Download as pdf or txt
Download as pdf or txt
You are on page 1of 241

Copyright Undertaking

This thesis is protected by copyright, with all rights reserved.

By reading and using the thesis, the reader understands and agrees to the following terms:

1. The reader will abide by the rules and legal ordinances governing copyright regarding the
use of the thesis.

2. The reader will use the thesis for the purpose of research or private study only and not for
distribution or further reproduction or any other purpose.

3. The reader agrees to indemnify and hold the University harmless from and against any loss,
damage, cost, liability or expenses arising from copyright infringement or unauthorized
usage.

IMPORTANT
If you have reasons to believe that any materials in this thesis are deemed not suitable to be
distributed in this form, or a copyright owner having difficulty with the material being included in
our database, please contact lbsys@polyu.edu.hk providing details. The Library will look into
your claim and consider taking remedial action upon receipt of the written requests.

Pao Yue-kong Library, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong

http://www.lib.polyu.edu.hk
CHEMOSELECTIVE CYSTEINE
MODIFICATION OF PEPTIDES AND
PROTEINS USING 1H-ISOINDOLIUM-

BASED ELECTRON-DEFICIENT ALLENES

O WA YI

MPhil

The Hong Kong Polytechnic University


2021
The Hong Kong Polytechnic University
Department of Applied Biology and Chemical Technology

Chemoselective Cysteine Modification of Peptides


and Proteins using 1H-Isoindolium-based Electron-

Deficient Allenes

O Wa Yi

A thesis submitted in partial fulfilment of the requirements for


the degree of Master of Philosophy

September 2020
CERTIFICATE OF ORIGINALITY

I hereby declare that this thesis is my own work and that, to the best of my knowledge

and belief, it reproduces no material previously published or written, nor material that

has been accepted for the award of any other degree or diploma, except where due

acknowledgement has been made in the text.

_______________________

O WA YI

I
ABSTRACT

Site-selective modification of peptides and proteins has been recognized as a

powerful tool for biological studies. It is of ongoing interest to develop new

modification strategies with high selectivity and efficiency under mild reaction

conditions. Cysteine with low abundance and high nucleophilicity is an ideal residue

for selective modification. A novel approach for cysteine modification of peptides and

proteins using 1H-isoindolium-based electron-deficient allenes was developed.

A novel electrophile-mediated cascade cyclization/iodination of propargylamine-

based 1,6-diyne has been developed for the synthesis of a new class of spiro 1H-

isoindoliums 3. Upon treatment with base, the alkyne moieties of 1H-isoindoliums were

transformed to allene functionalities to give the corresponding allene products. The

application of this new class of 1H-isoindolium-based electron-deficient allenes to

cysteine modification of peptides and proteins was presented. By treatment of cysteine-

containing peptides with 5 equivalents of allene reagents in aqueous media at 25 °C for

4 h, cysteine-modified peptides were obtained in good conversions up to 96% as

confirmed by LC–MS/MS analysis. Allene reagents with different functional groups

were compatible with this modification and the modified products were stable towards

excessive thiol-containing reagents and reducing agents. Further application of this

modification was demonstrated in protein modification and labeling. Fluorescent tags

could be labeled on the cysteine-containing proteins by sequential modification using

the azide-alkyne click reaction.

II
Based on the success of this newly developed cysteine modification method, a one-

step procedure for preparation of electron deficient allene reagents for bioconjugation

was established to enhance the efficiency of the modification. By reaction of

propargylamine-based 1,6-diynes with iodine monochloride for the formation of in situ

generated 1H-isoindoliums, followed by the addition of 1 equivalent of reducing agent,

stock solutions of in situ generated electron deficient allene reagents with high stability

could be prepared for direct use in bioconjugation.

With 1 equivalent of the in situ generated allene reagents, the modification of

peptides could proceed with high efficiency up to 99% conversion in 4 h and the

performance was comparable with the isolated allene reagents. The in situ generated

allene reagents could be further applied to protein modification and labeling. The results

demonstrated that this approach could serve as an alternative method for preparation of

allene reagents and be utilized in the newly developed cysteine modification using

electron-deficient allenes.

III
ACKNOWLEDGEMENT

I would like to express my deepest gratitude to my supervisor, Dr. Man-Kin Wong

for his supervision, support and encouragement as well as invaluable advices for my

research study in these two years.

I would like to convey my greatest appreciation to my group members: Dr. Ka-

Yan Karen Kung, Dr. Jian-Fang Cui, Mr. Bin Yang, Mr. Jie-Ren Deng, Mr. Jia-Jun

Jiang, Mr. Wai-Ming Yip, and Miss Hoi-Yi Sit for their precious guidance and patience

in providing useful and constructive recommendations for my research. I would also

like to express my sincere thanks to colleagues in Shenzhen Research Institute: Dr.

Qiong Yu and Mr. Li Wu Huang for their efforts and assistance in synthesis of some

compounds for this project.

Besides, I would like to thank all the staffs and technicians of the Department of

Applied Biology and Chemical Technology in the Hong Kong Polytechnic University

for their genuine help and technical support, especially Dr. Pui-Kin So and Dr. Melody

Yee-Man Wong for training and assistance in mass spectrometry.

I sincerely thank Dr. Man-Kin Wong and his research group again. It is my honor

to work and discuss with them in these two years and I will never forget this valuable

and memorable experience.

IV
TABLE OF CONTENT

CERTIFICATE OF ORIGINALITY I

ABSTRACT II

ACKNOWLEDGEMENT IV

TABLE OF CONTENT V

LIST OF FIGURES X

LIST OF TABLES XVII

LIST OF ABBREVIATIONS XVIII

Chapter 1 Introduction 1

1.1 Allene Chemistry 1

1.1.1 Synthesis of Allenes 3

1.1.2 Reactions of Allenes 10

1.2 Chemical Modification of Peptides and Proteins 16

1.2.1 Lysine and N-Terminal Modification 17

1.2.2 Modification of Other Natural Amino Acids 23

1.2.3 Modification of Unnatural Amino Acids 27

1.3 Cysteine Modification 30

1.3.1 Classical Methods 32

1.3.2 Transition Metal-Free Modifications 35

1.3.3 Transition Metal-Mediated Modifications 38

V
Chapter 2 Cysteine Modification of Peptides and Proteins using isolated

Electron-Deficient Allenes 40

2.1 Introduction 40

2.2 Results and Discussion 42

2.2.1 Modification of Cysteine-Containing Peptides Using Isolated Electron-

Deficient Allenes 42

2.2.1.1 Synthesis of Electron-Deficient Allenes 42

2.2.1.2 Optimization of Modification Reaction Conditions 45

2.2.1.3 Model Reaction of Cysteine Modification by Electron-Deficient

Allenes 49

2.2.1.4 Investigation of Cysteine Selectivity of the Modification 50

2.2.1.5 Investigation of Scope of Allene reagents 53

2.2.1.6 Stability Study of Allene-modified Bioconjugates 58

2.2.2 Modification of Cysteine-Containing Proteins Using Isolated Electron-

Deficient Allenes 59

2.2.2.1 Modification of BSA Protein with Allene Reagents 59

2.2.2.2 Control Experiments with Non-Free Cysteine-Containing Proteins 63

2.2.2.3 Fluorescent Labeling of Proteins 64

2.3 Conclusion 66

2.4 Suggestions for Future Research 67

VI
Chapter 3 Cysteine Modification of Peptides and Proteins using in situ

Generated Electron-Deficient Allenes 69

3.1 Introduction 69

3.2 Results and Discussion 70

3.2.1 Modification of Cysteine-Containing Peptides Using in situ Generated

Electron-Deficient Allenes 70

3.2.1.1 Optimization of Reaction Conditions 70

3.2.1.2 Time Course Experiments of Cysteine Modification of Peptides by in

situ Generated Allene Reagents 75

3.2.1.3 Investigation of Scope of in situ Generated Allene Reagents 77

3.2.2 Modification of Cysteine-Containing Proteins Using in situ Generated

Electron-Deficient Allenes 81

3.3 Conclusion 86

Chapter 4 Experimental Section 87

4.1 General procedures 87

4.2 Literature Reference of Compounds 89

4.3 Calculation of Peptide Conversion 89

4.4 Calculation of Protein Conversion 90

4.5 Synthesis of Propargylamine-based 1,6-diynes 2 90

4.5.1 Synthesis of Coumarin-derived Propargylamine-based 1,6-diyne 2g 94

4.6 Synthesis of 1H-isoindoliums 3 97

4.7 Synthesis of Electron-Deficient Allenes 4 103

VII
4.8 Synthesis of Model Compound 7 108

4.9 Modification of Peptides Using Isolated Electron-Deficient Allene Reagents 4

110

4.10 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using Isolated Electron-Deficient Allene Reagents 4 at Different pH Values 110

4.11 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using Isolated Electron-Deficient Allene Reagents 4 at Different Temperatures 110

4.12 Studies on the Stability of the Modified Peptide 5a 111

4.13 Modification of Proteins Using Isolated Electron-Deficient Allene Reagents 4

111

4.14 Sequential Modification of Proteins via Huisgen Azide-Alkyne Cycloaddition

111

4.15 SDS-PAGE Analysis of Native and Modified Proteins 111

4.16 Preparation of in situ Generated Electron-Deficient Allene Reagents 4 112

4.17 Modification of Peptides Using in situ Generated Electron-Deficient Allene

Reagents 4 112

4.18 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using in situ Generated Electron-Deficient Allene Reagents 4 at Different pH

Values 113

4.19 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using in situ Generated Electron-Deficient Allene Reagents 4 at Different

Temperatures 113

VIII
4.20 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using in situ Generated Electron-Deficient Allene Reagents 4a-d 114

4.21 Modification of Proteins Using in situ Generated Electron-Deficient Allene

Reagents 4 114

4.22 MS Spectra 115

4.23 NMR Spectra 172

References 204

IX
LIST OF FIGURES

Chapter 1 Introduction

Figure 1.1 (a) Structure of allene; (b) Description of π bonding of allene. 1

Figure 1.2 Examples of naturally occurring allenic compounds. 2

Figure 1.3 Examples of allenic pharmacologically active compounds. 2

Chapter 2 Cysteine Modification of Peptides and Proteins using isolated Electron-

Deficient Allenes

Figure 2.1 The synthesis of allene 4a. 43

Figure 2.2 MS/MS spectrum of cysteine-modified peptide 5a. 45

Figure 2.3 Time course experiments of the formation of cysteine modified peptide 5a

at different pH values. 47

Figure 2.4 Time course experiments of the formation of cysteine modified peptide 5a

at different temperatures. 48

Figure 2.5 MS/MS spectrum of cysteine-modified peptide 6a. 52

Figure 2.6 MS/MS spectrum of cysteine-modified peptide 8a. 52

Figure 2.7 MS/MS spectrum of cysteine-modified peptide 9a. 53

Figure 2.8 Mass spectrum of (a) BSA-1; (b) BSA-2 and (c) BSA-3. 61

X
Figure 2.9 Molecular feature extraction spectrum of cysteine modified

KGLVLIAFSQYLQQCPF after chymotrypsin digestion of BSA-1. 62

Figure 2.10 Molecular feature extraction spectrum of cysteine modified

SQYLQQCPFDEHVKLVNELTEF after chymotrypsin digestion of BSA-1. 62

Figure 2.11 Mass spectrum of RNaseA after reaction with 4a. 63

Figure 2.12 Mass spectrum of lysozyme after reaction with 4a. 63

Figure 2.13 Mass spectrum of rhodamine-labeled BSA-4. 65

Figure 2.14 SDS-PAGE analysis of BSA, allene-modified BSA-3 and rhodamine-

labelled BSA-4. 65

Chapter 3 Cysteine Modification of Peptides and Proteins using in situ Generated

Electron-Deficient Allenes

Figure 3.1 LC–MS analysis of in situ generated allene reagent (TIC chromatogram).

71

Figure 3.2 LC–MS analysis of side product 5a' (TIC chromatogram). 73

Figure 3.3 MS/MS analysis of side product 5a'. 74

Figure 3.4 Time course experiments of the formation of cysteine modified peptide 5a

at (a) different pH values; (b) different temperatures. 76

Figure 3.5 Time course experiment of formation of modified peptides 5a-d. 81

Figure 3.6 Mass spectrum of (a) BSA-6; (b) BSA-7; (c) BSA-8 and (d) BSA-9. 83

XI
Figure 3.7 Mass spectrum of fluorescent-labeled BSA-10. 84

Figure 3.8 SDS-PAGE analysis of native BSA and BSA-10. 85

Chapter 4 Experimental Section

Figure 4.1 Extracted ion chromatogram of native peptide AYEMWCFHQR 1a. 115

Figure 4.2 Mass spectrum of native peptide AYEMWCFHQR 1a. 115

Figure 4.3 MS/MS spectrum of native peptide AYEMWCFHQR 1a. 116

Figure 4.4 Extracted ion chromatogram of cysteine-modified peptide 5a. 116

Figure 4.5 Mass spectrum of cysteine-modified peptide 5a. 117

Figure 4.6 MS/MS spectrum of cysteine-modified peptide 5a. 117

Figure 4.7 Extracted ion chromatogram of cysteine-modified peptide 5b. 118

Figure 4.8 Mass spectrum of cysteine-modified peptide 5b. 118

Figure 4.9 MS/MS spectrum of cysteine-modified peptide 5b. 119

Figure 4.10 Extracted ion chromatogram of cysteine-modified peptide 5c. 120

Figure 4.11 Mass spectrum of cysteine-modified peptide 5c. 120

Figure 4.12 MS/MS spectrum of cysteine-modified peptide 5c. 121

Figure 4.13 Extracted ion chromatogram of cysteine-modified peptide 5d. 122

Figure 4.14 Mass spectrum of cysteine-modified peptide 5d. 122

Figure 4.15 MS/MS spectrum of cysteine-modified peptide 5d. 123

XII
Figure 4.16 Extracted ion chromatogram of cysteine-modified peptide 5e. 124

Figure 4.17 Mass spectrum of cysteine-modified peptide 5e. 124

Figure 4.18 MS/MS spectrum of cysteine-modified peptide 5e. 125

Figure 4.19 Extracted ion chromatogram of cysteine-modified peptide 5f. 126

Figure 4.20 Mass spectrum of cysteine-modified peptide 5f. 126

Figure 4.21 MS/MS spectrum of cysteine-modified peptide 5f. 127

Figure 4.22 Extracted ion chromatogram of cysteine-modified peptide 5g. 128

Figure 4.23 Mass spectrum of cysteine-modified peptide 5g. 128

Figure 4.24 MS/MS spectrum of cysteine-modified peptide 5g. 129

Figure 4.25 Extracted ion chromatogram of cysteine-modified peptide 5h. 130

Figure 4.26 Mass spectrum of cysteine-modified peptide 5h. 130

Figure 4.27 MS/MS spectrum of cysteine-modified peptide 5h. 131

Figure 4.28 Extracted ion chromatogram of native peptide STSSSCNLSK 1b. 132

Figure 4.29 Mass spectrum of native peptide STSSSCNLSK 1b. 132

Figure 4.30 MS/MS spectrum of native peptide STSSSCNLSK 1b. 133

Figure 4.31 Extracted ion chromatogram of cysteine-modified peptide 6a. 133

Figure 4.32 Mass spectrum of cysteine-modified peptide 6a. 134

Figure 4.33 MS/MS spectrum of cysteine-modified peptide 6a. 134

Figure 4.34 Extracted ion chromatogram of cysteine-modified peptide 6b. 135

XIII
Figure 4.35 Mass spectrum of cysteine-modified peptide 6b. 135

Figure 4.36 MS/MS spectrum of cysteine-modified peptide 6b. 136

Figure 4.37 Extracted ion chromatogram of cysteine-modified peptide 6c. 137

Figure 4.38 Mass spectrum of cysteine-modified peptide 6c. 137

Figure 4.39 MS/MS spectrum of cysteine-modified peptide 6c. 138

Figure 4.40 Extracted ion chromatogram of cysteine-modified peptide 6d. 139

Figure 4.41 Mass spectrum of cysteine-modified peptide 6d. 139

Figure 4.42 MS/MS spectrum of cysteine-modified peptide 6d. 140

Figure 4.43 Extracted ion chromatogram of cysteine-modified peptide 6e. 141

Figure 4.44 Mass spectrum of cysteine-modified peptide 6e. 141

Figure 4.45 MS/MS spectrum of cysteine-modified peptide 6e. 142

Figure 4.46 Extracted ion chromatogram of cysteine-modified peptide 6f. 143

Figure 4.47 Mass spectrum of cysteine-modified peptide 6f. 143

Figure 4.48 MS/MS spectrum of cysteine-modified peptide 6f. 144

Figure 4.49 Extracted ion chromatogram of cysteine-modified peptide 6g. 145

Figure 4.50 Mass spectrum of cysteine-modified peptide 6g. 145

Figure 4.51 MS/MS spectrum of cysteine-modified peptide 6g. 146

Figure 4.52 Extracted ion chromatogram of cysteine-modified peptide 6h. 147

Figure 4.53 Mass spectrum of cysteine-modified peptide 6h. 147

XIV
Figure 4.54 MS/MS spectrum of cysteine-modified peptide 6h. 148

Figure 4.55 Extracted ion chromatogram of native peptide KSTFC 1c. 149

Figure 4.56 Mass spectrum of native peptide KSTFC 1c. 149

Figure 4.57 MS/MS spectrum of native peptide KSTFC 1c. 150

Figure 4.58 Extracted ion chromatogram of cysteine-modified peptide 8a. 150

Figure 4.59 Mass spectrum of cysteine-modified peptide 8a. 151

Figure 4.60 MS/MS spectrum of cysteine-modified peptide 8a. 151

Figure 4.61 Extracted ion chromatogram of native peptide CSKFR 1d. 152

Figure 4.62 Mass spectrum of native peptide CSKFR 1d. 152

Figure 4.63 MS/MS spectrum of native peptide CSKFR 1d. 153

Figure 4.64 Extracted ion chromatogram of cysteine-modified peptide 9a. 154

Figure 4.65 Mass spectrum of cysteine-modified peptide 9a. 154

Figure 4.66 MS/MS spectrum of cysteine-modified peptide 9a. 155

Figure 4.67 Extracted ion chromatogram of native peptide STSSSANLSK 1e. 156

Figure 4.68 Mass spectrum of native peptide STSSSANLSK 1e. 156

Figure 4.69 MS/MS spectrum of native peptide STSSSANLSK 1e. 157

Figure 4.70 Extracted ion chromatogram of native peptide STSSSHNLSK 1f. 158

Figure 4.71 Mass spectrum of native peptide STSSSHNLSK 1f. 158

Figure 4.72 MS/MS spectrum of STSSSHNLSK 1f. 159

XV
Figure 4.73 Extracted ion chromatogram of native peptide AYEMWSFHQR 1g. 160

Figure 4.74 Mass spectrum of native peptide AYEMWSFHQR 1g. 160

Figure 4.75 MS/MS spectrum of native peptide AYEMWSFHQR 1g. 161

Figure 4.76 Extracted ion chromatogram of native peptide WSKFR 1h. 162

Figure 4.77 Mass spectrum of native peptide WSKFR 1h. 162

Figure 4.78 MS/MS spectrum of native peptide WSKFR 1h. 163

Figure 4.79 Extracted ion chromatogram of native peptide YSKFR 1i. 164

Figure 4.80 Mass spectrum of native peptide YSKFR 1i. 164

Figure 4.81 MS/MS spectrum of native peptide YSKFR 1i. 165

Figure 4.82 Extracted ion chromatogram of native peptide PSKFR 1j. 166

Figure 4.83 Mass spectrum of native peptide PSKFR 1j. 166

Figure 4.84 MS/MS spectrum of native peptide PSKFR 1j. 167

Figure 4.85 Extracted ion chromatogram of native peptide GSKFR 1k. 168

Figure 4.86 Mass spectrum of native peptide GSKFR 1k. 168

Figure 4.87 MS/MS spectrum of native peptide GSKFR 1k. 169

Figure 4.88 Extracted ion chromatogram of native peptide ISKFR 1l. 170

Figure 4.89 Mass spectrum of native peptide ISKFR 1l. 170

Figure 4.90 MS/MS spectrum of native peptide ISKFR 1l. 171

XVI
LIST OF TABLES

Chapter 2 Cysteine Modification of Peptides and Proteins using isolated Electron-

Deficient Allenes

Table 2.1 Optimization of modification reaction conditions. 46

Table 2.2 Screening of aqueous solutions of the modification. 48

Table 2.3 Investigation of the cysteine selectivity. 51

Table 2.4 Investigation of the scope of electron-deficient allenes. 54

Table 2.5 Modification of cysteine-containing peptide STSSSCNLSK 1b with

electron-deficient allenes. 56

Table 2.6 Modification of cysteine-containing protein BSA with electron-deficient

allenes. 60

Chapter 3 Cysteine Modification of Peptides and Proteins using in situ Generated

Electron-Deficient Allenes

Table 3.1 Optimization of modification conditions. 72

Table 3.2 Modification of cysteine-containing peptide AYEMWCFHQR 1a with in

situ generated electron-deficient allenes. 77

Table 3.3 Modification of cysteine-containing peptide STSSSCNLSK 1b with in situ

generated electron-deficient allenes. 79

XVII
LIST OF ABBREVIATIONS

keto-ABNO 9-azabicyclo[3.3.1]nonane-3-one-N-oxy

AcOH Acetic acid

ADCs Antibody-drug conjugates

BBS Borate buffered saline

BSA Bovine serum albumin

°C Celsius

CID Collision-induced dissociation

CuAAC Copper-catalyzed azide alkyne cycloaddition

Cys Cysteine

Da Dalton

Dha dehydroalanine

DOTA 1,4,7,10-tetraazacyclodecane1,4,7,10-tetraacetic acid

DTT Dithiothreitol

equiv Equivalent(s)

ee Enantiomeric excess

2-EBA 2-ethynylbenzaldehydes

ETD Electron transfer dissociation

FBDP 4-formylbenzene diazonium hexafluorophosphate

FDA U.S. Food and Drug Administration

FRET Förster resonance energy transfer

h Hour(s)

XVIII
HMPA Hexamethylphosphoramide

His Histidine

HPLC High Performance Liquid Chromatography

IEDDA Inverse electron-demand Diels–Alder cycloadditions

iPr Isopropyl

LC–MS Liquid chromatography–mass spectrometry

Lys Lysine

min minute(s)

mM millimolar

μM micromolar

MBH Morita–Baylis–Hillman

MDA malondialdehyde

Met Methionine

MSH O-mesi-tylenesulfonylhydroxylamine

MS/MS Tandem mass spectrometry

NHS N-hydroxysuccinimide

OPA ortho-phthalaldehyde

OTf Trifluoromethanesulfonate

2-PCA 2-pyridinecarboxyaldehyde

PEG Polyethylene glycol

PLP pyridoxal-5'-phosphate

Pro Proline

PTMs Post-translational modifications

XIX
PBS Phosphate-buffered saline

Phe Phenylalanine

pSer phosphoserine

QTOF Quadrupole Time-Of-Flight

R Generic functional groups

ReACT Redox-activated chemical tagging

RGD Arginylglycylaspartic acid

RNase Ribonuclease

SDS-PAGE Sodium dodecyl sulfate polyacrylamide gel electrophoresis

Ser Serine

SPAAC Strain-promoted azide alkyne cycloaddition

TCEP Tris(2-carboxyethyl)phosphine

TMV Tobacco mosaic virus

TNF Tumor necrosis factor

Tris Tris(hydroxymethyl)aminomethane

Trp Tryptophan

Tyr Tyrosine

UAAs Unnatural amino acids

UCB Union Chimique Belge

UV Ultraviolet

XX
Chapter 1 Introduction

1.1 Allene Chemistry

Allenes are the simplest class of cumulenes with the general formula R2C=C=CR2,

in which two π bonds are orthogonal to each other with an ideal C–C–C angle of 180o.

For an unsubstituted allene, the central carbon is sp-hybridized while the two terminal

carbons are sp2 hybridized. The remaining two p orbitals on the central carbon overlap

with the p orbitals on the terminal carbons to produce two π bonds that are

perpendicular to each other (Figure 1.1). The overall configuration of allenes resembles

to an elongated tetrahedron.1 Therefore, allenes can be chiral when there are having two

different substituents on each of the terminal carbon atoms.

Figure 1.1 (a) Structure of allene; (b) Description of π bonding of allene.

The unique structure of allenes provides various advantages for organic synthesis.

Allenes have higher reactivity compared with their alkenyl and alkynyl analogs, which

allows controllable selectivity under mild reaction conditions. Besides, the unsaturation

of allenes provides excellent flexibility and opportunity to perform multistep reactions.

The intrinsic axial chirality also allows allenes to be implemented for synthesis of

optically active compounds.2

1
Moreover, many natural products contain allene moieties. Nowadays, there are

about 150 natural products with known allenic or cumulenic structures, which shows

that allenes represent important structural elements for different classes of compounds.3

The majority of naturally occurring allenes can be divided into three classes: linear

allenes, allenic carotinoids and terpenoids, and bromoallenes, in which most of them

are chiral compounds showing interesting biological activities (Figure 1.2). This unique

property leads to the inspiration of introducing allenic moieties to pharmacologically

active compounds, such as steroids, prostaglandins and nucleosides, which can function

as enzyme inhibitors, cytotoxic or antiviral agents (Figure 1.3).

Figure 1.2 Examples of naturally occurring allenic compounds.3

Figure 1.3 Examples of allenic pharmacologically active compounds.3

2
The unique features of allenes have proven themselves to be valuable structural

building blocks of complex molecules, revealing novel applications in organic

synthesis, natural product synthesis and pharmaceutical chemistry. Therefore, it is of

importance to explore new methodologies for the synthesis and transformations of

allenes.

1.1.1 Synthesis of Allenes

In the past decades, various methods for synthesis of allenes have been developed.

Allenes can be prepared by using alkenes, conjugated enyes or alkynes as starting

materials through classical reaction types, including addition, elimination, substitution

and transition-metal catalyzed reactions.

Using alkenes as starting materials, 1,2-elimination represents a common type of

reactions for preparation of allenes (Scheme 1.1). For example, upon treatment with n-

butyl lithium, 2-bromo-3,3-difluoroallylic acetates could be converted to 1,1-

difluoroallenes through lithium-bromide exchange followed by 1,2-elimination of

lithium acetate to afford mono- and disubstituted 1,1-difluoroallenes (Scheme 1.2).4

Scheme 1.1 Synthesis of allenes by 1,2-elimination.

3
Scheme 1.2 Synthesis of 1,1-difluoroallenes via 1,2-elimination of lithium acetate.

Besides, Satoh and co-workers reported the synthesis of functionalized allenes by

the reaction of 1-chlorovinyl sulfoxides with iPrMgCl, which gave magnesium

carbenoids through sulfoxide-magnesium exchange, followed by the treatment with 2-

lithio-5-methoxyfuran (Scheme 1.3a).5a When the magnesium carbenoids were trapped

by lithium enolates of α-chloroalkanoates, fully substituted 2,3-allenonates could be

obtained (Scheme 1.3b).5b Moreover, Ogasawara et al. developed a SN2’ substitution

of 2-bromo-1,3-dienes with soft nucleophiles in the presence of catalytic

palladium/bisphosphine complexes, which resulted in the formation of α-functionalized

allenes. With the use of chiral phosphine on the palladium center, (alkylidene-p-

allyl)palladium species served as a key intermediate for highly enantioselective

preparation of axial chiral allenic compounds (Scheme 1.4).

Scheme 1.3 Synthesis of allenes conjugated with α,β-unsaturated esters with

magnesium alkylidene carbenoids.

4
Scheme 1.4 Palladium-catalyzed SN2' substitution of 2-bromo-1,3-dienes.

Conjugated enynes have also been widely used for allene synthesis, in which 1,4-

addition represented an efficient approach for preparation of multi-substituted

functionalized allenes. For example, Zhang and co-workers reported a highly

functionalized 1,2-allenes could be produced base-catalyzed Michael addition to

electron-deficient 1,3-conjugated enynes (Scheme 1.5).6a The regionselectivity of the

reaction was controlled by the choice of nucleophiles. This approach was further

developed to Pd-catalyzed three-component tandem Michael addition/cross-coupling

reactions, in which the reaction between electron-deficient enynes, nucleophiles and

aryl halides resulted in the formation of tetrasubstituted allenes (Scheme 1.6).6b

Scheme 1.5 Synthesis of allenes by base-catalyzed Michael addition.

5
Scheme 1.6 Synthesis of allenes by Pd0-catalyzed three-component tandem Michael

addition/cross-coupling reaction.

Besides, Ma and co-workers developed a highly selective protocol for preparation

of 1,1-diaryl allenes or 1,3-diaryl allenes by a tunable 1,3-lithium shift of

propargylic/allenylic lithium formed by the conjugate addition of 4-aryl-3-butyn-1-

enes, followed by sequential transmetalation and Pd(0)-catalyzed cross coupling with

aryl halides.7 The regioselectivity of the reaction was controlled by the addition of

HMPA, which promoted complete 1,3-lithium shift of the formed organolithium

species (Scheme 1.7).

Scheme 1.7 Synthesis of highly selective allenes through tunable 1,3-lithium shift.

Alkynes, which are the isomers of allenes, are the most widely used starting

materials for synthesis of allenes. Isomerization of alkynes in the presence of base is

one of the earliest methods for synthesis of allenes (Scheme 1.8).8 Generally, treatment

6
of strong bases such as alkali metal amide with simple alkynes at high temperatures

could lead to the formation of allenes. For alkynes coordinated to a transition metal or

consisting of α-hydrogen activated by alkene, arene, carbonyl or oxygen / nitrogen /

sulfur / phosphorus-containing groups, weaker bases could be used for the

isomerization.

Scheme 1.8 Isomerization of alkynes to allenes in the presence of base.

SN2’-substitution reaction of propargylic compounds is one of the common

methods for the synthesis of allenes and various synthetic strategies of allenes from

propargylic alcohols have been developed (Scheme 1.9). For example, haloallenes

served as useful substrates for substitution reactions and transition metal-catalyzed

cross coupling reactions. Different methods were developed for the synthesis of

haloallenes, such as the reaction of aryl-substituted propargyl alcohols with N-

halosuccinimide and triphenylphosphine developed by Bao,9a or bromination of 1-

arylpropargylic alcohol through Appel-type reaction developed by Konakahara.9b Also,

Ready and co-workers reported an efficient and stereospecific approach for

synthesizing allenes from propargylic alcohols.9c Cp2Zr(H)Cl was applied to the anti-

SN2’-type reductive substitution of the in situ generated magnesium or zinc alkoxides

of propargylic alcohols, giving allenes with high optical purities.

7
Scheme 1.9 Examples of allene synthesis from propargylic alcohols.

The success of propargyl alcohols in allene synthesis prompted Che and Wong to

use propargylamines for the synthesis of allenes (Scheme 1.10). In 2008, Che and Wong

reported that chiral 1,3-diphenyl propargylamines could be efficiently transformed into

chiral allenes using KAuCl4 as catalysts in CH3CN at 40 °C.10a High product yields (up

to 93% yield) and excellent enantioselectivities (up to 97% ee) were resulted.

Scheme 1.10 Gold-catalyzed synthesis of axially chiral allenes.

Furthermore, a silver(I)-mediated enantioselective synthesis of axially chiral

allenes was developed in 2010 (Scheme 1.11).10b With the optimized conditions, a

variety of axially chiral allenes were synthesized in high product yields up to 95% and

high enantioselectivities of 98–99% ee. The use of microwave irradiation allowed the

8
completion of the reaction in a much shorter time than that required under conventional

thermal conditions.

Scheme 1.11 Silver(I)-mediated synthesis of axially chiral allenes.

Recently, our group discovered a novel electrophile-mediated cascade

cyclization/iodination of propargylamine-based 1,6-diynes to generate spiro 1H-

isoindoliums, which can be transformed to the corresponding 1H-isoindolium-based

electron-deficient allenes upon treatment with triethylamine. To the best of our

knowledge, the synthesis of these kinds of electron deficient allenes has not been

reported.

H PF6 H PF6
R2 R2 I R2 2 I R2 2
N R R
1) I2 (1.05 equiv.), CH3CN, r.t., 1 h N Et3N (1 equiv.) N

CH2Cl2, 1 h R1
R3 2) AgPF6 (1.1 equiv.), r.t., 5 mins R1
R1
H
R3 R3
Propargylamine-based
1H-isoindoliums 3 Electron-deficient allenes 4
1,6-diynes 2

Scheme 1.12 Electrophile-mediated cascade cyclization / iodination of

propargylamine-based 1,6-diynes and transformation into 1H-isoindolium-based

electron-deficient allenes.

9
1.1.2 Reactions of Allenes

The high degree of unsaturation and a readily accessible π-bond system allow

allenes to undergo various types of reactions, such as electrophilic addition,

nucleophilic addition, radical addition and transition metal-catalyzed reactions. Allenes

can act as nucleophiles or electrophiles, depending on the substituents at the terminal

carbons (Scheme 1.13).

Scheme 1.13 Property of allenes with different substituents.

For electrophilic addition, allenes undergo the usual electrophilic addition

reactions, such as halogen and addition of hydrogen halides, but the regioselectivity is

different from the generally accepted order of cation stability (Scheme 1.14).

The protonation of an unsubstituted allene can lead to either the 2-propenyl cation

or the allyl cation. The resulting product for the addition of HBr to an unsubstituted

allene is 3-bromopropene, suggesting that the formation of an intermediate vinyl

carbocation is more favorable, as opposed to the relatively higher stability of allylic

carbocation than vinylic or primary carbocation. This regiochemistry can be explained

by the fact that the two adjacent π-bonds in the allene are perpendicular to each other.

For formation of stabilized allyl cation in allene, a 90° rotation of the C–C bond joining

the carbocation and the double bond should take place, which requires higher activation

10
energy.11 As a result, the formation of 2-propenyl cation by protonation at a terminal

carbon is kinetically favored. However, the presence of stabilizing groups, such as

phenyl and dialkyl groups, at terminal carbon atom of the allene can change the

regiochemistry.

Scheme 1.14 Regiochemistry of electrophilic addition of allenes.

For nucleophilic addition, the presence of electron-withdrawing groups leads to

the electron deficiency of the inner C=C double bond, which induces nucleophilic

addition at the central carbon atom. Generally, the addition of nucleophiles, such as

alcohols, phenols and thiols, is carried out in the presence of base and results in the

formation of two types of products (Scheme 1.15).12

In early examples of nucleophilic addition to electron-deficient allenes, the

formation of the non-conjugated product C is kinetically-controlled, while the

conjugated product D is the result of a thermodynamically-controlled reaction.12 Firstly,

the intermediate A is formed after nucleophilic attack on the central carbon atom of

allene, which requires a torsion of 90° to merge into the allylic carbanion B.

Intermediate A can only lead to product C by proton transfer, while the protonation of

B can result in both C and D.

11
Scheme 1.15 General mechanism of nucleophilic addition to allenes12

For radical addition of allenes, addition of different types of radicals, such as

carbon-, nitrogen-, phosphorus-, sulfur-and selenium-centered, to allenes has been

experimentally and theoretically studied.13 The intermolecular addition of a radical to

an unsymmetrically substituted allene can occur at the two terminal carbon atoms or

the central carbon (Scheme 1.16).

Radical attack at C1 or C3 results in a pair of radical intermediates III and IV. The

vinyl radical intermediate II or IV will isomerize from a π-type configuration to the

more energetically favored σ-type radicals III and V. Radical attack on C2 results in a

π-type alkyl radical VI or VIII, which rotates around the C2–C3 bond of VI or the C1–

C2 bond of VIII for the formation of resonance-stabilized allylic radical VII. The

preferred site for radical addition to allene depends on three factors: (1) the degree of

substitution; (2) the nature of the attacking radical and (3) reaction parameters such as

temperature and initial reactant concentrations.12 In general, selective α-addition to

12
allene occurs for unsubstituted allene, regardless of the nature of the radical, while

substituted allenes favor β-addition of radicals, resulting in the formation of stabilized

allylic intermediates.

Y R3
R3
X
X R4
R4 R2 R1
trapping
R2 R1 reagent Y
R1 R3
1 2 3 R3 X
a b g R4 R2 R1 R4 Y R4
X R2 R4
I II X X
R3 R3
R2 R1 R2 R1

III

R1 R1 Y

X X
R2 R2
R1 R3 R4 trapping R3 R4
R1
1 2 3 R3 X reagent Y
R4 R4
R2 R2 R3 R2
X R2 Y
IV X
I R1 X
4 R1
R3 R R3 R
4

R3 R 3 R4
R1 R1 R4
R1 R1
1 2 3 R3 R3 Y
R4
R4 R2 X
R2 R2 X trapping R2 X
X reagent Y
I VI
1
R1 R2 R
R1 R1 R3
1 2 3 R3 R2 R3 Y
R3
R4 R2
R2 X R4 X R4 X R4
X
I VII
VIII

X = C, F, Cl, Br, N, O, S, Se, Sn, etc.; Y = H, I, Se, S, etc.

Scheme 1.16 Regio- and stereoselectivity in intermolecular radical addition

of allenes.13

Transition metal-catalyzed transformations of allenes have also been widely

developed due to their high π-coordination ability towards transition metals. In

principle, transition metal-catalyzed cross-couplings of allenes with halogen or metal

substituents at one of the terminal carbons or substitution reactions of α-functionalized

allenes give products with unchanged cumulene π-system (Scheme 1.17a).12 Transition

13
metal-catalyzed couplings at the central carbon are often accompanied by subsequent

reactions which rule out the cumulene system in the final products (Scheme 1.17b).

Scheme 1.17 Types of reaction modes of transition metal-catalyzed

coupling of allenes.12

Among the transition metal-catalyzed reactions of allenes, palladium-catalyzed

reactions have gained considerable attention over the last two decades because of their

wide applicability in organic synthesis. Allenes readily undergo carbopalladation in two

pathways depending on the regiochemistry of the insertion reaction. In Path I, the

reaction produces a sp2 C-Pd species, while Path II results in a π-allylpalladium species

(Scheme 1.18a).14

a) b) R
R1
Pd
R3
Path I
Pd(0) + RX b- H elimination
R R2 R4
R Pd

R HPdX
R
RPdX R1 CH2R3 base
Path II
Pd R
+ R2 Pd R4
Pd Pd(0)
R1 CH2R3 L L

R
R2 R4 R1
CH2R3
Nu R2
Nu R4
+
R
R1 CH2R3

R4
R2 Nu

Scheme 1.18 Carbopalladation of allenes.14

14
In most cases, Pd firstly connects to RX, followed by insertion reaction into allene

to form a π-allylpalladium species which can undergo β-hydride elimination or

nucleophilic substitution (Scheme 1.18b).14 Apart from intermolecular reactions,

allenes with a nucleophilic functionality can undergo Pd-catalyzed cyclization to form

cyclic skeletons, which provides possible pathways for synthesis of natural products

(Scheme 1.19).15

Scheme 1.19 Pd-catalyzed cyclization of allenes.15

15
1.2 Chemical Modification of Peptides and Proteins

Proteins are a major class of biopolymers that play important roles in different

biological events in living organisms. Therefore, study of proteins is one of the

important research areas in science. Post-translational protein modifications (PTMs),

which refers to the modification of proteins after transcription and translation, is often

responsible for the diverse structures of proteins found in nature. Naturally occurring

protein PTMs play important roles in tuning the properties of proteins and modulating

their biological functions.16

Similar to the natural approach of expanding the function of biomolecules through

PTMs, the development of novel bioconjugation approaches provides new

opportunities for chemical biologists to modulate structure and function of

biomolecules through synthetic chemistry. Therefore, chemical modification of

peptides and proteins has emerged as an invaluable tool for biological studies and

development of targeted therapeutics. For example, fluorescent labeling of proteins

allows in vitro and in vivo imaging analysis of the structure, function and localization

of proteins in biological environments.17 The addition of polyethylene glycol (PEG)

can increase the circulation half-life of a therapeutic protein.18 Antibody–drug

conjugates (ADCs) consisting of a drug linked to a tumor-selective antibody have also

recently shown considerable promise as anticancer agents.19

Despite a wide range of modification methods developed, the progress in

bioconjugation is still limited. In comparison with organic transformations which can

be conducted in organic solvent and under comparatively harsh conditions, chemical

16
reactions for protein modification must tolerate biological ambient conditions at or

below 37 °C with moderate pH (pH 6–8) in aqueous solvent. Unlike standard reaction

conditions of organic synthesis at tens to hundreds mM concentrations, protein

modifications are generally conducted at μM concentrations or less. Therefore,

chemoselective synthesis in conventional organic chemistry cannot be generally

applied to modification of biomolecules. Another challenge for protein modification is

to achieve modification with high efficiency and site selectivity. In order to avoid

purification steps, modification methods at a single site with high conversion and

minimal generation of side products are desired.20, 21 Therefore, it is of great importance

to develop complementary reactions for site-selective protein modification.

1.2.1 Lysine and N-Terminal Modification

The use of chemical reagents that react with primary amines is one of the most

versatile techniques for peptide and protein modification. There are two groups of

primary amino groups in protein: the α-amino group on the N-terminus of polypeptide

chains and ε-amino groups on lysine residue (Lys, K). The N-terminal amine has pKa

≈8 while that of a lysine ε-amine has pKa ≈10, so they are protonated at physiological

pH and predominantly exposed on the surface protein tertiary structures, resulting in

easy accessibility to the modification reagents. Deprotonated primary amines are highly

nucleophilic which favor the modification in peptides and proteins. However,

protonation can significantly reduce their reactivity. Therefore, it is important to control

the pH when conducting the N-terminal or lysine modifications.

17
Depending on the reaction conditions, selective modification of either N-termini

or lysine residues can be achieved by using different chemical reagents. Classical

approaches of modifying the amino groups in protein includes using activated esters,

isothiocyanates and through reductive amination of aldehydes (Scheme 1.20). Among

these methods, N-hydroxysuccinimide activated esters represent the most popular

amine-specific reagents for protein bioconjugation and labelling. NHS-mediated

modifications work within mild pH ranges from 7 to 8, which is lower than that for

isothiocyanates (pH 9–9.5), which is favorable for modifying alkaline-sensitive

proteins. Although the possible side reactions between NHS-activated esters with

tyrosine, histidine, serine and threonine were reported, the higher reaction rate of NHS

towards free amine groups overcomes the drawbacks of these side reactions.22

Scheme 1.20 Classical methods of lysine modification.

Apart from the classical methods, various novel approaches for lysine

modification have been developed in the past decades. For example, Tanaka and co-

18
workers developed a novel lysine-based labeling of biomolecules based on a rapid 6π-

azaelectrocyclization (Scheme 1.21).23 By reaction with unsaturated aldehyde probes,

1,4,7,10-tetraazacyclodecane1,4,7,10-tetraacetic acid (DOTA) as a metal chelating

agent and fluorescent groups could be efficiently introduced to lysine residues. The

reaction was selective to lysine residues at the protein surfaces, while it was less

sensitive to internal lysines and N-terminal amines.

Scheme 1.21 Lysine modification through 6π-azaelectrocyclization.

In 2016, Li and co-workers reported the chemoselective lysine modification

methods using ortho-phthalaldehyde (OPA) and its derivatives through the formation

of phthalimidines (Scheme 1.22).24 This approach could be applied to generation of

PEGylated L-asparaginase with resistance to proteolytic degradation by trypsin and

Glu-C and retained 70% and 40% of activities, respectively.

Scheme 1.22 Lysine modification using ortho-phthalaldehyde (OPA).

In 2018, Bernardes and co-workers demonstrated chemo- and regioselective

modification of lysine using a sulfonyl acrylate reagent, which was highly selective to

19
the ε-amino group of the most reactive lysine in the presence of other nucleophilic

amino acids such as cysteine (Scheme 1.23). The chemoselectivity of the modification

was attributed to the transient hydrogen bonding of the lysine amine with the sulfone

moiety of the reagents. This reaction was successfully employed to proteins with

diverse structures, including a full-length IgG antibody.25

Scheme 1.23 Lysine modification using sulfonyl acrylate reagents.

In 1960s, an N-terminal conjugation method through transamination was

developed by Dixon and co-workers.26 Through transamination of the terminal residue,

a carbonyl group could be introduced into proteins for further functionalization by

formation of hydrazone or oxime bonds.27 However, this method was not applicable

due to the harsh reaction conditions. Francis and co-workers re-examined this approach

and developed the pyridoxal-5'-phosphate (PLP)-mediated transamination for

modifying the N-terminal residues of peptides (Scheme 1.24).28a By introducing

reactive ketone or aldehyde groups, further functionalization could be proceeded

through oxime or hydrazone formation. The modification method was highly efficient

and selective to the N-terminus of peptides and proteins. However, it was found that

there were side reactions between some residues (His, Trp, Lys, and Pro) and PLP.

20
Scheme 1.24 N-terminal modification through PLP-mediated transamination

followed by oxime ligation.

Considering the lower pKa of N-terminal α-amino group (pKa ~ 8) than that of

lysine ε-amino groups (pKa ~ 10), pH-controlled N-terminal selective modification

methods were developed. In 2006, Che and Wong discovered an N-terminal α amino

group ligation of peptides with alkyne moieties using [Mn(2,6-Cl2TPPCl)] as catalyst

and Oxone as terminal oxidant (Scheme 1.25).29a Detailed studies revealed that the in

situ-generated ketenes were the key intermediates for the modification, which could be

isolated for direct selective N-terminal modification.29b

Scheme 1.25 N-terminal modification of peptides with ketenes.

In 2015, Francis and co-workers developed a promising approach for one-step N-

terminal selective modification of proteins using 2-pyridinecarboxyaldehyde (2-PCA)

derivatives (Scheme 1.26).28b The reaction between 2-PCA and the N-terminal residue

21
formed an N-terminal imine, followed by a nucleophilic attack of the adjacent amide

nitrogen on its electrophilic carbon, resulting in the formation of cyclic imidazolidinone

products.

In 2017, Chou and co-workers reported an N-terminal modification method of

peptides and proteins through reductive alkylation with aldehyde derivatives (Scheme

1.27).30 The modification was efficient with all N-terminal residues with an exception

of N-terminal cysteines, which resulted in thiazolidine derivatives. The modified

peptide and protein derivatives with bioorthogonal groups allowed further

functionalizations, such as fluorophores or biotins, for biological studies.

Scheme 1.26 N-terminal modification using 2-PCA derivatives.

Scheme 1.27 N-terminal modification using aldehyde derivatives.

Most recently, our group reported an N-terminal selective modification method of

peptides and proteins using 2-ethynylbenzaldehydes (2-EBA).31 By performing the

reaction in slightly acidic phosphate buffer using 2-ethynyl-4-hydroxy-5-

methoxybenzaldehyde, modification of a library of peptides XSKFR (X = either one of

22
20 natural amino acids) resulted in good-to-excellent N-terminal selectivity up to >99:1

(Scheme 1.28). N-terminal modification was also applied to protein modification,

including lysozyme, ribonuclease A and a therapeutic recombinant Bacillus caldovelox

arginase mutant (BCArg mutant).

Scheme 1.28 N-terminal selective modification of peptides and proteins

using 2-EBA.

1.2.2 Modification of Other Natural Amino Acids

Tyrosine (Tyr, Y) is a choice of amino acid for site-specific modification of

proteins because of its lower abundance on the surface of proteins compared with Lys

(Scheme 1.29). In 2004, Francis and co-workers have demonstrated the chemoselective

azo coupling reaction using diazonium compounds to modify Tyr residues on the

exterior and interior surfaces of the tobacco mosaic virus (TMV) and bacteriophage

MS2 respectively.32a-b In 2006, they reported the use of π-allylpalladium complexes for

selective Tyr O- alkylation with a rhodamine dye and lipophilic moieties.32c The same

group also developed a three component Mannich-type reaction between Tyr residues,

aldehyde and aniline derivatives, which successfully modified solvent-exposed Tyr in

lysozyme, RNase A, and chymotrypsinogen A.32d-f

23
Apart from Francis’s group, Barbas and co-workers also developed two Tyr

modification strategies. In 2010, they reported an aqueous ene-type reaction for Tyr

modification using cyclic diazodicarboxamides, in which therapeutic antibody

herceptin was successfully labeled with an integrin binding cyclic RGD peptide under

optimized reaction conditions.33a In 2012, a novel 4-formylbenzene diazonium

hexafluorophosphate (FBDP) was developed for tyrosine-selective modification of

peptides and proteins, which was also applicable to labeling of the antibody

trastuzumab and cell surface labeling.33b

Scheme 1.29 Examples of tyrosine modification.

24
Tryptophan (Trp, W) is another for site-selective modification as it is a low

abundance amino acid with 1.1% frequency but exists in the primary sequence of

approximately 90% of native proteins (Scheme 1.30).34, 35 Some metal-mediated and

transition metal free modifications of Trp have been reported. For example, Francis and

co-workers reported rhodium carbenoids for Trp modification in 2004, yielding N-

alkylated and 2-alkylated products under acidic conditions (pH 1.5–3.5).36a The

protocol was improved to be performed at mild pH (pH 6–8) by replacing H2NOH

buffer with tert-BuNHOH.36b In 2007, Lindner and co-workers reported a condensation

reaction of malondialdehyde (MDA) with the indole nitrogen of tryptophan residues in

peptides under strongly acidic conditions.37 The resulting acroleinyl label at the indole

nitrogen could be further converted to hydrazone using hydrazide compounds, while

hydrazines led to cleavage of MDA adduct and releasing the free indole group.

In 2016, Kanai and co-workers reported a transition-metal free Trp-selective

bioconjugation method using an organic radical compound, 9-

azabicyclo[3.3.1]nonane-3-one-N-oxy (keto-ABNO) with NaNO2.38 Treatment of

peptides with the reagents in acidic aqueous solutions (0.1 – 0.5% AcOH) at room

temperature within 30 min gave keto-ABNO adducts which resulted from C–O bond

formation at the 3-position of Trp residues. This strategy was further applied to

modifications of lysozyme, myoglobin, concanavalin A, BSA and β2-microglobulin

antibody.

25
Scheme 1.30 Examples of tryptophan modification.

Methionine (Met, M) is the second rarest amino acid in proteins and has limited

surface-accessibility. Therefore, it is considered as a potential handle for highly

selective modification.39 There are limited numbers of Met modification methods

reported (Scheme 1.31). In 2012, Deming and co-workers developed a “click”-type

reaction for modification of methionine-containing peptides based on the

chemoselective alkylation of thioether groups in Met.40a This method allowed

introduction of a broad range of functional groups, such as carbohydrates, alkenes and

alkynes. The resulting sulfonium product could be reversibly dealkylated by addition

of nucleophiles, such as 2-mercaptoethanol, thiourea, 2-mercaptopyridine and

glutathione.40b

In 2017, Toste, Chang and co-workers developed a general and versatile Met-

selective bioconjugation method using redox-activated chemical tagging (ReACT).41

26
The modification was based on the oxidative sulfur imidation reaction between Met

and oxaziridine derivatives. The reagents enabled specific Met labeling from a single

protein to whole proteome level under mild biocompatible conditions. The broad utility

of the ReACT technology was demonstrated by synthesis of antibody–drug conjugates

and identification of hyper-reactive Met residues in whole proteomes.

Scheme 1.31 Examples of methionine modification.

1.2.3 Modification of Unnatural Amino Acids

The incorporation of unnatural amino acids (UAAs) to protein allows introduction

of chemical handles for subsequent bioorthogonal reactions. Different functional

groups, such as azide, alkyne, alkene and tetrazine can be chemically or genetically

incorporated in biomolecules, which can then further conjugate with specially designed

bioorthogonal probes.42 Various bioorthogonal reactions to target biomolecules have

been developed, including Staudinger ligation, copper-catalyzed azide alkyne

27
cycloaddition (CuAAC), strain-promoted azide alkyne cycloaddition (SPAAC), inverse

electron-demand Diels–Alder cycloadditions (IEDDA) and photoclick cycloaddition

(Scheme 1.32).

Scheme 1.32 Bioorthogonal reactions for protein labeling.

The Staudinger ligation between azides and triarylphosphines was demonstrated

by Bertozzi and co-workers.43a This technique was initially used for modification of

azido-glycoproteins on the surface of live cells, which was then further extended to the

modification of azide-incorporated proteins.43b-c Another example of azide

functionality was presented in CuAAC developed by Sharpless43d and Meldal, 43e which

was the Huisgen 1,3-dipolar-cycloaddition with alkynes in the presence of Cu(I)

catalysts. Because of the high specificity and fast reaction rate, CuAAC has been widely

28
used for protein modification. However, cytotoxicity caused by Cu(I) metal has

remained a limitation for modification of live cells, which led to the exploration of

alternative cycloaddition-type reactions. In 2004, Bertozzi’s group developed SPAAC,

in which highly strained cyclooctynes could react with azide-‘tagged’ glycoproteins at

room temperature without ligands or catalysts.44a After further studies for improving

the reaction rates, SPAAC has been broadly used in live mammalian cells and

animals.44b-d

Besides, Fox and co-workers developed the use of IEDDA reactions for

bioconjugation, which involved the cycloaddition between tetrazine dienophiles and

dienes.45a Because of the extremely fast reaction rate and high selectivity, IEDDA has

been widely used for modifications and fluorescent labeling of proteins in live

mammalian cells.45b-d

An alternative approach of cycloaddition-type modification of proteins has been

reported by Lin. Tetrazole-modified protein can undergo [3+2]-cycloadditions with

unactivated alkenes under irradiation with UV light.46a Attachment of alkenes to

proteins followed by treatment with tetrazoles proceeded with similar efficiency, which

has been used to for modification of alkenyl-UAAs such as homoallylglycine46b and

cyclopropenes.46c

Most recently, our group reported a novel approach for selective modification of

alkyne-linked peptides and proteins through alkynylation reaction of 6-membered ring

cyclometalated gold(III) complexes [Au(C^N)Cl2] with terminal alkynes (Scheme

1.33).47 After insertion of terminal alkyne to peptide, modification using 6-membered

29
ring cyclometalated gold (III) complexes afforded alkynylation product in > 99%

conversion. This approach could be further applied to fluorescent labeling of alkyne-

linked proteins using dansyl and BODIPY-linked gold(III) complexes.

Scheme 1.33 Chemoselective modification of alkyne-linked peptides and proteins by

cyclometalated gold(III) (C^N) complex-mediated alkynylation.

1.3 Cysteine Modification

Among the 20 natural amino acids, cysteine (Cys, C) residue has a unique

reactivity. Under physiological conditions, it has a high propensity to form the

nucleophilic thiolate ion (Scheme 1.34).48 The high nucleophilicity of the thiol moiety

offers advantages for developing modification approaches with high efficiency and

excellent site-selectivity.

Scheme 1.34 Formation of thiolate in cysteine residues.48

30
Conventional protein bioconjugation methods rely on reactions at nucleophilic

amino acids, particularly lysine or cysteine residues. However, the high abundance of

lysine residues on protein surfaces leads to the difficulties in the development of

selective labeling of a single lysine residue. In comparison, cysteine has a relatively

low natural abundance (~1.9%) and can be readily introduced by direct mutagenesis,

which allows the control of modification at a predetermined site.18 Therefore, cysteine

is considered as an ideal residue for selective modification and wide range of cysteine

modification strategies have been developed over the past decades.

Cysteine can be easily modified with suitable electrophiles, such as α-

halocarbonyls and Michael acceptors. However, some of the resulting bioconjugates

suffer the stability problems, in which they may undergo retro-Michael additions and

other thiol-promoted exchange reactions.17 These limitations prompt the development

of a number of metal-free and transition metal-mediated cysteine modifications.

Moreover, cysteine-selective functionalization of peptides and proteins occupies a

remarkable place in the field of bioconjugation as all of the recently FDA approved

ADCs are developed based on conjugation with cysteine residues in the polypeptide

chain.49 Therefore, it is of importance to explore new strategies of cysteine modification

for the generation of functional bioconjugates.

31
1.3.1 Classical Methods

Classical methods of cysteine modification include substitution with haloalkyl

substrates, Michael Addition with maleimides, disulfide formation and reaction of Dha

with thiols.

One of the earliest methods for modification of cysteine residues is by SN2

substitution with α-halocarbonyl and haloalkyl substrates (Scheme 1.35). Particularly,

reagents with iodide and bromide groups are commonly used. As early as 1935,

iodoacetamide was used to modify and study cysteine residues in proteins.50 A variety

of examples using such reagents for alkylating or arylating cysteine residues in proteins

have been reported over the past decades. For example, Davis and co-workers reported

that the p-iodobenzyl cysteine moiety on peptides and proteins could act as a coupling

partner or Suzuki–Miyaura coupling with phenylboronic acids.51 However, it was

reported that the haloalkyl substitution has a potential cross-reactivity with other amino

acids containing nucleophilic side chains, such as lysine and histidine.

Scheme 1.35 Modification of cysteine through nucleophilic substitution.

32
Another conventional approach of cysteine modification is the use of maleimides

(Scheme 1.36). Maleimides are Michael acceptors, which react with cysteine thiolates

to form thiosuccinimide bonds. The conjugation reaction between maleimide and the

cysteine residue in proteins was first reported by Moore and Ward in 1956, which

described the use of bis-maleimides as crosslinking agents of bovine plasma albumin

and wool keratin. To date, the use of maleimides still remains the preferred method of

cysteine modification of proteins for a number of reasons. Maleimides are high

selectivity towards thiol groups and the maleimides-thiols coupling reaction can be

performed in biocompatible conditions. Moreover, the maleimide moiety can be easily

functionalized with different conjugation partners such as fluorophores. This

conjugation reaction is also widely used in medicinal chemistry, in which antibody-

based therapeutics formed by maleimide-thiol coupling is a rapidly growing class of

biotherapeutic drugs. For example, Cimzia, a PEGylated anti-TNF construct developed

by UCB, has been approved by FDA as therapeutics drugs for Crohn’s disease and

arthritis.52 Despite its wide applications in research and industry fields, it still remains

problem of instability of thiosuccinimide adducts, which are known to undergo retro-

Michael additions or other thiol exchange reactions under physiological conditions.

Scheme 1.36 Modification of cysteine with maleimide reagents.

33
Disulfide bond formation is important for folding and maintaining tertiary

structural integrity in proteins. It has also been used as a common approach for chemical

modification of cysteine-containing proteins with other thiol-containing moieties. The

formation of disulfide bonds can be simply promoted by air oxidation, but have a

drawback of protein dimer formation. To prevent this, activated reagents RSX (X =

SO2R’, SeR’, I, Br, Cl) are usually used (Scheme 1.37). For example, Ellman reagent,

5,5’-dithiobis(2-nitrobenzoic acid), is a common reagent for activating cysteine thiols

on proteins,53 which have also been used for antibody–drug conjugates formation.54

Scheme 1.37 Common methods for disulfide formation.

Cysteine as a precursor to dehydroalanine (Dha) is of high importance for

chemical protein modification. Dha is a naturally occurring amino acid formed by

serine (Ser) dehydration or phosphoserine (pSer) elimination. There are various

methods for inserting dehydroalanine into peptides and proteins using serine as a

precursor. However, serine has a high natural abundance and often requires harsh

conditions for modifications, which causes difficulties in site-selectivity and their

applicability to a wide range of proteins.55 This leads to the development of method for

34
transforming cysteine to Dha.56 For example, in 2008, Davis and co-workers reported

an oxidative amidation / Cope-type elimination of cysteine to Dha using O-mesi-

tylenesulfonylhydroxylamine (MSH) (Scheme 1.38).57 Chemically, Dha is an α,β-

unsaturated carbonyl moiety that can undergo Michael-type addition reactions with

nucleophiles, which allow further functionalization of proteins.

Scheme 1.38 Transformation of cysteine to Dha in protein using MSH.

1.3.2 Transition Metal-Free Modifications

Apart from conventional approaches, different new methods for cysteine

modification were developed.

In 2009, Che and Wong first reported the use of electron-deficient alkynes,

including alkynoic amides, esters and alkynones, to perform modification of cysteine-

containing peptides and proteins through the formation of vinyl sulfide linkages

(Scheme 1.39).58a It was proposed that the modification proceeded through the

nucleophilic addition of thiols to the electron-deficient alkynes. The reaction was

facilitated in solutions at alkaline pH (pH 7.0–9.0), presumably due to the increased

concentration of thiolate ions in alkaline solution. The vinyl sulfide linkage was stable

towards TCEP while the alkynone-modified peptides could be cleaved to regenerate

the unmodified peptides by treatment with thiols.

35
Scheme 1.39 Modification of cysteine-containing peptides and proteins using

electron-deficient alkynes.

Based on this thiol-specific cleavage reaction, this cysteine modification was

further employed to develop a highly selective FRET-based fluorescent probe for

detection of cysteine and homocysteine.58b-c The probe was able to be applied to cellular

imaging. Besides, the introduction of luminescent iridium(III) complexes to the FRET-

based probes for the detection of thiols was also demonstrated.58d

In 2012, our group reported the amine-catalyzed Morita–Baylis–Hillman (MBH)

bioconjugation reaction (Scheme 1.40). With the formation of the β-hydroxyl-α-

methylene-carbonyl moiety, multifunctional modification of cysteine-containing

peptides and proteins with fluorescent probes/ biotin tags could be performed in slightly

basic aqueous media (pH 8.0).59

Scheme 1.40 Multifunctional modification of cysteine-containing peptides and

proteins by Morita–Baylis–Hillman reaction.

36
In 2016, Pentelute and co-workers reported a π-clamp-mediated site-selective

conjugation with perfluoroaromatic reagents (Scheme 1.41).60 A four-amino-acid

sequence (Phe-Cys-Pro-Phe) called the “π-clamp” tuned the reactivity of a cysteine

thiol to recognize the perfluoroaromatic reaction partner, which allowed selective

modification of one cysteine residue in proteins with multiple endogenous cysteine

residues, including antibodies and cysteine-containing enzymes.

Scheme 1.41 π-Clamp-mediated site-specific cysteine conjugation.

In 2019, Bernardes and co-workers developed an efficient and irreversible

cysteine-selective bioconjugation method through quaternization of vinyl- and alkynyl

pyridines (Scheme 1.42).61 Quaternization of the nitrogen of vinyl and alkynyl

pyridines transformed these molecules to reactive electrophiles towards thiols, allowing

selective bioconjugation of cysteine-containing protein. Also, the introduction of a +1

charge to the overall net charge of proteins by the quaternized vinyl pyridinium reagent

could modulate the electrophoretic mobility of proteins.

Scheme 1.42 Cysteine bioconjugation method through quaternization of vinyl- and

alkynyl pyridines.

37
1.3.3 Transition Metal-Mediated Modifications

Transition metal-mediated transformations are widespread in organic synthesis.

Due to the stringent requirements for bioconjugation and the existence of multiple

functional groups in biomolecules, only a few examples of transition-metal based

bioconjugation methods have been reported. Cysteine residue has metal binding ability,

in which the thiolate groups are capable of binding to metal ions such as iron, zinc and

cadmium.62 This unique property leads to the development of metal-mediated

modifications of cysteine.

Transition metals can serve as mediators for selective modifications at cysteine.

In 2014, our group first reported a gold-mediated cysteine modification method using

cyclometalated gold(III) complexes with bidentate C,N-donor ligands and ancillary

ligands (Scheme 1.43).63a The cyclometallated gold(III) complexes could selectively

modify the cysteine residues of peptides to form gold–cysteine adducts in > 90%

conversion in 2 h at room temperature. Increasing the temperature to 37 °C resulted in

reductive elimination of the gold–cysteine adducts to give alkylated cysteine.

Scheme 1.43 Cysteine modification using cyclometalated gold(III) complexes via

ligand controlled C–S bond formation.

38
In 2015, Buchward, Pentelute and co-workers reported a cysteine modification

method using organometallic palladium reagents (Scheme 1.44).64 Treatment of low

micromolar concentrations of palladium(II) complexes with cysteine-containing

peptides gave C–S reductive elimination products in excellent conversions within 5 min.

This strategy could be used for functionalization of unprotected peptides, such as

fluorescent tags and affinity labels. Further applications in peptide stapling and

construction of antibody–drug conjugates (ADCs) were also demonstrated.

Scheme 1.44 Cysteine modification using palladium(II) complexes.

Most recently, our group developed a novel isoxazolinium-mediated cysteine

modification of peptides and proteins (Scheme 1.45).63b The use of a stoichiometric

amount of isoxazolinium reagents generated in situ from a catalytic amount of silver

salts can efficiently modify cysteine-containing peptides and proteins. The modified

bioconjugate was stable towards thiol-containing reagents, reducing reagents and

oxidizing reagents. Under irradiation of UV-A light, the phenylacyl thioether linkage

bearing an alkyne moiety of can be rapidly cleaved to form a thioaldehyde moiety,

which could be converted back to cysteine by reduction.

Scheme 1.45 Isoxazolinium-mediated cysteine modification of peptides and proteins.


39
Chapter 2 Cysteine Modification of Peptides and Proteins using

isolated Electron-Deficient Allenes

2.1 Introduction

In 2013, our group first reported the gold-mediated selective cysteine modification

of peptides (Scheme 2.1).65 Gold compounds were effective in activating the π-system

of allenes for addition reactions with nucleophiles. In the presence of AuCl–AgOTf, a

diversity of allenes were successfully coupled with the thiol group of cysteine-

containing peptides to afford hydroxy vinyl thioethers instead of the commonly

observed 1,2 / 1,3-adducts. Another example of thiol-allene coupling was reported by

Loh and co-workers, which utilized the C-substituted terminal allenamide moieties for

cysteine modification of peptides (Scheme 2.2).66 The strategy was based on the 1,4-

Michael addition of the thiol group of cysteine residues to allenamides. This

modification method was further applied to fluorescent labeling of proteins. Despite

these examples, comprehensive study on the applicability of allenes on protein

modification remains largely elusive.

Scheme 2.1 Gold-mediated selective cysteine modification of peptides.

40
Scheme 2.2 Cysteine modification using allenamides.

Recently, our group disclosed a novel electrophile-mediated cascade cyclization /

iodination of propargylamine-based 1,6-diynes to generate spiro 1H-isoindolium,

which can be transformed to the corresponding 1H-isoindolium-based electron

deficient allenes upon treatment with triethylamine.

Given the electrophilic allenes and strong nucleophilicity of cysteine residues in

peptides and proteins, it is envisioned that the novel electron-deficient allenes would

be a useful handle for bioconjugation of cysteine-containing peptides and proteins. It is

hypothesized that the 1H-isoindolium-based electron deficient allenes would have

excellent selectivity and reactivity in metal-free cysteine modification. A detailed study

of cysteine modification of peptides and proteins using electron-deficient allenes was

conducted.

Reaction conditions were first optimized by screening with different amounts of

reagents, buffer media and temperature. Then, detailed studies of modification products

were performed by employing electron-deficient allene derivatives and various

cysteine-containing peptides. Moreover, the stability of modified peptides was

examined by treatment with excessive thiol-containing reagents and reducing agents.

41
The optimized modification protocol was then utilized for modification of proteins

with free cysteine residues, and the modification was confirmed by chymotrypsin

digestion followed by LC–MS analysis. This method would be applied for fluorescent

labeling of proteins.

2.2 Results and Discussion

2.2.1 Modification of Cysteine-Containing Peptides Using Isolated Electron-

Deficient Allenes

2.2.1.1 Synthesis of Electron-Deficient Allenes

The new class of electron deficient allenes was synthesized by electrophile-

mediated cascade cyclization / iodination of propargylamine-based 1,6-diynes 2, in

which propargylamine-based 1,6-diynes 2 could be synthesized by gold(III)-catalyzed

three-component coupling reaction of aldehydes, amines and alkynes.67

Scheme 2.3 General scheme for synthesis of propargylamine-based 1,6-diynes 2.

42
Treatment of propargylamine-based 1,6-diynes 2 with iodine generated the spiro

1H-isoindoliums 3, which were very sensitive to base. Upon treatment with

triethylamine, the alkyne moieties was easily isomerized to the corresponding 1H-

isoindolium-based electron-deficient allenes 4.

H PF6 H PF6
R2 R2 I R2 2 I R2 2
N R R
1) I2 (1.05 equiv.), CH3CN, r.t., 1 h N Et3N (1 equiv.) N

CH2Cl2, 1 h R1
R3 2) AgPF6 (1.1 equiv.), r.t., 5 mins R1
R1
H
R3 R3
Propargylamine-based
1H-isoindoliums 3 Electron-deficient allenes 4
1,6-diynes 2

Scheme 2.4 General scheme for synthesis of electron-deficient allenes.

Besides, ICl also worked well as an electrophile affording the corresponding

products 2 and this method is practical and easily scalable. Electron-deficient allene 3a

could be also obtained from 2a with 87% yield under basic conditions in a one-pot

reaction.

Figure 2.1 The synthesis of allene 4a.

43
The reaction mechanism for this novel iodine-mediated cascade iodination /

cyclization was proposed. Electrophilic addition of iodine to the terminal alkyne of 2

gives iodonium salt A. Then, the nitrogen atom of A attacks the iodonium moiety to

afford quaternary ammonium salt B. Subsequent anion exchange with AgPF6 gives

alkyne 3. Under basic conditions, isomerization of alkyne B to allene C followed by

anion exchange with AgPF6 provides allene 4.

I
H I H
R2 PF6
R2 R2 I 2 I R2 R2
N N R 2 I 2
I2 N R AgPF6 N R
R3
R1 R1 R1 R1
H AgI
R3 R3 Et3N
2 A B R3 3

Et3NH

H PF6
H I H I
I R2 I R2 I R2
2
N R AgPF6 N R
2
N R
2

R1 R1 R1
H H Et3N
AgI H NEt3
R3 R3 R3
4 D C

Scheme 2.5 Proposed reaction mechanism.

Compounds 4b, e-f were prepared by our group members and used for the reaction

condition screening of the bioconjugation studies in the following section. Given the

reaction screening results, I have designed and synthesized functionalized electron-

deficient allenes 4g-h for cysteine-selective peptide and protein labeling.

44
2.2.1.2 Optimization of Modification Reaction Conditions

Cysteine-containing peptide AYEMWCFHQR 1a and allene reagent 4a were

employed as substrates for reaction condition screening. By treatment of peptide

AYEMWCFQHR 1a (0.1 mM) with allene reagent 4a (1 equivalent) in potassium

phosphate buffer pH 8.0 buffer/CH3CN (9:1) at 25 °C for 4 h, modified peptide 5a was

afforded in 92% conversion with all other amino acid residues remaining intact as

confirmed by LC–MS/MS analysis (Table 2.1, Entry 1). Increasing the loading of allene

from 1 to 5 equivalents led to an improvement of the conversion to 95% (Entries 5-8).

Figure 2.2 MS/MS spectrum of cysteine-modified peptide 5a.

45
Table 2.1 Optimization of modification reaction conditions.a

H PF6
I O Cl O
H
N I
PF6
N H

H
4a S
AYEMWCFHQR Cl
Potassium phosphate buffer/ CH3CN (9:1) AYEMW FHQR
1a N
25 ºC, 4 h H
O

5a

Entry 4a (equiv.) pH values Temperature (°C) Conversion (%)b

1 1 8.0 25 92
2 1 5.8 25 19
3 1 6.5 25 36
4 1 7.4 25 74
5 2 8.0 25 93
6 3 8.0 25 94
7 4 8.0 25 95
8 5 8.0 25 95
9 5 8.0 37 96
10 5 5.8 25 62
11 5 6.5 25 90
12 5 7.4 25 93
a
Conditions of the modifications: treatment of AYEMWCFHQR 1a (0.1 mM) with
allene reagent 4a in 50 mM potassium phosphate buffer/CH3CN (9:1) at different pH
values and temperature for 4 h. b Conversion of the modification was determined by
LC–MS analysis.

46
Time course analysis was performed by conducting the reactions at different pH

values. When potassium phosphate buffer with different pH values (5.8, 6.5, 7.4 and

8.0) were used for the bioconjugation, lower conversion was resulted under slightly

acidic conditions (Figure 2.3). The conversions could be improved by increasing the

pH values of buffer (Table 2.1, Entries 1-4). This observation could be explained by

the pKa value of the thiol group of cysteine (pKa = ~ 8.5). Basic conditions could

promote the deprotonation of the thiol group to the thiolate group, resulting in stronger

nucleophilicity that facilitated the modification.

Time course experiment of the formation of modified peptide 5a at different

temperatures was also performed (Figure 2.4). The reaction proceeded relatively slow

at 4 °C. When the reaction was performed at 37 °C, modified peptide 5a in > 70%

conversion was observed after 30 min. Comparable results were obtained when the

modification was conducted at room temperature or 37 °C for 4 h (Table 2.1, Entry 9),

suggesting that the reaction proceeded efficiently at room temperature or 37 °C.

100
pH 5.8 buffer/ CH3CN (9:1)
80 pH 6.5 buffer/ CH3CN (9:1)
Conversion (%)

pH 7.4 buffer/ CH3CN (9:1)


60 pH 8.0 buffer/ CH3CN (9:1)

40

20

0
0 1 2 3 4
Time (h)

Figure 2.3 Time course experiments of the formation of cysteine modified peptide 5a

at different pH values.

47
100
4 °C
80 25 °C

Conversion (%)
37 °C
60

40

20

0
0 1 2 3 4
Time (h)

Figure 2.4 Time course experiments of the formation of cysteine modified peptide 5a

at different temperatures.

With the optimized conditions (using 5 equivalents of allene reagent 4a), the

modification proceeded smoothly under potassium phosphate, Tris-HCl, ammonium

bicarbonate, borate buffered saline (BBS) or sodium borate buffer medium, giving

conversions in 92-96% (Table 2.2).

Table 2.2 Screening of aqueous solutions of the modification.a

H PF6
I O Cl O
H
N I
PF6
N H

H
4a S
AYEMWCFHQR Cl
aqueous solution/ CH3CN (9:1) AYEMW FHQR
1a N
25 ºC, 4 h H
O

5a

Entry Aqueous solutions Conversion (%)b


1 50 mM Potassium phosphate buffer (pH 8.0) 95
2 50 mM Tris-HCl (pH 8.0) 96

48
3 50 mM NH4HCO3 buffer (pH 8.0) 92
4 50 mM BBS (pH 8.0) 95
5 50 mM Sodium borate buffer (pH 8.0) 96
a
Conditions of the modifications: treatment of AYEMWCFHQR 1a (0.1 mM) with
allene reagent 4a in aqueous solutions/CH3CN (9:1) at 25 °C for 4 h. b Conversion of
the modification was determined by LC–MS analysis.

2.2.1.3 Model Reaction of Cysteine Modification by Electron-Deficient Allenes

To investigate the structure of electron-deficient allene-modified peptide 5a, a

model study was conducted using benzyl mercaptan as substrate for the thiol-allene

reaction. Firstly, 1H-isoindoliums 3a was mixed with triethylamine for 1 h to generate

electron-deficient allene 4a (1.1 equiv.) under basic condition. After that, benzyl

mercaptan was added to the reaction mixture for further reaction of 3 h at room

temperature in H2O/CH3CN (1:3). After the reaction, the corresponding vinyl thioether

product 7 was isolated in 63% yield and the structure of product was confirmed by 1H

NMR analysis.

O
H
SH I PF6
N H
H PF6 H PF6
O I O Cl
I Et3N (1 equiv.) (1 equiv.)
N N
S Cl
CH3CN, r.t., 1 h CH3CN/H2O 3:1, r.t., 3 h

Cl H

3a 4a
7

Scheme 2.5 Model reaction of thiol with electron-deficient allenes.

49
2.2.1.4 Investigation of Cysteine Selectivity of the Modification

To examine the cysteine selectivity of the modification, another cysteine-

containing peptide STSSSCNLSK 1b, which consisted of different residues compared

with AYEMWCFHQR 1a, was treated with allene reagent 4a to afford modified

peptide 6a in 96% conversion (Table 2.3, Entry 2). From the MS/MS analysis, only the

cysteine residue on the peptides was modified with all other residues remaining intact

(Figure 2.5). The result indicated that this modification had high chemoselectivity

towards the thiol moiety of cysteine residue in the presence of other nucleophilic

residues, such as methionine, tryptophan, histidine etc.

Treatment of C-terminal cysteine-containing peptide KSTFC 1c and N-terminal

cysteine-containing peptide CSKFR 1d with allene reagent 4a gave modified peptide

8a and 9a in 93% and 94% conversion, respectively. LC–MS/MS analysis showed that

modification was on the cysteine residue with other residues remaining intact (Figures

2.6 and 2.7). Control experiments using peptides 1e-l without free cysteine residue gave

no modification (Table 2.3, Entries 5-13). This showed that the modification was highly

selective to cysteine residues.

50
Table 2.3 Investigation of the cysteine selectivity.a

H PF6 O
H
I O Cl I
N PF6
N H

H S
4a Cl
Peptides
pH 8.0 potassium N
1a-l H
phosphate buffer/ CH3CN (9:1), O
25 ºC, 4 h
5a, 6a, 8a and 9a

H O H O H O
I PF6 I PF6 I PF6
N H N H N H

S Cl S Cl S Cl
STSSS NLSK KSTF OH SKFR
N N H2N
H O H O O
6a 8a 9a

Entry Peptides Conversion (%)b


1 AYEMWCFHQR 1a 95
2 STSSSCNLSK 1b 96
3 KSTFC 1c 93
4 CSKFR 1d 94
5 STSSSANLSK 1e 0
6 STSSSHNLSK 1f 0
7 AYEMWSFHQR 1g 0
8 WSKFR 1h 0
9 YSKFR 1i 0
10 PSKFR 1j 0
12 GSKFR 1k 0
13 ISKFR 1l 0
a
Conditions of the modifications: treatment of peptides 1a-l (0.1 mM) with allene
reagent 4a in 50 mM pH 8.0 potassium phosphate buffer/CH3CN (9:1) at 25 °C for 4
h. b Conversion of the modification was determined by LC–MS analysis.

51
Figure 2.5 MS/MS spectrum of cysteine-modified peptide 6a.

Figure 2.6 MS/MS spectrum of cysteine-modified peptide 8a.


52
Figure 2.7 MS/MS spectrum of cysteine-modified peptide 9a.

2.2.1.5 Investigation of Scope of Allene reagents

The scope of allene reagents for the bioconjugation was investigated. Electron-

deficient allenes 5b-h were screened with cysteine containing peptide

AYEMWCFHQR 1a (Table 2.4). Allene reagents containing a wide range of para-

substituted phenyl groups at R1 position, such as methyl and halogen, gave modified

peptides in excellent conversion between 89% and 95% (Entries 1-6). Different amine

components at R2 were also compatible with the modification (Entries 6-7). The results

showed that the modification was well tolerated with allene reagents with different

substituents.

53
For functional labeling of the cysteine-containing peptides, coumarin-derived

allene 4g was employed for the modification and gave the corresponding modified

peptides 5g in 31% conversion (Entry 7). Modification with allene reagent 4h bearing

the alkyne moiety, which could be applied for sequential modification using click

reaction was also conducted, giving the formation of 5h in 94% conversion (Entry 8).

Bioconjugation of another cysteine-containing peptide STSSSCNLSK 1b with the

allene reagents was also performed. Modified peptides 6b-h were obtained in good to

excellent conversion between 59% and 96% (Table 2.5), indicating the high efficiency

of this modification.

Table 2.4 Investigation of the scope of electron-deficient allenes.a

H PF6
I R2 2 H
R I R2 PF6
N
R2
R1 N H

H R1
R3 R3
S
4
AYEMWCFHQR AYEMW FHQR
pH 8.0 potassium N
1a H
phosphate buffer/ CH3CN (9:1), O
25 ºC, 4 h
5a-h

Entry Allene Modified peptide Conversion (%)b

1 5a 95

54
2 5b 93

3 5c 91

4 5d 93

5 5e 89

6 5f 90

7 5g 31

55
8 5h 94

a
Conditions of the modifications: treatment of AYEMWCFHQR 1a (0.1 mM) with
allene reagent 4 in 50 mM pH 8.0 potassium phosphate buffer/CH3CN (9:1) at 25 °C
for 4 h. b Conversion of the modification was determined by LC–MS analysis.

Table 2.5 Modification of cysteine-containing peptide STSSSCNLSK 1b with

electron-deficient allenes.a

H PF6
I R2 2
R H
N I R2 PF6
R1 R2
N H
H
R1
R3
4 R3
S
STSSSCNLSK
pH 8.0 potassium STSSS NLSK
1b N
phosphate buffer/ CH3CN (9:1), H
O
25 ºC, 4 h
6a-h

Entry Allene Modified peptide Conversion (%)b

1 6a 96

2 6b 96

56
3 6c 93

4 6d 87

5 6e 84

6 6f 96

7 6g 59

8 6h 93

57
a
Conditions of the modifications: treatment of STSSSCNLSK 1b (0.1 mM) with
allene reagents 4 in 50 mM pH 8.0 potassium phosphate buffer/CH3CN (9:1) at 25 °C
for 4 h. b Conversion of the modification was determined by LC–MS analysis.

2.2.1.6 Stability Study of Allene-modified Bioconjugates

Previous studies have shown that some cysteine-selective reagents could be

removed after bioconjugation. For example, thiosuccinimide adducts formed by the

modification of cysteine with maleimides are now known to undergo retro-Michael

additions or thiol exchange reactions which cause cleavage of bioconjugates.68 This

would result in formation of heterogenous conjugate mixtures, limiting the applications

in in vivo studies.69 Therefore, stability study of the modified bioconjugates was

conducted for the allene-modified bionconjugates.

Investigations were conducted by treatment of modified peptide 5a with excessive

thiol-containing reagents and reducing reagents. Treatment of modified peptide 5a with

500 equivalents of L-cysteine and DL-homocysteine, respectively, showed that the

modified peptide 5a still remained intact after 3 h as confirmed by LC–MS analysis.

Besides, treatment of the modified peptide 5a with 500 equivalents of common

reducing reagents, DTT and sodium ascorbate, also showed no interference with the

modified bioconjugates.

These results indicated that the modified bioconjugates were stable toward

environments with thiol-containing reagents and common reducing reagents. Therefore,

the modification was regarded as a stable and irreversible cysteine modification.

58
2.2.2 Modification of Cysteine-Containing Proteins Using Isolated Electron-

Deficient Allenes

2.2.2.1 Modification of BSA Protein with Allene Reagents

The electron-deficient allenes were further employed to protein modification.

Bovine serum albumin (BSA) with a single free cysteine residue were utilized for

bioconjugation. Treatment of BSA (0.1 mM) with allene reagents 4a, 4d and 4h (10

equivalents) in PBS pH 7.4 buffer/CH3CN (9:1) at 25 °C for 4 h afforded modified

proteins in 56-70% conversion by LC–MS analysis (Table 2.6).

To identify the site of modification in protein, chymotrypsin digestion was

performed on the allene- modified protein BSA-1. By LC–MS analysis, two incomplete

digested peptide fragments, Lys20-Phe36 ([M+3H]3+ m/z 806.7) and Ser28-Phe49

([M+3H]3+ m/z 1044.1), were obtained (Figures 2.9 and 2.10). The expected mass of

this two peptide fragments from the native protein should be 1954.0 Da and 2666.3 Da

respectively. From the molecular feature extraction spectrum of modified protein BSA-

1, the mass of the two peptide fragments was shifted to 2417.1 Da and 3132.3 Da,

respectively. The increase in the mass of this two cysteine-containing fragment peptides

after modification indicated the incorporation of one molecule of 4a (M.W. = 462) into

the BSA peptide sequence. However, control experiment of chymotrypsin digestion of

native BSA protein is needed to confirm the mass shift of the two peptide fragments

and LC–MS/MS analysis of the modified peptide fragments is required to show that the

59
modification was on free cysteine residue Cys34 of BSA with other residues remaining

intact.

Table 2.6 Modification of cysteine-containing protein BSA with electron-deficient

allenes.a

H PF6 H
I R2 2 PF6
I R2
R2 R
N N H
R1
R1
SH S
H R3
R3
4

pH 7.4 PBS buffer,


25 ºC, 4 h
protein modified protein

Entry Allene Modified protein Conversion (%)b

1 BSA-1 65

2 BSA-2 70

3 BSA-3 56

a
Conditions of the modifications: treatment of BSA protein (0.1 mM) with allene
reagent 4 in 50 mM pH 7.4 PBS buffer/CH3CN (9:1) at 25 °C for 4 h. b Conversion of
the modification was determined by LC–MS analysis.

60
(a)

(b)

(c)

Figure 2.8 Mass spectrum of (a) BSA-1; (b) BSA-2 and (c) BSA-3.

61
Figure 2.9 Molecular feature extraction spectrum of cysteine modified

KGLVLIAFSQYLQQCPF after chymotrypsin digestion of BSA-1.

Figure 2.10 Molecular feature extraction spectrum of cysteine modified

SQYLQQCPFDEHVKLVNELTEF after chymotrypsin digestion of BSA-1.

62
2.2.2.2 Control Experiments with Non-Free Cysteine-Containing Proteins

Control experiments were conducted by treatment of non-free cysteine-containing

proteins, RNaseA and lysozyme in PBS pH 7.4 buffer at room temperature for 4 h. The

result showed that no modification was found. This result suggested that allene reagents

were highly selective to the cysteine residues on proteins and could be applied to

cysteine modification of proteins with high efficiency (Figures 2.11 and 2.12).

Figure 2.11 Mass spectrum of RNaseA after reaction with 4a.

Figure 2.12 Mass spectrum of lysozyme after reaction with 4a.

63
2.2.2.3 Fluorescent Labeling of Proteins

For the fluorescent labeling of proteins, alkyne handles were first incorporated on

the free cysteine residues of BSA using allene reagent 4h. After washing out the

excessive reagents, the alkyne-functionalized proteins BSA-3 were labeled with red

fluorescent rhodamine dye 10 by azide-alkyne Huisgen cycloadditions in 40%

conversion (Scheme 2.6 and Figure 2.13). SDS-PAGE analysis revealed that,

rhodamine-labeled protein BSA-4 were observed using fluorescent analysis, while no

fluorescent signal was observed for native BSA or alkyne-functionalized protein BSA-

3 (Figure 2.14). Employing Coomassie blue staining on the same gel, deep blue color

signals of native, alkyne-functionalized as well as rhodamine-labeled proteins were

observed, indicating that the fluorescent tag has been labeled on the proteins using the

allene-mediated cysteine modification and a sequential azide-alkyne click reaction.

PF6
H H O
I O I
PF6
N N H

SH
H S
4h

pH 7.4 PBS buffer,


protein 25 ºC, 4 h

N O N
N O N

O
O
N
H O
N N
I O
N PF6
N3
N H
O
Rhodamine azide 10,
S
CuSO4, TBTA, TCEP N
N N
pH 7.4 PBS buffer,
37 ºC, 1 h

Scheme 2.6 Modification of protein with alkyne-functionalized allene 4h and

sequential copper(I)-catalyzed azide-alkyne cycloadditions with rhodamine azide.

64
Figure 2.13 Mass spectrum of rhodamine-labeled BSA-4.

Figure 2.14 SDS-PAGE analysis of BSA, allene-modified BSA-3 and rhodamine-

labelled BSA-4.

65
2.3 Conclusion

In summary, a metal-free chemoselective cysteine modification using 1H-

isoindolium-based electron-deficient allenes was developed. We have developed a

novel electrophile-mediated cascade cyclization / iodination of propargylamine-based

1,6-diyne 2 for synthesis of a new class of spiro 1H-isoindoliums 3. Upon treatment

with base, the alkyne moieties of 1H-isoindoliums 2 were transformed to allene

functionalities to give the corresponding allene products 3. This new class of electron

deficient allenes could be employed for cysteine modification of peptides and proteins

through the interaction between allene and the nucleophilic thiol moiety on cysteine

residue.

A comprehensive study on this newly developed bioconjugation reaction was

presented. The modification could proceed with high efficiency (up to 96% conversion

in 4 h) and high chemoselectivity toward cysteine residue. Allene reagents with

different functional groups were compatible with this modification. The modified

products were stable towards excessive thiol-containing reagents and reducing agents.

This modification was further applied to protein modification and labeling.

Fluorescent tags could be labeled on the cysteine-containing proteins by sequential

modification using the azide-alkyne click reaction. This newly developed modification

would become a complementary method towards the currently used cysteine

modification protocols.

66
2.4 Suggestions for Future Research

In the present work, a highly selective cysteine modification of peptides and

proteins using 1H-isoindolium-based electron-deficient allenes was developed. For

peptide modification, the two cysteine-containing peptides 1a and 1b used for the study

are linear peptides consisting of different amino acid residues. The results showed that

the modification could proceed smoothly in the presence of amino acids with bulky or

aromatic side chains, including alanine, tyrosine, and tryptophan. In the future, it is

envisioned that the modification can be further applied to the modification of cyclic

peptides for the potential development of therapeutic agents.70 Besides, the

modification was well tolerated with allene reagents with different substituents. Based

on this results, it is suggested that other functionalized allene reagents can be prepared,

such as PEG probes for improving the water solubility and biocompatibility, so as to

further expand their applications on selective labelling of cysteine-containing peptides.

For protein modification, upon chymotrypsin digestion of allene-modified BSA

protein, the cysteine-containing peptide fragments were found, but the analysis and

identification of modification site was not yet completed. Therefore, it is suggested that

LC–MS/MS analysis should be performed in the future so as to confirm the location of

modification. Apart from chymotrypsin, it is suggested that the choice of protease for

protein digestion can be further explored, such as trypsin, Glu-C, or a combination of

multiple proteases, so as to optimize the digestion protocol for LC–MS/MS analysis.

Moreover, as protein digestion produces relatively long peptide fragments which may

be more difficult to interpret, the conditions for MS/MS analysis of tryptic peptide

67
fragments can be further optimized. For example, apart from performing MS/MS with

collision-induced dissociation (CID), electron transfer dissociation (ETD) may be an

alternative and effective approach for analysis of peptide fragments while the fragile

modification remaining intact. By using different fragmentation methods for MS/MS

analysis, more information of peptide fragmentation can be obtained.

68
Chapter 3 Cysteine Modification of Peptides and Proteins using in situ

Generated Electron-Deficient Allenes

3.1 Introduction

A detailed study of cysteine modification of peptides and proteins using electron-

deficient allenes was presented in Chapter 2, in which the electron deficient allene

reagents were prepared by multi-step chemical synthesis. Based on the success of

development of the modification, we would like to enhance the efficiency of this

modification by developing a simple, one-step procedure for preparation of electron

deficient allenes.

The investigation was initiated by preparation of in situ generated electron

deficient allene reagents by the reaction between propargylamine-based 1,6-diynes and

iodine monochloride, which resulted in the formation of in situ generation of 1H-

isoindolium. With the addition of reducing agent for quenching of reaction, the reaction

mixture could become a stock solution of in situ generated electron deficient allene with

high stability for direct use in bioconjugation.

Reaction conditions were optimized by screening different amount of allene

reagents with reducing agents. Then, studies of the modification were performed by

employing in situ generated electron-deficient allene reagents prepared from

propargylamine-based 1,6-diynes with different substituents. Moreover, the

modification was further applied to protein modification.

69
3.2 Results and Discussion

3.2.1 Modification of Cysteine-Containing Peptides Using in situ Generated

Electron-Deficient Allenes

3.2.1.1 Optimization of Reaction Conditions

To prepare the stock solution of in situ generated electron-deficient allene reagents,

a mixture of propargylamine-based 1,6-diyne 2 and iodine monochloride (ICl, 1.05

equivalents) was stirred in room temperature for 1 h to allow cyclization / iodination

for in situ generation of 1H-isoindolium 3 (Scheme 3.1). After one hour, the reaction

mixture was further diluted to 5 mM in CH3CN. Under basic reaction conditions, the

1H-isoindoliums 3 could isomerize to electron-deficient allenes 4 which react with the

thiol groups on cysteine residues.


H
R2 R2 I R2 2 Cl
N R
N
1) ICl (1.05 equiv.), CH3CN, r.t., 1 h

R3 2) reducing agent R1
R1
R3
2 3

Scheme 3.1 General procedure of preparation of in situ generated allene reagents.

As an initial attempt, propargylamine-based 1,6-diyne 2a was mixed with ICl for

the formation of in situ generated 1H-isoindolium 3a. After the reaction for 1 h,

reducing agent was added to the reaction mixture so as to quench the excess ICl. The

purpose of adding reducing agent was to prevent unwanted oxidation of cysteine

residues on peptides. The cysteine sulfhydryl group is nucleophilic and easily oxidized,

70
which can be readily converted to sulfenic acids and disulfides under oxidizing

conditions.48 Iodine monochloride is an oxidizing agent which potentially causes

oxidation of cysteine, which can block the target residue and hamper the modification

of peptides with allene reagents. Therefore, the addition of reducing agent in the

reaction mixtures could protect cysteine from oxidation.

The stock solution was directly used for bioconjugation with cysteine-containing

peptide AYEMWCFHQR 1a. Firstly, different common reducing agents, including

tris(2-carboxyethyl)phosphine (TCEP), dithiothreitol (DTT), sodium borohydride and

sodium ascorbate, were used for selection of reducing agent and optimization of

reaction conditions (Table 3.1, Entries 1-4). From the mass spectrum of the stock

solution, formation of 1H-isoindolium could be observed with m/z = 462.0129 (Figure

3.1). The stock solution showed excellent stability and could be stored at –20 °C for

repeated use.

Figure 3.1 LC–MS analysis of in situ generated allene reagent (TIC chromatogram).

With 1 equivalent of TCEP, treatment of cysteine-containing peptide

AYEMECFHQR 1a (0.1 mM) with in situ generated electron-deficient allene reagent

71
4a (1 equivalent) in pH 8.0 potassium phosphate buffer/CH3CN (9:1) at 25 °C for 4 h

afforded modified peptides 5a in 98% conversion. Increasing the amount of in situ

generated allene reagent 4a from 1 to 5 equivalents gave comparable results of

modification. Increasing the amount of TCEP lowered the overall conversion of

modification (Table 3.1, Entry 5). Additionally, formation of side product 5a’ was

observed when increasing the TCEP loading to 2 equivalents (Figure 3.2).

Table 3.1 Optimization of modification conditions.a

O
H
O Cl
N I
N
1) ICl (1.05 equiv.), CH3CN, r.t., 1 h
2) reducing agent
Cl
Cl
3a
2a

H O
I
Cl
N H
AYEMWCFHQR 1a
S
pH 8.0 potassium phosphate Cl
buffer/ CH3CN (9:1) AYEMW FHQR
N
25 ºC, 4 h H
O

5a

Propargylamine- Amount of
Conversion
Entry based 1,6-diynes Reducing Agent Reducing
(%)b
(equiv.) Agent (equiv.)
1 1 TCEP 1 98
2 1 NaBH4 1 95
3 1 DTT 1 95
4 1 Sodium ascorbate 1 93
5 1 TCEP 2 89
6 2 TCEP 1 99
7 3 TCEP 1 99

72
8 4 TCEP 1 99
9 5 TCEP 1 99
a
Conditions of the modifications: treatment of AYEMWCFHQR 1a (0.1 mM) with
different amounts of in situ generated allene reagent 4a prepared from
propargylamine-based 1,6-diyne 2a in 50 mM pH 8.0 potassium phosphate
b
buffer/CH3CN (9:1) at 25 °C for 4 h. Conversion of the modification was
determined by LC–MS analysis.

H Cl
I O O O
H
N I
Cl N
N H
Cl
+
S
Cl S
5 mM stock solution in CH3CN
with TCEP (2 equiv.) AYEMW FHQR
N AYEMW FHQR
AYEMWCFHQR 1a H N
O H Cl
O

5a (89% conversion) 5a’ (8% conversion)

Scheme 3.2 Formation of side product 5a'.

Figure 3.2 LC–MS analysis of side product 5a' (TIC chromatogram).

Control experiment was conducted by treatment of cysteine-containing peptide

AYEMWCFHQR 1a with 10 equivalents of propargylamine-based 1,6-diyne 2a, which

showed the formation of 5a’ as the single product in 30% conversion. The cysteine
73
residue on the peptides was modified while all other residues remained intact as shown

in LC–MS/MS analysis (Figure 3.3). This result revealed that product 5a’ was

suggested to form by cysteine modification through thiol-yne reaction between the thiol

group on cysteine and the alkyne group on the propargylamine-based 1,6-diyne 2a,

which could possibly be proceeded through radical processes or under base-catalyzed

conditions as reported in previous literature examples.58a, 71

Figure 3.3 MS/MS analysis of side product 5a'.

Although the modification of propargylamine-based 1,6-diynes on cysteine was

observed, the reactivity was far lower than that of electron-deficient allene reagents.

74
Besides, the formation of product 5a’ could be minimized (< 2% conversion) under

optimized reaction conditions. Therefore, this observation was assumed to have little

effect on the selectivity of the modification.

3.2.1.2 Time Course Experiments of Cysteine Modification of Peptides by in situ

Generated Allene Reagents

To compare the performance of bioconjugation using the in situ generated allene

reagents with that using the isolated electron-deficient allene reagents, time course

experiments of the formation of modified peptide 5a at different pH values and

temperatures were performed. At different pH values, bioconjugation of peptide with

the in situ generated allene reagents gave comparable results with that conducted using

the isolated reagents, in which better conversion could be obtained in buffer with higher

pH values. The reaction rates were lower in slightly acidic media (pH 5.8 and pH 6.5)

at the first one hour of the reaction, which could be attributed to the low conversion of

1H-isoindoliums 3 to electron-deficient allenes 4 at these pH values. The isomerization

of 1H-isoindoliums 3 to electron-deficient allenes 4 should be promoted under basic

reaction conditions. At slightly acidic reaction conditions, the reagents remained in

their alkyne forms, which was less reactive than allenes towards thiol group on cysteine.

As a result, lower reaction rates were observed when the bioconjugation was conducted

at pH 5.8 and 6.5.

When the reaction was performed at room temperature or 37 °C, modified peptide

5a in > 80% conversion was observed at the first hour of the reaction. The conjugation

75
could also proceed smoothly at 4 °C, giving approximately 80% conversion after 4 h.

The result revealed that the use of the in situ generated allene reagents generally gave

improved conversion of modified peptides compared with the isolated allene reagents.

This could be explained by the presence of TCEP in the in situ generated allene reagents,

which could prevent the cysteine-containing peptides from dimerization. Therefore, the

cysteine residues on peptides could remain active for reaction with the allene reagents,

resulting in higher conversion of modified peptides.

(a) 100

80 pH 5.8 buffer/ CH3CN (9:1)


Conversion (%)

pH 6.5 buffer/ CH3CN (9:1)


60 pH 7.4 buffer/ CH3CN (9:1)
pH 8.0 buffer/ CH3CN (9:1)
40

20

0
0 1 2 3 4
Time (h)

(b)
100
4 °C
80 25 °C
Conversion (%)

37 °C
60

40

20

0
0 1 2 3 4
Time (h)

Figure 3.4 Time course experiments of the formation of cysteine modified peptide
5a at (a) different pH values; (b) different temperatures.

76
3.2.1.3 Investigation of Scope of in situ Generated Allene Reagents

The optimized reaction conditions of employing 1 equivalent of in situ generated

allene reagent bearing 1 equivalent of TCEP and pH 8.0 potassium phosphate buffer at

25 °C for 4 h was utilized for expansion of the scope with other in situ generated allene

reagents prepared from different propargylamine-based 1,6-diynes 2. Two cysteine-

containing peptides, AYEMWCFHQR 1a and STSSSCNLSK 1b, were treated with in

situ generated allene reagents 4a-d, which differed in the substituents on the para-

substituted phenyl ring at the R1 position. Modified peptides 5a-d and 6a-d were

obtained in 96-98% and 98-99% conversion, respectively, suggesting that the

modification was compatible with in situ generated allene reagents with different

substituents. Fluorescent allene reagent 4g prepared from coumarin-derived

propargylamine-based 1,6-diyne 2g was also applied to the modification, giving

modified peptide 5g and 6g in 52% and 85% conversion, respectively.

Table 3.2 Modification of cysteine-containing peptide AYEMWCFHQR 1a with in

situ generated electron-deficient allenes.a

H
R2 R2 I R2 2 Cl
N R
N
1) ICl (1.05 equiv.), CH3CN, r.t., 1 h

R3 2) TCEP (1 equiv.) R1
R1
R3
2 3

H
I R2 Cl
R2
N H

AYEMWCFHQR 1a R1
R3 S
pH 8.0 potassium phosphate
buffer/ CH3CN (9:1) AYEMW FHQR
N
25 ºC, 4 h H
O

5a-d, g

77
Propargylamine-based
Entry Modified peptide Conversion (%)b
1,6-diynes 2

1 5a 98

2 5b 96

3 5c 96

4 5d 98

5 5g 52

a
Conditions of the modifications: treatment of AYEMWCFHQR 1a (0.1 mM) with
in situ generated allene reagents 4 prepared from different propargylamine-based 1,6-
diynes 2 in 50 mM pH 8.0 potassium phosphate buffer/CH3CN (9:1) at 25 °C for 4 h.
b
Conversion of the modification was determined by LC–MS analysis.

78
Table 3.3 Modification of cysteine-containing peptide STSSSCNLSK 1b with in situ

generated electron-deficient allenes.a


H
R2 R2 I R2 2 Cl
N R
N
1) ICl (1.05 equiv.), CH3CN, r.t., 1 h

R3 2) TCEP (1 equiv.) R1
R1
R3
2 3

H
I R2 2 Cl
R
N H
STSSSCNLSK 1b R1
R3 S
pH 8.0 potassium phosphate
buffer/ CH3CN (9:1) STSSS NLSK
N
25 ºC, 4 h H
O
6a-d, g

Propargylamine-based
Entry Modified peptide Conversion (%)b
1,6-diynes 2

1 6a 98

2 6b 98

3 6c 99

79
4 6d 99

5 6g 85

a
Conditions of the modifications: treatment of STSSSCNLSK 1b (0.1 mM) with in
situ generated allene reagents 4 prepared from different propargylamine-based 1,6-
diynes 2 in 50 mM pH 8.0 potassium phosphate buffer/CH3CN (9:1) at 25 °C for 4
h. b Conversion of the modification was determined by LC–MS analysis.

For in situ generated allene reagents 4a-d prepared from propargylamine-based

1,6-diynes 2a-d, the substituents on the phenyl ring showed increased electron

withdrawing property, in which 2d (-CN) > 2c (-CF3) > 2a (-Cl) > 2b (-H). Time course

experiment of the formation of their corresponding bioconjugates 5a-d was conducted

to examine the effect of electron withdrawing group on the efficiency of the

modification. All reagents gave excellent conversion to desired products after

modification for 4 h, but there was difference in reaction rate among the reagents at the

beginning of the modification. At the first half hour of the reaction, reagents with

stronger electron withdrawing substituents gave higher conversion to modified peptides,

in which 5d (-CN) > 5c (-CF3) > 5a (-Cl) > 5b (-H). This observation could be

potentially explained by the increased electrophilicity of allene reagents caused by the

80
presence of strong electron withdrawing groups, which facilitated the nucleophilic

addition of thiol group and resulting in higher reaction rate of modification.

Figure 3.5 Time course experiment of formation of modified peptides 5a-d.

3.2.2 Modification of Cysteine-Containing Proteins Using in situ Generated

Electron-Deficient Allenes

The in situ generated electron-deficient allene reagents were further employed to

protein modification. Bovine serum albumin (BSA) with a single free cysteine residue

were used for bioconjugation with in situ generated allene reagents 4a-d. With 1

equivalent of allene reagent 4a, modified protein BSA-1 was obtained in 28%

conversion by LC–MS analysis. When the loading of allene reagent was increased from

1 to 5 equivalents, both single-site and multisite modifications of protein were observed

and the results were consistent with four allene reagents 4a-d.

81
Taking modified protein BSA-6 as the example (Figure 3.6a), two bioconjugates

corresponding to single-site modification were found. The peak with m/z 66765.96 and

m/z 66892.00 indicated that one molecule of 2a (M.W.= 335) and 4a (M.W.= 462) was

incorporated into the protein, respectively. For multisite modification, six

bioconjugates were found. Random incorporation of one molecule of 4a and one

molecule of 2a (m/z 67227.51), two molecules of 4a (m/z 67354.50), two molecules of

4a and one molecule of 2a (m/z 67690.64), three molecules of 4a (m/z 67817.08), three

molecules of 4a and one molecule of 2a (m/z 68153.05), or four molecules of 4a (m/z

68281.26) was observed from the mass spectrum, indicating that the modification

occurred on not only the free cysteine Cys34, but also the cysteine residues produced

from reduced disulfide bonds. This observation was possibly attributed to the presence

of TCEP in the in situ generated allene reagents, leading to disulfide bond reduction

and resulting in multiple free cysteine residues available for modification.

H
R2 R2 I R2 2 Cl
N R
N
1) ICl (1.05 equiv.), CH3CN, r.t., 1 h

R3 2) TCEP (1 equiv.) R1
R1
R3
2 3
H
SH I R2 2 Cl
R
N H

R1
R3 S
protein

pH 7.4 PBS buffer, 25 ºC, 4 h

modified protein

Scheme 3.3 Modification of cysteine-containing protein BSA with in situ generated

electron-deficient allenes.a

82
(a)

(b)

(c)

(d)

Figure 3.6 Mass spectrum of (a) BSA-6; (b) BSA-7; (c) BSA-8 and (d) BSA-9.

83
Bioconjugation of BSA with 10 equivalents of fluorescent allene reagent 4g was

performed to give modified protein BSA-10. Two single-site modifications were

observed, including incorporation of one molecule of 4g and 2g in 8% and 19%

conversion, respectively (Figure 3.7). The high conversion of 4g-modified protein than

2g-modified protein could be attributed to their difference in steric bulkiness.

Propargylamine-based 1,6-diyne 2g was an uncharged compound that could access to

the charged protein surface more effectively than the charged and bulkier electron

deficient allene 4g. As a result, the reaction of protein with 2g was more favorable.

According to the SDS-PAGE analysis, modified protein BSA-10 was observed by

fluorescent analysis while no fluorescent signal was observed for native BSA (Figure

3.8). Employing Coomassie blue staining on the same gel, deep blue color signals of

native protein and BSA-10 were observed, indicating that the fluorescent tag has been

labeled on the protein.

Figure 3.7 Mass spectrum of fluorescent-labeled BSA-10.

84
Figure 3.8 SDS-PAGE analysis of native BSA and BSA-10.

Disulfide bond formation is essential for the folding, activity and stability of

proteins. The modification of proteins with the in situ generated allene reagents resulted

in multisite modified bioconjugates, which indicated the cleavage of protein disulfide

bonds. In this regard, LC–MS/MS analysis is required to determine the sites of

modification and Circular Dichroism can be used to examine if the modifications alter

the secondary and tertiary structures of the proteins. Despite the formation of

heterogenous products, this modification method still showed high selectivity towards

cysteine residues in proteins.

85
3.3 Conclusion

In summary, a cysteine modification using in situ generated electron-deficient

allenes was developed. Based on the newly developed cysteine modification method

using isolated 1H-isoindolium-based electron deficient allene reagents presented in

Chapter 2, a one-step procedure for preparation of electron deficient allene reagents for

bioconjugation was established.

The in situ generated electron deficient allene reagents were prepared by reaction

of propargylamine-based 1,6-diynes 2 with iodine monochloride, resulting in the

formation of in situ generation of 1H-isoindoliums 3. With 1 equivalent of reducing

agent, the stable reaction mixtures could serve as stock solutions of in situ generated

electron deficient allene reagents for direct use in bioconjugation.

The modification of peptides using the in situ generated allene reagents could

proceed with high efficiency (up to 99% conversion in 4 h) and the results were

comparable with those obtained using the isolated allene reagents. These findings

showed that this approach could serve as an alternative method for preparation of allene

reagents and be utilized in the newly developed cysteine modification using electron-

deficient allenes.

For protein modification and labeling, single-site modification was resulted to give

cysteine-modified protein in 28% conversion using 1 equivalent of allene reagent, while

multisite modification was observed by increasing the amount of reagent.

86
Chapter 4 Experimental Section

4.1 General procedures

All reagents were commercially available and used without further purification. Milli-

Q® water used as reaction solvent in peptide and protein modification and LC–MS was

deionised using a Milli-Q® Gradient A10 system (Millipore, Billerica, USA). Flash

column chromatography was performed using silica gel 60 (230-400 mesh ASTM) with

ethyl acetate/n-hexane or methanol/dichloromethane as eluent. 1H and 13


C NMR

spectra were recorded on a Bruker DPX-400, Varian Unity Inova 400 NB spectrometers.

All chemical shifts are quoted on the scale in ppm using TMS or residual solvent as the

internal standard. Coupling constants (J) are reported in Hertz (Hz) with the following

splitting abbreviations: s = singlet, br s = broad singlet, d = doublet, dd = double doublet,

t = triplet and m = multiplet. All the mass spectra were obtained on an ESI source of

Agilent 6540 Ultra High Definition (UHD) Accurate-Mass Q-TOF LC/MS systems in

the positive ion mode.

LC–MS analyses for peptide identification were performed by using an Agilent 6540

UHD Accurate-Mass Q-TOF LC/MS system equipped with an ion spray source and an

Agilent 1290 Infinity LC, using an Agilent ZORBAX RRHD SB300-C18 (1.8 μm, 2.1

x 100 mm) column. 3 μL of sample was injected with a flow rate of the flow rate was

0.2 mL/min. Mobile phase A was made of Milli-Q® water containing 0.1% formic acid.

Mobile phase B was made of HPLC grade methanol containing 0.1% formic acid. The

initial conditions for separation were 5% B for 3 min, followed by a linear gradient to

87
95% B by 17 min. The composition was maintained for 10 min, followed by a linear

gradient to 5% B by 1 min. The composition was maintained for 4 min. Operating

conditions optimized for the detection of reaction mixture were the following: Gas

temperature: 300 °C, Drying gas: 8 L/min, Nebulizer: 35 psig, Sheath gas temperature:

270 °C, Sheath gas flow: 11 L/min, VCap: 3500 V, Nozzle voltage: 1000 V.

LC–MS analyses for protein identification were performed by using an Agilent 6540

UHD Accurate-Mass Q-TOF LC/MS system equipped with an ion spray source and an

Agilent 1290 Infinity LC, using an Agilent ZORBAX RRHD SB300-C3 (1.8 μm, 2.1

x 100 mm) column. 3 μL of sample was injected with a flow rate of 0.2 mL/min. Mobile

phase A was made of Milli-Q® water containing 0.1% formic acid. Mobile phase B was

made of HPLC grade methanol containing 0.1% formic acid. The initial conditions for

separation were 5% B for 3 min, followed by a linear gradient to 95% B by 17 min.

The composition was maintained for 10 min, followed by a linear gradient to 5% B by

1 min. The composition was maintained for 4 min. Operating conditions optimized for

the detection of reaction mixture were the followings: Gas temperature: 300 °C, Drying

gas: 8 L/min, Nebulizer: 35 psig, Sheath gas temperature: 350 °C, Sheath gas flow: 11

L/min, VCap: 3500 V, Nozzle voltage: 1000 V.

88
4.2 Literature Reference of Compounds

M. A. Cinellu, A. Zucca, S. Stoccoro, G.


Minghetti, M. Manassero and M.
Sansoni, J. Chem. Soc., Dalton Trans.,
1995, 2865–2872.

G. Charron, M. M. Zhang, J. S. Yount, J.


Wilson, A. S. Raghavan, E. Shamir and
H. C. Hang, J. Am. Chem. Soc., 2009,
131, 4967–4975.

M. L. Mason, R. F. Lalisse, T. J.
Finnegan, C. M. Hadad, D. A. Modarelli,
J. R. Parquette, Langmuir, 2019, 35,
12460−12468.

4.3 Calculation of Peptide Conversion

The crude reaction mixture of unmodified peptide (starting) and modified peptide

(product) was subjected to LC–MS analysis with elution time of 35 min. After data

processing by MassHunter, peptide conversion at different time intervals was

determined by measuring the relative peak intensities of starting and product in the

mass spectrum as follows:

𝑃𝑒𝑝𝑡𝑖𝑑𝑒 𝐶𝑜𝑛𝑣𝑒𝑟𝑠𝑖𝑜𝑛 (%) =

𝑅𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝑃𝑒𝑎𝑘 𝐼𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑆𝑡𝑎𝑟𝑡𝑖𝑛𝑔


(1 − ) × 100%
𝑅𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝑃𝑒𝑎𝑘 𝐼𝑛𝑡𝑒𝑠𝑖𝑡𝑖𝑒𝑠 𝑜𝑓 𝑆𝑡𝑎𝑟𝑡𝑖𝑛𝑔 𝑎𝑛𝑑 𝑃𝑟𝑜𝑑𝑢𝑐𝑡

89
4.4 Calculation of Protein Conversion

The crude reaction mixture of unmodified protein (starting) and modified protein

(product) was subjected to LC–MS analysis with elution time of 35 min. After data

processing by MassHunter, protein conversion at different time intervals was

determined by measuring the relative peak intensities of starting and product in the

mass spectrum as follows:

𝑃𝑟𝑜𝑡𝑒𝑖𝑛 𝐶𝑜𝑛𝑣𝑒𝑟𝑠𝑖𝑜𝑛 (%) =

𝑅𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝑃𝑒𝑎𝑘 𝐼𝑛𝑡𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑆𝑡𝑎𝑟𝑡𝑖𝑛𝑔


(1 − ) × 100%
𝑅𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝑃𝑒𝑎𝑘 𝐼𝑛𝑡𝑒𝑠𝑖𝑡𝑖𝑒𝑠 𝑜𝑓 𝑆𝑡𝑎𝑟𝑡𝑖𝑛𝑔 𝑎𝑛𝑑 𝑃𝑟𝑜𝑑𝑢𝑐𝑡

4.5 Synthesis of Propargylamine-based 1,6-diynes 2

To synthesize i, a mixture of aldehyde (10 mmol, 1 equiv.), amine (12 mmol, 1.2 equiv.),

alkyne (15 mmol, 1.5 equiv.) and gold catalyst (2 mol%) in ClCH2CH2Cl was stirred at

60 °C overnight. After the reaction, the resulting mixture was concentrated under

90
reduced pressure and purified by flash column chromatography using EtOAc/hexane

(1:3) as eluent to give product i.

To synthesize 2, a mixture of i (5 mmol, 1 equiv.) and K2CO3 (5 mmol, 1 equiv.) in 10

mL of CH3OH was stirred at room temperature for 30 min. After the reaction, the

reaction mixture was diluted with 20 mL of H2O. The mixture was then extracted with

CH2Cl2 (20 mL × 3). The solvent was removed by rotary evaporation and purified by

flash column chromatography using EtOAc/hexane (1:3) as eluent to give product 2.

Pale yellow solid, 62% yield.

1H NMR (400 MHz, CDCl3) δ 7.70 (d, J = 7.6 Hz, 1H), 7.57 – 7.52 (m, 1H), 7.44 –

7.34 (m, 3H), 7.28 (dd, J = 8.8, 7.3 Hz, 3H), 5.17 (s, 1H), 3.76 – 3.65 (m, 4H), 3.34 (s,

1H), 2.72 – 2.59 (m, 4H).

13C NMR (100 MHz, CDCl3) δ 140.38, 134.44, 133.41, 133.15, 128.83, 128.77, 128.72,

127.92, 122.68, 121.53, 86.88, 86.74, 82.18, 81.85, 67.18, 59.96, 50.26.

HRMS (ESI): [M+H]+ Calcd. for [C21H19ClNO]+ 336.1150, found 336.1157.

91
Pale yellow solid, 40% yield.

1H NMR (400 MHz, CDCl3) δ 7.74 (d, J = 7.7 Hz, 1H), 7.58 – 7.53 (m, 1H), 7.52 –

7.46 (m, 2H), 7.41 – 7.35 (m, 1H), 7.34 – 7.25 (m, 4H), 5.19 (s, 1H), 3.76 – 3.66 (m,

4H), 3.34 (s, 1H), 2.76 – 2.61 (m, 4H).

13C NMR (100 MHz, CDCl3) δ 140.64, 133.37, 131.94, 128.93, 128.70, 128.45, 128.42,

127.84, 123.09, 122.69, 88.05, 85.61, 82.09, 81.94, 67.23, 59.97, 50.24.

HRMS (ESI): [M+H]+ Calcd. for [C21H20NO]+ 302.1539, found 302.1546.

Pale yellow solid, 58% yield.

1H NMR (400 MHz, CDCl3) δ 7.70 (d, J = 7.7 Hz, 1H), 7.58 (s, 5H), 7.41 – 7.35 (m,

1H), 7.30 (t, J = 7.5 Hz, 1H), 5.21 (s, 1H), 3.75 – 3.66 (m, 4H), 3.35 (s, 1H), 2.67 (ddt,

J = 20.2, 10.7, 5.1 Hz, 4H).

13C NMR (100 MHz, CDCl3) δ 140.17, 133.49, 132.21, 130.39, 128.81, 128.77, 128.03,

126.87, 125.46, 125.42, 125.39, 125.35, 122.75, 122.72, 88.41, 86.73, 82.27, 82.25,

81.80, 67.19, 59.99, 50.30.

HRMS (ESI): [M+H]+ Calcd. for [C22H19F3NO]+ 370.1413, found 370.1418.

92
Yellow solid, 40% yield.

1H NMR (400 MHz, CDCl3) δ 7.69 – 7.63 (m, 1H), 7.60 (d, J = 8.4 Hz, 2H), 7.55 (d,

J = 7.8 Hz, 3H), 7.38 (td, J = 7.6, 1.3 Hz, 1H), 7.29 (td, J = 7.5, 1.2 Hz, 1H), 5.20 (s,

1H), 3.69 (pt, J = 5.5, 2.8 Hz, 4H), 3.35 (s, 1H), 2.73 – 2.58 (m, 4H).

13C NMR (100 MHz, CDCl3) δ 133.50, 132.46, 132.16, 128.77, 128.70, 128.09, 82.34,

67.12, 60.00, 50.29.

HRMS (ESI): [M+H]+ Calcd. for [C22H19N2O]+ 327.1492, found 327.1497.

Pale yellow solid, 40% yield.

1H NMR (400 MHz, CDCl3) δ 7.70 (dd, J = 7.8, 1.3 Hz, 1H), 7.55 (dd, J = 7.6, 1.4 Hz,

1H), 7.43 (s, 4H), 7.38 (td, J = 7.6, 1.5 Hz, 1H), 7.31 – 7.26 (m, 1H), 5.19 (s, 1H), 3.70

(dt, J = 5.8, 3.8 Hz, 4H), 3.34 (s, 1H), 3.17 (s, 1H), 2.74 – 2.59 (m, 4H).

13C NMR (100 MHz, CDCl3) δ 140.39, 133.43, 132.18, 131.84, 128.87, 128.75, 127.94,

123.55, 122.70, 122.12, 87.82, 87.49, 83.36, 82.18, 81.87, 67.21, 60.02, 50.28.

HRMS (ESI): [M+H]+ Calcd. for [C23H20NO]+ 326.1539, found 326.1539.

93
4.5.1 Synthesis of Coumarin-derived Propargylamine-based 1,6-diyne 2g

To synthesize ii, a mixture of 7-(diethylamino)coumarin-3-carboxylic acid (2 mmol, 1

equiv.), N-hydroxysuccinimide (2.6 mmol, 1.3 equiv.), 1-ethyl-3-(3-

dimethylaminopropyl)-carbodiimide (EDCI) (2.6 mmol, 1.3 equiv.) and 4-

dimethylaminopyridine (DMAP) (2 mmol, 1 equiv.) in CH2Cl2 (15 mL) was stirred in

room temperature for 8 h. After the reaction, the reaction mixture was diluted with 30

mL of CH2Cl2. The mixture was washed with H2O (30 mL × 2), and then dried over

anhydrous MgSO4. The solvent was removed by rotary evaporation to give residue

which was purified by flash column chromatography using CH2Cl2/EtOAc (10:1) as

eluent to give product ii.

To synthesize iii, a mixture of compound ii (1.57 mmol, 1 equiv.), 4-amino-1-Boc-

piperidine (1.57 mmol, 1 equiv.) and DMAP (0.31 mmol, 0.2 equiv.) in CH2Cl2 (15 mL)

94
was stirred at room temperature for 8 h. After the reaction, the solvent was removed by

rotary evaporation. The mixture was re-dissolved by TFA/CH2Cl2 (1:1, 5 mL:5 mL)

and was stirred in room temperature for 1 h. After the reaction, the reaction mixture

was diluted with 20 mL of H2O. The mixture was neutralized to pH 8.0 by 1 M NaOH

solution. The mixture was extracted with CH2Cl2 (30 mL × 3), and then dried over

anhydrous MgSO4. The solvent was removed by rotary evaporation to give residue

which was purified by flash column chromatography using CH2Cl2/CH3OH (10:1) as

eluent to give product iii.

Yellow Powder; 91% isolated yield.

1H NMR (400 MHz, CDCl3) δ 8.80 (d, J = 7.6 Hz, 1H), 8.69 (s, 1H), 7.43 (d, J = 9.0

Hz, 1H), 6.64 (dd, J = 9.0, 2.4 Hz, 1H), 6.50 (d, J = 2.3 Hz, 1H), 4.02 – 4.12 (m, 1H),

3.45 (q, J = 7.1 Hz, 4H), 3.10 (dd, J = 9.0, 3.7 Hz, 2H), 2.80 – 2.65 (m, 2H), 2.01 (d, J

= 9.6 Hz, 2H), 1.56 – 1.38 (m, 2H), 1.24 (t, J = 7.1 Hz, 6H).

3) δ 162.88, 162.58, 157.82, 152.73, 148.23, 131.26, 110.34,


13C NMR (100 MHz, CDCl

110.09, 108.55, 96.77, 46.13, 45.22, 44.51, 31.84, 12.57.

HRMS (ESI): [M+H]+ Calcd. for [C19H26N3O3]+ 344.1969, found 344.1996.

To synthesize iv, a mixture of 2-[(trimethylsilyl)ethynyl]benzaldehyde (10 mmol, 1

equiv.), iii (12 mmol, 1.2 equiv.), 1-chloro-4-ethynylbenzene (15 mmol, 1.5 equiv.) and

95
gold catalyst (2 mol%) in ClCH2CH2Cl was stirred at 60 °C overnight. After the

reaction, the resulting mixture was concentrated under reduced pressure and purified

by flash column chromatography using EtOAc/CH2Cl2 (1:4) as eluent to give product

iv.

To synthesize 2g, a mixture of iv (5 mmol, 1 equiv.) and K2CO3 (5 mmol, 1 equiv.) in

10 mL of CH3OH was stirred at room temperature for 30 min. After the reaction, the

reaction mixture was diluted with 20 mL of H2O. The mixture was then extracted with

CH2Cl2 (20 mL × 3). The solvent was removed by rotary evaporation and purified by

flash column chromatography using EtOAc/CH2Cl2 (1:4) as eluent to give product 2g.

Pale yellow solid, 38% yield.

1H NMR (400 MHz, CDCl3) δ 8.79 (d, J = 7.8 Hz, 1H), 8.68 (s, 1H), 7.70 (d, J = 7.6

Hz, 1H), 7.57 – 7.52 (m, 1H), 7.46 – 7.34 (m, 4H), 7.31 – 7.27 (m, 2H), 6.63 (dd, J =

9.0, 2.3 Hz, 1H), 6.49 (d, J = 2.1 Hz, 1H), 5.22 (s, 1H), 4.00 (td, J = 13.8, 6.8 Hz, 1H),

3.44 (q, J = 7.1 Hz, 4H), 3.35 (s, 1H), 2.92 (dd, J = 29.3, 11.5 Hz, 2H), 2.69 – 2.56 (m,

1H), 2.53 – 2.42 (m, 1H), 2.07 – 1.91 (m, 2H), 1.78 (s, 1H), 1.74 – 1.51 (m, 2H), 1.23

(t, J = 7.1 Hz, 6H).

96
13C NMR (100 MHz, CDCl3) δ 162.84, 162.46, 157.76, 152.62, 148.07, 141.18, 134.27,

133.23, 131.20, 128.75, 127.73, 122.69, 121.75, 110.67, 110.03, 108.56, 96.74, 87.01,

82.07, 59.84, 47.06, 46.66, 45.21, 32.11, 12.60.

HRMS (ESI): [M+H]+ Calcd. for [C36H35ClN3O3]+ 592.2361, found 592.2377.

4.6 Synthesis of 1H-isoindoliums 3

A mixture of 2 (1.0 mmol), electrophile (1.05 mmol, 1.05 equiv.) in 5 mL of CH3CN

was stirred at room temperature for 1 h, and then AgPF6 (1.1 mmol, 1.1 equiv.) was

added to the reaction mixture and stirred for 5 min. The resulting precipitate (AgI) was

separated by suction filtration and washed by CH2Cl2. The filtrate was concentrated

under vacuum and the residue was purified by flash column chromatography using

CH2Cl2/CH3OH (50:1) to give product 3.

Grey solid, 90% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.60 (d, J = 7.7 Hz, 1H), 8.26 (s, 1H), 7.84 – 7.68

(m, 3H), 7.59 – 7.53 (m, 2H), 7.50 – 7.44 (m, 2H), 6.99 (s, 1H), 4.42 (td, J = 10.4, 2.8

97
Hz, 1H), 4.32 (td, J = 14.2, 2.4 Hz, 2H), 4.20 – 3.99 (m, 3H), 3.87 – 3.77 (m, 1H), 3.66

(d, J = 12.1 Hz, 1H).

13C NMR (100 MHz, d6-DMSO) δ 146.60, 136.25, 135.29, 133.80, 132.60, 130.21,

129.08, 129.00, 125.90, 124.82, 118.45, 92.86, 80.29, 79.03, 65.67, 62.94, 61.96, 61.72,

56.44.

HRMS (ESI): [M-PF6]+ Calcd. for [C21H18ClINO]+ 462.0116, found 462.0114.

Grey solid, 81% yield.

1H NMR (600 MHz, d6-DMSO) δ 8.60 (d, J = 8.5 Hz, 1H), 8.25 (s, 1H), 7.80 – 7.68

(m, 3H), 7.50 (dd, J = 25.8, 6.9 Hz, 3H), 7.42 (t, J = 7.5 Hz, 2H), 6.99 (s, 1H), 4.40 (t,

J = 11.6 Hz, 1H), 4.32 (t, J = 14.0 Hz, 2H), 4.19 – 4.11 (m, 1H), 4.07 (d, J = 12.9 Hz,

1H), 4.01 (d, J = 13.5 Hz, 1H), 3.81 (t, J = 11.5 Hz, 1H), 3.65 (d, J = 12.9 Hz, 1H).

13C NMR (150 MHz, d6-DMSO) δ 146.57, 136.40, 132.57, 132.01, 130.44, 130.17,

129.05, 128.84, 125.86, 124.76, 119.50, 94.01, 79.31, 79.02, 65.79, 62.83, 61.92, 61.69,

56.38, 30.68.

HRMS (ESI): [M-PF6]+ Calcd. for [C21H19INO]+ 428.0506, found 428.0505.

98
White solid, 40% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.60 (d, J = 7.4 Hz, 1H), 8.27 (s, 1H), 7.82 – 7.73

(m, 7H), 7.03 (s, 1H), 4.48 – 4.38 (m, 1H), 4.36 – 4.25 (m, 2H), 4.14 (t, J = 12.2 Hz,

2H), 4.02 (d, J = 13.6 Hz, 1H), 3.86 – 3.73 (m, 1H), 3.67 (d, J = 12.7 Hz, 1H).

13C NMR (100 MHz, d6-DMSO) δ 146.55, 136.04, 132.86, 132.58, 130.27, 129.07,

125.84, 125.70, 125.66, 125.63, 125.59, 124.82, 123.85, 122.35, 92.25, 81.62, 79.15,

65.45, 63.07, 61.95, 61.68, 56.54.

19
F NMR (376 MHz, d6-DMSO) δ -62.55, -70.21, -72.10.

HRMS (ESI): [M-PF6]+ Calcd. for [C22H18F3INO]+ 496.0380, found 496.0388.

White solid, 53% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.60 (d, J = 8.7 Hz, 1H), 8.27 (s, 1H), 7.91 (d, J =

8.3 Hz, 2H), 7.75 (t, J = 10.1 Hz, 5H), 7.03 (s, 1H), 4.42 (t, J = 10.8 Hz, 1H), 4.35 –

4.25 (m, 2H), 4.13 (t, J = 10.1 Hz, 2H), 4.01 (d, J = 13.4 Hz, 1H), 3.80 (t, J = 11.7 Hz,

1H), 3.66 (d, J = 12.2 Hz, 1H).

99
13C NMR (100 MHz, d6-DMSO) δ 146.54 (s), 135.93 (s), 132.67 (d, J = 18.6 Hz),

130.27 (s), 129.07 (s), 125.83 (s), 124.82 (s), 124.33 (s), 118.06 (s), 112.55 (s), 92.16

(s), 82.81 (s), 79.13 (s), 65.42 (s), 63.10 (s), 61.95 (s), 61.67 (s), 56.55 (s).

HRMS (ESI): [M-PF6]+ Calcd. for [C22H18IN2O]+ 453.0458, found 453.0461.

Brown solid, 93% yield.

1H NMR (600 MHz, CD3CN) δ 8.64 (d, J = 7.8 Hz, 1H), 7.78 (s, 1H), 7.75 (t, J = 7.5

Hz, 1H), 7.72 – 7.67 (m, 2H), 7.55 – 7.49 (m, 2H), 7.15 (t, J = 8.8 Hz, 2H), 6.43 (s,

1H), 4.26 (d, J = 13.8 Hz, 1H), 4.24 – 4.15 (m, 2H), 4.14 – 4.02 (m, 3H), 3.75 (ddd, J

= 14.3, 10.7, 3.9 Hz, 1H), 3.58 (dd, J = 13.3, 2.1 Hz, 1H).

13C NMR (150 MHz, CD3CN) δ 164.24 (d, J = 250.5 Hz), 147.64, 136.67, 135.26 (d,

J = 8.8 Hz), 133.45, 130.93, 129.83, 127.33, 125.42, 116.82 (d, J = 3.5 Hz), 116.68 (d,

J = 22.7 Hz), 94.90, 78.31, 75.85, 67.58, 63.24, 62.72, 62.60, 57.15.

19F NMR (565 MHz, CD3CN) δ -72.12, -73.37, -109.10.

HRMS (ESI): [M-PF6]+ Calcd. for [C21H18FINO]+ 446.0412, found 446.0410.

100
White solid, 85% yield.

1H NMR (600 MHz, d6-DMSO) δ 8.59 (d, J = 7.8 Hz, 1H), 8.09 (s, 1H), 7.81 – 7.62

(m, 3H), 7.37 (d, J = 7.9 Hz, 2H), 7.21 (d, J = 7.8 Hz, 2H), 6.76 (s, 1H), 4.20 (t, J =

13.3 Hz, 1H), 4.00 (d, J = 11.9 Hz, 1H), 3.63 (t, J = 12.0 Hz, 1H), 3.55 (d, J = 12.0 Hz,

1H), 2.35 – 2.20 (m, 4H), 2.13 (s, 2H), 1.83 (d, J = 12.7 Hz, 2H), 1.68-1.57 (m, 1H).

13C NMR (150 MHz, d6-DMSO) δ 147.57, 140.38, 136.93, 132.44, 131.80, 129.98,

129.46, 129.30, 125.73, 124.71, 116.64, 93.03, 79.40, 77.80, 64.88, 64.54, 57.56, 21.56,

21.34, 21.07, 20.18.

HRMS (ESI): [M-PF6]+ Calcd. for [C23H23IN]+440.0870, found 440.0837.

Yellow solid, 67% yield.

1H NMR (400 MHz, CDCl3) δ 9.12 (d, J = 6.1 Hz, 1H), 8.53 – 8.47 (m, 2H), 7.76 –

7.70 (m, 2H), 7.66 – 7.61 (m, 1H), 7.55 (d, J = 7.5 Hz, 1H), 7.43 (d, J = 8.5 Hz, 2H),

7.38 – 7.34 (m, 2H), 7.28 (s, 2H), 6.65 – 6.61 (m, 1H), 6.45 – 6.39 (m, 2H), 4.13 (d, J

= 13.4 Hz, 1H), 4.06 – 3.95 (m, 2H), 3.92 – 3.81 (m, 1H), 3.43 (d, J = 6.9 Hz, 4H), 2.74

101
(td, J = 10.6, 4.5 Hz, 1H), 2.60 (t, J = 10.6 Hz, 1H), 2.49 – 2.42 (m, 1H), 2.37 – 2.26

(m, 1H), 1.22 – 1.20 (m, 6H).

13C NMR (100 MHz, CDCl3) δ 163.56, 163.36, 157.85, 153.16, 148.45, 147.63, 136.79,

136.07, 133.95, 133.68, 133.28, 131.62, 130.35, 129.34, 129.11, 126.74, 125.35,

118.38, 110.55, 109.29, 108.49, 96.57, 94.93, 78.49, 73.80, 69.12, 60.49, 54.40, 45.31,

41.93, 26.90, 26.41, 12.60.

HRMS (ESI): [M-PF6]+ Calcd. for [C36H34ClIN3O3]+ 718.1328, found 718.1346.

White solid, 70% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.59 (d, J = 8.1 Hz, 1H), 8.25 (s, 1H), 7.76 – 7.72

(m, 2H), 7.57 – 7.49 (m, 4H), 6.99 (s, 1H), 4.40 (d, J = 17.0 Hz, 2H), 4.35 – 4.25 (m,

2H), 4.08 (td, J = 30.1, 27.2, 12.7 Hz, 4H), 3.86 – 3.74 (m, 1H), 3.65 (d, J = 12.6 Hz,

1H).

13C NMR (100 MHz, d6-DMSO) δ 146.52, 136.20, 132.53, 132.22, 131.97, 130.17,

129.03, 125.81, 124.75, 123.50, 119.87, 93.19, 83.66, 82.59, 81.06, 79.04, 65.68, 62.90,

61.91, 61.66, 56.43.

HRMS (ESI): [M-PF6]+ Calcd. for [C23H19INO]+ 452.0506, found 452.0510.

102
4.7 Synthesis of Electron-Deficient Allenes 4

A mixture of 3 (0.5 mmol) and Et3N (0.5 mmol, 1 equiv) in 5 mL of CH2Cl2 (5 mL)

was stirred at room temperature for 1 h. The product was separated by suction filtration

and washed by CH2Cl2 to give the product 4.

White solid, 88% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.79 (dd, J = 6.8, 1.9 Hz, 1H), 8.44 (s, 1H), 8.03 (s,

1H), 7.78 – 7.71 (m, 2H), 7.71 – 7.65 (m, 2H), 7.56 (s, 1H), 7.54 (s, 1H), 7.48 – 7.42

(m, 1H), 4.49 (t, J = 11.6 Hz, 1H), 4.35 (d, J = 12.5 Hz, 1H), 4.28 – 4.01 (m, 7H).

13C NMR (100 MHz, d6-DMSO) δ 195.44, 148.62, 135.10, 132.96, 130.97, 130.76,

130.63, 129.52, 129.20, 129.01, 124.82, 122.30, 121.35, 112.79, 78.09, 64.18, 63.20,

60.86, 60.76.

HRMS (ESI): [M-PF6]+ Calcd. For [C21H18ClINO]+ 462.0116, found 462.0153.

103
White solid, 85% yield.

1H NMR (600 MHz, d6-DMSO) δ 8.88 – 8.65 (m, 1H), 8.45 (s, 1H), 8.02 (s, 1H), 7.77

– 7.70 (m, 2H), 7.66 (dd, J = 4.3, 3.1 Hz, 2H), 7.53 – 7.42 (m, 4H), 4.53 (t, J = 12.4 Hz,

1H), 4.36 (d, J = 12.1 Hz, 1H), 4.30 – 4.04 (m, 6H).

13C NMR (150 MHz, d6-DMSO) δ 195.22, 148.66, 132.95, 130.87, 130.75, 130.49,

130.17, 129.51, 129.07, 128.95, 124.83, 122.23, 121.23, 113.80, 78.01, 64.11, 63.32,

60.87, 60.73.

HRMS (ESI): [M-PF6]+ Calcd. for [C21H19INO]+ 428.0506, found 428.0500.

White solid, 14% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.79 (d, J = 8.4 Hz, 1H), 8.45 (s, 1H), 8.13 (s, 1H),

7.91 – 7.80 (m, 4H), 7.74 (tt, J = 8.3, 3.9 Hz, 2H), 7.51 – 7.44 (m, 1H), 4.47 (t, J = 11.1

Hz, 1H), 4.37 (d, J = 12.5 Hz, 1H), 4.30 – 4.07 (m, 6H).

104
13C NMR (100 MHz, d6-DMSO) δ 196.30, 148.57, 134.55, 132.93, 131.05, 130.44,

129.71, 129.04, 126.18, 124.81, 122.32, 121.46, 112.64, 78.20, 64.28, 63.17, 60.79,

54.89.

19
F NMR (376 MHz, d6-DMSO) δ -62.26, -70.21, -72.10.

HRMS (ESI): [M-PF6]+ Calcd. for [C22H18F3INO]+ 496.0380, found 496.0389.

White solid, 75% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.79 (d, J = 7.9 Hz, 1H), 8.45 (s, 1H), 8.11 (s, 1H),

7.95 (d, J = 8.2 Hz, 2H), 7.85 (d, J = 8.2 Hz, 2H), 7.74 (p, J = 7.3 Hz, 2H), 7.46 (d, J =

7.9 Hz, 1H), 4.49 – 4.33 (m, 2H), 4.30 – 4.06 (m, 6H).

13C NMR (100 MHz, d6-DMSO) δ 196.69 (s), 148.55 (s), 135.07 (s), 133.19 (s), 132.92

(s), 131.08 (s), 130.36 (s), 129.69 (s), 129.05 (s), 124.81 (s), 122.34 (s), 121.56 (s),

118.41 (s), 112.70 (s), 112.40 (s), 78.16 (s), 64.30 (s), 63.09 (s), 60.83 (d, J = 6.6 Hz).

HRMS (ESI): [M-PF6]+ Calcd. for [C22H18IN2O]+ 453.0458, found 453.0462.

105
White solid, 83% yield.

1H NMR (600 MHz, d6-DMSO) δ 8.80 (s, 1H), 8.44 (s, 1H), 8.03 (s, 1H), 7.73 (s, 4H),

7.45 (s, 1H), 7.33 (s, 2H), 4.50 (t, J = 12.2 Hz, 1H), 4.35 (d, J = 12.4 Hz, 1H), 4.29-

4.02 (m, 6H).

13C NMR (150 MHz, d6-DMSO) δ 194.92, 163.24 (d, J = 248.5 Hz), 148.60, 132.91,

131.39, 131.36 (d, J = 8.5 Hz), 130.79 (d, J = 23.1 Hz), 131.33, 130.86, 130.71, 128.96,

124.79, 122.24, 121.26, 116.54 (d, J = 21.9 Hz), 112.78, 77.94, 64.12, 63.15, 60.83,

60.73.

19F NMR (565 MHz, d6-DMSO) δ -69.46, -70.42, -108.79.

HRMS (ESI): [M-PF6]+ Calcd. For [C21H18FINO]+ 446.0412, found 446.0370.

White solid, 91% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.84 – 8.73 (m, 1H), 8.33 (s, 1H), 7.91 (s, 1H), 7.77

– 7.66 (m, 2H), 7.51 (d, J = 8.0 Hz, 2H), 7.44 – 7.36 (m, 1H), 7.29 (d, J = 8.0 Hz, 2H),

106
4.25 (d, J = 11.4 Hz, 1H), 4.10 – 3.89 (m, 3H), 2.46 (dd, J = 13.2, 8.1 Hz, 1H), 2.11 (d,

J = 11.2 Hz, 1H), 1.88 (d, J = 12.0 Hz, 2H), 1.81 – 1.62 (m, 2H).

13C NMR (100 MHz, d6-DMSO) δ 195.64, 149.45, 140.31, 132.91, 131.02, 130.69,

130.08, 128.92, 128.83, 127.45, 124.67, 122.08, 120.68, 112.42, 77.17, 66.27, 65.06,

21.01, 20.46, 20.26, 19.03.

HRMS (ESI): [M-PF6]+ Calcd. for [C23H23IN]+ 440.0870, found 440.0862.

Yellow solid, 70% yield.

1H NMR (400 MHz, CDCl3) δ 9.25 (d, J = 6.0 Hz, 1H), 8.78 (d, J = 7.3 Hz, 1H), 8.53

(s, 1H), 8.02 (s, 1H), 7.80 (s, 1H), 7.59 (q, J = 7.5 Hz, 3H), 7.43 (s, 2H), 7.41 (s, 2H),

7.39 (s, 2H), 6.68 – 6.63 (m, 1H), 6.44 (s, 1H), 4.33 – 4.28 (m, 1H), 4.21 (d, J = 6.7 Hz,

1H), 4.10 (s, 1H), 3.99 (d, J = 8.6 Hz, 1H), 3.43 (d, J = 6.9 Hz, 5H), 2.54 – 2.32 (m,

4H), 1.23 (d, J = 5.8 Hz, 6H).

13C NMR (100 MHz, CDCl3) δ 194.48, 163.65, 157.88, 153.22, 149.21, 148.41, 137.06,

133.69, 131.66, 131.33, 130.47, 130.35, 130.27, 130.13, 128.67, 128.17, 126.04,

124.55, 122.92, 115.35, 110.62, 109.39, 108.53, 96.67, 73.05, 63.41, 62.50, 45.36,

41.01, 25.60, 25.49, 12.63.

HRMS (ESI): [M-PF6]+ Calcd. for [C36H34ClIN3O3]+ 718.1328, found 718.1345.

107
White solid, 90% yield.

1H NMR (400 MHz, d6-DMSO) δ 8.81 – 8.76 (m, 1H), 8.43 (s, 1H), 8.04 (s, 1H), 7.77

– 7.69 (m, 2H), 7.66 (d, J = 8.1 Hz, 2H), 7.57 (d, J = 8.1 Hz, 2H), 7.47 – 7.41 (m, 1H),

4.47 (t, J = 11.5 Hz, 1H), 4.35 (d, J = 19.5 Hz, 2H), 4.17 (ddt, J = 45.9, 16.3, 9.0 Hz,

6H).

13C NMR (100 MHz, d6-DMSO) δ 195.77, 148.53, 132.86, 132.56, 130.87, 130.63,

130.54, 129.17, 124.73, 122.22, 121.28, 113.10, 82.94, 77.98, 64.11, 63.13, 60.79,

60.68.

HRMS (ESI): [M-PF6]+ Calcd. for [C23H19INO]+ 452.0506, found 452.0510.

4.8 Synthesis of Model Compound 7

H PF6 H PF6
O I O Cl
I Et3N (1 equiv.)
N N
CH3CN, r.t., 1 h

Cl H

3a 4a
O
H
SH I PF6
N H

(1 equiv.)
S Cl
CH3CN/H2O 3:1, r.t., 3 h

108
A mixture of 1H-isoindolium 3a (0.2 mmol) and Et3N (0.2 mmol, 1 equiv) in 3 mL of

CH3CN was stirred at room temperature for 1 h. Then, benzyl mercaptan (0.2 mmol, 1

equiv) and 1 mL of H2O was added to the mixture and stirred for 3 h. After the reaction,

the reaction mixture was diluted with 3 mL of CH2Cl2. The mixture was then extracted

with CH2Cl2 (3 mL × 2). The solvent was removed by rotary evaporation and purified

by flash column chromatography using CH2Cl2/CH3OH (50:1) as the eluent to give the

product in 63% isolated yield.

Yellow solid, 63% yield.

1H NMR (400 MHz, CD3CN) δ 8.68 (d, J = 6.9 Hz, 1H), 7.77 – 7.64 (m, 3H), 7.62 –

7.55 (m, 3H), 7.48 (d, J = 8.4 Hz, 2H), 7.34 – 7.25 (m, 3H), 7.20 (d, J = 6.6 Hz, 2H),

7.15 (s, 1H), 6.44 (s, 1H), 3.94 – 3.84 (m, 2H), 3.81 (d, J = 12.8 Hz, 1H), 3.77 – 3.64

(m, 2H), 3.48 – 3.36 (m, 2H), 3.32 (d, J = 13.1 Hz, 1H), 3.12 (t, J = 12.3 Hz, 2H).

13C NMR (100 MHz, CD3CN) δ 149.72, 139.23, 137.64, 136.97, 136.16, 134.16,

133.78, 132.41, 131.93, 131.70, 131.19, 131.00, 130.59, 130.10, 129.22, 127.75,

126.51, 74.18, 73.48, 66.24, 63.73, 63.38, 55.55, 38.65.

HRMS (ESI): [M-PF6]+ Calcd. for [C28H26ClINOS]+ 586.0463, found 586.0461.

109
4.9 Modification of Peptides Using Isolated Electron-Deficient Allene Reagents 4

A mixture of 10 µL of peptides 1 (1 mM in H2O), 5 µL of allene reagents 4 (10 mM in

CH3CN), 5 µL of CH3CN and 80 µL of 50 mM pH 8.0 potassium phosphate buffer was

treated in a 1.5 mL Eppendorf tube at 25 ºC for 4 h. The modified product was

characterized by LC–MS and LC–MS/MS analysis.

4.10 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using Isolated Electron-Deficient Allene Reagents 4 at Different pH Values

A mixture of 20 μL of peptides 1a (1 mM in H2O), 2 μL of allene reagents 4a (10 mM

in CH3CN), 18 μL of CH3CN and 160 μL of 50 mM potassium phosphate buffer with

different pH values was treated in a 1.5 mL Eppendorf tube at 25 ºC for 0–4 h. At each

time point, 25 μL of the resulting mixture was collected and diluted with 25 µL of Milli-

Q® water. The resulting mixture was characterized by LC–MS and LC–MS/MS

analysis to determine the conversion.

4.11 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using Isolated Electron-Deficient Allene Reagents 4 at Different Temperatures

A mixture of 20 μL of peptides 1a (1 mM in H2O), 2 μL of allene reagents 4a (10 mM

in CH3CN), 18 μL of CH3CN and 160 μL of 50 mM pH 8.0 potassium phosphate buffer

was treated in a 1.5 mL Eppendorf tube at different temperatures for 0–4 h. At each

time point, 25 μL of the resulting mixture was collected and diluted with 25 µL of Milli-

Q® water. The resulting mixture was characterized by LC–MS and LC–MS/MS

analysis to determine the conversion.

110
4.12 Studies on the Stability of the Modified Peptide 5a

A mixture of 25 μL of modified peptide 5a (0.1 mM in 50 mM pH 8.0 PBS buffer/

CH3CN (9:1)) and 25 μL of thiol-containing reagents (L-cysteine, DL-homocysteine,

and dithiothreitol (DTT)), reducing reagent (sodium ascorbate) (50 mM in H2O) was

treated in a 1.5 mL Eppendorf tube at 25 ºC for 3 h. The resulting mixture was

characterized by LC–MS and LC–MS/MS analysis to determine the conversion.

4.13 Modification of Proteins Using Isolated Electron-Deficient Allene Reagents 4

A mixture of 10 μL of proteins (BSA, insulin, RNaseA or lysozyme) (1 mM in 50 mM

pH 7.4 PBS buffer), 10 μL of allene reagents (10 mM in CH3CN), 80 μL of 50 mM pH

7.4 PBS buffer was treated in a 1.5 mL Eppendorf tube at 25 ºC for 4 h. The modified

product was purified by Bio-Rad Bio-Spin® 6 column prior to LC–MS analysis.

4.14 Sequential Modification of Proteins via Huisgen Azide-Alkyne Cycloaddition

A mixture of 50 μL of alkyne-functionalized proteins BSA-3 (0.1 mM in 50 mM PBS

buffer), 5 μL of rhodamine azide 10 (50 mM in DMSO), 5 μL of TBTA (5 mM in

DMSO), 5 μL of TCEP (5 mM in H2O), 5 μL of CuSO4 solution (5 mM in H2O) and

30 μL of 50 mM PBS buffer was treated in a 1.5 mL Eppendorf tube at 37 ºC for 1 h.

The modified product was characterized by LC–MS analysis.

4.15 SDS-PAGE Analysis of Native and Modified Proteins

The native BSA and modified proteins (BSA-3, BSA-4, BSA-10) (10 μL) was mixed

with 2X loading buffer (10 μL) in a 0.5 mL Eppendorf tube and then boiled for 5 min.

111
Samples were analysed by SDS-PAGE by loading a sample of the boiled solution (10

μL) in each lane of a 12% SDS-PAGE gel and running in a SE 250 Mini-Vertical Unit

(Amersham, USA) at 125 V at room temperature for 110 min. After SDS-PAGE

separation, the fluorescence of gel was scanned by Azure C600 gel-imaging system.

The gel was stained with Coomassie Blue G250 for 10 min, and destained in the

methanol/acetic acid solution (45% methanol, 10% acetic acid) for 10 min. The

destained gel was scanned by Azure C600 gel-imaging system.

4.16 Preparation of in situ Generated Electron-Deficient Allene Reagents 4

A mixture of propargylamine-based 1,6-diyne 2 (5 μmol) and iodine monochloride

(5.25 μmol, 0.26 μL) in 500 μL of CH3CN was stirred at room temperature for 1 h.

After that, the reaction mixture was diluted with 450 μL of CH3CN and 50 μL of TCEP

solution (100 mM in H2O) was added to the mixture. The reagent was stored at –20 °C

for repeated use.

4.17 Modification of Peptides Using in situ Generated Electron-Deficient Allene

Reagents 4

A mixture of 10 µL of peptides 1 (1 mM in H2O), 2 µL of in situ generated allene

reagents 4 (5 mM in CH3CN), 8 µL of CH3CN and 80 µL of 50 mM pH 8.0 potassium

phosphate buffer was treated in a 1.5 mL Eppendorf tube at 25 ºC for 4 h. The modified

product was characterized by LC–MS and LC–MS/MS analysis.

112
4.18 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using in situ Generated Electron-Deficient Allene Reagents 4 at Different pH

Values

A mixture of 20 μL of peptides 1a (1 mM in H2O), 4 μL of in situ generated allene

reagents 4a (5 mM in CH3CN), 16 μL of CH3CN and 160 μL of 50 mM potassium

phosphate buffer with different pH values was treated in a 1.5 mL Eppendorf tube at

25 ºC for 0–4 h. At each time point, 25 μL of the resulting mixture was collected and

diluted with 25 µL of Milli-Q® water. The resulting mixture was characterized by LC–

MS and LC–MS/MS analysis to determine the conversion.

4.19 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using in situ Generated Electron-Deficient Allene Reagents 4 at Different

Temperatures

A mixture of 20 μL of peptides 1a (1 mM in H2O), 4 μL of in situ generated allene

reagents 4a (5 mM in CH3CN), 16 μL of CH3CN and 160 μL of 50 mM pH 8.0

potassium phosphate buffer was treated in a 1.5 mL Eppendorf tube at different

temperatures for 0–4 h. At each time point, 25 μL of the resulting mixture was collected

and diluted with 25 µL of Milli-Q® water. The resulting mixture was characterized by

LC–MS and LC–MS/MS analysis to determine the conversion.

113
4.20 Time Course Studies on the Modification of Peptide AYEMWCFHQR 1a

Using in situ Generated Electron-Deficient Allene Reagents 4a-d

A mixture of 20 μL of peptides 1a (1 mM in H2O), 4 μL of the in situ generated allene

reagents 4a-d (5 mM in CH3CN), 16 μL of CH3CN and 160 μL of 50 mM pH 8.0

potassium phosphate buffer was treated in a 1.5 mL Eppendorf tube at 25 ºC for 0–4 h.

At each time point, 25 μL of the resulting mixture was collected and diluted with 25 µL

of Milli-Q® water. The resulting mixture was characterized by LC–MS and LC–

MS/MS analysis to determine the conversion.

4.21 Modification of Proteins Using in situ Generated Electron-Deficient Allene

Reagents 4

A mixture of 5 μL of BSA protein (1 mM in 50 mM pH 7.4 PBS buffer), 10 μL of

allene reagents (5 mM in CH3CN), 85 μL of 50 mM pH 7.4 PBS buffer was treated in

a 1.5 mL Eppendorf tube at 25 ºC for 4 h. The modified product was purified by Bio-

Rad Bio-Spin® 6 column prior to LC–MS analysis.

114
4.22 MS Spectra

Figure 4.1 Extracted ion chromatogram of native peptide AYEMWCFHQR 1a.

Figure 4.2 Mass spectrum of native peptide AYEMWCFHQR 1a.

115
b-ions 72 235 364 495 681 784 931 1068 1196

H 2N A Y E M W C F H Q R COOH

1299 1136 1007 876 690 587 440 303 175 y-ions

Figure 4.3 MS/MS spectrum of native peptide AYEMWCFHQR 1a.

Figure 4.4 Extracted ion chromatogram of cysteine-modified peptide 5a.

116
Figure 4.5 Mass spectrum of cysteine-modified peptide 5a.

Figure 4.6 MS/MS spectrum of cysteine-modified peptide 5a.

117
Figure 4.7 Extracted ion chromatogram of cysteine-modified peptide 5b.

Figure 4.8 Mass spectrum of cysteine-modified peptide 5b.

118
Figure 4.9 MS/MS spectrum of cysteine-modified peptide 5b.

119
Figure 4.10 Extracted ion chromatogram of cysteine-modified peptide 5c.

Figure 4.11 Mass spectrum of cysteine-modified peptide 5c.

120
Figure 4.12 MS/MS spectrum of cysteine-modified peptide 5c.

121
Figure 4.13 Extracted ion chromatogram of cysteine-modified peptide 5d.

Figure 4.14 Mass spectrum of cysteine-modified peptide 5d.

122
Figure 4.15 MS/MS spectrum of cysteine-modified peptide 5d.

123
Figure 4.16 Extracted ion chromatogram of cysteine-modified peptide 5e.

Figure 4.17 Mass spectrum of cysteine-modified peptide 5e.

124
Figure 4.18 MS/MS spectrum of cysteine-modified peptide 5e.

125
Figure 4.19 Extracted ion chromatogram of cysteine-modified peptide 5f.

Figure 4.20 Mass spectrum of cysteine-modified peptide 5f.

126
Figure 4.21 MS/MS spectrum of cysteine-modified peptide 5f.

127
Figure 4.22 Extracted ion chromatogram of cysteine-modified peptide 5g.

Figure 4.23 Mass spectrum of cysteine-modified peptide 5g.

128
Figure 4.24 MS/MS spectrum of cysteine-modified peptide 5g.

129
Figure 4.25 Extracted ion chromatogram of cysteine-modified peptide 5h.

Figure 4.26 Mass spectrum of cysteine-modified peptide 5h.

130
Figure 4.27 MS/MS spectrum of cysteine-modified peptide 5h.

131
NH 2

H 3C OH OH SH
O H O H O H O H O
H 2N N N N N O
N N N N N
H O H O H O H O H OH
OH OH OH NH 2 OH
O
STSSSCNLSK

Figure 4.28 Extracted ion chromatogram of native peptide STSSSCNLSK 1b.

Figure 4.29 Mass spectrum of native peptide STSSSCNLSK 1b.

132
b-ions 88 189 276 363 450 667 780 867 1241

H 2N S T S S S C N L S K COOH

926 825 738 651 564 461 347 234 147 y-ions

Figure 4.30 MS/MS spectrum of native peptide STSSSCNLSK 1b.

Figure 4.31 Extracted ion chromatogram of cysteine-modified peptide 6a.

133
Figure 4.32 Mass spectrum of cysteine-modified peptide 6a.

Figure 4.33 MS/MS spectrum of cysteine-modified peptide 6a.

134
Figure 4.34 Extracted ion chromatogram of cysteine-modified peptide 6b.

Figure 4.35 Mass spectrum of cysteine-modified peptide 6b.

135
Figure 4.36 MS/MS spectrum of cysteine-modified peptide 6b.

136
Figure 4.37 Extracted ion chromatogram of cysteine-modified peptide 6c.

Figure 4.38 Mass spectrum of cysteine-modified peptide 6c.

137
Figure 4.39 MS/MS spectrum of cysteine-modified peptide 6c.

138
Figure 4.40 Extracted ion chromatogram of cysteine-modified peptide 6d.

Figure 4.41 Mass spectrum of cysteine-modified peptide 6d.

139
Figure 4.42 MS/MS spectrum of cysteine-modified peptide 6d.

140
Figure 4.43 Extracted ion chromatogram of cysteine-modified peptide 6e.

Figure 4.44 Mass spectrum of cysteine-modified peptide 6e.

141
Figure 4.45 MS/MS spectrum of cysteine-modified peptide 6e.

142
Figure 4.46 Extracted ion chromatogram of cysteine-modified peptide 6f.

Figure 4.47 Mass spectrum of cysteine-modified peptide 6f.

143
Figure 4.48 MS/MS spectrum of cysteine-modified peptide 6f.

144
Figure 4.49 Extracted ion chromatogram of cysteine-modified peptide 6g.

Figure 4.50 Mass spectrum of cysteine-modified peptide 6g.

145
Figure 4.51 MS/MS spectrum of cysteine-modified peptide 6g.

146
Figure 4.52 Extracted ion chromatogram of cysteine-modified peptide 6h.

Figure 4.53 Mass spectrum of cysteine-modified peptide 6h.

147
Figure 4.54 MS/MS spectrum of cysteine-modified peptide 6h.

148
Figure 4.55 Extracted ion chromatogram of native peptide KSTFC 1c.

Figure 4.56 Mass spectrum of native peptide KSTFC 1c.

149
Figure 4.57 MS/MS spectrum of native peptide KSTFC 1c.

Figure 4.58 Extracted ion chromatogram of cysteine-modified peptide 8a.

150
Figure 4.59 Mass spectrum of cysteine-modified peptide 8a.

Figure 4.60 MS/MS spectrum of cysteine-modified peptide 8a.

151
Figure 4.61 Extracted ion chromatogram of native peptide CSKFR 1d.

Figure 4.62 Mass spectrum of native peptide CSKFR 1d.

152
Figure 4.63 MS/MS spectrum of native peptide CSKFR 1d.

153
Figure 4.64 Extracted ion chromatogram of cysteine-modified peptide 9a.

Figure 4.65 Mass spectrum of cysteine-modified peptide 9a.

154
Figure 4.66 MS/MS spectrum of cysteine-modified peptide 9a.

155
Figure 4.67 Extracted ion chromatogram of native peptide STSSSANLSK 1e.

Figure 4.68 Mass spectrum of native peptide STSSSANLSK 1e.

156
b-ions 88 198 276 363 450 521 635 748 835

H 2N S T S S S A N L S K COOH

981 894 793 706 619 532 347 234 147 y-ions

Figure 4.69 MS/MS spectrum of native peptide STSSSANLSK 1e.

157
Figure 4.70 Extracted ion chromatogram of native peptide STSSSHNLSK 1f.

Figure 4.71 Mass spectrum of native peptide STSSSHNLSK 1f.

158
b-ions 88 189 276 363 450 587 673 786 873

H 2N S T S S S H N L S K COOH

960 859 772 685 598 461 347 234 147 y-ions

Figure 4.72 MS/MS spectrum of STSSSHNLSK 1f.

159
Figure 4.73 Extracted ion chromatogram of native peptide AYEMWSFHQR 1g.

Figure 4.74 Mass spectrum of native peptide AYEMWSFHQR 1g.

160
b-ions 72 235 364 495 681 768 915 1052 1180

H 2N A Y E M W S F H Q R COOH

1283 1120 991 860 674 587 440 303 175 y-ions

Figure 4.75 MS/MS spectrum of native peptide AYEMWSFHQR 1g.

161
Figure 4.76 Extracted ion chromatogram of native peptide WSKFR 1h.

Figure 4.77 Mass spectrum of native peptide WSKFR 1h.

162
Figure 4.78 MS/MS spectrum of native peptide WSKFR 1h.

163
Figure 4.79 Extracted ion chromatogram of native peptide YSKFR 1i.

Figure 4.80 Mass spectrum of native peptide YSKFR 1i.

164
Figure 4.81 MS/MS spectrum of native peptide YSKFR 1i.

165
Figure 4.82 Extracted ion chromatogram of native peptide PSKFR 1j.

Figure 4.83 Mass spectrum of native peptide PSKFR 1j.

166
Figure 4.84 MS/MS spectrum of native peptide PSKFR 1j.

167
Figure 4.85 Extracted ion chromatogram of native peptide GSKFR 1k.

Figure 4.86 Mass spectrum of native peptide GSKFR 1k.

168
Figure 4.87 MS/MS spectrum of native peptide GSKFR 1k.

169
Figure 4.88 Extracted ion chromatogram of native peptide ISKFR 1l.

Figure 4.89 Mass spectrum of native peptide ISKFR 1l.

170
Figure 4.90 MS/MS spectrum of native peptide ISKFR 1l.

171
4.23 NMR Spectra

1H NMR

13C NMR

172
1H NMR

13C NMR

173
1H NMR

13C NMR

174
1H NMR

13C NMR

175
19F NMR

176
1H NMR

13C NMR

177
1H NMR

13C NMR

178
19F NMR

179
1H NMR

13C NMR

180
1H NMR

13C NMR

181
1H NMR

13C NMR

182
1H NMR

13C NMR

183
1H NMR

13C NMR

184
1H NMR

13C NMR

185
19F NMR

186
1H NMR

13C NMR

187
1H NMR

13C NMR

188
19F NMR

189
1H NMR

13C NMR

190
1H NMR

13C NMR

191
1H NMR

13C NMR

192
1H NMR

13C NMR

193
1H NMR

13C NMR

194
1H NMR

13C NMR

195
19F NMR

196
1H NMR

13C NMR

197
1H NMR

13C NMR

198
19F NMR

199
1H NMR

13C NMR

200
1H NMR

13C NMR

201
1H NMR

13C NMR

202
1H NMR

13C NMR

203
References

1. Soriano, E.; Fernández, I., Allenes and computational chemistry: from bonding

situations to reaction mechanisms. Chem. Soc. Rev. 2014, 43 (9), 3041-3105.

2. Yu, S.; Ma, S., Allenes in Catalytic Asymmetric Synthesis and Natural Product

Syntheses. Angew. Chem. Int. Ed. Engl. 2012, 51 (13), 3074-3112.

3. Hoffmann-Röder, A.; Krause, N., Synthesis and Properties of Allenic Natural

Products and Pharmaceuticals. Angew. Chem. Int. Ed. Engl. 2004, 43 (10), 1196-

1216.

4. Yokota, M.; Fuchibe, K.; Ueda, M.; Mayumi, Y.; Ichikawa, J., Facile

Synthesis of 1,1-Difluoroallenes via the Difluorovinylidenation of Aldehydes and

Ketones. Org. Lett. 2009, 11 (17), 3994-3997.

5. (a) Mori, N.; Obuchi, K.; Katae, T.; Sakurada, J.; Satoh, T., Alkenylation of

thiophenes and furans at the 2-position and a synthesis of allenes conjugated with

α,β-unsaturated ester with magnesium alkylidene carbenoids. Tetrahedron 2009,

65 (17), 3509-3517; (b) Satoh, T.; Kaneta, H.; Matsushima, A.; Yajima, M., A

new synthesis of β,γ-unsaturated esters and allenic esters with construction of a

carbon–carbon bond between α- and β-positions by the reaction of magnesium

alkylidene carbenoids with lithium ester enolates. Tetrahedron Lett. 2009, 50

(46), 6280-6285.

204
6. (a) Yu, X.; Ren, H.; Xiao, Y.; Zhang, J., Efficient Assembly of Allenes, 1,3‐

Dienes, and 4H‐Pyrans by Catalytic Regioselective Nucleophilic Addition to

Electron‐Deficient 1,3‐Conjugated Enynes. Chem. Eur. J. 2008, 14 (28), 8481-

8485; (b) Xiao, Y.; Zhang, J., Tetrasubstituted allenes by Pd0-catalyzed three-

component tandem Michael addition/cross-coupling reaction. Chem. Commun.

2010, 46 (5), 752-754.

7. Ma, S.; He, Q.; Jin, X., Tunable 1,3-Lithium Shift of Propargylic/Allenylic

Lithiums Formed by Conjugate Addition of 4-Aryl-3-butyn-1-enes with

Organolithiums and the Subsequent Transmetalation/Pd(0)-Catalyzed Coupling

Reaction with Aryl Halides. Synlett 2005, 2005 (3), 514-516.

8. (a) Fenández, I.; Monterde, M. I.; Plumet, J., On the base-induced isomerization

of cyclic propargylamides to cyclic allenamides. Tetrahedron Lett. 2005, 46

(36), 6029-6031; (b) Fotsing, J. R.; Banert, K., First Propargyl Azides

Bearing Strong Acceptor Substituents and Their Effective Conversion into Allenyl

Azides: Influence of the Electronic Effects of Substituents on the Reactivity of

Propargyl Azides. Eur. J. Org. Chem. 2005, 2005 (17), 3704-3714; (c) Lepore, S.

D.; Khoram, A.; Bromfield, D. C.; Cohn, P.; Jairaj, V.; Silvestri, M. A., Studies

on the Manganese-Mediated Isomerization of Alkynyl Carbonyls to Allenyl

Carbonyls. J. Org. Chem. 2005, 70 (18), 7443-7446; (d) Brossat, M.; Heck, M.-

P.; Mioskowski, C., Bistrimethylsilylpropargylic Ether: A Versatile Ambident

Synthon to Access Substituted Allenyne Ethers and α-Substituted Bispropargylic

Alcohols. J. Org. Chem. 2007, 72 (15), 5938-5941.

205
9. (a) Du, X.; Dai, Y.; He, R.; Lu, S.; Bao, M., New Application of N-

Halosuccinimide/PPh3 for the Halogenation of Propargyl Alcohols to Haloallenes.

Synth. Comm. 2009, 39 (21), 3940-3949; (b) Sakai, N.; Maruyama, T.;

Konakahara, T., Chemoselective Isomerization of Secondary-Type

Propargylic Alcohols to Propargylic/Allenic Bromides, and Brominated Dienes

with Appel-Type Reaction Conditions. Synlett 2009, 2009 (13), 2105-2108; (c) Pu,

X.; Ready, J. M., Direct and Stereospecific Synthesis of Allenes via Reduction of

Propargylic Alcohols with Cp2Zr(H)Cl. J. Am. Chem. Soc. 2008, 130 (33), 10874-

10875.

10. (a) Lo, V. K.-Y.; Wong, M.-K.; Che, C.-M., Gold-Catalyzed Highly

Enantioselective Synthesis of Axially Chiral Allenes. Org. Lett. 2008, 10 (3), 517-

519; (b) Lo, V. K.-Y.; Zhou, C.-Y.; Wong, M.-K.; Che, C.-M., Silver(I)-

mediated highly enantioselective synthesis of axially chiral allenes under thermal

and microwave-assisted conditions. Chem. Commun. 2010, 46 (2), 213-215.

11. Taylor, D. R., The Chemistry of Allenes. Chem. Rev. 1967, 67 (3), 317-359.

12. Johnson, R. P., Modern Allene Chemistry, Volumes 1−2 Edited by Norbert Krause

(University of Dortmund, Germany) and A. Stephen K. Hashmi (University of

Stuttgart, Germany). Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim. 2004.

13. Liu, L.; Ward, R. M.; Schomaker, J. M., Mechanistic Aspects and Synthetic

Applications of Radical Additions to Allenes. Chem. Rev. 2019, 119 (24),

12422-12490.

206
14. Ma, S., Carbopalladation of Allenes. In Handbook of Organopalladium Chemistry

for Organic Synthesis, Negishi, E., Ed.; Wiley: New York, 2002; pp 1491-1521.

15. Ma, S., Palladium-Catalyzed Two- or Three-Component Cyclization of

Functionalized Allenes. In Palladium in Organic Synthesis, Tsuji, J., Ed.

Springer Berlin Heidelberg: Berlin, Heidelberg, 2005; pp 183-210.

16. Krall, N.; da Cruz, F. P.; Boutureira, O.; Bernardes, G. J. L., Site-selective

protein-modification chemistry for basic biology and drug development. Nat.

Chem. 2015, 8 (2), 103-113.

17. Xue, L.; Karpenko, I. A.; Hiblot, J.; Johnsson, K., Imaging and manipulating

proteins in live cells through covalent labeling. Nat. Chem. Biol. 2015, 11 (12),

917- 923.

18. Boutureira, O.; Bernardes, G. J. L., Advances in Chemical Protein Modification.

Chem. Rev. 2015, 115 (5), 2174-2195.

19. Chudasama, V.; Maruani, A.; Caddick, S., Recent advances in the construction

of antibody–drug conjugates. Nat. Chem. 2016, 8 (2), 114-119.

20. Gunnoo, S. B.; Madder, A., Chemical Protein Modification through Cysteine.

ChemBioChem 2016, 17 (7), 529-553.

21. Tamura, T.; Hamachi, I., Chemistry for Covalent Modification of

Endogenous/Native Proteins: From Test Tubes to Complex Biological Systems. J.

Am. Chem. Soc. 2019, 141 (7), 2782-2799.

207
22. Koniev, O.; Wagner, A., Developments and recent advancements in the field of

endogenous amino acid selective bond forming reactions for bioconjugation.

Chem. Soc. Rev. 2015, 44 (15), 5495-5551.

23. Tanaka, K.; Fujii, Y.; Fukase, K., Site‐Selective and Nondestructive Protein

Labeling through Azaelectrocyclization‐Induced Cascade Reactions.

ChemBioChem 2008, 9 (15), 2392-2397.

24. Tung, C. L.; Wong, C. T. T.; Fung, E. Y. M.; Li, X., Traceless and

Chemoselective Amine Bioconjugation via Phthalimidine Formation in Native

Protein Modification. Org. Lett. 2016, 18 (11), 2600-2603.

25. Matos, M. J.; Oliveira, B. L.; Martínez-Sáez, N.; Guerreiro, A.; Cal, P. M. S.

D.; Bertoldo, J.; Maneiro, M.; Perkins, E.; Howard, J.; Deery, M. J.;

Chalker, J. M.; Corzana, F.; Jiménez-Osés, G.; Bernardes, G. a. J. L., Chemo-

and Regioselective Lysine Modification on Native Proteins. J. Am. Chem. Soc.

2018, 140 (11), 4004-4017.

26. (a) Dixon, H. B., Transamination of peptides. Biochem. J. 1964, 92 (3), 661-666;

(b) Dixon, H. B.; Moret, V., Removal of the N-terminal residue of a protein after

transamination. Biochem. J. 1965, 94 (2), 463-469; (c) Dixon, H. B.; Weitkamp,

L. R., Conversion of the N-terminal serine residue of corticotrophin into glycine.

Biochem. J. 1962, 84 (3), 462-468.

27. Kalia, J.; Raines, R. T., Hydrolytic Stability of Hydrazones and Oximes. Angew.

Chem. Int. Ed. Engl. 2008, 47 (39), 7523-7526.

208
28. (a) Gilmore, J. M.; Scheck, R. A.; Esser‐Kahn, A. P.; Joshi, N. S.; Francis,

M. B., N‐Terminal Protein Modification through a Biomimetic Transamination

Reaction. Angew. Chem. Int. Ed. Engl. 2006, 45 (32), 5307-5311; (b) MacDonald,

J. I.; Munch, H. K.; Moore, T.; Francis, M. B., One-step site-specific

modification of native proteins with 2-pyridinecarboxyaldehydes. Nat. Chem. Biol.

2015, 11 (5), 326-331.

29. (a) Chan, A. O.-Y.; Ho, C.-M.; Chong, H.-C.; Leung, Y.-C.; Huang, J.-S.;

Wong, M.-K.; Che, C.-M., Modification of N-Terminal α-Amino Groups of

Peptides and Proteins Using Ketenes. J. Am. Chem. Soc. 2012, 134 (5), 2589-2598;

(b) Chan, W.-K.; Ho, C.-M.; Wong, M.-K.; Che, C.-M., Oxidative Amide

Synthesis and N-Terminal α-Amino Group Ligation of Peptides in Aqueous

Medium. J. Am. Chem. Soc. 2006, 128 (46), 14796-14797.

30. Chen, D.; Disotuar, M. M.; Xiong, X.; Wang, Y.; Chou, D. H.-C., Selective

N-terminal functionalization of native peptides and proteins. Chem. Sci. 2017, 8

(4), 2717-2722.

31. Deng, J.-R.; Lai, N. C.-H.; Kung, K. K.-Y.; Yang, B.; Chung, S.-F.; Leung, A. S.-

L.; Choi, M.-C.; Leung, Y.-C.; Wong, M.-K., N-Terminal selective modification

of peptides and proteins using 2-ethynylbenzaldehydes. Commun. Chem. 2020, 3

(1), 1-9.

32. (a) Hooker, J. M.; Kovacs, E. W.; Francis, M. B., Interior Surface Modification

of Bacteriophage MS2. J. Am. Chem. Soc. 2004, 126 (12), 3718-3719; (b) Joshi,

209
N. S.; Whitaker, L. R.; Francis, M. B., A Three-Component Mannich-Type

Reaction for Selective Tyrosine Bioconjugation. J. Am. Chem. Soc. 2004, 126 (49),

15942-15943; (c) McFarland, J. M.; Joshi, N. S.; Francis, M. B.,

Characterization of a Three-Component Coupling Reaction on Proteins by

Isotopic Labeling and Nuclear Magnetic Resonance Spectroscopy. J. Am. Chem.

Soc. 2008, 130 (24), 7639-7644; (d) Romanini, D. W.; Francis, M. B., Attachment

of Peptide Building Blocks to Proteins Through Tyrosine Bioconjugation.

Bioconjugate Chem. 2008, 19 (1), 153-157; (e) Schlick, T. L.; Ding, Z.;

Kovacs, E. W.; Francis, M. B., Dual-Surface Modification of the Tobacco

Mosaic Virus. J. Am. Chem. Soc. 2005, 127 (11), 3718-3723; (f) Tilley, S. D.;

Francis, M. B., Tyrosine-selective protein alkylation using pi-allylpalladium

complexes. J. Am. Chem. Soc. 2006, 128 (4), 1080-1081.

33. (a) Ban, H.; Gavrilyuk, J.; Barbas, C. F., Tyrosine Bioconjugation through

Aqueous Ene-Type Reactions: A Click-Like Reaction for Tyrosine. J. Am. Chem.

Soc. 2010, 132 (5), 1523-1525; (b) Gavrilyuk, J.; Ban, H.; Nagano, M.;

Hakamata, W.; Barbas, C. F., Formylbenzene Diazonium Hexafluorophosphate

Reagent for Tyrosine-Selective Modification of Proteins and the Introduction of a

Bioorthogonal Aldehyde. Bioconjugate Chem. 2012, 23 (12), 2321-2328.

34. Moelbert, S.; Emberly, E.; Tang, C., Correlation between sequence

hydrophobicity and surface‐exposure pattern of database proteins. Protein Sci.

2004, 13 (3), 752-762.

210
35. Gevaert, K.; Damme, P. V.; Martens, L.; Vandekerckhove, J., Diagonal

reverse-phase chromatography applications in peptide-centric proteomics: Ahead

of catalogue-omics? Anal. Biochem. 2005, 345 (1), 18-29.

36. (a) Antos, J. M.; Francis, M. B., Selective Tryptophan Modification with Rhodium

Carbenoids in Aqueous Solution. J. Am. Chem. Soc. 2004, 126 (33), 10256-10257;

(b) Antos, J. M.; McFarland, J. M.; Iavarone, A. T.; Francis, M. B.,

Chemoselective Tryptophan Labeling with Rhodium Carbenoids at Mild pH. J.

Am. Chem. Soc. 2009, 131 (17), 6301-6308.

37. Foettinger, A.; Melmer, M.; Leitner, A.; Lindner, W., Reaction of the Indole

Group with Malondialdehyde: Application for the Derivatization of Tryptophan

Residues in Peptides. Bioconjugate Chem. 2007, 18 (5), 1678-1683.

38. Seki, Y.; Ishiyama, T.; Sasaki, D.; Abe, J.; Sohma, Y.; Oisaki, K.; Kanai, M.,

Transition Metal-Free Tryptophan-Selective Bioconjugation of Proteins. J.

Am. Chem. Soc. 2016, 138 (34), 10798-10801.

39. Stipanuk, M. H., SULFUR AMINO ACID METABOLISM: Pathways for

Production and Removal of Homocysteine and Cysteine. Annu. Rev. Nutr. 2004,

24 (1), 539-577.

40. (a) Kramer, J. R.; Deming, T. J., Preparation of Multifunctional and Multireactive

Polypeptides via Methionine Alkylation. Biomacromolecules 2012, 13 (6), 1719-

1723; (b) Kramer, J. R.; Deming, T. J., Reversible chemoselective tagging and

211
functionalization of methionine containing peptides. Chem. Commun. 2013,

49 (45), 5144.

41. Lin, S.; Yang, X.; Jia, S.; Weeks, A. M.; Hornsby, M.; Lee, P. S.; Nichiporuk, R.

V.; Iavarone, A. T.; Wells, J. A.; Toste, F. D.; Chang, C. J., Redox-based reagents

for chemoselective methionine bioconjugation. Science 2017, 355 (6325), 597-

602.

42. (a) Lee, K. J.; Kang, D.; Park, H.-S., Site-Specific Labeling of Proteins Using

Unnatural Amino Acids. Mol. Cells 2019, 42 (5), 386-396; (b) Spicer, C. D.; Davis,

B. G., Selective chemical protein modification. Nat. Commun. 2014, 5 (1), 4740.

43. (a) Saxon, E., Cell Surface Engineering by a Modified Staudinger Reaction.

Science 2000, 287 (5460), 2007-2010; (b) Kristi, L. K.; Eliana, S.; David, A. T.;

Carolyn, R. B., Incorporation of azides into recombinant proteins for

chemoselective modification by the Staudinger ligation. Proc. Natl. Acad. Sci.

U.S.A. 2001, 99 (1), 19-24; (c) Tsao, M.-L.; Tian, F.; Schultz, P. G., Selective

Staudinger Modification of Proteins Containing p-Azidophenylalanine.

ChemBioChem 2005, 6 (12), 2147-2149; (d) Rostovtsev, V. V.; Green, L. G.;

Fokin, V. V.; Sharpless, K. B., A Stepwise Huisgen Cycloaddition Process:

Copper(I)‐Catalyzed Regioselective “Ligation” of Azides and Terminal Alkynes.

Angew. Chem. Int. Ed. Engl. 2002, 41 (14), 2596-2599; (e) Tornøe, C. W.;

Christensen, C.; Meldal, M., Peptidotriazoles on Solid Phase: [1,2,3]-Triazoles by

Regiospecific Copper(I)-Catalyzed 1,3-Dipolar Cycloadditions of Terminal

Alkynes to Azides. J. Org. Chem. 2002, 67 (9), 3057-3064.

212
44. (a) Agard, N. J.; Prescher, J. A.; Bertozzi, C. R., A Strain-Promoted [3 + 2]

Azide−Alkyne Cycloaddition for Covalent Modification of Biomolecules in

Living Systems. J. Am. Chem. Soc. 2004, 126 (46), 15046-15047; (b) Jeremy, M.

B.; Jennifer, A. P.; Scott, T. L.; Nicholas, J. A.; Pamela, V. C.; Isaac, A. M.;

Anderson, L.; Julian, A. C.; Carolyn, R. B., Copper-free click chemistry for

dynamic in vivo imaging. Proc. Natl. Acad. Sci. U.S.A. 2007, 104 (43), 16793-

16797; (c) Laughlin, S. T.; Baskin, J. M.; Amacher, S. L.; Bertozzi, C. R., In

Vivo Imaging of Membrane-Associated Glycans in Developing Zebrafish. Science

2008, 320 (5876), 664-667; (d) Mbua, N. E.; Guo, J.; Wolfert, M. A.; Steet,

R.; Boons, G. J., Strain‐Promoted Alkyne–Azide Cycloadditions (SPAAC) Reveal

New Features of Glycoconjugate Biosynthesis. ChemBioChem 2011, 12 (12),

1912-1921.

45. (a) Blackman, M. L.; Royzen, M.; Fox, J. M., Tetrazine Ligation: Fast

Bioconjugation Based on Inverse-Electron-Demand Diels−Alder Reactivity. J.

Am. Chem. Soc. 2008, 130 (41), 13518-13519; (b) Lang, K.; Davis, L.; Torres-

Kolbus, J.; Chou, C.; Deiters, A.; Chin, J. W., Genetically encoded norbornene

directs site-specific cellular protein labelling via a rapid bioorthogonal reaction.

Nat. Chem. 2012, 4 (4), 298-304; (c) Lang, K.; Chin, J. W., Bioorthogonal

Reactions for Labeling Proteins. ACS Chem. Biol. 2014, 9 (1), 16-20; (d) Nikić, I.;

Kang, J. H.; Girona, G. E.; Aramburu, I. V.; Lemke, E. A., Labeling proteins

on live mammalian cells using click chemistry. Nat. Protoc. 2015, 10 (5), 780-791.

213
46. (a) Song, W.; Wang, Y.; Qu, J.; Madden, M. M.; Lin, Q., A Photoinducible

1,3-Dipolar Cycloaddition Reaction for Rapid, Selective Modification of

Tetrazole-Containing Proteins. Angew. Chem. Int. Ed. Engl. 2008, 47 (15), 2832-

2835; (b) Song, W.; Wang, Y.; Qu, J.; Lin, Q., Selective Functionalization of a

Genetically Encoded Alkene-Containing Protein via “Photoclick Chemistry” in

Bacterial Cells. J. Am. Chem. Soc. 2008, 130 (30), 9654-9655; (c) Yu, Z.; Pan, Y.;

Wang, Z.; Wang, J.; Lin, Q., Genetically Encoded Cyclopropene Directs Rapid,

Photoclick-Chemistry-Mediated Protein Labeling in Mammalian Cells. Angew.

Chem. Int. Ed. Engl. 2012, 51 (42), 10600-10604.

47. Ko, H.-M.; Deng, J.-R.; Cui, J.-F.; Kung, K. K.-Y.; Leung, Y.-C.; Wong, M.-K.,

Selective modification of alkyne-linked peptides and proteins by cyclometalated

gold(III) (C^N) complex-mediated alkynylation. Bioorg. Med. Chem. 2020, 28

(7), 115375.

48. Poole, L. B., The basics of thiols and cysteines in redox biology and chemistry.

Free Radic. Biol. Med. 2015, 80, 148-157.

49. Ochtrop, P.; Hackenberger, C. P. R., Recent advances of thiol-selective

bioconjugation reactions. Curr. Opin. Chem. Biol. 2020, 58, 28-36.

50. Goddard, D. R.; Michaelis, L., Derivatives of Keratin. J. Biol. Chem. 1935, 112.

51. Chalker, J. M.; Wood, C. S. C.; Davis, B. G., A Convenient Catalyst for Aqueous

and Protein Suzuki−Miyaura Cross-Coupling. J. Am. Chem. Soc. 2009, 131 (45),

16346-16347.

214
52. Goel, N.; Stephens, S., Certolizumab Pegol. mAbs 2010, 2 (2), 137-147.

53. Ellman, G. L., Tissue sulfhydryl groups. Arch. of Biochem. Biophys. 1959, 82 (1),

70-77.

54. (a) Bernardes, G. J. L.; Casi, G.; Trüssel, S.; Hartmann, I.; Schwager, K.;

Scheuermann, J.; Neri, D., A Traceless Vascular-Targeting Antibody–Drug

Conjugate for Cancer Therapy. Angew. Chem. Int. Ed. Engl. 2012, 51 (4), 941-

944; (b) List, T.; Casi, G.; Neri, D., A chemically defined trifunctional antibody-

cytokine-drug conjugate with potent antitumor activity. Mol. Cancer Ther. 2014,

13 (11), 2641-2652.

55. Dadová, J.; Galan, S. R. G.; Davis, B. G., Synthesis of modified proteins via

functionalization of dehydroalanine. Curr. Opin. Chem. Biol. 2018, 46, 71-81.

56. Chalker, J. M.; Gunnoo, S. B.; Boutureira, O.; Gerstberger, S. C.; Fernández-

González, M.; Bernardes, G. J. L.; Griffin, L.; Hailu, H.; Schofield, C. J.; Davis,

B. G., Methods for converting cysteine to dehydroalanine on peptides and proteins.

Chem. Sci. 2011, 2 (9), 1666.

57. Bernardes, G. J. L.; Chalker, J. M.; Errey, J. C.; Davis, B. G., Facile

Conversion of Cysteine and Alkyl Cysteines to Dehydroalanine on Protein

Surfaces: Versatile and Switchable Access to Functionalized Proteins. J. Am.

Chem. Soc. 2008, 130 (15), 5052-5053.

215
58. (a) Shiu, H. Y.; Chan, T. C.; Ho, C. M.; Liu, Y.; Wong, M. K.; Che, C. M.,

Electron‐Deficient Alkynes as Cleavable Reagents for the Modification of

Cysteine‐Containing Peptides in Aqueous Medium. Chem. Eur. J. 2009, 15 (15),

3839-3850; (b) Shiu, H.-Y.; Chong, H.-C.; Leung, Y.-C.; Wong, M.-K.; Che, C.-

M., A Highly Selective FRET-Based Fluorescent Probe for Detection of Cysteine

and Homocysteine. Chem. Eur. J. 2010, 16 (11), 3308-3313; (c) Shiu, H.-Y.;

Wong, M.-K.; Che, C.-M., "Turn-on" FRET-based luminescent iridium(III)

probes for the detection of cysteine and homocysteine. Chem Commun (Camb)

2011, 47 (15), 4367; (d) Shiu, H.-Y.; Chong, H.-C.; Leung, Y.-C.; Zou, T.; Che,

C.-M., Phosphorescent proteins for bio-imaging and site selective bio-conjugation

of peptides and proteins with luminescent cyclometalated iridium(III) complexes.

Chem. Commun. 2014, 50 (33), 4375.

59. Li, G.-L.; Kung, K. K.-Y.; Zou, L.; Chong, H.-C.; Leung, Y.-C.; Wong, K.-H.;

Wong, M.-K., Multifunctional bioconjugation by Morita-Baylis-Hillman reaction

in aqueous medium. Chem. Commun. 2012, 48 (29), 3527.

60. Zhang, C.; Welborn, M.; Zhu, T.; Yang, N. J.; Santos, M. S.; Van Voorhis, T.;

Pentelute, B. L., π-Clamp-mediated cysteine conjugation. Nat. Chem. 2015, 8 (2),

120-128.

61. Matos, M. J.; Navo, C. D.; Hakala, T.; Ferhati, X.; Guerreiro, A.; Hartmann, D.;

Bernardim, B.; Saar, K. L.; Compañón, I.; Corzana, F.; Knowles, T. P. J.; Jiménez-

Osés, G.; Bernardes, G. J. L., Quaternization of Vinyl/Alkynyl Pyridine Enables

216
Ultrafast Cysteine-Selective Protein Modification and Charge Modulation.

Angew. Chem. Int. Ed. Engl. 2019, 58 (20), 6640-6644.

62. Giles, N. M.; Watts, A. B.; Giles, G. I.; Fry, F. H.; Littlechild, J. A.; Jacob,

C., Metal and Redox Modulation of Cysteine Protein Function. Chem. Biol. 2003,

10 (8), 677-693.

63. (a) Kung, K. K.-Y.; Ko, H.-M.; Cui, J.-F.; Chong, H.-C.; Leung, Y.-C.; Wong,

M.-K., Cyclometalated gold(III) complexes for chemoselective cysteine

modification via ligand controlled C-S bond-forming reductive elimination. Chem.

Commun. 2014, 50 (80), 11899-11902; (b) Deng, J.-R.; Chung, S.-F.; Leung, A.

S.-L.; Yip, W.-M.; Yang, B.; Choi, M.-C.; Cui, J.-F.; Kung, K. K.-Y.; Zhang, Z.;

Lo, K.-W.; Leung, Y.-C.; Wong, M.-K., Chemoselective and photocleavable

cysteine modification of peptides and proteins using isoxazoliniums. Commun.

Chem. 2019, 2 (1), 1-10.

64. Vinogradova, E. V.; Zhang, C.; Spokoyny, A. M.; Pentelute, B. L.; Buchwald, S.

L., Organometallic palladium reagents for cysteine bioconjugation. Nature 2015,

526 (7575), 687-691.

65. Chan, A. O.-Y.; Tsai, J. L.-L.; Lo, V. K.-Y.; Li, G.-L.; Wong, M.-K.; Che,

C.-M., Gold-mediated selective cysteine modification of peptides using allenes.

Chem. Commun. 2013, 49 (14), 1428.

217
66. Abbas, A.; Xing, B.; Loh, T. P., Allenamides as Orthogonal Handles for Selective

Modification of Cysteine in Peptides and Proteins. Angew. Chem. Int. Ed. Engl.

2014, 53 (29), 7491-7494.

67. Lo, V. K.-Y.; Liu, Y.; Wong, M.-K.; Che, C.-M., Gold(III) Salen Complex-

Catalyzed Synthesis of Propargylamines via a Three-Component Coupling

Reaction. Org. Lett. 2006, 8 (8), 1529-1532.

68. Ravasco, J. M. J. M.; Faustino, H.; Trindade, A.; Gois, P. M. P., Bioconjugation

with Maleimides: A Useful Tool for Chemical Biology. Chem. Eur. J. 2019, 25

(1), 43-59.

69. Cal, P. M. S. D.; Bernardes, G. J. L.; Gois, P. M. P., Cysteine-Selective Reactions

for Antibody Conjugation. Angew. Chem. Int. Ed. Engl. 2014, 53 (40), 10585-

10587.

70. Dougherty, P. G.; Sahni, A.; Pei, D., Understanding Cell Penetration of Cyclic

Peptides, Chem. Rev. 2019, 119 (17), 10241-10287.

71. (a) Conte, M. L.; Staderini, S.; Marra, A.; Sanchez-Navarro, M.; Davis, B.

G.; Dondoni, A., Multi-molecule reaction of serum albumin can occur through

thiol-yne coupling. Chem. Commun. 2011, 47 (39), 11086-11088; (b) Massi, A.;

Nanni, D., Thiol–yne coupling: revisiting old concepts as a breakthrough for up-

to-date applications. Org. Biomol. Chem. 2012, 10 (19), 3791-3807.

218

You might also like