<<

Bifacial PNA in Nucleic Folding, Ligation and in vitro Selection

Dissertation

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy

in the Graduate School of the Ohio State University

By

Xijun Piao, B.S.

Graduate Program in Chemistry

The Ohio State University

2016

Dissertation Committee:

Dr. Dennis Bong, Advisor

Dr. Jovica Badjic

Dr. Jonathan R. Parquette

i

Copyright by

Xijun Piao

2016

ii

Abstract

This dissertation summarizes three research projects regarding the fundamental studies and applications of bifacial Peptide (bPNA) – nucleic acid hybrid triplex system. Chapter 2 describes the syntheses of a series of bPNAs with different length and the cooperative folding of thymine-rich deoxyoligonucleotides by bPNA into binary, ternary and quaternary complexes. Chapter 3 demonstrates that template effect exists in abiotic bPNA – nucleic acid triplex system in that nucleic acid templates acyl group transfer, bPNA ligation and extension. Chapter 4 presents the application of bPNA – nucleic acid triplex as a replacement of native duplex in constructing a reversibly-structured RNA library for in vitro selection against stapled highly helical HIV-1 Rev and also explores the concept of target-directed evolution of RNA library.

In Chapter 2, we demonstrate -displaying bPNA recognizes thymine-rich DNA in predictable and multifaceted ways that allow binding affinity, structure stability and stoichiometry to be tuned through simple bPNA length modification and matching with

DNA length. The longer recognition interface results in the increase of melting temperature and enthalpy change for dissociation, as well as decrease of dissociation constant. 10mer bPNA can tolerate a high number of “mismatches” on thymine tracts and still yields triplex formation whose melting temperature correlates directly with thymine content.

Interestingly, when a DNA host has more T − T sites than melamine sites on bPNA, two or three bPNAs can bind to a single DNA, resulting in ternary and quaternary complexes that have higher thermal stability than the binary (1:1) bPNA − DNA complex, suggestive of cooperative multisite binding. In contrast, when two bPNAs of different lengths bind to ii the same DNA host, a ternary complex is formed with two melting transitions, corresponding to independent melting of each bPNA component from the complex. This set of data serve as the foundation for our future research.

In Chapter 3, we demonstrate DNA- and RNA-templated chemical transformation of bifacial peptide nucleic acid (bPNA) fragments directed by an abiotic thymine/uracil – melamine – thymine/uracil triplex hybrid interface. Watson-Crick base-pairing is widely studied in nucleic acid-templated chemistry, however very few reports are about non-

Watson-Crick recognition. Triplex hybridization of reactive bPNA fragments with DNA template is shown to catalyze acryl group transfer, bPNA native chemical ligation and length-controlled bPNA extension. Gratifyingly, RNA-templated oxidative coupling of bPNA fragments is found to result in the emergence of engineered cleavage.

These data demonstrate that nucleic acid template effect exists in abiotic melamine- thymine/uracil bifacial recognition and establish a connection between engineered and native reaction sites.

In Chapter 4, we report the use of bPNA – nucleic acid triplex hybridization to construct a reversibly-structured RNA library for in vitro selection against side chain-stapled HIV-1

Rev peptides. Bifacial PNA – nucleic acid triplex stem structurally replaces native duplex stem evolved from random region and forces the library into a stem-loop structure, leaving the precious random for the evolution of binding motifs instead of non-binding duplex stem. This bPNA-assisted reversibly-structured RNA library also facilitates the problematic reverse transcription of selected RNA by loosening the extensive secondary structures. Although many peptide targets have been used for in vitro selection, a stapled peptide is not seen and known to enhance helicity and binding affinity. In vitro selection against stapled original Rev peptide yields RNA aptamers showing the conserved core element presented in native Rev Responsive Element (RRE) RNA. A key iii asparagine on the stapled Rev peptide is mutated to two synthetic -displaying amino that direct the evolution of random based on likely hydrogen- bonding preference. This ongoing project integrates engineered bPNA – nucleic acid triplex stem into in vitro selection and may solve the problem of reverse transcription of in vitro selection. Further single mutations on stapled peptide targets demonstrate the concept of directed evolution of RNA library.

iv

Dedication

Dedicated to my wife and parents

v

Acknowledgements

I am very grateful to my PhD advisor, Dr. Dennis Bong. We have worked together for six years and he not only helped me enrich my knowledge in organic chemistry and biophysics, but also made me become a more confident and competent researcher. His patience, persistence and trust led us to complete our very challenging in vitro selection project, in which I learned nucleic acid chemistry from the very beginning and improved my problem- solving ability. His goal for research originality is highly admirable and everything I learned in his lab would become a great fortune for my future research.

I would also like to thank Dr. Badjic and Dr. Parquette for serving my candidacy and defense committee, as well as writing reference letters in support of my postdoctoral researcher application.

I also highly appreciate the helpful advice and discussions provided by friendly labmates, as well as productive collaborations.

Finally, I am very grateful and indebted to my family. My parents, Mingtao Piao and

Huiqing Gu, have been giving me unconditional love and tremendous support all the time throughout the years I was far away from home. My wife Ying Yu and my parents-in-law have been giving me endless love and indispensable trust since my marriage and I believe their love, understanding and support will make me become a better, more responsible and supportive husband to my family.

vi

Vita

2003 – 2006……………………………………. Anshan No.1 Middle School, Anshan, China

2006 – 2010………………………………………… B.S., Fudan University, Shanghai, China

2010 – 2016...... ……Graduate Teaching / Research Assistant, The Ohio State University

Publications during PhD

(1) Xijun Piao, Xin Xia, Jie Mao, and Dennis Bong*, “Peptide Ligation and RNA

Cleavage via an Abiotic Template Interface” J. Am. Chem. Soc. 2015, 137, 3751-3754.

(2) Xijun Piao, Xin Xia, and Dennis Bong*, “Bifacial Peptide Nucleic Acid Directs

Cooperative Folding and Assembly of Binary, Ternary, and Quaternary DNA

Complexes” Biochemistry, 2013, 52, 6313-6323.

(3) Xin Xia, Xijun Piao, and Dennis Bong*, “Bifacial Peptide Nucleic Acid as an

Allosteric Switch for and Ribozyme Function” J. Am. Chem. Soc. 2014, 136,

7265-7268.

(4) Xin Xia, Xijun Piao, Kurt Fredrick, and Dennis Bong*, “Bifacial PNA Complexation

Inhibits Enzymatic Access to DNA and RNA” ChemBioChem, 2014, 15, 31-36.

(5) Yingying Zeng, Yaowalak Pratumyot, Xijun Piao and Dennis Bong*, "Discrete

assembly of synthetic peptide-DNA triplex structures from polyvalent melamine -

thymine bifacial recognition" J. Am. Chem. Soc. 2012, 134, 832-835.

Fields of Study

Major field: Chemistry vii

Table of Contents

Abstract…………………………………………………………………………………………....ii

Acknowledgement…………………………………………………………………………….…vi

Vita………………………………………………………………………………………………..vii

Table of Contents………………………………………………………………………….…….vii

List of Figures…………………………………………………………………………………...xiv

List of Tables………………………………………………………………………………...... xviii

Chapter 1. Structures, Recognition and Template Effect of Nucleic Acids ...... 1

1.1 Nucleic acid structures ...... 2

1.1.1 DNA duplex ...... 3

1.1.2 DNA triplex ...... 4

1.1.3 Nucleic acid G-quadruplex ...... 6

1.1.4 RNA folding ...... 7

1.2 Modifications on nucleic acid to enhance triplex formation...... 8

1.3 Peptide Nucleic Acid (PNA) – Native on pseudopeptide backbone . 10

1.3.1 Chemical structure of PNA ...... 10

1.3.2 PNA-native nucleic acid duplex and triplex ...... 11

1.3.3 Biological and medicinal applications of PNA ...... 12

1.4 Targeting pairs with Janus-Wedge insertion strategy ...... 12

1.4.1 Janus-Wedge Insertion of a single nucleobase pair by small molecules ...... 13

1.4.2 Single Janus-Wedge insertion on biopolymer scaffold ...... 14 viii

1.4.3 Multivalent insertion of Janus-Wedge heterocycles into nucleobase pairs ..... 14

1.4.4 Triazine used to target nucleobase or nucleobase pairs and bifacial Peptide

Nucleic Acid (bPNA) ...... 16

1.5 Nucleic acid template effect on nonenzymatic reactions ...... 19

1.6 References ...... 22

Chapter 2. Bifacial Peptide Nucleic Acid Directs Cooperative Folding and Assembly of

Binary, Ternary, and Quaternary DNA Complexes ...... 36

2.1 Introduction ...... 37

2.2 Results and Discussion ...... 41

2.2.1 1:1 bPNA-DNA complexes ...... 41

2.2.2 Binding affinity and enthalpy of bPNA-DNA complexes ...... 43

2.2.3 Tolerance of “mismatch” sites in (EM*)10 DNA complexation ...... 46

2.2.4 bPNA-DNA higher-order assembly ...... 51

2.3 Conclusion ...... 55

2.4 Materials and Equipments ...... 56

2.5 Synthesis ...... 57

2.5.1 Synthesis of Fmoc-Lys(M*)-OH ...... 57

2.5.2 Solid phase synthesis of bPNA ...... 63

2.5.3 HPLC and Maldi-TOF mass spectra of the synthetic peptides ...... 64

2.6 Experimental Methods ...... 70

2.7 Acknowledgement ...... 71

2.8 References ...... 71

ix

Chapter 3. Peptide Ligation and RNA Cleavage via an Abiotic Template Interface ...... 78

3.1 Introduction ...... 79

3.2 Results and Discussion ...... 81

3.2.1 Acyl transfer catalyzed by bPNA-DNA hybridaztion...... 81

3.2.2 Native chemical ligation templated by bPNA-DNA hybridization ...... 82

3.2.3 Length-controlled bPNA extension ...... 85

3.2.4 Ribozyme-templated bPNA oxidative coupling and activation of ribozyme self-

cleavage function ...... 86

3.3 Conclusion ...... 89

3.4 Materials and Equipments ...... 89

3.5 Sequences ...... 90

3.6 Synthesis ...... 92

3.6.1 ABA-Gly thioester ...... 92

3.6.2 bPNA Synthesis ...... 101

3.7 Experimental Methods ...... 113

3.7.1 DNA Complexation ...... 113

3.7.2 UV melting ...... 113

3.7.3 Electrophoretic mobility shift assay (EMSA) ...... 113

3.7.4 Acyl small-molecule transfer ...... 114

3.7.5 Native chemical ligation ...... 114

3.7.6 Cysteine-free direct aminolysis ...... 114

3.7.7 Disulfide ligation ...... 115

3.7.8 Ellman’s test for dicysteine bPNA extension ...... 115

x

3.7.9 Product isolation from gel ...... 115

3.7.10 Gel quantification ...... 115

3.7.11 RNA transcription and purification ...... 116

3.7.12 Ribozyme cleavage with bPNA ...... 116

3.7.13 U3 Ribozyme-templated bPNA disulfide oxidation ...... 116

3.7.14 In situ bPNA oxidation and U3 Ribozyme cleavage ...... 117

3.8 Acknowledgement ...... 117

3.9 References ...... 117

Chapter 4. In vitro Selection of a bPNA-Assisted Reversibly-Structured RNA Library against Stapled HIV-1 Rev Peptides ...... 124

4.1 Introduction ...... 125

4.1.1 In vitro selection and aptamers ...... 125

4.1.2 Two major limitations in in vitro selection ...... 127

4.1.3 A novel strategy simultaneously overcomes the two limitations ...... 128

4.1.4 HIV-1 Rev peptide and it use in in vitro selection...... 129

4.1.5 Stapled peptides enhance α-helicity and binding affinity ...... 130

4.1.6 A bPNA-assisted reversibly-structured RNA library for in vitro selection against

stapled HIV-1 Rev peptides and directed evolution of aptamers ...... 131

4.2 Results and Discussion ...... 133

4.2.1 Stapled Rev peptides enhance helicity and binding affinity ...... 133

4.2.2 General procedure of in vitro selection ...... 136

4.2.3 Significant enrichment is observed after 6th round ...... 137

4.2.4 Sequences of interest are identified and share similarities with native RRE 139

xi

4.2.5 Single mutation on Rev peptide directs the evolution of RNA library ...... 142

4.3 Conclusion ...... 143

4.4 Materials and Equipments ...... 143

4.5 Sequences ...... 144

4.5.1 Stapled Rev peptide sequences ...... 145

4.5.2 Nucleic acid sequences ...... 145

4.6 Synthesis ...... 146

4.6.1 Synthesis of Fmoc-azidoalanine ...... 146

4.6.2 Synthesis of Fmoc-Lys(Biotin)-OH ...... 147

4.6.3 Synthesis of Fmoc-Dap(M*)-OH ...... 149

4.6.4 Synthesis of Fmoc-Dap(N*)-OH ...... 150

4.6.5 Rev and stapling ...... 153

4.7 Experimental Procedures ...... 157

4.7.1 Transcription of N25 RNA library ...... 158

4.7.2 Annealing of N25 RNA library with bPNA ...... 159

4.7.3 Pre-selection of annealed N25 RNA library ...... 160

4.7.4 In vitro selection against target ...... 161

4.7.5 Pull-down of binding aptamer ...... 161

4.7.6 Wash the beads with high buffer ...... 161

4.7.7 Denaturation and collection of binding aptamers ...... 161

4.7.8 Reverse transcription of selected RNA into complementary DNA ...... 162

4.7.9 PCR amplification of cDNA ...... 163

4.8 Acknowledgement ...... 164

xii

4.9 References ...... 164

Appendix A: Analysis of High-Throughput Sequencing Raw Data ...... 175

A.1 Installation ...... 176

A.2 Download raw data ...... 176

A.3 Removal of 3’-constant region on T-tract sequences ...... 176

A.4 Removal of 5’ constant region ...... 178

A.5 Removal of two T10 tracts ...... 179

A.6 Analysis of trimmed sequences ...... 181

A.7 References ...... 183

References ...... 184

Chapter 1 ...... 184

Chapter 2 ...... 197

Chapter 3 ...... 203

Chapter 4 ...... 209

Appendix A ...... 219

xiii

List of Figures

Figure 1.1 Chemical structures of nucleic acids………………………………….….………...3

Figure 1.2 Watson-Crick base pairing and DNA duplex………………………………..…....4

Figure 1.3 Nucleobase triplets and their orientations………………………………………...6

Figure 1.4 G-quadruplex structures………………………………………………………….....7

Figure 1.5 Secondary and tertiary structural elements in RNA……………………………...8

Figure 1.6 Examples of modifications on nucleic acid to enhance triplex formation………9

Figure 1.7 Peptide nucleic acid structure and its hybridization with native DNA………….10

Figure 1.8 PNA binding modes for targeting double-stranded DNA……………………….11

Figure 1.9 Janus-Wedge insertion into nucleobase pairs…………………………………..13

Figure 1.10 Examples of triazine-bearing sequences……………………………………….15

Figure 1.11 Bifacial melamine-thymine recognition…………………………………………18

Figure 1.12 Examples of nucleic acid template effect………………………………………20

Figure 2.1 Illustration of melamine-thymine interaction directing DNA and bPNA assembly and CD spectra………….……………………………………………………………………...38

Figure 2.2 Normalized fluorescence titration curve for Cbf-(EM*)n-dTnC4Tn……………..41

Figure 2.3 Characterization of binding affinity……………………………………………….43

Figure 2.4 Invasion of triplex (dT10C4T10+H2N-(EM*)10G) with dA10G4A10…………………44

Figure 2.5 Fluorescence titration curve for bPNA with mismatched DNA hairpin………..48

Figure 2.6 Normalized fluorescence titration curve for higher-order assembly…………..50

Figure 2.7 Illustration of higher-order bPNA-DNA assembly………………………………..52

Figure 2.8 Analysis of binary and ternary complexes of bPNA and dT15C4T15……………52

xiv

Figure 2.9 1H NMR of Fmoc-Lys(M*)-OH…………………………………………………….57

Figure 2.10 13C NMR of Fmoc-Lys(M*)-OH…………………………………………………..58

Figure 2.11 Characterization of 4mer bPNA Cbf-(β-A)-(EM*)4G……………………...... 60

Figure 2.12 Characterization of 5mer bPNA Cbf-(β-A)-(EM*)5G…….…………………….61

Figure 2.13 Characterization of 6mer bPNA Cbf-(β-A)-(EM*)6G………….……………….62

Figure 2.14 Characterization of 8mer bPNA Cbf-(β-A)-(EM*)8G………………..…………63

Figure 2.15 Characterization of 10mer bPNA Cbf-(β-A)-(EM*)10G…………………………64

Figure 2.16 Characterization of 10mer bPNA Aba-(β-A)-(EM*)10G…..……………………65

Figure 3.1 bPNA molecular recognition and nucleic acid templates………………………76

Figure 3.2 DNA-templated thioester exchange between bPNAs…………………………..78

Figure 3.3 DNA-templated native chemical ligation…………………………………………80

Figure 3.4 DNA-templated bPNA extension………………………………………………….82

Figure 3.5 Nucleic acid-templated oxidative coupling of bPNA thiols……………………..84

Figure 3.6 1H NMR of product 1………………………………………………………………..90

Figure 3.7 1H NMR of Product 2…………………………………………………………….…91

Figure 3.8 1H NMR for ABA-Gly thioester…………………………………………………….92

Figure 3.9 13C NMR of ABA-Gly thioester…………………………………………………….93

Figure 3.10 Characterization of bPNA 1………………………………………………………96

Figure 3.11 Characterization of bPNA 2………………………………………………………97

Figure 3.12 Characterization of bPNA 3………………………………………………………98

Figure 3.13 Characterization of bPNA 4………………………………………………………99

Figure 3.14 Characterization of bPNA 5…………………………………………………….100

Figure 3.15 Characterization of bPNA 6…………………………………………………….101

Figure 3.16 Characterization of bPNA 7…………………………………………………….102

Figure 3.17 Characterization of bPNA 8…………………………………………………….103 xv

Figure 3.18 Characterization of bPNA 9…………………………………………………….104

Figure 3.19 Characterization of bPNA 10…………………………………………………..105

Figure 4.1 General procedure of in vitro selection (SELEX)….…………………………..119

Figure 4.2 Illustration of bPNA-assisted reversibly-structured RNA library construction.121

Figure 4.3 Short Rev peptide adopts adaptive binding to its RNA aptamers…………...124

Figure 4.4 Structure of isolated short hairpin Rev responsive element (RRE), HIV-1 Rev peptide and their interaction………………………………………………………………….126

Figure 4.5 α-helix of Rev peptides…………………………………………………………..128

Figure 4.6 Procedure of in vitro selection of bPNA-assisted reversibley-structured RNA library.……………………………………………………………………………….………….129

Figure 4.7 UV melting of RNA library at different round with bPNA………………………131

Figure 4.8 Unique sequence percentage of each pool in three selections……………….131

Figure 4.9 Predicted folding of top aptamers of original stapled Rev peptide and native

RRE…………………………………………………………………………………………….134

Figure 4.10 Predicted folding of top sequences in Cluster 3 and possible bifacial recognition..……………………………………………………………………………………135

Figure 4.11 Characterization of stapled original HIV-1 Rev peptide…………………….146

Figure 4.12 Characterization of stapled melamine-displaying HIV-1 Rev peptide……..147

Figure 4.13 Characterization of stapled -displaying HIV-1 Rev peptide……..148

Figure 4.14 2% Agarose gel image for PCR duplexes in the 10th round of selection….156

Figure A.1 Removal of 3’ constant flanking region and adapter...... 169

Figure A.2 Removal of 5’ constant flanking region…………………………………………170

Figure A.3 Removal of one T-tract…………………………………………………………..171

Figure A.4 Removal of the other T-tract…………………………………………………….172

Figure A.5 Converting from fastq to fasta file……………………………………………….173 xvi

Figure A.6 Clustering the aptamers…………………………………………………………174

Figure A.7 Comparing the enrichment between different rounds………………………..175

xvii

List of Tables

a Table 2.1 Thermodynamic data for [(EM*)n-dTnC4Tn] ………………………………..…….42

Table 2.2 Tm of Cbf-10 complexes with T-rich 24-nt DNA…………………………………..46

Table 2.3 Tm’s of higher order bPNA-DNA assembly………………………………………..49

Table 4.1 Populated sequences identified from three in vitro selections………………..133

Table 4.2 In vitro selection condition.……………………………………………………….149

Table 4.3 Transcription condition for 1st round………..…………………………………….151

Table 4.4 Annealing condition………………………………………………………………..152

Table 4.5 Pre-selection condition..…………………………………………………………..152

Table 4.6 Reverse transcription condition…………………………………………………..154

Table 4.7 Manual hot start PCR condition…………………………………………………..155

Table A.1 Two representative sequences…………………………………………………..167

Table A.2 Representative sequence………………………………………………………...169

Table A.3 Representative sequence change……………………………………………….170

Table A.4 N25 region of interest………………………………………………………...…..172

xviii

Chapter 1. Structures, Recognition and Template Effect of Nucleic Acids

1

This dissertation describes and summarizes the fundamental studies and applications of hybrid triplexes formed by synthetic bifacial peptide nucleic acid (bPNA) and native nucleic acids (DNA, RNA). Related background of nucleic acid structures, enhanced triplex formed from modified oligonucleotides, peptide nucleic acid, Janus-Wedge insertion, bPNA and nucleic acid template effect is thus discussed in the introduction. A two-way communication is established in Chapter 2 and Chapter 3: bPNA folds DNA or RNA into a triplex stem-loop structure, while DNA or RNA catalyzes bPNA ligation. In Chapter 4, we test a novel strategy of constructing a bPNA-assisted reversibly-structured RNA library for in vitro selection against stapled original and point-mutated highly helical HIV-1 Rev peptides to overcome two major limitations exist in in vitro selection and explore the concept of directed evolution of nucleic acid.

1.1 Nucleic acid structures

Native nucleic acid includes deoxyribonucleic acid (DNA) and ribonucleic acid (RNA).

Nucleic acid is also known as polynucleotides while each is composed of one of the four nucleobases (, , adenine and thymine in DNA; cytosine, guanine, adenine and uracil in RNA), a monosaccharide (2- in DNA; in

RNA) and a group. Nucleobase forms N-glycosidic bond between atom and 1’-OH of the and phosphate group forms ester bond with 5’-OH of the sugar. Nucleotides are then bonded to each other linearly through 3’-5’ phosphodiester linkage and the name of polynucleotide indicates that nucleic acid is of highly repetitive covalent conjugation of nucleotides.1-3 The strand shows a 5’ to 3’ polarity.

2

Figure 1.1 Chemical structures of nucleic acids. Adapted from reference 3

1.1.1 DNA duplex

DNA tends to form helical duplex due to non-covalent Watson-Crick base pairing which states that adenine on one chain is specifically paired with thymine on the other chain, while guanine is always paired with cytosine on different chains.4 Thus, the two complementary strands are held together in an anti-parallel orientation. At least three forms of DNA duplex are found in nature, A-form, B-form and Z-form. And the most abundant B-form DNA is a right-handed helix with high periodicity. It is 23.7 Å wide and

34 Å per helical turn (~10.5 base pairs) about the axis in solution.5 A-form DNA helix is also right-handed but compressed along the axis, while Z-form DNA is left-handed and elongated along the axis. When forming a double-stranded helix, two types of grooves are

3 generated, major groove and minor groove. Major groove is wide and rich in chemical information which can be used to distinguish different base pairs. Minor groove is narrow and traditionally considered to contain less chemical information. However, Dervan and co-workers are able to differentiate base pairs using synthetic polyamides in DNA minor groove.6-9

Figure 1.2 Watson-Crick base pairing and DNA duplex. Base pairing with an exemplary

numbering (Left); DNA duplex of different forms (Right). Adapted from Molecular Biology

of the Gene, 5th edition, James D. Watson, published by Pearson Education, Inc.

1.1.2 DNA triplex

Nucleic acid triplex is formed when a third strand of binds to a preformed

DNA duplex through Hoogsteen or reversed Hoogsteen base pairing in the major 4 groove.10 Hoogsteen base pairing is different from classic Watson-Crick base pairing in that N7 of the base is involved in base pairing instead of N1. A third strand of homopyrimidine can bind to a double-stranded helix composed of a homopurine strand and a homopyrimidine strand via Hoogsteen base pairing, based on T-A*T and C-G*C+ triplex base pairing (* denotes Hoogsteen base pairing; - denotes Watson-Crick base pairing; C+ denotes a protonated cytosine).11,12 A third strand of homopurine13,14 can also bind to a duplex composed of a homopurine strand and a homopyrimidine strand via reversed Hoogsteen base pairing pattern, based on T-A*A and C-G*G triplex base pairing.

In Hoogsteen base pairing, the third strand is parallel to the homopurine strand, whereas they are antiparallel in the reversed Hoogsteen base pairing.15 DNA triplex generally forms in non-physiological conditions, such as lower pH (<6).16 As seen in C-G*C+ triplet, cytosine on the additional strand needs to be protonated to form hydrogen bonds with guanine. Furthermore, divalent cation17 and polyamine18 are helpful to shield electrostatic repulsion between the three highly negatively-charged oligonucleotide strands.19,20

Although DNA triplex often forms at non-physiological conditions, some homopurine- homopyrimidine DNA duplex that is a potential targeting site for triplex formation is also found in vivo.21-23 Marky and co-workers found that formation of DNA triplex was accompanied by a favored enthalpic change and an unfavored entropic change, indicative of an enthalpy and entropy compensation. Comparison of the thermodynamic profiles of these triplexes yielded enthalpic contributions of -24 kcal/mol, -23 kcal/mol, and -22 kcal/mol for the formation of TAT/TAT, TAT/CGC+, and CGC+/CGC+ base triplet stack.24

5

Figure 1.3 Nucleobase triplets and their orientations. (Left) Hoogsteen base-pairing;

(Right) reverse Hoogsteen base-pairing. Solid lines, purine strands; stippled lines,

strands; vertical lines, Watson-Crick hydrogen bonds; diamonds, Hoogsteen

or reverse Hoogsteen hydrogen bonds. Adapted from reference 11.

1.1.3 Nucleic acid G-quadruplex

G-quadruplex is possible in both DNA and RNA when the sequence is rich in guanine.

Four guanine nucleobases can form a square planar structure through Hoogsteen hydrogen bonding, and several layers can stack on top of each other to form a G- quadruplex. This structure is further stabilized by the presence of potassium cation which is coordinated in the center of G-quadruplex.25,26 Different from nucleic acid triplex, G-

6 quadruplex forms at physiological conditions and is found in vivo.27,28 A well-known example is human telomeric repeat d(GGTTAG), which forms G-quadruplexes to decrease the activity of telomerase and fail the maintenance of telomere length.29 As shown in Figure 1.4, G-quadruplex can be either intramolecular or intermolecular.

Figure 1.4 G-quadruplex structures. Cited from reference 15.

1.1.4 RNA folding

Quite distinguished from DNA, a single-stranded RNA can actually fold itself into very complex structures. Indeed, a RNA single strand frequently folds back on itself to form paired regions due to base pairing. Several non-paired nucleotides may be looped out when adjacent sequences form complementary base pairs or wobble base pairs, resulting in hairpin, bulge and loop structures.30,31 Paired RNA region is similar with A-form DNA.32

In contrast to these stem-loop structures, pseudoknot33 indicates that non-contiguous regions form complex structures. Due to the flexibility (freedom of rotation) gained from unpaired region, RNA molecules have a number of complex tertiary structures that are not seen in DNA. Indeed, many functional forms of RNA single strands frequently require specific tertiary structures, such as tRNA34 and ribozyme.35

7

Figure. 1.5 Secondary and tertiary structural elements in RNA. Cited from reference 31.

1.2 Modifications on nucleic acid to enhance triplex formation

As discussed above, nucleic acid triplex structure is of interest, however, native nucleic acid triplex often forms at non-physiological conditions, which limits its in vivo applications.

Researchers thus modified nucleobase and backbone to favor the formation of triplex at physiological conditions. Modifications on sugar or phosphate reduce negative charges on RNA and DNA, lowering electrostatic repulsion between strands and stabilizing triplex.36,37 2’-aminoethyl (EA) introduces positive charges on sugar to neutralize negative charges on phosphate backbone, leading to a more stable triplex structure.38 Among these two types of modifications, (LNA) is quite famous, 2’ oxygen and 4’ are linked within one ribose and locks ribose in the 3’-endo conformation.39 This preorganized conformation increases nucleobase stacking, resulting in more thermo- stable duplex and triplex with other oligonuleotide strands.40-42 Modifications on nucleobase can also enhance the formation of nucleic acid triplex. This usually remains

8

Watson-Crick or Hoogsteen recognition interface intact, but modifies the non-bonding side.

For example, by adding 8’ amino group, guanine and adenine can form more stable triplex with other native nucleobases.43 5’-methylated cytosine elevates the pKa of protonated

N3 and favors the formation of C-G*C+ triplet in DNA.44-46 6-thioguanine favors the formation of triplex over quadruplex and GA homodimer due to O→S modification which cripples the traditional trivalent hydrogen bonds and disfavors the formations of competing structures.47 In addition, a class of synthetic nucleobases interact with matched native nucleobase pairs simultaneously, thereby enhancing the stability of formed triplex.48

2-aminoethyl Locked Nucleic Acid (LNA)

8-aminoguanine Interaction with CG pair

Figure 1.6 Examples of modifications on nucleic acid to enhance triplex formation. Cited from reference 38, 39 and 48.

9

1.3 Peptide Nucleic Acid (PNA) – Native nucleobases on pseudopeptide backbone

1.3.1 Chemical structure of PNA

Different from modifications on phosphate group and sugar shown in last section, peptide nucleic acid (PNA) displays native nucleobases on a pseudopeptide chain, poly(N-(2- aminoethyl)-glycine).49,50 PNA thus mimics nucleic acid due to predictable base-pairing, and differs from nucleic acid in losing backbone enzymatic accessibility.51 All the base- pairing rules discussed in previous sections applies to PNA, including Watson-Crick base pairing and Hoogsteen base pairing.52,53 The native nucleobases are linked to secondary on the backbone and the repetitive structure mimics oligonucleotides greatly: three bonds between base and backbone while six bonds within a repeating monomer. PNA is found to bind to its complementary oligonucleotide stronger than its native counterpart, and this is attributed to the lack of negative charges on the backbone which reduces electrostatic repulsion upon hybridization.49,51 The poly(N-(2-aminoethyl)-glycine) chain is not easily accessible to and , making PNA resistant to enzymatic degradation.54

Figure 1.7 Peptide nucleic acid structure and its hybridization with native DNA. Cited

from reference 51.

10

1.3.2 PNA-native nucleic acid duplex and triplex

The backbone and amide linker to the base show constrained flexibility that is important to base-pairing, and reduction of the amide is detrimental to nucleic acid binding. Due to the neutral backbone, PNA reduces the electrostatic repulsion between strands, resulting in a more thermostable duplex and triplex. Indeed, PNA can invade an existing DNA duplex, forming a hybrid duplex with its complementary strand. In PNA-DNA and PNA-

RNA duplexes, PNA adapts the A-form and B-form helices preferred by RNA and DNA, respectively.51 However, a homopyrimidine PNA usually invades double-stranded DNA,

55-58 forming PNA2-DNA triplex instead of an expected PNA-DNA2 adduct. This strand displacement occurs due to the very high stability of PNA2-DNA triplex.

Figure 1.8 PNA binding modes for targeting double-stranded DNA. Cited from reference

51.

1.3.3 Biological and medicinal applications of PNA

Since synthetic PNA can recognize DNA sequence of interest and invade the existing double-stranded DNA, it may be used to target a particular sequence,59,60 interfering with

11 biochemical processes, such as replication, transcription and translation.61 Once PNA invasion occurs, a PNA-DNA duplex or PNA2-DNA triplex will form, blocking enzymatic read-through of this DNA strand. This is called antigene strategy which mainly interferes with transcription process.53,62,63 If PNA forms stable complex with mRNA and inhibits the translation process, this is the antisense strategy.64-66 Other biological and medicinal applications of PNA include PCR amplification enhancement of small allelic products, determination of telomere size, and single mutation analysis.61 Overall, the wide biological and medicinal applications of PNA originates from the higher thermostability of

PNA-nucleic acid complex which disables the activity of corresponding process.

1.4 Targeting nucleobase pairs with Janus-Wedge insertion strategy

1.4.1 Janus-Wedge Insertion of a single nucleobase pair by small molecules

To target a nucleobase pair, one needs to design a molecule bearing two recognition interfaces that will interact with each of the two nucleobases, Lehn and co-workers first developed the concept of wedge-like heterocycle that recognize cytosine and uracil mismatched pair in chloroform.67 However, a single insertion is not strong enough to be detected in water. Zimmerman and co-workers designed an intercalation-assisted melamine heterocycle that targets T-T or U-U mismatch in CTG or CUG trinucleotide repeat of myotonic dystrophy type 1 sequence, respectively. With the help of acridine, a strong nucleobase intercalator, melamine was able to recognize T-T or U-U mismatch with

68 a Kd at high nanomolecular range (390 nM) in aqueous buffers. This example demonstrates the potential use of such a strategy in diagnostics and therapeutics.69,70

Synthetic Janus compounds for the recognition of G-U mismatched nucleobase pair were also reported.71

12

a

b c

Figure 1.9 Janus-Wedge insertion into nucleobase pairs. (a) The concept of Janus-

Wedge insertion into matched A-T and C-G nucleobase pairs. (b) Example of C-U

mismatched nucleobase pair insertion. (c) Acridine-assisted melamine insertion into T-T

and U-U mismatched pairs. Cited from reference 67 and 68.

1.4.2 Single Janus-Wedge insertion on biopolymer scaffold

Researchers also insert a single synthetic nucleobase mimic that targets matched or mismatched nucleobase pair onto a biopolymer scaffold. Theoretically, a mutation may be detected by the slightly enhanced binding affinity caused by the bifacial recognition nucleobase analogue. Tor72 and Perrin73 designed bifacial nucleobase mimic and inserted it into a deoxyoligonucleotide, respectively. Instead of forming bifacial recognition with A

13 and T simultaneously, their synthetic nucleobases can only pair with A or T at one time, working as a surrogate for T and A in a DNA duplex. This may suggest that targeting a matched nucleobase pair with Janus-Wedge strategy at one single site may be difficult.

1.4.3 Multivalent insertion of Janus-Wedge heterocycles into nucleobase pairs

Different from insertion into a single nucleobase pair, assembly of multiple Janus-Wedge heterocycles onto a biopolymer backbone will provide dramatic increase of binding affinity in targeting nucleobase pairs by multivalent bifacial recognition. Two main categories of biopolymer backbone are DNA and PNA. As discussed in PNA section, PNA with a neutral backbone will give a higher stability of formed triplex hybrid, and this may enable a Janus-

Wedge insertion of perfectly matched nucleobase pairs (A-T and C-G) instead of mismatched pairs, as PNA provides extra energy to break preformed hydrogen bonds.

McLaughlin and co-workers integrated multiple synthetic Janus-Wedge bifacial recognition heterocycles onto a PNA backbone that successfully targeted C-T mismatch nucleobase pairs via triplex formation.74 A single matched C-G nucleobase pair may be wedged-like inserted when it was in the middle of unpaired C-T mismatches.75 Bong and co-workers assembled multiple melamine heterocycles on different scaffolds, such as peptide,76 peptoid77 and .78 These synthetic melamine-displaying molecules have been demonstrated to target T-T or U-U mismatched nucleobase pairs efficiently, inhibit enzymatic activities79 and activate nucleic acid functions.80

14

1.4.4 Triazine used to target nucleobase or nucleobase pairs and bifacial Peptide

Nucleic Acid (bPNA)

Figure 1.10 Examples of triazine-bearing sequences. (a) -displaying

PNA; (b) diaminotriazine-displaying peptide; (c) dioxotriazine-displaying peptide; (d)

melamine-displaying peptide; (e) melamine-displaying peptoid; (f) melamine-

displaying polyacrylate. Adapted from reference 76, 77, 78, 81 and 82.

Triazine can provide hydrogen bonds on three recognition interfaces and can be used in

Janus-Wedge insertion strategy. Triazine has been successfully attached to PNA,81 polyamide,82 peptide,76,82,83 peptoid77,82 and synthetic polymer78 backbone with different linker lengths ranging from 0 atoms to 4 atoms and at varied positions, including endocyclic carbon, endocyclic nitrogen and exocyclic nitrogen. Attaching triazine to a DNA backbone at exocyclic nitrogen is proved to be difficult due to isomerization and

15 hydrolysis.84 Upon covalently linked onto a backbone, one of the three recognition interfaces is blocked, leaving this heterocycle a bifacial recognition motif. Ganesh and co- worker replaced one or two thymine nucleobases of T8 PNA with cyanuric acid. Cyanuric acid was found a stabilizing effect when inserted in the middle of the sequence or at C- terminus.81 Krishnamurthy and Eschenmoser synthesized multiple triazine-bearing sequences on varied backbones with short carbon linkers (0 atoms to 2 atoms).

AcNH T Interestingly, they found 2,4-dioxotriazine-displaying AspGlu( OO)16 bound to corresponding pairing partner poly-adenine sequences dramatically weaker than that

AcNH T of poly-diaminotriazine sequence AspGlu( NN)16 bound to its corresponding partners.

They assumed the electronic property of 2,4-dioxotriazine nucleus disfavored such pairing system rather than a steric incompatibility. Missing one exocyclic amino group resulted in losing one recognition interface, thus their triazine-bearing sequences mainly formed 1:1 duplex with single-stranded oligonucleotide.82 Given the fact that diaminotriazine pairs well with thymine mismatch, Bong and co-workers successfully displayed multiple melamine

(2,4,6-triamino-1,3,5-triazine) heterocycles on different supporting backbones, including peptide, peptoid and . This heterocycle clearly offered two recognition interfaces when attached to backbone. They found melamine-bearing chains formed triplex with unstructured polythymine oligonucleotides in a ratio of 1:2 with strong affinity. If two poly(dT) regions are connected by a non-intervening loop, it formed triplex stem-loop structure with a peptide-DNA ratio of 1:1. The well-studied melamine-displaying peptide system was thus termed bifacial Peptide Nucleic Acid (bPNA). Surprisingly, no duplex was detected during this process. This was further supported by the fact that monofacial triazine prepared by methylating one of the exocyclic amine did not pair with poly(dT) sequence. The linker length played a crucial role on the binding affinity, generally four carbon atoms and three carbon atoms showed much better than one or two carbon atom 16 linker. This may be explained by the decreased electrostatic repulsion between negative charges on the backbone when forming triplex. This is different from what Krishnamurthy and Eschenmoser reported in which 0 carbon atom to 2 carbon atom linkers resulted in moderate to robust duplex formation. Less repulsion in a duplex than that of a triplex could account for this different observations. This is also comparable to the native DNA duplex and triplex system and the latter one usually forms in relatively “harsh” conditions.

Surprisingly, given the same linker length and the same numbers of melamine units on one sequence, a peptoid backbone resulted in a significantly lower binding affinity, whereas peptide backbone yielded robust triplex formation. Despite of irregular structure, the synthetic polyacrylate-supported melamine sequence yielded a robust triplex formation, comparable to that of peptide backbone. Yet the origin of this backbone difference was unclear.76-78,83 Bifacial PNA has abiotic nucleobase and native α-peptide backbone, which is structurally different from conventional PNA that has native nucleobases and non-native polypeptide backbone. Bifacial PNA is also operationally distinguished from conventional PNA in that bPNA forms triplex with non-interacting and unstructured oligonucleotides in an associative fashion, whereas conventional PNA displaces pre-formed duplex in a dissociative manner. Unlike the modified oligonucleotides discussed in Session 1.2 that bind to a pre-formed native duplex and strengthen the triplex formation at physiological conditions, bPNA generates triplex from non-interacting single-stranded oligonucleotides under physiological conditions.

17

Figure 1.11 Bifacial melamine-thymine recognition. Synthetic bPNA shows a 1:2

binding of 10mer bPNA to (dT)10 and a 1:1 binding of 10mer bPNA to dT10C4T10. Cited

from reference 76.

1.5 Nucleic acid template effect on nonenzymatic reactions

Template effect exists widely in molecular biology, as evidenced by replication, transcription, reverse transcription and translation. In any of them, a single strand of nucleic acid serves as the template and relevant polymerase or ribosome reads through the template sequence and brings together the complementary nucleotide triphosphate or aminoacyl-tRNA based on Watson-Crick base pairing to synthesize the complementary strand of nucleic acid or make a chain of polypeptide directed by mRNA. These processes involve several indispensable that catalyze the ligation of nucleotides or amino 18 acids. Mimicking what nature does, researchers take the advantage of nucleic acid template effect and perform in vitro non-enzymatic oligonucleotide ligation,85,86 PNA ligation,87-89 small-molecule synthesis,90,91 peptide ligation,92,93 reporter transfer94 for the purpose of prebiotic study,85,95,96 drug screening,90,91 mutation diagnosis97 and so on. Von

Kiedrowski studied a self-replicating system in which a hexadeoxynucleotide assembled two trideoxynucleotides through base-pairing and catalyzed the ligation of the two trideoxynucleotides via nucleophilic attack of a pre-activated phosphate group, resulting in a native phosphate linkage.98,99 Orgel and co-workers demonstrated the feasibility of template effect and information transfer between PNA, RNA and DNA.100-102 Szostack and co-workers pioneered non-enzymatic primer extension to study RNA self-replication.103,104

Liu and co-workers developed a method for reaction discovery105 and selection of macrocycles90 encoded by DNA library. Liu’s work is an excellent combination of nucleic acid template effect and in vitro selection of nucleic acid library.

In most examples of nucleic acid-templated reactions, one single strand of oligonucleotide serves as the template and the recognition is promoted by Watson-Crick base-pairing.

However, other recognition patterns can also be used for templated reactions. For example, Nicolaou and co-workers employed DNA duplex as the template for a self- replication system.106,107 Ladame and co-workers designed DNA-templated fluorogenic reaction for the synthesis of trimethine cyanine dyes to monitor the formation of G- quadruplex.108 A number of aqueous compatible chemical reactions can be used for nucleic acid-templated reactions and thus the resulted linkers are versatile, such as native phosphate109,101 and amide formed by direct aminolysis110 or native chemical ligation,111 as well as phosphorothioate112 / phosphoramide113 resulted from nucleophilic substitution, secondary amine prepared by reductive amination,114,115 olefin yielded by Wittig reaction.116,117 19

a

b

Figure 1.12 Examples of nucleic acid template effect. (a) Templated ligation in DNA,

RNA and PNA. (b) DNA-templated group transfer to facilitate the synthesis of

macrocycle library and in vitro selection of macrocycle library. Cited from reference 98

and 100.

These nucleic acid-templated reactions usually occur efficiently at low micromolar to nanomolar range, while regular organic reactions occur at milimolar range. The reason

20 behind nucleic acid template effect could be accounted by effective molarity: the template assembles reactants onto itself, arranging to the favored orientations and proximity, as well as increasing the effective molarity of each reactant.118 Therefore, reactions take place efficiently at significantly low absolute concentrations.

1.6 References

(1) Crick, F. H. C.; Watson, J. D. The Complementary Structure of Deoxyribonucleic

Acid. Proceedings of the Royal Society of London A: Mathematical, Physical and

Engineering Sciences 1954, 223, 80–96.

(2) Wilkins, M. H. F.; Stokes, A. R.; Wilson, H. R. Molecular Structure of Deoxypentose

Nucleic Acids. Nature 1953, 171, 738–740.

(3) Rana, T. M. Illuminating the Silence: Understanding the Structure and Function of

Small . Nat. Rev. Mol. Biol. 2007, 8, 23–36.

(4) Watson, J. D.; Crick, F. H. C.; Others. Molecular Structure of Nucleic Acids. Nature

1953, 171, 737–738.

(5) Wang, J. C. Helical Repeat of DNA in Solution. Proc. Natl. Acad. Sci. U. S. A. 1979,

76, 200–203.

(6) Dervan, P. B. Molecular Recognition of DNA by Small Molecules. Bioorg. Med.

Chem. 2001, 9, 2215–2235.

(7) White, S.; Szewczyk, J. W.; Turner, J. M.; Baird, E. E.; Dervan, P. B. Recognition of

the Four Watson-Crick Base Pairs in the DNA Minor Groove by Synthetic Ligands.

Nature 1998, 391, 468–471.

(8) Renneberg, D.; Dervan, P. B. Imidazopyridine / Pyrrole and Hydroxybenzimidazole

/ pyrrole Pairs for DNA Minor Groove Recognition. J. Am. Chem. Soc. 2003, 125,

5707–5716. 21

(9) Dervan, P. B.; Edelson, B. S. Recognition of the DNA Minor Groove by Pyrrole-

Imidazole Polyamides. Curr. Opin. Struct. Biol. 2003, 13, 284–299.

(10) François, J. C.; Saison-Behmoaras, T.; Hélène, C. Sequence-Specific Recognition

of the Major Groove of DNA by Oligodeoxynucleotides via Triple Helix Formation.

Footprinting Studies. Nucleic Acids Res. 1988, 16, 11431–11440.

(11) Frank-Kamenetskii, M. D.; Mirkin, S. M. Triplex DNA Structures. Annu. Rev. Biochem.

1995, 64, 65–95.

(12) Duca, M.; Vekhoff, P.; Oussedik, K.; Halby, L.; Arimondo, P. B. The Triple Helix: 50

Years Later, the Outcome. Nucleic Acids Res. 2008, 36, 5123–5138.

(13) Arimondo, P. B.; Barcelo, F.; Sun, J. S.; Maurizot, J. C.; Garestier, T.; Hélène, C.

Triple Helix Formation by (G,A)-Containing Oligonucleotides: Asymmetric Sequence

Effect. Biochemistry 1998, 37, 16627–16635.

(14) Beal, P. A.; Dervan, P. B. Second Structural Motif for Recognition of DNA by

Oligonucleotide-Directed Triple-Helix Formation. Science 1991, 251, 1360–1363.

(15) Song, G.; Ren, J. Recognition and Regulation of Unique Nucleic Acid Structures by

Small Molecules. Chem. Commun. 2010, 46, 7283–7294.

(16) Lyamichev, V. I.; Mirkin, S. M.; Frank-Kamenetskii, M. D.; Cantor, C. R. A Stable

Complex between Homopyrimidine and the Homologous Regions of

Duplex . Nucleic Acids Res. 1988, 16, 2165–2178.

(17) Blume, S. W.; Lebowitz, J.; Zacharias, W.; Guarcello, V.; Mayfield, C. A.; Ebbinghaus,

S. W.; Bates, P.; Jones, D. E., Jr; Trent, J.; Vigneswaran, N.; Miller, D. M. The

Integral Divalent Cation within the Intermolecular Purine*purine. Pyrimidine

Structure: A Variable Determinant of the Potential for and Characteristics of the

Triple Helical Association. Nucleic Acids Res. 1999, 27, 695–702.

22

(18) Thomas, T.; Thomas, T. J. Selectivity of Polyamines in Triplex DNA Stabilization.

Biochemistry 1993, 32, 14068–14074.

(19) Takabatake, T.; Asada, K.; Uchimura, Y.; Ohdate, M.; Kusukawa, N. The Use of

Purine-Rich Oligonucleotides in Triplex-Mediated DNA Isolation and Generation of

Unidirectional Deletions. Nucleic Acids Res. 1992, 20, 5853–5854.

(20) Pilch, D. S.; Levenson, C.; Shafer, R. H. Structural Analysis of the (dA) 10.2 (dT) 10

Triple Helix. Proceedings of the National Academy of Sciences 1990, 87, 1942–

1946.

(21) Goñi, J. R.; de la Cruz, X.; Orozco, M. Triplex‐forming Oligonucleotide Target

Sequences in the Human Genome. Nucleic Acids Res. 2004, 32, 354–360.

(22) Goñi, J. R.; Vaquerizas, J. M.; Dopazo, J.; Orozco, M. Exploring the Reasons for the

Large of Triplex-Forming Oligonucleotide Target Sequences in the Human

Regulatory Regions. BMC Genomics 2006, 7, 63.

(23) Durland, R. H.; Kessler, D. J.; Gunnell, S.; Duvic, M.; Pettitt, B. M.; Hogan, M. E.

Binding of Triple Helix Forming Oligonucleotides to Sites in Gene Promoters.

Biochemistry 1991, 30, 9246–9255.

(24) Soto, A. M.; Loo, J. Energetic Contributions for the Formation of TAT/TAT,

TAT/CGC+, and CGC+/CGC+ Base Triplet Stacks. J. Am. Chem. Soc. 2002, 124,

14355–14363.

(25) Bochman, M. L.; Paeschke, K.; Zakian, V. A. DNA Secondary Structures: Stability

and Function of G-Quadruplex Structures. Nat. Rev. Genet. 2012, 13, 770–780.

(26) Burge, S.; Parkinson, G. N.; Hazel, P.; Todd, A. K.; Neidle, S. Quadruplex DNA:

Sequence, Topology and Structure. Nucleic Acids Res. 2006, 34, 5402–5415.

(27) Todd, A. K.; Johnston, M.; Neidle, S. Highly Prevalent Putative Quadruplex

Sequence Motifs in Human DNA. Nucleic Acids Res. 2005, 33, 2901–2907. 23

(28) Huppert, J. L.; Balasubramanian, S. G-Quadruplexes in Promoters throughout the

Human Genome. Nucleic Acids Res. 2007, 35, 406–413.

(29) Paeschke, K.; Simonsson, T.; Postberg, J.; Rhodes, D.; Lipps, H. J. Telomere End-

Binding Proteins Control the Formation of G-Quadruplex DNA Structures in Vivo.

Nat. Struct. Mol. Biol. 2005, 12, 847–854.

(30) Tinoco, I., Jr; Bustamante, C. How RNA Folds. J. Mol. Biol. 1999, 293, 271–281.

(31) Tian, B.; Bevilacqua, P. C.; Diegelman-Parente, A.; Mathews, M. B. The Double-

Stranded-RNA-Binding Motif: Interference and Much More. Nat. Rev. Mol. Cell Biol.

2004, 5, 1013–1023.

(32) Salazar, M.; Fedoroff, O. Y.; Miller, J. M.; Ribeiro, N. S.; Reid, B. R. The DNA Strand

in DNA.cntdot.RNA Hybrid Duplexes Is Neither B-Form nor A-Form in Solution.

Biochemistry 1993, 32, 4207–4215.

(33) Theimer, C. A.; Blois, C. A.; Feigon, J. Structure of the Human Telomerase RNA

Pseudoknot Reveals Conserved Tertiary Interactions Essential for Function. Mol.

Cell 2005, 17, 671–682.

(34) Ishitani, R.; Nureki, O.; Nameki, N.; Okada, N.; Nishimura, S.; Yokoyama, S.

Alternative Tertiary Structure of tRNA for Recognition by a Posttranscriptional

Modification Enzyme. Cell 2003, 113, 383–394.

(35) Fedor, M. J.; Williamson, J. R. The Catalytic Diversity of RNAs. Nat. Rev. Mol. Cell

Biol. 2005, 6, 399–412.

(36) Michel, T.; Debart, F.; Heitz, F.; Vasseur, J.-J. Highly Stable DNA Triplexes Formed

with Cationic Phosphoramidate Pyrimidine Alpha-Oligonucleotides. Chembiochem

2005, 6, 1254–1262.

(37) Michel, T.; Martinand-Mari, C.; Debart, F.; Lebleu, B.; Robbins, I.; Vasseur, J.-J.

Cationic Phosphoramidate Alpha-Oligonucleotides Efficiently Target Single- 24

Stranded DNA and RNA and Inhibit Hepatitis C Virus IRES-Mediated Translation.

Nucleic Acids Res. 2003, 31, 5282–5290.

(38) Seidman, M. M.; Puri, N.; Majumdar, A.; Cuenoud, B.; Miller, P. S.; Alam, R. The

Development of Bioactive Triple Helix-Forming Oligonucleotides. Ann. N. Y. Acad.

Sci. 2005, 1058, 119–127.

(39) Petersen, M.; Wengel, J. LNA: A Versatile Tool for Therapeutics and Genomics.

Trends Biotechnol. 2003, 21, 74–81.

(40) Højland, T.; Kumar, S.; Babu, B. R.; Umemoto, T.; Albaek, N.; Sharma, P. K.; Nielsen,

P.; Wengel, J. LNA (locked Nucleic Acid) and Analogs as Triplex-Forming

Oligonucleotides. Org. Biomol. Chem. 2007, 5, 2375–2379.

(41) Brunet, E.; Alberti, P.; Perrouault, L.; Babu, R.; Wengel, J.; Giovannangeli, C.

Exploring Cellular Activity of Locked Nucleic Acid-Modified Triplex-Forming

Oligonucleotides and Defining Its Molecular Basis. J. Biol. Chem. 2005, 280, 20076–

20085.

(42) Sun, B.-W.; Babu, B. R.; Sørensen, M. D.; Zakrzewska, K.; Wengel, J.; Sun, J.-S.

Sequence and pH Effects of LNA-Containing Triple Helix-Forming Oligonucleotides:

Physical Chemistry, Biochemistry, and Modeling Studies. Biochemistry 2004, 43,

4160–4169.

(43) Aviñó, A.; Cubero, E.; González, C.; Eritja, R.; Orozco, M. Antiparallel Triple Helices.

Structural Characteristics and Stabilization by 8-Amino Derivatives. J. Am. Chem.

Soc. 2003, 125, 16127–16138.

(44) Xodo, L. E.; Manzini, G.; Quadrifoglio, F.; van der Marel, G. A.; van Boom, J. H.

Effect Pf 5-Methylcytosine on the Stability of Triple-Stranded DNA—a

Thermodynamic Study. Nucleic Acids Res. 1991, 19, 5625–5631.

25

(45) Strobel, S. A.; Dervan, P. B. Single-Site Enzymatic Cleavage of Yeast Genomic DNA

Mediated by Triple Helix Formation. Nature 1991, 350, 172–174.

(46) Lee, J. S.; Woodsworth, M. L.; Latimer, L. J.; Morgan, A. R. Poly(pyrimidine).

Poly(purine) Synthetic DNAs Containing 5-Methylcytosine Form Stable Triplexes at

Neutral pH. Nucleic Acids Res. 1984, 12, 6603–6614.

(47) Spacková, N.; Cubero, E.; Sponer, J.; Orozco, M. Theoretical Study of the Guanine

--> 6-Thioguanine Substitution in Duplexes, Triplexes, and Tetraplexes. J. Am.

Chem. Soc. 2004, 126, 14642–14650.

(48) Malnuit, V.; Duca, M.; Benhida, R. Targeting DNA Base Pair Mismatch with Artificial

Nucleobases. Advances and Perspectives in Triple Helix Strategy. Org. Biomol.

Chem. 2011, 9, 326–336.

(49) Nielsen, P. E.; Egholm, M.; Buchardt, O. Peptide Nucleic Acid (PNA). A DNA Mimic

with a Peptide Backbone. Bioconjug. Chem. 1994, 5, 3–7.

(50) Nielsen, P. E.; Haaima, G. Peptide Nucleic Acid (PNA). A DNA Mimic with a

Pseudopeptide Backbone. Chem. Soc. Rev. 1997, 26, 73–78.

(51) Nielsen, P. E. Peptide Nucleic Acid. A Molecule with Two Identities. Acc. Chem. Res.

1999, 32, 624–630.

(52) Egholm, M.; Buchardt, O.; Christensen, L.; Behrens, C.; Freier, S. M.; Driver, D. A.;

Berg, R. H.; Kim, S. K.; Norden, B.; Nielsen, P. E. PNA Hybridizes to Complementary

Oligonucleotides Obeying the Watson-Crick Hydrogen-Bonding Rules. Nature 1993,

365, 566–568.

(53) Larsen, H. J.; Nielsen, P. E. Transcription-Mediated Binding of Peptide Nucleic Acid

(PNA) to Double-Stranded DNA: Sequence-Specific Suicide Transcription. Nucleic

Acids Res. 1996, 24, 458–463.

26

(54) Elayadi, A. N.; Corey, D. R. Application of PNA and LNA Oligomers to Chemotherapy.

Curr. Opin. Investig. Drugs 2001, 2, 558–561.

(55) Bentin, T.; Larsen, H. J.; Nielsen, P. E. Combined Triplex/duplex Invasion of Double-

Stranded DNA by “Tail-Clamp” Peptide Nucleic Acid. Biochemistry 2003, 42,

13987–13995.

(56) Hansen, M. E.; Bentin, T.; Nielsen, P. E. High-Affinity Triplex Targeting of Double

Stranded DNA Using Chemically Modified Peptide Nucleic Acid Oligomers. Nucleic

Acids Res. 2009, 37, 4498–4507.

(57) Nielsen, P. E.; Egholm, M.; Berg, R. H.; Buchardt, O. Sequence-Selective

Recognition of DNA by Strand Displacement with a Thymine-Substituted Polyamide.

Science 1991, 254, 1497–1500.

(58) Betts, L.; Josey, J. A.; Veal, J. M.; Jordan, S. R. A Nucleic Acid Triple Helix Formed

by a Peptide Nucleic Acid-DNA Complex. Science 1995, 270, 1838–1841.

(59) Norton, J. C.; Piatyszek, M. A.; Wright, W. E.; Shay, J. W.; Corey, D. R. Inhibition of

Human Telomerase Activity by Peptide Nucleic Acids. Nat. Biotechnol. 1996, 14,

615–619.

(60) Demidov, V.; Frank-Kamenetskii, M. D.; Egholm, M.; Buchardt, O.; Nielsen, P. E.

Sequence Selective Double Strand DNA Cleavage by Peptide Nucleic Acid (PNA)

Targeting Using S1. Nucleic Acids Res. 1993, 21, 2103–2107.

(61) Ray, A.; Nordén, B. Peptide Nucleic Acid (PNA): Its Medical and Biotechnical

Applications and Promise for the Future. The FASEB Journal 2000, 14, 1041–1060.

(62) Boffa, L. C.; Morris, P. L.; Carpaneto, E. M.; Louissaint, M.; Allfrey, V. G. Invasion

of the CAG Triplet Repeats by a Complementary Peptide Nucleic Acid Inhibits

Transcription of the Androgen Receptor and TATA-Binding Genes and

27

Correlates with Refolding of an Active Nucleosome Containing a Unique AR Gene

Sequence. J. Biol. Chem. 1996, 271, 13228–13233.

(63) Knudsen, H.; Nielsen, P. E. Antisense Properties of Duplex- and Triplex-Forming

PNAs. Nucleic Acids Res. 1996, 24, 494–500.

(64) Cogoi, S.; Codognotto, A.; Rapozzi, V.; Xodo, L. E. Antigene Property of PNA

Conjugated to the Nuclear Localization Signal Peptide. Nucleotides

Nucleic Acids 2005, 24, 971–974.

(65) Cutrona, G.; Carpaneto, E. M.; Ulivi, M.; Roncella, S.; Landt, O.; Ferrarini, M.; Boffa,

L. C. Effects in Live Cells of a c-Myc Anti-Gene PNA Linked to a Nuclear Localization

Signal. Nat. Biotechnol. 2000, 18, 300–303.

(66) Tonelli, R.; Purgato, S.; Camerin, C.; Fronza, R.; Bologna, F.; Alboresi, S.; Franzoni,

M.; Corradini, R.; Sforza, S.; Faccini, A.; et al. Anti-Gene Peptide Nucleic Acid

Specifically Inhibits MYCN Expression in Human Neuroblastoma Cells Leading to

Cell Growth Inhibition and Apoptosis. Mol. Cancer Ther. 2005, 4, 779–786.

(67) Branda, N.; Kurz, G.; Lehn, J.-M. JANUS WEDGES: A New Approach towards

Nucleobase-Pair Recognition. Chem. Commun. 1996, 2443–2444.

(68) Arambula, J. F.; Ramisetty, S. R.; Baranger, A. M.; Zimmerman, S. C. A Simple

Ligand That Selectively Targets CUG Trinucleotide Repeats and Inhibits MBNL

Protein Binding. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 16068–16073.

(69) Jahromi, A. H.; Honda, M.; Zimmerman, S. C.; Spies, M. Single-Molecule Study of

the CUG Repeat--MBNL1 Interaction and Its Inhibition by Small Molecules. Nucleic

Acids Res. 2013, 41, 6687–6697.

(70) Nguyen, L.; Lee, J.; Wong, C.-H.; Zimmerman, S. C. Small Molecules That Target

the Toxic RNA in Myotonic Dystrophy Type 2. ChemMedChem 2014, 9, 2455–2462.

28

(71) Artigas, G.; Marchán, V. Synthesis of Janus Compounds for the Recognition of G-

U Mismatched Nucleobase Pairs. J. Org. Chem. 2013, 78, 10666–10677.

(72) Shin, D.; Tor, Y. Bifacial as a Surrogate for Both T and A in Duplex DNA.

J. Am. Chem. Soc. 2011, 133, 6926–6929.

(73) Largy, E.; Liu, W.; Hasan, A.; Perrin, D. M. Base-Pairing Behavior of a Carbocyclic

Janus-AT Nucleoside Analogue Capable of Recognizing A and T within a DNA

Duplex. Chembiochem 2013, 14, 2199–2208.

(74) Chen, D.; Meena, M.; Sharma, S. K.; McLaughlin, L. W. Formation and Stability of

a Janus-Wedge Type of DNA Triplex. J. Am. Chem. Soc. 2004, 126, 70–71.

(75) Chen, H.; McLaughlin, L. W. A Janus-Wedge DNA Triplex with A-W1-T and G-W2-

C Base Triplets. J. Am. Chem. Soc. 2008, 130, 13190–13191.

(76) Zeng, Y.; Pratumyot, Y.; Piao, X.; Bong, D. Discrete Assembly of Synthetic Peptide-

DNA Triplex Structures from Polyvalent Melamine-Thymine Bifacial Recognition. J.

Am. Chem. Soc. 2012, 134, 832–835.

(77) Mao, J.; Bong, D. Synthesis of DNA-Binding Peptoids. Synlett 2015, 26, 1581–1585.

(78) Zhou, Z.; Xia, X.; Bong, D. Synthetic Polymer Hybridization with DNA and RNA

Directs Nanoparticle Loading, Silencing Delivery, and Aptamer Function. J. Am.

Chem. Soc. 2015, 137, 8920–8923.

(79) Xia, X.; Piao, X.; Fredrick, K.; Bong, D. Bifacial PNA Complexation Inhibits Enzymatic

Access to DNA and RNA. Chembiochem 2014, 15, 31–36.

(80) Xia, X.; Piao, X.; Bong, D. Bifacial Peptide Nucleic Acid as an Allosteric Switch for

Aptamer and Ribozyme Function. J. Am. Chem. Soc. 2014, 136, 7265–7268.

(81) Vysabhattar, R.; Ganesh, K. N. Cyanuryl Peptide Nucleic Acid: Synthesis and DNA

Complexation Properties. Tetrahedron Lett. 2008, 49, 1314–1318.

29

(82) Mittapalli, G. K.; Reddy, K. R.; Xiong, H.; Munoz, O.; Han, B.; De Riccardis, F.;

Krishnamurthy, R.; Eschenmoser, A. Mapping the Landscape of Potentially

Primordial Informational Oligomers: Oligodipeptides and Oligodipeptoids Tagged

with as Recognition Elements. Angew. Chem. Int. Ed. 2007, 46, 2470–

2477.

(83) Piao, X.; Xia, X.; Bong, D. Bifacial Peptide Nucleic Acid Directs Cooperative Folding

and Assembly of Binary, Ternary, and Quaternary DNA Complexes. Biochemistry

2013, 52, 6313–6323.

(84) Hysell, M.; Siegel, J. S.; Tor, Y. Synthesis and Stability of Exocyclic Triazine

Nucleosides. Org. Biomol. Chem. 2005, 3, 2946–2952.

(85) Kozlov, I. A.; Pitsch, S.; Orgel, L. E. Oligomerization of Activated D- and L-Guanosine

Mononucleotides on Templates Containing D- and L-Deoxycytidylate Residues.

Proceedings of the National Academy of Sciences 1998, 95, 13448–13452.

(86) Gao, K.; Orgel, L. E. Polyphosphorylation and Non-Enzymatic Template-Directed

Ligation of Oligonucleotides. Orig. Life Evol. Biosph. 2000, 30, 45–51.

(87) Rosenbaum, D. M.; Liu, D. R. Efficient and Sequence-Specific DNA-Templated

Polymerization of Peptide Nucleic Acid . J. Am. Chem. Soc. 2003, 125,

13924–13925.

(88) Heemstra, J. M.; Liu, D. R. Templated Synthesis of Peptide Nucleic Acids via

Sequence-Selective Base-Filling Reactions. J. Am. Chem. Soc. 2009, 131, 11347–

11349.

(89) Roloff, A.; Seitz, O. Bioorthogonal Reactions Challenged: DNA Templated Native

Chemical Ligation during PCR. Chem. Sci. 2012, 4, 432–436.

30

(90) Gartner, Z. J.; Tse, B. N.; Grubina, R.; Doyon, J. B.; Snyder, T. M.; Liu, D. R. DNA-

Templated Organic Synthesis and Selection of a Library of Macrocycles. Science

2004, 305, 1601–1605.

(91) Gartner, Z. J.; Liu, D. R. The Generality of DNA-Templated Synthesis as a Basis for

Evolving Non-Natural Small Molecules. J. Am. Chem. Soc. 2001, 123, 6961–6963.

(92) Erben, A.; Grossmann, T. N.; Seitz, O. DNA-Triggered Synthesis and Bioactivity of

Proapoptotic Peptides. Angew. Chem. Int. Ed. 2011, 50, 2828–2832.

(93) McKee, M. L.; Evans, A. C.; Gerrard, S. R.; O’Reilly, R. K.; Turberfield, A. J.; Stulz,

E. Peptidomimetic Bond Formation by DNA-Templated Acyl Transfer. Org. Biomol.

Chem. 2011, 9, 1661–1666.

(94) Grossmann, T. N.; Seitz, O. DNA-Catalyzed Transfer of a Reporter Group. J. Am.

Chem. Soc. 2006, 128, 15596–15597.

(95) Joyce, G. F. Nonenzymatic Template-Directed Synthesis of Informational

Macromolecules. Cold Spring Harb. Symp. Quant. Biol. 1987, 52, 41–51.

(96) Bolli, M.; Micura, R.; Eschenmoser, A. Pyranosyl-RNA: Chiroselective Self-Assembly

of Base Sequences by Ligative Oligomerization of Tetranucleotide-2’,3'-

Cyclophosphates (with a Commentary Concerning the Origin of Biomolecular

Homochirality). Chem. Biol. 1997, 4, 309–320.

(97) Xu, Y.; Karalkar, N. B.; Kool, E. T. Nonenzymatic Autoligation in Direct Three-Color

Detection of RNA and DNA Point Mutations. Nat. Biotechnol. 2001, 19, 148–152.

(98) von Kiedrowski, G. A Self-Replicating Hexadeoxynucleotide. Angew. Chem. Int. Ed

Engl. 1986, 25, 932–935.

(99) Sievers, D.; von Kiedrowski, G. Self-Replication of Complementary Nucleotide-

Based Oligomers. Nature 1994, 369, 221–224.

31

(100) Böhler, C.; Nielsen, P. E.; Orgel, L. E. Template Switching between PNA and RNA

Oligonucleotides. Nature 1995, 376, 578–581.

(101) Schmidt, J. G.; Nielsen, P. E.; Orgel, L. E. Information Transfer from Peptide Nucleic

Acids to RNA by Template-Directed Syntheses. Nucleic Acids Res. 1997, 25, 4797–

4802.

(102) Schmidt, J. G.; Christensen, L.; Nielsen, P. E.; Orgel, L. E. Information Transfer from

DNA to Peptide Nucleic Acids by Template-Directed Syntheses. Nucleic Acids Res.

1997, 25, 4792–4796.

(103) Rohatgi, R.; Bartel, D. P.; Szostak, J. W. Kinetic and Mechanistic Analysis of

Nonenzymatic, Template-Directed Oligoribonucleotide Ligation. J. Am. Chem. Soc.

1996, 118, 3332–3339.

(104) Rajat Rohatgi; David P. Bartel, †.; Szostak*, J. W. Nonenzymatic, Template-

Directed Ligation of Oligoribonucleotides Is Highly Regioselective for the Formation

of 3‘−5’ Phosphodiester Bonds. J. Am. Chem. Soc. 1996, 118, 3340–3344.

(105) Kanan, M. W.; Rozenman, M. M.; Sakurai, K.; Snyder, T. M.; Liu, D. R. Reaction

Discovery Enabled by DNA-Templated Synthesis and in Vitro Selection. Nature

2004, 431, 545–549.

(106) Li, T.; Nicolaou, K. C. Chemical Self-Replication of Palindromic Duplex DNA. Nature

1994, 369, 218–221.

(107) Li, T.; Weinstein, D. S.; Nicolaou, K. The Chemical End-Ligation of Homopyrimidine

Oligodeoxyribonucleotides within a DNA Triple Helix. Chem. Biol. 1997, 4, 209–214.

(108) Meguellati, K.; Koripelly, G.; Ladame, S. DNA-Templated Synthesis of Trimethine

Cyanine Dyes: A Versatile Fluorogenic Reaction for Sensing G-Quadruplex

Formation. Angew. Chem. Int. Ed. 2010, 49, 2738–2742.

32

(109) Izgu, E. C.; Fahrenbach, A. C.; Zhang, N.; Li, L.; Zhang, W.; Larsen, A. T.; Blain, J.

C.; Szostak, J. W. Uncovering the Thermodynamics of Monomer Binding for RNA

Replication. J. Am. Chem. Soc. 2015, 137, 6373–6382.

(110) Bruick, R. K.; Dawson, P. E.; Kent, S. B.; Usman, N.; Joyce, G. F. Template-Directed

Ligation of Peptides to Oligonucleotides. Chem. Biol. 1996, 3, 49–56.

(111) Ficht, S.; Dose, C.; Seitz, O. As Fast and Selective as Enzymatic Ligations:

Unpaired Nucleobases Increase the Selectivity of DNA-Controlled Native Chemical

PNA Ligation. Chembiochem 2005, 6, 2098–2103.

(112) Xu, Y.; Kool, E. T. Chemical and Enzymatic Properties of Bridging 5’-S-

Phosphorothioester Linkages in DNA. Nucleic Acids Res. 1998, 26, 3159–3164.

(113) Izgu, E. C.; Oh, S. S.; Szostak, J. W. Synthesis of Activated 3′-Amino-3′-Deoxy-2-

Thio-Thymidine, a Superior Substrate for the Nonenzymatic Copying of Nucleic Acid

Templates. Chem. Commun. 2016, 52, 3684–3686.

(114) Goodwin, J. T.; Lynn, D. G. Template-Directed Synthesis: Use of a Reversible

Reaction. J. Am. Chem. Soc. 1992, 114, 9197–9198.

(115) Xiaoyu Li; Zheng-Yun J. Zhan; Rachel Knipe; Lynn, D. G. DNA-Catalyzed

Polymerization. J. Am. Chem. Soc. 2002, 124, 746–747.

(116) McKee, M. L.; Milnes, P. J.; Bath, J.; Stulz, E.; O’Reilly, R. K.; Turberfield, A. J.

Programmable One-Pot Multistep Organic Synthesis Using DNA Junctions. J. Am.

Chem. Soc. 2012, 134, 1446–1449.

(117) McKee, M. L.; Milnes, P. J.; Bath, J.; Stulz, E.; Turberfield, A. J.; O’Reilly, R. K.

Multistep DNA-Templated Reactions for the Synthesis of Functional Sequence

Controlled Oligomers. Angew. Chem. Int. Ed Engl. 2010, 49, 7948–7951.

33

(118) Li, X.; Liu, D. R. DNA-Templated Organic Synthesis: Nature’s Strategy for

Controlling Chemical Reactivity Applied to Synthetic Molecules. Angew. Chem. Int.

Ed. 2004, 43, 4848–4870.

34

Chapter 2. Bifacial Peptide Nucleic Acid Directs Cooperative Folding and

Assembly of Binary, Ternary, and Quaternary DNA Complexes

35

2.1 Introduction

Triplex formed from nucleic acid or its analogue has inspired decades of research aimed at understanding triplex formation1 and targeting the information presented in the nucleic acid duplex.2,3 Nucleic acid triplex formation usually involves Hoogsteen base-pairing of a single-stranded oligonucleotide to a preformed duplex paired by Watson-Crick recognition.

In particular, a homopurine or homopyrimidine oligonucleotide is inserted into the major groove of a duplex formed by a homopurine and homopyrimidine oligonucleotides. Typical forms of nucleobase triplet include T-A*T, C-G*C+, T-A*A and C-G*G. Triplex formation often occurs under non-physiological pH and salt conditions, because cytosine in the third strand needs to be protonated under acidic pH to form Hoogsteen hydrogen bonds with guanine. Higher salt concentration and divalent salts4,5 are usually necessary to shield the electrostatic repulsion of the highly negatively-charged backbones. We recently reported an alternative strategy for a new class of DNA triplex formation involving synthetic melamine-displaying peptide and polythymine oligonuclotide that did not require any prior

6 secondary structures. Two unstructured (dT)10 were assembled onto a bridging melamine-displaying peptide to form a 2:1 DNA-peptide heterotriplex via melamine- thymine bifacial recognition. When the two (dT)10 strands were connected by a non- interacting loop, a 1:1 DNA-peptide triplex stem-loop structure formed with higher affinity compared to its non-looped counterpart. We therefore term it bifacial peptide nucleic acid

(bPNA). Unlike DNA triplex or triplex involving native nucleobases, bPNA does not self- assemble, use Hoogsteen base-pairing or require direct interaction between DNA strands and instead targets unstructured T-tracts for binding. Bifacial PNA is an artificial peptide strand that binds and folds a DNA partner through the use of melamine as a synthetic nucleobase mimic that hydrogen bonds with two thymine bases on their Watson-Crick binding faces. This bifacial recognition concept is similar to intramolecular DNA triplexes3 36 and synthetic Janus-wedge bifacial nucleobase mimics,7-9 as well as the targeting of short, single-stranded oligopurines with pyrimidine “clamps” or cyclic DNA10 with the distinction that no prior secondary structure, covalent cyclization, or Mg2+ is needed. Different from conventional PNA with a non-native polypeptide backbone and native nucleobase content, bPNA has an α-peptide structure derived from native amino acids and a non-native bifacial melamine nucleobase mimic. Bifacial PNA is also operationally distinguished from conventional PNA in that bPNA is an associative agent that brings noninteracting oligothymidine DNA strands together in 1:2 bPNA-DNA triplex, whereas PNA is often used as a dissociative agent that invades existing nucleic acid (NA) duplexes to produce 2:1

PNA-NA triplex hybrids.11-14

The melamine bPNAs are easily prepared via reaction of the ε-amino group of lysine with diaminochlorotriazine to yield an amino acid derivative for standard solid-phase peptide synthesis.15 The facility of direct amino substitution on chlorotriazine rings has led to considerable investigation of melamine as a diversity platform for combinatorial chemistry and dendrimer synthesis,16,17 and triazines themselves have been suggested as products of prebiotic chemistry,18-21 forming readily from cyano derivatives.22-26 The synthetic triaminotriazine (melamine) provides a divalent site for thymine recognition when monoderivatized on an exocyclic nitrogen. Melamine and its derivatives have been extensively studied in organic solvents and the solid-state as an assembly partner for cyanuric acid derivatives,27-29 and more recently in aqueous-phase assemblies.30-41

Heterocycle recognition of melamine and cyanurates is directed by hydrogen bond recognition, but driven by exothermic base-stacking akin to DNA-recognition.39 Co-

37 a b

c

Figure 2.1 Illustration of melamine-thymine interaction directing DNA and bPNA assembly and CD spectra. a) The structure of bPNA with N- and C- capping groups shown. Cbf = 5(6)-carboxyfluorescein, ABA = 4-acetamidobenzamide. b) Schematics of

1:1 bPNA-DNA hairpin triplex structures. c) CD spectra of single-stranded dTnC4Tn and the corresponding bPNA-DNA complex. The concentration is 5 μM.

38 crystallization of melamine with thymine, uracil, and 5-fluorouracil (5-FU)37 from provided structural evidence of the expected multifacial hydrogen bonding patterns between triazine and nucleobase in the solid state. Furthermore, Zimmerman and co-workers had demonstrated that melamine-acridine conjugates selectively intercalated into T-T and U-U mismatch sites in solution,42,43 indicative of bifacial melamine recognition.

Robust exothermic binding of 5-fluorouracil to polyacrylate-displayed melamine was found to trigger condensation of soluble polymer into 5-FU loaded nanoparticles, suggesting that melamine-cyanurate type recognition may be generally extended to include interactions with thymine/uracil.37 Indeed, Eschenmoser and Krishnamurthy prepared triazine- displayed peptide/peptoid that interacts with native DNA/RNA to form duplexes.23,24 They have demonstrated that diaminotriazine-displaying α-peptides function as novel PNAs that form stable peptide-DNA/RNA duplexes. We speculated that replacement of the aliphatic

CH2 linkage to the heterocycle in diaminotriazine with an NH to yield a triaminotriazine

(melamine) side chain derivative would create a bifacial recognition nucleus for thymine and switch the assembly from duplex to triplex structures. This was demonstrated in our

6 earlier study using 10mer bPNA and (dT)10, as discussed above. In this study, we focus our examination on melamine-thymine driven assembly of bPNA-DNA complexes as a function of interface length in the context of the hairpin-triplex stem motif. The dependence of bPNA-DNA binding on the length of the interface was investigated by synthesis of

(EM*)n peptides with n = 4, 5, 6, 8, and 10. These peptides bound to their dTnC4Tn partners with the expected 1:1 hairpin triplex stoichiometry, and complex stability decreased with n. Notably, length matching of the DNA and peptide was critical in determining stoichiometry; longer DNA strands were found to host more than one peptide, supporting the notion that a single melamine repeat unit (EM*) is required to interact with each T-T pair. The recognition interface tolerated a high number of “mismatches” and indicated half- 39 site, or monofacial, recognition between melamine and thymine may occur if only 1 complementary nucleobase was available.

2.2 Results and Discussion

2.2.1 1:1 bPNA-DNA complexes

Bifacial Peptide Nucleic Acid (bPNA) was prepared by standard solid phase synthesis with the general form (EM*)n, wherein M* denotes melamine-lysine residue and E represents glutamic acid. N indicates the number of melamine heterocycles. Glutamic acid was chosen to occupy alternate positions to achieve similar sequence pattern shown by

Eschenmoser23,24 and Ghadiri21, respectively. In addition, acidic side chains are deprotonated in physiological conditions, improving of bPNA and minimizing nonspecific electrostatic interaction when hybridizing with nucleic acid. DNA complexation of 5(6)-Carboxyfluorescein-labeled bPNA (Cbf-bPNA) showed fluorescence quenching and increased fluorescence anisotropy. (EM*)n showed a clear 1:1 complexation with dTnC4Tn by fluorescence titration with the exception of (EM*)4 which exhibited weakest binding and was difficult to determine a clear stoichiometry under the same conditions

(Figure 2.2). Hybridization with bPNA induced a similar structural change in DNA as qualitatively judged by circular dichroism (CD). A large positive Cotton effect at ~280 nm corresponding to unstructured DNA was transformed into a strong negative peak at ~260 nm (Fig. 2.1c). Bifacial PNA itself showed a negligible signal on CD.

40

Figure 2.2 Normalized fluorescence titration curve for Cbf-(EM*)n-dTnC4Tn. The concentration of fluorescent peptide was 1 µM in each titration. (a) n=4, (b) n=5, (c) n=6,

(d) n=8, (e) n=10. Excitation: 492 nm, emission: 525 nm. Titration was done at 25 C.

41

2.2.2 Binding affinity and enthalpy of bPNA-DNA complexes

a Table 2.1 Thermodynamic data for [(EM*)n-dTnC4Tn]

Longer recognition interface generally results in stronger binding as quantitatively judged by UV melting, Kd determination and differential scanning calorimetry (DSC). In particular,

(EM*)n-dTnC4Tn triplexes were melted at different temperatures ranging from 19 ºC to 57

ºC along with the increase of (EM*) units. This melting temperature trend corresponded to enthalpy change of dissociation. A heat of 293 kcal/mol was needed to break (EM*)10- dT10C4T10 complex, whereas the ΔHd of (EM*)4-dT4C4T4 was undetectable under the same conditions. Fluorescence anisotropy was used to obtain dissociation constant (Kd) by fitting to a 1:1 binding model. The longest bPNA (n=10) yielded a Kd of ~2.6 nM. This bPNA-DNA hybridization is robust to survive dA10G4A10 invasion as evidenced by the fact that the intensity of (EM*)10-dT10C4T10 complex did not change despite of increasing amount of dA10G4A10. This is also supported by the distribution of (EM*)10-dT10C4T10 triplex and dT10C4T10-dA10G4A10 duplex when equal amounts of bPNA and two oligonucleotides were denatured and allowed to cool slowly (Lane 1, Figure 2.4). As summarized in Table

1, melting temperatures of bPNA-DNA triplexes drop significantly from n = 10 to 4, though 42 a b

c d

Figure 2.3 Characterization of binding affinity. a) Normalized first derivatives of thermal melts (UV) of the 1:1 (EM*)n-dTnC4Tn complexes at 5 μM. b) Binding isotherms of complex formation fit to a 1:1 binding model (R2 ≥ 0.96) with 25 nM bPNA concentration. c) DSC upscans of the 1:1 (EM*)n-dTnC4Tn complexes at 25 μM. d) Thermodynamic parameters for the complex dissociation (as labeled) and Tm (●) on the right axis as a function of repeat unit.

43

1 2 3 4 5 6 7 8 9 10 11 12 13

Figure 2.4 Invasion of triplex (dT10C4T10+H2N-(EM*)10G) with dA10G4A10. Gel was

stained with SYBR Gold. Lane 1) H2N-(EM*)10G+dT10C4T10+dA10G4A10, 250 nM for

each. 2) dA10G4A10, 500 nM. 3) dA10G4A10+H2N-(EM*)10G, 500 nM for each. 4)

dT10C4T10, 500 nM. 5) DNA duplex (dT10C4T10 + dA10G4A10), 250 nM. 6) Triplex

(dT10C4T10+H2N-(EM*)10G), 500 nM. 7) Triplex (500 nM) + dA10G4A10 (125 nM). 8) Triplex

(500 nM) + dA10G4A10 (250 nM). 9) Triplex (500 nM) + dA10G4A10(375 nM). 10) Triplex

(500 nM) + dA10G4A10(500 nM). 11) Triplex (500 nM) + dA10G4A10(750 nM). 12) Triplex

(500 nM) + dA10G4A10(1000 nM). 13) Triplex (500 nM) + dA10G4A10(2500 nM). The

sample in Lane 1 and the DNA duplex were heated at 95 C for 5 min followed by slowly

cooled to room temperature, respectively. Triplex(dT10C4T10+H2N-(EM*)10G) and all

other samples were heated at 37 C for 1 h, then slowly cooled to room temperature.

Kd increases are less substantial, suggestive of strong enthalpy-entropy compensation.

This was borne out by differential scanning calorimetry analysis, which indicated enthalpically driven44,45 bPNA-DNA base-stacking became markedly less exothermic as the recognition interface decreased. Near perfect enthalpy-entropy compensation was observed as the length of the interface varied, with both heat of binding and entropic cost 44 of assembly sharply decreasing with the number of (EM*) repeat units46,47 The enthalpic benefit of bPNA-DNA binding vanished with an (EM*) repeat of four, as reflected by the significant loss of DNA affinity as bPNA length is decreased from 5 to 4.

2.2.3 Tolerance of “mismatch” sites in (EM*)10 DNA complexation

Although the selectivity of (EM*)10 bPNA for polythymidine DNA relative to the other native

DNA nucleobases has been previously established, the tolerance of non-thymine nucleobases at the recognition interface was unknown. We systematically replaced selected thymine nucleobases with cytosine and tested the affinity of Cbf-10 for a series of 24-nt oligonucleotides based on dT10C4T10, as a function of thymine content. Eighteen

DNA binding partners for Cbf-10 with thymine content decreasing relative to dT10C4T10 were evaluated by Tm measurement to assay the binding affinity (Table 2.2). Notably, complex formation was detectable even with 40% replacement of T for C, though thermal stability dropped to ~19 ºC with this high number of mismatches. Thermal stability of the complexes showed good linear correlation with thymine content (Figure 2.5d). Selected low melting complexes were studied by fluorescence quenching titration to confirm the 1:1 bPNA-DNA stoichiometry of a hairpin triplex, despite the high number of mismatches

(Figure 2.5a-c). Inspection of complex Tm as a function of T→C substitution demonstrates bPNA can bind to single thymine half-sites, presented as T-C or C-T mismatch sites formed when the 24-nt DNA is folded into a hairpin structure. Indeed, two T→C substitutions that yield half-sites (entry 3, Tm = 54.1 ºC) are more thermally stable than substitutions that do not (entry 5, T m = 49.8 ºC). This effect may also be observed when comparing entries 6 and 11, which have identical thymine fraction (67%) yet are separated by 6 ºC in melting temperature. Again, the higher melting complex formed in entry 6

45 distributes the T→C substitutions exclusively into half-sites (T-C, C-T pairs), whereas entry 11 presents two contiguous C-C mismatch sites, effectively dividing the

Table 2.2 Tm of Cbf-10 complexes with T-rich 24-nt DNA

46 original dT10 tract into two T4 tracts separated by CC linkers. Similarly, the favorable presence of T-C, C-T half-sites is seen when comparing entries 12 and 14, both of which are 58 % T. Thus, it appears that clustered T→C substitutions are more damaging to bPNA recognition than distributed substitutions; the latter pattern allows the entire peptide to participate in full site (T-T) or T-C and C-T half-site recognition. With a total of six T-T sites, Cbf-10 binding to entry 16 is analogous to the [Cbf-6 · dT6C4T6] complex

(Table 2.1), which in fact has much higher thermal stability (Tm= 38 ºC). Lower thermal stability relative to [6 · dT 6C4T6] may be a result of the greater entropic cost of complexing larger macromolecules with more degrees of freedom, as well as a

“dangling end” effect in which the terminal bases are not buried on both sides of the T- tract in entry 16.

47

d

Figure 2.5 Fluorescence titration curve for bPNA with mismatched DNA hairpin. The concentration of fluorescent bPNA 10 was 1 µM, excitation: 492 nm, emission: 525 nm.

Titration was done at 25 ºC. (a) dTTCTTCTTCTC4TCTTCTTCTT, Tm was 33.4 ºC, (b) dTTTTCCCTTTC4TTTCCCTTTT, Tm was 29.9 ºC, (c) dTCTTCTTCTTC4TCTCTTCTTT,

Tm was 34.0 ºC, (d) Melting temperature dependence on thymine content.

48

2.2.4 bPNA-DNA higher-order assembly

Table 2.3 Tm’s of higher order bPNA-DNA assembly

The clear dependence of complex stability on the melamine-thymine interface length suggested that higher-order bPNA-DNA assembly may be possible by altering the interface length of bPNA or DNA. We set out to test this notion via two different models:

1) Inserting multiple copies of the same bPNA into a long DNA host; 2) Inserting bPNA fragments of different length into a long DNA host. In the first type of assembly, a long oligonucleotide, dT18C4T18, assembled two copies of Cbf-bPNA when n = 6-10 into ternary complexes, while three Cbf-4 and Cbf-5 were inserted into the same DNA host, resulting in quaternary complexes (Figure 2.6 and 2.7). This type of higher-order assembly showed a single transition with up to 10 ºC increase on melting temperature compared to length- matched binary bPNA-DNA complexation. These stability increases are indicative of strongly cooperative multiplex binding that releases all bound bPNA at once without detectable intermediate complexes.

49

Figure 2.6 Normalized fluorescence titration curve for higher-order assembly. The concentration of Cbf-10 was 1 µM. (a) n=4, dT18C4T18 (b) n=5, dT18C4T18 (c) n=6, dT18C4T18 (d) n=8, dT18C4T18 (e) n=10, dT18C4T18 (f) n=4, dT10C4T10.

50

However, this conclusion does not fit the second type of higher-order assembly in which

(EM*)4 and (EM*)10 were inserted into dT15C4T15. The reason to use dT15C4T15 rather than dT18C4T18 was that dT15C4T15 was not long enough to host two (EM*)10, but only one.

Replacement of the fluorephore with ABA on bPNA 10 permitted the formation of a 1:1 complex between ABA-10 and dT15C4T15 that would be spectroscopically invisible in a

fluorescence quenching titration with Cbf-4. The [ABA-10 · dT15C4T15] complex would thus be preorganized to accept an additional bPNA binding partner in the five unbound T-T sites. Gratifyingly, fluorescence quenching titration of Cbf-4 against preformed [ABA-

10 · dT15C4T15] revealed a clear transition to a 1:1 stoichiometry of interaction with the complex, providing evidence for formation of a 1:1:1 complex of [ABA-10 · Cbf-

4 · dT15C4T15]. This heterotriplex showed two distinct melting temperatures, corresponding to the separate thermally driven dissociations of the two bPNA fragments. A similar titration with Cbf-4 and dT15C4T15 indicated that a 3:1 bPNA-DNA binding. Further investigation by DSC revealed that dissociation of [ABA-10 · dT15C4T15] required 281 kcal/mol, similar to what was observed on analysis of [Cbf-10 · dT10C4T10] by DSC (293 kcal/mol). The heterotrimer [ABA-10 · Cbf-4 · dT15C4T15] exhibited two melting transitions with absorption of 6 kcal/mol in the first transition, followed by 294 kcal/mol in the second transition. These findings are again consistent with the binary, length-matched

[(EM*)n · dTnC4Tn] systems, wherein [Cbf-4 · dT4C4T4] melting did not yield detectable heat release; thus, a modest but detectable improvement in Cbf-4 binding could account for the

6 kcal/mol absorbance in the DSC spectrum. This was due to the pre-organization of [ABA-

10 · dT15C4T15] that lowered the entropic cost for binding to Cbf-4. Calorimetric analysis of

[(Cbf-4)3 · dT15C4T15] yielded the most striking finding: 330 kcal heat per mole of DNA was absorbed on complex dissociation! This value was nearly the same as a full-length bPNA binding with dT15C4T15 (an estimated dissociation enthalpy change of 341 kcal/mol). 51

Figure 2.7 Illustration of higher-order bPNA-DNA assembly

Figure 2.8 Analysis of binary and ternary complexes of bPNA and dT15C4T15. bPNA

Cbf-4 and ABA-10 were used, labeled as follows: (---) = [(Cbf-4)3 · dT15C4T15], (□) = [Cbf-

/ABA-10 · dT15C4T15], and (·) = [Cbf-/ABA-10 · Cbf-4 · dT15C4T15]. (Left) Normalized

fluorescence quenching plots indicating stoichiometry of binding of Cbf-4 as a function of increasing dT15C4T15 (---) and as a function of increasing preformed [ABA-

10 · dT15C4T15] complex (·). (Center) UV melts of dT15C4T15 complexes, as labeled.

(Right) Differential scanning calorimetry (upscans only) of the binary and ternary complexes of dT15C4T15.

52

This indicated that a very strong cooperative assembly of ternary and quaternary bPNA−DNA complexes exhibit enthalpy and thermal stability gains relative to the length- matched bPNA-DNA heterodimers.

2.3 Conclusion

We find optimal bPNA-DNA complexation when the ratio of melamine to thymine bases is

1:2, consistent with interface length-matching selection. Although enthalpy-entropy compensation yielded a slow increase in Kd as the recognition interface was shortened, dissociation enthalpy change and thermal stability exhibited strongly linear positive correlations with the number of melamine binding sites in bPNA and thymine content of the DNA interface. Complexation may thus be easily tuned by the length of the designed interface. Maximum binding efficacy was observed when the numbers of T-T hairpin sites and melamine peptide residues were the same, though a high number of mismatch sites and T-C / C-T half-sites were tolerated before complex Tm dropped below 37 ºC. As both

DNA and bPNA partner develop secondary structure on binding, complex formation is generally cooperative, with the entropic cost of folding paid for by the enthalpic benefit of base-stacking. A striking feature of this system is the high degree of cooperative multisite binding observed in the complexation of long T-tracts with short melamine peptides. This may be rationalized because the first peptide folds thymine-rich DNA into a hairpin structure, aligning additional T-T sites for further melamine peptide binding, resulting in highly cooperative binding. Similar to turn nucleation,48-55 the greatest entropic cost is expected to be associated with the initial formation of the dC4 loop structure to allow antiparallel presentation of the two T-rich hairpin strands; base-stacking fails to compensate for this cost around four triplex repeat units of bPNA-DNA. More structured synthetic multivalent systems that have decoupled folding from binding are known to bind 53 monovalent ligands independently rather than cooperatively.56,57 Overall, the melamine- thymine recognition motif may be utilized to predictably prepare stable binary, ternary and quaternary macromolecular bPNA-DNA complexes with thymine-rich single-stranded

DNA. The structural simplicity and synthetic accessibility of the bPNA strand, coupled with its robust polythymidine recognition properties, raises the intriguing possibility that triazine- bearing macromolecules may be used as biotechnology tools to direct the chemistry of native nucleic acids.

2.4 Materials and Equipments

Chemicals for amino acid derivative synthesis, solid phase peptide synthesis, purification and characterization were purchased from Sigma-Aldrich, Chem-Impex and AAPPTec without further purification unless otherwise specified. Rink Resin LS (100-200 mesh, 0.28 mmol/g) was purchased from Advanced ChemTech. All DNA oligomers were obtained from Sigma-Aldrich. Dulbecco’s Phosphate-Buffered Salines (DPBS 1×, no calcium, no magnesium) used as the binding buffer was purchased from Invitrogen. 40% acrylamide and bis-acrylamide solution (19:1) was purchased from Bio-Rad. TBE Buffer (10× solution) for gel electrophoresis was purchased from American Bioanalytical. SYBR Gold (10000×) was purchased from Invitrogen.

All DNA oligomers and synthetic peptides were quantified on UV-Vis HP 8543. MALDI-

TOF Mass spectra were acquired on Bruker Microflex MALDI-TOF instrument.

Electrospray mass spectroscopy was acquired on Bruker MicroTOF equipped with an source. instruments were provided by a grant from the Ohio BioProducts Innovation Center. NMR spectra were acquired on Bruker

Advance DPX 400 instrument. Solid phase peptide synthesis was performed on AATTPec

Apex 396 peptide synthesizer. All peptides were purified on RP-C18 preparative HPLC 54 column and the purity was confirmed on RP-C18 analytical column. UV melting was performed on Varian Cary-100 UV-Vis Spectrophotometer equipped with Cary

Temperature Controller. Fluorescence titration was performed on Perkin Elmer

Luminescence Spectrometer LS 50B equipped with PTP 1 Temperature Programmer.

Fluorescence anisotropy was performed on Molecular Devices SpectraMax M5. Circular

Dichroism spectra were obtained on Jasco J815 Circular Dichroism Spectrometer.

Differential scanning calorimetry was performed on Microcalorimeter VP-DSC.

Fluorescence gel images were acquired using a Typhoon Trio Variable Mode Imager

(Amersham Biosciences).

2.5 Synthesis

2.5.1 Synthesis of Fmoc-Lys(M*)-OH

Scheme 2.1 Synthesis route of Fmoc-Lys(M*)-OH

Fmoc-Lys(M*)-OH was synthesized based on slight modification of our recent publication.

Briefly, Boc-Lys-OH (4.94 g, 20 mmol) was stirred with 6-chloro-2,4-diamino-1,3,5-triazine

(3.49 g, 24 mmol) and NaOH (1.60 g, 40 mmol) in 35 ml of water at 85 ºC overnight. Solid

55 was removed from the solution by centrifuge and the solution was extracted with EtOAc six times. The aqueous solution was then dried under reduced pressure. Boc protecting group was removed in 15 ml TFA for 2 hours. Most TFA was removed under reduced pressure and the resulting sticky oil was diluted in 75 ml water and neutralized with

Na2CO3 solid carefully to pH 7.5-8, followed by adding 75 ml THF solution of Fmoc-OSu

(7.42 g, 22 mmol) and stirring overnight. The reaction was washed with EtOAC six times to remove unreacted Fmoc-OSu and then acidified with 1 M HCl solution to precipitate product. Off-white solid product (5.60 g, 58% yield) was collect through filtration and washed by water and EtOAc. 1H NMR (400 MHz, d6-DMSO) δ (ppm) 1.35 (d, J=2.8Hz,

2H), 1.50 (t, J=5.4Hz, 2H), 1.62-1.71 (m, 2H), 3.25 (d, J=2.4Hz, 2H), 3.93 (m, 1H), 4.22 (t,

J=5.4Hz, 1H), 4.27-4.30 (m, 2H), 7.31 (t, J=5.8Hz, 2H), 7.40 (t, J=6Hz, 2H), 7.63 (d, J =

8Hz, 1H), 7.72 (m, 2H), 7.88 (m, 6H), 8.09 (s, 1H). 13C NMR (100 MHz, d6-DMSO) δ (ppm)

23.48, 28.80, 30.95, 47.18, 54.30, 66.10, 120.61, 125.78, 127.56, 128.15, 141.23, 144.34,

156.69, 158.96, 174.45. Calculated Mass [M+H]: 478.2205. Found: 478.2382.

56

57

Figure 2.9 1H NMR of Fmoc-Lys(M*)-OH

57

58

Figure 2.10 13C NMR of Fmoc-Lys(M*)-OH

58

2.5.2 Solid phase synthesis of bPNA

Standard Fmoc chemistry was employed. DIC/HOBt were used as the coupling reagents.

And 20% piperidine in NMP was used for Fmoc deprotection. 95% TFA was used to cleave peptide from resin and remove the protecting groups. Resin was removed by filtration through cotton and the peptide was precipitated and washed by cold . Crude peptides were then dissolved in solvent A and purified by HPLC on a C18 reversed phase column using a gradient of 10–50% solvent B in 50 min (solvent A= 0.1% TFA in water, solvent B = 0.01% TFA in 45% acetonitrile, 45% isopropanol, 10% water). The UV detector was set at 238 nm. Purified peptides were lyophilized to dryness.

59

2.5.3 HPLC and Maldi-TOF mass spectra of the synthetic peptides

a

b

Figure 2.11 Characterization of 4mer bPNA Cbf-(β-A)-(EM*)4G. (a) HPLC trace on a

RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-TOF

Mass of Cbf-(β-A)-(EM*)4G. Calculated [M+H]: 1969.976, Found: 1968.735.

60 a

b

Figure 2.12 Characterization of 5mer bPNA Cbf-(β-A)-(EM*)5G. (a) HPLC trace on a

RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-TOF

Mass of Cbf-(β-A)-(EM*)5G. Calculated [M+H]: 2336.352, Found: 2337.701.

61 a

b

Figure 2.13 Characterization of 6mer bPNA Cbf-(β-A)-(EM*)6G. (a) HPLC trace on a

RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-TOF

Mass of Cbf-(β-A)-(EM*)6G. Calculated [M+H]: 2702.728, Found: 2702.209.

62 a

b

Figure 2.14 Characterization of 8mer bPNA Cbf-(β-A)-(EM*)8G. (a) HPLC trace on a

RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-TOF

Mass of Cbf-(β-A)-(EM*)8G. Calculated [M+H]: 3435.480, Found: 3436.586.

63 a

b

Figure 2.15 Characterization of 10mer bPNA Cbf-(β-A)-(EM*)10G. (a) HPLC trace on a

RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-TOF

Mass of Cbf-(β-A)-(EM*)10G. Calculated [M+H]: 4168.232 Found: 4167.691.

64 a

b

Figure 2.16 Characterization of 10mer bPNA Aba-(β-A)-(EM*)10G. (a) HPLC trace on a

RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-TOF

Mass of Aba-(β-A)-(EM*)10G. Calculated [M+H]: 3971.028 Found: 3970.596.

65

2.6 Experimental Methods

All measurements were carried out in triplicate using Dulbecco’s phosphate buffered saline (DPBS) at pH 7.4 without additional salt. Each peptide-DNA sample was incubated at 37 °C for one hour then cooled to 25 °C over one hour prior to measurement except where noted. Peptides and DNA concentrations were calculated from stock solutions using UV absorbance derived from melamine, ninhydrin test, ABA, Cbf, or nucleobases as appropriate.

Fluorescence quenching experiments were performed on a Perkin-Elmer luminescence spectrometer (LS-50B) equipped with a PTP-1 temperature programmer. Concentration of all peptides in fluorescence quenching titration experiments to determine binding stoichiometry was 1 μM. Fluorescence anisotropy (from Cbf−(EM*)n) based binding isotherms were measured on a Molecular Devices Spectramax M5. All binding isotherms were carried out at 25 nM peptide at 25 °C with the exception of Cbf-4 binding to dT4C4T4, which was carried out at 400 nM peptide at 8 °C on the LS-50B; samples were incubated in an ice bath for 1 h prior to measurement.

All measurements were carried out with a temperature change rate of 1 °C per minute.

Peptide −DNA complex melting was followed by UV absorbance changes on a Cary-100

UV-vis spectrophotometer equipped with an air-circulating temperature controller.

Peptide-DNA complex concentration was 5 μM for all UV samples. All DSC samples were measured on a Microcalorimeter VP-DSC at 25 μM peptide/DNA concentration with exceptions as noted. The average of at least three upscans is presented; the average of three downscans is also presented for selected complexes.

66

2.7 Acknowledgement

This work was supported by The Ohio State University.

2.8 References

(1) Felsenfeld, G.; Davies, D. R.; Rich, A. Formation of a three-stranded polynucleotide

molecule. J. Am. Chem. Soc. 1957, 79, 2023-2024.

(2) Chen, H.; Meena; McLaughlin, L. W. A Janus-Wedge DNA Triplex with A-W1-T G-

W2-C Base Triplets. J. Am. Chem. Soc. 2008, 130, 13190-13191.

(3) Soto, A. M.; Loo, J.; Marky, L. A. Energetic contributions for the formation of TAT/TAT,

TAT/CGC(+), CGC(+)/CGC(+) base triplet stacks. J. Am. Chem. Soc. 2002, 124,

14355-14363.

(4) Li, T. H.; Weinstein, D. S.; Nicolaou, K. C. The chemical end-ligation of

homopyrimidine oligodeoxyribonucleotides within a DNA triple helix, Chemistry &

Biology 1997, 4, 209-214.

(5) Li, H.; Franzini, R. M.; Bruner, C.; Kool, E. T. Templated Chemistry for Sequence-

Specific Fluorogenic Detection of Duplex DNA, Chembiochem, 2010, 11, 2132-2137.

(6) Zeng, Y.; Pratumyot, Y.; Piao, X.; Bong, D. Discrete assembly of synthetic peptide-

DNA triplex structures from polyvalent melamine-thymine bifacial recognition. J. Am.

Chem. Soc. 2012, 134, 832-835.

(7) Branda, N.; Kurz, G.; Lehn, J.-M. Janus wedges: a new approach towards

nucleobase-pair recognition. Chem. Commun. 1997, 1996, 2443-2444.

(8) Shin, D.; Tor, Y. Bifacial Nucleoside as a Surrogate for Both T A in Duplex DNA. J.

Am. Chem. Soc. 2011, 133, 6926-6929.

67

(9) Fenniri, H.; Packiarajan, M.; Vidale, K. L.; Sherman, D. M.; Hallenga, K.; Wood, K.

V.; Stowell, J. G. Helical rosette nanotubes: Design, self-assembly, characterization.

J. Am. Chem. Soc. 2001, 123, 3854-3855.

(10) Kool, E. T. Preorganization of DNA: Design Principles for Improving Nucleic Acid

Recognition by Synthetic Oligonucleotides. Chem. Rev. 1997, 97, 1473-1488.

(11) Nielsen, P. E.; Egholm, M.; Berg, R. H.; Buchardt, O. Sequence-selective recognition

of DNA by strand displacement with a thymine-substituted polyamide. Science 1991,

254, 1497-1500.

(12) Betts, L.; Josey, J. A.; Veal, J. M.; Jordan, S. R. A nucleic acid triple helix formed by

a peptide nucleic acid-DNA complex. Science 1995, 270, 1838-1841.

(13) Bentin, T.; Larsen, H. J.; Nielsen, P. E. Combined Triplex/Duplex Invasion of Double-

Stranded DNA by “Tail-Clamp” Peptide Nucleic Acid. Biochemistry 2003, 42, 13987-

13995.

(14) Kaihatsu, K.; Shah, R. H.; Zhao, X.; Corey, D. R. Extending Recognition by Peptide

Nucleic Acids (PNAs): Binding to Duplex DNA Inhibition of Transcription by Tail-

Clamp PNA-Peptide Conjugates. Biochemistry 2003, 42, 13996-14003.

(15) Zhang, W.; Nowlan, D. T.; Thomson, L. M.; Lackowski, W. M.; Simanek, E. E.

Orthogonal, Convergent Syntheses of Dendrimers Based on Melamine with One or

Two Unique Surface Sites for Manipulation. J. Am. Chem. Soc. 2001, 123, 8914-

8922.

(16) Blotny, G. Recent applications of 2, 4, 6-trichloro-1, 3, 5-triazine its derivatives in

organic synthesis. Tetrahedron 2006, 62, 9507-9522.

(17) Chen, H.-T.; Neerman, M. F.; Parrish, A. R.; Simanek, E. E. Cytotoxicity, Hemolysis,

Acute in Vivo Toxicity of Dendrimers Based on Melamine, Candidate Vehicles for

Drug Delivery. J. Am. Chem. Soc. 2004, 126, 10044-10048. 68

(18) Leman, L.; Orgel, L.; Ghadiri, M. R. Carbonyl sulfide-mediated prebiotic formation

of peptides. Science 2004, 306, 283-286.

(19) Li, X.; Liu, D. R. DNA-templated organic synthesis: nature’s strategy for controlling

chemical reactivity applied to synthetic molecules. Angew. Chem., Int. Ed. Engl.

2004, 43, 4848-4870.

(20) Nielsen, P. E. Peptide Nucleic Acids the Origin of Life. Chem. Biodiversity 2007, 4,

1996-2002.

(21) Ura, Y.; Beierle, J. M.; Leman, L. J.; Orgel, L. E.; Ghadiri, M. R. Self-assembling

sequence-adaptive peptide nucleic acids. Science 2009, 325, 73-77.

(22) Hysell, M.; Siegel, J. S.; Tor, Y. Synthesis stability of exocyclic triazine nucleosides.

Org. Biomol. Chem. 2005, 3, 2946-2952.

(23) Mittapalli, G. K.; Osornio, Y. M.; Guerrero, M. A.; Reddy, K. R.; Krishnamurthy, R.;

Eschenmoser, A. Mapping the landscape of potentially primordial informational

oligomers: oligodi-peptides tagged with 2,4-disubstituted 5-aminopyrimidines as

recognition elements. Angew. Chem., Int. Ed. 2007, 46, 2478-2484.

(24) Mittapalli, G. K.; Reddy, K. R.; Xiong, H.; Munoz, O.; Han, B.; De Riccardis, F.;

Krishnamurthy, R.; Eschenmoser, A. Mapping the landscape of potentially

primordial informational oligomers: oligodipeptides oligodipeptoids tagged with

triazines as recognition elements. Angew. Chem., Int. Ed. 2007, 46, 2470-2477.

(25) Ferris, J. P.; Sanchez, R. A.; Orgel, L. E. Studies in prebiotic synthesis: III. Synthesis

of from cyanoacetylene cyanate. J. Mol. Biol. 1968, 33, 693-704.

(26) Menor-Salvan, C.; Ruiz-Bermejo, D.; Guzman, M. I.; Osuna-Esteban, S.;

Veintemillas-Verdaguer, S. Synthesis of pyrimidines triazines in ice: implications for

the prebiotic chemistry of nucleobases. Chem. A-Eur. J., 2009, 15, 4411-4418.

69

(27) Whitesides, G. M.; Simanek, E. E.; Mathias, J. P.; Seto, C. T.; Chin, D.; Mammen,

M.; Gordon, D. M. Noncovalent Synthesis: Using Physical-Organic Chemistry To

Make Aggregates. Acc. Chem. Res. 1995, 28, 37-44.

(28) ten Cate, M. G. J.; Huskens, J.; Crego-Calama, M.; Reinhoudt, D. N. Thermodynamic

stability of hydrogen-bonded nanostructures: A calorimetric study. Chem. A-Eur. J.,

2004, 10, 3632-3639.

(29) Prins, L. J.; Reinhoudt, D. N.; Timmerman, P. Noncovalent synthesis using hydrogen

bonding. Angew. Chem., Int. Ed. 2001, 40, 2382-2426.

(30) Oshovsky, G. V.; Reinhoudt, D. N.; Verboom, W. Supramolecular chemistry in water.

Angew. Chem., Int. Ed. 2007, 46, 2366-2393.

(31) De Greef, T. F. A.; Smulders, M. M. J.; Wolffs, M.; Schenning, A.; Sijbesma, R. P.;

Meijer, E. W. Supramolecular Polymerization. Chem. Rev. 2009, 109, 5687-5754.

(32) Cafferty, B. J.; Gallego, I.; Chen, M. C.; Farley, K. I.; Eritja, R.; Hud, N. V. Efficient

Self-Assembly in Water of Long Noncovalent Polymers by Nucleobase Analogues.

J. Am. Chem. Soc. 2013, 135, 2447-2450.

(33) Ai, K.; Liu, Y.; Lu, L. Hydrogen-Bonding Recognition-Induced Color Change of Gold

Nanoparticles for Visual Detection of Melamine in Raw . J. Am.

Chem. Soc. 2009, 131, 9496-9497.

(34) Kawasaki, T.; Tokuhiro, M.; Kimizuka, N.; Kunitake, T. Hierarchical self-assembly of

chiral complementary hydrogen-bond networks in water: reconstitution of

supramolecular membranes. J. Am. Chem. Soc. 2001, 123, 6792-6800.

(35) Ariga, K.; Kunitake, T. Molecular Recognition at Air-Water Related Interfaces:

Complementary Hydrogen Bonding Multisite Interaction. Acc. Chem. Res. 1998, 31,

371-378.

70

(36) Marchi-Artzner, V.; Gulik-Krzywicki, T.; Guedeau-Boudeville, M.-A.; Gosse, C.;

Sanderson, J. M.; Dedieu, J.-C.; Lehn, J.-M. Selective adhesion, lipid exchange

membrane-fusion processes between vesicles of various sizes bearing

complementary molecular recognition groups. ChemPhysChem 2001, 2, 367-376.

(37) Zhou, Z.; Bong, D. Small-Molecule/Polymer Recognition Triggers Aqueous-Phase

Assembly Encapsulation. Langmuir 2013, 29, 144-150.

(38) Ma, M.; Bong, D. Directed peptide assembly at the lipid-water interface cooperatively

enhances membrane binding activity. Langmuir 2011, 27, 1480-1486.

(39) Ma, M.; Bong, D. Determinants of cyanuric acid melamine assembly in water.

Langmuir 2011, 27, 8841-8853.

(40) Ma, M.; Bong, D. Protein assembly directed by synthetic molecular recognition motifs.

Org. Biomol. Chem. 2011, 9, 7296-7299.

(41) Ma, M.; Paredes, A., Bong, D. Intra- intermembrane pairwise molecular recognition

between synthetic hydrogen-bonding phospholipids. J. Am. Chem. Soc. 2008, 130,

14456-14458.

(42) Jahromi, A. H.; Nguyen, L.; Fu, Y.; Miller, K. A.; Baranger, A. M.; Zimmerman, S. C.

A Novel CUG exp -MBNL1 Inhibitor with Therapeutic Potential for Myotonic

Dystrophy Type 1. ACS Chem. Biol. 2013, 8, 1037-1043.

(43) Arambula, J. F.; Ramisetty, S. R.; Baranger, A. M.; Zimmerman, S. C. A simple

ligand that selectively targets CUG trinucleotide repeats inhibits MBNL protein

binding. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 16068-16073.

(44) SantaLucia, J., Jr.; Hicks, D. The thermodynamics of DNA structural motifs. Annu.

Rev. Biophys. Biomol. Struct. 2004, 33, 415-440.

(45) Kool, E. T. Hydrogen bonding, base stacking steric effects in DNA replication. Annu.

Rev. Biophys. Biomol. Struct. 2003, 30, 1-22. 71

(46) Gilli, P.; Ferretti, V.; Gilli, G.; Borea, P. A. Enthalpy-entropy compensation in drug-

receptor binding. J. Phys. Chem. 1994, 98, 1515-1518.

(47) Dunitz, J. D. Win some, lose some: enthalpy-entropy compensation in weak

intermolecular interactions. Chem. Biol. 1995, 2, 709-712.

(48) Ghadiri, M. R.; Choi, C. Secondary structure nucleation in peptides. Transition metal

ion stabilized. alpha.-helices. J. Am. Chem. Soc. 1990, 112, 1630-1632.

(49) Ghadiri, M. R.; Fernholz, A. K. Peptide architecture. Design of stabled .alpha.-helical

metallopeptides via a novel exchange-inert ruthenium(III) complex. J. Am. Chem.

Soc. 1990, 112, 9633-9635.

(50) Cabezas, E.; Satterthwait, A. C. The hydrogen bond mimic approach: solid-phase

synthesis of a peptide stabilized as an α-helix with a hydrazone link. J. Am. Chem.

Soc. 1999, 121, 3862-3875.

(51) Ruan, F.; Chen, Y.; Itoh, K.; Sasaki, T.; Hopkins, P. B. Synthesis of peptides

containing unnatural, metal-ligating residues: aminodiacetic acid as a peptide side

chain. J. Org. Chem. 1991, 56, 4347-4354.

(52) Schafmeister, C. E.; Po, J.; Verdine, G. L. An All-Hydrocarbon Cross-Linking System

for Enhancing the Helicity Metabolic Stability of Peptides. J. Am. Chem. Soc. 2000,

122, 5891-5892.

(53) Chapman, R. N.; Dimartino, G.; Arora, P. S. A Highly Stable Short alpha-Helix

Constrained by a Main-Chain Hydrogen-Bond Surrogate. J. Am. Chem. Soc. 2004,

126, 12252-12253.

(54) Torres, O.; Yuksel, D.; Bernardina, M.; Kumar, K., Bong, D. Peptide tertiary structure

nucleation by side-chain crosslinking with metal complexation double “click”

cycloaddition. ChemBioChem 2008, 9, 1701-1705.

72

(55) Gellman, S. H. Minimal model systems for β –sheet secondary structure in proteins.

Curr. Opin. Chem. Biol. 1998, 2, 717-725.

(56) Mammen, M.; Choi, S.-K.; Whitesides, G. M. Polyvalent Interactions in Biological

Systems: Implications for Design Use of Multivalent Ligands Inhibitors. Angew.

Chem., Int. Ed. 1998, 37, 2754-2794.

(57) Rao, J.; Lahiri, J.; Isaacs, L.; Weis, R. M.; Whitesides, G. M. A trivalent system from

vancomycin.D-ala-D-Ala with higher affinity than avidin-biotin. Science 1998, 280,

708-711.

73

Chapter 3. Peptide Ligation and RNA Cleavage via an Abiotic Template Interface

74

3.1 Introduction

Nucleic acid recognition has been exploited to direct the chemistry of native and artificial1,2 macromolecules. Classic Watson-Crick base-pairing has been widely used to direct group transfer,3-5 peptide synthesis,6,7 nucleic acid ligation,8,9 drug selection,10,11 mutation or disease diagnosis,12 nucleic acids detection,13 ,14,15 self-replication16 and functional materials fabrication.17,18 In addition, base-pairing templates can code for synthesis of non-native scaffolds19-21 and multisite macromolecular modification.22-24

Hoogsteen base-pairing can also be involved in nucleic acid-templated chemistry in that a third strand of oligonucleotide interacts with a preformed duplex,25 or a circular DNA26 usually in non-physiological conditions.27,28 A lot of work has been done on exploring templated chemistry using native nucleobase-pairing system, including Watson-Crick and

Hoogsteen base-pairing of A, T/U, C and G. However, very few work has been reported on studying non-native base-pairing system. We have previously reported that melamine- displaying bifacial peptide nucleic acid (bPNA)29,30 docked two unstructured T/U tracts to form obligate triplex hybrid through thymine-melamine-thymine or uracil-melamine-uracil bifacial recognition. This abiotic base triple interface between bifacial peptide nucleic acid

(bPNA) and nucleic acids provides an opportunity to explore alternative template topologies. Although symmetric31-35 bifacial recognition36-39 of this type has no cognate in extant biology, amino and oxo 2,4,6-substituted triazines40,41 recapitulate the Watson-Crick hydrogen-bonding patterns, fueling the speculative notion of triazine-derived precursors42 to DNA and RNA. We have been seeking to establish a two-way communication between synthetic bPNA and native nucleic acid. On one hand, bPNA can assemble and fold nucleic acid, triggering DNA and RNA chemistry.43,44 On the other hand, we are curious what nucleic acid can do to bPNA upon hybridization. We report herein that hybridization can likewise trigger bPNA chemistry. Single-stranded and partially structured DNA/RNA 75

Figure 3.1 bPNA molecular recognition and nucleic acid templates. (Top) Molecular

structure of bifacial peptide nucleic acid (bPNA) and thymine-melamine-thymine or

uracil-melamine-uracil bifacial recognition. (Bottom) The four template topologies

shown catalyze thioester exchange, fragment ligation, chain extension, and oxidative

coupling with activation of ribozyme self-cleavage function with bPNA substrates (dark

lines).

topologies were found to serve as templates to catalyze bPNA coupling and length- controlled chain extension (oligomerization) of bi-reactive bPNAs. Furthermore, integration of a template site into a ribozyme fold renders RNA self-cleavage dependent on oxidative ligation of bPNA; this could serve as a blueprint for chemically sensitive nucleic acid switches45,46 and gates47-49 with applications in DNA/RNA nanotechnology.50,51

76

Overall, these data demonstrate readout and transformation of non-native macromolecules through an abiotic template interface in DNA/RNA template topologies that are not accessible via native base-pairing.

3.2 Results and Discussion

3.2.1 Acyl transfer catalyzed by bPNA-DNA hybridaztion

An n-mer of bPNA has the general form (EM*)n wherein M* is lysine sidechain modified with melamine (Figure 3.1). Triplex hybridization of bPNA with single-stranded DNA of the general form dTnC4Tn results in triplex stem loop (hairpin) structures with n thymine- melamine-thymine (TMT) base triples. Binary, ternary and quaternary bPNA-DNA

30 complexes can be formed with 1, 2 or 3 bPNAs bound to a single dTnC4Tn DNA strand.

For instance, dT15C4T15 binds 10mer and 4mer bPNA simultaneously to produce a heterotrimeric complex. Though 4mer hybridization is weak, 10mer bPNA preorganizes

DNA into a hairpin conformation, which facilitates binding of the lower affinity 4mer bPNA.

We hypothesized that dT15C4T15 could catalyze S→S acyl transfer between 10mer and

4mer bPNA thiols. A carboxyfluorescein-labeled 10mer bPNA thiol (10) and 4mer bPNA

(ABA) thioester (1) were prepared as substrates (Figure 3.2). Indeed, a new fluorescent

10mer was evident on denaturing PAGE, indicative of strongly template-dependent acyl transfer from 1 to 10 (Figure 3.2). Acyl transfer was undetectable without template under these conditions (200 nM bPNA). Band excision, extraction and mass spectroscopic analysis confirmed acyl transfer from 1 to 10. Regardless of increasing concentration of acyl donor, transfer yield ranged from 40-50%. This is consistent with the notion that all acyl transfer occurs on-template instead of off-template.

77

a b

Figure 3.2 DNA-templated thioester exchange between bPNAs. (a) Illustration of

hairpin folding of 10mer thiol (10, Cbf-β(EM*)10C) to dT15C4T15 followed by 4mer

thioester (1, ABA-G-Mpa-M*(EM*)3G) binding and transfer of R1 (ABA-G).

Cbf=carboxyfluorescein; ABA=4-acetamidobenzamide; Mpa=mercaptopropionamide;

β=β-alanine. Unless noted, bPNAs are C-terminal amides. (b) (Top) Representative

denaturing PAGE indicating acylated (<) and free thiol 10 (*) with increasing DNA

template. (Lower) Gel quantification. Error bars are from triplicate data.

3.2.2 Native chemical ligation templated by bPNA-DNA hybridization

Successful catalysis of S→S acyl transfer in a 10mer bPNA-preorganized hairpin template prompted investigation of more weakly organized templates using only reactive 4mer bPNAs. DNA hairpin templates were studied as catalysts for native chemical ligation52,53 of two bPNA 4mer fragments. Nominal background ligation of 4mer bPNAs 2

(fluorescently-labeled C-terminal thioester) and 3 (N-terminal cysteine) was observed by

8 M denaturing PAGE over several days. Remarkably, a strong template effect was observed with T10 hairpin DNA template, despite the modest binding affinity of 4mer bPNA (Figure 3.3). Increased concentration of DNA template dramatically enhanced ligation rate. Under optimal conditions (1000 nM), the template effect produced an 78 approximate 2500-fold rate enhancement over background. Consistent with the key role of the TMT interface, ligation yield dropped sharply with T→C substitutions in the DNA template. As expected, the reaction profile is consistent with product inhibition,54 further underscored by competitive inhibition with 10mer unreactive bPNA (Figure 3.3). A product-

DNA complex can also be detected as the reaction progresses by the melting temperature which shifted from 25 °C to 51.5 °C, the latter is similar to the thermal stability observed with an 8-mer bPNA-DNA complex. Notably, optimum reaction is observed at the Tm of the 4mer-DNA complex (25°C), with decreased reaction rate at higher (35°C) and lower

(4°C, 15°C) temperatures, suggestive of the importance of dynamic complexation. Higher temperature disfavored the formation of high-order complex whereas lower temperature froze the formed complex to inhibit orientation changes of each bPNA fragments. In addition, longer T15 and T18 hairpin DNA templates were found to yield significantly higher reaction rate than T10 hairpin DNA template, due to the fact that more 4mer bPNA fragments were held together at the same time.30 It is also likely that the longer templates that can accommodate more than two pairs of bPNA become preorganized following the first coupling, leading to enhanced binding and catalysis of subsequent fragments. To directly probe the effect of pre-structuring, duplex-organized DNA templates were tested in bPNA native chemical ligation. Indeed, ligation rates of bPNAs 2 and 3 increased dramatically as the T10 tracts were buttressed by one (T10-duplex) and two duplexes (T10-

(duplex)2). Duplex presentation of the T-tracts decreases the entropic cost of bPNA triplex hybridization, leading to increased template efficiency. Notably, the T-tract bulge presented by T10-(duplex)2 is a unique DNA template topology that takes advantage of the associative nature of bPNA targeting. Increased effective molarity55 of the fragments in the bPNA-DNA ternary complex can readily account for this rate acceleration.

79 a b

c d

e

Figure 3.3 DNA-templated native chemical ligation. (a) Illustration of DNA-templated native chemical ligation of bPNA fragments 2 and 3. (b) Exemplary gel image of

T10C4T10-templated native chemical ligation with product (<) and starting material (*) indicated. DNA concentration was 1000 nM. (c) Melting temperature shift. d) Ligation rate over time as a function of template concentration indicated (left) and 10mer (EM*)10 bPNA as competitive inhibitor, concentration indicated (right). e) Ligation rates increased with longer hairpin DNA template (left) and pre-organized templates (right) with the illustration of pre-organized templates shown on the side. Lines are drawn to guide the eye.

80

3.2.3 Length-controlled bPNA extension

DNA-templated native chemical ligation suggested the possibility of chain extension

(oligomerization) through multiple on-template couplings of bi-reactive bPNAs. This appeared reasonable as the thermal stability of bPNA-DNA complexes increases from binary to ternary to quaternary. This concept is similar to PNA chain extension20 and poly-

T formation19,56 on single-stranded deoxynucleotide based on Watson-Crick base-pairing via rapid reductive amination. Rapid cyclization57,58 of bPNAs bearing N-terminal cysteine and C-terminal thioester functionality prompted investigation of cysteine-free amide coupling of bPNAs. Though direct peptide aminoacylation with thioesters is low yielding reaction in aqueous milieu, amino acid side chains can greatly influence reaction rate and yield.59 Accordingly, a 4mer bPNA (4) fitted with N-terminal glycine and C-terminal histidine thioester was prepared for templated-chain extension. The reaction mixture was spiked with 5 mol% thioester 2 to fluorescently label the products for PAGE analysis. While background coupling was insignificant, ligation was observed on incubation with dTnC4Tn hairpin templates (n = 8, 10, 15, 18), with 25-50% overall conversion wherein higher yields corresponded to the longer templates. Furthermore, longer ligation products were observed with longer templates. For n = 8 and 10, dimer was the dominant outcome, consistent with the notion that two 4mer bPNAs could fit on the template at once. Longer and tetramer bPNA products from two and three on-template couplings were clearly detected as major products with the dT15C4T15 and dT18C4T18 templates, commensurate with higher-order complex formation. The identity of the oligomers was confirmed by band isolation and MALDI-MS. Though fluorescence labeling was used to image the gel, isolated bands yielded masses corresponding to the unlabeled, thioester hydrolyzed population. Thus, on-template, direct aminolysis of thioester fragments from ternary,

81 quaternary, and apparent quinternary bPNA-DNA complexes leads to bPNA chain extension by virtue of the length-matching abiotic TMT interface.

Figure 3.4 DNA-templated bPNA extension. (Top) Peptide chain extension of 4 with

dTnC4Tn DNA. (Lower) 8M urea denaturing PAGE of extension reactions with hairpin

template T-tract indicated. Product bands were identified as (a) hydrolyzed 4, (b) 4, (c)

dimer, (d) trimer, and (e) tetramer by MALDI-MS (c-e are C-terminal acids).

3.2.4 Ribozyme-templated bPNA oxidative coupling and activation of ribozyme self-cleavage function

In addition to acceleration of amide bond coupling, oxidation of bPNA dithiol 5 with T10-

(duplex)2 was also significantly accelerated over background. While thiol oxidation is more facile than amide bond formation under these conditions, the duplex-constrained template limited products formed to dimeric and trimeric extension; in contrast, a wide range of

82 oxidation products were formed off template. Unlike amide bond chain extension, three bPNA fragments may be oxidatively coupled on a T10 template; this is perhaps due to the increased flexibility of the disulfide linkage. Successful catalysis with partially folded DNA templates prompted investigation of Un-template loops imbedded within RNA folds. This notion was tested using a minimal type I hammerhead ribozyme60 in which stem III was

44 replaced with an rU10CACAU10 loop (U3-ribozyme). It was initially thought that the constrained U-loop would exhibit orientational bias with respect the RNA template; thus, two 4mer bPNAs bearing N-terminal (6) and C-terminal (7) thiols were prepared and studied. However, both thiols and their mixture, gave identical template-enhanced oxidation profiles with U3-ribozyme. Interestingly, the U3-ribozyme was able to achieve a similar yield of oxidation at lower catalyst loading (12%) compared to the DNA (50%), suggesting a higher exchange rate off the RNA loop template, though more quantitative measurements are needed to pinpoint the origin of this difference. The U3-ribozyme sequence has ablated self-cleavage activity due to the loss of stem III structure. We have previously demonstrated that duplex stems in aptamers and ribozyme folds can be structurally replaced with bPNA triplex hybrid stems when base-pairing sequences are replaced with T/U tracts. This allows bPNA to be used as an allosteric switch for both aptamer affinity and ribozyme catalysis. We hypothesized that native nucleic acid function could report on the coupling of short bPNA fragments. This notion was tested using oxidative thiol coupling since amide bond ligation occurs on a time scale similar to RNA degradation. While 4mer bPNA (EM*)4 and bPNAs 5, 6, and 7 only weakly activate cleavage of the U3-ribozyme, oxidative coupling produces an ~8mer bPNA disulfide product that binds more tightly to the template and strongly activates function. Oxidation conditions are more concentrated than those for ribozyme cleavage; thus, reactions were

83 started by dilution of partially oxidized samples into ribozyme cleavage conditions with 1 mM Mg2+. PAGE analysis of the reaction indicated the formation of two RNA products

a b

c

Figure 3.5 Nucleic acid-templated oxidative coupling of bPNA thiols. (a) Oxidation of

bPNA 5 alone (○) and with 50 mol% T10-(duplex)2 DNA (●), followed by 8 M Urea

denaturing PAGE. (b) Oxidation of bPNAs 6 and 7 alone (○) and with 12 mol% U3-

ribozyme (●), followed by Ellman’s test. (c) Illustration of U3-ribozyme cleavage upon

oxidation of bPNA thiols 6 and 7 and addition of Mg2+, and denaturing PAGE of RNA

cleavage triggered by 4mer bPNA (EM*)4 and 10% oxidized 6 and 7, with full-length

RNA (*), tRNA (<), and ribozyme (≪) cleavage products indicated, along with gel

quantification. Lines are drawn to guide the eye.

84 upon 10% bPNA oxidation, which were identified as the tRNA fusion and the hammerhead ribozyme components. Positive control experiments using 8mer bPNA (EM*)8 and fully oxidized and purified mixed disulfides of 6 and 7 gave higher yields of self-cleavage.

Ribozyme catalytic activity therefore may be used to report on fragment oxidative coupling, indicating two-way communication between an engineered abiotic template site and a native RNA self-cleavage site. This connection makes possible functional selection61,62 and optimization of template and redox-switchable .

3.3 Conclusion

These data collectively demonstrate that effective molarity increases on DNA and RNA templates can catalyze acyl transfer and oxidative coupling as well as chain extension of bPNA fragments. Insertion of template sites into folded nucleic acids is uniquely achieved through the thymine-melamine-thymine or uracil-melamine-uracil triplex interface with bPNA. Native nucleic acid function can thus be linked with engineered reactivity through allostery and template effects using the abiotic TMT interface. It is anticipated that facile inclusion of partially folded template topologies in nucleic acid directed chemistry will have use in DNA/RNA nanotechnology.

3.4 Materials and Equipments

Chemicals for amino acid derivative synthesis, solid phase peptide synthesis, purification and characterization were purchased from Sigma-Aldrich, Chem-Impex, VWR and

AAPPTec without further purification unless otherwise specified. Rink Resin LS (100-200 mesh, 0.28 mmol/g) and Fmoc-2-Cl-Trt Resin SS (100-200 mesh, 0.72 mmol/g) were purchased from Advanced ChemTech and Chempep. All DNA oligomers were obtained from Sigma-Aldrich. Dulbecco’s Phosphate-Buffered Salines (DPBS 1 ×, no calcium, no 85 magnesium) used as the binding buffer was purchased from Invitrogen. 40% acrylamide and bis-acrylamide solution (19:1) was purchased from Bio-Rad. TBE Buffer (10 ×) for gel electrophoresis was purchased from American Bioanalytical. MEGAshortscript T7 kit and

MEGAclear purification kit for RNA purification were purchased from Invitrogen. SYBR

Gold was purchased from Invitrogen.

The concentrations of all DNA oligomers and synthetic peptides were determined by UV absorbance using a UV-Vis HP 8543. MALDI-TOF Mass spectra were acquired on Bruker

Microflex MALDI-TOF instrument. Electrospray mass spectroscopy (ESI) was acquired on

Bruker MicroTOF equipped with an electrospray ionization source. Mass spectrometry instruments were provided by a grant from the Ohio BioProducts Innovation Center. NMR spectra were acquired on Bruker Advance DPX 400 instrument. Solid phase peptide synthesis was performed on AAPPTec Apex 396 peptide synthesizer. All peptides were purified on RP-C18 preparative HPLC column and the purity was confirmed on RP-C18 analytical column. UV melting was performed on Varian Cary-100 UV-Vis

Spectrophotometer equipped with Cary Temperature Controller. Fluorescence gel images were acquired using a Typhoon Trio Variable Mode Imager (Amersham Biosciences).

3.5 Sequences

bPNA Sequence

1 N-terminal 4M* peptide thioester, 1 ABA-G-COS-CH2CH2-M*(EM*)3G-CONH2

2 C-terminal 4M* peptide thioester, 2 Cbf-(β-A)-(EM*)4G-COS-CH2CH2SO3Na

3 N-Cysteine 4M* peptide, 3 H2N-C-M*-(EM*)3G-CONH2

86

4 C-terminal 4M* peptide thioester, 4 H2N-G-M*-(EM*)3H-COS-CH2CH2SO3Na

5 N,C-dicysteine 4M* peptide, 5 CH3CO-C-M*(EM*)3-C-CONH2

6 N-thiol (EM*)4 peptide, 6 HS-CH2CH2-(EM*)4G-CONH2

7 C-cysteine (EM*)4 peptide, 7 H2N-(EM*)4-C-CONH2

8 N-Cbf (EM*)4 Cysteine peptide Cbf-(β-A)-(EM*)4C-CONH2

9 DNA template inhibitor ABA-(β-A)-(EM*)10G-CONH2

10 N-Cbf (EM*)10 cysteine peptide, 10 Cbf-(β-A)-(EM*)10C-CONH2

DNA Sequence

T10-duplex 5'-ACTGACGACTGACGTTTTTTTTTT-3'

3'-TGACTGCTGACTGCTTTTTTTTTT-5'

T10-(duplex)2 5'-ACTGACGTTTTTTTTTTGCTGACG-3'

3'-TGACTGCTTTTTTTTTTCGACTGC-5'

T15-(duplex)2 5'-ACTGACGACTTTTTTTTTTTTTTTACTCGA-3'

3'-TGACTGCTGTTTTTTTTTTTTTTTTGAGCT-5'

T8-hairpin 5'-TTTTTTTTCCCCTTTTTTTT-3'

T10-hairpin 5'-TTTTTTTTTTCCCCTTTTTTTTTT-3'

T15-hairpin 5'-TTTTTTTTTTTTTTTCCCCTTTTTTTTTTTTTTT-3'

T18-hairpin 5'-TTTTTTTTTTTTTTTTTTCCCCTTTTTTTTTTTTTTTTTT-3'

87

T10 5'-TTTTTTTTTT-3'

Mismatched template 5'-TTTTTCTTTTCCCCTTTTCTTTTT-3'

Mismatched template 5'-TTTTCCTTTTCCCCTTTTCCTTTT-3'

Mismatched template 5'-TTCTCTCTCTCCCCTTTTCTTTTT-3'

Mismatched template 5'-TTCTTTCTTTCCCCTTTCTTTCTT-3'

3.6 Synthesis

Scheme 3.1 Synthesis of ABA-Gly thioester

H2N-Gly-OEt.HCl 1) NaOH CH3CONH COOH CH3CONH CONHCH2COOCH2CH3 EDCI, TEA 2) HCl ABA 1 SH

CH3CONH CONHCH2COOH CH3CONH CONHCH2COSCH2 EDCI, TEA 2 ABA-Gly thioester

3.6.1 ABA-Gly thioester

4-Acetamidobenzoic acid (1.79 g, 10 mmol, ABA) was reacted with H-Gly-OEt salt (1.54 g, 11 mmol) in the presence of EDCI (2.5 g, 13 mmol) and

TEA (4.33 ml, 31 mmol). This reaction was done in CH2Cl2 (35 ml), and the solution became cloudy in 12 hours, the white solid collected through filtration was confirmed to be

1 the product 1. H NMR (400 MHz, CD3OD) δ (ppm): 1.28 (t, J=7.2Hz, 3H), 2.15 (s, 3H),

4.09 (s, 2H), 4.21 (q, J=7.2 Hz, 2H), 7.67 (d, J=7.6 Hz, 2H), 7.82 (d, J=8.4 Hz, 2H).

Calculated [M+H]: 265.1190; Found [M+H]: 265.0936.

88

Product 1 (1.00 g, 3.78 mmol) was dissolved in 10 ml methanol followed by adding 10ml

1M NaOH solution. Ethyl protecting group was removed in 1 hour. Most of methanol was rotavaporated before neutralization with 1 M HCl. The resulting solution was stored at -20 oC overnight. A lot of white crystalline solid (Product 2) appeared, and was collected

1 through filtration. H NMR (400 MHz, CD3OD) δ (ppm): 2.15 (s, 3H), 4.08 (s, 2H), 7.67 (d,

J=8.4 Hz, 2H) 7.83 (d, J=8.8 Hz, 2H). Calculated [M+H]: 237.0877; Found [M+H]:

237.0658.

Product 2 (235 mg, 1 mmol) was reacted benzyl mercaptan (0.1 ml, 0.85 mmol) in the presence of EDCI (287 mg, 1.5 mmol) and TEA (0.2 ml, 1.43 mmol) in CH2Cl2 (10 ml).

After 12 hours, the reaction was washed with 1 M HCl three times. The organic layer was dried over anhydrous Na2SO4 and removed by rotavaporation. The product (ABA-Gly thioester) was purified by silica chromatography (MeOH: DCM=1 : 50). 1H NMR (400 MHz,

CD3OD) δ (ppm): 2.15 (s, 3H), 4.15 (s, 2H), 4.26 (s, 2H), 7.25 (m, 5H), 7.68 (d, 2H, J=8.0

13 Hz), 7.83 (d, 2H, J=8.0 Hz). C NMR (125 MHz, CD3OD) δ (ppm): 23.97, 33.47, 50.34,

104.79, 120.29, 128.27, 129.40, 129.57, 129.91, 139.00, 143.57, 198.97. Calculated

[M+H]: 343.1118; Found [M+H]: 343.0801.

89

7.832 7.811 7.685 7.663 4.831 4.217 4.199 4.089 3.309 2.147 1.297 1.279 1.261

90

8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 ppm

2.01 1.99 2.01 1.99 3.03 3.00 Figure 3.6 1H NMR of product 1

90

7.838 7.816 7.681 7.660 4.842 4.082 3.310 3.307 2.147

91

8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 ppm

1.98 1.99 1.97 3.00

Figure 3.7 1H NMR of Product 2

91

7.839 7.817 7.692 7.670 7.297 7.277 7.261 7.241 7.221 7.204 4.833 4.265 4.147 3.313 3.309 2.146

92

8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 1.5 1.0 0.5 ppm

2.00 2.00 4.98 1.98 1.98 3.02

Figure 3.8 1H NMR for ABA-Gly thioester

92

198.97 143.57 139.00 129.91 129.57 129.40 128.27 120.29 104.79 50.34 33.47 23.97

93

210 200 190 180 170 160 150 140 130 120 110 100 90 80 70 60 50 40 30 20 10 0 ppm

Figure 3.9 13C NMR of ABA-Gly thioester

93

3.6.2 bPNA Synthesis

N-terminal bPNA thioester (1) N-terminal peptide thioester 1, ABA-G-COS-

CH2CH2M*(EM*)3G was synthesized through a solid-phase thioester exchange. The precursor CH3COSCH2CH2-CONH-M*(EM*)3G was synthesized on Rink amide resin using stardard Fmoc chemistry. The S-acetyl protecting group was removed by 0.4 M

NaOH solution (DMF/Water, 1:1). The beads then were shaken with ABA-Gly thioester (5 equiv.) in the presence of TEA and PPh3 for 1.5 h twice. The N-terminal thioester was cleaved in 95% TFA and 5% water, precipitated by cold ether, followed by HPLC purification using a gradient of 10% solvent B (45% Acetonitrile, 45% isopropanol, 10% water, 0.01% TFA) and 90% solvent A (100% water and 0.1% TFA) to 100% solvent B over 60 minutes.

C-terminal bPNA thioesters (2) & (4) C-terminal peptide thioesters (2 and 4) were synthesized on Fmoc-2-Cl Trityl Resin following a reference.63 The solution of first Fmoc protected amino acid and TEA was shaken with the resin without coupling reagents followed by blocking unreacted sites with MeOH. The last amino acid has to be Boc protected or Cbf labeled. Fully protected peptides (C-terminus carboxylic acid) were obtained by treating resins with DCM/TFE/AcOH = 8:1:1 for 2 hours. Then, resins were filtered off and large excess hexane was added to the solution. All solvents were removed under reduced pressure. Dioxane was added to the resulting solids and was freeze-dried overnight. The protected peptides were reacted with 20 equiv. of sodium 2- sulfanylethanesulfonate (MESNa) in the presence with 20 equiv. of DIC and 5 mol %

DMAP. The reaction was monitored by Maldi-Tof Mass. DMF was removed by rotavaporator once no starting protected peptides were detected in 72 hours. All protecting groups were then removed by 95% TFA/ 5% water. Peptide thioesters were precipitated by cold diethyl ether and purified by HPLC using a gradient of 10% solvent B (45% 94

Acetonitrile, 45% isopropanol, 10% water, 0.01% TFA) and 90% solvent A (100% water and 0.1% TFA) to 100% solvent B over 60 minutes.

All other bPNAs Peptides were synthesized on Rink amide resin using standard Fmoc

SPPS, and were cleaved in a solution of TFA/water/m-Cresol/Thioanisole/EDT

(82.5/5/5/5/2.5), followed by HPLC purification using a gradient of 10% solvent B (45%

Acetonitrile, 45% isopropanol, 10% water, 0.01% TFA) and 90% solvent A (100% water and 0.1% TFA) to 100% solvent B over 60 minutes.

95 a

b [a.u.] Intens. 1250

1000 1715.603

750

500

250

857.592

0 1000 1500 2000 2500 3000 3500 4000 4500 m/z Figure 3.10 Characterization of bPNA 1. (a) HPLC trace of bPNA 1 ABA-G-

COSCH2CH2-M*(EM*)3G-CONH2 on a RP-C18 analytical column, monitored by a UV-

Vis detector at 238 nm. (b) MALDI-TOF Mass of 1: Calculated [M+H]: 1717.816, Found:

1715.603.

96 a

4000 Intens. [a.u.] Intens.

3000

b

2000

2096.341

1000

0 1600 1800 2000 2200 2400 2600 2800 m/z

Figure 3.11 Characterization of bPNA 2. (a) HPLC trace of 2 Cbf-(β-A)-(EM*)4G-

COSCH2CH2SO3Na on a RP-C18 analytical column using a gradient of 10-50% solvent

B over 50 minutes, monitored by a UV-Vis detector at 238 nm. (b) MALDI-TOF 2:

Calculated [M-Na+H]: 2094.130, Found: 2096.341.

97 a

x104

2.0 Intens. [a.u.] Intens.

b 1.5

1514.653

1.0

0.5

0.0 1000 1200 1400 1600 1800 2000 2200 2400 m/z

Figure 3.12 Characterization of bPNA 3. (a) HPLC trace of 3 H2N-CM*(EM*)3G-CONH2 on a RP-C18 analytical column using a gradient of 10-50% solvent B over 50 minutes, monitored by a UV-Vis detector at 238 nm. (b) MALDI-TOF Mass of 3: Calculated

[M+H]: 1514.622, Found: 1514.653.

98

a Intens. [a.u.] Intens.

3000 b

2500

2000 1673.568

1500

1000

500

0 1000 1200 1400 1600 1800 2000 2200 m/z

Figure 3.13 Characterization of bPNA 4. (a) HPLC trace of 4. H2N-GM*(EM*)3H-

COSCH2CH2SO3Na on a RP-C18 analytical, monitored by a UV-Vis detector at 238 nm.

(b) MALDI-TOF Mass of 4: Calculated [M-Na+H]: 1672.785, Found: 1673.568.

99 a

C(EM)8 0:F11 MS Raw Intens. [a.u.] Intens.

1250 b

1000

1602.908

750

500

250

0 500 1000 1500 2000 2500 3000 3500 4000 m/z

Figure 3.14 Characterization of bPNA 5. (a) HPLC trace of 5 Ac-CM*(EM*)3C-CONH2 on a RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-

TOF Mass of 5: Calculated [M+H]:1602.759, Found:1602.908.

100 a

disulfide 0:H12 MS Raw

1000 Intens. [a.u.] Intens.

800 b

600 1628.679

400

200

0 500 1000 1500 2000 2500 3000 3500 4000 m/z

Figure 3.15 Characterization of bPNA 6. (a) HPLC trace of 6 HS-CH2CH2-(EM*)4G-

CONH2 on a RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b)

MALDI-TOF Mass of 6: Calculated [M+H]: 1628.725, Found: 1628.679.

101 a

EM4C 0:G7 MS Raw Intens. [a.u.] Intens.

3000

b

2000

1586.856

1000

0 1000 1500 2000 2500 3000 3500 4000 4500 m/z

Figure 3.16 Characterization of bPNA 7. (a) HPLC trace of 7 H2N-(EM*)4C-CONH2 on a RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-

TOF Mass of 7: Calculated [M+H]:1586.693, Found:1586.856.

102 a

EM4C 0:F11 MS Raw Intens. [a.u.] Intens.

6000

b

4000

2017.235

2000

0 500 1000 1500 2000 2500 3000 m/z

Figure 3.17 Characterization of bPNA 8. (a) HPLC trace of Cbf-(β-A)-(EM*)4C-CONH2 on a RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b) MALDI-

TOF Mass Calculated [M+H]:2016.077, Found: 2017.235.

103 a

* Aba-EM10\0_C11\1\1SRef

6000[a.u.] Intens.

5000 b

4000

3000 3970.596

2000

1000

0 1000 1500 2000 2500 3000 3500 4000 4500 m/z

Figure 3.18 Characterization of bPNA 9. (a) HPLC trace of 10mer ABA-(β-A)-(EM*)10G-

CONH2 on a RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b)

MALDI-TOF Mass of Aba-(β-A)-(EM*)10G-CONH2. Calculated [M+H]: 3971.028, Found:

3970.596.

104 a

201406013 0:G2 MS Raw Intens. [a.u.] Intens. 300

250

200 b

150 4213.571

100

50 2105.422

0 1000 1500 2000 2500 3000 3500 4000 4500 m/z

Figure 3.19 Characterization of bPNA 10. (a) HPLC trace of 10 Cbf-(β-A)-(EM*)10C-

CONH2 on a RP-C18 analytical column, monitored by a UV-Vis detector at 238 nm. (b)

MALDI-TOF Mass of 10: Cbf-(β-A)-(EM*)10C-CONH2. Calculated [M+H]: 4214.352,

Found: 4213.571.

105

3.7 Experimental Methods

3.7.1 DNA Complexation

1 × DPBS (No divalent cations, pH 7.4) was used as the binding buffer. Preformed double- stranded DNA templates with T-T mismatches were annealed at 95 C for 5 minutes and cooled slowly to room temperature overnight. Single-stranded DNA templates were used without annealing.

3.7.2 UV melting

2 µM C-terminal Cbf thioester 2, 2 µM N-Cysteine peptide 3 and 2 µM T10C4T10 were incubated at 25 C for 48 h. Control: 4 µM C-terminal Cbf thioester 2, 2 µM T10C4T10 were incubated at 25 C for 48 h. 1 × DPBS contained 1 mM MESNa and 1.2 mM TCEP to maintain a reducing environment. Absorbance at 260 nm was monitored from 4 C to 90

C. The temperature increase rate was 1 C / minute.

3.7.3 Electrophoretic mobility shift assay (EMSA)

General: 20% acrylamide, 8 M Urea and 0.5 × TBE were used under 220 V to separate different products. 20 µl samples were mixed with 10 µl loading buffer (95% formamide,

0.025% bromophenol blue, 0.025% xylene cyanol, 4.95% 0.25 M EDTA solution) and heated at 95 C for 5 minutes and cooled on ice prior to loading. 10 µl of each sample was loaded. Gel running was stopped once the front dye reached the bottom. Gel was scanned under green fluorescence channel (Excitation: 488 nm, Emission: 526 nM).

106

3.7.4 Acyl small-molecule transfer

The concentration of Cbf labeled (EM*)10 with a free thiol at C-terminal was 200 nM. The concentrations of N-terminal ABA thioester 1 (0 nM to 4000 nM) and DNA template (0 nM to 1600 nM) were varied. Cbf-(β-A)-(EM*)10C-CONH2 (10) was incubated with T15C4T15 for

30 minutes at 35 C followed by adding N-terminal ABA thioester 1. The resulting solution was equilibrated for 4 hours. 1.2 mM TCEP (No MESNa) was included in 1 × DPBS buffer to maintain a reducing environment.

3.7.5 Native chemical ligation

200 nM C-terminal Cbf thioester 2, 200 nM N-cysteine peptide 3 and DNA template were mixed at desired reaction temperatures. 1 × DPBS contained 1 mM MESNa and 1.2 mM

TCEP to maintain a reducing environment. Variables: template concentration (0 nM to

1000 nM), temperature (25 C, 15 C and 35 C), inhibitor concentration (0 nM to 1000 nM), hairpin template length (T10, T15 and T18) and pre-structured T10 template (T10,

T10C4T10, one-end duplex T10 and two-end duplex T10) and mismatched templates. 20 µl samples were fast frozen by dry ice at desired time points and kept at -20 oC before EMSA.

3.7.6 Cysteine-free direct aminolysis

5% (200 nM) Cbf-thioester 2, 95% (3.8 µM) unlabeled C-terminal 4M* peptide thioester 4 and 2 µM DNA template were mixed at 35 C for 96 hours. 2 µM of different DNA templates

(T8C4T8, T10C4T10, T15C4T15 and T18C4T18) were used. 20 µl samples were fast frozen by dry ice at desired time points and kept at -20 C before EMSA.

107

3.7.7 Disulfide ligation

90% (3.6 µM) of N,C-dicysteine bPNA 5 [CH3CO-CM*(EM*)3C-CONH2] and 10% (400 nM) of Cbf-(β-A)-(EM*)4C-CONH2 were used. The amount of DNA template was 2 µM. The ligation was carried out at room temperature in 24 hours. 20 µl samples were fast frozen by dry ice at desired time points and kept at -20 C before EMSA.

3.7.8 Ellman’s test for dicysteine bPNA extension

Background oxidation sample was prepared with 40 μM dicysteine peptide 5 in 1 x DPBS buffer. DNA-templated oxidation sample was prepared with 40 μM dicysteine peptide 5 and 20 μM two-end T10 duplex DNA template. Ellman’s testing solution was made of 0.1 mM DTNB (5,5'-dithiobis-(2-nitrobenzoic acid), Ellman’s reagent), 2.5 mM sodium acetate and 8 M urea in 0.1 M Tris buffer (pH=8.0). 40 μl of sample was mixed with 40 μl of

Ellman’s testing solution and then subjected to UV-Vis measurement. Optical absorbance at 412 nm was measure at 0h, 3h, 6h, 9h, 12h, 18h and 24h.

3.7.9 Product isolation from gel

The masses of the ligation products were confirmed by excising the band and extracting it with a mixture of 50% water, 50% acetonitrile and 0.1% TFA (10 µl) for 4 hours. 20 µl of

10 µM reaction was used in this part. 1 µl of the extracted solution was used for Maldi-Tof

Mass analysis.

3.7.10 Gel quantification

Gel images were quantified using ImageQuant 5.0. A unit background intensity was obtained by averaging unit intensities of five randomly-picked empty spots. The intensities of product bands and reactant bands were obtained by subtracting background intensities 108 from overall intensities of selected spots. Percentage yield was the ratio of product band intensity to the sum of product band intensity, byproduct intensity and reactant intensity on the same lane.

3.7.11 RNA transcription and purification

DNA oligonucleotides containing the desired RNA sequences were constructed to imbed

T7 promoter sequence at 3’ end. Transcription assays were performed using

MEGAshortscript T7 kit, and sequentially purified using MEGAclear purification kit. The purity of transcribed RNAs was analyzed through denaturing electrophoresis.

3.7.12 Ribozyme cleavage with bPNA

500 nM U-Ribozyme constructs were mixed with 4 μM of bPNA in 1 x Tris-Cl (pH 7.4, 50 mM) buffer. Cleavage reaction was initiated by adding 1 mM Mg2+ for U3 cleavage or

10mM Mg2+ for U23. Aliquots of reaction mixtures were taken out at 15, 30, 60, 120, 180 minutes and quenched with 500 mM EDTA. Cleavage products were analyzed through

14% denaturing PAGE (8 M urea, 1 x TBE) under 350 V. Gels were stained with SYBR®

Gold. Data was quantified using Equation 1.

kt Equation 1: [product]  [U3ribozyme]*(1 exp )

3.7.13 U3 Ribozyme-templated bPNA disulfide oxidation

25 μM U-ribozyme constructs were mixed with 200 μM reduced 4mer bPNA (6 and 7) in

1 x Tris-Cl (pH 7.4, 50 mM), the mixture was incubated at 37 °C. Oxidation progress was monitored and quantified through Ellman’s Test. Control experiment with only 200 μM

4mer bPNA (6 and 7) was performed simultaneously and quantified using the same method. 109

3.7.14 In situ bPNA oxidation and U3 Ribozyme cleavage

4mer bPNA (6 and 7) oxidation in the presence of U3 ribozyme was performed as described above. When oxidation percentage reached 10%, 30% and 50%, portions of oxidation mixture was diluted 50-fold and cleavage reaction was initiated with the addition of 1 mM Mg2+, aliquots of reaction mixtures were taken out at 15, 30, 60, 120, 180 minutes and quenched with 500 mM EDTA. Cleavage products were analyzed through denaturing electrophoresis and gels were stained with SYBR Gold. Data was quantified using

Equation 1.

3.8 Acknowledgement

Research was supported in part by NSF-DMR. Facilities were provided by The Ohio State

University.

3.9 References

(1) Schmidt, J. G.; Christensen, L.; Nielsen, P. E.; Orgel, L. E. Information transfer from

DNA to peptide nucleic acids by template-directed syntheses, Nucleic Acids Res.

1997, 25, 4792-4796.

(2) Goodwin, J. T.; Mehta, A. K.; Lynn, D. G. Digital and analog chemical evolution, Acc.

Chem. Res. 2012, 45, 2189-2199.

(3) Grossmann, T. N.; Seitz, O. DNA-catalyzed transfer of a reporter group, J. Am. Chem.

Soc. 2006, 128, 15596-15597

(4) McKee, M. L.; Evans, A. C.; Gerrard, S. R.; O’Reilly, R. K.; Turberfield, A. J.; Stulz,

E. Peptidomimetic bond formation by DNA-templated acyl transfer, Org. Biomol.

Chem. 2011, 9, 1661-1666.

110

(5) McKee, M. L.; Milnes, P. J.; Bath, J.; Stulz, E.; Turberfield, A. J.; O’Reilly, R. K.

Multistep DNA-templated reactions for the synthesis of functional sequence

controlled oligomers, Angew. Chem., Int. Ed. 2010, 49, 7948-7951

(6) Erben, A.; Grossmann, T. N.; Seitz, O. DNA-triggered synthesis and bioactivity of

proapoptotic peptides, Angew. Chem. Int. Ed. 2011, 50, 2828-2832.

(7) He, Y.; Liu, D. R. A sequential strand-displacement strategy enables efficient six-

step DNA-templated synthesis, J. Am. Chem. Soc. 2011, 133, 9972-9975.

(8) Gao, K.; Orgel, L. E. Polyphosphorylation and Non-Enzymatic Template-Directed

Ligation of Oligonucleotides, Origins of Life and Evolution of the Biosphere 2000,

30, 45-51.

(9) Schmidt, J. G.; Nielsen, P. E.; Orgel, L. E. Information transfer from peptide nucleic

acids to RNA by template-directed syntheses, Nucleic Acids Res. 1997, 25, 4797-

4802.

(10) Gartner, Z. J.; Tse, B. N.; Grubina, R.; Doyon, J. B.; Snyder, T. M.; Liu, D. R. DNA-

templated organic synthesis and selection of a library of macrocycles, Science 2004,

305, 1601-1605.

(11) Tse, B. N.; Snyder, T. M.; Shen, Y. H.; Liu, D. R. Translation of DNA into a library of

13 000 synthetic small-molecule macrocycles suitable for in vitro selection, J. Am.

Chem. Soc. 2008, 130, 15611-15626.

(12) Xu, Y. Z.; Karalkar, N. B.; Kool, E. T. Nonenzymatic autoligation in direct three-color

detection of RNA and DNA point mutations, Nat. Biotechnol. 2001, 19, 148-152.

(13) Silverman, A. P.; Kool, E. T. Detecting RNA and DNA with templated chemical

reactions, Chem. Rev. 2006, 106, 3775-3789.

(14) Chandrasekar, J.; Silverman, S. K. Catalytic DNA with phosphatase activity, Proc.

Natl. Acad. Sci., U.S.A. 2013, 110, 5315-5320. 111

(15) Parker, D. J.; Xiao, Y.; Aguilar, J. M.; Silverman, S. K. DNA catalysis of a normally

disfavored RNA hydrolysis reaction, J. Am. Chem. Soc. 2013, 135, 8472-8475.

(16) Von Kiedrowski, G. A self-replicating hexadeoxynucleotide, Angew. Chem., Int. Ed.

1986, 25, 932-935.

(17) Schreiber, R.; Kempter, S.; Holler, S.; Schuller, V.; Schiffels, D.; Simmel, S. S.;

Nickels, P. C.; Liedl, T. DNA origami-templated growth of arbitrarily shaped metal

nanoparticles, Small, 2011, 7, 1795-1799.

(18) Hopkins, D. S.; Pekker, D.; Goldbart, P. M.; Bezryadin, A. Quantum interference

device made by DNA templating of superconducting nanowires, Science 2005, 308,

1762-1765.

(19) Li, Z. Y.; Zhang, Z. Y. J.; Knipe, R.; Lynn, D. G. DNA-catalyzed polymerization, J.

Am. Chem. Soc. 2002, 124, 746-747.

(20) Rosenbaum, D. M.; Liu, D. R. Efficient and sequence-specific DNA-templated

polymerization of peptide nucleic acid aldehydes, J. Am. Chem. Soc. 2003, 125,

13924-13925.

(21) Lo, P. K.; Sleiman, H. F. Nucleobase-templated polymerization: copying the chain

length and polydispersity of living polymers into conjugated polymers, J. Am. Chem.

Soc. 2009, 131, 4182-4183.

(22) Ura, Y.; Beierle, J. M.; Leman, L. J.; Orgel, L. E.; Ghadiri, M. R. Self-assembling

sequence-adaptive peptide nucleic acids, Science 2009, 325, 73-77.

(23) Heemstra, J. M.; Liu, D. R. Templated synthesis of peptide nucleic acids via

sequence-selective base-filling reactions, J. Am. Chem. Soc. 2009, 131, 11347-

11349.

112

(24) McHale, R.; O’Reilly, R. K. Nucleobase containing synthetic polymers: advancing

biomimicry via controlled synthesis and self-assembly, Macromolecules 2012, 45,

7665-7675.

(25) Lee, S. H.; Wang, S.; Kool, E. T. Templated chemistry for monitoring damage and

repair directly in duplex DNA, Chem. Comm. 2012, 48, 8069-8071.

(26) Selvasekaran, J.; Turnbull, K. D. Chemical ligation of oligodeoxyribonucleotides on

circular DNA templates, Nucleic Acids Res. 1999, 27, 624-627.

(27) Li, T. H.; Weinstein, D. S.; Nicolaou, K. C. The chemical end-ligation of

homopyrimidine oligodeoxyribonucleotides within a DNA triple helix, Chemistry &

Biology 1997, 4, 209-214.

(28) Li, H.; Franzini, R. M.; Bruner, C.; Kool, E. T. Templated Chemistry for Sequence-

Specific Fluorogenic Detection of Duplex DNA, Chembiochem, 2010, 11, 2132-2137.

(29) Zeng, Y.; Pratumyot, Y.; Piao, X.; Bong, D. Discrete assembly of synthetic peptide-

DNA triplex structures from polyvalent melamine-thymine bifacial recognition, J. Am.

Chem. Soc. 2012, 134, 832-835

(30) Piao, X.; Xia, X.; Bong, D. Bifacial peptide nucleic acid directs cooperative folding

and assembly of binary, ternary, and quaternary DNA complexes, Biochemistry

2013, 52, 6313-6323.

(31) Felsenfeld, G.; Davies, D. R.; Rich, A. Formation of a three-stranded polynucleotide

molecule, J. Am. Chem. Soc. 1957, 79, 2023-2024.

(32) Pilch, D. S.; Levenson, C.; Shafer, R. H. Structural analysis of the (dA) 10.2 (dT) 10

triple helix, Proc. Natl. Acad. Sci. U.S.A. 1990, 87, 1942-1946.

(33) Betts, L.; Josey, J. A.; Veal, J. M.; Jordan, S. R. A nucleic acid triple helix formed by

a peptide nucleic acid-DNA complex, Science 1995, 270, 1838-1841.

113

(34) Soto, A. M.; Loo, J.; Marky, L. A. Energetic contributions for the formation of TAT/TAT,

TAT/CGC+, and CGC+/CGC+ base triplet stacks, J. Am. Chem. Soc. 2002, 124,

14355-14363

(35) Duca, M.; Vekhoff, P.; Oussedik, K.; Halby, L.; Arimondo, P. B. The triple helix: 50

years later, the outcome, Nucleic Acids Res. 2008, 36, 5123-5138.

(36) Shin, D.; Tor, Y. Bifacial nucleoside as a surrogate for both T and A in duplex DNA,

J. Am. Chem. Soc. 2011, 133, 6926-6929.

(37) Branda, N.; Kurz, G.; Lehn, J.-M. JANUS WEDGES: a new approach towards

nucleobase-pair recognition, Chem. Commun. 1996, 2443-2444.

(38) Chen, H.; Meena; McLaughlin, L. W. A Janus-Wedge DNA triplex with A-W1-T and

G-W2-C base triplets, J. Am. Chem. Soc. 2008, 130, 13190-13191.

(39) Largy, E.; Liu, W.; Hasan, A.; Perrin, D. M. Base-pairing behavior of a carbocyclic

Janus-AT nucleoside analogue capable of recognizing A and T within a DNA duplex,

ChemBioChem 2013, 14, 2199 − 2208.

(40) Mittapalli, G. K.; Reddy, K. R.; Xiong, H.; Munoz, O.; Han, B.; De Riccardis, F.;

Krishnamurthy, R.; Eschenmoser, A. Mapping the landscape of potentially

primordial informational oligomers: oligodipeptides and oligodipeptoids tagged with

triazines as recognition elements, Angew. Chem., Int. Ed. 2007, 46, 2470-2477.

(41) Hysell, M.; Siegel, J. S.; Tor, Y. Synthesis and stability of exocyclic triazine

nucleosides, Org. Biomol. Chem. 2005, 3, 2946-2952.

(42) Menor-Salvan, C.; Ruiz-Bermejo, D.; Guzman, M. I.; Osuna-Esteban, S.;

Veintemillas-Verdaguer, S. Synthesis of pyrimidines and triazines in ice:

implications for the prebiotic chemistry of nucleobases, Chem.- A Eur. J. 2009, 15,

4411-4418.

114

(43) Xia, X.; Piao, X.; Fredrick, K.; Bong, D. Bifacial PNA complexation inhibits enzymatic

access to DNA and RNA, ChemBioChem 2013, 15, 31-36.

(44) Xia, X.; Piao, X.; Bong, D. Bifacial PNA as an allosteric switch for aptamer and

ribozyme function, J. Am. Chem. Soc. 2014, 76, 7265-7268.

(45) Liedl, T.; Olapinski, M.; Simmel, F. C. A surface-bound DNA switch driven by a

chemical oscillator, Angew. Chem., Int. Ed. 2006, 45, 5007-5010.

(46) Krishnan, Y.; Simmel, F. C. Nucleic acid based molecular devices, Angew. Chem.,

Int. Ed. 2011, 50, 3124-3156.

(47) Chen, X.; Wang, Y.; Liu, Q.; Zhang, Z.; Fan, C.; He, L. Construction of molecular

logic gates with a DNA-cleaving , Angew. Chem. Int. Ed. 2006, 118,

1791-1794.

(48) Yoshida, W.; Yokobayashi, Y. Photonic boolean logic gates based on DNA aptamers,

Chem. Commun. 2007, 195-197.

(49) Saghatelian, A.; Volcker, N. H.; Guckian, K. M.; Lin, V. S.-Y.; Ghadiri, M. R. DNA-

based photonic logic gates: AND, NAND, and INHIBIT, J. Am. Chem. Soc. 2003,

125, 346-347.

(50) Guo, P. The emerging field of RNA nanotechnology, Nat. Nano. 2010, 5, 833-842.

(51) Chakraborty, S.; Mehtab, S.; Krishnan, Y. The predictive power of synthetic nucleic

acid technologies in RNA biology, Acc. Chem. Res. 2014, 47, 1710-1719.

(52) Dawson, P. E.; Muir, T. W.; Clarklewis, I.; Kent, S. B. H. Synthesis of proteins by

native chemical ligation, Science 1994, 266, 776-779.

(53) Roloff, A.; Seitz, O. Bioorthogonal reactions challenged: DNA templated native

chemical ligation during PCR, Chem. Sci. 2013, 4, 432-436.

(54) Dose, C.; Ficht, S.; Seitz, O. Reducing product inhibition in DNA-template-controlled

ligation reactions, Angew. Chem., Int. Ed. 2006, 45, 5369-5373. 115

(55) Li, X.; Liu, D. R. DNA-templated organic synthesis: nature's strategy for controlling

chemical reactivity applied to synthetic molecules, Angew. Chem., Int. Ed. 2004, 43,

4848-4870.

(56) Zhan, Z. Y. J.; Lynn, D. G. Chemical amplification through template-directed

synthesis, J. Am. Chem. Soc. 1997, 119, 12420-12421.

(57) Li, Y.; Yongye, A.; Giulianotti, M.; Martinez-Mayorga, K.; Yu, Y.; Houghten, R. A.

Synthesis of cyclic peptides through direct aminolysis of peptide thioesters

catalyzed by imidazole in aqueous organic solutions, J. Comb. Chem. 2009, 11,

1066-1072.

(58) White, C. J.; Yudin, A. K. Contemporary strategies for peptide macrocyclization, Nat.

Chem. 2011, 3, 509-524.

(59) Payne, R. J.; Ficht, S.; Greenberg, W. A.; Wong, C.-H. Cysteine-free peptide and

glycopeptide ligation by direct aminolysis, Angew. Chem., Int. Ed. 2008, 120, 4483-

4487

(60) Murray, J. B.; Terwey, D. P.; Maloney, L.; Karpeisky, A.; Usman, N.; Beigelman, L.;

Scott, W. G. The structural basis of hammerhead ribozyme self-cleavage, Cell 1998,

92, 665-673.

(61) Tuerk, C.; Gold, L. Systematic evolution of ligands by exponential enrichment: RNA

ligands to bacteriophage T4 DNA polymerase, Science 1990, 249, 505-510.

(62) Ellington, A. D.; Szostak, J. W. In vitro selection of RNA molecules that bind specific

ligands, Nature 1990, 346, 818-822.

(63) Von Eggelkraut-Gottanka, R.; Klose, A.; Beck-Sickinger, A. G.; Beyermann, M.

Peptide αthioester formation using standard Fmoc-chemistry, Tetrahedron Letters

2003, 44, 3551-3554.

116

Chapter 4. In vitro Selection of a bPNA-Assisted Reversibly-Structured RNA

Library against Stapled HIV-1 Rev Peptides

117

4.1 Introduction

4.1.1 In vitro selection and aptamers

In 1990, Szostak,1 Gold2 and Joyce3 independently developed a similar in vitro selection procedure that allowed simultaneous screening of a highly diverse library of nucleic acid against varied targets for different functionalities. This procedure is also known as in vitro evolution or Systematic Evolution of Ligands by EXponential enrichment (SELEX).

Generally, a library of up to 1015 individual oligonucleotides are incubated with a certain target, such as protein, at a relatively mild condition (high concentration, low temperature, long incubation time). This library is then selected for a certain functionality, such as binding, reactivity et al. The survived oligonucleotides are pulled down from the solution and amplified for the next round selection in which the selection stringency is increased.

In this way, the starting library of oligonucleotides for each next round is enriched compared to that of last round. Gradually, the selection condition becomes harsh (low concentration, high temperature, short incubation time), and 6-20 rounds of selection are commonly required though a significant enrichment may be observed around the 6th or 7th round analyzed by high-throughput sequencing.4,5 Finally, the enriched library is sequenced and the selected oligonucleotides are subject to post-SELEX analysis. The evolved oligonucleotide is widely known as aptamer. The binding of aptamer to its target is based on a comprehensive effect of shape complementarity, base stacking, electrostatic interaction and hydrogen bonding.6,7 In vitro selection is a useful technology for biotechnological and pharmaceutical research to discover new aptamers with desired functionalities. Theoretically, in vitro selection can be used to yield aptamer against any molecules with high affinity and specificity, although targets with negative charges or hydrophobic properties are known to be more difficult. However, difficult targets may not

118

Figure 4.1 General procedure of in vitro selection (SELEX). Cited from reference 6.

be a problem with the development of unnatural hydrophobic or cationic nucleobases.6 A reported hydrophobic nucleobase-contained aptamer for VEGF (vascular endothelial grow factor) showed high binding affinity at picomolar.8 Furthermore, in vitro selection is powerful to differentiate chiral molecules9-11 and recognize a different epitope of a target molecule.12 Hundreds of publications have reported in vitro selections against a variety of targets, including cell,13,14 protein,15-18 peptide,19,20 nucleic acid,21 small molecule22-27 and even metal ion.28 Among different types of targets, protein is the most popular and important as selected aptamer usually shows nanomolar affinity and has the potential to block protein active sites, serving as an inhibitor for disease. The first and so far only aptamer therapeutic product approved by FDA is Pegaptanib which is a modified anti- human VEGF aptamer and used as the medicinal active component for the treatment of

119 neovascular age-related macular degeneration.29 Other applications of aptamers include aptamer-based affinity chromatography, , bioimaging and drug delivery.30-32

4.1.2 Two major limitations in in vitro selection

Regardless of the success of in vitro selection and wide applications of aptamers, the selection procedure is limited by several factors. First of all, the size of library is actually limited by chemical synthesis of completely randomized oligonucleotides. Oligonucleotide synthesizer is able to synthesize up to 1015 different DNA molecules, which covers about

25 random positions (425=1.125 × 1015). This number is significant if one considers that a mouse can possibly produce 109-1011 different antibodies. However, many reported in vitro selections use much more random positions than 25. The chemical synthesis limitation and low coverage (40 random position will theoretically need 440 = 2 mole of oligonucleotides to cover one copy of the library) of the initial library input may result in a dramatic loss of potential aptamers.33 Joyce and co-workers developed a strategy to create more randomized positions by adding Mn2+ and changing the concentration of deoxynucleotide triphosphate during PCR.34,35 In this way, Taq polymerase was forced to create mutations to cover more random positions. However, this strategy would not apply to the more broadly binding affinity selection for the possible reasons below: 1) it would have the potential to cause mutations on constant flanking primer regions to cause extra random positions; 2) mutations on constant primer binding region would lower the PCR efficiency and a number of oligonucleotides would not be amplified during PCR. Second, especially for RNA in vitro selection, the reverse transcription step to convert selected

RNA into first-strand cDNA for PCR amplification is known to be problematic when RNA shows extensive secondary and tertiary structures.33,36,37 However, these highly-structured molecules are usually of interest. 120

4.1.3 A novel strategy simultaneously overcomes the two limitations

Figure 4.2 Illustration of bPNA-assisted reversibly-structured RNA library construction.

Melamine heterocycles on bPNA bifacially recognize uracil nucleobases, resulting in a

triplex stem to pre-structre the RNA library.

We previously reported that synthetic bPNA formed triplex stem-loop structure38,39 with thymine-rich DNA and uracil-rich RNA and this triplex stem can structurally replace native duplex stem while the function of nucleic acid remained intact.40 We herein hypothesize that bPNA can be used to construct a reversibly-structured RNA library for in vitro selection which would overcome the two major problems mentioned above. First, many reported

121 aptamers contain duplex stems evolved from the randomized region but do not engage in binding.14,41,42 The binding motifs often come from bulge and loop. Indeed, researchers can truncate, replace or modify the duplex stem while the binding affinity is not affected.

Our bPNA would just trigger the pre-formation of supporting stem, avoiding the “waste” of precious random region and rendering the upper limit 25 random nucleotide positions to evolve for binding to target. By pre-installing the supporting stem, the random region is minimized to cover full library by several copies easily. Second, bPNA is abiotic and not part of the library and can be readily removed before reverse transcription, resulting in a loosely-structured RNA library. Consequently, reverse transcription may have a better chance to succeed.

4.1.4 HIV-1 Rev peptide and it use in in vitro selection

HIV-1 Rev is an important RNA-binding protein that transports gag-pal and env mRNAs from the nucleus to the cytoplasm of infected cells.43-46 A helical 17-amino acid arginine- rich sequence on Rev interacts with Rev responsive element (RRE) RNA located in an intron downstream of env gene. On one hand, researchers demonstrated this short 17- amino acid peptide bound to RRE with specificity similar to the intact Rev protein. On the other hand, biochemical and structural analyses identified a 66-nucleotide stem (domain

II) that was responsible for the high-affinity Rev-binding on RRE47 and demonstrated 17- amino acid Rev peptide still bound to 66-nucleotide RRE domain II with specificity. 48-51

Out of the interest in HIV-1 Rev protein and Rev peptide, in vitro selections against them were performed previously. Szostak and co-workers constructed a RNA library containing

66 random position, mimicking the domain II and selected against Rev protein. Gratifyingly, a conserved core recognition element was identified from in vitro selection that showed the same sequence as the wild-type domain II.52 Ellington and co-workers further shrank 122 the RNA pool size by only randomizing the nucleotides that were known to be involved in the interaction with Rev protein and peptide, thereby creating a pre-structured RNA library in which the two small random regions were separated by a constant loop.53 Interestingly, in vitro selection of RNA library against Rev peptide yielded two classes of aptamers: one shared conserved core element with native RRE and the other one did not. In the binding with the first class of RNA aptamer, Patel and co-workers found Rev peptide adopted helical conformation as expected in the native binding system, however, an extended conformation was identified in the binding with the second class of RNA aptamers. This finding shows the short Rev peptide loses its helical secondary structure when taken out of protein and can adopt adaptive binding to different RNA aptamers (Fig. 4.3).54

4.1.5 Stapled peptides enhance α-helicity and binding affinity

The α-helix is an important secondary structure in proteins and responsible for many protein-protein and protein-nucleic acid interaction. When taken out of context, short peptides will lose this structure, for the reason that internal hydrogen bonds are not strong enough. Researchers thus reinforced the α-helix of small peptide by stapling its side chains that were not involved in binding.55,56 Stapling shows nucleation effect that even one single stapling at the end of peptide will trigger α-helix formation and enhance the helicity for the entire peptide. Stapling usually occurs at i and i+4 positions, covering one helical turn. Double stapling or stapling i and i+7 is also reported. Different stapling strategies have been reported, including ring-closing metathesis,57,58 lactamization,59-61 cycloaddition62,63 or double-cycloaddition64,65 and thioether formation.66 Stapling greatly improves pharmacologic performance of peptides in terms of target binding affinity, proteolytic stability and cell permeability.67,68 More importantly, stapled peptides have been shown safe, efficacious, and biological pathway specific in animal models of human 123 cancer.69,70 Stapled HIV-1 Rev peptide showed higher binding affinity than non-stapled wide-type Rev peptide55 and inhibitory effect on Rev-RRE interaction.71

Figure 4.3 Short Rev peptide adopts adaptive binding to its RNA aptamers. Left:

Aptamer conserves core binding element; Right: Aptamer shares no similarity with

native RRE IIB. Cited from reference 54.

4.1.6 A bPNA-assisted reversibly-structured RNA library for in vitro selection against stapled HIV-1 Rev peptides and directed evolution of aptamers

Combining the RNA library construction and the target design, a RNA library of a random

N25 buttressed by two U10 regions was used in in vitro selection against stapled highly helical HIV-1 Rev peptide. We hypothesized the stapled highly helical Rev peptide would prefer RNA aptamers that conserved core binding element with the native RRE RNA during in vitro selection. Furthermore, the interaction between 66-nucleotide RRE domain 124

II and 17-amino acid HIV-1 Rev peptide has already been well established. NMR study showed that a non-Watson-Crick G-A nucleobase pair on RRE domain II interacted with an indispensable asparagine on Rev peptide through coplanar hydrogen bonding.72 This key interaction inspired us to design a directed in vitro selection through the mutation of this asparagine amino acid. To test this notion, we prepared two analogues to the stapled original Rev peptide: melamine-displaying stapled Rev peptide and ammeline-displaying stapled Rev peptide. Both melamine and ammeline are nucleobase mimics and known to interact with nucleobase pairs, for example, melamine interacts a U-U pair. Prior to in vitro selection, we demonstrated that these single-mutated stapled Rev peptides did not bind to native RRE IIB at the same condition, further confirming the key interaction between asparagine and G-A pair. Thus, three parallel in vitro selections of a reversibly-structured

N25 RNA library against original and artificial nucleobase-displaying stapled Rev peptides were performed to test the proposal that synthetic bPNA-assisted reversibly-structured

RNA library would benefit in vitro selection, stapled highly helical HIV-1 Rev peptide would prefer aptamers conserving core binding element as the native RRE RNA and single mutation on the target would direct the evolution of RNA library. The selected RNA aptamers of each round were analyzed by high-throughput sequencing. Although conventional cloning and Sanger sequencing is still prevalent nowadays to decipher the in vitro selection results, next generation high-throughput sequencing provides the information for each single sequence, resulting in a much more comprehensive and powerful analysis for in vitro selection.4,5,73,74

125

Isolated RRE IIB

17-mer HIV-1 Rev peptide sequence:

TRQARRNRRRRWRERQR

Figure 4.4 Structure of isolated short hairpin Rev responsive element (RRE), HIV-1 Rev

peptide and their interaction. Boxed nucleobases and bold amino acids contact directly

and cannot be replaced; underlined amino acids make electrostatic interaction with RNA

and can be replaced with lysine; aspargine in red interacts with key non-Watson-Crick G-

A pair on the same planar via hydrogen bonding. Adapted from Reference 71.

4.2 Results and Discussion

4.2.1 Stapled Rev peptides enhance helicity and binding affinity

To staple HIV-1 Rev peptide, two amino acids were replaced by azidoalanine at i and i+4 positions of C-terminus and followed by double cycloaddition with 1,5-hexadiyne. This

126 surface is known not to be involved in binding with RRE.75 The N-terminus is succinylated to further increase helicity of peptide.

To compare the helicity enhancement, four peptides with different stapling and succinylation properties were synthesized based on the original Rev peptide. It was clear that stapled succinylated Rev peptide showed significantly high helicity, which was comparable to a double stapled Rev peptide (92% helicity reported by Fairlie).55

Succinylated non-stapled Rev peptide and stapled non-succinylated Rev peptide showed similar helicity that was comparable to succinlyated Rev peptide (51% helicity reported by

Frankel).51 The non-stapled non-succinylated Rev peptide showed about 10% helicity compared to Frankel’s work. Single site mutation occurred on the only asparagine of original Rev peptide (M* denotes melamine-displaying amino acid; N* denotes ammeline- displaying amino acid). Together with the stapled succinylated Rev peptide, they all showed high helicity with maxima at 190 nm and double minima at 208 and 222 nm.

Gratifyingly, stapled Rev peptide showed higher binding affinity compared to non-stapled

Rev peptide, as evidenced by the shift on lane 2 of each gel image. When the ratio of peptide to RNA was 1:2, half of the RNA bound to peptide and product band moved slower than RNA band due to increased molecular weight and decreased negative charge.

Comparing stapled Rev peptide and non-stapled Rev peptide, they have the same number of positive charges, indicating enhanced binding affinity is not caused by charge change but helicity increase.

127

a b

c Stapled Rev peptide Non-stapled original Rev peptide

Peptide 0 0.5 1 2 4 8 0 0.5 1 2 4 8

Figure 4.5 α-helix of Rev peptides. (a) Stapling and succinylation on α-helix of original

Rev peptide. (b) CD spectra of stapled succinylated original and mutated Rev peptides.

(c) Stapled Rev peptide showed higher binding affinity than non-stapled wild-type Rev peptide (the numbers indicated peptide equivalent, RRE concentration was 500 nM, gel was stained by SYBR gold).

128

4.2.2 General procedure of in vitro selection

Figure 4.6 Procedure of in vitro selection of bPNA-assisted reversibley-structured RNA

library.

First of all, bPNA was mixed with RNA library in a ratio of 1:1 in 1 × annealing buffer and annealed at 95 °C for 5 minutes. The mixture was slowly cooled to room temperature over

2 hours. In this way, 10mer bPNA formed triplex stem with the two U10 tracts and left the

N25 random region in the loop to evolve for binding. The annealed library was then subject 129 to pre-selection with streptavidin beads which was used later to pull down binding aptamers. After pre-selection, the survived complexes separated from the beads were incubated with corresponding stapled Rev peptide for desired time and at desired temperature, followed by pulling down with the same amount of streptavidin beads. The beads were then washed with 1 × high salt wash buffer for 2 – 6 times to remove weak binders. The beads were heated at 95 °C for 5 minutes in 4 M guanidinium thiocyanate

(GuSCN) solution. This denaturation condition was strong enough to remove bPNA from

RNA, and RNA aptamers from the targets. A spin column was perfect to remove beads and retain stapled Rev peptides, bPNA, GuSCN in the column, while aptamers were eluted. Followed by the completion of selection, survived RNA aptamers were converted into complementary DNA by reverse transcription followed by PCR for amplification whose products were transcribed by T7 polymerase into enriched RNA library for next round in vitro selection. The pulldown yields in the 1st to 6th round were usually between 0.5% to

0.1%, and increased to 2.0% - 7.1% in 7th to 10th round. The PCR products of each round of selection were sent to high-throughput sequencing.

4.2.3 Significant enrichment is observed after 6th round

Along with the progress of in vitro selection, random nucleobases got paired and loose structures became more compact. This can be monitored by UV melting. RNA pool 0 subject to Round 1 selection annealed with bPNA showed only one relatively large transition due to the binding to bPNA. A new smaller transition was observed at 20 C prior to the Round 4 selection, however it may shift into higher temperature range, making it undetectable. Starting the Round 7, a transition at about 80 C appeared, indicative the formation of highly-structured aptamers.

130

Figure 4.7 UV melting of RNA library at different round with bPNA. From left to right:

Pool 0, Pool 3 and Pool 6 prior to 1st round, 4th round and 7th round selection,

respectively.

High-throughput sequencing results also revealed significant enrichment starting from

Round 6 as evidenced by the unique sequence percentage in the pool. The dramatic decrease of unique sequence percentage indicates the increased reads per million

RPM) for each sequence, for example, the RPM of the top sequence of the second selection increased from 18 in the 1st round to 397181 in the 10th round.

Figure 4.8 Unique sequence percentage of each pool in three selections. Red line:

original stapled Rev peptide; Blue line: M* stapled Rev peptide; Green line: N* stapled

Rev peptide.

131

4.2.4 Sequences of interest are identified and share similarities with native RRE

To simplify the high-throughput sequencing data analysis, flanking constant primer regions and T10-tracts are removed using Cutadapt,76 only the N25 random region was used for analysis by Fastaptamer.77 Two major clusters (Cluster 1 and Cluster 2) were identified from the 1st selection against stapled original Rev peptide. Interestingly, the most populated single sequence had about 81% of the total reads of the 10th pool. The aptamers selected against stapled original Rev peptide had very good specificity, as they showed very low RPM in the 2nd and 3rd selections. Since only one amino acid is different on the three targets, this good specificity (over 1000-fold) may indicate the key amino acid is involved in binding. In the 2nd selection, only one major cluster (Cluster 3) was identified.

Interestingly, two top sequences in this cluster had similarly high RPM. The specificity of these two sequences were not as good as the sequences identified in the 1st selection, as they also had over ten thousand of RPM in the 3rd selection. Several clusters were found in the 3rd selection, including the major Cluster 3. However, the exact sequences of the top aptamers were not the same between the 2nd and 3rd selections. This may be caused by the slight difference on the side chain of the synthetic amino acids. Top sequences identified from the 3rd selection showed good specificity. When examining the enrichment of each round in the three selections, the top aptamers found in that selection usually started to enrich at 5th-6th round and only enriched in that selection instead of the other two, indicative of great data specificity between different selections (Table 4.1).

The foldings of three top aptamers in the 1st selection were predicted using online UNA program provided by IDT DNA, inc. Current folding prediction program would not be able to consider abiotic bPNA-RNA triplex, instead a dummy well-paired duplex stem was used to replace bPNA-RNA triplex for folding prediction purpose. They all folded into imperfect hairpins which were promising for binding to helical Rev peptide.54,77-80 Most gratifyingly, 132

Table 4.1 Populated sequences identified from three in vitro selections

SELEX Cluster Top Sequences (N25) Reads per million (RPM)

Original Melamine Ammeline

Original 1 CUAAUCUAUAGCAAUUAUGUGUCGC 808774 633 30

AAUCUAUAGCAAUUAUGUGUCGC 29906 35 -

2 GGAGUAAACCAACAAGUGAAAGGGC 23041 27 -

GGAGUAAACCAACGUUGUAAGGGC 4191 - -

Melamine 3 AAAGUAAAAACGACUAAUGAUCACC - 445681 11099

AAAGCUAAAACGACUAACGAUCACC - 417148 16451

133

Ammeline 3 AAAGAUAAAACGACUAAAGAUCACC - 81 397181

CAAGUAAAAACGACUAAUGAUCACC - 955 151153

4 ACAGUUGACGGAUCUGAGAUAUUCC - 8 77684

ACUGUUGACGGGUCUAAGACAUUCC - 37 65080

5 GAGCAGAACUUGUUUAAACCACCAC - - 38040

GAGUUAAAACGACUGAUGAUCACC - 8 15334

133 folding of top sequence in cluster 2 (Figure 4.9b) conserved core binding element as the native RRE IIB and IA (Figure 4.9c), the latter one was a recently discovered additional binding site of HIV-1 Rev peptide by Frankel.78

a b

c

Figure 4.9 Predicted folding of top aptamers of original stapled Rev peptide and native

RRE. (a) Top sequence in Cluster 1; (b) Top sequence in Cluster 2; (c) Native RRE

RNA sequence (adapted from reference 78).

134

4.2.5 Single mutation on Rev peptide directs the evolution of RNA library

a b c

d

Figure 4.10 Predicted folding of top sequences in Cluster 3 and possible bifacial

recognition. (a) Predicted folding with dummy duplex stem; (b) Predicted folding of top

sequence of Cluster 3 in 2nd selection without dummy duplex stem; (c) Predicted folding

of top sequence of Cluster 3 in 3rd selection without dummy duplex stem; (d) Possible

bifacial recognition between synthetic triazine nucleobase mimics and U-U mismatch.

135

The folding prediction of Cluster 3 with dummy duplex stem had many unpaired adenines in the loop, which reminded us this cluster may fold itself into a duplex without incorporating bPNA, as triplex formation and duplex formation were competing process.

Actually, the high-throughput sequencing data of Cluster 3 revealed a high percentage of shortened T-tract on 5’ end. This would lower the binding affinity to bPNA and facilitate the self duplex formation. Indeed, the folding predictions of three top sequences in Cluster

3 without dummy duplex stem showed similar imperfect duplex hairpin, featuring a 7-nt loop, a bulge and a U-U mismatch. It may be possible that the melamine and ammeline interacted with U-U mismatch in a fashion of bifacial recognition (Figure 4.10d).

4.3 Conclusion

Abiotic bPNA is used to construct a reversibly-structured RNA library for in vitro selection against stapled highly helical HIV-1 Rev peptides and high-throughput sequencing reveals aptamers conserved core binding element as the native RRE RNA. Single mutation of key amino acid on the target is demonstrated to direct the evolution of RNA library. Abiotic bPNA-RNA triplex stem is possible to be replaced by evolved native duplex stem.

4.4 Materials and Equipments

Chemicals for amino acid derivative synthesis, solid phase peptide synthesis, purification and characterization were purchased from Sigma-Aldrich, Chem-Impex, VWR, Alfa Aesar and AAPPTec without further purification unless otherwise specified. Rink Resin LS (100-

200 mesh, 0.28 mmol/g) from Advanced ChemTech. DNA template and primers were obtained from IDT DNA. 40% acrylamide and bis-acrylamide solution (19:1) and ammonium persulfate were purchased from Amresco. TBE Buffer (10 ×) for gel electrophoresis was purchased from American Bioanalytical. TEMED was from Bio-Rad 136

SYBR Gold was purchased from Invitrogen. T7 polymerase, Taq polymerase and M-MuLv reverse transcriptase and their corresponding 10 × reaction buffer were purchased from

New England BioLabs. Streptavidin agarose and 5.0 µm centrifugal filters were obtained from EMD Millipore. MiniElute PCR purification kit was from Qiagen.

The concentration of all DNA oligonucleotides and synthetic peptides was determined by

UV absorbance using Nanodrop. MALDI-TOF Mass spectra were acquired on Bruker

Microflex MALDI-TOF instrument. Electrospray mass spectroscopy (ESI) was acquired on

Bruker MicroTOF equipped with an electrospray ionization source. Mass spectrometry instruments were provided by a grant from the Ohio BioProducts Innovation Center. NMR spectra were acquired on Bruker Advance DPX 400 instrument. Solid phase peptide synthesis was performed on AAPPTec Apex 396 peptide synthesizer. All peptides were purified on RP-C18 preparative HPLC column and the purity was confirmed on RP-C18 analytical column. UV melting was performed on Varian Cary-100 UV-Vis

Spectrophotometer equipped with Cary Temperature Controller. Gel images were acquired using a Typhoon Trio Variable Mode Imager (Amersham Biosciences). PCR was done Eppendorf Mastercycler Personal 5883.

4.5 Sequences

4.5.1 Stapled Rev peptide sequences

Stapled original Rev peptide

Stapled melamine Rev peptide

Stapled ammeline Rev peptide

137

All peptides are succinylated on N-terminus and amidated on C-terminus. The peptides

are also biotinylated for pulldown purpose. Az denotes azidoalanine and the side chains

are stapled by double cycloaddition with 1,5-hexadiyne. N denotes key amino acid,

asparagine; M* represents melamine-displaying L-2,3-diaminopropionic acid; N*

represents ammeline-displaying L-2,3-diaminopropionic acid. Melamine and ammeline

are monoderivatized on exocyclic nitrogen atom.

4.5.2 Nucleic acid sequences

DNA template and primer sequences. Underlined is T7 promoter region.

DNA template ordered from IDT:

5'- TCG TCG TAT GTG AAT TCA GC-A10N25A10-G CGG ATC CTA CCC TAT AGT

GAG TCG TAT TA -3'

Transcribed N25 RNA library:

5’-GGG UAG GAU CCG C-U10N25U10-GC UGA AUU CAC AUA CGA CGA-3’

Reverse primer: 5'- TCG TCG TAT GTG AAT TCA GCA AAA AAA AAA -3'

Forward primer: 5’-TAA TAC GAC TCA CTA TAG GGT AGG ATC CGC-3’

138

4.6 Synthesis

4.6.1 Synthesis of Fmoc-azidoalanine

Scheme 4.1 Synthesis of Fmoc-azidoalanine

Fmoc-azidoalanine was synthesized according to published literatures.63,75 Briefly, 3g of

NaN3 was dissolved in a mixture of 10 mL H2O and 20 mL dichloromethane and stirred vigorously at 0 °C while adding 1.5 mL triflic anhydride dropwise. The heterogeneous solution was stirred for 2 hours at room temperature. Water layer was extracted with DCM

10 mL × 3. DCM layers were then combined and washed once with 15 mL saturated

NaHCO3. To a mixture of 0.9 g Boc-Dap-OH, 0.92 g K2CO3 and 30 mg CuSO4 in 20 mL

H2O and 30 mL MeOH was added the resulting DCM layer slowly. The heterogeneous solution was stirred at room temperature overnight. After removing organic solvent under reduced pressure, pH was adjusted to about 3 by concentrated hydrochloric acid, followed by extraction with DCM three times. DCM layer was then dried over Na2SO4 and condensed, yielding crude Boc-azidoalanine. It was used without further purification.

5 mL TFA was used to remove Boc protecting group for 1 hour, followed by blowing TFA to sticky oil. Deprotected azidoalanine was readily precipitated by adding cold diethyl ether.

The solid was collected by centrifuge and dissolved in 30 mL water. After adjusting pH to

~7.5, 1.48 g Fmoc-OSu dissolved in 30 mL THF was added to this aqueous solution. This 139 solution was stirred at room temperature overnight. Unreacted Fmoc-OSu was removed by ethyl acetate extraction. The resulting aqueous solution was then acidified to pH 2~3 by 1 M hydrochloric acid followed by ethyl acetate extraction. After drying and removing organic solvent, the crude product was purified by column chromatography in 2% and 98% DCM, yielding white powder, Fmoc-azidoalanine. 1H NMR (400 M, DMSO- d6) δ (ppm): 3.614 (d, J=6.4, 2H), 4.236-4.322 (m, 4H), 7.325 (t, J=7.6Hz, 2H), 7.420 (t,

J=7.2Hz, 2H), 7.731 (d, J=7.6Hz, 2H), 7.895 (t, J=7.8Hz, 2H), 7.919 (s, 1H), 13.065 (s,

1H). Calculated [M+Na]: 375.1069; Found [M+Na]: 375.0898.

4.6.2 Synthesis of Fmoc-Lys(Biotin)-OH

Scheme 4.2 Preparation of Fmoc-Lys(Biotin)-OH

140

2.44 g Biotin was stirred with 1.4 g N-hydroxysuccinimide (1.2 equiv.) and 2.3 g EDC (1.2 equiv) in 30 mL DMF overnight. DMF was removed under reduced pressure and the resulting white solid was washed twice with 1 M HCl, yielding 82 % Biotin-NHS. 1H NMR

(400 M, DMSO-d6) δ (ppm): 1.476-1.685 (m, 6H), 2.596 (S, 1H), 2.670 (t, J=7.6Hz, 2H),

2.729-2.889 (m, 5H), 3.082-3.128 (m, 1H), 4.131-4.161 (dd, J=8.0Hz, 4.4Hz, 1H), 4.290-

4.321 (dd, J=8.0Hz, 4.4Hz, 1H) 6.418 (m, 2H).

3.12 g Biotin-NHS was reacted with 2.37 g Boc-Lys-OH (1.05 equiv.) in the presence of

1.9 mL trimethylamine in 30 mL DMF overnight. DMF was removed under reduced pressure and the resulting white solid was washed twice with pH 3 HCl, yielding 75% Boc-

Lys(Biotin)-OH. 1H NMR (400 M, DMSO-d6) δ (ppm): 1.285-1.637 (m, 21H), 2.036 (t,

J=7.6Hz, 2H), 2.556 (d, J=12.4Hz, 1H), 2.798-2.842 (dd, J=12.4Hz, 5.2Hz, 1H), 2.971-

3.017 (q, J=5.6Hz, 2H), 3.084-3.101 (q, J=2.4Hz, 1H), 3.816 (m, 1H), 4.109-4.139 (m, 1H),

4.284-4.316 (m, 1H), 6.346 (s, 1H), 6.408 (s, 1H), 6.997 (d, J=8.0Hz, 1H), 7.732 (t,

J=5.6Hz, 1H), 12.398 (s, 1H). Calculated [M+Na]: 495.2253; Found [M+Na]: 495.2015

0.52 g Boc-Lys(Biotin)-OH was reacted with TFA for 1 hour to remove Boc protecting group. TFA was then removed by reduced pressure, followed by dissolving the resulting sticky oil into 15 mL water and neutralizing it with Na2CO3 solid to bring pH to ~ 7.5. To the solution above was added 0.44 g Fmoc-OSu in 15 mL THF. The solution was stirred overnight. Unreacted Fmoc-OSu was readily extracted with ethyl acetate three times, followed by acidifying the aqueous solution with 1 M HCl. White precipitate was collected, washed with 1 M HCl and dried to yield Fmoc-Lys(Biotin)-OH. 1H NMR (400 M, DMSO- d6) δ (ppm): 1.277-1.389 (m, 12H), 2.041 (t, J=7.6Hz, 2H), 2.570 (d, J=12.4Hz, 1H), 2.783-

2.826 (dd, J=12.4Hz, 4.8Hz, 1H), 3.009-3.078 (m, 3H), 3.912 (m, 1H), 4.097-4.127 (dd,

J=7.6Hz, 4.4Hz, 1H) 4.221-4.307 (m, 4H), 6.432 (br, 2H), 7.346 (td, J=7.6Hz, 1.2Hz, 2H),

141

7.416 (t, J=7.6Hz, 2H), 7.617 (d, J=8Hz, 2H), 7.727 (d, J=7.6Hz, 2H), 7.775 (t, J=5.2Hz,

1H), 7.890 (d, J=7.2Hz, 2H). Calculated [M+H]: 595.2592; Found [M+H]: 595.2189

4.6.3 Synthesis of Fmoc-Dap(M*)-OH

Scheme 4.3 Synthesis of Fmoc-Dap(M*)-OH

0.5 g Boc-Dap-OH was reacted with 0.41 g 2,4-diamino-6-chloro-1,3,5-triazine in the presence of 0.2 g NaOH in 10 mL water at 85 °C overnight. Triazine was removed by filtration. The aqueous layer was washed three times with ethyl acetate to remove dissolved triazine. Aqueous layer was then dried under reduced pressure. The crude product, Boc-Dap(M*)-OH was used without further purification. Boc protecting group was removed in TFA in 2 hours. After removing TFA under reduced pressure, the oily product was dissolved in 20 mL water and neutralized to pH ~7.5 by Na2CO3 solid. To the solution was added 0.88 g Fmoc-OSu in 20 mL THF. After overnight stirring, unreacted Fmoc-OSu was extracted three times by ethyl acetate. The resulting aqueous solution was acidified by 1 M HCl to pH 1-2. Heavy white precipitate was observed, collected through filtration, washed with 1 M HCl and dried, yielding Fmoc-Dap(M*)-OH (Yield: 65%). 1H NMR (400

142

M, DMSO-d6) δ (ppm): 3.470 (q, J=6.8Hz, 1H), 3.574 (q, J=6.8Hz, 1H), 4.053-4.109 (q,

J=6.8Hz, 1H), 4.197-4.269 (m, 3H), 6.510-6.757 (m, 4H), 7.333 (t, J=7.2Hz, 2H), 7.411 (t,

J=7.2Hz, 2H), 7.515 (d, J=7.2Hz, 1H), 7.698 (d, J=7.6Hz, 2H), 7.886 (d, J=7.6Hz, 2H). 13C

NMR (100 MHz, DMSO-d6) δ (ppm) 40.94, 46.57, 54.34, 65.78, 120.07, 125.29, 127.12,

127.60, 140.66, 143.74, 143.82, 155.84, 173.10. Calculated [M+H]: 436.1735; Found

[M+H]: 436.1511

4.6.4 Synthesis of Fmoc-Dap(N*)-OH

Scheme 4.4 Synthesis of Fmoc-Dap(N*)-OH

0.45 g Boc-Dap-OH was reacted with 0.42 g 2-amino-4,6-dichloro-1,3,5-triazine in presence of 0.37 g NaHCO3 in 10 mL water at 50 °C overnight. Triazine was removed by filtration. The aqueous layer was washed three times with ethyl acetate to remove dissolved triazine. Aqueous layer was then dried under reduced pressure. The crude product, Boc-Dap(N*)-OH was used without further purification. The removal of Boc protecting group and the hydrolysis of chloride on triazine were done by reacting with TFA at 50 °C for two hours. After removing TFA under reduced pressure, the oily product was 143 dissolved in 20 mL water and neutralized to pH ~7.5 by Na2CO3 solid. To the solution was added 0.82 g Fmoc-OSu in 20 mL THF. After overnight stirring, unreacted Fmoc-OSu was extracted three times by ethyl acetate. The resulting aqueous solution was acidified by 1

M HCl till heavy white precipitate was observed. The white precipitate was collected through filtration, washed with 1 M HCl and dried, yielding Fmoc-Dap(N*)-OH (Yield: 52%).

1H NMR (400 M, DMSO-d6) δ (ppm): 3.507-3.543 (m, 1H), 3.552-3.591 (m, 1H), 4.204-

4.281 (m, 4H), 7.323 (t, J=7.6Hz, 2H), 7.409 (t, J=7.6Hz, 2H), 7.658-7.707 (m, 3H), 7.773

(m, 1H), 7.885 (d, J=7.6Hz, 2H). 13C NMR (100 MHz, DMSO-d6) δ (ppm): 41.49, 47.05,

53.94, 66.31, 120.57, 125.71, 127.58, 128.11, 141.17, 144.25, 156.49, 161.06, 172.47.

Calculated [M+H]: 437.1575; Found [M+H]: 437.1321

144

4.6.5 Rev peptide synthesis and stapling

Standard Fmoc chemistry was used to synthesize HIV-1 Rev peptides. 166 mg Rink amide resin (0.28 mmol/g) was used. Notably, H-Lys(Biotin)-OH, azidoalanine, H-Dap(M*)-OH and H-Dap(N*)-OH did not require side-chain protection. N-terminus was succinylated by shaking resins with 5 equiv. of succinic anhydride in the presence of 5 equiv. of trimethylamine for 1 hour in DMF. Non-stapled Rev peptides were cleaved and deprotected in a cocktail of 88% TFA, 5% H2O, 5% m-cresol and 2% triisopropylsilane (4 mL total volume) and precipitated in cold ether. Non-stapled Rev peptides were purified by preparative C-18 reverse phase HPLC and lyophilized.

Purified non-stapled Rev peptides were stapled using double cycloaddition described previously.63,76 Briefly, several milligrams of non-stapled Rev peptide bearing azide at i and i+4 positions were dissolved in degassed 100 mM pH 7.9 Tris-HCl buffer followed by adding 1,5-hexadiyne in pentane/DMF (1.1 equiv.), CuSO4 solution (1 equiv.) THPTA

(Tris(3-hydroxypropyltriazolylmethyl)amine, 1 equiv.) and sodium ascorbate (3 equiv.).

The clear solution was mildly shaken for 1 hour. Stapled Rev peptides were then purified by preparative C-18 reverse phase HPLC and the purity was confirmed on analytical C-

18 reverse phase HPLC. Detection wavelength was 214 nm.

145

Succinilyation 0:G4 MS Raw Intens. [a.u.] Intens.

1500 a

1000

3052.172

500

1526.116

0 1500 1750 2000 2250 2500 2750 3000 3250 3500 m/z b

Figure 4.11 Characterization of stapled original Rev peptide. (a) Maldi-Tof Mass spectrum, calculaed [M+H]: 3051.441; Found [M+H]: 3052.172. (b) HPLC trace for stapled original Rev peptide.

146

Stapled M REV 0:H12 MS Raw 1500

a Intens. [a.u.] Intens.

1250

1000

3129.593

750

500

250

0 1000 1500 2000 2500 3000 3500 m/z b

Figure 4.12 Characterization of stapled melamine-displaying Rev peptide. (a) Maldi-Tof

Mass spectrum, calculated [M+H]: 3131.878; found [M+H]: 3129.593. (b) HPLC trace for stapled melamine-displaying Rev peptide.

147

Stapled M REV 0:E4 MS Raw a

1200 Intens. [a.u.] Intens.

1000

800 3130.851

600

1565.440 400

200

0 1000 1500 2000 2500 3000 3500 m/z b

Figure 4.13 Characterization of stapled ammeline-displaying Rev peptide. (a) Maldi-Tof mass spectrum, calculated [M+H]: 3132.862; found [M+H]: 3130.851. (b) HPLC trace for stapled ammeline-displaying Rev peptide.

148

4.7 Experimental Procedures

10 rounds of iterative in vitro selection were done with increased stringency, as summarized in Table 4.1. Session 4.7.1 - 4.7.10 take 1st round as an example. For later rounds, the scale and conditions were changed accordingly. 1 × binding buffer contains

10 mM HEPES-KOH and NaCl of different concentration as indicated, pH 7.5. Wash buffer contains 10 mM HEPES-KOH and 1 M NaCl, pH 7.5.

Table 4.2 In vitro selection condition

RNA Rev peptide NaCl Incubation time

Round concentration concentration concentration and temperature wash

1st 1 µM 1 µM 100 mM 60 min, 0 °C 2

2nd 0.8 µM 0.8 µM 100 mM 60 min, 0 °C 4

3rd 0.6 µM 0.6 µM 100 mM 60 min, 0 °C 6

4th 0.6 µM 0.6 µM 400 mM 60 min, 0 °C 6

5th 0.5 µM 0.5 µM 400 mM 60 min, 0 °C 6

6th 0.5 µM 0.5 µM 1 M 60 min, 0 °C 6

7th 0.4 µM 0.4 µM 1 M 60 min, 0 °C 6

8th 0.4 µM 0.4 µM 1 M 10 min, RT 6

9th 0.4 µM 0.3 µM 1 M 0 min, RT 6

10th 0.4 µM 0.2 µM 1 M 0 min, RT 6

149

4.7.1 Transcription of N25 RNA library

4 nmol of DNA template annealed with T7 promoter in 1 × annealing buffer was transcribed into RNA library using 1 × HEPES high yielding buffer (10 × HEPES buffer includes 1 M

HEPES-KOH, 100 mM MgCl2, 20 mM Spemidine hydrochloride, 400 mM DTT, pH 7.5) in

1.5 hours. Heavy white precipitate formed during transcription was dissolved by adding

100 µL 500 mM EDTA solution prior to adding 200 µL 5 M NH4Ac followed by precipitation with 5.5 mL ethanol. The resulting white precipitate was stored at -20 °C for 2 hours before purification by 15% 8M urea PAGE. The RNA band was visualized by UV shadowing on

TLC plate. The correct band was excised and eluted with RNase-free water overnight at room temperature. Solution was then separated from gel solid and precipitated with 10% volume of 5 M NH4Ac solution and 2.5 volume ethanol. The solution was store at -20 °C for 2 hours or 10 minutes on dry ice before centrifuge at 13200 rpm at 4 °C. White RNA precipitate was washed once with 200 µL 70% ethanol and centrifuge again. RNA library was finally dried in spin vacuum for 20 minutes at room temperature. RNA solid was readily dissolved in water and the concentration was measured by Nanodrop.

150

Table 4.3 Transcription condition for 1st round

Reagents Volume Final concentration

H2O 4000 µL -

10 × buffer 800 µL 1 ×

0.1 M DTT 800 µL 10 mM

0.5 M MgCl2 320 µL 20 mM

25 mM NTP mixture 960 µL 3 mM

50% 960 µL 6%

T7 polymerase 80 µL -

50 µM DNA template 80 µL 500 nM

4.7.2 Annealing of N25 RNA library with bPNA

N25 RNA library was annealed with (EM*)10 bPNA in a ratio of 1:1 in 1 × annealing buffer

(5 × buffer contains 50 mM Tris-HCl, 100 mM NaCl and 5 mM EDTA, pH 8.0) at 95 °C and then slowly cooled to room temperature in 2 hours.

151

Table 4.4 Annealing condition

Reagents Volume

H2O 642 µL

5 × annealing buffer 200 µL

bPNA (33 µM) 120 µL

N25 RNA library (29 µM) 138 µL

4.7.3 Pre-selection of annealed N25 RNA library

Annealed RNA library was selected against streptavidin agarose beads in case that any streptavidin aptamer presents in the pool. The heterogeneous solution was rotated end- to-end for 40 minutes at room temperature. The solution was separated from beads through 5.0 µm centrifugal filter.

Table 4.5 Pre-selection condition

Reagents Volume

Annealed RNA library 1000 µL

10 × binding buffer 400 µL

H2O 2350 µL

Streptavidin beads 150 µL

152

4.7.4 In vitro selection against target

Elute collected from pre-selection was incubated with stapled Rev peptide for 1 hour on ice. The ratio of annealed RNA library and target was 1:1. For the first round against stapled original Rev peptide, 12.8 µL of 313 µM peptide was added.

4.7.5 Pull-down of binding aptamer

150 µL streptavidin beads were added to the solution and end-to-end rotation was done for 40 minutes. The amount of streptavidin was at least 3 times more than the amount of stapled Rev peptides which was biotinylated at C-terminus. Regardless of binding with aptamer or not, all stapled Rev peptides in the solution should be pulled down by streptavidin beads. The binding aptamers associated with stapled Rev peptides were also pulled down.

4.7.6 Wash the beads with high salt buffer

The beads were washed with high salt buffer for 2 minutes and 500 µL of wash buffer was applied for end-to-end rotation. Beads was collected. Wash was repeated accordingly.

4.7.7 Denaturation and collection of binding aptamers

The beads were incubated with 100 µL 4 M guanidinium thiocyanate (GuSCN) and heated at 95 °C for 5 minutes. The resulting mixture was loaded to a Bio-Rad P-30 spin column whose buffer was exchanged with water. The spinning time followed manufacturer's instruction. This desalting procedure efficiently removed GuSCN, bPNA, stapled Rev peptide and streptavidin beads, eluting only binding aptamers which survived buffer wash.

The elute solution was reduced to about 20 µL by spin vacuum at room temperature.

153

4.7.8 Reverse transcription of selected RNA into complementary DNA

For the first round selection against original stapled Rev peptide, the pulldown yield was

0.49%, which was about 19.45 pmol RNA. They were annealed with 1.5 equiv. of reverse primer in 1 × annealing buffer at 75 °C for 5 minutes and then on ice for 5 minutes, followed by adding reverse transcription buffer, dNTP mix and M-MuLv reverse transcriptase. The reaction was carried out at 42 °C for 1 hour. The enzyme was deactivated at 90 °C for 10 minutes.

Table 4.6 Reverse transcription condition

Reagents Volume Final concentration

1) annealing

Selected RNA library 24 µL 0.56 µM

10 × annealing buffer 3.5 µL 1 ×

Reverse primer (5.35 µM) 5.45 µL 0.83 µM

H2O 2.05 µL -

2) reverse transcription

Anneal complex 35 µL 0.39 µM

10 mM dNTP mixture 2.5 µL 0.5 mM for each dNTP

5 × reverse transcription buffer 10 µL 1 ×

M-MuLv reverse transcriptase 2 µL -

154

4.7.9 PCR amplification of cDNA

PCR amplification of library is known to be a problem during SELEX. According to several publications,83,84 16 rounds amplification was ideal in terms of product yield and byproduct formation. To avoid significant primer dimer formation, manual hot start was used. Briefly,

Taq polymerase was separated from the other reagents and heated at 95 °C simultaneously for 2 minutes, followed by transferring 30 µL solution from tube 1 to tube

2 which contained 20 µL solution. The denaturation was done at 95 °C for 30 seconds, annealing was done at 52 °C for 1 minute and polymerization was done at 68 °C for 20 seconds. The PCR product was then purified by Qiagen MiniElute PCR purification kit and ready for transcription. 2% agarose gel is used to check the quality of PCR to ensure no obvious byproducts formed.

Table 4.7 Manual hot start PCR condition

Reagents Tube 1 Tube 2

H2O 458 µL 355 µL

10 × strand polymerization buffer 60 µL 40 µL

10 mM dNTP mixture 20 µL -

Forward primer (9,94 µM) 20.12 µL -

Reverse primer (10.7 µM) 18.68 µL -

cDNA 25 µL -

Taq polymerase - 5 µL

155

It is known that iterative selection will cause mutations on the length of the library. To check the quality of the PCR duplex products before sending to high-throughput sequencing, PCR duplex from the last round selection against each target were compared on 2% agarose gel with 50bp DNA ladder, stained by ethidium bromide. They still showed a good match with the initial length, 95 bp.

Figure 4.14 2% Agarose gel image for PCR duplexes in the 10th round of selection.

4.8 Acknowledgement

The authors are indebted to Ying Yu for her generous computer assistance.

4.9 References

(1) Ellington, A. D.; Szostak, J. W. In Vitro Selection of RNA Molecules That Bind

Specific Ligands. Nature 1990, 346, 818–822.

(2) Tuerk, C.; Gold, L. Systematic Evolution of Ligands by Exponential Enrichment:

RNA Ligands to Bacteriophage T4 DNA Polymerase. Science 1990, 249, 505–510.

(3) Joyce, G. F. Amplification, Mutation and Selection of Catalytic RNA. Gene 1989, 82,

83–87.

156

(4) Ameta, S.; Winz, M.-L.; Previti, C.; Jäschke, A. Next-Generation Sequencing

Reveals How RNA Catalysts Evolve from Random Space. Nucleic Acids Res. 2014,

42, 1303–1310.

(5) Dupont, D. M.; Larsen, N.; Jensen, J. K. Characterisation of Aptamer–target

Interactions by Branched Selection and High-Throughput Sequencing of SELEX

Pools. Nucleic acids 2015.

(6) Stoltenburg, R.; Reinemann, C.; Strehlitz, B. SELEX--a (r)evolutionary Method to

Generate High-Affinity Nucleic Acid Ligands. Biomol. Eng. 2007, 24, 381–403.

(7) Patel, D. J. Structural Analysis of Nucleic Acid Aptamers. Curr. Opin. Chem. Biol.

1997, 1, 32–46.

(8) Kimoto, M.; Yamashige, R.; Matsunaga, K.-I.; Yokoyama, S.; Hirao, I. Generation of

High-Affinity DNA Aptamers Using an Expanded Genetic Alphabet. Nat. Biotechnol.

2013, 31, 453–457.

(9) Michaud, M.; Jourdan, E.; Villet, A.; Ravel, A.; Grosset, C.; Peyrin, E. A DNA

Aptamer as a New Target-Specific Chiral Selector for HPLC. J. Am. Chem. Soc.

2003, 125, 8672–8679.

(10) Shoji, A.; Kuwahara, M.; Ozaki, H.; Sawai, H. Modified DNA Aptamer That Binds the

(R)-Isomer of a Thalidomide Derivative with High Enantioselectivity. J. Am. Chem.

Soc. 2007, 129, 1456–1464.

(11) Famulok, M.; Szostak, J. W. Stereospecific Recognition of Tryptophan Agarose by

in Vitro Selected RNA. J. Am. Chem. Soc. 1992, 114, 3990–3991.

(12) Jenison, R. D.; Gill, S. C.; Pardi, A.; Polisky, B. High-Resolution Molecular

Discrimination by RNA. Science 1994, 263, 1425–1429.

157

(13) Shangguan, D.; Li, Y.; Tang, Z.; Cao, Z. C.; Chen, H. W.; Mallikaratchy, P.; Sefah,

K.; Yang, C. J.; Tan, W. Aptamers Evolved from Live Cells as Effective Molecular

Probes for Cancer Study. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 11838–11843.

(14) Fang, X.; Tan, W. Aptamers Generated from Cell-SELEX for Molecular Medicine: A

Chemical Biology Approach. Acc. Chem. Res. 2010, 43, 48–57.

(15) Bock, L. C.; Griffin, L. C.; Latham, J. A.; Vermaas, E. H.; Toole, J. J. Selection of

Single-Stranded DNA Molecules That Bind and Inhibit Human Thrombin. Nature

1992, 355, 564–566.

(16) Jellinek, D.; Green, L. S.; Bell, C.; Janjić, N. Inhibition of Receptor Binding by High-

Affinity RNA Ligands to Vascular Endothelial Growth Factor. Biochemistry 1994, 33,

10450–10456.

(17) Ruckman, J.; Green, L. S.; Beeson, J.; Waugh, S.; Gillette, W. L.; Henninger, D. D.;

Claesson-Welsh, L.; Janjic, N. 2′-Fluoropyrimidine RNA-Based Aptamers to the 165-

Amino Acid Form of Vascular Endothelial Growth Factor (VEGF165) Inhibition of

Receptor Binding and VEGF-Induced Vascular Permeability through Interactions

Requiring the Exon 7-Encoded Domain. J. Biol. Chem. 1998, 273, 20556–20567.

(18) Tuerk, C.; MacDougal, S.; Gold, L. RNA Pseudoknots That Inhibit Human

Immunodeficiency Virus Type 1 Reverse Transcriptase. Proc. Natl. Acad. Sci. U. S.

A. 1992, 89, 6988–6992.

(19) Williams, K. P.; Liu, X. H.; Schumacher, T. N.; Lin, H. Y.; Ausiello, D. A.; Kim, P. S.;

Bartel, D. P. Bioactive and Nuclease-Resistant L-DNA Ligand of Vasopressin. Proc.

Natl. Acad. Sci. U. S. A. 1997, 94, 11285–11290.

(20) Ylera, F.; Lurz, R.; Erdmann, V. A.; Fürste, J. P. Selection of RNA Aptamers to the

Alzheimer’s Disease Amyloid Peptide. Biochem. Biophys. Res. Commun. 2002, 290,

1583–1588. 158

(21) Scarabino, D.; Crisari, A.; Lorenzini, S.; Williams, K.; Tocchini-Valentini, G. P. tRNA

Prefers to Kiss. EMBO J. 1999, 18, 4571–4578.

(22) Ellington, A. D.; Szostak, J. W. Selection in Vitro of Single-Stranded DNA Molecules

That Fold into Specific Ligand-Binding Structures. Nature 1992, 355, 850–852.

(23) Wilson, C.; Szostak, J. W. Isolation of a Fluorophore-Specific DNA Aptamer with

Weak Redox Activity. Chem. Biol. 1998, 5, 609–617.

(24) Sassanfar, M.; Szostak, J. W. An RNA Motif That Binds ATP. Nature 1993, 364,

550–553.

(25) Famulok, M.; Szostak, J. W. In Vitro Selection of Specific Ligand-Binding Nucleic

Acids. Angew. Chem. Int. Ed Engl. 1992, 31, 979–988.

(26) Wang, Y.; Rando, R. R. Specific Binding of Aminoglycoside Antibiotics to RNA.

Chem. Biol. 1995, 2, 281–290.

(27) Burke, D. H.; Hoffman, D. C.; Brown, A.; Hansen, M.; Pardi, A.; Gold, L. RNA

Aptamers to the Peptidyl Transferase Inhibitor Chloramphenicol. Chem. Biol. 1997,

4, 833–843.

(28) Hofmann, H. P.; Limmer, S.; Hornung, V.; Sprinzl, M. Ni2+-Binding RNA Motifs with

an Asymmetric Purine-Rich Internal Loop and a GA Base Pair. RNA 1997.

(29) Ng, E. W. M.; Shima, D. T.; Calias, P.; Cunningham, E. T.; Guyer, D. R.; Adamis, A.

P. Pegaptanib, a Targeted Anti-VEGF Aptamer for Ocular Vascular Disease. Nat.

Rev. Drug Discov. 2006, 5, 123–132.

(30) Blind, M.; Blank, M. Aptamer Selection Technology and Recent Advances. Mol. Ther.

Nucleic Acids 2015, 4, e223.

(31) Iliuk, A. B.; Hu, L.; Tao, W. A. Aptamer in Bioanalytical Applications. Anal. Chem.

2011, 83, 4440–4452.

159

(32) Song, K.-M.; Lee, S.; Ban, C. Aptamers and Their Biological Applications.

Sensors 2012, 12, 612–631.

(33) Klug, S. J.; Famulok, M. All You Wanted to Know about SELEX. Mol. Biol. Rep.

1994, 20, 97–107.

(34) Lehman, N.; Joyce, G. F. Evolution in Vitro of an RNA Enzyme with Altered Metal

Dependence. Nature 1993, 361, 182–185.

(35) Beaudry, A. A.; Joyce, G. F. Directed Evolution of an RNA Enzyme. Science 1992,

257, 635–641.

(36) Wu, W.; Henderson, L. E.; Copeland, T. D.; Gorelick, R. J.; Bosche, W. J.; Rein, A.;

Levin, J. G. Human Immunodeficiency Virus Type 1 Nucleocapsid Protein Reduces

Reverse Transcriptase Pausing at a Secondary Structure near the Murine Leukemia

Virus Polypurine Tract. J. Virol. 1996, 70, 7132–7142.

(37) Harrison, G. P.; Mayo, M. S.; Hunter, E.; Lever, A. M. Pausing of Reverse

Transcriptase on Retroviral RNA Templates Is Influenced by Secondary Structures

Both 5’ and 3' of the Catalytic Site. Nucleic Acids Res. 1998, 26, 3433–3442.

(38) Zeng, Y.; Pratumyot, Y.; Piao, X.; Bong, D. Discrete Assembly of Synthetic Peptide-

DNA Triplex Structures from Polyvalent Melamine-Thymine Bifacial Recognition. J.

Am. Chem. Soc. 2012, 134, 832–835.

(39) Piao, X.; Xia, X.; Bong, D. Bifacial Peptide Nucleic Acid Directs Cooperative Folding

and Assembly of Binary, Ternary, and Quaternary DNA Complexes. Biochemistry

2013, 52, 6313–6323.

(40) Xia, X.; Piao, X.; Bong, D. Bifacial Peptide Nucleic Acid as an Allosteric Switch for

Aptamer and Ribozyme Function. J. Am. Chem. Soc. 2014, 136, 7265–7268.

160

(41) Ruff, K. M.; Snyder, T. M.; Liu, D. R. Enhanced Functional Potential of Nucleic Acid

Aptamer Libraries Patterned to Increase Secondary Structure. J. Am. Chem. Soc.

2010, 132, 9453–9464.

(42) Xu, W.; Ellington, A. D. Anti-Peptide Aptamers Recognize Amino Acid Sequence

and Bind a Protein Epitope. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 7475–7480.

(43) Malim, M. H.; Hauber, J.; Le, S. Y.; Maizel, J. V.; Cullen, B. R. The HIV-1 Rev Trans-

Activator Acts through a Structured Target Sequence to Activate Nuclear Export of

Unspliced Viral mRNA. Nature 1989, 338, 254–257.

(44) Felber, B. K.; Hadzopoulou-Cladaras, M.; Cladaras, C.; Copeland, T.; Pavlakis, G.

N. Rev Protein of Human Immunodeficiency Virus Type 1 Affects the Stability and

Transport of the Viral mRNA. Proc. Natl. Acad. Sci. U. S. A. 1989, 86, 1495–1499.

(45) Emerman, M.; Vazeux, R.; Peden, K. The Rev Gene Product of the Human

Immunodeficiency Virus Affects Envelope-Specific RNA Localization. Cell 1989, 57,

1155–1165.

(46) Hammarskjöld, M. L.; Heimer, J.; Hammarskjöld, B.; Sangwan, I.; Albert, L.; Rekosh,

D. Regulation of Human Immunodeficiency Virus Env Expression by the Rev Gene

Product. J. Virol. 1989, 63, 1959–1966.

(47) Kjems, J.; Brown, M.; Chang, D. D.; Sharp, P. A. Structural Analysis of the

Interaction between the Human Immunodeficiency Virus Rev Protein and the Rev

Response Element. Proc. Natl. Acad. Sci. U. S. A. 1991, 88, 683–687.

(48) Kjems, J.; Calnan, B. J.; Frankel, A. D.; Sharp, P. A. Specific Binding of a Basic

Peptide from HIV-1 Rev. EMBO J. 1992, 11, 1119–1129.

(49) Battiste, J. L.; Tan, R.; Frankel, A. D.; Williamson, J. R. Binding of an HIV Rev

Peptide to Rev Responsive Element RNA Induces Formation of Purine-Purine Base

Pairs. Biochemistry 1994, 33, 2741–2747. 161

(50) Tan, R.; Frankel, A. D. Costabilization of Peptide and RNA Structure in an HIV Rev

Peptide-RRE Complex. Biochemistry 1994, 33, 14579–14585.

(51) Tan, R.; Chen, L.; Buettner, J. A.; Hudson, D.; Frankel, A. D. RNA Recognition by

an Isolated Alpha Helix. Cell 1993, 73, 1031–1040.

(52) Bartel, D. P.; Zapp, M. L.; Green, M. R.; Szostak, J. W. HIV-1 Rev Regulation

Involves Recognition of Non-Watson-Crick Base Pairs in Viral RNA. Cell 1991, 67,

529–536.

(53) Giver, L.; Bartel, D.; Zapp, M.; Pawul, A.; Green, M.; Ellington, A. D. Selective

Optimization of the Rev-Binding Element of HIV-1. Nucleic Acids Res. 1993, 21,

5509–5516.

(54) Ye, X.; Gorin, A.; Frederick, R.; Hu, W.; Majumdar, A.; Xu, W.; McLendon, G.;

Ellington, A.; Patel, D. J. RNA Architecture Dictates the Conformations of a Bound

Peptide. Chem. Biol. 1999, 6, 657–669.

(55) Harrison, R. S.; Shepherd, N. E.; Hoang, H. N.; Ruiz-Gómez, G.; Hill, T. A.; Driver,

R. W.; Desai, V. S.; Young, P. R.; Abbenante, G.; Fairlie, D. P. Downsizing Human,

Bacterial, and Viral Proteins to Short Water-Stable Alpha Helices That Maintain

Biological Potency. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 11686–11691.

(56) Lau, Y. H.; de Andrade, P.; Wu, Y.; Spring, D. R. Peptide Stapling Techniques

Based on Different Macrocyclisation Chemistries. Chem. Soc. Rev. 2014, 44, 91–

102.

(57) Blackwell, H. E.; Grubbs, R. H. Highly Efficient Synthesis of Covalently Cross-Linked

Peptide Helices by Ring-Closing Metathesis. Angew. Chem. Int. Ed. 1998, 37,

3281–3284.

162

(58) Christian E. Schafmeister; Julia Po; Verdine*, G. L. An All-Hydrocarbon Cross-

Linking System for Enhancing the Helicity and Metabolic Stability of Peptides. J. Am.

Chem. Soc. 2000, 122, 5891–5892.

(59) Felix, A. M.; Heimer, E. P.; Wang, C.-T.; Lambros, T. J.; Fournier, A.; Mowles, T. F.;

Maines, S.; Campbell, R. M.; Wegrzynski, B. B.; Toome, V.; et al. Synthesis,

Biological Activity and Conformational Analysis of Cyclic GRF Analogs. Int. J. Pept.

Protein Res. 1988, 32, 441–454.

(60) Taylor, J. W. The Synthesis and Study of Side-Chain Lactam-Bridged Peptides.

Biopolymers 2002, 66, 49–75.

(61) Shepherd, N. E.; Hoang, H. N.; Abbenante, G.; Fairlie, D. P. Single Turn Peptide

Alpha Helices with Exceptional Stability in Water. J. Am. Chem. Soc. 2005, 127,

2974–2983.

(62) Scrima, M.; Le Chevalier-Isaad, A.; Rovero, P.; Papini, A. M.; Chorev, M.; D’Ursi, A.

M. CuI-Catalyzed Azide–Alkyne Intramolecular I-to-(i+4) Side-Chain-to-Side-Chain

Cyclization Promotes the Formation of Helix-Like Secondary Structures. Eur. J. Org.

Chem. 2010, 2010, 446–457.

(63) Kawamoto, S. A.; Coleska, A.; Ran, X.; Yi, H.; Yang, C.-Y.; Wang, S. Design of

Triazole-Stapled BCL9 α-Helical Peptides to Target the β-catenin/B-Cell

CLL/lymphoma 9 (BCL9) Protein-Protein Interaction. J. Med. Chem. 2012, 55,

1137–1146.

(64) Torres, O.; Yüksel, D.; Bernardina, M.; Kumar, K.; Bong, D. Peptide Tertiary

Structure Nucleation by Side-Chain Crosslinking with Metal Complexation and

Double “Click” Cycloaddition. Chembiochem 2008, 9, 1701–1705.

(65) Lau, Y. H.; de Andrade, P.; Quah, S.-T.; Rossmann, M.; Laraia, L.; Sköld, N.; Sum,

T. J.; Rowling, P. J. E.; Joseph, T. L.; Verma, C.; et al. Functionalised Staple 163

Linkages for Modulating the Cellular Activity of Stapled Peptides. Chem. Sci. 2014,

5, 1804–1809.

(66) Brunel, F. M.; Dawson, P. E. Synthesis of Constrained Helical Peptides by Thioether

Ligation: Application to Analogs of gp41. Chem. Commun. 2005, 2552–2554.

(67) Verdine, G. L.; Hilinski, G. J. Stapled Peptides for Intracellular Drug Targets.

Methods Enzymol. 2012, 503, 3–33.

(68) Walensky, L. D.; Bird, G. H. Hydrocarbon-Stapled Peptides: Principles, Practice,

and Progress. J. Med. Chem. 2014, 57, 6275–6288.

(69) Moellering, R. E.; Cornejo, M.; Davis, T. N.; Del Bianco, C.; Aster, J. C.; Blacklow,

S. C.; Kung, A. L.; Gilliland, D. G.; Verdine, G. L.; Bradner, J. E. Direct Inhibition of

the NOTCH Transcription Factor Complex. Nature 2009, 462, 182–188.

(70) Walensky, L. D.; Kung, A. L.; Escher, I.; Malia, T. J.; Barbuto, S.; Wright, R. D.;

Wagner, G.; Verdine, G. L.; Korsmeyer, S. J. Activation of Apoptosis in Vivo by a

Hydrocarbon-Stapled BH3 Helix. Science 2004, 305, 1466–1470.

(71) Mills, N. L.; Daugherty, M. D.; Frankel, A. D.; Guy, R. K. An Alpha-Helical

Peptidomimetic Inhibitor of the HIV-1 Rev-RRE Interaction. J. Am. Chem. Soc. 2006,

128, 3496–3497.

(72) Battiste, J. L.; Mao, H.; Rao, N. S.; Tan, R.; Muhandiram, D. R.; Kay, L. E.; Frankel,

A. D.; Williamson, J. R. Alpha Helix-RNA Major Groove Recognition in an HIV-1 Rev

Peptide-RRE RNA Complex. Science 1996, 273, 1547–1551.

(73) Schütze, T.; Wilhelm, B.; Greiner, N.; Braun, H.; Peter, F.; Mörl, M.; Erdmann, V. A.;

Lehrach, H.; Konthur, Z.; Menger, M.; et al. Probing the SELEX Process with next-

Generation Sequencing. PLoS One 2011, 6, e29604.

164

(74) Ogawa, N.; Biggin, M. D. High-Throughput SELEX Determination of DNA

Sequences Bound by Transcription Factors in Vitro. Methods Mol. Biol. 2012, 786,

51–63.

(75) Harrison, R. S.; Shepherd, N. E.; Hoang, H. N.; Beyer, R. L.; Ruiz-Gómez, G.; Kelso,

M. J.; Mei Kok, W.; Hill, T. A.; Abbenante, G.; Fairlie, D. P. Helical Cyclic

Pentapeptides Constrain HIV-1 Rev Peptide for Enhanced RNA Binding.

Tetrahedron 2014, 70, 7645–7650.

(76) Martin, M. Cutadapt Removes Adapter Sequences from High-Throughput

Sequencing Reads. EMBnet.journal 2011, 17, 10–12.

(77) Alam, K. K.; Chang, J. L.; Burke, D. H. FASTAptamer: A Bioinformatic Toolkit for

High-Throughput Sequence Analysis of Combinatorial Selections. Mol. Ther.

Nucleic Acids 2015, 4, e230.

(78) Daugherty, M. D.; D’Orso, I.; Frankel, A. D. A Solution to Limited Genomic Capacity:

Using Adaptable Binding Surfaces to Assemble the Functional HIV Rev Oligomer

on RNA. Mol. Cell 2008, 31, 824–834.

(79) Ye, X.; Gorin, A.; Ellington, A. D.; Patel, D. J. Deep Penetration of an Alpha-Helix

into a Widened RNA Major Groove in the HIV-1 Rev Peptide-RNA Aptamer Complex.

Nat. Struct. Biol. 1996, 3, 1026–1033.

(80) Leclerc, F.; Cedergren, R.; Ellington, A. D. A 3-DIMENSIONAL MODEL OF THE

REV-BINDING ELEMENT OF HIV-1 DERIVED FROM ANALYSES OF APTAMERS.

Nat. Struct. Biol. 1994, 1, 293–300.

(81) Link, A. J.; Vink, M. K. S.; Tirrell, D. A. Presentation and Detection of Azide

Functionality in Bacterial Cell Surface Proteins. J. Am. Chem. Soc. 2004, 126,

10598–10602.

165

(82) Lau, Y. H.; Wu, Y.; de Andrade, P.; Galloway, W. R. J. D.; Spring, D. R. A Two-

Component “Double-Click” Approach to Peptide Stapling. Nat. Protoc. 2015, 10,

585–594.

(83) Shao, K.; Ding, W.; Wang, F.; Li, H.; Ma, D.; Wang, H. Emulsion PCR: A High

Efficient Way of PCR Amplication of Random DNA Libraries in Aptamer Selection,

Plos ONE 2011, 6, e24910.

(84) Musheev, M. U.; Krylov, S. N. Selection of Aptamers by systematic evolution of

Ligands by Expotential Enrichment: Addressing the Polymerase Chain Reaction

Issue, Analytica Chimica Acta 2006, 564, 91-96.

166

Appendix A: Analysis of High-Throughput Sequencing Raw Data

High-throughput sequencing is a powerful tool to decipher in vitro selection and it generally yields millions of reads that require programs to process and analyze these data. This in- house protocol describes how the sequencing results of our unique library is processed and analyzed. Two command-in-line softwares, Cutadapt1 and Fastaptamer2 are used for this purpose. Commands are typed in Terminal in MacPro. To elucidate how the command affects the sequence, two representative sequences from PE-100bp sequencing on

Illumina HiSeq 2500 are used as examples. TruSeq Universal Adapters were used by

OSU Comprehensive Cancer Center genomics.

In high-throughput sequencing, both sense and antisense strands are sequenced, so two distinctive types of sequences are found in the result. One has two T-tracts and the other has two A-tracts. There are also some sequences with only one or no T/A-tract. The sequences with two T-tracts are of interest and the other sequences are discarded in the process.

Table A.1 Two representative sequences

5’-TAATACGACTCACTATAGGGTAGGATCCGCTTTTTTTTTTACGGGGATAAAACA

GCGATCAATGCTTTTTTTTTTGCTGAATTCACATACGACGAAGATC-3’

5’-TCGTCGTATGTGAATTCAGCAAAAAAAAAACATTGGTAATCTGCCGGTGTGAAC

AAAAAAAAAAGCGGATCCTACCCTATAGTGAGTCGTATTAAGATCG-3’

167

A.1 Installation

Please follow the instructions in the handbook of the two programs to install software. It is easier to use OS X system.

A.2 Download raw data

The raw data are downloaded from Illumina basespace or vendor’s FTP. The raw data is

FASTQ file. Although paired-end sequencing was used to yield read 1 and read 2, only read 1 is processed and analyzed for the purpose of simplicity. This equals to single-read sequencing mode.

The length of our library is 95-bp, however, the sequencing result is 100-bp which includes five nucleotides from adenylation (adenylation adds one nucleotide, adenine) and TruSeq

Universal adapter. Please note there may be some length mutations, so the actual sequence length may not be exact 95-bp. If 150-bp is used in high-throughput sequencing, the 55-bp from adenylation and the adapters vendor used are included.

The screen shot of each command is listed to help you understand the process and one can also find more information on the screen shots, for example, how many sequences are processed.

A.3 Removal of 3’-constant region on T-tract sequences

What we are really interested in is the N25 random region between two T10 tracts. To remove 3’-constant flanking region from the desired sequences and discard the unwanted sequences, the basic command-line typed in Terminal is:

168

Figure A.1 Removal of 3’ constant flanking region and adapter

10-700894_S58_R1_001.fastq is the name of your input raw data file. 10_5removed.fastq is the name of new output file, and one can design any other name as long as it is convenient for you. After this command, the 3’ constant flanking regions are removed for all sequences bearing two T10 tracts. All untrimmed sequences are discarded, and these sequences include A10-tract sequences and wrong sequences.

Table A.2 Representative sequence

5’-TAATACGACTCACTATAGGGTAGGATCCGCTTTTTTTTTTACGGGGATAAAACA

GCGATCAATGCTTTTTTTTTT-3’

169

The trimmed sequences now are 75-bp in length. The command is 10% error-tolerated by default, although you can set a desired value. Error-tolerance is useful to remove adapters or constant flanking region that have a certain amount of mutations.

A.4 Removal of 5’ constant region

Similarly, 5’-constant flanking region can be removed, the basic command is:

Figure A.2 Removal of 5’ constant flanking region

This command will cut 5’-constant region and discard any untrimmed sequences.

Table A.3 Representative sequence change

5’-TTTTTTTTTTACGGGGATAAAACAGCGATCAATGCTTTTTTTTTT-3’

170

The sequences after last command is good enough to analyze, but the two T tracts can be removed to provide an even simpler sequence. It is always encouraged to check the output file after processing to ensure the results are what you expected. These FASTQ file can be opened by TEXT.

A.5 Removal of two T10 tracts

Figure A.3 Removal of one T-tract

This command will cut T10 region from either end once, so one will see two types of sequences, T10N25 or N25T10. This command can be done once more to remove T10 completely and filter off sequences that are longer than 27 or shorter than 23.

171

Figure A.4 Removal of the other T-tract

One to two extra T are found to be added to T10 tracts, so one can cut T15 instead of T10 to remove T tracts completely. N25 region will not be affected if it does not start with T. If the sequences start with T or have several T close to 5’, these Ts will also be removed.

This may be a problem.

Table A.4 N25 region of interest

5’-ACGGGGATAAAACAGCGATCAATGC-3’

The way to trim raw data is flexible, for example, one can remove A-tract sequence prior to trim 3’-constant flanking region. A good way to check the effect of a command is to open the output file and make sure the sequences are trimmed as you expected.

172

A.6 Analysis of trimmed sequences

Fastaptamer is used to analyze the trimmed sequences. FASTQ file needs to be converted to FASTA file. With this command, one is able to know the number of the total sequences in this pool.

Figure A.5 Converting from fastq to fasta file

FASTA file is ready to be used for clustering which will generate clusters of similar sequences. Levenshtein edit distance is specified by the command of -d 7, other numbers can be set depending on how many nucleobase mutations you want a cluster can tolerate.

Clustering is very time-consuming and tens of hours may not be sufficient. Thus, it is necessary to set a filter that only clusters the sequences showing over certain times. -f number is how filter is set. Different filters can be used, such as 2, 10, 50 and 100. The

173 tsv file generated summarizes the total reads of a cluster and the number of sequences in this cluster. And individual sequences can be found in corresponding FASTA file.

Figure A.6 Clustering the aptamers

Fastaptamer_enrich command can be used to compare the reads of the same aptamer in different rounds of selection, indicating the progress of in vitro selection. Assuming fastaptamer_count and fastaptamer_cluster have been done on the files from all rounds of selection. For example, you want to see the enrichment from round 9 to round 10.

174

Figure A.7 Comparing the enrichment between different rounds

In the generated tsv file, you can easily find the reads of sequence in round 9 and the reads of corresponding cluster in round10, as well as an enrichment fold. This data is helpful to check the enrichment of individual sequences or certain clusters.

A.7 References

(1) Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing

reads, EMBnet.journal, 2011, 17.1, 10-12.

(2) Alam, K. K.; Chang, J. L.; Burke, D. H. FASTApamer: a bioinformatic toolkit for high-

throughput sequence analysis of combinatorial selections. Molecular Therapy-

Nucleic Acids, 2015, 4, e230.

175

References

Chapter 1

(1) Crick, F. H. C.; Watson, J. D. The Complementary Structure of Deoxyribonucleic

Acid. Proceedings of the Royal Society of London A: Mathematical, Physical and

Engineering Sciences 1954, 223, 80–96.

(2) Wilkins, M. H. F.; Stokes, A. R.; Wilson, H. R. Molecular Structure of Deoxypentose

Nucleic Acids. Nature 1953, 171, 738–740.

(3) Rana, T. M. Illuminating the Silence: Understanding the Structure and Function of

Small RNAs. Nat. Rev. Mol. Cell Biol. 2007, 8, 23–36.

(4) Watson, J. D.; Crick, F. H. C.; Others. Molecular Structure of Nucleic Acids. Nature

1953, 171, 737–738.

(5) Wang, J. C. Helical Repeat of DNA in Solution. Proc. Natl. Acad. Sci. U. S. A. 1979,

76, 200–203.

(6) Dervan, P. B. Molecular Recognition of DNA by Small Molecules. Bioorg. Med.

Chem. 2001, 9, 2215–2235.

(7) White, S.; Szewczyk, J. W.; Turner, J. M.; Baird, E. E.; Dervan, P. B. Recognition of

the Four Watson-Crick Base Pairs in the DNA Minor Groove by Synthetic Ligands.

Nature 1998, 391, 468–471.

176

(8) Renneberg, D.; Dervan, P. B. Imidazopyridine / Pyrrole and Hydroxybenzimidazole

/ pyrrole Pairs for DNA Minor Groove Recognition. J. Am. Chem. Soc. 2003, 125,

5707–5716.

(9) Dervan, P. B.; Edelson, B. S. Recognition of the DNA Minor Groove by Pyrrole-

Imidazole Polyamides. Curr. Opin. Struct. Biol. 2003, 13, 284–299.

(10) François, J. C.; Saison-Behmoaras, T.; Hélène, C. Sequence-Specific Recognition

of the Major Groove of DNA by Oligodeoxynucleotides via Triple Helix Formation.

Footprinting Studies. Nucleic Acids Res. 1988, 16, 11431–11440.

(11) Frank-Kamenetskii, M. D.; Mirkin, S. M. Triplex DNA Structures. Annu. Rev. Biochem.

1995, 64, 65–95.

(12) Duca, M.; Vekhoff, P.; Oussedik, K.; Halby, L.; Arimondo, P. B. The Triple Helix: 50

Years Later, the Outcome. Nucleic Acids Res. 2008, 36, 5123–5138.

(13) Arimondo, P. B.; Barcelo, F.; Sun, J. S.; Maurizot, J. C.; Garestier, T.; Hélène, C.

Triple Helix Formation by (G,A)-Containing Oligonucleotides: Asymmetric Sequence

Effect. Biochemistry 1998, 37, 16627–16635.

(14) Beal, P. A.; Dervan, P. B. Second Structural Motif for Recognition of DNA by

Oligonucleotide-Directed Triple-Helix Formation. Science 1991, 251, 1360–1363.

(15) Song, G.; Ren, J. Recognition and Regulation of Unique Nucleic Acid Structures by

Small Molecules. Chem. Commun. 2010, 46, 7283–7294.

(16) Lyamichev, V. I.; Mirkin, S. M.; Frank-Kamenetskii, M. D.; Cantor, C. R. A Stable

Complex between Homopyrimidine Oligomers and the Homologous Regions of

Duplex DNAs. Nucleic Acids Res. 1988, 16, 2165–2178.

(17) Blume, S. W.; Lebowitz, J.; Zacharias, W.; Guarcello, V.; Mayfield, C. A.; Ebbinghaus,

S. W.; Bates, P.; Jones, D. E., Jr; Trent, J.; Vigneswaran, N.; Miller, D. M. The

Integral Divalent Cation within the Intermolecular Purine*purine. Pyrimidine 177

Structure: A Variable Determinant of the Potential for and Characteristics of the

Triple Helical Association. Nucleic Acids Res. 1999, 27, 695–702.

(18) Thomas, T.; Thomas, T. J. Selectivity of Polyamines in Triplex DNA Stabilization.

Biochemistry 1993, 32, 14068–14074.

(19) Takabatake, T.; Asada, K.; Uchimura, Y.; Ohdate, M.; Kusukawa, N. The Use of

Purine-Rich Oligonucleotides in Triplex-Mediated DNA Isolation and Generation of

Unidirectional Deletions. Nucleic Acids Res. 1992, 20, 5853–5854.

(20) Pilch, D. S.; Levenson, C.; Shafer, R. H. Structural Analysis of the (dA) 10.2 (dT) 10

Triple Helix. Proceedings of the National Academy of Sciences 1990, 87, 1942–

1946.

(21) Goñi, J. R.; de la Cruz, X.; Orozco, M. Triplex‐forming Oligonucleotide Target

Sequences in the Human Genome. Nucleic Acids Res. 2004, 32, 354–360.

(22) Goñi, J. R.; Vaquerizas, J. M.; Dopazo, J.; Orozco, M. Exploring the Reasons for the

Large Density of Triplex-Forming Oligonucleotide Target Sequences in the Human

Regulatory Regions. BMC Genomics 2006, 7, 63.

(23) Durland, R. H.; Kessler, D. J.; Gunnell, S.; Duvic, M.; Pettitt, B. M.; Hogan, M. E.

Binding of Triple Helix Forming Oligonucleotides to Sites in Gene Promoters.

Biochemistry 1991, 30, 9246–9255.

(24) Soto, A. M.; Loo, J. Energetic Contributions for the Formation of TAT/TAT,

TAT/CGC+, and CGC+/CGC+ Base Triplet Stacks. J. Am. Chem. Soc. 2002, 124,

14355–14363.

(25) Bochman, M. L.; Paeschke, K.; Zakian, V. A. DNA Secondary Structures: Stability

and Function of G-Quadruplex Structures. Nat. Rev. Genet. 2012, 13, 770–780.

(26) Burge, S.; Parkinson, G. N.; Hazel, P.; Todd, A. K.; Neidle, S. Quadruplex DNA:

Sequence, Topology and Structure. Nucleic Acids Res. 2006, 34, 5402–5415. 178

(27) Todd, A. K.; Johnston, M.; Neidle, S. Highly Prevalent Putative Quadruplex

Sequence Motifs in Human DNA. Nucleic Acids Res. 2005, 33, 2901–2907.

(28) Huppert, J. L.; Balasubramanian, S. G-Quadruplexes in Promoters throughout the

Human Genome. Nucleic Acids Res. 2007, 35, 406–413.

(29) Paeschke, K.; Simonsson, T.; Postberg, J.; Rhodes, D.; Lipps, H. J. Telomere End-

Binding Proteins Control the Formation of G-Quadruplex DNA Structures in Vivo.

Nat. Struct. Mol. Biol. 2005, 12, 847–854.

(30) Tinoco, I., Jr; Bustamante, C. How RNA Folds. J. Mol. Biol. 1999, 293, 271–281.

(31) Tian, B.; Bevilacqua, P. C.; Diegelman-Parente, A.; Mathews, M. B. The Double-

Stranded-RNA-Binding Motif: Interference and Much More. Nat. Rev. Mol. Cell Biol.

2004, 5, 1013–1023.

(32) Salazar, M.; Fedoroff, O. Y.; Miller, J. M.; Ribeiro, N. S.; Reid, B. R. The DNA Strand

in DNA.cntdot.RNA Hybrid Duplexes Is Neither B-Form nor A-Form in Solution.

Biochemistry 1993, 32, 4207–4215.

(33) Theimer, C. A.; Blois, C. A.; Feigon, J. Structure of the Human Telomerase RNA

Pseudoknot Reveals Conserved Tertiary Interactions Essential for Function. Mol.

Cell 2005, 17, 671–682.

(34) Ishitani, R.; Nureki, O.; Nameki, N.; Okada, N.; Nishimura, S.; Yokoyama, S.

Alternative Tertiary Structure of tRNA for Recognition by a Posttranscriptional

Modification Enzyme. Cell 2003, 113, 383–394.

(35) Fedor, M. J.; Williamson, J. R. The Catalytic Diversity of RNAs. Nat. Rev. Mol. Cell

Biol. 2005, 6, 399–412.

(36) Michel, T.; Debart, F.; Heitz, F.; Vasseur, J.-J. Highly Stable DNA Triplexes Formed

with Cationic Phosphoramidate Pyrimidine Alpha-Oligonucleotides. Chembiochem

2005, 6, 1254–1262. 179

(37) Michel, T.; Martinand-Mari, C.; Debart, F.; Lebleu, B.; Robbins, I.; Vasseur, J.-J.

Cationic Phosphoramidate Alpha-Oligonucleotides Efficiently Target Single-

Stranded DNA and RNA and Inhibit Hepatitis C Virus IRES-Mediated Translation.

Nucleic Acids Res. 2003, 31, 5282–5290.

(38) Seidman, M. M.; Puri, N.; Majumdar, A.; Cuenoud, B.; Miller, P. S.; Alam, R. The

Development of Bioactive Triple Helix-Forming Oligonucleotides. Ann. N. Y. Acad.

Sci. 2005, 1058, 119–127.

(39) Petersen, M.; Wengel, J. LNA: A Versatile Tool for Therapeutics and Genomics.

Trends Biotechnol. 2003, 21, 74–81.

(40) Højland, T.; Kumar, S.; Babu, B. R.; Umemoto, T.; Albaek, N.; Sharma, P. K.; Nielsen,

P.; Wengel, J. LNA (locked Nucleic Acid) and Analogs as Triplex-Forming

Oligonucleotides. Org. Biomol. Chem. 2007, 5, 2375–2379.

(41) Brunet, E.; Alberti, P.; Perrouault, L.; Babu, R.; Wengel, J.; Giovannangeli, C.

Exploring Cellular Activity of Locked Nucleic Acid-Modified Triplex-Forming

Oligonucleotides and Defining Its Molecular Basis. J. Biol. Chem. 2005, 280, 20076–

20085.

(42) Sun, B.-W.; Babu, B. R.; Sørensen, M. D.; Zakrzewska, K.; Wengel, J.; Sun, J.-S.

Sequence and pH Effects of LNA-Containing Triple Helix-Forming Oligonucleotides:

Physical Chemistry, Biochemistry, and Modeling Studies. Biochemistry 2004, 43,

4160–4169.

(43) Aviñó, A.; Cubero, E.; González, C.; Eritja, R.; Orozco, M. Antiparallel Triple Helices.

Structural Characteristics and Stabilization by 8-Amino Derivatives. J. Am. Chem.

Soc. 2003, 125, 16127–16138.

180

(44) Xodo, L. E.; Manzini, G.; Quadrifoglio, F.; van der Marel, G. A.; van Boom, J. H.

Effect Pf 5-Methylcytosine on the Stability of Triple-Stranded DNA—a

Thermodynamic Study. Nucleic Acids Res. 1991, 19, 5625–5631.

(45) Strobel, S. A.; Dervan, P. B. Single-Site Enzymatic Cleavage of Yeast Genomic DNA

Mediated by Triple Helix Formation. Nature 1991, 350, 172–174.

(46) Lee, J. S.; Woodsworth, M. L.; Latimer, L. J.; Morgan, A. R. Poly(pyrimidine) .

Poly(purine) Synthetic DNAs Containing 5-Methylcytosine Form Stable Triplexes at

Neutral pH. Nucleic Acids Res. 1984, 12, 6603–6614.

(47) Spacková, N.; Cubero, E.; Sponer, J.; Orozco, M. Theoretical Study of the Guanine

--> 6-Thioguanine Substitution in Duplexes, Triplexes, and Tetraplexes. J. Am.

Chem. Soc. 2004, 126, 14642–14650.

(48) Malnuit, V.; Duca, M.; Benhida, R. Targeting DNA Base Pair Mismatch with Artificial

Nucleobases. Advances and Perspectives in Triple Helix Strategy. Org. Biomol.

Chem. 2011, 9, 326–336.

(49) Nielsen, P. E.; Egholm, M.; Buchardt, O. Peptide Nucleic Acid (PNA). A DNA Mimic

with a Peptide Backbone. Bioconjug. Chem. 1994, 5, 3–7.

(50) Nielsen, P. E.; Haaima, G. Peptide Nucleic Acid (PNA). A DNA Mimic with a

Pseudopeptide Backbone. Chem. Soc. Rev. 1997, 26, 73–78.

(51) Nielsen, P. E. Peptide Nucleic Acid. A Molecule with Two Identities. Acc. Chem. Res.

1999, 32, 624–630.

(52) Egholm, M.; Buchardt, O.; Christensen, L.; Behrens, C.; Freier, S. M.; Driver, D. A.;

Berg, R. H.; Kim, S. K.; Norden, B.; Nielsen, P. E. PNA Hybridizes to Complementary

Oligonucleotides Obeying the Watson-Crick Hydrogen-Bonding Rules. Nature 1993,

365, 566–568.

181

(53) Larsen, H. J.; Nielsen, P. E. Transcription-Mediated Binding of Peptide Nucleic Acid

(PNA) to Double-Stranded DNA: Sequence-Specific Suicide Transcription. Nucleic

Acids Res. 1996, 24, 458–463.

(54) Elayadi, A. N.; Corey, D. R. Application of PNA and LNA Oligomers to Chemotherapy.

Curr. Opin. Investig. Drugs 2001, 2, 558–561.

(55) Bentin, T.; Larsen, H. J.; Nielsen, P. E. Combined Triplex/duplex Invasion of Double-

Stranded DNA by “Tail-Clamp” Peptide Nucleic Acid. Biochemistry 2003, 42,

13987–13995.

(56) Hansen, M. E.; Bentin, T.; Nielsen, P. E. High-Affinity Triplex Targeting of Double

Stranded DNA Using Chemically Modified Peptide Nucleic Acid Oligomers. Nucleic

Acids Res. 2009, 37, 4498–4507.

(57) Nielsen, P. E.; Egholm, M.; Berg, R. H.; Buchardt, O. Sequence-Selective

Recognition of DNA by Strand Displacement with a Thymine-Substituted Polyamide.

Science 1991, 254, 1497–1500.

(58) Betts, L.; Josey, J. A.; Veal, J. M.; Jordan, S. R. A Nucleic Acid Triple Helix Formed

by a Peptide Nucleic Acid-DNA Complex. Science 1995, 270, 1838–1841.

(59) Norton, J. C.; Piatyszek, M. A.; Wright, W. E.; Shay, J. W.; Corey, D. R. Inhibition of

Human Telomerase Activity by Peptide Nucleic Acids. Nat. Biotechnol. 1996, 14,

615–619.

(60) Demidov, V.; Frank-Kamenetskii, M. D.; Egholm, M.; Buchardt, O.; Nielsen, P. E.

Sequence Selective Double Strand DNA Cleavage by Peptide Nucleic Acid (PNA)

Targeting Using Nuclease S1. Nucleic Acids Res. 1993, 21, 2103–2107.

(61) Ray, A.; Nordén, B. Peptide Nucleic Acid (PNA): Its Medical and Biotechnical

Applications and Promise for the Future. The FASEB Journal 2000, 14, 1041–1060.

182

(62) Boffa, L. C.; Morris, P. L.; Carpaneto, E. M.; Louissaint, M.; Allfrey, V. G. Invasion

of the CAG Triplet Repeats by a Complementary Peptide Nucleic Acid Inhibits

Transcription of the Androgen Receptor and TATA-Binding Protein Genes and

Correlates with Refolding of an Active Nucleosome Containing a Unique AR Gene

Sequence. J. Biol. Chem. 1996, 271, 13228–13233.

(63) Knudsen, H.; Nielsen, P. E. Antisense Properties of Duplex- and Triplex-Forming

PNAs. Nucleic Acids Res. 1996, 24, 494–500.

(64) Cogoi, S.; Codognotto, A.; Rapozzi, V.; Xodo, L. E. Antigene Property of PNA

Conjugated to the Nuclear Localization Signal Peptide. Nucleosides Nucleotides

Nucleic Acids 2005, 24, 971–974.

(65) Cutrona, G.; Carpaneto, E. M.; Ulivi, M.; Roncella, S.; Landt, O.; Ferrarini, M.; Boffa,

L. C. Effects in Live Cells of a c-Myc Anti-Gene PNA Linked to a Nuclear Localization

Signal. Nat. Biotechnol. 2000, 18, 300–303.

(66) Tonelli, R.; Purgato, S.; Camerin, C.; Fronza, R.; Bologna, F.; Alboresi, S.; Franzoni,

M.; Corradini, R.; Sforza, S.; Faccini, A.; et al. Anti-Gene Peptide Nucleic Acid

Specifically Inhibits MYCN Expression in Human Neuroblastoma Cells Leading to

Cell Growth Inhibition and Apoptosis. Mol. Cancer Ther. 2005, 4, 779–786.

(67) Branda, N.; Kurz, G.; Lehn, J.-M. JANUS WEDGES: A New Approach towards

Nucleobase-Pair Recognition. Chem. Commun. 1996, 2443–2444.

(68) Arambula, J. F.; Ramisetty, S. R.; Baranger, A. M.; Zimmerman, S. C. A Simple

Ligand That Selectively Targets CUG Trinucleotide Repeats and Inhibits MBNL

Protein Binding. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 16068–16073.

(69) Jahromi, A. H.; Honda, M.; Zimmerman, S. C.; Spies, M. Single-Molecule Study of

the CUG Repeat--MBNL1 Interaction and Its Inhibition by Small Molecules. Nucleic

Acids Res. 2013, 41, 6687–6697. 183

(70) Nguyen, L.; Lee, J.; Wong, C.-H.; Zimmerman, S. C. Small Molecules That Target

the Toxic RNA in Myotonic Dystrophy Type 2. ChemMedChem 2014, 9, 2455–2462.

(71) Artigas, G.; Marchán, V. Synthesis of Janus Compounds for the Recognition of G-

U Mismatched Nucleobase Pairs. J. Org. Chem. 2013, 78, 10666–10677.

(72) Shin, D.; Tor, Y. Bifacial Nucleoside as a Surrogate for Both T and A in Duplex DNA.

J. Am. Chem. Soc. 2011, 133, 6926–6929.

(73) Largy, E.; Liu, W.; Hasan, A.; Perrin, D. M. Base-Pairing Behavior of a Carbocyclic

Janus-AT Nucleoside Analogue Capable of Recognizing A and T within a DNA

Duplex. Chembiochem 2013, 14, 2199–2208.

(74) Chen, D.; Meena, M.; Sharma, S. K.; McLaughlin, L. W. Formation and Stability of

a Janus-Wedge Type of DNA Triplex. J. Am. Chem. Soc. 2004, 126, 70–71.

(75) Chen, H.; McLaughlin, L. W. A Janus-Wedge DNA Triplex with A-W1-T and G-W2-

C Base Triplets. J. Am. Chem. Soc. 2008, 130, 13190–13191.

(76) Zeng, Y.; Pratumyot, Y.; Piao, X.; Bong, D. Discrete Assembly of Synthetic Peptide-

DNA Triplex Structures from Polyvalent Melamine-Thymine Bifacial Recognition. J.

Am. Chem. Soc. 2012, 134, 832–835.

(77) Mao, J.; Bong, D. Synthesis of DNA-Binding Peptoids. Synlett 2015, 26, 1581–1585.

(78) Zhou, Z.; Xia, X.; Bong, D. Synthetic Polymer Hybridization with DNA and RNA

Directs Nanoparticle Loading, Silencing Delivery, and Aptamer Function. J. Am.

Chem. Soc. 2015, 137, 8920–8923.

(79) Xia, X.; Piao, X.; Fredrick, K.; Bong, D. Bifacial PNA Complexation Inhibits Enzymatic

Access to DNA and RNA. Chembiochem 2014, 15, 31–36.

(80) Xia, X.; Piao, X.; Bong, D. Bifacial Peptide Nucleic Acid as an Allosteric Switch for

Aptamer and Ribozyme Function. J. Am. Chem. Soc. 2014, 136, 7265–7268.

184

(81) Vysabhattar, R.; Ganesh, K. N. Cyanuryl Peptide Nucleic Acid: Synthesis and DNA

Complexation Properties. Tetrahedron Lett. 2008, 49, 1314–1318.

(82) Mittapalli, G. K.; Reddy, K. R.; Xiong, H.; Munoz, O.; Han, B.; De Riccardis, F.;

Krishnamurthy, R.; Eschenmoser, A. Mapping the Landscape of Potentially

Primordial Informational Oligomers: Oligodipeptides and Oligodipeptoids Tagged

with Triazines as Recognition Elements. Angew. Chem. Int. Ed. 2007, 46, 2470–

2477.

(83) Piao, X.; Xia, X.; Bong, D. Bifacial Peptide Nucleic Acid Directs Cooperative Folding

and Assembly of Binary, Ternary, and Quaternary DNA Complexes. Biochemistry

2013, 52, 6313–6323.

(84) Hysell, M.; Siegel, J. S.; Tor, Y. Synthesis and Stability of Exocyclic Triazine

Nucleosides. Org. Biomol. Chem. 2005, 3, 2946–2952.

(85) Kozlov, I. A.; Pitsch, S.; Orgel, L. E. Oligomerization of Activated D- and L-Guanosine

Mononucleotides on Templates Containing D- and L-Deoxycytidylate Residues.

Proceedings of the National Academy of Sciences 1998, 95, 13448–13452.

(86) Gao, K.; Orgel, L. E. Polyphosphorylation and Non-Enzymatic Template-Directed

Ligation of Oligonucleotides. Orig. Life Evol. Biosph. 2000, 30, 45–51.

(87) Rosenbaum, D. M.; Liu, D. R. Efficient and Sequence-Specific DNA-Templated

Polymerization of Peptide Nucleic Acid Aldehydes. J. Am. Chem. Soc. 2003, 125,

13924–13925.

(88) Heemstra, J. M.; Liu, D. R. Templated Synthesis of Peptide Nucleic Acids via

Sequence-Selective Base-Filling Reactions. J. Am. Chem. Soc. 2009, 131, 11347–

11349.

(89) Roloff, A.; Seitz, O. Bioorthogonal Reactions Challenged: DNA Templated Native

Chemical Ligation during PCR. Chem. Sci. 2012, 4, 432–436. 185

(90) Gartner, Z. J.; Tse, B. N.; Grubina, R.; Doyon, J. B.; Snyder, T. M.; Liu, D. R. DNA-

Templated Organic Synthesis and Selection of a Library of Macrocycles. Science

2004, 305, 1601–1605.

(91) Gartner, Z. J.; Liu, D. R. The Generality of DNA-Templated Synthesis as a Basis for

Evolving Non-Natural Small Molecules. J. Am. Chem. Soc. 2001, 123, 6961–6963.

(92) Erben, A.; Grossmann, T. N.; Seitz, O. DNA-Triggered Synthesis and Bioactivity of

Proapoptotic Peptides. Angew. Chem. Int. Ed. 2011, 50, 2828–2832.

(93) McKee, M. L.; Evans, A. C.; Gerrard, S. R.; O’Reilly, R. K.; Turberfield, A. J.; Stulz,

E. Peptidomimetic Bond Formation by DNA-Templated Acyl Transfer. Org. Biomol.

Chem. 2011, 9, 1661–1666.

(94) Grossmann, T. N.; Seitz, O. DNA-Catalyzed Transfer of a Reporter Group. J. Am.

Chem. Soc. 2006, 128, 15596–15597.

(95) Joyce, G. F. Nonenzymatic Template-Directed Synthesis of Informational

Macromolecules. Cold Spring Harb. Symp. Quant. Biol. 1987, 52, 41–51.

(96) Bolli, M.; Micura, R.; Eschenmoser, A. Pyranosyl-RNA: Chiroselective Self-Assembly

of Base Sequences by Ligative Oligomerization of Tetranucleotide-2’,3'-

Cyclophosphates (with a Commentary Concerning the Origin of Biomolecular

Homochirality). Chem. Biol. 1997, 4, 309–320.

(97) Xu, Y.; Karalkar, N. B.; Kool, E. T. Nonenzymatic Autoligation in Direct Three-Color

Detection of RNA and DNA Point Mutations. Nat. Biotechnol. 2001, 19, 148–152.

(98) von Kiedrowski, G. A Self-Replicating Hexadeoxynucleotide. Angew. Chem. Int. Ed

Engl. 1986, 25, 932–935.

(99) Sievers, D.; von Kiedrowski, G. Self-Replication of Complementary Nucleotide-

Based Oligomers. Nature 1994, 369, 221–224.

186

(100) Böhler, C.; Nielsen, P. E.; Orgel, L. E. Template Switching between PNA and RNA

Oligonucleotides. Nature 1995, 376, 578–581.

(101) Schmidt, J. G.; Nielsen, P. E.; Orgel, L. E. Information Transfer from Peptide Nucleic

Acids to RNA by Template-Directed Syntheses. Nucleic Acids Res. 1997, 25, 4797–

4802.

(102) Schmidt, J. G.; Christensen, L.; Nielsen, P. E.; Orgel, L. E. Information Transfer from

DNA to Peptide Nucleic Acids by Template-Directed Syntheses. Nucleic Acids Res.

1997, 25, 4792–4796.

(103) Rohatgi, R.; Bartel, D. P.; Szostak, J. W. Kinetic and Mechanistic Analysis of

Nonenzymatic, Template-Directed Oligoribonucleotide Ligation. J. Am. Chem. Soc.

1996, 118, 3332–3339.

(104) Rajat Rohatgi; David P. Bartel, †.; Szostak*, J. W. Nonenzymatic, Template-

Directed Ligation of Oligoribonucleotides Is Highly Regioselective for the Formation

of 3‘−5‘ Phosphodiester Bonds. J. Am. Chem. Soc. 1996, 118, 3340–3344.

(105) Kanan, M. W.; Rozenman, M. M.; Sakurai, K.; Snyder, T. M.; Liu, D. R. Reaction

Discovery Enabled by DNA-Templated Synthesis and in Vitro Selection. Nature

2004, 431, 545–549.

(106) Li, T.; Nicolaou, K. C. Chemical Self-Replication of Palindromic Duplex DNA. Nature

1994, 369, 218–221.

(107) Li, T.; Weinstein, D. S.; Nicolaou, K. The Chemical End-Ligation of Homopyrimidine

Oligodeoxyribonucleotides within a DNA Triple Helix. Chem. Biol. 1997, 4, 209–214.

(108) Meguellati, K.; Koripelly, G.; Ladame, S. DNA-Templated Synthesis of Trimethine

Cyanine Dyes: A Versatile Fluorogenic Reaction for Sensing G-Quadruplex

Formation. Angew. Chem. Int. Ed. 2010, 49, 2738–2742.

187

(109) Izgu, E. C.; Fahrenbach, A. C.; Zhang, N.; Li, L.; Zhang, W.; Larsen, A. T.; Blain, J.

C.; Szostak, J. W. Uncovering the Thermodynamics of Monomer Binding for RNA

Replication. J. Am. Chem. Soc. 2015, 137, 6373–6382.

(110) Bruick, R. K.; Dawson, P. E.; Kent, S. B.; Usman, N.; Joyce, G. F. Template-Directed

Ligation of Peptides to Oligonucleotides. Chem. Biol. 1996, 3, 49–56.

(111) Ficht, S.; Dose, C.; Seitz, O. As Fast and Selective as Enzymatic Ligations:

Unpaired Nucleobases Increase the Selectivity of DNA-Controlled Native Chemical

PNA Ligation. Chembiochem 2005, 6, 2098–2103.

(112) Xu, Y.; Kool, E. T. Chemical and Enzymatic Properties of Bridging 5’-S-

Phosphorothioester Linkages in DNA. Nucleic Acids Res. 1998, 26, 3159–3164.

(113) Izgu, E. C.; Oh, S. S.; Szostak, J. W. Synthesis of Activated 3′-Amino-3′-Deoxy-2-

Thio-Thymidine, a Superior Substrate for the Nonenzymatic Copying of Nucleic Acid

Templates. Chem. Commun. 2016, 52, 3684–3686.

(114) Goodwin, J. T.; Lynn, D. G. Template-Directed Synthesis: Use of a Reversible

Reaction. J. Am. Chem. Soc. 1992, 114, 9197–9198.

(115) Xiaoyu Li; Zheng-Yun J. Zhan; Rachel Knipe; Lynn*, D. G. DNA-Catalyzed

Polymerization. J. Am. Chem. Soc. 2002, 124, 746–747.

(116) McKee, M. L.; Milnes, P. J.; Bath, J.; Stulz, E.; O’Reilly, R. K.; Turberfield, A. J.

Programmable One-Pot Multistep Organic Synthesis Using DNA Junctions. J. Am.

Chem. Soc. 2012, 134, 1446–1449.

(117) McKee, M. L.; Milnes, P. J.; Bath, J.; Stulz, E.; Turberfield, A. J.; O’Reilly, R. K.

Multistep DNA-Templated Reactions for the Synthesis of Functional Sequence

Controlled Oligomers. Angew. Chem. Int. Ed Engl. 2010, 49, 7948–7951.

188

(118) Li, X.; Liu, D. R. DNA-Templated Organic Synthesis: Nature’s Strategy for

Controlling Chemical Reactivity Applied to Synthetic Molecules. Angew. Chem. Int.

Ed. 2004, 43, 4848–4870.

Chapter 2

(1) Felsenfeld, G.; Davies, D. R.; Rich, A. Formation of a three-stranded polynucleotide

molecule. J. Am. Chem. Soc. 1957, 79, 2023-2024.

(2) Chen, H.; Meena; McLaughlin, L. W. A Janus-Wedge DNA Triplex with A-W1-T G-

W2-C Base Triplets. J. Am. Chem. Soc. 2008, 130, 13190-13191.

(3) Soto, A. M.; Loo, J.; Marky, L. A. Energetic contributions for the formation of TAT/TAT,

TAT/CGC(+), CGC(+)/CGC(+) base triplet stacks. J. Am. Chem. Soc. 2002, 124,

14355-14363.

(4) Li, T. H.; Weinstein, D. S.; Nicolaou, K. C. The chemical end-ligation of

homopyrimidine oligodeoxyribonucleotides within a DNA triple helix, Chemistry &

Biology 1997, 4, 209-214.

(5) Li, H.; Franzini, R. M.; Bruner, C.; Kool, E. T. Templated Chemistry for Sequence-

Specific Fluorogenic Detection of Duplex DNA, Chembiochem, 2010, 11, 2132-2137.

(6) Zeng, Y.; Pratumyot, Y.; Piao, X.; Bong, D. Discrete assembly of synthetic peptide-

DNA triplex structures from polyvalent melamine-thymine bifacial recognition. J. Am.

Chem. Soc. 2012, 134, 832-835.

(7) Branda, N.; Kurz, G.; Lehn, J.-M. Janus wedges: a new approach towards

nucleobase-pair recognition. Chem. Commun. 1997, 1996, 2443-2444.

(8) Shin, D.; Tor, Y. Bifacial Nucleoside as a Surrogate for Both T A in Duplex DNA. J.

Am. Chem. Soc. 2011, 133, 6926-6929.

189

(9) Fenniri, H.; Packiarajan, M.; Vidale, K. L.; Sherman, D. M.; Hallenga, K.; Wood, K.

V.; Stowell, J. G. Helical rosette nanotubes: Design, self-assembly, characterization.

J. Am. Chem. Soc. 2001, 123, 3854-3855.

(10) Kool, E. T. Preorganization of DNA: Design Principles for Improving Nucleic Acid

Recognition by Synthetic Oligonucleotides. Chem. Rev. 1997, 97, 1473-1488.

(11) Nielsen, P. E.; Egholm, M.; Berg, R. H.; Buchardt, O. Sequence-selective recognition

of DNA by strand displacement with a thymine-substituted polyamide. Science 1991,

254, 1497-1500.

(12) Betts, L.; Josey, J. A.; Veal, J. M.; Jordan, S. R. A nucleic acid triple helix formed by

a peptide nucleic acid-DNA complex. Science 1995, 270, 1838-1841.

(13) Bentin, T.; Larsen, H. J.; Nielsen, P. E. Combined Triplex/Duplex Invasion of Double-

Stranded DNA by “Tail-Clamp” Peptide Nucleic Acid. Biochemistry 2003, 42, 13987-

13995.

(14) Kaihatsu, K.; Shah, R. H.; Zhao, X.; Corey, D. R. Extending Recognition by Peptide

Nucleic Acids (PNAs): Binding to Duplex DNA Inhibition of Transcription by Tail-

Clamp PNA-Peptide Conjugates. Biochemistry 2003, 42, 13996-14003.

(15) Zhang, W.; Nowlan, D. T.; Thomson, L. M.; Lackowski, W. M.; Simanek, E. E.

Orthogonal, Convergent Syntheses of Dendrimers Based on Melamine with One or

Two Unique Surface Sites for Manipulation. J. Am. Chem. Soc. 2001, 123, 8914-

8922.

(16) Blotny, G. Recent applications of 2, 4, 6-trichloro-1, 3, 5-triazine its derivatives in

organic synthesis. Tetrahedron 2006, 62, 9507-9522.

(17) Chen, H.-T.; Neerman, M. F.; Parrish, A. R.; Simanek, E. E. Cytotoxicity, Hemolysis,

Acute in Vivo Toxicity of Dendrimers Based on Melamine, Candidate Vehicles for

Drug Delivery. J. Am. Chem. Soc. 2004, 126, 10044-10048. 190

(18) Leman, L.; Orgel, L.; Ghadiri, M. R. Carbonyl sulfide-mediated prebiotic formation

of peptides. Science 2004, 306, 283-286.

(19) Li, X.; Liu, D. R. DNA-templated organic synthesis: nature’s strategy for controlling

chemical reactivity applied to synthetic molecules. Angew. Chem., Int. Ed. Engl.

2004, 43, 4848-4870.

(20) Nielsen, P. E. Peptide Nucleic Acids the Origin of Life. Chem. Biodiversity 2007, 4,

1996-2002.

(21) Ura, Y.; Beierle, J. M.; Leman, L. J.; Orgel, L. E.; Ghadiri, M. R. Self-assembling

sequence-adaptive peptide nucleic acids. Science 2009, 325, 73-77.

(22) Hysell, M.; Siegel, J. S.; Tor, Y. Synthesis stability of exocyclic triazine nucleosides.

Org. Biomol. Chem. 2005, 3, 2946-2952.

(23) Mittapalli, G. K.; Osornio, Y. M.; Guerrero, M. A.; Reddy, K. R.; Krishnamurthy, R.;

Eschenmoser, A. Mapping the landscape of potentially primordial informational

oligomers: oligodi-peptides tagged with 2,4-disubstituted 5-aminopyrimidines as

recognition elements. Angew. Chem., Int. Ed. 2007, 46, 2478-2484.

(24) Mittapalli, G. K.; Reddy, K. R.; Xiong, H.; Munoz, O.; Han, B.; De Riccardis, F.;

Krishnamurthy, R.; Eschenmoser, A. Mapping the landscape of potentially

primordial informational oligomers: oligodipeptides oligodipeptoids tagged with

triazines as recognition elements. Angew. Chem., Int. Ed. 2007, 46, 2470-2477.

(25) Ferris, J. P.; Sanchez, R. A.; Orgel, L. E. Studies in prebiotic synthesis: III. Synthesis

of pyrimidines from cyanoacetylene cyanate. J. Mol. Biol. 1968, 33, 693-704.

(26) Menor-Salvan, C.; Ruiz-Bermejo, D.; Guzman, M. I.; Osuna-Esteban, S.;

Veintemillas-Verdaguer, S. Synthesis of pyrimidines triazines in ice: implications for

the prebiotic chemistry of nucleobases. Chem. A-Eur. J., 2009, 15, 4411-4418.

191

(27) Whitesides, G. M.; Simanek, E. E.; Mathias, J. P.; Seto, C. T.; Chin, D.; Mammen,

M.; Gordon, D. M. Noncovalent Synthesis: Using Physical-Organic Chemistry To

Make Aggregates. Acc. Chem. Res. 1995, 28, 37-44.

(28) ten Cate, M. G. J.; Huskens, J.; Crego-Calama, M.; Reinhoudt, D. N. Thermodynamic

stability of hydrogen-bonded nanostructures: A calorimetric study. Chem. A-Eur. J.,

2004, 10, 3632-3639.

(29) Prins, L. J.; Reinhoudt, D. N.; Timmerman, P. Noncovalent synthesis using hydrogen

bonding. Angew. Chem., Int. Ed. 2001, 40, 2382-2426.

(30) Oshovsky, G. V.; Reinhoudt, D. N.; Verboom, W. Supramolecular chemistry in water.

Angew. Chem., Int. Ed. 2007, 46, 2366-2393.

(31) De Greef, T. F. A.; Smulders, M. M. J.; Wolffs, M.; Schenning, A.; Sijbesma, R. P.;

Meijer, E. W. Supramolecular Polymerization. Chem. Rev. 2009, 109, 5687-5754.

(32) Cafferty, B. J.; Gallego, I.; Chen, M. C.; Farley, K. I.; Eritja, R.; Hud, N. V. Efficient

Self-Assembly in Water of Long Noncovalent Polymers by Nucleobase Analogues.

J. Am. Chem. Soc. 2013, 135, 2447-2450.

(33) Ai, K.; Liu, Y.; Lu, L. Hydrogen-Bonding Recognition-Induced Color Change of Gold

Nanoparticles for Visual Detection of Melamine in Raw Milk Infant Formula. J. Am.

Chem. Soc. 2009, 131, 9496-9497.

(34) Kawasaki, T.; Tokuhiro, M.; Kimizuka, N.; Kunitake, T. Hierarchical self-assembly of

chiral complementary hydrogen-bond networks in water: reconstitution of

supramolecular membranes. J. Am. Chem. Soc. 2001, 123, 6792-6800.

(35) Ariga, K.; Kunitake, T. Molecular Recognition at Air-Water Related Interfaces:

Complementary Hydrogen Bonding Multisite Interaction. Acc. Chem. Res. 1998, 31,

371-378.

192

(36) Marchi-Artzner, V.; Gulik-Krzywicki, T.; Guedeau-Boudeville, M.-A.; Gosse, C.;

Sanderson, J. M.; Dedieu, J.-C.; Lehn, J.-M. Selective adhesion, lipid exchange

membrane-fusion processes between vesicles of various sizes bearing

complementary molecular recognition groups. ChemPhysChem 2001, 2, 367-376.

(37) Zhou, Z.; Bong, D. Small-Molecule/Polymer Recognition Triggers Aqueous-Phase

Assembly Encapsulation. Langmuir 2013, 29, 144-150.

(38) Ma, M.; Bong, D. Directed peptide assembly at the lipid-water interface cooperatively

enhances membrane binding activity. Langmuir 2011, 27, 1480-1486.

(39) Ma, M.; Bong, D. Determinants of cyanuric acid melamine assembly in water.

Langmuir 2011, 27, 8841-8853.

(40) Ma, M.; Bong, D. Protein assembly directed by synthetic molecular recognition motifs.

Org. Biomol. Chem. 2011, 9, 7296-7299.

(41) Ma, M.; Paredes, A., Bong, D. Intra- intermembrane pairwise molecular recognition

between synthetic hydrogen-bonding phospholipids. J. Am. Chem. Soc. 2008, 130,

14456-14458.

(42) Jahromi, A. H.; Nguyen, L.; Fu, Y.; Miller, K. A.; Baranger, A. M.; Zimmerman, S. C.

A Novel CUG exp -MBNL1 Inhibitor with Therapeutic Potential for Myotonic

Dystrophy Type 1. ACS Chem. Biol. 2013, 8, 1037-1043.

(43) Arambula, J. F.; Ramisetty, S. R.; Baranger, A. M.; Zimmerman, S. C. A simple

ligand that selectively targets CUG trinucleotide repeats inhibits MBNL protein

binding. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 16068-16073.

(44) SantaLucia, J., Jr.; Hicks, D. The thermodynamics of DNA structural motifs. Annu.

Rev. Biophys. Biomol. Struct. 2004, 33, 415-440.

(45) Kool, E. T. Hydrogen bonding, base stacking steric effects in DNA replication. Annu.

Rev. Biophys. Biomol. Struct. 2003, 30, 1-22. 193

(46) Gilli, P.; Ferretti, V.; Gilli, G.; Borea, P. A. Enthalpy-entropy compensation in drug-

receptor binding. J. Phys. Chem. 1994, 98, 1515-1518.

(47) Dunitz, J. D. Win some, lose some: enthalpy-entropy compensation in weak

intermolecular interactions. Chem. Biol. 1995, 2, 709-712.

(48) Ghadiri, M. R.; Choi, C. Secondary structure nucleation in peptides. Transition metal

ion stabilized. alpha.-helices. J. Am. Chem. Soc. 1990, 112, 1630-1632.

(49) Ghadiri, M. R.; Fernholz, A. K. Peptide architecture. Design of stabled .alpha.-helical

metallopeptides via a novel exchange-inert ruthenium(III) complex. J. Am. Chem.

Soc. 1990, 112, 9633-9635.

(50) Cabezas, E.; Satterthwait, A. C. The hydrogen bond mimic approach: solid-phase

synthesis of a peptide stabilized as an α-helix with a hydrazone link. J. Am. Chem.

Soc. 1999, 121, 3862-3875.

(51) Ruan, F.; Chen, Y.; Itoh, K.; Sasaki, T.; Hopkins, P. B. Synthesis of peptides

containing unnatural, metal-ligating residues: aminodiacetic acid as a peptide side

chain. J. Org. Chem. 1991, 56, 4347-4354.

(52) Schafmeister, C. E.; Po, J.; Verdine, G. L. An All-Hydrocarbon Cross-Linking System

for Enhancing the Helicity Metabolic Stability of Peptides. J. Am. Chem. Soc. 2000,

122, 5891-5892.

(53) Chapman, R. N.; Dimartino, G.; Arora, P. S. A Highly Stable Short alpha-Helix

Constrained by a Main-Chain Hydrogen-Bond Surrogate. J. Am. Chem. Soc. 2004,

126, 12252-12253.

(54) Torres, O.; Yuksel, D.; Bernardina, M.; Kumar, K., Bong, D. Peptide tertiary structure

nucleation by side-chain crosslinking with metal complexation double “click”

cycloaddition. ChemBioChem 2008, 9, 1701-1705.

194

(55) Gellman, S. H. Minimal model systems for β –sheet secondary structure in proteins.

Curr. Opin. Chem. Biol. 1998, 2, 717-725.

(56) Mammen, M.; Choi, S.-K.; Whitesides, G. M. Polyvalent Interactions in Biological

Systems: Implications for Design Use of Multivalent Ligands Inhibitors. Angew.

Chem., Int. Ed. 1998, 37, 2754-2794.

(57) Rao, J.; Lahiri, J.; Isaacs, L.; Weis, R. M.; Whitesides, G. M. A trivalent system from

vancomycin.D-ala-D-Ala with higher affinity than avidin-biotin. Science 1998, 280,

708-711.

Chapter 3

(1) Schmidt, J. G.; Christensen, L.; Nielsen, P. E.; Orgel, L. E. Information transfer from

DNA to peptide nucleic acids by template-directed syntheses, Nucleic Acids Res.

1997, 25, 4792-4796.

(2) Goodwin, J. T.; Mehta, A. K.; Lynn, D. G. Digital and analog chemical evolution, Acc.

Chem. Res. 2012, 45, 2189-2199.

(3) Grossmann, T. N.; Seitz, O. DNA-catalyzed transfer of a reporter group, J. Am. Chem.

Soc. 2006, 128, 15596-15597

(4) McKee, M. L.; Evans, A. C.; Gerrard, S. R.; O’Reilly, R. K.; Turberfield, A. J.; Stulz,

E. Peptidomimetic bond formation by DNA-templated acyl transfer, Org. Biomol.

Chem. 2011, 9, 1661-1666.

(5) McKee, M. L.; Milnes, P. J.; Bath, J.; Stulz, E.; Turberfield, A. J.; O’Reilly, R. K.

Multistep DNA-templated reactions for the synthesis of functional sequence

controlled oligomers, Angew. Chem., Int. Ed. 2010, 49, 7948-7951

(6) Erben, A.; Grossmann, T. N.; Seitz, O. DNA-triggered synthesis and bioactivity of

proapoptotic peptides, Angew. Chem. Int. Ed. 2011, 50, 2828-2832. 195

(7) He, Y.; Liu, D. R. A sequential strand-displacement strategy enables efficient six-

step DNA-templated synthesis, J. Am. Chem. Soc. 2011, 133, 9972-9975.

(8) Gao, K.; Orgel, L. E. Polyphosphorylation and Non-Enzymatic Template-Directed

Ligation of Oligonucleotides, Origins of Life and Evolution of the Biosphere 2000,

30, 45-51.

(9) Schmidt, J. G.; Nielsen, P. E.; Orgel, L. E. Information transfer from peptide nucleic

acids to RNA by template-directed syntheses, Nucleic Acids Res. 1997, 25, 4797-

4802.

(10) Gartner, Z. J.; Tse, B. N.; Grubina, R.; Doyon, J. B.; Snyder, T. M.; Liu, D. R. DNA-

templated organic synthesis and selection of a library of macrocycles, Science 2004,

305, 1601-1605.

(11) Tse, B. N.; Snyder, T. M.; Shen, Y. H.; Liu, D. R. Translation of DNA into a library of

13 000 synthetic small-molecule macrocycles suitable for in vitro selection, J. Am.

Chem. Soc. 2008, 130, 15611-15626.

(12) Xu, Y. Z.; Karalkar, N. B.; Kool, E. T. Nonenzymatic autoligation in direct three-color

detection of RNA and DNA point mutations, Nat. Biotechnol. 2001, 19, 148-152.

(13) Silverman, A. P.; Kool, E. T. Detecting RNA and DNA with templated chemical

reactions, Chem. Rev. 2006, 106, 3775-3789.

(14) Chandrasekar, J.; Silverman, S. K. Catalytic DNA with phosphatase activity, Proc.

Natl. Acad. Sci., U.S.A. 2013, 110, 5315-5320.

(15) Parker, D. J.; Xiao, Y.; Aguilar, J. M.; Silverman, S. K. DNA catalysis of a normally

disfavored RNA hydrolysis reaction, J. Am. Chem. Soc. 2013, 135, 8472-8475.

(16) Von Kiedrowski, G. A self-replicating hexadeoxynucleotide, Angew. Chem., Int. Ed.

1986, 25, 932-935.

196

(17) Schreiber, R.; Kempter, S.; Holler, S.; Schuller, V.; Schiffels, D.; Simmel, S. S.;

Nickels, P. C.; Liedl, T. DNA origami-templated growth of arbitrarily shaped metal

nanoparticles, Small, 2011, 7, 1795-1799.

(18) Hopkins, D. S.; Pekker, D.; Goldbart, P. M.; Bezryadin, A. Quantum interference

device made by DNA templating of superconducting nanowires, Science 2005, 308,

1762-1765.

(19) Li, Z. Y.; Zhang, Z. Y. J.; Knipe, R.; Lynn, D. G. DNA-catalyzed polymerization, J.

Am. Chem. Soc. 2002, 124, 746-747.

(20) Rosenbaum, D. M.; Liu, D. R. Efficient and sequence-specific DNA-templated

polymerization of peptide nucleic acid aldehydes, J. Am. Chem. Soc. 2003, 125,

13924-13925.

(21) Lo, P. K.; Sleiman, H. F. Nucleobase-templated polymerization: copying the chain

length and polydispersity of living polymers into conjugated polymers, J. Am. Chem.

Soc. 2009, 131, 4182-4183.

(22) Ura, Y.; Beierle, J. M.; Leman, L. J.; Orgel, L. E.; Ghadiri, M. R. Self-assembling

sequence-adaptive peptide nucleic acids, Science 2009, 325, 73-77.

(23) Heemstra, J. M.; Liu, D. R. Templated synthesis of peptide nucleic acids via

sequence-selective base-filling reactions, J. Am. Chem. Soc. 2009, 131, 11347-

11349.

(24) McHale, R.; O’Reilly, R. K. Nucleobase containing synthetic polymers: advancing

biomimicry via controlled synthesis and self-assembly, Macromolecules 2012, 45,

7665-7675.

(25) Lee, S. H.; Wang, S.; Kool, E. T. Templated chemistry for monitoring damage and

repair directly in duplex DNA, Chem. Comm. 2012, 48, 8069-8071.

197

(26) Selvasekaran, J.; Turnbull, K. D. Chemical ligation of oligodeoxyribonucleotides on

circular DNA templates, Nucleic Acids Res. 1999, 27, 624-627.

(27) Li, T. H.; Weinstein, D. S.; Nicolaou, K. C. The chemical end-ligation of

homopyrimidine oligodeoxyribonucleotides within a DNA triple helix, Chemistry &

Biology 1997, 4, 209-214.

(28) Li, H.; Franzini, R. M.; Bruner, C.; Kool, E. T. Templated Chemistry for Sequence-

Specific Fluorogenic Detection of Duplex DNA, Chembiochem, 2010, 11, 2132-2137.

(29) Zeng, Y.; Pratumyot, Y.; Piao, X.; Bong, D. Discrete assembly of synthetic peptide-

DNA triplex structures from polyvalent melamine-thymine bifacial recognition, J. Am.

Chem. Soc. 2012, 134, 832-835

(30) Piao, X.; Xia, X.; Bong, D. Bifacial peptide nucleic acid directs cooperative folding

and assembly of binary, ternary, and quaternary DNA complexes, Biochemistry

2013, 52, 6313-6323.

(31) Felsenfeld, G.; Davies, D. R.; Rich, A. Formation of a three-stranded polynucleotide

molecule, J. Am. Chem. Soc. 1957, 79, 2023-2024.

(32) Pilch, D. S.; Levenson, C.; Shafer, R. H. Structural analysis of the (dA) 10.2 (dT) 10

triple helix, Proc. Natl. Acad. Sci. U.S.A. 1990, 87, 1942-1946.

(33) Betts, L.; Josey, J. A.; Veal, J. M.; Jordan, S. R. A nucleic acid triple helix formed by

a peptide nucleic acid-DNA complex, Science 1995, 270, 1838-1841.

(34) Soto, A. M.; Loo, J.; Marky, L. A. Energetic contributions for the formation of TAT/TAT,

TAT/CGC+, and CGC+/CGC+ base triplet stacks, J. Am. Chem. Soc. 2002, 124,

14355-14363

(35) Duca, M.; Vekhoff, P.; Oussedik, K.; Halby, L.; Arimondo, P. B. The triple helix: 50

years later, the outcome, Nucleic Acids Res. 2008, 36, 5123-5138.

198

(36) Shin, D.; Tor, Y. Bifacial nucleoside as a surrogate for both T and A in duplex DNA,

J. Am. Chem. Soc. 2011, 133, 6926-6929.

(37) Branda, N.; Kurz, G.; Lehn, J.-M. JANUS WEDGES: a new approach towards

nucleobase-pair recognition, Chem. Commun. 1996, 2443-2444.

(38) Chen, H.; Meena; McLaughlin, L. W. A Janus-Wedge DNA triplex with A-W1-T and

G-W2-C base triplets, J. Am. Chem. Soc. 2008, 130, 13190-13191.

(39) Largy, E.; Liu, W.; Hasan, A.; Perrin, D. M. Base-pairing behavior of a carbocyclic

Janus-AT nucleoside analogue capable of recognizing A and T within a DNA duplex,

ChemBioChem 2013, 14, 2199 − 2208.

(40) Mittapalli, G. K.; Reddy, K. R.; Xiong, H.; Munoz, O.; Han, B.; De Riccardis, F.;

Krishnamurthy, R.; Eschenmoser, A. Mapping the landscape of potentially

primordial informational oligomers: oligodipeptides and oligodipeptoids tagged with

triazines as recognition elements, Angew. Chem., Int. Ed. 2007, 46, 2470-2477.

(41) Hysell, M.; Siegel, J. S.; Tor, Y. Synthesis and stability of exocyclic triazine

nucleosides, Org. Biomol. Chem. 2005, 3, 2946-2952.

(42) Menor-Salvan, C.; Ruiz-Bermejo, D.; Guzman, M. I.; Osuna-Esteban, S.;

Veintemillas-Verdaguer, S. Synthesis of pyrimidines and triazines in ice:

implications for the prebiotic chemistry of nucleobases, Chem.- A Eur. J. 2009, 15,

4411-4418.

(43) Xia, X.; Piao, X.; Fredrick, K.; Bong, D. Bifacial PNA complexation inhibits enzymatic

access to DNA and RNA, ChemBioChem 2013, 15, 31-36.

(44) Xia, X.; Piao, X.; Bong, D. Bifacial PNA as an allosteric switch for aptamer and

ribozyme function, J. Am. Chem. Soc. 2014, 76, 7265-7268.

(45) Liedl, T.; Olapinski, M.; Simmel, F. C. A surface-bound DNA switch driven by a

chemical oscillator, Angew. Chem., Int. Ed. 2006, 45, 5007-5010. 199

(46) Krishnan, Y.; Simmel, F. C. Nucleic acid based molecular devices, Angew. Chem.,

Int. Ed. 2011, 50, 3124-3156.

(47) Chen, X.; Wang, Y.; Liu, Q.; Zhang, Z.; Fan, C.; He, L. Construction of molecular

logic gates with a DNA-cleaving deoxyribozyme, Angew. Chem. Int. Ed. 2006, 118,

1791-1794.

(48) Yoshida, W.; Yokobayashi, Y. Photonic boolean logic gates based on DNA aptamers,

Chem. Commun. 2007, 195-197.

(49) Saghatelian, A.; Volcker, N. H.; Guckian, K. M.; Lin, V. S.-Y.; Ghadiri, M. R. DNA-

based photonic logic gates: AND, NAND, and INHIBIT, J. Am. Chem. Soc. 2003,

125, 346-347.

(50) Guo, P. The emerging field of RNA nanotechnology, Nat. Nano. 2010, 5, 833-842.

(51) Chakraborty, S.; Mehtab, S.; Krishnan, Y. The predictive power of synthetic nucleic

acid technologies in RNA biology, Acc. Chem. Res. 2014, 47, 1710-1719.

(52) Dawson, P. E.; Muir, T. W.; Clarklewis, I.; Kent, S. B. H. Synthesis of proteins by

native chemical ligation, Science 1994, 266, 776-779.

(53) Roloff, A.; Seitz, O. Bioorthogonal reactions challenged: DNA templated native

chemical ligation during PCR, Chem. Sci. 2013, 4, 432-436.

(54) Dose, C.; Ficht, S.; Seitz, O. Reducing product inhibition in DNA-template-controlled

ligation reactions, Angew. Chem., Int. Ed. 2006, 45, 5369-5373.

(55) Li, X.; Liu, D. R. DNA-templated organic synthesis: nature's strategy for controlling

chemical reactivity applied to synthetic molecules, Angew. Chem., Int. Ed. 2004, 43,

4848-4870.

(56) Zhan, Z. Y. J.; Lynn, D. G. Chemical amplification through template-directed

synthesis, J. Am. Chem. Soc. 1997, 119, 12420-12421.

200

(57) Li, Y.; Yongye, A.; Giulianotti, M.; Martinez-Mayorga, K.; Yu, Y.; Houghten, R. A.

Synthesis of cyclic peptides through direct aminolysis of peptide thioesters

catalyzed by imidazole in aqueous organic solutions, J. Comb. Chem. 2009, 11,

1066-1072.

(58) White, C. J.; Yudin, A. K. Contemporary strategies for peptide macrocyclization, Nat.

Chem. 2011, 3, 509-524.

(59) Payne, R. J.; Ficht, S.; Greenberg, W. A.; Wong, C.-H. Cysteine-free peptide and

glycopeptide ligation by direct aminolysis, Angew. Chem., Int. Ed. 2008, 120, 4483-

4487

(60) Murray, J. B.; Terwey, D. P.; Maloney, L.; Karpeisky, A.; Usman, N.; Beigelman, L.;

Scott, W. G. The structural basis of hammerhead ribozyme self-cleavage, Cell 1998,

92, 665-673.

(61) Tuerk, C.; Gold, L. Systematic evolution of ligands by exponential enrichment: RNA

ligands to bacteriophage T4 DNA polymerase, Science 1990, 249, 505-510.

(62) Ellington, A. D.; Szostak, J. W. In vitro selection of RNA molecules that bind specific

ligands, Nature 1990, 346, 818-822.

(63) Von Eggelkraut-Gottanka, R.; Klose, A.; Beck-Sickinger, A. G.; Beyermann, M.

Peptide αthioester formation using standard Fmoc-chemistry, Tetrahedron Letters

2003, 44, 3551-3554.

Chapter 4

(1) Ellington, A. D.; Szostak, J. W. In Vitro Selection of RNA Molecules That Bind

Specific Ligands. Nature 1990, 346, 818–822.

(2) Tuerk, C.; Gold, L. Systematic Evolution of Ligands by Exponential Enrichment:

RNA Ligands to Bacteriophage T4 DNA Polymerase. Science 1990, 249, 505–510. 201

(3) Joyce, G. F. Amplification, Mutation and Selection of Catalytic RNA. Gene 1989, 82,

83–87.

(4) Ameta, S.; Winz, M.-L.; Previti, C.; Jäschke, A. Next-Generation Sequencing

Reveals How RNA Catalysts Evolve from Random Space. Nucleic Acids Res. 2014,

42, 1303–1310.

(5) Dupont, D. M.; Larsen, N.; Jensen, J. K. Characterisation of Aptamer–target

Interactions by Branched Selection and High-Throughput Sequencing of SELEX

Pools. Nucleic acids 2015.

(6) Stoltenburg, R.; Reinemann, C.; Strehlitz, B. SELEX--a (r)evolutionary Method to

Generate High-Affinity Nucleic Acid Ligands. Biomol. Eng. 2007, 24, 381–403.

(7) Patel, D. J. Structural Analysis of Nucleic Acid Aptamers. Curr. Opin. Chem. Biol.

1997, 1, 32–46.

(8) Kimoto, M.; Yamashige, R.; Matsunaga, K.-I.; Yokoyama, S.; Hirao, I. Generation of

High-Affinity DNA Aptamers Using an Expanded Genetic Alphabet. Nat. Biotechnol.

2013, 31, 453–457.

(9) Michaud, M.; Jourdan, E.; Villet, A.; Ravel, A.; Grosset, C.; Peyrin, E. A DNA

Aptamer as a New Target-Specific Chiral Selector for HPLC. J. Am. Chem. Soc.

2003, 125, 8672–8679.

(10) Shoji, A.; Kuwahara, M.; Ozaki, H.; Sawai, H. Modified DNA Aptamer That Binds the

(R)-Isomer of a Thalidomide Derivative with High Enantioselectivity. J. Am. Chem.

Soc. 2007, 129, 1456–1464.

(11) Famulok, M.; Szostak, J. W. Stereospecific Recognition of Tryptophan Agarose by

in Vitro Selected RNA. J. Am. Chem. Soc. 1992, 114, 3990–3991.

(12) Jenison, R. D.; Gill, S. C.; Pardi, A.; Polisky, B. High-Resolution Molecular

Discrimination by RNA. Science 1994, 263, 1425–1429. 202

(13) Shangguan, D.; Li, Y.; Tang, Z.; Cao, Z. C.; Chen, H. W.; Mallikaratchy, P.; Sefah,

K.; Yang, C. J.; Tan, W. Aptamers Evolved from Live Cells as Effective Molecular

Probes for Cancer Study. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 11838–11843.

(14) Fang, X.; Tan, W. Aptamers Generated from Cell-SELEX for Molecular Medicine: A

Chemical Biology Approach. Acc. Chem. Res. 2010, 43, 48–57.

(15) Bock, L. C.; Griffin, L. C.; Latham, J. A.; Vermaas, E. H.; Toole, J. J. Selection of

Single-Stranded DNA Molecules That Bind and Inhibit Human Thrombin. Nature

1992, 355, 564–566.

(16) Jellinek, D.; Green, L. S.; Bell, C.; Janjić, N. Inhibition of Receptor Binding by High-

Affinity RNA Ligands to Vascular Endothelial Growth Factor. Biochemistry 1994, 33,

10450–10456.

(17) Ruckman, J.; Green, L. S.; Beeson, J.; Waugh, S.; Gillette, W. L.; Henninger, D. D.;

Claesson-Welsh, L.; Janjic, N. 2′-Fluoropyrimidine RNA-Based Aptamers to the 165-

Amino Acid Form of Vascular Endothelial Growth Factor (VEGF165) Inhibition of

Receptor Binding and VEGF-Induced Vascular Permeability through Interactions

Requiring the Exon 7-Encoded Domain. J. Biol. Chem. 1998, 273, 20556–20567.

(18) Tuerk, C.; MacDougal, S.; Gold, L. RNA Pseudoknots That Inhibit Human

Immunodeficiency Virus Type 1 Reverse Transcriptase. Proc. Natl. Acad. Sci. U. S.

A. 1992, 89, 6988–6992.

(19) Williams, K. P.; Liu, X. H.; Schumacher, T. N.; Lin, H. Y.; Ausiello, D. A.; Kim, P. S.;

Bartel, D. P. Bioactive and Nuclease-Resistant L-DNA Ligand of Vasopressin. Proc.

Natl. Acad. Sci. U. S. A. 1997, 94, 11285–11290.

(20) Ylera, F.; Lurz, R.; Erdmann, V. A.; Fürste, J. P. Selection of RNA Aptamers to the

Alzheimer’s Disease Amyloid Peptide. Biochem. Biophys. Res. Commun. 2002, 290,

1583–1588. 203

(21) Scarabino, D.; Crisari, A.; Lorenzini, S.; Williams, K.; Tocchini-Valentini, G. P. tRNA

Prefers to Kiss. EMBO J. 1999, 18, 4571–4578.

(22) Ellington, A. D.; Szostak, J. W. Selection in Vitro of Single-Stranded DNA Molecules

That Fold into Specific Ligand-Binding Structures. Nature 1992, 355, 850–852.

(23) Wilson, C.; Szostak, J. W. Isolation of a Fluorophore-Specific DNA Aptamer with

Weak Redox Activity. Chem. Biol. 1998, 5, 609–617.

(24) Sassanfar, M.; Szostak, J. W. An RNA Motif That Binds ATP. Nature 1993, 364,

550–553.

(25) Famulok, M.; Szostak, J. W. In Vitro Selection of Specific Ligand-Binding Nucleic

Acids. Angew. Chem. Int. Ed Engl. 1992, 31, 979–988.

(26) Wang, Y.; Rando, R. R. Specific Binding of Aminoglycoside Antibiotics to RNA.

Chem. Biol. 1995, 2, 281–290.

(27) Burke, D. H.; Hoffman, D. C.; Brown, A.; Hansen, M.; Pardi, A.; Gold, L. RNA

Aptamers to the Peptidyl Transferase Inhibitor Chloramphenicol. Chem. Biol. 1997,

4, 833–843.

(28) Hofmann, H. P.; Limmer, S.; Hornung, V.; Sprinzl, M. Ni2+-Binding RNA Motifs with

an Asymmetric Purine-Rich Internal Loop and a GA Base Pair. RNA 1997.

(29) Ng, E. W. M.; Shima, D. T.; Calias, P.; Cunningham, E. T.; Guyer, D. R.; Adamis, A.

P. Pegaptanib, a Targeted Anti-VEGF Aptamer for Ocular Vascular Disease. Nat.

Rev. Drug Discov. 2006, 5, 123–132.

(30) Blind, M.; Blank, M. Aptamer Selection Technology and Recent Advances. Mol. Ther.

Nucleic Acids 2015, 4, e223.

(31) Iliuk, A. B.; Hu, L.; Tao, W. A. Aptamer in Bioanalytical Applications. Anal. Chem.

2011, 83, 4440–4452.

204

(32) Song, K.-M.; Lee, S.; Ban, C. Aptamers and Their Biological Applications.

Sensors 2012, 12, 612–631.

(33) Klug, S. J.; Famulok, M. All You Wanted to Know about SELEX. Mol. Biol. Rep.

1994, 20, 97–107.

(34) Lehman, N.; Joyce, G. F. Evolution in Vitro of an RNA Enzyme with Altered Metal

Dependence. Nature 1993, 361, 182–185.

(35) Beaudry, A. A.; Joyce, G. F. Directed Evolution of an RNA Enzyme. Science 1992,

257, 635–641.

(36) Wu, W.; Henderson, L. E.; Copeland, T. D.; Gorelick, R. J.; Bosche, W. J.; Rein, A.;

Levin, J. G. Human Immunodeficiency Virus Type 1 Nucleocapsid Protein Reduces

Reverse Transcriptase Pausing at a Secondary Structure near the Murine Leukemia

Virus Polypurine Tract. J. Virol. 1996, 70, 7132–7142.

(37) Harrison, G. P.; Mayo, M. S.; Hunter, E.; Lever, A. M. Pausing of Reverse

Transcriptase on Retroviral RNA Templates Is Influenced by Secondary Structures

Both 5’ and 3' of the Catalytic Site. Nucleic Acids Res. 1998, 26, 3433–3442.

(38) Zeng, Y.; Pratumyot, Y.; Piao, X.; Bong, D. Discrete Assembly of Synthetic Peptide-

DNA Triplex Structures from Polyvalent Melamine-Thymine Bifacial Recognition. J.

Am. Chem. Soc. 2012, 134, 832–835.

(39) Piao, X.; Xia, X.; Bong, D. Bifacial Peptide Nucleic Acid Directs Cooperative Folding

and Assembly of Binary, Ternary, and Quaternary DNA Complexes. Biochemistry

2013, 52, 6313–6323.

(40) Xia, X.; Piao, X.; Bong, D. Bifacial Peptide Nucleic Acid as an Allosteric Switch for

Aptamer and Ribozyme Function. J. Am. Chem. Soc. 2014, 136, 7265–7268.

205

(41) Ruff, K. M.; Snyder, T. M.; Liu, D. R. Enhanced Functional Potential of Nucleic Acid

Aptamer Libraries Patterned to Increase Secondary Structure. J. Am. Chem. Soc.

2010, 132, 9453–9464.

(42) Xu, W.; Ellington, A. D. Anti-Peptide Aptamers Recognize Amino Acid Sequence

and Bind a Protein Epitope. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 7475–7480.

(43) Malim, M. H.; Hauber, J.; Le, S. Y.; Maizel, J. V.; Cullen, B. R. The HIV-1 Rev Trans-

Activator Acts through a Structured Target Sequence to Activate Nuclear Export of

Unspliced Viral mRNA. Nature 1989, 338, 254–257.

(44) Felber, B. K.; Hadzopoulou-Cladaras, M.; Cladaras, C.; Copeland, T.; Pavlakis, G.

N. Rev Protein of Human Immunodeficiency Virus Type 1 Affects the Stability and

Transport of the Viral mRNA. Proc. Natl. Acad. Sci. U. S. A. 1989, 86, 1495–1499.

(45) Emerman, M.; Vazeux, R.; Peden, K. The Rev Gene Product of the Human

Immunodeficiency Virus Affects Envelope-Specific RNA Localization. Cell 1989, 57,

1155–1165.

(46) Hammarskjöld, M. L.; Heimer, J.; Hammarskjöld, B.; Sangwan, I.; Albert, L.; Rekosh,

D. Regulation of Human Immunodeficiency Virus Env Expression by the Rev Gene

Product. J. Virol. 1989, 63, 1959–1966.

(47) Kjems, J.; Brown, M.; Chang, D. D.; Sharp, P. A. Structural Analysis of the

Interaction between the Human Immunodeficiency Virus Rev Protein and the Rev

Response Element. Proc. Natl. Acad. Sci. U. S. A. 1991, 88, 683–687.

(48) Kjems, J.; Calnan, B. J.; Frankel, A. D.; Sharp, P. A. Specific Binding of a Basic

Peptide from HIV-1 Rev. EMBO J. 1992, 11, 1119–1129.

(49) Battiste, J. L.; Tan, R.; Frankel, A. D.; Williamson, J. R. Binding of an HIV Rev

Peptide to Rev Responsive Element RNA Induces Formation of Purine-Purine Base

Pairs. Biochemistry 1994, 33, 2741–2747. 206

(50) Tan, R.; Frankel, A. D. Costabilization of Peptide and RNA Structure in an HIV Rev

Peptide-RRE Complex. Biochemistry 1994, 33, 14579–14585.

(51) Tan, R.; Chen, L.; Buettner, J. A.; Hudson, D.; Frankel, A. D. RNA Recognition by

an Isolated Alpha Helix. Cell 1993, 73, 1031–1040.

(52) Bartel, D. P.; Zapp, M. L.; Green, M. R.; Szostak, J. W. HIV-1 Rev Regulation

Involves Recognition of Non-Watson-Crick Base Pairs in Viral RNA. Cell 1991, 67,

529–536.

(53) Giver, L.; Bartel, D.; Zapp, M.; Pawul, A.; Green, M.; Ellington, A. D. Selective

Optimization of the Rev-Binding Element of HIV-1. Nucleic Acids Res. 1993, 21,

5509–5516.

(54) Ye, X.; Gorin, A.; Frederick, R.; Hu, W.; Majumdar, A.; Xu, W.; McLendon, G.;

Ellington, A.; Patel, D. J. RNA Architecture Dictates the Conformations of a Bound

Peptide. Chem. Biol. 1999, 6, 657–669.

(55) Harrison, R. S.; Shepherd, N. E.; Hoang, H. N.; Ruiz-Gómez, G.; Hill, T. A.; Driver,

R. W.; Desai, V. S.; Young, P. R.; Abbenante, G.; Fairlie, D. P. Downsizing Human,

Bacterial, and Viral Proteins to Short Water-Stable Alpha Helices That Maintain

Biological Potency. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 11686–11691.

(56) Lau, Y. H.; de Andrade, P.; Wu, Y.; Spring, D. R. Peptide Stapling Techniques

Based on Different Macrocyclisation Chemistries. Chem. Soc. Rev. 2014, 44, 91–

102.

(57) Blackwell, H. E.; Grubbs, R. H. Highly Efficient Synthesis of Covalently Cross-Linked

Peptide Helices by Ring-Closing Metathesis. Angew. Chem. Int. Ed. 1998, 37,

3281–3284.

207

(58) Christian E. Schafmeister; Julia Po; Verdine*, G. L. An All-Hydrocarbon Cross-

Linking System for Enhancing the Helicity and Metabolic Stability of Peptides. J. Am.

Chem. Soc. 2000, 122, 5891–5892.

(59) Felix, A. M.; Heimer, E. P.; Wang, C.-T.; Lambros, T. J.; Fournier, A.; Mowles, T. F.;

Maines, S.; Campbell, R. M.; Wegrzynski, B. B.; Toome, V.; et al. Synthesis,

Biological Activity and Conformational Analysis of Cyclic GRF Analogs. Int. J. Pept.

Protein Res. 1988, 32, 441–454.

(60) Taylor, J. W. The Synthesis and Study of Side-Chain Lactam-Bridged Peptides.

Biopolymers 2002, 66, 49–75.

(61) Shepherd, N. E.; Hoang, H. N.; Abbenante, G.; Fairlie, D. P. Single Turn Peptide

Alpha Helices with Exceptional Stability in Water. J. Am. Chem. Soc. 2005, 127,

2974–2983.

(62) Scrima, M.; Le Chevalier-Isaad, A.; Rovero, P.; Papini, A. M.; Chorev, M.; D’Ursi, A.

M. CuI-Catalyzed Azide–Alkyne Intramolecular I-to-(i+4) Side-Chain-to-Side-Chain

Cyclization Promotes the Formation of Helix-Like Secondary Structures. Eur. J. Org.

Chem. 2010, 2010, 446–457.

(63) Kawamoto, S. A.; Coleska, A.; Ran, X.; Yi, H.; Yang, C.-Y.; Wang, S. Design of

Triazole-Stapled BCL9 α-Helical Peptides to Target the β-catenin/B-Cell

CLL/lymphoma 9 (BCL9) Protein-Protein Interaction. J. Med. Chem. 2012, 55,

1137–1146.

(64) Torres, O.; Yüksel, D.; Bernardina, M.; Kumar, K.; Bong, D. Peptide Tertiary

Structure Nucleation by Side-Chain Crosslinking with Metal Complexation and

Double “Click” Cycloaddition. Chembiochem 2008, 9, 1701–1705.

(65) Lau, Y. H.; de Andrade, P.; Quah, S.-T.; Rossmann, M.; Laraia, L.; Sköld, N.; Sum,

T. J.; Rowling, P. J. E.; Joseph, T. L.; Verma, C.; et al. Functionalised Staple 208

Linkages for Modulating the Cellular Activity of Stapled Peptides. Chem. Sci. 2014,

5, 1804–1809.

(66) Brunel, F. M.; Dawson, P. E. Synthesis of Constrained Helical Peptides by Thioether

Ligation: Application to Analogs of gp41. Chem. Commun. 2005, 2552–2554.

(67) Verdine, G. L.; Hilinski, G. J. Stapled Peptides for Intracellular Drug Targets.

Methods Enzymol. 2012, 503, 3–33.

(68) Walensky, L. D.; Bird, G. H. Hydrocarbon-Stapled Peptides: Principles, Practice,

and Progress. J. Med. Chem. 2014, 57, 6275–6288.

(69) Moellering, R. E.; Cornejo, M.; Davis, T. N.; Del Bianco, C.; Aster, J. C.; Blacklow,

S. C.; Kung, A. L.; Gilliland, D. G.; Verdine, G. L.; Bradner, J. E. Direct Inhibition of

the NOTCH Transcription Factor Complex. Nature 2009, 462, 182–188.

(70) Walensky, L. D.; Kung, A. L.; Escher, I.; Malia, T. J.; Barbuto, S.; Wright, R. D.;

Wagner, G.; Verdine, G. L.; Korsmeyer, S. J. Activation of Apoptosis in Vivo by a

Hydrocarbon-Stapled BH3 Helix. Science 2004, 305, 1466–1470.

(71) Mills, N. L.; Daugherty, M. D.; Frankel, A. D.; Guy, R. K. An Alpha-Helical

Peptidomimetic Inhibitor of the HIV-1 Rev-RRE Interaction. J. Am. Chem. Soc. 2006,

128, 3496–3497.

(72) Battiste, J. L.; Mao, H.; Rao, N. S.; Tan, R.; Muhandiram, D. R.; Kay, L. E.; Frankel,

A. D.; Williamson, J. R. Alpha Helix-RNA Major Groove Recognition in an HIV-1 Rev

Peptide-RRE RNA Complex. Science 1996, 273, 1547–1551.

(73) Schütze, T.; Wilhelm, B.; Greiner, N.; Braun, H.; Peter, F.; Mörl, M.; Erdmann, V. A.;

Lehrach, H.; Konthur, Z.; Menger, M.; et al. Probing the SELEX Process with next-

Generation Sequencing. PLoS One 2011, 6, e29604.

209

(74) Ogawa, N.; Biggin, M. D. High-Throughput SELEX Determination of DNA

Sequences Bound by Transcription Factors in Vitro. Methods Mol. Biol. 2012, 786,

51–63.

(75) Harrison, R. S.; Shepherd, N. E.; Hoang, H. N.; Beyer, R. L.; Ruiz-Gómez, G.; Kelso,

M. J.; Mei Kok, W.; Hill, T. A.; Abbenante, G.; Fairlie, D. P. Helical Cyclic

Pentapeptides Constrain HIV-1 Rev Peptide for Enhanced RNA Binding.

Tetrahedron 2014, 70, 7645–7650.

(76) Martin, M. Cutadapt Removes Adapter Sequences from High-Throughput

Sequencing Reads. EMBnet.journal 2011, 17, 10–12.

(77) Alam, K. K.; Chang, J. L.; Burke, D. H. FASTAptamer: A Bioinformatic Toolkit for

High-Throughput Sequence Analysis of Combinatorial Selections. Mol. Ther.

Nucleic Acids 2015, 4, e230.

(78) Daugherty, M. D.; D’Orso, I.; Frankel, A. D. A Solution to Limited Genomic Capacity:

Using Adaptable Binding Surfaces to Assemble the Functional HIV Rev Oligomer

on RNA. Mol. Cell 2008, 31, 824–834.

(79) Ye, X.; Gorin, A.; Ellington, A. D.; Patel, D. J. Deep Penetration of an Alpha-Helix

into a Widened RNA Major Groove in the HIV-1 Rev Peptide-RNA Aptamer Complex.

Nat. Struct. Biol. 1996, 3, 1026–1033.

(80) Leclerc, F.; Cedergren, R.; Ellington, A. D. A 3-DIMENSIONAL MODEL OF THE

REV-BINDING ELEMENT OF HIV-1 DERIVED FROM ANALYSES OF APTAMERS.

Nat. Struct. Biol. 1994, 1, 293–300.

(81) Link, A. J.; Vink, M. K. S.; Tirrell, D. A. Presentation and Detection of Azide

Functionality in Bacterial Cell Surface Proteins. J. Am. Chem. Soc. 2004, 126,

10598–10602.

210

(82) Lau, Y. H.; Wu, Y.; de Andrade, P.; Galloway, W. R. J. D.; Spring, D. R. A Two-

Component “Double-Click” Approach to Peptide Stapling. Nat. Protoc. 2015, 10,

585–594.

(83) Shao, K.; Ding, W.; Wang, F.; Li, H.; Ma, D.; Wang, H. Emulsion PCR: A High

Efficient Way of PCR Amplication of Random DNA Libraries in Aptamer Selection,

Plos ONE 2011, 6, e24910.

(84) Musheev, M. U.; Krylov, S. N. Selection of Aptamers by systematic evolution of

Ligands by Expotential Enrichment: Addressing the Polymerase Chain Reaction

Issue, Analytica Chimica Acta 2006, 564, 91-96.

Appendix A

(1) Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing

reads, EMBnet.journal, 2011, 17.1, 10-12.

(2) Alam, K. K.; Chang, J. L.; Burke, D. H. FASTApamer: a bioinformatic toolkit for high-

throughput sequence analysis of combinatorial selections. Molecular Therapy-

Nucleic Acids, 2015, 4, e230.

211