). (200) Ning, X.; Guo, J.; Wolfert, M. A.; Boons, G.-J. Visualizing Metabolically- Labeled Glycoconjugates of Living Cells by Copper-Free and Fast Huisgen Cycloadditions. Angew. Chem. Int. Ed Engl. 2008, 47 (12), 2253–2255. (201) Debets, M. F.; Berkel, S. S. van; Schoffelen, S.; Rutjes, F. P. J. T.; Hest, J. C. M. van; Delft, F. L. van. Aza-Dibenzocyclooctynes for Fast and Efficient Enzyme PEGylation via Copper-Free (3+2) Cycloaddition. Chem. Commun. 2010, 46 (1), 97–99.

149

(202) Jewett, J. C.; Sletten, E. M.; Bertozzi, C. R. Rapid Cu-Free Click Chemistry with Readily Synthesized Biarylazacyclooctynones. J. Am. Chem. Soc. 2010, 132 (11), 3688–3690. (203) Yoshimura, T.; Tsukurimichi, E.; Iizuka, Y.; Mizuno, H.; Isaji, H.; Shimasaki, C. The Ei Reaction of Substituted Threo- and Erythro-1-Phenylethyl Phenyl Sulfoxides. Bull. Chem. Soc. Jpn. 1989, 62 (6), 1891–1899. (204) Sletten, E. M.; de Almeida, G.; Bertozzi, C. R. A Homologation Approach to the Synthesis of Difluorinated Cycloalkynes. Org. Lett. 2014, 16 (6), 1634–1637. (205) Chang, P. V.; Prescher, J. A.; Sletten, E. M.; Baskin, J. M.; Miller, I. A.; Agard, N. J.; Lo, A.; Bertozzi, C. R. Copper-Free Click Chemistry in Living Animals. Proc. Natl. Acad. Sci. U. S. A. 2010, 107 (5), 1821–1826. (206) Zhao, Y.; Truhlar, D. G. Hybrid Meta Density Functional Theory Methods for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions: The MPW1B95 and MPWB1K Models and Comparative Assessments for Hydrogen Bonding and van Der Waals Interactions. J. Phys. Chem. A 2004, 108 (33), 6908– 6918. (207) Hogg, D. R. Chemistry of Sulphenic Acids and . In Sulfenic Acids and Derivatives (1990); Patai, S., Ed.; John Wiley & Sons, Ltd., 1990; pp 361–402. (208) Freeman, F.; Bui, A.; Dinh, L.; Hehre, W. J. Dehydrative Cyclocondensation Mechanisms of Hydrogen Thioperoxide and of Alkanesulfenic Acids. J. Phys. Chem. A 2012, 116 (30), 8031–8039. (209) Freeman, F.; Adesina, I. T.; La, J. L.; Lee, J. Y.; Poplawski, A. A. Conformers of Cysteine and Cysteine Sulfenic Acid and Mechanisms of the Reaction of Cysteine Sulfenic Acid with 5,5-Dimethyl-1,3-Cyclohexanedione (Dimedone). J. Phys. Chem. B 2013, 117 (50), 16000–16012. (210) de la Vega, J. G.; Miguel, B. Basis Sets for Computational Chemistry. Introd. Adv. Top. Comput. Chem. Ed. Montero Dıaz R Bader 41–80. (211) Zhao, Y.; González-García, N.; Truhlar, D. G. Benchmark Database of Barrier Heights for Heavy Atom Transfer, Nucleophilic Substitution, Association, and Unimolecular Reactions and Its Use to Test Theoretical Methods. J. Phys. Chem. A 2005, 109 (9), 2012–2018. (212) Zhao, Y.; Schultz, N. E.; Truhlar, D. G. Design of Density Functionals by Combining the Method of Constraint Satisfaction with Parametrization for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2006, 2 (2), 364–382. (213) Zhao, Y.; Truhlar, D. G. Exploring the Limit of Accuracy of the Global Hybrid Meta Density Functional for Main-Group Thermochemistry, Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2008, 4 (11), 1849–1868.

150

(214) 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. 2007, 120 (1–3), 215–241. (215) Peverati, R.; Truhlar, D. G. Improving the Accuracy of Hybrid Meta-GGA Density Functionals by Range Separation. J. Phys. Chem. Lett. 2011, 2 (21), 2810– 2817. (216) Mardirossian, N.; Head-Gordon, M. How Accurate Are the Minnesota Density Functionals for Noncovalent Interactions, Isomerization Energies, Thermochemistry, and Barrier Heights Involving Molecules Composed of Main-Group Elements? J. Chem. Theory Comput. 2016, 12 (9), 4303–4325. (217) Roux, M. V.; Temprado, M.; Jiménez, P.; Dávalos, J. Z.; Notario, R.; Martín- Valcárcel, G.; Garrido, L.; Guzmán-Mejía, R.; Juaristi, E. Thermochemistry of 1,3- Dithiacyclohexane 1-Oxide (1,3-Dithiane Sulfoxide): Calorimetric and Computational Study. J. Org. Chem. 2004, 69 (16), 5454–5459. (218) Juaristi, E.; Notario, R.; Roux, M. V. Calorimetric and Computational Study of Sulfur-Containing Six-Membered Rings. Chem. Soc. Rev. 2005, 34 (4), 347. (219) Denis, P. A. Basis Set Requirements for Sulfur Compounds in Density Functional Theory: A Comparison between Correlation-Consistent, Polarized-Consistent, and Pople-Type Basis Sets. J. Chem. Theory Comput. 2005, 1 (5), 900–907. (220) Petersen, H.; Meier, H. Synthese von Cycloocteninen. Chem. Ber. 1980, 113 (7), 2383–2397. (221) Stöckmann, H.; Neves, A. A.; Stairs, S.; Ireland-Zecchini, H.; Brindle, K. M.; Leeper, F. J. Development and Evaluation of New Cyclooctynes for Cell Surface Glycan Imaging in Cancer Cells. Chem. Sci. R. Soc. Chem. 2010 2011, 2 (5), 932– 936. (222) Scholl, M.; Ding, S.; Lee, C. W.; Grubbs, R. H. Synthesis and Activity of a New Generation of Ruthenium-Based Olefin Metathesis Catalysts Coordinated with 1,3- Dimesityl-4,5-Dihydroimidazol-2-Ylidene Ligands. Org. Lett. 1999, 1 (6), 953–956. (223) Garber, S. B.; Kingsbury, J. S.; Gray, B. L.; Hoveyda, A. H. Efficient and Recyclable Monomeric and Dendritic Ru-Based Metathesis Catalysts. J. Am. Chem. Soc. 2000, 122 (34), 8168–8179. (224) Morra, N. A.; Pagenkopf, B. L. Gram Scale Synthesis of the C(1)–C(9) Fragment of Amphidinolide C. Tetrahedron 2013, 69 (40), 8632–8644. (225) Conrad, J. C.; Eelman, M. D.; Silva, J. A. D.; Monfette, S.; Parnas, H. H.; Snelgrove, J. L.; Fogg, D. E. Oligomers as Intermediates in Ring-Closing Metathesis. J. Am. Chem. Soc. 2007, 129 (5), 1024–1025.

151

(226) Belletire, J. L.; Rauh, R. J. Semicarbazide. In Encyclopedia of Reagents for Organic Synthesis; John Wiley & Sons, Ltd, 2001. (227) Sletten, E. M.; Bertozzi, C. R. A Hydrophilic Azacyclooctyne for Cu-Free Click Chemistry. Org. Lett. 2008, 10 (14), 3097–3099. (228) Hagendorn, T.; Bräse, S. A New Route to Dithia- and Thiaoxacyclooctynes via Nicholas Reaction. RSC Adv. 2014, 4 (30), 15493–15495. (229) Rauk, A.; Yu, D.; Armstrong, D. A. Oxidative Damage to and by Cysteine in Proteins: An Ab Initio Study of the Radical Structures, CH, SH, and CC Bond Dissociation Energies, and Transition Structures for H Abstraction by Thiyl Radicals. J. Am. Chem. Soc. 1998, 120 (34), 8848–8855. (230) Schöneich, C. Cysteine Residues as Catalysts for Covalent Peptide and Protein Modification: A Role for Thiyl Radicals? Biochem. Soc. Trans. 2011, 39 (5), 1254– 1259. (231) Wright, J. S.; Shadnia, H.; Chepelev, L. L. Stability of Carbon-Centered Radicals: Effect of Functional Groups on the Energetics of Addition of Molecular Oxygen. J. Comput. Chem. 2009, 30 (7), 1016–1026. (232) Baskin, J. M.; Prescher, J. A.; Laughlin, S. T.; Agard, N. J.; Chang, P. V.; Miller, I. A.; Lo, A.; Codelli, J. A.; Bertozzi, C. R. Copper-Free Click Chemistry for Dynamic in Vivo Imaging. Proc. Natl. Acad. Sci. 2007, 104 (43), 16793–16797. (233) McKay, C. S.; Finn, M. G. Click Chemistry in Complex Mixtures: Bioorthogonal Bioconjugation. Chem. Biol. 2014, 21 (9), 1075–1101. (234) McGarry, D. J.; Shchepinova, M. M.; Lilla, S.; Hartley, R. C.; Olson, M. F. A Cell-Permeable Biscyclooctyne As a Novel Probe for the Identification of Protein Sulfenic Acids. ACS Chem. Biol. 2016.

152

SCHOLASTIC VITA BORN: January 11, 1985

Charlotte, North Carolina

EDUCATION:

2017 Ph.D. Candidate in Chemistry Wake Forest University Winston Salem, NC, 27109

2011 M.S. in Chemistry Furman University Greenville, SC, 29613

2010 B.S. in Chemistry University of North Carolina Asheville Asheville, NC, 28804

EXPERIENCE:

2015-2016 Research Assistant Wake Forest University

2014-2015 PhD Scientist Intern GlaxoSmithKline

2012-2014 Research Assistant Wake Forest University

2011-2012 Teaching Assistant Wake Forest University

HONORS AND AWARDS:

2016 Outstanding Senior Graduate Student Wake Forest University

2012-2014 NIH Training Grant & Research Fellowship Structural and Computational Biophysics T32 GM 095440 Wake Forest University

2010 Most Valuable Senior University of North Carolina Asheville

153

PROFESSIONAL MEMBERSHIPS:

2009-present American Chemical Society

PUBLICATIONS

1. Keceli, G.; Wu, H.; Devarie-Baez, N. O.; Poole T. H.; Tsang, A. W.; Poole, L. B.; Xian, M.; King, S. B.; Furdui, C. M. Selective Thiol-Blocking Reagents Enhance Quantitative Capabilities in Redox Proteomics. Nature Communications Revisions Submitted

2. Poole, T. H.; Reisz, J. A.; Zhao, W. L.; Poole, L. B.; Furdui, C. M.; and King, S. B. Strained cycloalkynes as new protein sulfenic acid traps. Journal of the American Chemical Society 2014, 136(17), 6167−70.

3. Graham, T. J. A.; Poole, T. H.; Reese, C. N.; Goess, B. C. Regioselective Semihydrogenation of Dienes. The Journal of Organic Chemistry 2011, 76, 4132-4138.

PRESENTATIONS

1. Poole, T. H.; Reisz, J. A.; Poole, L. B.; Furdui, C. M.; and King, S. B. “Sulfenic Acid Determination Via Small Molecule Probes” Thiol-Based Redox Regulation & Signaling, Girona, Spain, July 2014

2. Poole, T. H.; Reisz, J. A.; Poole, L. B.; Furdui, C. M.; and King, S. B. “Sulfenic Acid Determination Via Small Molecule Probes” National Organic Chemistry Symposium, Seattle, WA, June 2013

3. Poole, T. H.; Graham, T. J. A.; Reese, C. N.; Goess, B. C. “Regioselective semihydrogenation of dienes” Southwest & Southeast Meeting of the ACS, Division of Organic Chemistry Oral Presentation, New Orleans, LA, December 2010

4. Poole, T. H., Abuadas. A, M., Holt, H. L. “Synthesis of Novel Indole-Based Combretastatin and Phenstatin Analogs” Southeastern Meeting of the ACS, Sci-MiX Poster Session, San Juan, PR October 2009

5. Holt, H. L., Poole, T. H., Abuadas. A, M., Oliver, R. A., ”Methodology Towards the Synthesis of Potentially Cytotoxic Heterocycles” Southeastern Meeting of the ACS, Poster Session, Nashville, TN October 2009 154