+ All Categories
Home > Documents > Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

Date post: 11-Sep-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
168
Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y. Wang Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Graduate School of Arts and Sciences COLUMBIA UNIVERSITY 2015
Transcript
Page 1: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells

Tracy Y. Wang

Submitted in partial fulfillment of the

requirements for the degree of

Doctor of Philosophy

in the Graduate School of Arts and Sciences

COLUMBIA UNIVERSITY

2015

Page 2: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

© 2015

Tracy Y. Wang

All rights reserved

Page 3: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

ABSTRACT

Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells

Tracy Y. Wang

By tagging biomolecules with bright and photostable fluorophores, chemical tags enable single

molecule (SM) detection and imaging for the study of biological mechanism. One of these tags.

the trimethoprim chemical tag (TMP-tag), labels biomolecules using the high affinity interaction

between E.coli dihydrofolate reductase and fluorescent derivatives of the antibiotic trimethoprim.

The TMP-tag is one of the few chemical tags that has enabled live cell SM imaging. In this work,

I present the development of the TMP-tag as a versatile tool for SM imaging. First, I establish

that the TMP-tag is a robust tool for labeling proteins with organic fluorophores that enable SM

detection and imaging. I then examine the properties of novel organic fluorophores that expand

the palette of dyes that can be used with SM imaging. I also investigate fluorescent

nanodiamonds for expanded imaging and sensing capabilities in live cells using the TMP-tag.

Finally, I explore applications of the TMP-tag to label and image ion channels for the

identification of pharmacological chaperones as therapeutics for protein mistrafficking diseases.

Together, these studies highlight the versatility of the TMP-tag, furthering our ability to study

biomolecules under challenging imaging and biological conditions.

Page 4: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

i

Table of Contents

List of Figures iv

List of Tables vi

Acknowledgements vii

1.1 Chapter Overview 2

1.2 Introduction 3

1.3 Single Molecule Fluorescence Microscopy 5

1.4 Single Molecule Fluorophores 11

1.4.1 Fluorescent Proteins 12

1.4.2 Organic Dyes 15

1.4.3 Fluorescent Nanomaterials 18

1.5 Chemical Tags for Fluorescent Protein Labeling 21

1.5.1 Peptide Chemical Tags 21

1.5.2 Protein Chemical Tags 23

1.5.3 Trimethoprim-Based Chemical Tags 25

1.6 Single Molecule Imaging under Live Cell Conditions with Chemical Tags 27

1.7 Outlook 28

1.8 References 29

2.1 Chapter Outlook 41

2.2 Introduction 42

2.3 Experimental Methods 45

2.3.1 Chemical Synthesis 45

2.3.2 Protein Expression, Purification, Labeling and Biotinylation 47

2.3.3 Single Molecule Methods 48

Chapter 1 Fluorescent Labeling with Chemical Tags for Single Molecule Imaging 1

Chapter 2 Covalent Trimethoprim Chemical Tags For Single Molecule Imaging 40

Page 5: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

ii

2.3.4 Ensemble Methods 49

2.4 Results 50

2.4.1 Single Molecule Total Photon Output 50

2.4.2 Quantum Yield and Ensemble Photostability Lifetime 54

2.5 Discussion 57

2.6 Conclusion and Outlook 60

2.7 NMR Spectra of Synthesized TMP-Fluorophores 61

2.8 References 65

3.1 Chapter Outlook 70

3.2 Introduction 70

3.3 Results 72

3.4 Conclusion and Outlook 82

3.5 Supporting Information 83

3.5.1 General Experimental Methods 83

3.5.2 Preparation of Hydroxyindolines and Iodoindolines 83

3.5.3 Coupling of Phenols and Aryl Iodides. 89

3.5.4 4-Nitrobenzenediazonium Tetrafluoroborate Reactions 95

3.5.5 Diazene-Diaryl Ether Cyclization to Oxazine 97

3.5.6 Tandem Friedel-Crafts Acylation/Cyclization 100

3.5.7 Dye Characterization 103

3.6 References 103

4.1 Chapter Overview 107

Chapter 3 Oxazine and Xanthene Fluorophores Synthesized from a Common Diaryl

Intermediate 69

Chapter 4 Development of Targeted Fluorescent Nanodiamonds for Cell Imaging 106

Page 6: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

iii

4.2 Introduction 107

4.2.1 Types and Synthesis of Nanodiamonds 108

4.2.2 Nanodiamond Nitrogen Vacancy (NV-) Centers 109

4.2.3 Chemical Modification of the Nanodiamond Surface 112

4.2.4 Nanodiamond Labeling with the TMP-tag 113

4.3 Experimental Methods 115

4.3.1 Size Separation 115

4.3.2 Chemical Conjugation 116

4.3.3 Nanodiamond Characterization 117

4.3.4 Cell Imaging 118

4.4 Results and Discussion 119

4.4.1 Size Separation 119

4.4.2 Nanodiamond Surface Modification 122

4.4.3 Nanodiamond Cell Labeling and Imaging 127

4.5 Conclusions and Outlook 131

4.6 References 132

5.1 Chapter Overview 138

5.2 Introduction 138

5.2.1 Cellular Trafficking of hERG Ion Channels 141

5.2.2 Visualizing hERG Ion Channel Trafficking with the TMP-tag 143

5.3 Materials and Methods 146

5.4 Results and Discussion 148

5.5 Conclusion and Outlook 155

5.6 References 155

Chapter 5 Development of Imaging Ion Channel Trafficking Assays 137

Page 7: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

iv

List of Figures

Figure 1-1: Schematics of common single molecule microscopy methods .................................... 7

Figure 1-2: Chromophores of fluorescent proteins. ...................................................................... 13

Figure 1-3: Protein based chemical tags used in live cell imaging ............................................... 24

Figure 2-1: The covalent trimethoprim chemical tag ................................................................... 43

Figure 2-2: Scheme of fluorophore modification by A-TMP-tag. ................................................ 46

Figure 2-3: Scheme of biotin-Alexa647 synthesis ........................................................................ 47

Figure 2-4 : Demonstration of covalent fluorophore labeling of eDHFR:L28C .......................... 48

Figure 2-5: Single molecule photon fluxes and survival lifetimes ............................................... 51

Figure 2-6: Total photon output .................................................................................................... 53

Figure 2-7: Quantum yield and photostability lifetime ................................................................. 55

Figure 2-8: Alexa647-A-TMP 1H NMR ....................................................................................... 61

Figure 2-9: Atto655-A-TMP 1H NMR ......................................................................................... 62

Figure 2-10: Atto680-A-TMP 1H NMR ....................................................................................... 63

Figure 2-11: Cy3-A-TMP 1H NMR .............................................................................................. 64

Figure 2-12: Biotin-Alexa647 1H NMR ....................................................................................... 65

Figure 3-1: Retrosynthetic analysis for oxazine and xanthene fluorophores ................................ 72

Figure 3-2: Reaction sequence for conversion of diaryl ethers to oxazine dyes. ......................... 76

Figure 3-3: Absorbance and fluorescence spectra of oxazine and xanthene derivatives .............. 81

Figure 3-4: Scheme of palladium catalyzed coupling between a phenol and aryl triflate. ........... 88

Figure 4-1: Diamond nitrogen vacancy (NV-) defect center ...................................................... 111

Figure 4-2: Changes to HPHT ND appearance during synthesis and oxidation......................... 113

Figure 4-3: Fluorescent ND protein labeling strategy with the TMP-tag. .................................. 115

Page 8: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

v

Figure 4-4: Transmission Electron Microscopy of Size Separated Nanodiamonds ................... 120

Figure 4-5: Schemes for nanodiamond functionalization ........................................................... 124

Figure 4-6: FT-IR of nanodiamond after surface modification treatments................................. 125

Figure 4-7: FT-IR of TMP-functionalized nanodiamonds ......................................................... 128

Figure 4-8: Extracellular protein labeling with NDs using the TMP-tag in live cells ................ 129

Figure 4-9: HEK293T cells with TMP-ND aggregates .............................................................. 130

Figure 5-1: Mechanism of pharmacological chaperone assisted protein folding. ...................... 140

Figure 5-2: Imaging hERG trafficking using the TMP-tag ........................................................ 145

Figure 5-3: Immunofluorescence labeling of hERG ion channels .............................................. 149

Figure 5-4: Live cell images of hERG fusion proteins ............................................................... 151

Figure 5-5: Live cell images of hERG-eDHFR mutants ............................................................ 153

Figure 5-6: Live cell images of hERG-GFP mutants.................................................................. 154

Page 9: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

vi

List of Tables

Table 1-1: Structures and examples of popular fluorophores for single molecule imaging ......... 15

Table 2-1: Single molecule photon flux and survival lifetime ..................................................... 51

Table 2-2: Single molecule total photon output ............................................................................ 54

Table 2-3: Ensemble photophysical properties in PBS buffer ...................................................... 56

Table 3-1: Copper(I)-catalyzed couplings between phenols and aryl iodides to furnish diaryl

ethers. ............................................................................................................................................ 75

Table 3-2: Synthesis of substituted oxazine dyes ......................................................................... 77

Table 3-3: Tandem catalytic Friedel–Crafts acylation/cylization reaction for the synthesis of

xanthene fluorophores. .................................................................................................................. 79

Table 3-4: Spectral properties of fluorescent dyes in H2O ........................................................... 82

Table 4-1: Size and distribution of NDs separated by centrifugation determined using DLS and

TEM analysis .............................................................................................................................. 121

Table 5-1 : Primers for cloning the hERG gene for Gibson Assembly ...................................... 147

Table 5-2 : Primers for mutagenesis of the hERG gene ............................................................. 147

Page 10: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

vii

Acknowledgements

I've been extraordinarily fortunate to have worked with many amazing people over the

course of my studies. Every one of them has helped move along this path and I will always be

grateful for their support.

First and foremost, I would like to acknowledge my thesis advisor, Dr. Virginia Cornish,

for her incredible support during my graduate studies. I also would like to thank my committee

members, Dr. Wei Min, Dr. Laura Kaufman, Dr. Scott Snyder, Dr. Mary Sever, and Dr. Howard

Hang. I would also like to thank my collaborators, Dr. Aaron Hoskins, Dr. Jeff Gelles, and Dr.

Dirk Englund. I would also like to express my gratitude to my previous research advisors for

preparing me with the experiences and skills to accomplish this work, Dr. Elizabeth Hillman, Dr.

Michael Hearn, Dr. Dora Carrico-Moniz, Dr. Don Elmore, Dr. Mala Radakrishnan and Tucker

Crum.

I would like to give special thanks to my fellow researchers who have served as my

mentors, especially to Dr. Zhixing Chen and Dr. Casey Brown. I would like to thank Dr.

Chaoran Jing, Dr. Yongjun Li, Dr. Rohitha SriRamaratnam, Dr. Larry Friedman, Dr. Keewook

Paeng, Dr. Stefan Jockush, and Dr. Abe Wolcott. I also owe thanks to both the past and current

members of the Cornish laboratory, Marie Harton, Mia Shandell, Caroline Patenode, Andrew

Anzalone, Gabriella Sanguinetti, Miguel Jimenez, Andy Ng, Ehud Herbst, Bertrand Adanve,

Jamie Brisbois, Dr. Nili Ostrov, Dr. Laura Wingler, Dr. Mike Englander, Dr. Matt Merguerian,

Dr. Sonja Billerbeck, Dr. Dante Romanini, Millicent Olawale, Amanda Olivo, Corey Perez, and

Heather Horgan.

To my wonderful and loving family, I would like to thank all of you for helping me arrive

at where I am today. Thank you, Mom, for pushing me to be to achieve my dreams. Thank you,

Page 11: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

viii

Dad, for inspiring me to pursue a PhD and to study what I love. Thank you, Jeffrey and Eric, for

reminding me of what the important things are in life.

Finally, I owe my deepest gratitude to my best friend and partner, Alexander G. Eng, for

riding the highs and lows with me on this fantastic journey. Together with OcieMeow and the

support committee, we keep it real and make it fun.

Page 12: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

1

Chapter 1

Fluorescent Labeling with Chemical Tags for Single Molecule Imaging

Page 13: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

2

1.1 Chapter Overview

The development of methods for single molecule (SM) microscopy and imaging has

revolutionized the study of biological systems. Using fluorescently labeled biomolecules, SM

imaging can provide unique insights into molecular mechanisms that are otherwise obscured in

ensemble measurements. Great strides have been made in overcoming the technical demands of

detecting fluorescence from individual molecules amid background and signal noise. Single

molecule methods are becoming increasingly routine for examining biological systems in vitro.

However, in vivo studies, especially intracellular studies that take place within the complex and

dynamic environment of the living cell, require more sophisticated techniques for single

molecule imaging in these demanding environments.

Historically, fluorescent proteins are used to label biomolecules, but their relatively low

total photon outputs, in comparison to organic fluorophores, impose greater challenges in

detection and imaging of single molecules. In contrast, chemical tags label proteins with high

performance fluorophores that are suitable for live cell SM imaging. With chemical tags, a

protein of interest is modified to include a polypeptide that is subsequently modified by an

organic fluorophore. The specificity afforded by genetic encoding of the polypeptide, combined

with the photophysical properties of organic fluorophores, give the chemical tags distinct

advantages over other labeling modalities for SM imaging. Thus, this chapter will detail the

development of the microscopy and labeling methods that enable single SM in live cells. We

highlight the ability of chemical tags to overcome obstacles in fluorophore detection and protein

labeling, emphasizing their utility for improved biological imaging.

Page 14: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

3

1.2 Introduction

SM studies provide valuable mechanistic and kinetic information about biological

processes. Chemical reactions arise from the stochastic behavior of single molecules, where the

state of a molecule undergoing a chemical reaction is determined by random probability

distributions for each step of that reaction. While ensemble measurements probe multiple

molecules simultaneously, the result is an average formed by contributions of each molecule's

state and can suggest that the collective of molecules assumes the same state. Consequently,

reactions in ensemble measurements smoothly progress from beginning to end state, despite the

fact that individual molecules go through discrete states. By measuring and tracking individual

molecules through a reaction with SM fluorescence imaging, the presence of transient states,

sub-populations, and other heterogeneities that have significant impacts on reaction dynamics

and mechanics can be revealed. Probing the behavior of single molecules by imaging has

emerged as a powerful tool for studying biological systems, and has been used to study the

mechanisms of myosin motor protein motility, cholesterol oxidase enzymatic turnover, and

tRNA translocation during ribosomal translation, among others.1-6

While providing tremendous biological insights, the majority of SM fluorescence

imaging investigations are performed using reconstituted systems in vitro that are not reflective

of the crowded, compartmentalized, and steady state conditions of the living cell. There are

examples of SM studies performed in live cells, but the technical challenges of fluorophore

detection and selective biomolecule labeling are magnified in the cellular environment. 7-10

Multiple metabolites exhibit autofluorescence, creating high background noise that obscures

signal from single molecules.11

In addition, the diversity of biological functional groups can

cause nonspecific fluorophore binding and labeling.12

Combined advancements in microscopy,

Page 15: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

4

fluorophore development, and protein labeling strategies can overcome these hurdles and enable

routine live cell SM imaging for studying mechanistic biology.

Advancement of microscopy methods was critical for the development of SM

fluorescence imaging in biological systems. SM microscopy was originally limited by equipment

detection sensitivity, restricting experiments to non-biologically relevant conditions in

specialized laboratories.13,14

Improvements in optics significantly eased SM detection and

increased the accessibility of SM methods to study biological processes. The application of

optical sectioning became critical for reducing background fluorescence that could easily

overwhelm signal, while additional methods were developed to ensure resolvability of individual

molecules. Dozens of microcopy techniques have been developed for SM imaging and the

success of these techniques requires fluorophores with sustained fluorescence or specialized

photophysical properties under experimental conditions.15

Organic dyes are the preferred fluorophores for SM imaging due to their higher photon

outputs relative to fluorescent proteins (FPs). Photon output is critical because fluorophores must

remain bright and fluorescent over the experimental time course for adequate detection and

resolvability.16

While FPs are widely used in biological imaging, their stability and brightness

are generally not well suited for SM imaging. Organic dyes offer more attractive photophysical

properties as well as the advantage of being smaller than FPs and less likely to interfere with

native protein function. Despite the advantages of organic fluorophores, they are still ultimately

vulnerable to photodestruction. As a result, fluorescent nanomaterials with superior brightness

and photostability are also being investigated as candidates for SM imaging. Nanodiamonds in

particular are emerging as desirable and biologically inert fluorophores with interesting optical

properties for imaging and sensing applications.17

With a diversity of high performance organic

Page 16: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

5

dyes and fluorescent nanomaterials, methods to selectively label proteins with fluorophores

becomes critical for SM imaging.

Another emergent technology for advancing SM imaging is the use of chemical tags to

label macromolecules of interest. Chemical tags can facilitate live cell SM imaging by providing

a modular technology that tags proteins with high photon output fluorophores with high

specificity. By genetically fusing a protein to a polypeptide that is modified by a cell-permeable

fluorophore, these agents allow selective labeling of proteins with molecular probes in live cells

without the use of microinjection or other invasive permeabilization methods. For SM imaging,

the specificity of the chemical tags is essential for preventing nonspecific binding that could be

misinterpreted as actual data. Chemical tags based on fluorescent ligand affinity or reactivity

with peptides and proteins have been developed and used for live cell imaging.18

Among these

tags, the trimethoprim chemical tag (TMP-tag) has the greatest engineered diversity and has been

used in a broad range of biological imaging and sensing applications.19-25

In addition, the TMP-

tag is one of the few chemical tags that have been successfully used for live cell single molecule

imaging.26-28

The robustness of the TMP-tag and other chemical tags under live cell conditions

illustrates their potential in elucidating mechanisms of complex biological systems using SM

fluorescence microscopy and imaging.

1.3 Single Molecule Fluorescence Microscopy

Single molecule (SM) fluorescence microscopy requires that the signal from an

individual molecule is both detectable and resolvable from its background. The first

measurement of SM absorption and fluorescence by Moerner in 1989 and Orrit in 1990,

respectively, illustrate the initial challenges of detecting single molecules using equipment with

limited sensitivity. In both experiments, pentacene was immobilized within a p-terphenyl host

Page 17: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

6

crystal and measurements were performed at liquid helium temperatures.13,29

Cryogenic

temperatures were critical for suppressing photodestruction and increasing absorption cross

section. While room temperature detection of single molecule fluorescence was later achieved in

1993 by Betzig and later by others, the conditions from those experiments were not amenable to

studying biological systems.30-32

In 1996, Moerner's imaging of single Nile Red molecules and Cy5 labeled proteins in

aqueous solutions opened opportunities for SM imaging in biological systems. Using a

polyacrylamide gel to slow diffusion, single molecules were imaged and their photophysical

properties were characterized.33

Moerner used the same set-up to study the photophysics of the

green fluorescent protein (GFP), further highlighting the ability to perform SM imaging of

immobilized fluorophores in aqueous solutions.34

Although alternative methods for single

molecule immobilization, such as biotin-streptavidin, are currently more preferred, Moerner's

work serves the basis for in vitro study of single molecules in biological systems.35

Combined

with the development of high sensitivity photodetectors and optics have made SM detection and

imaging in aqueous solutions a relatively routine feat for research laboratories.

The success of Moerner's SM imaging in aqueous solution and other SM studies relies on

the use of optical sectioning to limit background fluorescence. By limiting illumination to

specific planes within the sample, out-of-plane fluorescence is significantly reduced, enabling

clear resolution of signals from single molecules. One of the more popular techniques is total

internal reflection (TIRF) microscopy, where fluorescence illumination is limited to an

evanescent wavefront that is produced as a result of total internal reflection (Figure 1-1a).36,37

With TIRF, only fluorophores within a few hundred nanometers of the slide surface exhibit

fluorescence, making it a popular tool for in vitro studies where molecules can be restricted and

Page 18: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

7

immobilized to the slide surface.35

Most notably, TIRF was used image ATP turnover of myosin

in one of the first SM imaging experiments that probed mechanistic biology.4 Because TIRF

limited fluorescence to the surface, co-localized fluorescence from a Cy3-ATP conjugate with an

immobilized Cy5-labeled myosin directly corresponded to the binding and hydrolysis of ATP.

While TIRF can also be used in live cells, it is limited to the study of membrane proteins and has

limited applications for imaging intracellular proteins.38

Figure 1-1: Schematics of common single molecule microscopy methods

a) Total internal reflection (TIRF) microscopy produces an evanescent wavefront from total internal reflection,

illuminating fluorophores close to the surface. b) Highly inclined and laminated sheet (HiLo) microscopy,

illuminates fluorophores a few microns from the surface using refraction. c) Fluorescence speckle microscopy

(FSM) uses dilute labels to resolve individual fluorophores. d) Photoactivable localization microscopy (PALM) and

stochastic optical reconstruction microscopy (STORM) use stochastic fluorophore activate to image a subset of

individually resolved fluorophores in each image. Figure adapted from Coehlo et al, 2013.39

Alternative microscopy techniques can achieve optical sectioning for in vivo imaging

deeper inside cells. Light sheet microscopy and its variations, which use a thin and focused light

Page 19: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

8

sheet for fluorescence illumination, are among the most widely used. Highly inclined and

laminated optical sheet (HiLo) microscopy generates an inclined sheet of light by refraction off

the slide surface (Figure 1-1b). HiLo is capable illumination microns from the surface and has

been used to image nuclear pore complexes.40

For thicker samples, both selective plan

illumination microscopy (SPIM) and light sheet fluorescence microscopy (LSFM) illuminate

samples with a sheet of light and collect fluorescence signal orthogonally to the illumination to

image at depths of 50-200 microns.41,42

Although additional techniques using confocal

microscopy have also been developed for SM imaging, TIRF and HiLo remain the most

commonly used for live cell imaging due to their relatively straightforward optical set-ups. 43

Overall, SM microscopy with optical sectioning improves detection against background, but

successful SM imaging also requires that detected fluorophores are spatially resolved from one

another.

The major barrier to individual fluorophore resolvability is the diffraction limit of light.

The maximum resolution of a point in optical microscopy is approximately 200 nm as

determined by Abbe's diffraction limit.44

As a result, fluorophores that are located within 200 nm

of one another appear as a single spot and another cannot be visually resolved from one another.

In vitro studies can manipulate the spacing of fluorophores through the use of low concentrations

and immobilization such that a single spot corresponds to one molecule.35

In addition, a number

of in vitro and extracellular methods that further localize the spatial position of fluorophores

beyond the diffraction limit for super resolution imaging have also been developed.45-53

However, a limited number of techniques have risen to meet the challenge of resolving

individual molecules among many in live cells.

Page 20: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

9

In one of the first demonstrations of live cell intracellular SM imaging, fluorescence

speckle microscopy (FSM) employed fluorescently labeled actin at very dilute concentrations to

resolve single actin proteins within filaments in lamellipodia (Figure 1-1c). Using concentrations

of 1 labeled molecule to 10,000 unlabeled molecules, single actin was tracked through filament

formation, demonstrating basal polymerization and depolymeraization throughout the filament as

well as polyermization concentrated at filament tips.54

However, one of the major drawbacks of

FSM is the use of microinjection or other cell permeabilization strategies to ensure sufficiently

low concentrations of labeled proteins for SM resolution.55

As a result, other SM strategies for

live cell imaging have been developed that are less invasive and more suitable to studying

heterogeneous protein interactions.

The groundbreaking technique to bring SM imaging to live cells was the development of

photoactivatable localization microscopy (PALM) and stochastic optical reconstruction

microscopy (STORM).56-58

In both techniques, the fluorescence of only a subset of fluorophores

is activated such that single fluorophores can be resolved (Figure 1-1d). The spatial location of

each fluorophore is localized based on its point spread function, providing resolution beyond the

diffraction limit called super resolution imaging. Multiple images with different subsets of

fluorophores are taken and the final image is reconstructed from these images. While

conceptually similar, PALM was developed using photoactivatable FPs while STORM used

photoswitchable organic fluorophores. Both methods have been used in live cell SM imaging,

with PALM imaging of single EosFP labeled adhesion complexes with 60 nm resolution and

STORM imaging of single Cy3-Alexa647 labeled clathrin light chain proteins within clathrin

coated pits with 30 nm resolution. 27,59

Page 21: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

10

The widespread use of PALM is currently limited by the availability of suitable

fluorescent proteins. Because PALM relies on the stochastic activation of fluorescence, only

photoactivatable or photoswitchable FPs can be used. While reversible photoswitchable proteins

could be used for PALM, the majority of those proteins use the same the same or similar

wavelength for fluorescence emission and switching and cannot be easily used with PALM.60

Irreversible photoactivatable or photoswitchable FPs allow controlled activation with orthogonal

wavelengths, but requires that the FPs be photobleached before subsequent acquisitions and

preclude multiple measurements of the same molecule. As a result, initial PALM studies were

limited to FPs that were sufficiently bright and photostable for detection and localization, but

could not be so photostable as to delay subsequent acquisitions.59

The recent development of a

reversibly photoswitchable GFP, Dreiklang, with decoupled fluorescence and switching

facilitates the use of PALM in live cells.43

Further developments in FP technology can increase

the applications of PALM.

On the other hand, STORM uses photoswitching behavior of organic fluorophores for

SM imaging with better resolution and speed compared to PALM. Originally, STORM used

fluorophore activator-reporter pairs of Cy3- Cy5 that required cycling between different

wavelengths for photoswitching, achieving resolutions of 20 nm (Rust).57

STORM was also

developed for three dimensional (3D) imaging, using optical astigmatism to determine axial and

lateral fluorophore positions.61

Both STORM and 3D STORM achieved resolutions of 25 nm and

30 nm and frame rates of 0.5 second and 1 second, respectively, in live cells.27

In addition, a

variation of STORM using individual fluorophores rather than fluorophore pairs was also

developed (dSTORM) and used in live cell imaging of H2B nuclear histone proteins.26,58,62

Overall, the use of organic dyes, which are brighter than FPs, permits faster fluorophore imaging

Page 22: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

11

and higher resolution images, demonstrating the advantages of using high performance

fluorophores with specialized properties.63

Developments in microscopy have greatly improved the ability to perform SM imaging,

but all these methods require bright and photostable fluorophores for SM detection and

resolution. For live cell imaging, the intracellular environment and protein mobility necessitates

the use of fluorophores that are at least as bright and photostable as, if not more than,

fluorophores used for in vitro studies. The diffusion of proteins creates additional challenges for

SM imaging, by reducing the amount of time a protein spends within a diffraction-limited area

and consequently lowering detectable signal for imaging. Currently, freely diffusing cytosolic

proteins are nearly impossible to image at the SM level, limiting live cell SM imaging to proteins

with restricted mobility. To expand the ability to image more mobile proteins within the live

cell, fluorophores must have photon outputs that enable sufficient detection over cellular

background on the shorter timescales during which proteins remain within a diffraction limited

area. While there are many different types of fluorophores that can be used for biological

imaging, few have the requisite properties that make them compatible with both live cells and

experimental conditions for SM imaging.

1.4 Single Molecule Fluorophores

One of the major barriers to widespread use of SM imaging in live cells is the lack of

suitable fluorophores. The development and discovery of fluorophores with superior

photophysical properties is key for expanding SM imaging in live cells. The critical requirement

for fluorophores is high total photon output, which is a product of photon flux and the

photostability. Photon flux, the rate of photon emission, is a measure of brightness, while the

photostability is a measure of the time a fluorophore can remain fluorescent before

Page 23: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

12

photobleaching. So, fluorophores for biological imaging should be both bright and long-lasting

under aqueous conditions. For SM imaging in live cells, the requirement for high photon output

is heightened by the need to overcome background cellular autofluorescence while maintaining

photostability over the timescale of the experiment. However, even fluorophores with sufficient

photon outputs must meet other conditions for successful imaging.

Experimental demands place additional restrictions on fluorophore choice based on

photophysical behaviors, spectral properties, and cellular compatibility. For microscopy

requiring photoswitching, fluorophores must exhibit this behavior under live cell conditions. In

addition, fluorophore spectral properties must be compatible with a given microscopy set-up and

resolvable from other fluorophores if multiple colors are used. Red fluorescence is generally

preferred for cell imaging because of reduced background fluorescence and cellular

photodamage under red illumination.64

The fluorophore itself should also be non-cytotoxic and

be sufficiently small as to not interfere with mechanisms it is probing. Despite these many

considerations, the total photon output remains the most important criteria for SM imaging. To

date, a diverse set of different fluorophores have been applied in both in vitro and in vivo live

cell SM imaging, and the development of new fluorophores holds additional promise for

expanding this palette.

1.4.1 Fluorescent Proteins

The discovery and use of FPs was an incredible breakthrough for fluorescent imaging in

biological systems. First isolated from Aequorea victoria, the green fluorescent protein (GFP) is

a 238 amino acid protein (27kDa) that houses a 4-p-hydroxybene-5-imidazolinone chromophore,

formed from the oxidation of serine, tyrosine, and glycine residues, in its beta barrel core (Figure

1-2).65

After its discovery, GFP was quickly adopted as a genetically encoded label for in vivo

Page 24: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

13

studies of protein expression and localization in E. coli and C. elegans.66

Later engineering of the

FPs by Roger Tsien and others have produced a toolbox of FPs with different spectral properties

for a wide range of cell imaging applications.67

One such protein, enhanced yellow fluorescent

protein (EYFP) was one of the first fluorophores used for live cell SM molecule in bacteria cells.

As a reporter for gene expression, the detection of single EYFPs synthesized by E. coli revealed

the stochastic nature of protein translation.8 Favored FPs for SM imaging include EYFP for its

fast chromophore maturation time, enhanced green fluorescent protein (EGFP) for its photon

output, and TagRFP for its red fluorescence.8,68

The ability to genetically encode FPs gives them

near perfect labeling specificity, making them a popular choice for live cell SM imaging.

Figure 1-2: Chromophores of fluorescent proteins.

The fluorophore in green fluorescent protein (GFP) is formed from oxidation of tyrosine, serine, and glycine.

Threonine is sometimes used in place of serine. Different FP variants have also been engineered, each with different

chromophores dependent on the amino acid residues and their positioning within the beta barrel core. Figure adapted

from Wang et al, 2014.18

The development of photoactivatable, photoconvertible, and photoswitching FPs has

been particularly useful for SM super resolution imaging with PALM and other related

Page 25: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

14

methods.69,70

When exposed to certain wavelengths, these FPs can be classified by whether they

switch from dark to bright states (photoactivatable) or if they switch from one color to another

(photoconvertible) and also if they can switch between states reversibly (photoswitchable). The

first live cell PALM used photoactivatable GFP to image the distribution of hemagglutinin on the

cell membrane, revealing the formation of irregular clusters.71

Photoconvertible FPs have also be

used in live cell PALM, notably Dendra2 without a fusion protein and EosFP tagged adhesion

complexes.59,72

Two color live cell PALM was later carried out using photoactivatable GFP and

mCherry to image transferrin and clathrin light chains within clathrin coated pits.73

Although it

has not yet been used in live cells, the development of Dreiklang, a photoswitchable GFP protein

that has little crosstalk between the fluorescence and switching wavelengths, holds promise for

greater applications of photoswitchable FPs in PALM.74

The engineering of FPs with

increasingly appealing photophysical properties has opened new opportunities for live cell SM

imaging.

Despite the popularity of FPs for SM imaging, they are limited by their generally lower

total photon output. FPs tend to photobleach within seconds under SM imaging conditions, with

the best FP, mCherry, exhibiting fluorescence up to 20 seconds.75

While the FPs could be used

for imaging biological processes on shorter timescales, extending their application to a even a

full minute is extremely difficult. In addition, fluorescent proteins can suffer from poor folding,

slow and oxygen dependent chromophore maturation, and oligomerization, all of which can

cause undesirable experimental artifacts when tagging less robust or smaller proteins.67

In

particular, many investigations of ion channel function and trafficking are not amenable to FP

fusions.76

Most concerning, however, is that transgenic mice has demonstrated cardiovascular

and muscular defects when GFP is tagged to myosin, even though deleterious effects appear

Page 26: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

15

minimal at the cellular level.77,78

This raises the question as to whether the size of the FP fusions

allow them to accurately reflect native protein functions. Finally, despite the identification of

red-fluorescent proteins, there are few far red and near infrared fluorescent proteins ( > 600 nm

excitation and emission) that have the requisite brightness and photostability that is preferred for

cell imaging.79

To overcome these limitations with fluorescent proteins, organic fluorophores

with greater stability and more general applicability have been sought for application in live cell

imaging.

1.4.2 Organic Dyes

The small size and photophysical properties of organic fluorophores are advantageous for

SM imaging. Organic fluorophores have been utilized as live cell probes since 1963 when the

catalysis and fluorescence of fluorescein-galactose was first used to report -D-galactosidase

activity.80

Multiple families of organic fluorophores have been discovered and developed over

the years, each having different chemical and photophysical properties for biological imaging

applications.81

For single SM imaging, however, the xanthenes, cyanines, and oxazines are

among the most frequently used fluorophores (Table 1-1).

Table 1-1: Structures and examples of popular fluorophores for single molecule imaging

Core Structure Key Features Examples

O

Inexpensive

Popular for general bioimaging

TAMRA

Alexa488

Alexa594 Xanthene

+R2N NR2n

Very bright

Photoswitchable

Cy3

Cy5

Alexa647 Cyanine

N

O

Photostable

Photoswitchable

Red fluorescence

Atto655

Oxazine

Page 27: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

16

Xanthenes-based fluorophores, including fluoresceins and rhodamines, are characterized

by a core structure containing a heterocyclic ring system (Table 1-1). The early development of

fluorogenic fluorescein diacetates that were activated upon intracellular esterase activity

established these fluorophores as suitable for live cell imaging.82

However, fluorescein is a

notoriously poor SM fluorophore, despite its popularity in conventional fluorescence

microscopy. Rhodamines, on the other hand, have better photon output and have been used in a

number of SM imaging applications. The commercial fluorophore, Alexa488, is a common

substitute for fluorescein while Alexa594 is particularly popular for its red fluorescence.

Photocaged and photoswitchable rhodamines also have promising applications for live cell SM

imaging.79

Overall, the total photon output of xanthenes is not as high as other organic

fluorophores and there are few red and far-red xanthene fluorophores. Consequently, cyanine and

oxazine fluorophores are preferred for SM imaging.

Cyanines are characterized by a polymethine chain connecting two nitrogen groups

(Table 1-1). The fluorescence of cyanines is tunable based on the length of the methine chain,

with longer chains producing more red shifted fluorescence. Originally developed as membrane

voltage sensors for live cells, water soluble cyanine fluorophores exhibit high extinction

coefficients and are used in many SM and super resolution imaging experiments.83,84

The

commercial fluorophores, Cy3, Cy5, and Alexa647 were used in the first demonstrations of

STORM as well as in some of the first in vitro mechanistic SM studies of ATP turnover and

myosin motility.1,4,57,61

While their brightness is valuable for SM imaging, cyanines are

vulnerable to photobleaching. Photostability of cyanines can be enhanced by removing of

dissolved oxygen, adding proximal triplet state quenchers and rational chemical design reducing

Page 28: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

17

photoreactivity.85-87

The applications of cyanines in seminal SM studies shows the importance of

their brightness for detection and imaging.1,4,57,61

However, the need to use quenchers and

oxygen scavenger additives to prolong their photostability limits their application in live cells.

Oxazine fluorophores are characterized by a core structure containing a nitrogen and

oxygen heterocyclic ring system (Table 1-1). Similar to the cyanines, oxazines were initially

developed for imaging cell membranes with improved photostability.88,89

Oxazines are generally

bright and have exceptional photostability under aqueous conditions. In fact, Moerner's first

room temperature detection of single molecule fluorescence in solution used Nile Red, an

oxazine derivative.33

More recently, the commercially available oxazine, Atto655, has been

extensively studied for its photoswitching behavior, which is attractive for SM super resolution

imaging using STORM and dSTORM. Unlike cyanines and other fluorophores which require

photoactivation, deoxygenation, or conjugation to activator fluorophores for photoswitching,

Atto655 photoswitching is reversible, inherent to the molecule and can be catalyzed by thiol-

containing reducing agents found in cells.26,90-92

Finally, Atto655 and other oxazines have far red

fluorescence, which is favorable for cells. Their spectral properties, photoswitching, and

photostability give the oxazines tremendous potential as fluorophores that can enable routine live

cell SM imaging.

Fluorophores outside these three families have also been investigated as potentially

superior dyes for single molecule imaging. Dicyanodihydrofurans (DCDHF) have favorable

photophysical properties and also show sensitivity to local viscosity.93

In addition, they have

been employed as photoactivatable fluorophores for live cell SM imaging.94

With photostability

at least an order of magnitude greater than fluorescein, acceptable water solubility for SM

experiments, and availability of red-fluorescent derivatives, DCDHFs are appealing fluorophores

Page 29: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

18

for SM live cell imaging. Another class of fluorophores being explored for their exceptional

photostability are perylene dyes. While initial perylenes were completely water insoluble, their

attractive photophysical properties prompted the development of water soluble derivatives that

retained their excellent performance.95,96

Water soluble perylenes have been used to track single

enzymes in vitro and photoswitching behavior can be induced through the introduction of redox

reagents through similar mechanism as for Atto655.97,98

While both DCDHFs and perylenes have

not yet gained widespread popularity, their superior photon outputs indicate their potential for

surpassing cyanines and oxazines as fluorophores for SM imaging.

The photon output of organic fluorophores can overcome challenges in live cell SM

imaging. However, organic dyes are still limited by their eventual vulnerability to

photobleaching. The rational synthesis of dyes with improved photostability and the

development of new classes of fluorophores with superior properties will certainly help improve

the performance of organic dyes in SM imaging. At the same time, it is hypothesized that the

lack of significant strides in photostable organic fluorophores suggests that we are approaching

limits for small molecule photon output. As a result, there is an increased interest in using

fluorescent nanomaterials to meet the demands for SM fluorophores. Because fluorescent

nanodiamonds are incredibly resistant to photodestruction in comparison to organic

fluorophores, they can have great applications for SM imaging

1.4.3 Fluorescent Nanomaterials

Fluorescent nanomaterials, including noble metal nanoparticles, semiconductor quantum

dots (QDs), carbon nanotubes, and nanodiamonds (NDs), are promising new fluorescent

materials for SM imaging because of their high photon outputs and resistance to

photobleaching.99

Many of these materials exhibit fluorescence on timescales that are orders of

Page 30: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

19

magnitudes longer than those of organic fluorophores and fluorescent proteins. The adoption of

fluorescent nanomaterials in biological imaging has greater technical hurdles due to their

comparatively large sizes and cytotoxic properties, but significant strides have been made to

improve nanomaterial biocompatibility. QDs are among the most frequently used nanomaterial

for biological imaging due to their tunable and bright fluorescence. However, fluorescent

nanodiamonds demonstrate great promise with their non-toxicity and interesting optical sensing

capabilities.

QDs were among one of the first fluorescent nanomaterials used for biological imaging.

These cadmium selenium nanocrystals exhibit size dependent fluorescence arising from

electronic energy band gaps.100

While the QD itself is only a few nm in size, the coatings

required to retain their fluorescence and render them hydrophilic can increase their size to 20-30

nm.101

In addition, cadmium release from QDs has also raised concerns about cytotoxicity.102

Despite these drawbacks, QDs are appealing for live cell fluorescent probes because their

brightness and narrow excitation and emission bands permits multicolor imaging on a timescale

of weeks.103

QDs enabled imaging of diffusion and localization of single glycine receptors for 20

minutes.104

Surprisingly, the primary concern for this SM experiment was not the ability to detect

the fluorophore, but rather the toxicity of continuous light illumination for the cells. QDs

continue to have SM applications in live cells, but their use is typically restricted to extracellular

applications due to their size and membrane impermeability.105

The remarkable ease of SM

detection and imaging with QDs in comparison to organic fluorophores demonstrates how

superior fluorescent nanomaterials can truly facilitate SM imaging for research applications.

More recently, fluorescent NDs (FNDs) have drawn significant interest as a material for

biological imaging. NDs are nanometer sized diamond crystals that are biologically inert and can

Page 31: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

20

produce fluorescence from nitrogen vacancy center defects in the crystal lattice.106

Contained

within the lattice and shielded from the environment, nitrogen vacancy centers are remarkably

photostable and can be easily used for optical SM detection.107,108

Interest in FNDs for live cell

SM imaging originated from a demonstration of single FND imaging and tracking in HeLa cells

in 2007.109

In this study, 35 nm FNDs were taken up by cells with no observed toxic effects and

exhibited continuous fluorescence for at least 5 minutes, in comparison to the 10 seconds for

Alexa546 to photobleach. Interestingly, later reports of smaller 5 nm FNDs exhibited blinking

properties that could be useful for super resolution SM imaging.110

The FNDs also have special

optical properties that were demonstrated for differentiation between individual NDs and

magnetic field sensing in cells.111

While nanodiamonds have not yet been used to study

biological processes, their optical and biological properties show extraordinary promise for

expanding the ability to probe cellular behavior at the SM level.

Both organic fluorophores and FNDs have high photon outputs that are advantageous for

single molecule imaging. However, these fluorophores inherently lack the selectivity of

genetically encoded FPs for protein labeling. In order to tag biomolecules with organic

fluorophores or FNDs for SM imaging, creative strategies in chemical labeling must be

employed. Traditional biomolecular crosslinking strategies can label purified proteins, but

require the microinjection of these proteins for live cell studies, which is damaging to

cells.55,112,113

Bioorthogonal labeling chemistry, such as including azide-alkyne reactions and

others, can selectively label proteins in cells.12,114

However, the incorporation of appropriate

reactive handles for these bioorthogonal reactions require sophisticated techniques, such as

unnatural amino acid incorporation. Chemical tags bridge the gap between selective labeling and

Page 32: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

21

the ability to use superior fluorophores in live cells using genetic encoding of polypeptide

fusions and cell-permeable fluorophore ligands.

1.5 Chemical Tags for Fluorescent Protein Labeling

Chemical tags can selectively label biomolecules with a wide array of fluorophores with

SM imaging capabilities. The selectivity of fluorophore labeling is essential for SM experiments

because contributions from nonspecifically bound fluorophores are measured and not averaged

out. Particularly in experiments with low protein concentrations that are within the partition

coefficient for nonspecific labeling, there can be significant contributions for nonspecific

fluorophores. For chemical tags, specificity of labeling is afforded by genetic encoding of the

polypeptide and high affinity, selective ligand binding. Consequently, chemical tags can offer

selectivity similar to genetically encoded FPs and at similar, if not smaller, sizes. However, the

identification and development of both a protein or peptide and ligand that are orthogonal to

other biomolecular interactions in mammalian cells is challenging. Many research groups have

identified and developed chemical tags for protein labeling, but few have the robustness for

intracellular labeling in live cells. For those tags that can overcome these hurdles, they confer a

major benefit over the FPs in their ability to label proteins with organic fluorophores that have

significantly better photophysical properties.

1.5.1 Peptide Chemical Tags

The first report of a chemical tag designed for live cell imaging was the FlAsH tag,

developed by Roger Tsien in 1998. With FlAsH, protein is tagged with a six amino acid

sequence containing a tetracysteine motif (CCXXCC). A cell permeable biarsenical fluorescein

reagent is introduced and binds to the tetracysteine, rendering the label fluorescent.115

Further

development of this technology has provided a number of biarsenical fluorophores of different

Page 33: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

22

colors, with the original green-fluorescent FlAsH and red-fluorescent ReAsH being the most

frequently used.116-118

Most notably, the small size of the FlAsH and ReAsH tags were used to

label and image the transportation of small gap junction proteins that would have otherwise had

impacted function with FPs.119

Despite its small size and fluorgenicity, the FlAsH tag suffers

from nonspecific binding to cysteine-rich proteins due to the affinity between arsenic and

thiols.79

In addition, toxicity from the biarsenical reagents is also a concern. Other peptide-heavy

metal affinity tags suffer from similar problems and have comparatively worse selectivity and

binding affinities. Of note are both the oligo-asp tag (D4 tag) and his-tag, which bind to

fluorescent zinc complexes, and are limited to extracellular labeling. 120,121

A variation of peptide affinity based chemical tags are enzymatic peptide tags. In this

approach, a protein is tagged with a peptide sequence that is recognized and modified by an

enzyme to incorporate a fluorophore or a chemical handle that is subsequently modified by a

fluorophore. The Ting lab engineered a coumarin ligase based on an E. coli lipoic acid ligase,

which labels the lysine side chain of a 13 amino acid recognition sequence with coumarin for

highly specific live cell protein labeling.122

The one-step incorporation of a fluorophore offers

significant advantages over previous two-step technology in which an engineered ligase

incorporates a bioorthogonal handle that is subsequently modified.79

Unfortunately, engineered

ligases that can accommodate additional fluorophores have yet to be reported and the UV-to-blue

fluorescence of coumarins are not ideal for live cell imaging. Additional enzymatic peptide tags

have developed based on the activity of biotin ligase, human transglutaminase (Q-tag) and

bacterial sortase (SorTag), but their applications so far have been limited to extracellular

labeling.123-127

Enzymatic peptide tags show great promise because of their small size, but cross-

reactivity, condition dependent activity of the enzymes, and compatibility with multiple

Page 34: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

23

fluorophores can be problematic. The limitations of both affinity based and enzymatic peptide

tags have made protein based chemical tags the preferred option for live cell single molecule

imaging.

1.5.2 Protein Chemical Tags

Affinity-based protein chemical tags rely on the specific interaction between a protein

domain and a ligand. Developed by Kai Johnsson in 2003, the first protein chemical tag reported

for live cell imaging was the SNAP-tag. The target protein is tagged with an engineered mutant

of the human DNA repair protein O6-alkylguanine-DNA alkyltransferase (20kDa), which

irreversibly reacts to and binds with a fluorescent O6-benzylguanine substrate (Figure 1-3a). The

SNAP tag exhibits high selectivity and fast reactivity, while the formation of a covalent bond

between the tag and the fluorophore ensures stable labeling.128

A variation of the SNAP-tag,

called CLIP-tag, has also been developed that selectively uses orthogonal O2-benzylcytosine

conjugates (Figure 1-3b).129

The SNAP-tag has been used in a wide variety of both imaging and

sensing applications.130

In particular, the use of organic fluorophores with SNAP-tag has been

exploited for deep tissue imaging of protein half-life in mice and reduced background of

intracellular FRET between receptor proteins.131-133

While the SNAP-tag is one of the most

commonly used chemical tags, other tags have been developed and have also enjoyed success in

biological and live cell imaging.

Page 35: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

24

Figure 1-3: Protein based chemical tags used in live cell imaging

a) The SNAP-tag uses fluorescent O6-benzylguanine suicide substrates to label proteins tagged with a modified O

6-

alkylguanine-DNA alkyltransferase. b) The CLIP-tag uses fluorescent O6-benzylcytosine suicide substrates to label

proteins tagged with an orthogonally engineered SNAP-tag. c) The HaloTag uses fluorescent chloroalkane suicide

substrates to label proteins tagged with haloalkane dehalogenase. d) The TMP-tag uses fluorescent trimethoprim

conjugates to label proteins tagged with E. coli dihydrofolate reductase. Figure adapted from Wang et al, 2014 .18

The HaloTag is another popular chemical tag that has demonstrated intracellular labeling

capabilities.134

Using the same suicide enzyme substrate labeling mechanism as the SNAP-tag,

the HaloTag tags a protein with an engineered bacterial haloalkane dehalogenase (34kDa) that

forms a covalent attachment to fluorescent chloroalkane conjugates (Figure 1-3c). The HaloTag

enjoys fast reactivity and high specificity and has been used in a number of live cell imaging

applications.18,135,136

One interesting application of the HaloTag in protein expression and pull

down assays. The high selectivity and water solubility of the tag increases purity and reduces

Page 36: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

25

aggregates of insoluble proteins.137

The solubility of the HaloTag may be useful to tagging more

challenging proteins and its reactivity for fast fluorophore labeling is advantageous for live cell

imaging.

Overall, the SNAP-tag, CLIP-tag, and HaloTag are among the tags that have the greatest

use in biology research because their fluorescent substrates are commercially available.

However, additional protein-based chemical tags have been developed by other research groups.

The Kikuchi group has engineered two tags, the BL-tag and the PYP-tag for live cell imaging.

The BL-tag uses an engineered bacterial beta lactamase (29kDa) that binds to fluorescent beta-

lactam prodrug ligands.138

The PYP-tag(Hori) is based on the direct binding of a small water

soluble protein from bacteria (14 kDa) to 7-hydroxycoumarin-3-carboxylic acid derivatives and

is among one of the smallest protein based chemical tags.139

Finally, the cutinase tag (22 kDa)

which is based on fungal cutinase and has been used for extracellular protein labeling, may have

future applications for intracellular labeling due to the lack of native cutinases in mammalian

cells.140

While these tags meet the criteria of selectivity and affinity for live cell labeling with

fluorescent probes, most of these tags use one mechanism for ligand binding and reactivity to

label proteins. Creative engineering of different ligands can achieve fluorogenic substrates for

some tags, but none of these tags have been engineered with multiple labeling mechanisms that

can be exploited for a broad variety of applications.138,139,141

1.5.3 Trimethoprim-Based Chemical Tags

Of the chemical tags used for live cell imaging, the trimethoprim based chemical tags

(TMP-tags) offer the greatest versatility with multiple engineered variants. Developed by

Virginia "The Boss Lady" Cornish in 2005, target proteins are tagged with E. coli dihydrofolate

reductase (eDHFR), which subsequently binds to fluorescent trimethoprim (TMP) ligands

Page 37: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

26

(Figure 1-3d).19

The high affinity between eDHFR and TMP (KD = ca. 1nM) permits highly

specific labeling using low concentrations of fluorescent ligand, reducing background.

Developed as an antibiotic, the TMP ligand shows incredible orthogonality to mammalian

dihydrofolate reductases and is highly cell permeable. Early demonstrations of the TMP-tag

technology illustrated the compatibility of the tag for live cell imaging of multiple cell lines of

nuclear, cytosolic, and plasma membrane proteins. The high affinity of the TMP-tag allowed

noncovalent labeling that was at similar levels to the covalent labeling with the SNAP-tag.

A covalent TMP-tag was then engineered to further expand its versatility and labeling

capabilities. The first covalent TMP-tag used an engineered eDHFR mutant that contained a

cysteine residue proximal to the TMP binding pocket (eDHFR:L28C) and a trifunctional TMP-

fluorophore-acrylamide ligand. Upon binding of the TMP ligand to the eDHFR, the cysteine

undergoes a nucleophilic attack on the acrylamide group, producing a covalent bond.142

Further

optimization of the positioning between the TMP and acrylamide moiety improved the reactivity

of labeling to an 8 minute reaction half life. As a result, the covalent TMP-tag succeeded in

imaging both abundant and diffuse cytosolic proteins.20

The improvement of rendering labeling

covalent improved the versatility of the TMP-tag to study the diverse array of proteins within

living cells.

Finally, a fluorogenic TMP-tag has been developed to further reduce background signal

from unlabeled fluorophores. A trifunctional TMP-fluorophore-quencher was developed in

which the quencher molecule was linked to the TMP-fluorophore with a tosylate group. Binding

of the TMP-fluorophore-quencher to eDHFR:L28C causes the cysteine nucleophile to attack and

displace the tosylate linked quencher, activating fluorescence. The fluorogenic TMP-tag exhibits

a 20-fold increase in fluorescence upon binding, greatly reducing background signal for

Page 38: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

27

nonspecifically bound fluorophores that was previously observed with the covalent TMP-tag.21

Another fluorogenic variant of the TMP-tag has also been developed using increased

fluorescence lifetime of a fluorophore that is rotationally constricted due to binding with a

protein. However, this variation only produces a 2-fold increase in fluorescence.143

The

technological achievements of engineering the TMP-tag clearly illustrate its many potential

applications for not only live cell imaging, but also a range of biological studies.

The TMP-tag has been used to label proteins with both fluorophores and other molecules

to probe different types of biological activities. By labeling differentially localized proteins with

Cy3, the TMP-tag was used to measure local environmental viscosity within live cells based on

Cy3 fluorescence lifetime readout. Fluorescent lanthanides labels for live cell time resolved

FRET have installed on extracellular, unpurifiable proteins.24

Focal adhesion complexes have

been labeled with magnetic iron oxide nanoparticles and magnetically manipulated in live cells.25

The conjugation of a photosensitizer to myosin with the TMP-tag was also used for chromophore

assisted laser inactivation to study myosin in cytoskeletal coherence.22

Altogether, the TMP-tag

and other chemical tags provide important tools for both live cell imaging and other research

applications.

1.6 Single Molecule Imaging under Live Cell Conditions with Chemical Tags

The ability of chemical tags to conjugate fluorophores with exceptional photophysical

properties to proteins in live cells make the tags well suited for SM imaging. Single clathrin light

chain proteins within clathrin coated pits have been imaged with the SNAP-tag and HaloTag,

while single nuclear histone H2B proteins have been imaged with the SNAP-tag and the TMP-

tag.26-28,58

In addition, the SNAP-tag and HaloTag have been used in other live cell SM imaging

demonstrations that have investigated the performance of different organic fluorophores.94,144

Page 39: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

28

Recently, the SNAP-tag was also used to label and image single RNA polymerases in the cell

nuclei.145

The use of STORM, dSTORM and other SM microscopy techniques in these proof-of-

principle experiments illustrate the advantages of using organic fluorophores with chemical tags

for single and multi-color SM imaging in live cells. Additionally, the modularity of the chemical

tags has been exploited for labeling with fluorescent nanomaterials. The SNAP-tag and Halo-Tag

were used to attach QDs to extracellular Notch receptor and neurexin synaptic adhesion protein,

respectively, for demonstrations of comparatively straightforward live cell SM imaging.146,147

While applications of chemical tags to address SM mechanistic questions in live cells has been

limited, the success of the tags in these investigations firmly establishes their potential for

widespread SM imaging applications.

In particular, the TMP-tag and SNAP-tag are very promising for probing SM intracellular

biology as demonstrated by their use in deciphering the order of binding events in spliceosome

assembly.148

The spliceosome is a challenging system to reconstitute as it is composed of

approximately 100 core proteins.149

Therefore, in vitro examination of the spliceosome takes

place in whole cell extracts where the photon output of FPs is insufficient to overcome

autofluorescence.150

The TMP-tag and SNAP-tag were used to label spliceosome complex

proteins with organic fluorophores, enabling the multi-color imaging of complex protein binding

events that form functional spliceosomes. The live cell conditions of this seminal study confirms

that the chemical tags and their ability to use organic fluorophores lends superior capabilities for

probing complex macromolecular machinery of the cell.

1.7 Outlook

The advancement of SM imaging in live cells requires the ability to selectively label

biomolecules with fluorophores that can withstand demanding imaging conditions. SM imaging

Page 40: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

29

can reveal incredibly detailed mechanistic information that revolutionizes our understanding of

biology. The studies conducted so far have confirmed that molecular heterogeneity is

fundamental to biological mechanism and drive our curiosity to perform SM studies in additional

systems. Live cell SM imaging would allow us in depth study of complex cellular processes in

their native environments. While live cell SM imaging has been achieved, these experiments are

far from routine. Similar to how the first in vitro SM experiments were initially restricted to

dedicated laboratories due to photodetection limits, live cell SM imaging is currently faced with

limits in fluorophore performance that prevent its widespread adoption for studying biology.

Chemical tags allow photophysically desirable fluorophores to used for protein labeling,

offering great advantages for SM imaging. At the same time, though, current organic

fluorophores are inherently limited in their total photon output by photobleaching. Further

investigations are needed to ensure that the chemical tags are compatible with organic

fluorophores. The identification and development of new fluorophores with improved photon

output and bioorthogonality is also necessary to improve the palette of fluorophores for live cell

imaging. Finally, the robustness and size of chemical tags can be applied for studying more

fragile biological systems, such as ion channels. The concurrent exploration of chemical tag

applications with fluorophore development can enable routine SM imaging for mechanistic

studies in biology.

1.8 References

(1) Yildiz, A.; Forkey, J. N.; McKinney, S. A.; Ha, T.; Goldman, Y. E.; Selvin, P. R.

Science 2003, 300, 2061.

(2) Lu, H. P.; Xun, L.; Xie, X. S. Science 1998, 282, 1877.

(3) Blanchard, S. C.; Kim, H. D.; Gonzalez, R. L., Jr.; Puglisi, J. D.; Chu, S.

Proceedings of the National Academy of Sciences of the United States of America 2004, 101,

12893.

Page 41: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

30

(4) Funatsu, T.; Harada, Y.; Tokunaga, M.; Saito, K.; Yanagida, T. Nature 1995, 374,

555.

(5) Myong, S.; Rasnik, I.; Joo, C.; Lohman, T. M.; Ha, T. Nature 2005, 437, 1321.

(6) Selvin, P. R.; Ha, T. Single-molecule techniques, a laboratory manual; Cold

Spring Harbor Laboratory Press: Cold Spring Harbor, NY, 2008.

(7) Elf, J.; Li, G.-W.; Xie, X. S. Science 2007, 316, 1191.

(8) Yu, J.; Xiao, J.; Ren, X.; Lao, K.; Xie, X. S. Science 2006, 311, 1600.

(9) Yang, W.; Gelles, J.; Musser, S. M. Proceedings of the National Academy of

Sciences of the United States of America 2004, 101, 12887.

(10) Joo, C.; Fareh, M.; Narry Kim, V. Trends in biochemical sciences 2013, 38, 30.

(11) Aubin, J. E. Journal of Histochemistry & Cytochemistry 1979, 27, 36.

(12) Sletten, E. M.; Bertozzi, C. R. Angewandte Chemie International Edition 2009,

48, 6974.

(13) Orrit, M.; Bernard, J. Phys. Rev. Lett. 1990, 65, 2716.

(14) Weiss, S. Science 1999, 283, 1676.

(15) Biteen, J. S.; Thompson, M. A.; Tselentis, N. K.; Bowman, G. R.; Shapiro, L.;

Moerner, W. E. Nat Meth 2008, 5, 947.

(16) Juette, M. F.; Terry, D. S.; Wasserman, M. R.; Zhou, Z.; Altman, R. B.; Zheng,

Q.; Blanchard, S. C. Current Opinion in Chemical Biology 2014, 20, 103.

(17) Mochalin, V. N.; Shenderova, O.; Ho, D.; Gogotsi, Y. Nat Nano 2012, 7, 11.

(18) Wang, Z.; Ding, X.; Li, S.; Shi, J.; Li, Y. RSC Advances 2014, 4, 7235.

(19) Miller, L. W.; Cai, Y.; Sheetz, M. P.; Cornish, V. W. Nat Meth 2005, 2, 255.

(20) Chen, Z.; Jing, C.; Gallagher, S. S.; Sheetz, M. P.; Cornish, V. W. Journal of the

American Chemical Society 2012, 134, 13692.

(21) Jing, C.; Cornish, V. W. ACS chemical biology 2013, 8, 1704.

(22) Cai, Y.; Rossier, O.; Gauthier, N. C.; Biais, N.; Fardin, M.-A.; Zhang, X.; Miller,

L. W.; Ladoux, B.; Cornish, V. W.; Sheetz, M. P. Journal of Cell Science 2010, 123, 413.

Page 42: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

31

(23) Gatzogiannis, E.; Chen, Z.; Wei, L.; Wombacher, R.; Kao, Y.-T.; Yefremov, G.;

Cornish, V. W.; Min, W. Chemical Communications 2012, 48, 8694.

(24) Reddy, D. R.; Pedro Rosa, L. E.; Miller, L. W. Bioconjug Chem 2011, 22, 1402.

(25) Long, M. J.; Pan, Y.; Lin, H. C.; Hedstrom, L.; Xu, B. Journal of the American

Chemical Society 2011, 133, 10006.

(26) Wombacher, R.; Heidbreder, M.; van de Linde, S.; Sheetz, M. P.; Heilemann, M.;

Cornish, V. W.; Sauer, M. Nat Meth 2010, 7, 717.

(27) Jones, S. A.; Shim, S.-H.; He, J.; Zhuang, X. Nat Meth 2011, 8, 499.

(28) Wilmes, S.; Staufenbiel, M.; Lisse, D.; Richter, C. P.; Beutel, O.; Busch, K. B.;

Hess, S. T.; Piehler, J. Angewandte Chemie (International ed. in English) 2012, 51, 4868.

(29) Moerner, W. E.; Kador, L. Phys. Rev. Lett. 1989, 62, 2535.

(30) Betzig, E.; Chichester, R. J. Science 1993, 262, 1422.

(31) Ha, T.; Enderle, T.; Ogletree, D. F.; Chemla, D. S.; Selvin, P. R.; Weiss, S.

Proceedings of the National Academy of Sciences 1996, 93, 6264.

(32) Schmidt, T.; Schutz, G. J.; Baumgartner, W.; Gruber, H. J.; Schindler, H.

Proceedings of the National Academy of Sciences of the United States of America 1996, 93,

2926.

(33) Dickson, R. M.; Norris, D. J.; Tzeng, Y.-L.; Moerner, W. E. Science 1996, 274,

966.

(34) Dickson, R. M.; Cubitt, A. B.; Tsien, R. Y.; Moerner, W. E. Nature 1997, 388,

355.

(35) Roy, R.; Hohng, S.; Ha, T. Nat Meth 2008, 5, 507.

(36) Ambrose, E. J. Experimental Cell Research 1961, 8, 54.

(37) Axelrod, D. The Journal of Cell Biology 1981, 89, 141.

(38) Mashanov, G. I.; Tacon, D.; Knight, A. E.; Peckham, M.; Molloy, J. E. Methods

2003, 29, 142.

(39) Coelho, M.; Maghelli, N.; Tolic-Norrelykke, I. M. Integrative Biology 2013, 5,

748.

Page 43: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

32

(40) Tokunaga, M.; Imamoto, N.; Sakata-Sogawa, K. Nat Meth 2008, 5, 159.

(41) Cella Zanacchi, F.; Lavagnino, Z.; Perrone Donnorso, M.; Del Bue, A.; Furia, L.;

Faretta, M.; Diaspro, A. Nat Meth 2011, 8, 1047.

(42) Ritter, J. G.; Veith, R.; Veenendaal, A.; Siebrasse, J. P.; Kubitscheck, U. PLoS

ONE 2010, 5, e11639.

(43) Nie, S.; Chiu, D.; Zare, R. Science 1994, 266, 1018.

(44) Abbe, E. Journal of the Royal Microscopical Society 1883, 3, 790.

(45) Qu, X.; Wu, D.; Mets, L.; Scherer, N. F. Proceedings of the National Academy of

Sciences of the United States of America 2004, 101, 11298.

(46) Burnette, D. T.; Sengupta, P.; Dai, Y.; Lippincott-Schwartz, J.; Kachar, B.

Proceedings of the National Academy of Sciences 2011, 108, 21081.

(47) Hell, S. W.; Wichmann, J. Optics Letters 1994, 19, 780.

(48) Willig, K. I.; Kellner, R. R.; Medda, R.; Hein, B.; Jakobs, S.; Hell, S. W. Nat

Meth 2006, 3, 721.

(49) Pinaud, F.; Dahan, M. Proceedings of the National Academy of Sciences 2011,

108, E201.

(50) Gordon, M. P.; Ha, T.; Selvin, P. R. Proceedings of the National Academy of

Sciences of the United States of America 2004, 101, 6462.

(51) Churchman, L. S.; Okten, Z.; Rock, R. S.; Dawson, J. F.; Spudich, J. A.

Proceedings of the National Academy of Sciences of the United States of America 2005, 102,

1419.

(52) Sharonov, A.; Hochstrasser, R. M. Proceedings of the National Academy of

Sciences 2006, 103, 18911.

(53) Lew, M. D.; Lee, S. F.; Ptacin, J. L.; Lee, M. K.; Twieg, R. J.; Shapiro, L.;

Moerner, W. E. Proceedings of the National Academy of Sciences 2011, 108, E1102.

(54) Watanabe, N.; Mitchison, T. J. Science 2002, 295, 1083.

(55) Danuser, G.; Waterman-Storer, C. M. Annual review of biophysics and

biomolecular structure 2006, 35, 361.

Page 44: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

33

(56) Betzig, E.; Patterson, G. H.; Sougrat, R.; Lindwasser, O. W.; Olenych, S.;

Bonifacino, J. S.; Davidson, M. W.; Lippincott-Schwartz, J.; Hess, H. F. Science 2006, 313,

1642.

(57) Rust, M. J.; Bates, M.; Zhuang, X. Nature methods 2006, 3, 793.

(58) Klein, T.; Loschberger, A.; Proppert, S.; Wolter, S.; van de Linde, S.; Sauer, M.

Nat Meth 2011, 8, 7.

(59) Shroff, H.; Galbraith, C. G.; Galbraith, J. A.; Betzig, E. Nature methods 2008, 5,

417.

(60) Bourgeois, D.; Adam, V. IUBMB Life 2012, 64, 482.

(61) Huang, B.; Wang, W.; Bates, M.; Zhuang, X. Science 2008, 319, 810.

(62) Heilemann, M.; van de Linde, S.; Mukherjee, A.; Sauer, M. Angewandte Chemie

International Edition 2009, 48, 6903.

(63) Huang, B.; Bates, M.; Zhuang, X. Annual review of biochemistry 2009, 78, 993.

(64) Schneckenburger, H.; Weber, P.; Wagner, M.; Schickinger, S.; Richter, V.; Bruns,

T.; Strauss, W. S. L.; Wittig, R. Journal of Microscopy 2012, 245, 311.

(65) Tsien, R. Y. Annual review of biochemistry 1998, 67, 509.

(66) Chalfie, M.; Tu, Y.; Euskirchen, G.; Ward, W.; Prasher, D. Science 1994, 263,

802.

(67) Shaner, N. C.; Steinbach, P. A.; Tsien, R. Y. Nat Meth 2005, 2, 905.

(68) Gebhardt, J. C. M.; Suter, D. M.; Roy, R.; Zhao, Z. W.; Chapman, A. R.; Basu,

S.; Maniatis, T.; Xie, X. S. Nat Meth 2013, 10, 421.

(69) Lippincott-Schwartz, J.; Patterson, G. H. Trends in Cell Biology 2009, 19, 555.

(70) Nienhaus, K.; Ulrich Nienhaus, G. Chemical Society Reviews 2014, 43, 1088.

(71) Hess, S. T.; Gould, T. J.; Gudheti, M. V.; Maas, S. A.; Mills, K. D.; Zimmerberg,

J. Proceedings of the National Academy of Sciences 2007, 104, 17370.

(72) Chudakov, D. M.; Lukyanov, S.; Lukyanov, K. A. Nature protocols 2007, 2,

2024.

(73) Subach, F. V.; Patterson, G. H.; Manley, S.; Gillette, J. M.; Lippincott-Schwartz,

J.; Verkhusha, V. V. Nature methods 2009, 6, 153.

Page 45: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

34

(74) Brakemann, T.; Stiel, A. C.; Weber, G.; Andresen, M.; Testa, I.; Grotjohann, T.;

Leutenegger, M.; Plessmann, U.; Urlaub, H.; Eggeling, C.; Wahl, M. C.; Hell, S. W.; Jakobs, S.

Nat Biotech 2011, 29, 942.

(75) Seefeldt, B.; Kasper, R.; Seidel, T.; Tinnefeld, P.; Dietz, K.-J.; Heilemann, M.;

Sauer, M. Journal of Biophotonics 2008, 1, 74.

(76) Wang, W.; Takimoto, J. K.; Louie, G. V.; Baiga, T. J.; Noel, J. P.; Lee, K.-F.;

Slesinger, P. A.; Wang, L. Nat Neurosci 2007, 10, 1063.

(77) Agbulut, O.; Coirault, C.; Niederlander, N.; Huet, A.; Vicart, P.; Hagege, A.;

Puceat, M.; Menasche, P. Nat Meth 2006, 3, 331.

(78) Huang, W.-Y.; Aramburu, J.; Douglas, P. S.; Izumo, S. Nat Med 2000, 6, 482.

(79) Fernandez-Suarez, M.; Ting, A. Y. Nat Rev Mol Cell Biol 2008, 9, 929.

(80) Rotman, B.; Zderic, J. A.; Edelstein, M. Proceedings of the National Academy of

Sciences 1963, 50, 1.

(81) Haugland, R. P. In The handbook—a guide to fluorescent probes and labeling

technologies; 10th ed.; Molecular Probes: Eugene, OR, 2005.

(82) Rotman, B.; Papermaster, B. W. Proceedings of the National Academy of

Sciences of the United States of America 1966, 55, 134.

(83) Davila, H. V.; Salzberg, B. M.; Cohen, L. B.; Waggoner, A. S. Nature New

Biology 1973, 241, 159.

(84) Mujumdar, R. B.; Ernst, L. A.; Mujumdar, S. R.; Lewis, C. J.; Waggoner, A. S.

Bioconjugate Chemistry 1993, 4, 105.

(85) Levitus, M.; Ranjit, S. Quarterly Reviews of Biophysics 2011, 44, 123.

(86) Altman, R. B.; Terry, D. S.; Zhou, Z.; Zheng, Q.; Geggier, P.; Kolster, R. A.;

Zhao, Y.; Javitch, J. A.; Warren, J. D.; Blanchard, S. C. Nat Meth 2012, 9, 68.

(87) Samanta, A.; Vendrell, M.; Das, R.; Chang, Y.-T. Chemical Communications

2010, 46, 7406.

(88) Greenspan, P.; Mayer, E. P.; Fowler, S. D. The Journal of Cell Biology 1985, 100,

965.

(89) Kuznetsova, N. A.; Marinina, L. E.; Alekseeva, V. I.; Luk'yanets, E. A. J Appl

Spectrosc 1978, 29, 978.

Page 46: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

35

(90) Heilemann, M.; van de Linde, S.; Schüttpelz, M.; Kasper, R.; Seefeldt, B.;

Mukherjee, A.; Tinnefeld, P.; Sauer, M. Angewandte Chemie International Edition 2008, 47,

6172.

(91) Vogelsang, J.; Cordes, T.; Forthmann, C.; Steinhauer, C.; Tinnefeld, P.

Proceedings of the National Academy of Sciences 2009, 106, 8107.

(92) Kottke, T.; van de Linde, S.; Sauer, M.; Kakorin, S.; Heilemann, M. The Journal

of Physical Chemistry Letters 2010, 1, 3156.

(93) Willets, K. A.; Ostroverkhova, O.; He, M.; Twieg, R. J.; Moerner, W. E. Journal

of the American Chemical Society 2003, 125, 1174.

(94) Lee, H.-l. D.; Lord, S. J.; Iwanaga, S.; Zhan, K.; Xie, H.; Williams, J. C.; Wang,

H.; Bowman, G. R.; Goley, E. D.; Shapiro, L.; Twieg, R. J.; Rao, J.; Moerner, W. E. Journal of

the American Chemical Society 2010, 132, 15099.

(95) Kohl, C.; Weil, T.; Qu, J.; Müllen, K. Chemistry – A European Journal 2004, 10,

5297.

(96) Margineanu, A.; Hofkens, J.; Cotlet, M.; Habuchi, S.; Stefan, A.; Qu, J.; Kohl, C.;

Müllen, K.; Vercammen, J.; Engelborghs, Y.; Gensch, T.; De Schryver, F. C. The Journal of

Physical Chemistry B 2004, 108, 12242.

(97) Cordes, T.; Vogelsang, J.; Anaya, M.; Spagnuolo, C.; Gietl, A.; Summerer, W.;

Herrmann, A.; Müllen, K.; Tinnefeld, P. Journal of the American Chemical Society 2010, 132,

2404.

(98) Peneva, K.; Mihov, G.; Nolde, F.; Rocha, S.; Hotta, J.-i.; Braeckmans, K.;

Hofkens, J.; Uji-i, H.; Herrmann, A.; Müllen, K. Angewandte Chemie 2008, 120, 3420.

(99) Yao, J.; Yang, M.; Duan, Y. Chemical Reviews 2014, 114, 6130.

(100) Michalet, X.; Pinaud, F. F.; Bentolila, L. A.; Tsay, J. M.; Doose, S.; Li, J. J.;

Sundaresan, G.; Wu, A. M.; Gambhir, S. S.; Weiss, S. Science 2005, 307, 538.

(101) Medintz, I. L.; Uyeda, H. T.; Goldman, E. R.; Mattoussi, H. Nat Mater 2005, 4,

435.

(102) Walling, M. A.; Novak, J. A.; Shepard, J. R. International journal of molecular

sciences 2009, 10, 441.

(103) Jaiswal, J. K.; Mattoussi, H.; Mauro, J. M.; Simon, S. M. Nat Biotech 2003, 21,

47.

Page 47: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

36

(104) Dahan, M.; Levi, S.; Luccardini, C.; Rostaing, P.; Riveau, B.; Triller, A. Science

2003, 302, 442.

(105) Chang, J. C.; Rosenthal, S. J. Methods in molecular biology (Clifton, N.J.) 2013,

991, 149.

(106) Schrand, A. M.; Huang, H.; Carlson, C.; Schlager, J. J.; Ōsawa, E.; Hussain, S.

M.; Dai, L. The Journal of Physical Chemistry B 2006, 111, 2.

(107) Yu, S. J.; Kang, M. W.; Chang, H. C.; Chen, K. M.; Yu, Y. C. Journal of the

American Chemical Society 2005, 127, 17604.

(108) Gruber, A.; Dräbenstedt, A.; Tietz, C.; Fleury, L.; Wrachtrup, J.; Borczyskowski,

C. v. Science 1997, 276, 2012.

(109) Fu, C.-C.; Lee, H.-Y.; Chen, K.; Lim, T.-S.; Wu, H.-Y.; Lin, P.-K.; Wei, P.-K.;

Tsao, P.-H.; Chang, H.-C.; Fann, W. Proceedings of the National Academy of Sciences 2007,

104, 727.

(110) BradacC; GaebelT; NaidooN; Sellars, M. J.; TwamleyJ; Brown, L. J.; Barnard, A.

S.; PlakhotnikT; Zvyagin, A. V.; Rabeau, J. R. Nat Nano 2010, 5, 345.

(111) McGuinness, L. P.; YanY; StaceyA; Simpson, D. A.; Hall, L. T.; MaclaurinD;

PrawerS; MulvaneyP; WrachtrupJ; CarusoF; Scholten, R. E.; Hollenberg, L. C. L. Nat Nano

2011, 6, 358.

(112) Glazer, A. N. Annual review of biochemistry 1970, 39, 101.

(113) Feramisco, J. R. Proceedings of the National Academy of Sciences 1979, 76,

3967.

(114) Debets, M. F.; van Hest, J. C. M.; Rutjes, F. P. J. T. Organic & Biomolecular

Chemistry 2013, 11, 6439.

(115) Griffin, B. A.; Adams, S. R.; Tsien, R. Y. Science 1998, 281, 269.

(116) Cao, H.; Xiong, Y.; Wang, T.; Chen, B.; Squier, T. C.; Mayer, M. U. Journal of

the American Chemical Society 2007, 129, 8672.

(117) Bhunia, A. K.; Miller, S. C. Chembiochem : a European journal of chemical

biology 2007, 8, 1642.

(118) Adams, S. R.; Campbell, R. E.; Gross, L. A.; Martin, B. R.; Walkup, G. K.; Yao,

Y.; Llopis, J.; Tsien, R. Y. Journal of the American Chemical Society 2002, 124, 6063.

Page 48: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

37

(119) Gaietta, G.; Deerinck, T. J.; Adams, S. R.; Bouwer, J.; Tour, O.; Laird, D. W.;

Sosinsky, G. E.; Tsien, R. Y.; Ellisman, M. H. Science 2002, 296, 503.

(120) Ojida, A.; Honda, K.; Shinmi, D.; Kiyonaka, S.; Mori, Y.; Hamachi, I. Journal of

the American Chemical Society 2006, 128, 10452.

(121) Hauser, C. T.; Tsien, R. Y. Proceedings of the National Academy of Sciences of

the United States of America 2007, 104, 3693.

(122) Uttamapinant, C.; White, K. A.; Baruah, H.; Thompson, S.; Fernandez-Suarez,

M.; Puthenveetil, S.; Ting, A. Y. Proceedings of the National Academy of Sciences of the United

States of America 2010, 107, 10914.

(123) Mao, H.; Hart, S. A.; Schink, A.; Pollok, B. A. Journal of the American Chemical

Society 2004, 126, 2670.

(124) Popp, M. W.; Antos, J. M.; Grotenbreg, G. M.; Spooner, E.; Ploegh, H. L. Nat

Chem Biol 2007, 3, 707.

(125) Lin, C. W.; Ting, A. Y. Journal of the American Chemical Society 2006, 128,

4542.

(126) Howarth, M.; Takao, K.; Hayashi, Y.; Ting, A. Y. Proceedings of the National

Academy of Sciences of the United States of America 2005, 102, 7583.

(127) Sugimura, Y.; Hosono, M.; Wada, F.; Yoshimura, T.; Maki, M.; Hitomi, K. The

Journal of biological chemistry 2006, 281, 17699.

(128) Keppler, A.; Gendreizig, S.; Gronemeyer, T.; Pick, H.; Vogel, H.; Johnsson, K.

Nat Biotech 2003, 21, 86.

(129) Gautier, A.; Juillerat, A.; Heinis, C.; Correa, I. R., Jr.; Kindermann, M.; Beaufils,

F.; Johnsson, K. Chemistry & biology 2008, 15, 128.

(130) Corrêa Jr, I. R. Current Opinion in Chemical Biology 2014, 20, 36.

(131) Bojkowska, K.; Santoni de Sio, F.; Barde, I.; Offner, S.; Verp, S.; Heinis, C.;

Johnsson, K.; Trono, D. Chemistry & biology 2011, 18, 805.

(132) Roed, S. N.; Wismann, P.; Underwood, C. R.; Kulahin, N.; Iversen, H.; Cappelen,

K. A.; Schaffer, L.; Lehtonen, J.; Hecksher-Soerensen, J.; Secher, A.; Mathiesen, J. M.; Brauner-

Osborne, H.; Whistler, J. L.; Knudsen, S. M.; Waldhoer, M. Molecular and cellular

endocrinology 2014, 382, 938.

Page 49: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

38

(133) Maurel, D.; Comps-Agrar, L.; Brock, C.; Rives, M. L.; Bourrier, E.; Ayoub, M.

A.; Bazin, H.; Tinel, N.; Durroux, T.; Prezeau, L.; Trinquet, E.; Pin, J. P. Nature methods 2008,

5, 561.

(134) Los, G. V.; Encell, L. P.; McDougall, M. G.; Hartzell, D. D.; Karassina, N.;

Zimprich, C.; Wood, M. G.; Learish, R.; Ohana, R. F.; Urh, M.; Simpson, D.; Mendez, J.;

Zimmerman, K.; Otto, P.; Vidugiris, G.; Zhu, J.; Darzins, A.; Klaubert, D. H.; Bulleit, R. F.;

Wood, K. V. ACS chemical biology 2008, 3, 373.

(135) Jing, C.; Cornish, V. W. Accounts of chemical research 2011, 44, 784.

(136) Koudelakova, T.; Bidmanova, S.; Dvorak, P.; Pavelka, A.; Chaloupkova, R.;

Prokop, Z.; Damborsky, J. Biotechnology Journal 2013, 8, 32.

(137) Ohana, R. F.; Encell, L. P.; Zhao, K.; Simpson, D.; Slater, M. R.; Urh, M.; Wood,

K. V. Protein expression and purification 2009, 68, 110.

(138) Watanabe, S.; Mizukami, S.; Akimoto, Y.; Hori, Y.; Kikuchi, K. Chemistry

(Weinheim an der Bergstrasse, Germany) 2011, 17, 8342.

(139) Hori, Y.; Norinobu, T.; Sato, M.; Arita, K.; Shirakawa, M.; Kikuchi, K. Journal

of the American Chemical Society 2013, 135, 12360.

(140) Bonasio, R.; Carman, C. V.; Kim, E.; Sage, P. T.; Love, K. R.; Mempel, T. R.;

Springer, T. A.; von Andrian, U. H. Proceedings of the National Academy of Sciences of the

United States of America 2007, 104, 14753.

(141) Sun, X.; Zhang, A.; Baker, B.; Sun, L.; Howard, A.; Buswell, J.; Maurel, D.;

Masharina, A.; Johnsson, K.; Noren, C. J.; Xu, M. Q.; Correa, I. R., Jr. Chembiochem : a

European journal of chemical biology 2011, 12, 2217.

(142) Gallagher, S. S.; Sable, J. E.; Sheetz, M. P.; Cornish, V. W. ACS chemical biology

2009, 4, 547.

(143) Liu, W.; Li, F.; Chen, X.; Hou, J.; Yi, L.; Wu, Y. W. Journal of the American

Chemical Society 2014, 136, 4468.

(144) Bosch, Peter J.; Corrêa, Ivan R., Jr.; Sonntag, Michael H.; Ibach, J.; Brunsveld,

L.; Kanger, Johannes S.; Subramaniam, V. Biophysical Journal, 107, 803.

(145) Zhao, Z. W.; Roy, R.; Gebhardt, J. C. M.; Suter, D. M.; Chapman, A. R.; Xie, X.

S. Proceedings of the National Academy of Sciences 2014, 111, 681.

(146) Farlow, J.; Seo, D.; Broaders, K. E.; Taylor, M. J.; Gartner, Z. J.; Jun, Y.-w. Nat

Meth 2013, 10, 1203.

Page 50: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

39

(147) Liu, D. S.; Phipps, W. S.; Loh, K. H.; Howarth, M.; Ting, A. Y. ACS nano 2012,

6, 11080.

(148) Hoskins, A. A.; Friedman, L. J.; Gallagher, S. S.; Crawford, D. J.; Anderson, E.

G.; Wombacher, R.; Ramirez, N.; Cornish, V. W.; Gelles, J.; Moore, M. J. Science 2011, 331,

1289.

(149) Fabrizio, P.; Dannenberg, J.; Dube, P.; Kastner, B.; Stark, H.; Urlaub, H.;

Lührmann, R. Molecular Cell 2009, 36, 593.

(150) Crawford, D. J.; Hoskins, A. A.; Friedman, L. J.; Gelles, J.; Moore, M. J. RNA

(New York, N.Y.) 2008, 14, 170.

Page 51: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

40

Chapter 2

Covalent Trimethoprim Chemical Tags For Single Molecule Imaging

A portion of this chapter was published in Wang, Tracy Y.; Friedman, Larry J.; Gelles, J.; Min,

W.; Hoskins, Aaron A.; Cornish, Virginia W. Biophysical Journal 2014, 106, 272.

Page 52: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

41

2.1 Chapter Outlook

Fluorophore modularity with chemical tags is critical for their advantageous properties

over the FPs. With the TMP-tag, a protein can be labeled with virtually any fluorophore that can

be chemically conjugated to TMP. While enabling SM imaging with high performance

fluorophores, the TMP-tag and other chemical tags are presumed to not affect fluorophore

photophysical properties. With the demand for high photon output, it is imperative that the

chemical tags do not affect fluorophore brightness and photostability. The local chemical

environment can strongly affect fluorophore performance. Conjugation of a ligand to a

fluorophore or bringing the fluorophore in close proximity to a protein may adversely affect a

fluorophore that is being used with chemical tags. Unfortunately, the effect of the chemical tags

on fluorophore photon output had not been systematically investigated.

This chapter examines the effect that the covalent TMP-tag has on the performance of

common oxazine and cyanine SM imaging fluorophores. We examine the SM properties of

fluorophores with and without the TMP-tag under conditions that simulate those of the live cell.

In addition, we also investigate whether ensemble properties of fluorophores can act as indicators

of SM properties for screening fluorophores. We demonstrate that the covalent TMP tag

generally does not affect the total photon output fluorophores, establishing the benefit of using

chemical tags with organic fluorophores for SM live cell imaging. I am the main contributor to

this project. I designed and carried out the experiment to study both ensemble and single molecule

properties of fluorophores under the guidance of Wei Min, Aaron Hoskins, and Larry Friedman. I

wrote the chapter with contributions from Virginia Cornish and all other co-authors.

Page 53: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

42

2.2 Introduction

Single molecule (SM) imaging of biomolecules has transformed our ability to probe

biology.1-4

Over the past decade, SM imaging has provided insights into the molecular

mechanism of motor protein walking, conformational transitions of the ribosome, and structural

heterogeneity in enzyme catalysis, among others.5-12

However, these studies are often rely on

reconstituted pathways, which are not representative of native cell conditions and not suitable for

investigating biological processes such as spliceosome, replisome, and protein kinase signaling

activity.13-15

Broadly accessible SM imaging in live cells or whole-cell extracts would be

transformative for studying biological mechanism, but requires the ability to selectively label

proteins with bright and photostable fluorophores that can overcome background and cellular

autofluorescence. Fluorescent proteins are genetically encoded and inherently selective, but

generally lack sufficient photon output for SM imaging. While the best organic fluorophores

have nearly an order of magnitude greater photon output than the FPs, they lack labeling

selectivity.5,16-20

By combining genetic encoding with the advantages of organic fluorophores,

chemical tags can overcome the need for selective fluorescent labels in SM imaging.

Chemical tags fluorescently label proteins for high performance imaging applications by

genetically fusing proteins to a polypeptide that binds to an organic fluorophore. The modularity

of the chemical tags allows labeling with a wide variety of commercially available organic

fluorophores with similar selectivity as the FPs.21,22

Chemical tags, such as the TMP-tag, SNAP-

tag, and Halo-tag have been used extensively for biological imaging including live cell SM super

resolution imaging.23-28

Based on the high affinity interaction between trimethoprim (TMP) and

Escherichia coli dihydrofolate reductase (eDHFR), the covalent TMP-tag (A-TMP-tag), is

particularly well-suited for SM imaging (Figure 2-1A). 23

Target proteins are tagged with

Page 54: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

43

eDHFR that has an engineered cysteine nucleophile outside the TMP-binding pocket

(eDHFR:L28C) and covalently labeled by an acrylamide-TMP-fluorophore (A-TMP-

fluorophore) via a proximity-induced Michael reaction between the cysteine and the acrylamide

following TMP-eDHFR binding.29,30

The selectivity and fast labeling kinetics of the A-TMP-tag

has permitted high resolution live cell imaging of both nuclear and cytoplasmic proteins.30

Permanent labeling by the A-TMP-tag can expand the utility of the non-covalent TMP-tag,

which has established SM imaging capabilities in studies of spliceosome assembly in whole cell

yeast extracts and demonstrations of super resolution imaging of nuclear H2B in live cells.27,31

Combining the advantage of covalent labeling with the performance of the non-covalent TMP-

tag, the A-TMP-tag meets the rigorous demands for selectively labeling proteins with organic

fluorophores in for SM imaging.

Figure 2-1: The covalent trimethoprim chemical tag

(A) Schematic cartoon of the covalent trimethoprim chemical tag (A-TMP-tag). A target protein is tagged with

an E. coli dihydrofolate reductase cysteine mutant (eDHFR:L28C) and covalently bound to a cell-permeable

acrylamide-trimethoprim-fluorophore (A-TMP-fluorophore). (B) Schematic of fluorophores and their A-TMP-tag

conjugates examined in this investigation.

Broad application of the A-TMP-tag and other chemical tags for SM imaging is

dependent on the tags maintaining the photophysical properties of the organic fluorophores used

with them. Although the chemical tags are assumed to not impact the properties of organic

Page 55: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

44

fluorophores, the tags introduce alterations to both chemical structure and local environment that

can significantly impact a fluorophore's SM imaging capabilities. With the A-TMP-tag,

fluorophore conjugation to the electron-rich small molecule A-TMP may cause quenching by

intramolecular electron transfer, which has been previously observed with fluorophores

conjugated to O-benzylguanine in SNAP-tag.32,33

Furthermore the A-TMP-fluorophore is

covalently bound to the eDHFR:L28C protein. While protein binding may recover quenched

fluorescence, as in the case of SNAP-tag, the fluorophore may remain quenched by proximity to

electron-rich amino acid residues such as tryptophan, phenylalanine, tyrosine, and/or

histidine.34-37

These interactions and other effects caused by the A-TMP-tag, such as sterics, local

polarity and electrostatics, may considerably influence fluorophore photon output and resultant

SM imaging performance.

Systematic evaluation of the photophysical properties of fluorophores with chemical tags

is crucial to realizing the potential of the tags for SM imaging. To understand the impact

chemical and environmental modifications of the A-TMP-tag have on fluorophore SM imaging

performance, we measure both SM and ensemble properties of fluorophores, fluorophores

conjugated to A-TMP (A-TMP-fluorophore) and fluorophores conjugated to A-TMP bound to

eDHFR:L28C (tagged fluorophore) (Figure 2-1B). Because overcoming cellular

autofluorescence and background noise while minimizing photobleaching are major challenges

for biological SM imaging, we focus our investigation on properties corresponding to brightness

and photostability. We determine SM photon flux, survival lifetime and total photon output

under different buffer conditions that mimic the intracellular environment. Since SM

measurements are technically challenging and require specialized imaging equipment to perform,

we sought to measure ensemble properties that could serve as indicators of photon flux and

Page 56: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

45

survival lifetime. As a result, we also measure quantum yield and ensemble photostability

lifetime. While ensemble properties are not equivalents for SM properties, correlation between

these properties indicates that ensemble measurements can be adapted for more rapid screening

of fluorophores and chemical tags that are suitable for SM imaging. We focus our investigation

on two fluorophores, an oxazine, Atto655, and a cyanine, Alexa647, because these two

fluorophores were used with chemical tags to label nuclear H2B and clathrin coated pits in the

cytosol in demonstrations of live cell SM super resolution imaging.26,27

Studying these

fluorophores with the A-TMP-tag gives insights to the properties that allow for successful SM

imaging. We assess the impact of the A-TMP-tag on fluorophore SM imaging performance by

comparing the SM and ensemble properties of Atto655, Alexa647, and their A-TMP-tag

conjugates.

2.3 Experimental Methods

2.3.1 Chemical Synthesis

Chemical structures of fluorophores and A-TMP-fluorophores are shown in Figure 2-2.

Alexa647-NHS ester (Invitrogen), Atto655-NHS ester (Atto tec), Atto680-NHS ester (Atto tec),

and Cy3-NHS ester (GE Life Sciences) were used without further purification to characterize

unmodified fluorophores and synthesize A-TMP-fluorophore conjugates. Atto655-biotin was

purchased from Sigma and used without further purification for SM experiments.

A-TMP was synthesized as previously described.30

A-TMP-fluorophores were

synthesized by adding 1 mg of fluorophore NHS ester to 500 L DMF, 1 equivalent of A-TMP

and 5 L triethylamine (Figure 2-2). The mixture was stirred at RT for 16 hours before being

concentrated. A-TMP-fluorophores were purified by reverse phase HPLC.

Page 57: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

46

Figure 2-2: Scheme of fluorophore modification by A-TMP-tag.

A) Commercially available succinimidyl ester fluorophores react with acrylamide-trimethoprim-amine (A-TMP-

NH2) to form the A-TMP-fluorophore. B) Covalent attachment of A-TMP-Fl to E. coli dihydrofolate reductase

cysteine mutant (eDHFR:L28C) is initiated by the high affinity binding between eDHFR and TMP and followed by

the proximity induced reaction between the cysteine thiol and acrylamide, forming the tagged-fluorophore. C)

Chemical structures of Atto655, Alexa647, Atto680 and Cy3.

Alexa647-biotin for SM experiments was synthesized by adding 1 mg Alexa647 NHS

ester and 1 mg Amine-PEG2-biotin (Thermo Scientific) with 5 L triethylamine in DMF (Figure

2-3). The mixture was stirred at RT for 16 hours before being concentrated. The reaction was

purified by reverse phase HPLC.

Page 58: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

47

Figure 2-3: Scheme of biotin-Alexa647 synthesis

2.3.2 Protein Expression, Purification, Labeling and Biotinylation

The vector encoding eDHFR:L28C for E.coli over-expression and protein purification

has been previously published.29

Plasmids were expressed in BL21(DE3) pLysS cells

(Invitrogen). Cells were grown at 37°C to an OD600 of 0.6, induced with 0.4 mM IPTG for three

hours and purified using a nickel sepharose column (HisTrap HP, GE Life Sciences). The protein

was dialyzed in phosphate buffered saline (PBS) at 4°C, snap frozen and stored at −80°C.

For preparation of tagged-fluorophores, 200 M eDHFR:L28C in PBS was thawed at

4°C and incubated with 1mM A-TMP-fluorophore and 1mM NADPH (Sigma) for two hours

(Figure 2-4). Unlabeled A-TMP-fluorophore was separated from tagged-fluorophores using 7000

MWCO Zeba desalting spin column (Thermo Scientific) equilibrated with PBS.

Page 59: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

48

Figure 2-4 : Demonstration of covalent fluorophore labeling of eDHFR:L28C

Binding of eDHFR (18 kDa) with A-TMP-Fl results in a shift of approximately 1 kDa that can be fluorescently

visualized. A) Gel shift assay of eDHFR:L28C-A-TMP-Fl and B) Merged image of in-gel green fluorescence (Cy3)

and red fluorescence (Atto655, Alexa647) of labeled eDHFR:L28C. The fluorescence of Atto680 is too red-shifted

to be easily detected. Column 1: protein ladder. Column 2: wt eDHFR. Column 3: eDHFR:L28C. Column 4:

eDHFR:L28C with A-TMP-Cy3. Column 5: eDHFR:L28C with A-TMP-Alexa647. Column 6: eDHFR:L28C-A-

TMP Atto655. Column 7: eDHFR:L28C-A-TMP-Atto680.

The vector encoding of eDHFR:L28C-bioseq, for E. coli expression and protein

purification has been previously published.38

eDHFR:L28C expression and purification were

carried our as previously described.39

eDHFR:L28C-bioseq biotinylation was carried as

previously described.39,40

Biotinylated eDHFR:L28C was dialyzed in phosphate buffered saline

(PBS) at 4°C, snap frozen and stored at −80°C and thawed at 4°C before use in SM experiments.

For preparation of biotinylated tagged-fluorophores, 1 M biotinylated eDHFR:L28C with 2 M

A-TMP-Atto655 or A-TMP-Alexa647 and 2 M NADPH in PBS was incubated for two hours

and used for imaging without further separation.

2.3.3 Single Molecule Methods

Single molecule imaging of fluorophore-biotin and biotinylated tagged-fluorophores was

performed on a homebuilt TIRF microscope. Samples were immobilized on glass slides

Page 60: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

49

passivated with PEG and PEG-streptavidin Samples were excited with 633 nm laser at 250

Watt and imaged until at least 90% of molecules were photobleached. Photostability survival

time was calculated by plotting the individual traces of 300-500 individual molecules,

identifying the photobleaching time, and fitting those times to a maximum likelihood single

exponential fit of survival times. Photon flux was measured by picking 10-20 well resolved

single molecules and integrating the total Gaussian fluorophore signal over time. The

background, as determined by the average signal after photobleaching, was subtracted from the

average fluorescence signal. The resulting signal was converted to photon flux using the ADU

conversion factor that had been determined for that camera using the calibration protocol

previously described.41

Total photon output was calculated by multiplying photon flux by

survival lifetime.

2.3.4 Ensemble Methods

The quantum yields were determined using the comparative method.42

Aqueous solutions

of fluorophore, A-TMP-fluorophore, and tagged-fluorophore in PBS buffer were diluted to

absorbances less than 0.1 to prevent inner filter effects. Absorbance measurements and spectra

were measured using a Tecan Infinite 200 and fluorescence measurements and spectra were

obtained with a Horiba Scientific Fluorolog-3 spectrofluorometer. The slope of the plot of

fluorescence emission compared to absorbance was compared to that of a quantum yield

reference solution under the same excitation and collection conditions. This measurement was

repeated in triplicate for each fluorophore. Fluorescence quantum yield standards used were

cresyl violet in methanol and nile blue in methanol.

Photobleaching time constants were determined using a microdroplet photobleaching

assay, similar to one used by Tsien and coworkers.43

Aqueous microdroplets of fluorophore, A-

Page 61: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

50

TMP-fluorophore, and tagged-fluorophore in PBS buffer were created under mineral oil that had

been previously extracted with PBS buffer. The microdroplets were bleached using laser

excitation (532 nm or 633 nm) such that the laser beam incidence is greater than the size of

individual droplets.

The absorbance and emission spectra for fluorophores and their A-TMP-tag conjugates

were measured. We observe no differences between fluorophores and their A-TMP-tag

conjugates for all fluorophores.

2.4 Results

2.4.1 Single Molecule Total Photon Output

To examine the effect the A-TMP-tag has on SM imaging performance, we determine the

total photon output of fluorophores and tagged-fluorophores (Figure 2-1B). Using total internal

reflection (TIRF) microscopy, we measure SM photon flux and survival lifetime, which are used

to calculate total photon output. The need to biotinylate fluorophores for biotin-streptavidin

immobilization in TIRF microscopy limits our SM evaluation to fluorophores and the tagged-

fluorophore. Photon flux and survival lifetimes for fluorophores and tagged-fluorophores are

measured in PBS buffer, PBS buffer containing 10 mM β-mercaptoethylamine (MEA) to

reproduce the reducing environment of the cell, and PBS containing 40% whole cell yeast extract

(YE) to reproduce the macromolecular density of the cell. SM photon fluxes and survival

lifetimes for fluorophores and tagged-fluorophores for Atto655 and Alexa647 are displayed in

Figure 2-5 and Table 2-1.

Page 62: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

51

Figure 2-5: Single molecule photon fluxes and survival lifetimes

Fluorophores (Fl) and tagged-fluorophores (Tagged-Fl) for Atto655 and Alexa647 in PBS buffer, 10 mM β-

mercaptoethylamine (MEA), and 40% yeast extract (YE) with 250 W 633 nm laser illumination. A) SM photon

flux, which is determined by the number of detected photons. B) Average SM survival lifetime before

photobleaching.

Table 2-1: Single molecule photon flux and survival lifetime

Photon flux a (10

3 photons/s)

Survival lifetime (min)

PBS MEA YE PBS MEA YE

Atto655 5 ± 2 4.3 ± 0.8 4.6 ± 0.7 8.5 ± 0.5- 0.93 ± 0.03 4.0 ± 0.2- Tagged-Atto655 4.3 ± 0.6 4.2 ± 0.8- 3.4 ± 0.4- 5.8 ± 0.5- 1.43 ± 0.08 3.5 ± 0.2-

Alexa647 13 ± 2- 13 ± 2- 16 ± 2- 0.20 ± 0.01 0.097 ± 0.005 0.35 ± 0.02 Tagged-Alexa647 13 ± 2- 10.1 ± 0.7- 11 ± 1- 0.25 ± 0.02 0.13 ± 0.01 0.29 ± 0.02 a Photon flux is the number of photons detected by the camera.

Page 63: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

52

Differences in photon flux and survival lifetimes between fluorophores and tagged-

fluorophores are generally within fifty percent and vary between buffer conditions. The photon

fluxes of the tagged-fluorophores are either the same or lower than those of their counterparts. In

PBS, the tagged-fluorophores have no differences in photon flux from the fluorophores. On the

other hand, in yeast extract, the tagged-fluorophores have thirty percent lower photon fluxes than

those of the untagged fluorophores. In 10 mM MEA, tagged-Atto655 has the same photon flux

as that of Atto655 but tagged-Alexa647 has a twenty percent lower photon flux than that of

Alexa647.

The survival lifetimes of the tagged-fluorophores are either higher or lower than those of

the fluorophores. In PBS, tagged-Atto655 has a thirty percent shorter survival lifetime than that

of Atto655 while tagged-Alexa647 has a twenty percent longer survival lifetime than that of

Alexa647. In the presence of 10 mM MEA, the tagged-fluorophores have longer survival

lifetimes than those of the fluorophores, with fifty percent longer survival lifetime for tagged-

Atto655 and a thirty percent longer survival lifetime for tagged-Alexa647. In yeast extract, the

tagged-fluorophores have twenty and ten percent shorter survival lifetimes than those of the

fluorophores for Atto655 and Alexa647, respectively.

The fluorophores have greater differences in photon flux and survival lifetime between

buffer conditions than between the fluorophore and tagged-fluorophores in the same conditions.

Both Atto655 and Alexa647 have lower survival lifetimes in MEA than those in PBS, with the

survival lifetime of Atto655 in MEA nearly an order of magnitude less than that in PBS.

Although the survival lifetime of Atto655 in yeast extract is less than half of that in PBS, the

survival lifetime of Alexa647 in yeast extract is thirty percent longer than that in PBS. The

fluorophores have no differences in photon flux between the buffer conditions. Overall, the

Page 64: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

53

photon fluxes of Alexa647 are two to three times higher than those of Atto655 across the

investigated conditions. However, the survival lifetimes of Atto655 are nearly an order of

magnitude greater than those of Alexa647.

SM total photon flux for fluorophores and tagged-fluorophores for Atto655 and Alexa647

are displayed in Figure 2-6 and Table 2-2. Although the tagged fluorophores exhibit some

differences in photon flux and survival lifetime from the untagged fluorophores, the total photon

output of the tagged-fluorophores is the same or on the same order of magnitude as that of the

fluorophores. The total photon outputs of the tagged-Atto655 are the same as Atto655 across all

the investigated conditions. The total photon outputs of tagged-Alexa647 are the same as

Alexa647 in PBS and in 10 mM MEA. Alexa647 in yeast extract is the only fluorophore that

exhibits a difference in photon output between the fluorophore and the tagged-fluorophore, with

the tagged-Alexa647 having thirty percent less photon output than that of Alexa647. However,

the photon output of tagged-Alexa647 in yeast extract is the same as the photon output of both

Alexa647 and tagged-Alexa647 in PBS. The TMP-tag does not affect the overall SM imaging

performance of Atto655 and Alexa647.

Figure 2-6: Total photon output

Page 65: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

54

Fluorophores (Fl) and tagged-fluorophores (Tagged-Fl) for Atto655 and Alexa647 in PBS buffer, 10 mM β-

mercaptoethylamine (MEA), and 40% yeast extract (YE) with 250 W 633 nm laser illumination. Total photon

output is determined by multiplying photon flux by the average survival lifetime.

Table 2-2: Single molecule total photon output

Total photon output b (x10

5 photons)

PBS MEA YE Atto655 27 ± 9 2.4 ± 0.6 10 ± 2 Tagged-Atto655 15 ± 3 3.7 ± 0.9 -7 ± 1-

Alexa647 1.6 ± 0.3 0.8 ± 0.1 3.3 ± 0.5 Tagged-Alexa647 2.0 ± 0.4 0.7 ± 0.1 1.9 ± 0.3 b Total photon output is determined by multiplying photon flux by survival lifetime.

Overall, the total photon output of Atto655 is approximately an order of magnitude

greater than that of Alexa647. Atto655 has the greatest photon output in PBS, which is 2,700,000

± 900,000 photons. However, in yeast extract, Atto655 photon output is sixty percent less than

the output in PBS. Alexa647 has the greatest photon output in yeast extract at 330,000 ± 50,000

photons, which is nearly double the photon output in PBS. Both fluorophores have lowest photon

output in MEA with Atto655 having nearly an order of magnitude lower output than that in PBS

and Alexa647 having nearly half the output than that in PBS.

2.4.2 Quantum Yield and Ensemble Photostability Lifetime

We measure the quantum yields and photostability lifetimes of Atto655, Alexa647 and

their corresponding A-TMP-fluorophores and tagged-fluorophores (Figure 2-1B) to examine the

correlation between ensemble and SM properties. We also perform ensemble measurements on

an additional oxazine, Atto680, and cyanine, Cy3, to more broadly understand the impacts of the

A-TMP-tag on oxazines and cyanines. The ensemble properties of these fluorophores and their

A-TMP-tag conjugates in PBS buffer are shown in Figure 2-7 and Table 2-3.

Page 66: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

55

Figure 2-7: Quantum yield and photostability lifetime

Fluorophores (Fl), A-TMP-fluorophore (A-TMP-Fl) and tagged-fluorophores (Tagged-Fl) for Atto655 and

Alexa647 in PBS buffer. A) Quantum yield. B) Photostability lifetime is the half life of fluorescence signal due to

photobleaching.

Page 67: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

56

Table 2-3: Ensemble photophysical properties in PBS buffer

Quantum Yield ()

Photostability Lifetime (min)

ex

(nm) em (nm)

Atto655 0.21 ± 0.02 23 ± 3- 663 680 A-TMP-Atto655 0.08 ± 0.01 7.7 ± 0.8 663 680 A-TMP-tagged Atto655 0.23 ± 0.02 8.9 ± 0.3 663 680

Atto680 0.17 ± 0.02 37 ± 7- 680 702 A-TMP-Atto680 0.09 ± 0.01 20 ± 1- 680 702 A-TMP-tagged Atto680 0.19 ± 0.02 12.5 ± 0.7- 680 702 Alexa647 0.29 ± 0.03 0.148 ± 0.007 651 660 A-TMP-Alexa647 0.34 ± 0.03 0.35 ± 0.01 651 660 A-TMP-tagged Alexa647 0.36 ± 0.04 0.18 ± 0.01 651 660 Cy3 0.05 ± 0.01 12.8 ± 0.2 550 555 A-TMP-Cy3 0.11 ± 0.01 21 ± 2- 550 555 A-TMP-tagged Cy3 0.13 ± 0.02 14.4 ± 0.3 550 555

The two fluorophore classes exhibit trends in quantum yield and photostability lifetime

between the fluorophores and their A-TMP-tag conjugates. For Atto655 and Atto680, the tagged-

oxazines have the same quantum yield as the oxazines even though the A-TMP-oxazines have a

lower quantum yield. However, the photostability lifetimes of the A-TMP-oxazines and tagged-

oxazines are both shorter than those the oxazines. For the cyanines Alexa647 and Cy3, the A-

TMP-cyanines and the tagged-cyanines have greater or the same quantum yields as the cyanines.

Although the tagged-cyanines have shorter photostability lifetimes than those the A-TMP-

cyanines, they are still longer than those of the cyanines.

Overall, the tagged-fluorophores have the same if not better quantum yield than the

fluorophores. While the tagged-cyanines are more photostable than the cyanines, the tagged-

oxazines are less photostable than the oxazines. The A-TMP-tag does not affect the ensemble

photophysical properties of the cyanines or the quantum yield of oxazines, but does reduce the

Page 68: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

57

photostability of the oxazines. Alexa647 has the highest quantum yield, but is the least

photostable out of the examined fluorophores. Atto680, Atto655 and Cy3 are approximately an

order of magnitude more photostable than Alexa647, with Atto680 being the most photostable.

2.5 Discussion

These results establish that the A-TMP-tag complements the advantageous SM imaging

properties of Atto655 and Alexa647, facilitating SM imaging with these fluorophores. Based on

the SM examination of these fluorophores and their tagged-counterparts under buffers that mimic

the reducing potential and macromolecular density of the cell, we determine the A-TMP-tag has

buffer dependent effects on photon flux and survival lifetime. However, the A-TMP-tag does not

affect total photon output in most cases. Although the A-TMP-tag lowers Alexa647 photon

output in yeast extract, this output is the same as that for Alexa647 in PBS. With the best organic

fluorophores having photon outputs only approximately ten-fold greater than the best FPs, the A-

TMP-tag's preservation of Atto655 and Alexa647 photon output is significant for upholding

these fluorophores' SM imaging performances.5,16-20

These findings also reveal that Atto655 is

superior to Alexa647 for imaging on longer timescales, with survival lifetimes and total photon

outputs nearly an order of magnitude greater than those of Alexa647. At the same time, the

photon fluxes of Alexa647 are two to three times greater than those of Atto655, making

Alexa647 better suited for applications requiring positional localization.44

The comparative

advantages of survival lifetime and photon flux between these two fluorophores are not altered

by the A-TMP-tag, further indicating that the A-TMP-tag is valuable tool for labeling proteins

for SM imaging applications.

These findings reflect Atto655 and Alexa647 performance for SM imaging and highlight

the importance of fluorophore characterization under different conditions. Recent studies of

Page 69: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

58

fluorophore performance in various buffer conditions and in fixed cells have begun to provide

guidelines for selecting appropriate fluorophores for SM and super resolution imaging.45,46

Our

investigation goes a step further by characterizing fluorophores with chemical tags under

conditions that imitate those of the live cell. While we examine fluorophores in a reducing

environment and a macromolecularly dense environment, we do not include oxygen scavengers

because these conditions are not representative of the native, live cell. As a result, the measured

photon output of Alexa647 in this study is lower than that of Atto655, which is the opposite of

what has been previously observed with these fluorophores in low oxygen, super resolution SM

imaging conditions.46

While differences in photon output can vary between experimental set-ups,

Alexa647's lower photon output in this investigation is likely attributed to its sensitivity to photo-

oxidative bleaching as a cyanine along with the use of non-deoxygenated buffers.47,48

These

results suggest that Atto655 may be more advantageous for oxygenated conditions in live cell

imaging, but further investigation is needed to account for fluorophore interactions in cell

environments, such as non-specific binding, which may further impact imaging performance.

Ensemble evaluation of fluorophores gives valuable insights regarding the utility of

ensemble properties as indicators of SM imaging performance. As a measure of the probability

of fluorescence emission, quantum yield is not an ensemble equivalent of photon flux, which is

the rate of photon emission from the fluorophore. While photostability lifetime and survival

lifetime are both direct measures of photostability, differences in illumination intensities and

imaging conditions can affect these lifetimes as high illumination causes fluorophores to

experience more dark state transitions from which they are photobleached.43,49,50

However, we do

observe a correlation between the ensemble and SM properties as both Atto655 and Alexa647

experience no change in quantum yield and photon flux with the addition of the A-TMP-tag in

Page 70: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

59

PBS. In addition, the A-TMP-tag lowers both photostability lifetime and survival lifetime for

Atto655 while the tag raises Alexa647 photostability lifetime and survival lifetime. The ability to

measure ensemble properties without specialized equipment and the correlation between the

ensemble and SM properties indicates that these ensemble measurements can be applied for

wider screenings of chemically tagged fluorophores with advantageous properties that can be

later confirmed by SM investigation.

Ensemble measurements also allow separate examination of chemical and environmental

effects from the A-TMP-tag because the properties of fluorophores, A-TMP-fluorophore, and the

tagged-fluorophores can be measured individually. With the oxazines Atto655 and Atto680, we

observe that the A-TMP-oxazines have lower quantum yields, which is consistent with previous

observations of quenching by intramolecular electron transfer after conjugation to electron-rich

molecule in these fluorophores and other oxazines.32,34,51

We also observe that tagged-Atto655

and tagged-Atto680 have quantum yields that are unchanged for the original fluorophores, an

effect that was similarly observed with oxazines and the SNAP-tag, and may be caused by

inhibited electron transfer due to the interaction of A-TMP with the eDHFR binding pocket.32,33

For the cyanine Cy3, we observe stepwise increases in quantum yield from Cy3, to A-TMP-Cy3

and to tagged-Cy3, which is consistent with the tag causing steric inhibition of cis-trans

photoisomerization in the methinine chain, a mechanism of non-radiative decay.52,53

Although a

similar trend in quantum yield is not observed with the cyanine Alexa647, its longer methinine

chain has a lower rate of cis-trans isomerization that likely reduces the steric effects of the A-

TMP-tag on quantum yield.54,55

Examining fluorophores with chemical tags using ensemble

measurements can provide insights to their interactions and predict whether fluorophores in the

same chemical class will behave similarly.

Page 71: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

60

2.6 Conclusion and Outlook

Chemical tags are emerging as a strategic experimental tool for selectively labeling

proteins with organic fluorophores to meet the demand for fluorescent labels with high photon

outputs in SM imaging. As the chemical tags make SM imaging more accessible, there is a

growing need for characterization to not only understand the effects the chemical tags have on

fluorophores, but also identify chemically tagged fluorophores with suitable properties for SM

imaging. We examined both SM and ensemble photophysical properties of the fluorophores,

Atto655 and Alexa647, and their A-TMP-tag conjugates. We demonstrated that the A-TMP-tag

is an effective labeling reagent for SM imaging because it upholds Atto655 and Alexa647 total

photon output. Characterizing these commercially available fluorophores with the A-TMP-tag

provides photophysical benchmarks to compare and guide the selection of fluorophores and

chemical tags. As the capabilities of chemical tags evolve with the developments of new organic

fluorophores and fluorescent materials for biological imaging, this framework can continue to

evaluate and identify the most promising fluorophores and chemical tags for SM imaging. By

using SM imaging to decipher the complex dynamics of protein function and interaction in live

cells, the chemical tags have the potential to revolutionize the study of mechanistic biology.

Page 72: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

61

2.7 NMR Spectra of Synthesized TMP-Fluorophores

Figure 2-8: Alexa647-A-TMP

1H NMR

(400 MHz, Methanol-d4) δ 8.36 (s, 3H), 7.94 – 7.85 (m, 5H), 7.45 (t, 3H), 7.30 (s, 1H), 6.75 (d, 1H), 6.58 (s, 3H),

6.50 (d, 2H), 6.31 – 6.22 (m, 1H), 6.18 (d, 1H), 5.62 (d, 1H), 4.89 – 4.79 (m, 3H), 4.37 (s, 6H), 3.96 (t, 3H), 3.81 (s,

7H), 3.67 (s, 3H), 3.21 (t, 4H), 3.04 – 2.98 (m, 4H), 2.89 (d, 1H), 2.70 (dd, J = 22.5, 6.7 Hz, 2H), 2.47 (s, 1H), 2.29

– 2.15 (m, 8H), 2.04 (d, J = 7.8 Hz, 3H), 1.86 (t, J = 6.1 Hz, 2H), 1.73 (dd, J = 17.4, 5.4 Hz, 16H), 1.63 (s, 2H), 1.48

(s, 3H), 1.34 (d, 12H), 1.22 (s, 3H), 0.92 (d, 1H), 0.63 (s, 1H).

Page 73: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

62

Figure 2-9: Atto655-A-TMP

1H NMR

(400 MHz, Methanol-d4) of. δ 7.83 (dd, J = 1.8, 1.0 Hz, 1H), 7.50 (s, 1H), 7.28 (d, J = 1.0 Hz, 1H), 7.09 (s, 1H),

6.97 (s, 1H), 6.50 (d, J = 0.8 Hz, 2H), 6.32 – 6.12 (m, 2H), 5.63 (ddd, J = 9.9, 2.1, 0.8 Hz, 1H), 4.80 (td, J = 6.7, 1.5

Hz, 1H), 3.97 – 3.87 (m, 2H), 3.77 (s, 11H), 3.70 – 3.53 (m, 14H), 3.52 – 3.35 (m, 6H), 3.24 (td, J = 6.9, 1.4 Hz,

2H), 3.09 – 2.93 (m, 3H), 2.74 (dd, J = 15.0, 6.3 Hz, 1H), 2.67 – 2.51 (m, 2H), 2.43 (q, J = 6.4 Hz, 2H), 2.08 (q, J =

5.3 Hz, 6H), 1.90 – 1.67 (m, 8H), 1.55 (s, 3H), 1.47 – 1.29 (m, 9H).

Page 74: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

63

Figure 2-10: Atto680-A-TMP

1H NMR

(400 MHz, Methanol-d4) δ 7.84 (s, 1H), 7.44 (d, J = 1.3 Hz, 1H), 7.25 (d, J = 0.9 Hz, 1H), 7.04 (s, 1H), 6.92 (s, 1H),

6.47 (s, 2H), 6.31 – 6.11 (m, 2H), 6.03 (s, 1H), 5.62 (dd, J = 9.8, 2.1 Hz, 1H), 4.80 (t, J = 6.7 Hz, 1H), 4.05 (s, 2H),

3.94 (t, J = 5.7 Hz, 2H), 3.77 (s, 14H), 3.69 – 3.53 (m, 12H), 3.24 (td, J = 6.8, 1.9 Hz, 2H), 2.99 – 2.86 (m, 3H), 2.73

(dd, J = 15.0, 6.4 Hz, 1H), 2.62 (dd, J = 15.0, 7.1 Hz, 1H), 2.45 (t, J = 6.6 Hz, 2H), 2.18 (s, 2H), 2.10 – 2.04 (m,

5H), 1.92 – 1.79 (m, 4H), 1.72 (p, J = 6.4 Hz, 2H), 1.61 (s, 7H), 1.32 (s, 13H).

Page 75: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

64

Figure 2-11: Cy3-A-TMP

1H NMR

(400 MHz, Methanol-d4) δ 8.36 (s, 2H), 7.94 – 7.85 (m, 4H), 7.45 (t, 2H), 7.30 (s, 1H), 6.75 (d, 1H), 6.58 (s, 2H),

6.50 (d, 2H), 6.28 (dd,1H), 6.17 (d, 1H), 5.62 (d,1H), 5.37 (t, J = 4.8 Hz, 1H), 4.81 (s, 1H), 4.37 (s, 5H), 3.96 (t,2H),

3.81 (s, 6H), 3.67 (s, 2H), 3.63 – 3.52 (m, 8H), 3.51 – 3.42 (m, 6H), 3.22 (d,5H), 3.02 (d, J = 0.5 Hz, 1H), 2.89 (d, J

= 0.7 Hz, 1H), 2.66 (s, 1H), 2.47 (s, 1H), 2.29 – 2.15 (m, 6H), 2.04 (d,3H), 1.90 – 1.82 (m, 2H), 1.78 – 1.67 (m,

13H), 1.63 (s, 1H), 1.47 (d,2H), 1.39 – 1.29 (m, 10H), 1.22 (s, 3H).

Page 76: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

65

Figure 2-12: Biotin-Alexa647

1H NMR

(400 MHz, Methanol-d4) δ 8.27 (q, 2H), 7.86 – 7.76 (m, 4H), 7.37 (t, 2H), 6.45 – 6.37 (m, 2H), 4.49 – 4.37 (m, 1H),

4.28 (d, 5H), 3.67 – 3.39 (m, 10H), 3.28 (dd, J = 9.5, 5.5 Hz, 3H), 3.13 (dq, J = 9.3, 4.6 Hz, 1H), 2.94 – 2.75 (m,

5H), 2.63 (d, 1H), 2.19 – 2.07 (m, 8H), 1.97 (t, 2H), 1.71 – 1.63 (m, 9H), 1.61 – 1.47 (m, 2H), 1.37 (dq, J = 15.0, 7.2

Hz, 4H), 1.23 (d, 4H), 1.14 (t, 2H).

2.8 References

(1) Moerner, W. E. Proceedings of the National Academy of Sciences 2007, 104,

12596.

(2) Joo, C.; Balci, H.; Ishitsuka, Y.; Buranachai, C.; Ha, T. Annual review of

biochemistry 2008, 77, 51.

(3) Weiss, S. Science 1999, 283, 1676.

(4) Zhuang, X.; Bartley, L. E.; Babcock, H. P.; Russell, R.; Ha, T.; Herschlag, D.;

Chu, S. Science 2000, 288, 2048.

Page 77: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

66

(5) Yildiz, A.; Forkey, J. N.; McKinney, S. A.; Ha, T.; Goldman, Y. E.; Selvin, P. R.

Science 2003, 300, 2061.

(6) Fei, J.; Bronson, J. E.; Hofman, J. M.; Srinivas, R. L.; Wiggins, C. H.; Gonzalez,

R. L. Proceedings of the National Academy of Sciences 2009, 106, 15702.

(7) Lu, H. P.; Xun, L.; Xie, X. S. Science 1998, 282, 1877.

(8) Yu, J.; Xiao, J.; Ren, X.; Lao, K.; Xie, X. S. Science 2006, 311, 1600.

(9) Schuler, B.; Lipman, E. A.; Eaton, W. A. Nature 2002, 419, 743.

(10) Myong, S.; Rasnik, I.; Joo, C.; Lohman, T. M.; Ha, T. Nature 2005, 437, 1321.

(11) Elf, J.; Li, G.-W.; Xie, X. S. Science 2007, 316, 1191.

(12) Funatsu, T.; Harada, Y.; Tokunaga, M.; Saito, K.; Yanagida, T. Nature 1995, 374,

555.

(13) Crawford, D. J.; Hoskins, A. A.; Friedman, L. J.; Gelles, J.; Moore, M. J. RNA

(New York, N.Y.) 2008, 14, 170.

(14) Shen, Z.; Chakraborty, A.; Jain, A.; Giri, S.; Ha, T.; Prasanth, K. V.; Prasanth, S.

G. Molecular and Cellular Biology 2012, 32, 3107.

(15) Jain, A.; Liu, R.; Ramani, B.; Arauz, E.; Ishitsuka, Y.; Ragunathan, K.; Park, J.;

Chen, J.; Xiang, Y. K.; Ha, T. Nature 2011, 473, 484.

(16) Fernandez-Suarez, M.; Ting, A. Y. Nat Rev Mol Cell Biol 2008, 9, 929.

(17) Xia, T.; Li, N.; Fang, X. Annual Review of Physical Chemistry 2013, 64, 459.

(18) Schmidt, T.; Kubitscheck, U.; Rohler, D.; Nienhaus, U. Single Molecules 2002, 3.

(19) Yildiz, A.; Selvin, P. R. Accounts of chemical research 2005, 38, 574.

(20) Sako, Y.; Minoghchi, S.; Yanagida, T. Nat Cell Biol 2000, 2, 168.

(21) Wombacher, R.; Cornish, V. W. Journal of biophotonics 2011, 4, 391.

(22) Jing, C.; Cornish, V. W. Accounts of chemical research 2011, 44, 784.

(23) Miller, L. W.; Cai, Y.; Sheetz, M. P.; Cornish, V. W. Nat Meth 2005, 2, 255.

(24) Keppler, A.; Gendreizig, S.; Gronemeyer, T.; Pick, H.; Vogel, H.; Johnsson, K.

Nat Biotech 2003, 21, 86.

Page 78: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

67

(25) Los, G. V.; Encell, L. P.; McDougall, M. G.; Hartzell, D. D.; Karassina, N.;

Zimprich, C.; Wood, M. G.; Learish, R.; Ohana, R. F.; Urh, M.; Simpson, D.; Mendez, J.;

Zimmerman, K.; Otto, P.; Vidugiris, G.; Zhu, J.; Darzins, A.; Klaubert, D. H.; Bulleit, R. F.;

Wood, K. V. ACS chemical biology 2008, 3, 373.

(26) Jones, S. A.; Shim, S.-H.; He, J.; Zhuang, X. Nat Meth 2011, 8, 499.

(27) Wombacher, R.; Heidbreder, M.; van de Linde, S.; Sheetz, M. P.; Heilemann, M.;

Cornish, V. W.; Sauer, M. Nat Meth 2010, 7, 717.

(28) Wilmes, S.; Staufenbiel, M.; Lisse, D.; Richter, C. P.; Beutel, O.; Busch, K. B.;

Hess, S. T.; Piehler, J. Angewandte Chemie (International ed. in English) 2012, 51, 4868.

(29) Gallagher, S. S.; Sable, J. E.; Sheetz, M. P.; Cornish, V. W. ACS chemical biology

2009, 4, 547.

(30) Chen, Z.; Jing, C.; Gallagher, S. S.; Sheetz, M. P.; Cornish, V. W. Journal of the

American Chemical Society 2012, 134, 13692.

(31) Hoskins, A. A.; Friedman, L. J.; Gallagher, S. S.; Crawford, D. J.; Anderson, E.

G.; Wombacher, R.; Ramirez, N.; Cornish, V. W.; Gelles, J.; Moore, M. J. Science 2011, 331,

1289.

(32) St hr, K.; Siegberg, D.; Ehrhard, T.; peropoulos, .; z, S.; Schul eister, S.;

Pfeifer, A. C.; Bach ann, J.; ling ller, U.; Sourjik, V.; Herten, D.-P. Analytical Chemistry

2010, 82, 8186.

(33) Sun, X.; Zhang, A.; Baker, B.; Sun, L.; Howard, A.; Buswell, J.; Maurel, D.;

Masharina, A.; Johnsson, K.; Noren, C. J.; Xu, M. Q.; Correa, I. R., Jr. Chembiochem : a

European journal of chemical biology 2011, 12, 2217.

(34) Marmé, N.; Knemeyer, J.-P.; Sauer, M.; Wolfrum, J. Bioconjugate Chemistry

2003, 14, 1133.

(35) Chen, H.; Ahsan, S. S.; Santiago-Berrios, M. E. B.; Abruña, H. D.; Webb, W. W.

Journal of the American Chemical Society 2010, 132, 7244.

(36) Abe, R.; Ohashi, H.; Iijima, I.; Ihara, M.; Takagi, H.; Hohsaka, T.; Ueda, H.

Journal of the American Chemical Society 2011, 133, 17386.

(37) Buschmann, V.; Weston, K. D.; Sauer, M. Bioconjugate Chemistry 2002, 14, 195.

(38) Zhang, Z.; Rajagopalan, P. T. R.; Selzer, T.; Benkovic, S. J.; Hammes, G. G.

Proceedings of the National Academy of Sciences of the United States of America 2004, 101,

2764.

Page 79: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

68

(39) Antikainen, N. M.; Smiley, R. D.; Benkovic, S. J.; Hammes, G. G. Biochemistry

2005, 44, 16835.

(40) Miller, G. P.; Benkovic, S. J. Biochemistry 1998, 37, 6327.

(41) Friedman, L. J.; Chung, J.; Gelles, J. Biophysical Journal 2006, 91, 1023.

(42) Williams, A. T. R.; Winfield, S. A.; Miller, J. N. Analyst 1983, 108, 1067.

(43) Tsien, R. Y.; Ernst, L.; Waggoner, A. In Handbook of biological confocal

microscopy; 3rd ed.; Pawley, J. B., Ed.; Springer: New York, 2006, p 338.

(44) Huang, B.; Bates, M.; Zhuang, X. Annual review of biochemistry 2009, 78, 993.

(45) Kasper, R.; Heilemann, M.; Tinnefeld, P.; Sauer, M.; Popp, J., von Bally, G.,

Eds.; Optical Society of America: 2007; Vol. 6633, p 6633_71.

(46) Dempsey, G. T.; Vaughan, J. C.; Chen, K. H.; Bates, M.; Zhuang, X. Nat Meth

2011, 8, 1027.

(47) Dave, R.; Terry, D. S.; Munro, J. B.; Blanchard, S. C. Biophysical Journal 2009,

96, 2371.

(48) Aitken, C. E.; Marshall, R. A.; Puglisi, J. D. Biophysical Journal 2008, 94, 1826.

(49) Donnert, G.; Eggeling, C.; Hell, S. W. Nat Meth 2007, 4, 81.

(50) Humpolickova, J.; Benda, A.; Machan, R.; Enderlein, J.; Hof, M. Physical

Chemistry Chemical Physics 2010, 12, 12457.

(51) Zhu, R.; Li, X.; Zhao, X. S.; Yu, A. The Journal of Physical Chemistry B 2011,

115, 5001.

(52) Aramendia, P. F.; Negri, R. M.; Roman, E. S. The Journal of Physical Chemistry

1994, 98, 3165.

(53) Mishra, A.; Behera, R. K.; Behera, P. K.; Mishra, B. K.; Behera, G. B. Chemical

Reviews 2000, 100, 1973.

(54) Widengren, J.; Schwille, P. The Journal of Physical Chemistry A 2000, 104, 6416.

(55) Chibisov, A. K.; Zakharova, G. V.; Gorner, H. Journal of the Chemical Society,

Faraday Transactions 1996, 92, 4917.

Page 80: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

69

Chapter 3

Oxazine and Xanthene Fluorophores Synthesized from a Common Diaryl Intermediate

A portion of this chapter was published in Anzalone, A. V.; Wang, T. Y.; Chen, Z.; Cornish, V.

W. Angewandte Chemie International Edition 2013, 52, 650.

Page 81: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

70

3.1 Chapter Outlook

The photophysical properties of organic fluorophores used with chemical tags are critical

for achieving live cell SM imaging. However, there are few fluorophores that have the requisite

photon output for reliable SM detection and imaging. Additionally, preferred fluorophores for

SM imaging are difficult to synthesize and, consequently, very expensive. Not only is the

discovery of new fluorophores with attractive photophysical properties beneficial for expanding

the accessibility of SM imaging, but the improved synthetic routes can greatly improve the

efficiency of fluorophore production to reduce the costs. There is a need for reliable synthetic

strategies to produce organic fluorophores with SM imaging capabilities.

In this chapter, a new synthesis of oxazine and xanthene fluorophores with high yield is

presented. This work is capable of gram scale production of water soluble fluorophores with red,

far-red and infrared fluorescence that are compatible with cell imaging. In addition, this work

enables the synthesis of asymmetric oxazine and xanthenes that may have unique photophysical

properties. The robustness of this synthetic route enables further investigation and discovery of

organic fluorophores with ideal photophysical and chemical properties for live cell SM imaging.

Andrew Anzalone and I are the main contributors of this work. Andrew Anzalone designed and

synthesized the fluorophores with help from Zhixing Chen. I performed all photophysical

characterizations of the fluorophores. Andrew Anzalone wrote the paper with contributions from all

co-authors.

3.2 Introduction

Recent advances in fluorescence spectroscopy have driven the demand for dyes with

improved photophysical and fluorescence properties.1-3

In addition to their development as

cellular and single molecule imaging tools, engineered fluorescent dyes have also been

Page 82: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

71

developed as environmental sensors that can provide readouts of local viscosity, pH, solute

concentration, and electrical potential.4-8

Current and future developments in this field, especially

those relevant to single-molecule and cellular imaging, depend on the synthesis of customized

fluorescent dyes that emit in the red region of the visible spectrum, have high extinction

coefficients and quantum yields, and display high photostability.9,10

Amongst our best dyes for these purposes are those from the oxazine and xanthene

classes, exemplified by commercial compounds ATTO-655 and Alexa Fluor-594,

respectively.9,11,12

In the synthesis of derivatized or customized fluorophores, modification of

commercially available dyes is often limited due either to cumbersome functionalization of the

parent compounds or to prohibitive costs in obtaining sufficient quantities of dye for carrying out

the necessary synthetic steps. Thus, the de novo synthesis of fluorescent dyes from basic organic

building blocks is an essential aspect of technology development. Despite the obvious

importance of these molecules and the evident need for improved synthetic methodologies,

oxazines and xanthenes are still largely synthesized using methods reported decades or more ago

that do not take advantage of the efficiencies of modern chemical transformations.13-15

Here, we report a novel and scalable synthetic approach to the assembly of the widely

used oxazine and xanthene fluorophores through a common diaryl ether intermediate (Figure 3-

1). Taking advantage of recent developments in transition metal catalysis, we constructed

electronically activated diaryl ethers to serve as tethered nucleophiles, reacting with a range of

substrates to undergo cyclization to cationic fluorescent compounds (Figure 3-1, our work). Final

products were provided in good overall yields, were amenable to purification using standard

silica-based normal phase flash chromatography, and, significantly, could be prepared on the

gram scale.

Page 83: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

72

Figure 3-1: Retrosynthetic analysis for oxazine and xanthene fluorophores

Differences in approach between prior works and our work are highlighted. G=leaving group, X=H or RCO2−.

3.3 Results

As part of our ongoing research program to develop improved technologies for in vivo

super resolution imaging, we sought to prepare derivatives of the commonly utilized oxazine

ATTO-655.12

Due to the expense associated with obtaining large quantities of this dye and the

lack of commercial availability of other desirable analogs, we set out to synthesize the oxazines

following previously described methods.14,15

These prior works rely on coupling a pair of

aminophenol derivatives, one of which is substituted with an electrophilic nitroso or diazo

functionality, by heating the two components in an acidic medium (Figure 3-1, prior works).

Though the aminophenol intermediates are readily prepared, the final coupling and cyclization

reactions frequently result in low yields (~15%) and necessitate the use of preparative scale

reverse phase high performance liquid chromatography (prep-HPLC) in order to obtain material

of satisfactory purity. In our hands, the quantities obtained by this method proved insufficient to

reasonably carry out the remaining steps in the synthesis of our final targets, and we concluded

that scaling up the process to obtain the desired quantities was impractical. As a result, we

Page 84: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

73

concluded that the existing approach was not a viable synthetic route and chose to pursue

alternative synthetic strategies.

Within the oxazine core structure, we identified a previously unexplored diaryl ether

disconnection, which is exploited in the restrosynthetic analysis shown in Figure 3-1. We

envisioned that a step-wise coupling and cyclization process would proceed in higher overall

yield when compared to the classical coupling strategy and, more importantly, would yield fewer

byproducts in the final step. Thus, isolation and purification could be carried out using standard

flash chromatography techniques amenable to gram-scale operations. In pursuit of this synthetic

strategy, it was necessary to examine methods for synthesizing the diaryl ether in a versatile and

robust fashion.

In addition to the more traditional dihydroquinoline and tetrahydroquinoline derivatives,

we also chose to explore an indoline-based scaffold to study this synthetic sequence. The

precursor indoles provide a large number of building blocks from which to start, and we were

also interested in the photophysical properties of the resulting oxazine products, which are not

well described for the indoline class. Preparation of the indoline coupling partners proved to be

straightforward by Gribble reduction and alkylation of the commercially available indole starting

materials.16

The Ullmann ether synthesis poses a potential route to the diaryl ether structural motif by

copper(I) promoted reaction between phenols and aryl halides.17,18

Although attractive in theory,

the classical reaction conditions typically employ strong base, stoichiometric quantities of

copper, and heating at temperatures in excess of 200°C. Additionally, the reaction is known to be

highly substrate dependent and, when successful, commonly results in only modest yields. To

overcome these limitations and eliminate the harsh reaction conditions, we chose to explore a

Page 85: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

74

more contemporary method that utilizes palladium catalysis.19

Unexpectedly, coupling between

phenol 1c (Table 3-1) and its corresponding triflate resulted predominantly in reaction at C-5 of

the indoline in a Heck-type manner, producing the biaryl phenol derivative as the major product.

Thus, the palladium catalyzed coupling reaction does not appear to be a useful method for

synthesizing diaryl ethers when electron rich carbon nucleophiles, such as those that exist in our

system, are present.

We next turned to recent developments in ligand-assisted copper catalyzed coupling

reactions.20

Based on work reported by Buchwald and coworkers, we found that the coupling

between phenol 1b and 3-iodoaniline 2a (Table 3-1) provided the diaryl ether 3b in 82% yield

when carried out in the presence of catalytic copper iodide (10 mol %) and 2-picolinic acid (20

mol %) at 85°C in DMSO (Table 3-1, entry 2). This catalytic system offers the advantage of

orthogonality with the aniline functional group. As opposed to aryl iodides, aryl bromides were

found to be very poor substrates in our system, with nearly no product formation observed even

at elevated temperatures (180°C) and increased catalyst loadings (50 mol %). Nonetheless, aryl

iodides were readily prepared from the corresponding commercially available aryl bromides by

employing a copper promoted halogen exchange reaction.21

Couplings carried out according to

the aforementioned conditions provided a range of substituted diaryl ethers in good yields (Table

3-1). Following these coupling reactions, anilines 3a and 3b were transformed to their

dihydroquinoline derivatives via a modified Skraup reaction and subsequently alkylated and

reduced where appropriate.

Page 86: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

75

Table 3-1: Copper(I)-catalyzed couplings between phenols and aryl iodides to furnish

diaryl ethers.[a]

Entry Phenol Aryl iodide Diaryl ether Yield [%]

1

1a

2a

3a 90

2

1b

2a

3b

82

3

1c

2b

3c 88

4

1c

2c

3d 80

5

1c

2d

3e

77

6

1c

2e

3f

80

[a] Conditions: phenol (1.2 equiv), copper(I) iodide (10 ol %), 2-picolinic acid (20 ol %), 3PO4 (2.0 equiv),

DMSO, 85°C, 24 h. DMSO=dimethyl sulfoxide.

Having established a reliable means of synthesizing the critical diaryl ether intermediates,

we next explored conditions for converting these compounds to their corresponding oxazine

dyes. This transformation was readily accomplished by reaction of diaryl ether 3d with one

Page 87: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

76

equivalent of 4-nitrobenzene diazonium tetrafluoroborate, followed by heating the corresponding

diazene 4d with p-toluenesulfonic acid (TsOH) to 65°C in ethanol (Figure 3-2). The latter

process proceeded with near-quantitative conversion and following trivial silica based flash

chromatography to remove the liberated p-nitroaniline, we obtained the oxazine-tosylate salt in

high purity and 90% yield. Implementing this strategy, several substituted oxazine dyes with

various spectroscopic properties were synthesized in good to excellent yields from the

corresponding diazene compounds (Table 3-2). Of note, over one gram of oxazine 6f was

prepared by this method in 94% yield, demonstrating the scalability of our approach (Table 3-2,

entry 6). Additionally, 6g, the sulfonated analog of 6c resembling commercially available

ATTO-655 was synthesized in good yield by this method.

Figure 3-2: Reaction sequence for conversion of diaryl ethers to oxazine dyes.

Page 88: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

77

Table 3-2: Synthesis of substituted oxazine dyes.[a]

Entry Diazene[b,c]

Oxazine Yield [%]

1

4a

6a

74

2

4b/5b

6b

85

3

4c/5c

6c

87

4

4d

6d

90

5

4e/5e

6e

78

6[d]

4f/5f

6f

94

7

4g/5g

6g

83

[a] Conditions: TsOH (3.0 equiv), EtOH, 65°C, 4–8 h.

[b] Ar=p-NO2Ph.

[c] Diazenes were obtained and used as

regioisomeric mixtures to cyclize to a single oxazine product. [d]

Reaction conducted on a gram-scale.

Page 89: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

78

To further expand the scope of this methodology, we examined other popular fluorescent

dyes for similar diaryl ether disconnections and identified xanthene dyes, exemplified by

rhodamine and rosamine, as potential targets. The classical method for synthesizing these

aminoxanthenes calls for heating two equivalents of an aminophenol with one equivalent of a

carboxylic acid anhydride or aldehyde under acidic conditions (ZnCl2 or H2SO4) often at high

temperatures (Figure 3-1, prior works).13

When this strategy is employed with derivatized

analogs, the reactions typically result in modest to poor yields and make isolation of the product

difficult by conventional purification methods.22

We sought to prepare these fluorescent dyes

from our diaryl ethers in an analogous manner to the synthesis of the oxazine dyes. After

screening several Lewis acid catalysts and reaction conditions, we found that Ga(OTf)3 catalyzed

the tandem Friedel-Crafts acylationcyclization reaction between the aforementioned diaryl ethers

and aromatic acid chlorides to provide the corresponding xanthene dyes in modest to good yields

with substantial recoverable starting material (Table 3-3).23,24

Page 90: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

79

Table 3-3: Tandem catalytic Friedel–Crafts acylation/cylization reaction for the synthesis

of xanthene fluorophores. [a]

Entry R1

R2 Prod. Yield [%]

[b]

1 Et

7 a 56 (92)

2[c,e]

7 b 83

3[d]

Et

7 c 77 (82)

4 Et

7 d 48 (89)

5 Et

7 e 53 (93)

6[d]

Et

7 f 36 (81)

[a] Conditions: acid chloride (8 equiv), Ga(OTf)3 (15 ol %), MeNO2, 60 C, 4 Å olecular sieves (M.S.), 16 h.

[b]

Yields in brackets based on recovered starting material. [c]

No Lewis acid catalyst required, TFAA (2.2 equiv) in

CH2Cl2 at RT for 12 h. [d]

Acid chloride (5 equiv) was used. [e]

Reaction conducted on a gram-scale.

Supplementing the reaction with additional Ga(OTf)3 did not lead to further product

formation, suggesting that a byproduct formed in the course of the reaction was inhibiting

forward progression. Possibly, this was due to substrate inactivation by HCl, which is

generated in the course of the Friedel-Crafts acylation and from hydrolysis of the acid

chloride (water is generated in the cyclization step). Interestingly, in the case of the p-Me2N

derivative (Table 3-3, entry 3), nearly full consumption of starting material was observed,

Page 91: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

80

suggesting that the dimethylamino group may be capable of buffering the reaction to allow

for increased conversion. Higher conversions of starting material could be obtained by

resubjecting the crude product mixture (following workup) to the initial reaction conditions.

Of note, this synthetic strategy allows for the synthesis of asymmetrically functionalized

xanthene dyes, which are not accessible by the classical coupling strategy. Additionally, several

dyes were prepared in significantly higher yields when compared to traditional syntheses. For

example, compound 7b, an asymmetric xanthene, was synthesized in 83% yield simply by

reaction of the diaryl ether with 2.2 equivalents of trifluoroacetic anhydride at room temperature

without a Lewis acid catalyst. By comparison, rhodamine 700, a similar dye possessing the

trifluoromethyl substituent at the meso carbon, is reported to be prepared in only 5% yield from

aminophenols.25

Although this current work is limited to the synthesis of rosamines with fully

substituted anilines, it may be possible to synthesize mono- or non-alkylated analogs by minor

modification of this synthetic approach.

Figure 3-3 shows the absorption and fluorescence spectra for several of the oxazine and

xanthene dyes synthesized. Notably, the increased conjugation of oxazines 6a and 6b results

in significant red-shifted absorption and fluorescence. Similarly, xanthenes 7b, 7e and 7f

possessing electron withdrawing side chains also display red-shifted spectra. The

dimethylamino-substituted rosamine 7c displays pH dependent fluorescence, with a near 30-

fold increase in fluorescence intensity in acidic conditions (data not shown). Spectral

properties of the dyes in aqueous solution are provided in Table 3-4. In summary, we have

established a high-yielding, scalable synthetic route for the preparation of the widely used

oxazine and xanthene based fluorescent dyes from a common diaryl ether intermediate.

Compared to previously existing synthetic methodologies, our work provides a versatile means

Page 92: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

81

for preparing these fluorescent dyes and eliminates the need for tedious and expensive

purifications. Following this synthetic approach, a number of oxazine and xanthene fluorophores

were synthesized and characterized. With proper synthetic planning, we believe this to be a

general and widely applicable approach to the synthesis of derivatized oxazine and xanthene

dyes, and may facilitate the development of novel fluorophores and probes with unique

properties.

Figure 3-3: Absorbance and fluorescence spectra of oxazine and xanthene derivatives

a) Absorbance and b) fluorescence spectra of various oxazine and xanthene derivatives. Spectra were obtained in

H2O, with the exception of 7c, which was obtained in aqueous HCl (50 mM).

Page 93: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

82

Table 3-4: Spectral properties of fluorescent dyes in H2O [a]

Dye λ ax abs [nm] εmax [M−1 cm

−1] λfluor [nm] Fwhm

[b] [nm] Φf

6 a 703 50, 000 717 42 0.08

6 b 682 69, 000 696 44 0.09

6 c 664 67, 000 681 45 0.24

6 d 648 66, 000 661 43 0.11

6 g 663 97, 000 677 43 0.20

7 a 553 58, 000 576 49 0.19

7 b 616 61, 000 643 90 0.08

7 c 559 79, 000 585 53 0.14

7 d 555 60, 000 575 47 0.21

7 e 579 68, 000 602 44 0.27

7 f 581 57, 000 600 42 0.24 [a]

Measurements were taken in H2O, with the exception of 7c, which was measured in aqueous HCl (50 mM).

[b] Full-width at half-maximum height.

3.4 Conclusion and Outlook

The efficient and scalable synthesis of oxazine and xanthene fluorophores can

significantly improve the availability of fluorophores for SM imaging. High yield and purity can

reduce production costs of currently employed SM imaging fluorophores, such as Atto655. At

the same time, novel fluorophores with favorable SM imaging properties can be identified. The

fluorophores synthesized in this investigation exhibited high extinction coefficients and red to

infrared fluorescence in aqueous solutions, which are all desirable characteristics for live cell SM

imaging. Due to the relationship between SM photon output and quantum yield with extinction

coefficients established in the first chapter, these fluorophores exhibit promising potential for

good SM imaging performance. Further investigation into the SM properties of fluorophores

from this study and novel fluorophores may reveal outstanding dyes for live cell imaging.

Page 94: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

83

3.5 Supporting Information

3.5.1 General Experimental Methods

All reactions were carried out under an argon atmosphere. Anhydrous 1,4-dioxane,

dimethylsulfoxide, tetrahydrofuran, acetonitrile and dichloromethane were purchased as Sure

Seal™ bottles fro Sig a-Aldridge. Potassium phosphate tribasic was purchased from Sigma-

Aldrich and ground to a fine powder just prior to use. Sodium iodide powder was purchased

from Alfa Aesar and ground to a fine powder just prior to use. Copper(I) iodide (99.999%)

powder was purchased from Sigma-Aldrich. 6-hydroxyindole and 6-bromoindole were

purchased from Chem-Impex International. Gallium(III) trifluoromethanesulfonate was obtained

from Acros. All other reagents were commercially available and used without further

purification. Flash chromatography was performed using a Teledyne ISCO CombiFlash Rf

system. NMR spectra were recorded on Bruker DRX-300 and DRX-400 instruments. The

following abbreviations were used to describe multiplicities: s = singlet; d = doublet; t = triplet; q

= quartet; m = multiplet; br = broad. High-resolution mass spectrometry (HRMS) was performed

by the Columbia University Mass Spectroscopy Core Facility with a JOEL HX110 mass

spectrometer by means of fast atom bombardment (FAB). Absorbance measurements and spectra

were obtained with a Tecan Infinite 200 and fluorescence measurements and spectra were

obtained with a Horiba Scientific Fluorolog-3 spectrofluorometer. Measurements were

performed in quartz cuvettes with a 1-cm path length using solutions with absorbance under 0.1

to prevent inner filter and other non-linear effects.

3.5.2 Preparation of Hydroxyindolines and Iodoindolines

1,2,3,4-tetrahydroquinolin-7-ol (S1). 7-hydroxy-3,4-dihydroquinolin-2(1H)-one (4.90

g, 30.0 mmol) in a slurry of THF (125 mL) was cooled in an ice bath and treated with lithium

Page 95: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

84

aluminum hydride (1.82 g, 1.60 mmol). After complete addition, the reaction was heated to

reflux overnight. Upon cooling to room temperature, the reaction was quenched by the addition

of aqueous saturated ammonium chloride (50 mL) and extracted with ethyl acetate (3 x 40 mL).

The combined organic layers were washed with brine and dried over anhydrous Na2SO4, then

concentrated to provide S1 in 4.34 g as an orange solid (98%). This compound has been

previously characterized.15

1H NMR (400 MHz, CDCl3) δ 6.81 (d, J = 8.1 Hz, 1H), 6.12 (dd, J =

8.1, 2.5 Hz, 1H), 6.01 (d, J = 2.4 Hz, 1H), 3.46 (s, 1H), 3.29 (t, J = 5.7, 2H), 2.70 (t, J = 6.4 Hz,

2H), 1.93 (td, J = 11.4, 6.3 Hz, 2H).

1-ethyl-1,2,3,4-tetrahydroquinolin-7-ol (1b). A solution of S1 (1.00 g, 6.70 mmol) in

glacial acetic acid (25 mL) was treated with sodium borohydride (1.01 g, 26.8 mmol). After

stirring at room temperature for 2 hours, a TLC of the reaction indicated approximately 50%

consumption of starting material. At this time, acetaldehyde was added in small portions

(reaction T C’d after each addition) until consu ption of starting material was achieved. The

reaction was then concentrated under reduced pressure to a thick residue, which was diluted with

60 mL of saturated aqueous NaHCO3 and neutralized with solid NaHCO3. The resulting solution

was extracted with ethyl acetate (3 x 35 mL). The combined organic layers were washed with

brine (30 mL) and dried over anhydrous Na2SO4, then concentrated to a crude orange oil.

Purification by flash chromatography (5-20% ethyl acetate/hexanes) provided 1.09 g (92%) of

1b as a light orange crystalline solid. 1H NMR (400 MHz, CDCl3) δ 6.82 (d, J = 8.0 Hz, 1H),

6.24 (s, 1H), 6.12 (d, J = 8.0 Hz, 1H), 3.33 (q, J = 7.1 Hz, 2H), 3.28 (t, J = 5.7 Hz, 2H), 2.70 (t, J

= 6.4 Hz, 2H), 2.02 – 1.91 (m, 2H), 1.18 (t, J = 7.1 Hz, 3H). 13C NMR (300 MHz, CDCl3) δ

155.41, 146.33, 130.27, 115.73, 103.21, 98.89, 48.64, 45.92, 27.80, 22.84, 10.94. HRMS (FAB)

Calcd for C11H15NO+ [M]

+: 177.1154; found 177.1162.

Page 96: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

85

1-ethylindolin-6-ol (1c). 6-hydroxyindole (2.66 g, 20.0 mmol) was dissolved in glacial

acetic acid (65 mL) and treated with sodium borohydride (3.78 g, 100 mmol) in small portions at

room temperature. After 4 hours, the reaction was concentrated in vacuo to a thick residue,

which was then diluted with 60 mL of saturated aqueous NaHCO3 and neutralized with solid

NaHCO3. The resulting solution was extracted with ethyl acetate (35 mL, 3x). The combined

organic layers were washed with brine (30 mL) and dried over anhydrous Na2SO4, then

concentrated in vacuo and dried under high vacuum to provide an orange crude solid.

Purification by flash chromatography (5-20% ethyl acetate/hexanes) provided 2.09 g (64%) of 1c

as a tan solid. 1H NMR (400 MHz, CDCl3) δ 6.90 (d, J = 7.8 Hz, 1H), 6.10 (dd, J = 7.8, 2.1 Hz,

1H), 6.02 (d, J = 2.1 Hz, 1H), 4.87 (br s, 1H), 3.37 (t, J = 8.2 Hz, 2H), 3.12 (q, J = 7.2 Hz, 2H),

2.90 (t, J = 8.2 Hz, 2H), 1.20 (t, J = 7.2 Hz, 3H). 13C NMR (75 MHz, CDCl3) δ 156.28, 153.81,

125.12, 122.92, 104.95, 96.93, 53.32, 43.78, 28.14, 11.98. HRMS Calcd. for C10H13NO+ [M

+]:

163.0997; found 163.1007.

6-iodoindole (S2). According to the published protocol by Klapars and Buchwald, 21

an

oven dried, sealable glass tube was charged with a magnetic stirbar, 6-bromoindole (4.94 g, 25.2

mmol), freshly ground sodium iodide (7.55 g, 50.4 mmol), and copper(I) iodide (480 mg, 2.52

mmol). The vessel was then fitted with a rubber septum, evacuated under vacuum and backfilled

with argon. This process was repeated three times. The vessel was then charged with 1,4-dioxane

(25 ) followed b N,N’-dimethylethylenediamine (0.58 mL, 5.40 mmol) via syringe. The

rubber septum was removed and the reaction vessel immediately sealed tightly with a Teflon

screw cap and heated to 110°C for 22 hours. After cooling to room temperature, the reaction was

diluted with saturated aqueous NH4Cl (30 mL) and extracted with dichloromethane (4 x 25 mL).

The combined organic layers were washed with brine (30 mL) and dried over Na2SO4, then

Page 97: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

86

concentrated to a brown residue. The residue was triturated in hexanes and concentrated to

provide S2 (5.87 g, 99%) as a brown crystalline solid. This compound has been previously

characterized. 1H NMR (400 MHz, CDCl3) δ 8.18 (br s, 1H), 7.80 (s, 1H), 7.43 (s, 2H), 7.21 –

7.13 (t, J = 2.8 Hz, 2H), 6.56 (t, J = 2.2 Hz, 1H).

6-iodoindoline (2b). S2 (5.86 g, 24.1 mmol) was dissolved in glacial acetic acid (100

mL) and cooled in an ice bath just until the solution began to become partially frozen. At this

point, the solution was treated portion-wise with sodium cyanoborohydride (4.52 g, 72.3 mmol),

then allowed to warm to room temperature and stir for 3 hours. The reaction mixture was

concentrated to a thick residue, then diluted with 50 mL of saturated aqueous NaHCO3. The

resulting solution was extracted with ethyl acetate (3 x 40 mL). The combined organic layers

were washed with saturated aqueous NaHCO3 (30 mL) and brine (30 mL), then dried over

anhydrous Na2SO4 and concentrated to a brown residue. Purification by flash chromatography

(0-15% ethyl acetate/hexanes) afforded 2b (4.78 g, 81%) as a white crystalline solid. This

compound has been previously characterized. 26

1H NMR (400 MHz, CDCl3) δ 7.10-6.95 (m,

2H), 6.87 (d, J = 7.5 Hz, 1H), 3.59 (td, J = 8.4, 2.9 Hz, 2H), 3.01 (t, J = 8.4 Hz, 2H).

1-ethyl-6-iodoindoline (2c). S2 (729 mg, 3.00 mmol) was dissolved in glacial acetic acid

(15 mL) and treated with sodium borohydride (681 mg, 18.0 mmol) in small portions at room

temperature. After stirring for 1 hour, the reaction was treated with sodium cyanoborohydride

(377 mg, 6.00 mmol). The reaction was concentrated in vacuo to a thick residue, which was then

diluted with 25 mL of saturated aqueous NaHCO3. The resulting solution was extracted with

ethyl acetate (25 mL, 3x). The combined organic layers were washed with saturated aqueous

NaHCO3 and brine (30 mL), then dried over anhydrous Na2SO4 and concentrated in vacuo to a

brown residue. Purification by flash chromatography (0-10% ethyl acetate/hexanes) afforded 2c

Page 98: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

87

(590 mg, 76%) as a light yellow oil. 1H NMR (400 MHz, CDCl3) δ 6.97 (dd, J = 7.6, 1.1 Hz,

1H), 6.80 (d, J = 7.6 Hz, 1H), 6.78 (s, 1H), 3.38 (t, J = 8.2 Hz, 2H), 3.14 (q, J = 7.2 Hz, 2H),

2.93 (t, J = 8.2 Hz, 2H), 1.19 (t, J = 7.2 Hz, 3H). 13

C NMR (300 MHz, CDCl3) δ 154.13, 130.53,

126.44, 126.27, 116.06, 92.76, 52.52, 42.98, 28.49, 12.15. HRMS (FAB) Calcd for C10H12IN+

[M]+: 273.0014; found 273.0011.

1-(4-azidobutyl)-6-iodoindoline (2d). 2b (294 mg, 1.20 mmol) was dissolved in DMF

and treated with 1-azido- 4-iodobutane (405 mg, 1.80 mmol) and sodium carbonate (382 mg,

3.60 mmol). The reaction was heated to 60°C for 8 hours. After cooling to room temperature, the

reaction was diluted with water (25 mL) and extracted with ethyl acetate (3 x 25 mL). The

combined organic layers were washed with brine (30 mL) and dried over anhydrous Na2SO4,

then concentrated to a light brown oil. Purification by flash chromatography (0-5% ethyl

acetate/hexanes) provided 326 mg (80%) of 2d as a colorless oil. 1H NMR (400 MHz, CDCl3) δ

6.98 (dd, J = 7.6, 1.5 Hz, 1H), 6.81 (d, J = 7.6 Hz, 1H), 6.76 (d, J = 1.4 Hz, 1H), 3.49 – 3.29 (m,

4H), 3.08 (t, J = 6.8 Hz, 2H), 2.95 (t, J = 8.2 Hz, 2H), 1.77 – 1.65 (m, 4H). 13

C NMR (75 MHz,

CDCl3) δ 154.39, 130.26, 126.58, 126.37, 115.77, 92.81, 53.39, 51.70, 48.71, 28.57, 27.01,

25.02. HRMS (FAB) Calcd. For C12H15IN4+ [M

+]: 342.0341; found 342.0328.

ethyl 5-(6-iodoindolin-1-yl)pentanoate (2e). 2b (2.29 g, 9.34 mmol) was dissolved in

DMF (15 mL) and treated with diisopropylethylamine (2.44 mL, 14.0 mmol), ethyl 5-

bromovalerate (2.02 mL, 12.6 mmol) and potassium iodide (1.86 g, 11.2 mmol). The reaction

was heated to 60 °C for 12 hours. After cooling to room temperature, the reaction was diluted

with water (30 mL) and concentrated aqueous NH4Cl (10 mL), then extracted with ethyl acetate

(3 x 30 mL). The combined organic layers were dried over anhydrous Na2SO4 and concentrated

to a crude brown oil. Purification by flash chromatography (0-10% ethyl acetate/hexanes)

Page 99: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

88

provided 2e (3.10 g, 89%) as a beige oil. 1H NMR (400 MHz, CDCl3) δ 6.95 (dd, J = 7.6, 1.5 Hz,

1H), 6.79 (d, J = 7.6 Hz, 1H), 6.74 (d, J = 1.3 Hz, 1H), 4.17 (q, J = 7.1 Hz, 2H), 3.38 (t, J = 8.3

Hz, 2H), 3.06 (t, J = 7.2 Hz, 2H), 2.93 (t, J = 8.3 Hz, 2H), 2.39 (t, J = 7.2 Hz, 2H), 1.79 – 1.70

(m, 2H), 1.69-1.60 (m, 2H), 1.30 (t, J = 7.1 Hz, 3H). 13

C NMR (300 MHz, CDCl3) δ 173.79,

154.36, 130.26, 126.48, 126.27, 115.81, 92.74, 60.72, 53.32, 48.80, 34.42, 28.52, 27.08, 22.96,

14.68. HRMS (FAB) Calcd. For C15H20INO2 + [M

+]: 373.0539; found 373.0553.

Figure 3-4: Scheme of palladium catalyzed coupling between a phenol and aryl triflate.

Conditions: 1.2 eq.phenol, 2 mol% Pd(OAc)2, 3 mol% 2-(di-tert-butylphosphino)biphenyl, 2 eq. K3PO4, toluene,

100°C.

Product isolated in Figure 3-4. Structure was confirmed by 1H NMR and mass

spectrometry. The 1H NMR in CDCl3 shows an asymmetrical product with the expected

multiplicities for the aromatic protons (3 singlets, 2 doublets). An additional singlet was present

at 6.50 ppm (phenol –OH). The corresponding singlet is not present in the spectrum when

obtained in MeOD. Mass spectrometry (ESI) showed a correct mass of 309.5 (M+1

). 1H NMR

(400 MHz, CDCl3) δ 7.15 (d, J = 7.3 Hz, 1H), 6.97 (s, 1H), 6.69 (d, J = 7.1 Hz, 1H), 6.50 (s,

1H), 6.18 (s, 1H), 5.55 (s, 1H), 3.47-3.37 (m, 4H), 3.24 – 3.14 (m, 4H), 3.02 (t, J = 8.2 Hz, 2H),

2.95 (t, J = 8.1 Hz, 2H), 1.27-1.18 ( , 6H). 1H NMR (400 MHz, MeOD) δ 7.05 (d, J = 7.5 Hz,

1H), 6.93 (s, 1H), 6.78 (d, J = 7.5, 1H), 6.71 (s, 1H), 6.13 (s, 1H), 3.30 (td, J = 8.1, 1.8 Hz, 4H),

3.12 (dq, J = 12.0, 7.2 Hz, 4H), 2.92 (t, J = 8.1 Hz, 2H), 2.86 (t, J = 8.0 Hz, 2H), 1.20 (t, J = 7.2

Hz, 6H).

Page 100: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

89

3.5.3 Coupling of Phenols and Aryl Iodides.

According to the published protocol by Maiti and Buchwald,20

an oven dried, sealable

glass vessel was charged with a magnetic stirbar, the phenol (2.40 mmol), potassium phosphate

(4.00 mmol, 849 mg), copper(I) iodide (0.20 mmol, 38 mg), 2-picolinic acid (0.40 mmol, 49

mg), and the aryl iodide, if a solid (2.00 mmol). The vessel was then fitted with a rubber septum,

evacuated under vacuum and backfilled with argon. This process was repeated 3 times. The

vessel was then charged with DMSO (4.0 mL), or if the aryl iodide is a liquid, the vessel was

charged with the aryl iodide as a solution in DMSO. The rubber septum was removed and the

reaction vessel was immediately sealed tightly with a Teflon screw cap. The reaction was then

heated to 85°C for 16-24 hours. After cooling to room temperature, the reaction was diluted with

10 mL of water and extracted with ethyl acetate (25 mL, 4x). The combined organic layers were

washed with brine and dried over Na2SO4, then concentrated in vacuo to a crude residue.

Purification by flash chromatography (hexanes/ethyl acetate) afforded the diaryl ethers as

colorless oils, which were stored at -20°C under inert atmosphere (these compounds become

colored upon prolonged exposure to air at room temperature or storage as a solution in a

halogenated solvent, such as chloroform).

3,3'-oxydianiline (3a). Following the general procedure, 3-aminophenol (1a, 1.32 g, 12.0

mmol) and 3-iodoaniline (2a, 2.18 g, 10.0 mmol) were coupled to provide 3a (1.80 g, 90%) as a

dense tan crystalline solid. This compound has been previously characterized.20

1H NMR (400

MHz, CDCl3) δ 7.08 (t, J = 8.0 Hz, 2H), 6.42 (m, 4H), 6.34 (s, 2H), 3.67 (br s, 4H).

3-(1-ethyl-1,2,3,4-tetrahydroquinolin-7-yloxy)aniline (3b). Following the general

procedure, phenol 1b (1.20 g, 12.0 mmol) and 3-iodoaniline (2a, 1.23 g, 10.0 mmol) were

coupled to provide 3b (1.24 g, 82%) as a colorless oil. 1H NMR (400 MHz, CDCl3) δ 7.13 (t, J =

8.0 Hz, 1H), 6.94 (d, J = 8.0 Hz, 1H), 6.49 (ddd, J = 8.2, 2.2, 0.9 Hz, 1H), 6.44-6.38 (m, 3H),

Page 101: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

90

6.29 (dd, J = 8.0, 2.2 Hz, 1H), 3.70 (s, 2H), 3.39-3.29 (m, 4H), 2.80 (t, J = 6.4 Hz, 2H), 2.03 (dq,

J = 9.0, 6.2 Hz, 2H), 1.18 (t, J = 7.1 Hz, 3H). 13C NMR (300 MHz, CDCl3) δ 159.72, 156.50,

148.38, 146.53, 130.54, 130.12, 118.25, 109.77, 108.63, 106.63, 105.21, 102.69, 48.64, 45.84,

28.05, 22.80, 11.18. HRMS (FAB+) Calcd. For C17H20N2O

+ [M

+]: 268.1576; found 268.1583.

1-ethyl-6-(indolin-6-yloxy)indoline (3c). Following the general procedure, phenol 1c

(196 mg, 1.20 mmol) and aryl iodide 2b (245 mg, 1.00 mmol) were coupled to provide 3c (251

mg, 90%) as a colorless oil. 1H NMR (400 MHz, CDCl3) δ 7.05 (d, J = 7.9 Hz, 1H), 7.00 (d, J =

7.9 Hz, 1H), 6.40 (dd, J = 7.9, 2.2 Hz, 1H), 6.36 (d, J = 2.1 Hz, 1H), 6.31 (dd, J = 7.8, 2.2 Hz,

1H), 6.23 (d, J = 2.1 Hz, 1H), 4.32 (br s, 1H), 3.61 (t, J = 8.3 Hz, 2H), 3.41 (t, J = 8.3 Hz, 2H),

3.12 (q, J = 7.2 Hz, 2H), 3.03 (t, J = 8.2 Hz, 2H), 2.96 (t, J = 8.0 Hz, 2H), 1.20 (t, J = 7.2 Hz,

3H). 13C NMR (300 MHz, CDCl3) δ 158.29, 157.99, 154.12, 153.15, 125.23, 124.91, 109.08,

108.04, 104.17, 101.00, 99.64, 96.18, 53.19, 48.36, 43.37, 29.53, 28.32, 12.20. HRMS (FAB)

Calcd. For C18H20N2O+ [M

+]: 280.1576; found 280.1579.

6,6'-oxybis(1-ethylindoline) (3d). Following the general procedure, phenol 1c (412 mg,

2.53 mmol) and aryl iodide 2c (575 mg, 2.10 mmol) were coupled to provide 3d (520 mg, 80%)

as a clear colorless oil. 1H NMR (400 MHz, CDCl3) δ 6.98 (d, J = 7.8 Hz, 1H), 6.29 (d, J = 7.8

Hz, 1H), 6.22 (s, 1H), 3.40 (t, J = 8.2 Hz, 2H), 3.11 (q, J = 7.2 Hz, 2H), 2.95 (t, J = 8.2 Hz, 2),

1.18 (t, J = 7.2 Hz, 3H). 13

C NMR (75 MHz, CDCl3) δ 158.26, 154.02, 125.10, 124.85, 107.64,

99.27, 53.17, 43.33, 28.31, 12.19. HRMS (FAB) Calcd. For C20H24N2O+ [M

+]: 308.1889; found

308.1897.

1-(4-azidobutyl)-6-(1-ethylindolin-6-yloxy)indoline (3e). Following the general

procedure, the phenol 1c (166 mg, 1.02 mmol) and the aryl iodide 2d (290 mg, 0.848 mmol)

were coupled to provide 3e (245 mg, 77%) as a colorless oil. 1H NMR (400 MHz, CDCl3) δ 7.01

Page 102: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

91

(d, J = 3.5 Hz, 1H), 6.99 (d, J = 3.5 Hz, 1H), 6.32 (d, J = 2.3 Hz, 1H), 6.30 (t, J = 2.4 Hz, 1H),

6.24 (d, J = 2.1 Hz, 1H), 6.22 (d, J = 2.1 Hz, 1H), 3.43 (t, J = 8.2 Hz, 4H), 3.39 – 3.33 (m, 2H),

3.14 (q, J = 7.2 Hz, 2H), 3.10-3.05 (m, 2H), 2.98 (td, J = 8.1, 2.5 Hz, 4H), 1.78 – 1.67 (m, 4H),

1.21 (t, J = 7.2 Hz, 3H). 13

C NMR (300 MHz, CDCl3) δ 158.36, 158.18, 154.32, 154.05, 125.21,

124.92, 124.87, 124.72, 107.67, 107.54, 99.32, 98.86, 54.00, 53.17, 51.74, 48.98, 43.33, 28.36,

28.32, 27.02, 25.04, 12.19. HRMS (FAB) Calcd. For C22H27N5O+ [M

+]: 377.2216; found

377.2223.

ethyl 5-(6-(1-ethylindolin-6-yloxy)indolin-1-yl)pentanoate (3f). Following the general

procedure, the phenol 1c (1.60 g, 9.77 mmol) and the aryl iodide 2e (3.04, 8.14 mmol) were

coupled to provide 3f (2.67 g, 80%) as a colorless oil. 1H NMR (400 MHz, CDCl3) δ 6.98 (d, J =

7.8 Hz, 1H), 6.29 (d, J = 7.8 Hz, 1H), 6.22 (s, 1H), 3.40 (t, J = 8.2 Hz, 2H), 3.11 (q, J = 7.2 Hz,

2H), 2.95 (t, J = 8.2 Hz, 2), 1.18 (t, J = 7.2 Hz, 3H). 13

C NMR (75 MHz, CDCl3) δ 173.85,

158.28, 158.20, 154.34, 125.14, 124.83, 124.77, 107.69, 107.44, 99.33, 98.93, 60.66, 53.94,

53.15, 49.08, 43.35, 34.48, 28.32, 28.29, 27.16, 23.01, 14.65, 12.15. HRMS (FAB) Calcd. For

C25H32N2O3+ [M

+]: 408.2413; found 408.2408.

7,7'-oxybis(2,2,4-trimethyl-1,2-dihydroquinoline) (S3). 3a (400 mg, 2.00 mmol) was

dissolved in 2,2-dimethoxypropane (5 mL) and treated with p-toluenesulfonic acid monohydrate

and stirred at room temperature for 5 days. The reaction was quenched by the addition of 4% aq.

NaHCO3 (30 mL) and extracted with ethyl acetate (3 x 25 mL). The combined organic layers

were washed with brine, dried over anhydrous Na2SO4 and concentrated to a crude residue.

Purified by flash chromatography to provide 340 mg of S3 as a white crystalline solid (47%). 1H

NMR (400 MHz, CDCl3) δ 7.03 (d, J = 8.2 Hz, 1H), 6.35 (dd, J = 8.2, 1.6 Hz, 1H), 6.13 (d, J =

Page 103: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

92

1.6 Hz, 1H), 5.26 (s, 1H), 3.63 (br s, 1H), 2.01 (s, 3H), 1.28 (s, 6H). 13

C NMR (300 MHz,

CDCl3) δ 158.10, 145.01, 128.38, 127.32, 125.10, 117.36, 108.00, 103.55, 52.39, 31.59, 19.07.

HRMS (FAB+) Calcd. For C24H29N2O

+ [M

+1]: 361.2274; found 361.2280.

diethyl 5,5'-(7,7'-oxybis(2,2,4-trimethylquinoline-7,1(2H)-diyl))dipentanoate (S4). A

solution of S3 (270 mg, 0.75 mmol) in anhydrous acetonitrile (4 mL) was treated with sodium

iodide (112 mg, 0.75 mmol), sodium carbonate (476 mg, 4.50 mmol), and ethyl 5-bromovalerate

(0.60 mL, 3.75 mmol), then heated to reflux. After 24 hours, the reaction was charged with

additional ethyl 2-bromovalerate (0.30 mL, 1.9 mmol) and sodium carbonate (238 mg, 2.25

mmol). After an additional 24 hours of heating, TLC of the reaction indicated a ~3:1 mixture of

the desired bis-alkylated product to the mono-alkylated product. The reaction was then quenched

by the addition of water (30 mL) and extracted with ethyl acetate (3 x 25 mL). The combined

organic layers were washed with brine and dried over anhydrous Na2SO4 and concentrated to a

crude residue. Purified by flash chromatography to provide 230 mg of S4 as a colorless oil

(50%). 1H NMR (400 MHz, CDCl3) δ 6.98 (d, J =8.3 Hz, 2H), 6.25 (dd, J = 8.3, 2.1 Hz, 2H),

6.15 (d, J = 2.2 Hz, 2H), 5.17 (d, J = 1.2 Hz, 2H), 4.14 (q, J = 7.1 Hz, 4H), 3.17 (t, J = 7.0 Hz,

4H), 2.32 (t, J = 6.9 Hz, 4H), 1.97 (d, J = 1.1 Hz, 6H), 1.69 – 1.57 (m, 8H), 1.32 (s, 12H), 1.27

(t, J = 7.1 Hz, 6H). 13

C NMR (300 MHz, CDCl3) δ 173.75, 158.44, 145.65, 128.10, 127.83,

124.81, 118.69, 105.63, 102.00, 60.65, 57.26, 44.25, 34.46, 28.88, 28.29, 22.96, 19.13, 14.64.

HRMS (FAB+) Calcd. For C38H53N2O5

+ [M

+1]: 617.3949; found 617.3947.

7-(1-ethyl-1,2,3,4-tetrahydroquinolin-7-yloxy)-2,2,4-trimethyl-1,2-dihydroquinoline

(S5). A solution of S4 (1.20 g, 4.47 mmol) in acetone (80 mL) was treated with p-toluenesulfonic

acid monohydrate (213 mg, 1.12 mmol) and stirred at room temperature for 48 hours. The

reaction was then concentrated to a residue under reduced pressure, brought up 4% aq. NaHCO3

Page 104: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

93

(30 mL) and extracted with ethyl acetate (3 x 25 mL). The combined organics were washed with

brine and dried over anhydrous Na2SO4 and concentrated to a crude residue. Purified by flash

chromatography to provide 1.17 g of S5 as a white foamy solid (75%). 1H NMR (400 MHz,

CDCl3) δ 7.03 (d, J = 8.4 Hz, 1H), 6.91 (d, J = 8.0 Hz, 1H), 6.38 (d, J = 2.3 Hz, 1H), 6.35 (dd, J

= 8.4, 2.4 Hz, 1H), 6.27 (dd, J = 8.0, 2.3 Hz, 1H), 6.14 (d, J = 2.4 Hz, 1H), 3.71 (s, 1H), 3.37-

3.27 (m, 4H), 2.77 (t, J = 6.4 Hz, 2H), 2.06 – 1.99 (m, 5H), 1.31 (s, 6H), 1.16 (t, J = 7.1 Hz, 3H).

13C NMR (300 MHz, CDCl3) δ 159.07, 156.45, 146.44, 145.01, 130.03, 128.60, 126.92, 125.01,

118.14, 116.89, 107.08, 106.71, 102.71, 102.65, 52.36, 48.61, 45.81, 31.52, 28.02, 22.78, 19.07,

11.15. HRMS (FAB+) Calcd. For C23H28N2O

+ [M

+]: 348.2202; found 348.2186.

ethyl 5-(7-(1-ethyl-1,2,3,4-tetrahydroquinolin-7-yloxy)-2,2,4-trimethylquinolin-

1(2H)-yl)pentanoate (S6). According to the protocol used to synthesize S4, S5 (1.05 g, 3.00

mmol) was subjected to alkylation and purified by flash chromatography to yield 845 mg of S6

as a colorless oil (59%). 1H NMR (400 MHz, CDCl3) δ 6.97 (d, J = 8.3 Hz, 1H), 6.88 (d, J = 8.0

Hz, 1H), 6.36 (d, J = 2.3 Hz, 1H), 6.25 (dd, J = 4.3, 2.3 Hz, 1H), 6.23 (dd, J = 4.6, 2.3 Hz, 1H),

6.11 (d, J = 2.2 Hz, 1H), 5.15 (d, J = 1.3 Hz, 1H), 4.16 (q, J = 7.1 Hz, 2H), 3.35 – 3.25 (m, 4H),

3.15 (t, J = 7.2 Hz, 2H), 2.74 (t, J = 6.3 Hz, 2H), 2.31 (t, J = 7.1 Hz, 2H), 2.03 – 1.93 (m, 2H),

1.97 (s, 3H) 1.67-1.55 (m, 4H), 1.32 (s, 6H), 1.29 (t, J = 7.1 Hz, 3H), 1.13 (t, J = 7.1 Hz, 3H).

13C NMR (300 MHz, CDCl3) δ 173.80, 164.92, 159.31, 156.47, 146.37, 145.54, 129.94, 127.84,

124.81, 118.14, 118.02, 106.52, 104.81, 102.60, 101.17, 60.66, 57.25, 48.58, 45.78, 44.24, 34.47,

28.90, 28.28, 27.95, 22.94, 22.75, 19.13, 14.65, 11.13. HRMS (FAB+) Calcd. For C30H41N2O3

+

[M+1

]: 477.3112; found 477.3129.

ethyl-5-(7-(1-ethyl-1,2,3,4-tetrahydroquinolin-7-yloxy)-2,2,4-trimethyl-3,4-

dihydroquinolin-1(2H)-yl)pentanoate (S7). A round bottom flask charged with a magnetic

Page 105: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

94

stirbar, S6 (185 mg, 0.39 mmol) and ethyl acetate (6 mL) was evacuated under vacuum and

backfilled with argon gas. This process was repeated three times. The vessel was then charged

with 10% palladium on carbon (41 mg, 0.039 mmol), evacuated with vacuum, and backfilled

with hydrogen gas. This process was repeated five times, then the reaction left to stir overnight

under an atmosphere of hydrogen gas (balloon). The next morning, the reaction was filtered

through a pad of celite and concentrated under reduced pressure to provide 180 mg of S7 as a

colorless oil (97%). 1H NMR (400 MHz, CDCl3) δ 7.08 (d, J = 8.3 Hz, 1H), 6.91 (d, J = 8.0 Hz,

1H), 6.41 (s, 1H), 6.31 (d, J = 8.3 Hz, 1H), 6.28 (d, J = 8.0, 1H), 6.21 (s, 1H), 4.20 (q, J = 7.1

Hz, 2H), 3.40 – 3.25 (m, 5H), 3.08 - 2.97 (m, 1H), 2.95– 2.84 (m, 1H), 2.77 (t, J = 6.2 Hz, 2H),

2.35 (m, 2H), 2.05 – 1.97 (m, 2H), 1.77 (dd, J = 12.9, 4.5 Hz, 1H), 1.70 - 1.54 (m, 5H), 1.43 –

1.28 (m, 9H), 1.23 (s, 3H), 1.17 (t, J = 7.0 Hz, 3H). 13

C NMR (300 MHz, CDCl3) δ 173.91,

157.88, 156.64, 146.37, 146.30, 129.98, 126.73, 122.43, 117.89, 106.58, 104.62, 102.75, 101.60,

60.72, 54.84, 48.62, 47.52, 45.82, 45.31, 34.54, 30.08, 29.01, 28.00, 27.28, 25.50, 22.94, 22.81,

20.56, 14.74, 11.16. HRMS (FAB+) Calcd. For C30H42N2O3

+ [M

+]: 478.3195; found 478.3210.

(1-(5-ethoxy-5-oxopentyl)-7-(1-ethyl-1,2,3,4-tetrahydroquinolin-7-yloxy)-2,2-

dimethyl-1,2-dihydroquinolin-4-yl)methanesulfonate (S8) sodium salt. To a solution of

fuming sulfuric acid (20% SO3, 1.2 mL) in sulfuric acid (4 mL) cooled in a brine ice bath was

added S6 (954 mg, 2.00 mmol). The mixture was stirred on ice for 30 minutes then allowed to

warm to room temperature and stir for 12 hours at room temperature. The reaction was then

poured into a beaker of ice (50 grams) and the resulting solution, cooled in an ice bath, was

neutralized to pH 7 with 10% sodium hydroxide. The precipitated sodium sulfate solids were

filtered, and the resulting aqueous filtrate was concentrated under reduced pressure to white

solids. This material was boiled gently in 30 mL of ethanol, filtered, and the solids washed with

Page 106: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

95

additional hot ethanol. The filtrate was concentrated under reduced pressure and purified by flash

chromatography to provide 825 mg of S8 as a light pink oil (72%). 1H NMR (400 MHz, MeOD)

δ 7.28 (d, J = 8.4 Hz, 1H), 6.83 (d, J = 8.0 Hz, 1H), 6.27 (d, J = 1.8 Hz, 1H), 6.18 (ddd, J = 14.3,

8.2, 1.9 Hz, 2H), 5.98 (d, J = 1.7 Hz, 1H), 5.50 (s, 1H), 4.11 (t, J = 7.2 Hz, 2H), 3.88 (s, 2H),

3.25 (m, 4H), 3.10 (m, 2H), 2.68 (t, J = 6.2 Hz, 2H), 2.24 (t, J = 6.6 Hz, 2H), 1.96 – 1.87

(m,v2H), 1.78 – 1.64 (m, 2H), 1.64 – 1.56 (m, 2H), 1.33 (s, 6H), 1.25 (t, J = 7.1 Hz, 3H), 1.08 (t,

J = 7.0 Hz, 3H). 13C NMR (300 MHz, MeOD) δ 176.67, 174.15, 159.61, 156.29, 146.28, 145.58,

132.49, 129.71, 125.84, 124.63, 118.30, 116.59, 106.86, 104.70, 102.72, 100.70, 67.60, 61.55,

60.52, 57.03, 53.82, 48.28, 45.34, 43.91, 33.89, 33.62, 32.04, 28.77, 27.89, 27.67, 27.37, 22.56,

22.46, 21.54, 13.74, 10.10. HRMS (FAB+) Calcd. For C30H39N2O6SNa

2+ [M

+2Na]: 601.2324;

found 601.2307.

3.5.4 4-Nitrobenzenediazonium Tetrafluoroborate Reactions

The diaryl ether (1.00 mmol) was dissolved in methanol or ethanol (2 mL), cooled to 0°C

in an ice bath, then treated with aqueous 2N HCl (5 mL). After sufficient time to cool, the

reaction was treated with p-nitrobenzenediazonium tetrafluoroborate (1.00 mmol) in 5 mg

portions over 5 minutes, then stirred at 0°C for 1 hour. The reaction was then diluted with

dichloromethane (10 mL), quenched by the slow addition of saturated aq. NaHCO3 and

extracted with dichloromethane (3 x 20 mL). The combined organic layers were dried over

Na2SO4 and concentrated in vacuo to a deep red-purple residue. Flash chromatography (10-30%

ethyl acetate/hexanes) provided the diazene as a mixture of regioisomers in the case of

asymmetrical diaryl ether starting materials (both of which transform to the same oxazine

product upon cyclization). Full characterization of diazene 4d is provided.

Page 107: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

96

Diazene 4a. Following the general procedure, S4 (205 mg, 0.332 mmol) was reacted to

give 167 mg of 4a as a deep purple residue (66%).

Diazenes 4b and 5b. Following the general procedure, S6 (185 mg, 0.388 mmol) was

reacted to give 155 mg of a mixture of 4b and 5b as a deep purple residue (69%).

Diazenes 4c and 5c. Following the general procedure, S7 (180 mg, 0.376 mmol) was

reacted to give 170 mg of a mixture of 4c and 5c as a deep purple residue (78%).

Diazene 4d. Following the general procedure, 3d (230 mg, 0.756 mmol) was reacted to

give 270 mg of 4d as a deep red/bronze solid (79%). 1H NMR (400 MHz, CDCl3) δ 8.24 (d, J =

9.1 Hz, 2H), 7.79-7.71 (m, 3H), 7.00 (d, J = 7.6 Hz, 1H), 6.34 (d, J = 7.9 Hz, 1H), 6.30 (s, 1H),

6.06 (s, 1H), 3.67 (t, J = 8.2 Hz, 2H), 3.42 (t, J = 8.2Hz, 2H), 3.27 (t, J = 7.1 Hz, 2H), 3.13 (t, J =

7.2 Hz, 2H), 3.08 (t, J = 8.1 Hz, 2H), 2.97 (t, J = 8.2 Hz, 2H), 1.26- 1.14 (m, 6H). 13

C NMR (300

MHz, CDCl3) δ 160.66, 158.73, 158.04, 157.61, 154.21, 146.99, 135.85, 127.05, 125.56, 124.98,

122.91, 113.32, 107.55, 99.21, 96.07, 53.18, 51.95, 43.28, 41.59, 28.31, 26.97, 12.11.

HRMS(FAB+) Calcd. For C26H27N5O3

+ [M

+]: 457.2114; found 457.2131.

Diazenes 4e and 5e. Following the general procedure, 3e (240 mg, 0.636 mmol) was

reacted to give 237 mg of a mixture of 4e and 5e as a deep purple film (71%).

Diazenes 4f and 5f. Following the general procedure, 3f (1.42 g, 3.47 mmol) was reacted

to give 1.46 g of a mixture of 4f and 5f as a deep purple residue (76%).

Diazenes 4g and 5g. A round bottom flask charged with a magnetic stir-bar, S8 (150 mg,

0.26 mmol) and methanol (10 mL) was evacuated under vacuum and backfilled with argon gas.

This process was repeated three times. The vessel was then charged with 10% palladium on

carbon (30 mg, 0.026 mmol), evacuated with vacuum, and backfilled with hydrogen gas. This

process was repeated five times, then the reaction left to stir overnight under an atmosphere of

Page 108: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

97

hydrogen gas (balloon). The next morning, the reaction was filtered through a pad of celite and

concentrated under reduced pressure to tan oil. The resulting oil was then brought up in methanol

(2 mL) and cooled in an ice bath for 10 minutes. The solution was then treated with 2N aqueous

HCl (5 mL, pre-cooled in ice bath) followed by the slow addition of p-nitrobenzenediazonium

tetrafluoroborate (62 mg, 0.26 mmol). After stirring on ice for 1.5 hours, the reaction was diluted

with water (20 mL) and extracted directly without neutralization (DCM, 3 x 25 mL). The

combined organic layers were dried over Na2SO4 and concentrated in vacuo to a deep red-purple

residue. Flash chromatography (5-30% dichloromethane/methanol) provided the regioisomeric

mixture of the diazenes 4g and 5g (115 mg, 63%).

3.5.5 Diazene-Diaryl Ether Cyclization to Oxazine

The diazene diaryl ether (0.2 mmol) was dissolved in anhydrous ethanol (20 mL) and

treated with p-toluenesulfonic acid monohydrate (0.6 mmol). The deep red reaction mixture was

heated to 65°C – 70°C, becoming deep blue after a short period of heating. The reaction was

monitored by TLC (8% MeOH/DCM) and continued until complete conversion of starting

material to the product (typically ~4-8 hrs). After cooling to room temperature, the reaction was

treated with 4% NaHCO3 and extracted with dichloromethane (3 x 25 mL). The combined

organics were dried over anhydrous Na2SO4, then concentrated to a crude blue-red residue.

Products were purified by flash chromatography (silica gel) using a gradient of 0-10%

MeOH/DCM.

Oxazine 6a. Following the general procedure, 4a (91 mg, 0.119 mmol) was cyclized to

provide 70 mg of 6a as a purple film (74%). 1H NMR (400 MHz, MeOD) δ 7.70 (d, J = 8.1 Hz,

2H), 7.42 (s, 2H), 7.19 (d, J = 8.0 Hz, 2H), 6.74 (s, 2H), 5.85 (s, 2H), 4.18 (q, J = 7.1 Hz, 4H),

3.73 (br s, 4H), 2.48 (br s, 4H), 2.35 (s, 3H), 2.13 (s, 6H), 1.83 (br s, 8H), 1.59 (s, 12H), 1.28 (t, J

Page 109: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

98

= 7.1 Hz, 6H). 13C NMR (300 MHz, MeOD) δ 173.92, 153.45, 149.98, 142.77, 140.44, 135.13,

134.47, 128.71, 126.69, 125.96, 125.78, 124.72, 96.03, 61.76, 60.58, 33.42, 28.35, 27.41, 22.16,

20.36, 17.64, 13.66. HRMS (FAB) Calcd. for C38H50N3O5+

[M+]: 628.3745; found 628.3775.

Oxazine 6b. Following the general procedure, the mixture of regioisomeric diazenes 4b

and 5b (140 mg, 0.241 mmol) was cyclized to provide 135 mg of 6b as a deep blue film (85%).

1H NMR (400 MHz, MeOD) δ 7.70 (d,=J = 8.0 Hz, 2H), 7.43 (s, 1H), 7.40 (s, 1H), 7.19 (d, J =

8.0 Hz, 2H), 6.90 (s, 1H), 6.68 (s, 1H), 5.82 (s, 1H), 4.18 (q, J = 7.1 Hz, 2H), 3.80 – 3.64 (m,

6H), 2.94 (t, J = 5.9 Hz, 2H), 2.48 (br s, 2H), 2.35 (s, 3H), 2.11 (s, 3H), 2.09 – 2.02 (m, 2H), 1.82

(br s, 4H), 1.58 (s, 6H), 1.37 (t, J = 7.0 Hz, 3H), 1.28 (t, J = 7.1 Hz, 3H). 13

C NMR (300 MHz,

MeOD) δ 173.95, 154.47, 153.30, 149.94, 148.86, 142.75, 140.48, 135.05, 134.19, 130.69,

129.82, 128.72, 126.26, 125.96, 125.76, 124.74, 95.80, 95.31, 61.59, 60.58, 50.26, 46.16, 33.39,

28.40, 27.41, 27.35, 22.14, 20.93, 20.35, 17.64, 13.65, 10.79. HRMS (FAB) Calcd. For

C30H38N3O3+ [M

+]: 488.2908; found 488.2897.

Oxazine 6c. Following the general procedure, the mixture of regioisomeric diazenes 4c

and 5c (150 mg, 0.257 mmol) was cyclized to provide 148 mg of 6d as a deep blue film (87%).

1H NMR (400 MHz, MeOD) δ 7.68 (d, J = 8.1 Hz, 2H), 7.50 (s, 1H), 7.41 (s, 1H), 7.15 (d, J =

8.0 Hz, 2H), 6.88 (s, 1H), 6.66 (s, 1H), 4.17 (q, J = 7.1 Hz, 2H), 3.70 (t, J = 5.6 Hz, 4H), 3.57 –

3.44 (m, 1H), 2.98 (d, J = 6.5 Hz, 1H), 2.95 – 2.90 (m, 2H), 2.47 (br s, 2H), 2.31 (s, 3H), 2.10 –

1.97 (m, 3H), 1.79 (br s, 4H), 1.53 (s, 3H), 1.45 (s, 6H), 1.36 (t, J = 6.9 Hz, 3H), 1.28 (t, J = 7.1

Hz, 4H). 13C NMR (300 MHz, MeOD) δ 173.90, 154.75, 154.32, 149.04, 148.01, 142.96,

140.32, 135.02, 134.20, 133.45, 130.84, 129.89, 128.68, 128.13, 125.98, 96.56, 95.30, 60.56,

59.25, 50.34, 46.62, 44.47, 33.44, 28.40, 27.75, 27.40, 27.16, 25.26, 22.27, 20.90, 20.44, 18.49,

13.73, 10.90. HRMS (FAB) Calcd. For C32H38N3O3+ [M

+]: 490.3064; found 490.3079.

Page 110: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

99

Oxazine 6d. Following the general procedure, diazene 4d (108 mg, 0.236 mmol) was

cyclized to provide 105 mg of 6d as a deep blue film (90%). 1H NMR (400 MHz, MeOD) δ 7.71

(d, J = 8.2 Hz, 2H), 7.40 (s, 2H), 7.22 (d, J = 8.0 Hz, 2H), 6.68 (s, 2H), 4.03 (t, J = 7.2 Hz, 4H),

3.66 (q, J = 7.3 Hz, 4H), 3.29 (t J = 7.2 Hz, 4H), 2.36 (s, 3H), 1.35 (t, J = 7.3 Hz, 6H). 13

C NMR

(300 MHz, MeOD) δ 159.56, 150.80, 142.51, 140.64, 137.96, 134.66, 128.75, 125.96, 90.43,

52.73, 41.92, 26.09, 20.29, 11.16. HRMS (FAB) Calcd. for C20H22N3O+ [M

+]: 320.1757; found

320.1764.

Oxazine 6e. Following the general procedure, the mixture of regioisomeric diazenes 4e

and 5e (157 mg, 0.298 mmol) was cyclized to provide 130 mg of 6e as a dark blue film (78%).

1H NMR (400 MHz, MeOD) δ 7.51 – 7.41 (m, 2H), 6.74 (s, 1H), 6.72 (s, 1H), 4.05 (q, J = 6.6

Hz, 4H), 3.75-3.62 (m, 4H), 3.43 (t, J = 6.6 Hz, 2H), 3.31 (m, 4H), 1.92 – 1.80 (m, 2H), 1.77-

1.66 (m, 2H), 1.36 (t, J = 7.3 Hz, 3H). 13C NMR (300 MHz, CDCl3) δ 159.96, 159.73, 150.99,

150.79, 138.22, 137.65, 135.03, 134.58, 126.07, 90.55, 90.43, 53.25, 52.78, 51.08, 46.74, 41.94,

26.26, 26.14, 26.07, 24.40, 11.15. HRMS (FAB) Calcd. for C22H25N6O+ [M

+]: 389.2084; found

389.2090.

Oxazine 6f. Following the general procedure, the mixture of regioisomeric diazenes 4d

and 5d (1.05 g, 1.88 mmol) were cyclized to provide 1.04 g of 7f as a metallic red amorphous

solid (94%). 1H NMR (400 MHz, MeOD) δ 7.69 (d, J = 8.0 Hz, 2H), 7.24 (s, 2H), 7.16 (d, J =

8.0 Hz, 2H), 6.60 (s, 1H), 6.56 (s, 1H), 4.14 (q, J = 7.1 Hz, 2H), 3.97 (t, J = 6.3 Hz, 4H), 3.63 –

3.51 (m, 4H), 3.21 (br s, 4H), 2.42 (t, J = 6.9 Hz, 2H), 2.31 (s, 3H), 1.81 – 1.65 (m, 4H), 1.31 (t,

J = 7.2 Hz, 3H), 1.26 (t, J = 7.1 Hz, 3H). 13C NMR (300 MHz, MeOD) δ 173.95, 159.77, 159.43,

150.48, 150.37, 143.03, 140.38, 138.04, 137.62, 134.61, 134.32, 128.73, 125.98, 90.66, 90.52,

Page 111: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

100

60.54, 53.29, 52.80, 47.00, 42.04, 33.48, 26.50, 26.18, 26.14, 22.28, 20.41, 13.67, 11.35. HRMS

(FAB) Calcd. for C25H30N3O3+ [M

+]: 420.2282; found 420.2291.

Oxazine 6g. Following the general procedure, the mixture of regioisomeric diazenes 4g

and 5g (102 mg, 0.144 mmol) were cyclized to provide 68 mg of 7g as a deep blue residue

(83%). 1H NMR (400 MHz, MeOD) δ 7.69 (s, 1H), 7.41 (s, 1H), 6.92 (s, 1H), 6.64 (s, 1H), 4.18

(q, J = 7.1 Hz, 2H), 3.74 (m, 5H), 3.60 (dd, J = 14.0, 4.0 Hz, 1H), 3.53 – 3.39 (m, 2H), 3.04-2.91

(m, 3H), 2.59 (dd, J = 13.6, 4.3 Hz, 1H), 2.47 (m, 2H), 2.12 – 2.02 (m, 2H), 1.86 – 1.68 (m, 5H),

1.54 (s, 3H), 1.46 (s, 3H), 1.38 (t, J = 7.0 Hz, 3H), 1.28 (t, J = 7.1 Hz, 3H). 13

C NMR (300 MHz,

MeOD) δ 173.92, 155.10, 154.26, 149.14, 147.89, 142.73, 140.50, 135.47, 133.15, 131.29,

131.07, 130.16, 128.95, 128.72, 125.96, 96.65, 95.36, 60.57, 59.00, 55.06, 50.45, 46.60, 41.84,

33.42, 30.25, 28.26, 27.77, 27.37, 25.08, 22.27, 20.89, 20.33, 13.65, 10.89. HRMS (FAB+)

Calcd. for C30H40N3O6S+ [M

+1]: 570.2632; found 570.2629.

3.5.6 Tandem Friedel-Crafts Acylation/Cyclization

(7a, 7b-f). A small roundbottom flask or sealable vial was purged with argon and charged

with a magnetic stirbar, activated 4 Å molecular sieves, the diaryl ether (0.50 mmol), anhydrous

nitromethane (0.80 mL), the acid chloride (4.0 mmol for 7a, 7d, 7e; 2.5 mmol for 7c, 7f), and

gallium triflate (0.075 mmol). The reaction was heated to 60°C for 16 hours. After cooling to

RT, the reaction was diluted with dichloromethane (25 mL) and 4% aq. NaHCO3 (30 mL) and

extracted with dichloromethane (3 x 25 mL). The combined organic layers were dried over

anhydrous Na2SO4 and then concentrated to crude darkly colored residues. Products were

purified by flash chromatography (silica gel) using a gradient of 0-10% MeOH/DCM.

Xanthene 7a. Following the general procedure, diaryl ether 3d (154 mg, 0.500 mmol)

was reacted with benzoyl chloride (0.46 mL, 4.0 mmol) to provide 120 mg of 7a as a deep purple

Page 112: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

101

residue (56%) and 55 mg of recovered 3d (36%). 1H NMR (400 MHz, MeOD) δ 7.70-7.65 (m,

3H), 7.45-7.40 (m, 2H), 6.96 (t, J = 1.7 Hz, 2H), 6.72 (s, 2H), 3.92 (t, J = 7.8 Hz, 4H), 3.60 (q, J

= 7.2 Hz, 4H), 3.15 – 3.08 (m, 4H), 1.33 (t, J = 7.3 Hz, 6H). 13C NMR (300 MHz, MeOD) δ

159.69, 159.47, 154.79, 134.33, 133.47, 129.89, 129.43, 129.00, 123.22, 114.41, 90.55, 51.93,

41.27, 26.08, 10.92. HRMS (FAB) Calcd. for C27H27N2O+

[M+]: 395.2118; found 395.2122.

Xanthene 7c. Following the general procedure, diaryl ether 3d (154 mg, 0.500 mmol)

was reacted with p- (Me2N)-benzoyl chloride (460 mg, 2.5 mmol) to provide 182 mg of 7c as a

deep red residue (77%). 1H NMR (400 MHz, MeOD) δ 7.30 (d, J = 8.8 Hz, 2H), 7.19 (s, 2H),

7.02 (d, J = 8.8 Hz, 2H), 6.64 (s, 2H), 3.89 (t, J = 7.8 Hz, 4H), 3.57 (q, J = 7.2 Hz, 4H), 3.17 –

3.08 (m, 10H), 1.32 (t, J = 7.2 Hz, 6H). 13

C NMR (300 MHz, CDCl3) δ 159.55, 159.12, 151.82,

133.76, 131.33, 123.76, 121.12, 120.45, 114.34, 112.31, 90.50, 51.82, 41.21, 39.67, 26.16, 10.91.

HRMS (FAB) Calcd. for C29H32N3O+ [M

+]: 438.2540; found 438.2545.

Xanthene 7d. Following the general procedure, diaryl ether 3d (154 mg, 0.500 mmol)

was reacted with p-fluorobenzoyl chloride (0.47 mL, 4.0 mmol) to provide 106 mg of 7d as a

deep purple residue (48%) and 63 mg of recovered 3d (41%). 1H NMR (400 MHz, MeOD) δ

7.50 – 7.39 (m, 4H), 6.96 (s, 2H), 6.71 (s, 2H), 3.92 (t, J = 7.8 Hz, 4H), 3.61 (q, J = 7.2 Hz, 4H),

3.19 – 3.08 (m, 4H), 1.33 (t, J = 7.2 Hz, 6H). 13

C NMR (300 MHz, CDCl3) δ 159.65, 159.48,

153.65, 134.46, 131.84, 131.73, 129.45, 123.07, 116.24, 115.94, 114.50, 90.59, 51.95, 41.29,

26.10, 10.92. HRMS (FAB) Calcd. for C27H26N2FO+ [M

+]: 413.2024; found 413.2017.

Xanthene 7e. Following the general procedure, diaryl ether 3d (154 mg, 0.500 mmol)

was reacted with 2,3,4,5,6-pentafluorobenzoyl chloride (0.57 mL, 4.0 mmol) to provide 138 mg

of 7e as a green/blue solid (53%) and 61 mg of recovered 3d (40%). 1H NMR (400 MHz,

MeOD) δ 7.06 (s, 2H), 6.77 (s, 2H), 3.97 (t, J = 7.6 Hz, 4H), 3.64 (q, J = 7.2 Hz, 4H), 3.27 –

Page 113: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

102

3.10 (m, 4H), 1.35 (t, J = 7.3 Hz, 6H). 13C NMR (300 MHz, MeOD) δ 159.87, 159.45, 145.99,

142.80, 140.39, 137.53, 135.81, 121.90, 114.53, 107.73, 90.91, 52.21, 41.46, 26.09, 10.97.

HRMS (FAB) Calcd. for C27H22N2F5O+ [M

+]: 485.1647; found 485.1663.

Xanthene 7f. Following the general procedure, diaryl ether 3d (154 mg, 0.500 mmol)

was reacted with 1-methyl-1H-imidazole-2-carbonyl chloride (361 mg, 2.5 mmol) to provide 78

mg of 7e as a deep blue-purple solid (36%) and 68 mg of recovered 3d (44%). 1H NMR (400

MHz, MeOD) δ 7.95 (d, J = 1.9 Hz, 1H), 7.92 (d, J = 1.9 Hz, 1H), 6.85 – 6.81 (m, 4H), 4.01 (t, J

= 7.5 Hz, 4H), 3.80 (s, 3H), 3.67 (q, J = 7.3 Hz, 4H), 3.26 – 3.18 (m, 4H), 1.35 (t, J = 7.3 Hz,

6H). 13C NMR (300 MHz, CDCl3) δ 159.98, 159.36, 138.08, 136.83, 130.75, 125.95, 122.11,

120.92, 114.93, 91.53, 52.45, 41.63, 34.80, 26.12, 11.04. HRMS (FAB) Calcd. for C27H27N4O+

[M+]: 399.2179; found 399.2197.

Xanthene 7b. Diaryl ether 3f (1.15 g, 2.81 mmol) in anhydrous dichloromethane (30

mL) was treated dropwise via a pressure equalizing addition funnel with a solution of

trifluoroacetic anhydride (0.87 mL, 6.2 mmol) in 20 mL of DCM over the course of

approximately 2 hours. After addition was complete, the reaction was left to stir at room

temperature overnight. The next day, the deep blue reaction mixture was quenched by the

addition of 4% aqueous NaHCO3, then extracted with dichloromethane (3 x 30 mL). The

combined organic layers were dried over Na2SO4 and then concentrated under reduced pressure.

The crude product was purified by flash chromatography (0-10% methanol/dichloromethane) to

provide 1.40 g of 7b as a green/blue solid (83%). 1H NMR (400 MHz, MeOD) δ 7.72 (s, 2H),

6.76 (s, 1H), 6.74 (s, 1H), 4.15 (q, J = 7.1 Hz, 2H), 4.01 (td, J = 7.4, 3.7 Hz, 4H), 3.69 – 3.58 (m,

4H), 3.32-3.25 (m, 4H), 2.44 (t, J = 7.0 Hz, 2H), 1.86 – 1.68 (m, 4H), 1.35 (t, J = 7.3 Hz, 3H),

1.26 (t, J = 7.1 Hz, 3H). 13C NMR (300 MHz, CDCl3) δ 173.99, 159.95, 159.85, 159.51, 159.12,

Page 114: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

103

136.61, 136.27, 120.92, 120.84, 120.77, 111.85, 111.67, 91.49, 91.33, 60.53, 52.87, 52.36, 46.61,

41.61, 33.43, 26.42, 22.23, 13.52, 11.09. HRMS (FAB) Calcd. For C27H30F3N2O3+ [M

+]:

487.2203; found 487.2205.

3.5.7 Dye Characterization

All dyes were HPLC purified (C18 semi-prep, MeCN/H2O) prior to characterization.

Measurements were performed in DI water with the exception of 7c, which required 50 mM HCl

for fluorescence. Molar extinction coefficients were determined using the Beer-Lambert law by

measuring absorbances of known concentrations of dye solution. Fluorescence quantum yields

were determined using the comparative method using Nile Blue and zinc-phthalocyanine as

reference standards.27

In brief, fluorescence emission was compared to that of a quantum yield

reference solution under the same excitation and collection conditions.

3.6 References

(1) Tinnefeld, P.; Sauer, M. Angewandte Chemie International Edition 2005, 44,

2642.

(2) Huang, B.; Bates, M.; Zhuang, X. Annual review of biochemistry 2009, 78, 993.

(3) Fernandez-Suarez, M.; Ting, A. Y. Nat Rev Mol Cell Biol 2008, 9, 929.

(4) Haidekker, M. A.; Brady, T. P.; Lichlyter, D.; Theodorakis, E. A. Journal of the

American Chemical Society 2006, 128, 398.

(5) Charier, S.; Ruel, O.; Baudin, J.-B.; Alcor, D.; Allemand, J.-F.; Meglio, A.;

Jullien, L. Angewandte Chemie International Edition 2004, 43, 4785.

(6) Grynkiewicz, G.; Poenie, M.; Tsien, R. Y. The Journal of biological chemistry

1985, 260, 3440.

(7) Yu, F.; Li, P.; Li, G.; Zhao, G.; Chu, T.; Han, K. Journal of the American

Chemical Society 2011, 133, 11030.

Page 115: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

104

(8) Fluhler, E.; Burnham, V. G.; Loew, L. M. Biochemistry 1985, 24, 5749.

(9) Dempsey, G. T.; Vaughan, J. C.; Chen, K. H.; Bates, M.; Zhuang, X. Nat Meth

2011, 8, 1027.

(10) Fölling, J.; Belov, V.; Kunetsky, R.; Medda, R.; Schönle, A.; Egner, A.; Eggeling,

C.; Bossi, M.; Hell, S. W. Angewandte Chemie International Edition 2007, 46, 6266.

(11) Heilemann, M.; van de Linde, S.; Mukherjee, A.; Sauer, M. Angewandte Chemie

International Edition 2009, 48, 6903.

(12) Wombacher, R.; Heidbreder, M.; van de Linde, S.; Sheetz, M. P.; Heilemann, M.;

Cornish, V. W.; Sauer, M. Nat Meth 2010, 7, 717.

(13) Beija, M.; Afonso, C. A.; Martinho, J. M. Chem Soc Rev 2009, 38, 2410.

(14) Kanitz, A.; Hartmann, H. European Journal of Organic Chemistry 1999, 1999,

923.

(15) Pauff, S. M.; Miller, S. C. Organic letters 2011, 13, 6196.

(16) Gribble, G. W.; Lord, P. D.; Skotnicki, J.; Dietz, S. E.; Eaton, J. T.; Johnson, J.

Journal of the American Chemical Society 1974, 96, 7812.

(17) Ullmann, F.; Sponagel, P. Berichte der deutschen chemischen Gesellschaft 1905,

38, 2211.

(18) Scott Sawyer, J. Tetrahedron 2000, 56, 5045.

(19) Burgos, C. H.; Barder, T. E.; Huang, X.; Buchwald, S. L. Angewandte Chemie

International Edition 2006, 45, 4321.

(20) Maiti, D.; Buchwald, S. L. Journal of the American Chemical Society 2009, 131,

17423.

(21) Klapars, A.; Buchwald, S. L. Journal of the American Chemical Society 2002,

124, 14844.

(22) Mitronova, G. Y.; Belov, V. N.; Bossi, M. L.; Wurm, C. A.; Meyer, L.; Medda,

R.; Moneron, G.; Bretschneider, S.; Eggeling, C.; Jakobs, S.; Hell, S. W. Chemistry (Weinheim

an der Bergstrasse, Germany) 2010, 16, 4477.

(23) Olah, G. A.; Farooq, O.; Farnia, S. M. F.; Olah, J. A. Journal of the American

Chemical Society 1988, 110, 2560.

Page 116: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

105

(24) Prakash, G. K.; Mathew, T.; Olah, G. A. Accounts of chemical research 2012, 45,

565.

(25) Clunas, S.; Strory, J.; Rickard, J.; Horsley, D.; Harrington, C.; Wischik, C.;

Office, U. S. P. a. T., Ed.; Wista Laboratories Ltd.: United States, 2010

(26) Yamada, Y.; Akiba, A.; Arima, S.; Okada, C.; Yoshida, K.; Itou, F.; Kai, T.;

Satou, T.; Takeda, K.; Harigaya, Y. Chemical & pharmaceutical bulletin 2005, 53, 1277.

(27) Williams, A. T. R.; Winfield, S. A.; Miller, J. N. Analyst 1983, 108, 1067.

Page 117: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

106

Chapter 4

Development of Targeted Fluorescent Nanodiamonds for Cell Imaging

Page 118: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

107

4.1 Chapter Overview

Compared to organic dyes, fluorescent nanomaterials possess superior photophysical

properties and are ideal fluorophores for SM imaging. In particular, fluorescent nanodiamonds

(FNDs) are well suited for live cell imaging due to their biocompatibility. FNDs are

extraordinarily photostable, allowing fluorescence imaging on unprecedented time scales. In

addition, FNDs have other optical sensing capabilities that can be used for multi-color imaging,

probing molecular orientation, and measuring magnetic field. The application of FNDs in

biological studies would revolutionize our ability to investigate molecular mechanisms at the SM

level, but the adoption of these fluorophores in bioimaging has been limited by the ability to

chemically modify the FND surface for targeted labeling of biomolecules.

This chapter focuses on the development of FNDs for SM imaging in live cells. A

proposal to functionalize FNDs with TMP is presented, allowing the TMP-tag to label specific

biomolecules with FND fluorophores for fluorescence imaging applications. We explore

methods to size select for small NDs that are less likely to interfere with biomolecular function

during labeling. Next, we investigate chemical methods to modify the ND surface that would

enable straightforward covalent conjugation with TMP ligands. Finally, we attempt live cell

extracellular labeling and imaging of NDs using the TMP-tag. I am the main contributor of this

work. I synthesized and characterized the silanized NDs with assistance from Ophir Gaaton and Abe

Wolcott. Yongjun Li synthesized TMP-functionalized ND while I performed the live cell imaging

experiments of TMP-NDs.

4.2 Introduction

FNDs have emerged as an interesting class of material for biological imaging because of

their unique structural, chemical, biological, and optical properties. NDs are nanometer sized

Page 119: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

108

diamond crystals whose fluorescence arises from defect centers in the diamond crystal lattice. In

particular, the red fluorescence and other optical properties of the nitrogen vacancy (NV-) center

defect are appealing for live cell SM imaging. However, the natural occurrence of NV- center

defects in NDs too low for general fluorescent imaging applications.1,2

Consequently, methods

have been created to implant NV- centers and produce FNDs.

3-7 Because the NV

- center is

housed within the diamond lattice, manipulation of these fluorophores requires control over ND

chemical properties. Thus, techniques to modify the surface chemistry of NDs for targeted

biological labeling and applications have also been developed.8 However, the efficiency and

costs of these methods varies significantly based on the synthetic origin of the NDs due to

differences in size, shape, and lattice purity. For successful applications in live cell SM imaging,

NDs must be fluorescent, small, and functionalized for targeted labeling of biomolecules.

4.2.1 Types and Synthesis of Nanodiamonds

The first synthesized, and most commonly used, NDs were created by the denotation of

carbonaceous material with negative oxygen balance (oxygen content lower than necessary for

complete combustion) under high pressure.9 These NDs are called detonation nanodiamonds

(DNDs) or ultradisperse diamonds (UDDs). While first produced by Soviet researchers in the

1960s, DNDs did not gain broader awareness within the international scientific community until

researchers at Los Alamos reported their synthesis in the late 1980s.10,11

DNDs are typically

4-5 nm in size, making them smaller and possessing narrower distributions compared to NDs

synthesized from other methods. However, DNDs require extensive purification to remove metal

impurities produced from the ignition source and reaction chamber wall.12

In addition, DNDs are

especially prone to aggregation, requiring milling methods that introduce additional

contaminants to stabilized colloidal DND solutions.13,14

Most importantly, DNDs are synthesized

Page 120: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

109

with relatively few lattice impurities, which are critical for creating fluorescent NV- defect

centers. As a result, DNDs have been typically studied for non-fluorescent applications. DNDs

have been investigated for industrial applications, particularly as additives for lubricants and

composite materials.15,16

In addition, DNDs are being developed as carriers for poorly water

soluble drugs due to their high surface area, large adsorptive capacity, and biological inertness.17

After DNDs, the second most common type of NDs is formed using high pressure and

high temperature (HPHT) synthesis. Mimicking the conditions for natural diamond formation,

HPHT diamonds are produced by applying high pressure and temperature to a carbon source.

High energy ball milling of HPHT diamond microcrystals yield HPHT NDs of varying sizes.

The major advantage of HPHT NDs is that they can be synthesized to include high nitrogen

impurities by using starting material with high nitrogen concentrations.18

The high nitrogen

content of HPHT NDs allows for the implantation of NV- centers around existing nitrogen while

also increasing the frequency of naturally occurring NV- centers.

3 As a result, FNDs are usually

created from HPHT NDs. However, HPHT NDs have some notable drawbacks in comparison to

DNDs. The size distribution of HPHT NDs is not as narrow as DNDs and the yield of NDs

smaller than 10 nm is very low.19

Also, the high nitrogen content can create multiple NV- centers

per diamond, which can lead to interference with certain sensing capabilities.20

While there are

other methods of synthesizing high purity NDs with single NV- centers using chemical vapor

deposition, these methods are very costly.20,21

All in all, the relative ease and low cost of creating

NV- centers in HPHT NDs makes them the most common type of diamond used for FNDs.

4.2.2 Nanodiamond Nitrogen Vacancy (NV-) Centers

The methodology for the creation and implantation of NV- centers in HPHT NDs has

significantly improved the production yield of FND for fluorescence imaging applications.

Page 121: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

110

Nitrogen impurities must be accompanied by a vacancy in the diamond crystal lattice to form the

fluorescent NV- center. However, the nitrogen impurities in HPHT NDs are usually found in NV

-

centers. To generate the NV- center, vacancies in the diamond lattice surface are produced used

high energy electron or helium beam irradiation. Vacuum annealing of NDs migrates the

vacancies to nitrogen sites, where they create more stable NV- centers.

3-5 One of the challenge to

generating small (<10 nm) FNDs is that the probability of containing nitrogen and receiving

sufficient radiation for vacancy formation is lower for smaller crystals. However, FNDs as small

as 5 nm have been reported to contain NV- centers and are particularly useful for biological

applications due to their small size.22

The NV- center has unique optical properties beyond conventional fluorescence that can

be used for imaging and sensing. Containing six electrons, the NV- center is formed by three

unpaired lone electrons from the lattice vacancy, two nitrogen lone pairs electrons, and an

additional electron usually captured from the lattice from nitrogen donors (Figure 4-1a). The NV-

center produces infrared fluorescence centered at 680 nm after excitation from the triplet ground

state (3A) with 532 nm light to the triplet excited state (

3E) (Figure 4-1b,c).

23 Excitation from the

ms = ± 1 spin sublevel, however, has a greater likelihood of intersystem crossing to the singlet 1A

state, which nonradiatively decays to the 3A ms=0 state. As a result, resonant microwave

radiation that induces ms=0 to ms= ± 1 transitions reduces fluorescence by as much as 30%.24

Application of a magnetic field splits the ms= ±1 states, changing the resonant microwave

wavelength (Figure 4-1d). The ability to readout ms spin state transitions by reductions in

fluorescence is called optically detected magnetic resonance (ODMR) and can be used to

determine magnetic field strength.25

When used with a known magnetic field, the orientation of

the NV- center can also be determined.

26 Resonant microwave wavelength variation between

Page 122: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

111

NV- centers based on different strain in the lattice structure can be used to differentiate FNDs

and may be useful for multicolor imaging.26

Taken together, the fluorescence from NV- centers

can be used for both imaging and sensing applications in biological research.

Figure 4-1: Diamond nitrogen vacancy (NV

-) defect center

a) Atomic representation of the NV- center containing a nitrogen impurity and accompanying vacancy in the crystal

lattice structure. b) Fluorescence emission of a single NV- center. c) Energy level diagram of the NV- center. d)

Optically detected magnetic resonance spectra of a single NV center spin. Figure adapted from Balasubramaian et

al, 2013.23

Importantly, the NV- center is shielded by the diamond lattice, leaving it unaffected by

changes to the FND surface and surrounding environment. As a result, the optical properties of

the NV- center remain highly preserved even under intense irradiation. The NV

- center is

functionally immune to photobleaching and this remarkable photostability has enabled imaging

resolution under 5 nm using super resolution stimulated emission depletion (STED), the highest

resolution ever achieved using optical microscopy.23,27-30

In addition, the NV- center has an

unusually long fluorescence lifetime of nearly 15 ns in comparison to organic fluorophores of 1-

Page 123: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

112

4 ns, making it a useful marker for fluorescence lifetime imaging.31

Although smaller NDs have

higher probabilities of having NV- centers close to the surface that can be subjected to

environmentally induced blinking, this photophysical behavior could be exploited for super

resolution imaging using STORM.32

Overall, the surrounding environment and surface chemistry

of FNDs have relatively little impact on the fluorescence of NV- centers, making FNDs very

attractive fluorophores for SM imaging in live cells. Thus, methods to functionalize the ND

surface for targeted bioimaging have also been developed.

4.2.3 Chemical Modification of the Nanodiamond Surface

Successful application of FNDs for live cell SM imaging requires the ability to direct

FNDs to label specific biomolecules. At the same time, individual FNDs must also be stabilized

in solution as their tendency to aggregate could cause adverse impacts to biological systems and

affect the ability to perform SM imaging and measurements. Additionally, reduced aggregation

increases the accessibility of ND surfaces for covalent conjugation of targeting molecules. While

successful functionalization of HPHT ND surfaces has been achieved, the chemistry of DND

surfaces is more widely reported and exhibits greater reaction diversity.8 Interestingly, different

types of NDs have distinct surface properties and characteristics, limiting the applicability of

surface chemistry from DNDs to functionalize HPHT NDs.8,33,34

It is hypothesized that DNDs

are easier to functionalize because their smaller size and more spherical shape results in

comparatively more edges that host chemically reactive functional groups. Nevertheless, a

number of strategies to functionalize the HPHT ND surface have been developed. 22,35,36

Regardless of the surface chemistry being performed, nearly all chemical modification of

HPHT NDs begins with surface oxidation (Figure 4-2). This step removes outer graphitic

material and amorphous carbon that is formed during ball milling of ND production, enhancing

Page 124: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

113

FND fluorescence. In addition, oxidation instills various oxygen-containing functional groups

that can be used for chemistry, including ketones, alcohols, acid anhydrides and carboxylic

acids.36

Oxidation was first reported using washes with strong acids and bases, but more

environmentally friendly and convenient methods using air oxidation in tube furnaces have been

reported.36

Early reports determined that carboxylic acids were the predominant functional group

on the oxidized ND surface and could be covalently coupled to amines.37,38

Reduction of the

carboxylic moieties to hydroxyl groups also allows coupling to functionalized silane reagents

that can be later modified.39,40

Overall, steady progress is being made towards developing robust

surface modification chemistry and increasing the functionality of NDs for biological and other

applications.

Figure 4-2: Changes to HPHT ND appearance during synthesis and oxidation

Bulk HPHT diamond contains approxi atel 100−200 ppm of nitrogen that produces a yellow color. Ball milling

yields nanometer sized diamond crystals ranging in size from 5 to 50 nm with a black color originating from

graphitic and amorphous carbon produced during the milling process. Aerobic oxidation of HPHT NDs removes

graphitic and amorphous carbon, lightening the ND powder color with increasing temperatures. Figure adapted from

Wolcott et al, 2014.41

4.2.4 Nanodiamond Labeling with the TMP-tag

The advances in surface chemistry of HPHT NDs have made targeted cell labeling and

imaging with FNDs possible. While a number of studies have used molecules adsorbed to the

ND surface for targeted labeling, there are few examples using covalent attachments.4,42

The first

Page 125: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

114

report of covalent surface modification of NDs for cellular imaging was performed in 2007,

where 100 nm NDs were conjugated to human growth hormone and labeled extracellular growth

hormone receptors. While not using FNDs, the study used Raman imaging based on the carbon

sp3 peak at 1332 cm

-1 to visualize the NDs.

43 The use of FNDs for targeted fluorescent cell

imaging was reported one year later. FNDs were conjugated to transferrin and used to label

extracellular transferrin receptors.44

In both experiments, the carboxylic acids on the surface of

oxidized HPHT NDs had been covalently linked to their targeting biomolecules using EDC (1-

Ethyl-3-(3-dimethylaminopropyl)carbodiimide) coupling to amines. These initial promising

experiments demonstrated the compatibility of NDs for live cell imaging, but were limited by

targeting receptors based on the identity of the ND-conjugated molecule. Recently, streptavidin

functionalized FNDs were used to label biotin-antibodies via biotin-streptavidin binding. These

FND-antibody complexes were then used to label extracellular carbohydrate receptors for live

cell florescent imaging.45

While biotin-streptavidin labeling is robust and has high affinity, it is

poorly suited for direct cellular labeling due to cross-reactivity with endogenous biotin. By using

chemical tags, such as the TMP-tag, direct and single step biomolecule labeling of FNDs to

multitude of different extracellular and intracellular proteins could be achieved.

Page 126: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

115

Figure 4-3: Fluorescent ND protein labeling strategy with the TMP-tag.

Covalent functionalization of NDs with TMP ligands allows ND labeling of protein of interest (POI) with a eDHFR

fusion via the selective and high affinity binding between eDHFR and TMP. The photostability of NDs allow

labeled proteins to examine visualized with fluorescent SM imaging.

Labeling FNDs using the TMP-tag can improve the ability to use these fluorophores for

live cell imaging applications (Figure 4-3). By covalently linking FNDs to TMP, the FND can

label proteins of interest using an eDHFR fusion with the TMP-tag. Although the FND may be

relatively large compared to the protein, the high affinity binding between TMP-eDHFR

interaction should facilitate effective protein labeling. The TMP-tag has been used with magnetic

iron oxide nanoparticles for magnetic field sensing while other chemical tags have successfully

labeled proteins with nanomaterials for fluorescent imaging.46-48

Additionally, the solubility of

the TMP moiety may help stabilize FND in solution and reduce aggregation. The synthesis and

application of a TMP-functionalized FND for live cell imaging with the TMP-tag would

significantly broaden our ability to perform SM imaging and open opportunities for magnetic

sensing, orientation sensing, and molecular tracking.

4.3 Experimental Methods

4.3.1 Size Separation

Colloidal aqueous monocrystalline ND solution with size dispersity of 0-30 nm and

median size of 18 nm suspended were obtained from Microdiamant. NDs were diluted to a

Page 127: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

116

concentration of 5 mg/mL in water and centrifuged using Avanti JE or Proteome Lab XL-A

Beckman Coulter centrifuges. Suspended NDs in the supernatant were taken as the size separated

fraction. Sizes of NDs were analyzed using TEM images of 60-100 NDs and DLS of 1 mL of 0.1

mg/mL solutions. Size separation of NDs was also attempted using dialysis with 20K MWCO

Slide-A-Lyzer dialysis cassettes (Thermo Fisher) and 100K MWCO protein concentrators

(Pierce). Gel filtration was conducted using Sephacryl S200HR (GE Healthcare Life Sciences) in

a gravity flow column.

4.3.2 Chemical Conjugation

Monocrystalline ND powder with size dispersion of 0-30 nm and median size of 18 nm

were obtained from Microdiamant. 200 mg of NDs were oxidized in a tube furnace at 575°C for

2 hours or 475°C for 24 hours. NDs were then washed in 150 mL of 1:9 HNO3: H2SO4 at 80°C

for 24 hours, 1 M NaOH at 80°C for 24 hours, and 0.1 M HCl at 80°C for 24 hours. Synthesis of

hydroxylated and aminosilanized NDs was carried out according to literature procedures.8,38

Synthesis of alkyne coupled NDs was carried out using a modified procedure from the

literature.22

10 mg of NDs and 200 L of propargylamine (Sigma) were dissolved in 5 mL of

water with 8 mg of N-(3-dimethylaminopropyl)-N′-ethyl-carbodiimide hydrochloride (Sigma)

and 6 mg of N-hydroxysuccinimide (Sigma). The reaction was stirred at room temperature for 24

hours. NDs were washed thoroughly with DI water and centrifuged before use.

TMP-functionalized ND were prepared using unoxidized NDs. Boron reduction and

silanization with APTES was carried out as previously described.49

The resulting ND-NH2 (20

mg) was suspended in DMF (5 mL). The suspension was sonicated for 30 min. With stirring and

under N2, 20 mg of TMP-COOH was added to the suspension, followed by 20 mg of (1-Ethyl-3-

(3-dimethylaminopropyl)carbodiimide) (EDC·HCl), 2 mg of 1-Hydroxy-7-azabenzotriazole

Page 128: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

117

(HOAt) and 200 L N,N-Diisopropylethylamine (DIEA). The final suspension was stirred at

room temperature for 3 days under N2. Then the suspension was transferred to 1 mL centrifuge

tubes (5) and centrifuged at 10 k rpm for 20 min. The top layer (solvent) was removed and 1 mL

of DMF was added to each centrifuge tube. Stir or sonicate to re-disperse the precipitate in DMF.

Then the suspension was centrifuged again. The process was repeated for 5 times. Then 1 mL of

methanol was added to each centrifuge tube and centrifuged at 10 k rpm for 20 min. The process

was repeated 15 times. After removing methanol, the solid was dried under vacuum for 12 hours.

4.3.3 Nanodiamond Characterization

Dynamic light scattering and zeta potential measurements of ND solutions of 0.1 mg/mL

were measured in single-use polystrene cuvettes or single use zeta potential capillary cells

(Malvern Instruments) on a Malvern Zetasizer Nano-ZS. All measurements were taken at room

temperature using refractive index of bulk diamond for reference (n = 2.419). Measurements

were taken in triplicate and weighted by intensity to account for greater light scattering by larger

particles.

Infrared spectroscopy measurements was performed with 1 mg of ND mixed with KBr

(0.5 g, <8 wt %). The NDs and KBr were ground in a mortar and pestle inside a nitrogen

glovebox. The mixture was loaded into the high-temperature stage of a Thermo Nicolet 6700

FTIR with a Praying Mantis attachment. The chamber was sealed under partial vacuum (>600

Torr) and placed within the beam path of the DRIFTS cell. The chamber was then connected to

another vacuum pump fitted with an liquid N2 cold trap and evacuated to <10–4

Torr. KBr blanks

were taken in an identical manner for each temperature, and all data were processed by

algorithmic baseline correction followed by the Kubelka–Munk transformation so that spectral

intensity was proportional to concentration.

Page 129: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

118

Electron microscopy of NDs was performed by drying 0.001 mg/mL ND solutions on

holey carbon 300 mesh copper electron microscopy grids (Ted Pella). Images were taken on a

JEOL 100CX-II microscope.

4.3.4 Cell Imaging

A vector for pDisplay-eDHFR-GFP was created for co-labeling of extracellular TMP-ND

conjugates with GFP. The gene encoding GFP was amplified using PCR from the H2B-GFP

vector (Addgene Cat#11680) using primers 5' -GAACTCTCACAGCTATTGCTTTGAGA

TTCTGGAGCGGCGGGATCCACCGGTCGCCACCATG -3' and 5'- CATTCAGATCCTCT

TCTGAGATGAGTTTTTGTTCGTCGAC CTTGTACAGCTCGTCCATGC - 3'. The backbone

including the gene for pDisplay-eDHFR was amplified using PCR with the primers 5'-

CCGCCGCTCCAGAATCTCAA-3' and 5'- GTCGACGAACAAAAACTCATC-3' . The final

pDisplay-eDHFR-GFP was assembled by Gibson Assembly (New England Biolabs). 50

HEK 293T cells were cultured in Dulbecco's Modified Eagle Medium (DMEM)

w/ glutamine (Gibco #11995) with 10% v/v fetal bovine serum (FBS) and 1% v/v

Pen/Strep. All cells were maintained under 5% CO2 at 37°C. Cells were plated in 6 well plates at

500,000 cells per well 12 hours before transfection. Cells were transfected with the pDisplay-

eDHFR-GFP expression plas id (2 μg per well) using Xtre e Gene HP (Roche). 12 hours after

transfection, cells were replated in a 1:2 ratio onto 35 mm Fluorodish cell culture dishes. Cells

were incubated with 1 M of TMP-Alexa647 or 2 ng of TMP-NDs for 1 hour, followed by two

washes with fresh media before imaging. Confocal images were obtained using an Zeiss LSM

700 confocal microscope. Images were processed by Zeiss Zen image processing software.

Page 130: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

119

4.4 Results and Discussion

To develop FNDs as fluorophores for applications with the TMP-tag requires controlled

manipulation of FND size and surface characteristics. The ability to isolate small FNDs is critical

for selecting fluorophores whose size is less likely to impact the function of tagged

biomolecules. The surface of the FND must also be covalently modified to include TMP for

selective labeling with the TMP-tag. Ideally, the FNDs would be monofunctionalized to prevent

labeling and aggregation of multiple TMP-tagged targets around one fluorophore. However,

methods to ensure monofunctionalization were not explored due to the overall difficulty of

performing and confirming surface modifications. For our work, we used HPHT NDs without

implanted NV- centers due to their commercial availability. While these NDs did not undergo

high energy ionization and vacuum annealing during the NV- center production process, the

surface properties of NDs before and after NV- implantation are believed to be very similar. We

develop methods to isolate small NDs while concurrently exploring surface chemistry methods

to functionalize NDs with TMP. We also attempt live cell labeling of FNDs with the TMP-tag.

4.4.1 Size Separation

Biochemical methods of protein purification and separation were employed for size-

based separation of colloid unoxidized NDs ranging in size from 0-30 nm. Because complex

mixtures of proteins can be separated by size and molecular weight in biochemistry, these

methods had attractive potential for separating NDs with relatively large size distributions. First,

dialysis was attempted to select for NDs too large to diffuse across the porous membrane.

However, no changes in size distribution between undialyzed and dialyzed ND solutions were

observed. Gel filtration was then attempted to fractionate NDs by size, but also failed to produce

fractions with measurable differences in size distributions. Centrifugation and ultracentrifugation

Page 131: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

120

of NDs in solution produced two fractions consisting of smaller NDs suspended in the

supernatant and a pellet consisting of larger NDs. The looseness of the ND pellet prevented

complete separation between pelleted and supernatant NDs, but the supernatant NDs were easily

separable from the pellet and measured for size distribution using TEM images and DLS analysis

(Figure 4-4 and Table 4-1).

Figure 4-4: Transmission Electron Microscopy of Size Separated Nanodiamonds

Transmission electron microscopy of ND still in suspension after centrifugation, illustrating distributions of smaller

ND sizes with greater centrifugation. a) Starting ND material 0-25 nm, b-f) Size selected fractions after

centrifugation at b) 15,000 x g for 1 hour, c) 15,000 x g for 2 hours, d) 15,000 x g for 3 hours, e) 100,000 x g for 30

minutes, f) 100,000 x g for 1 hour.

Page 132: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

121

Table 4-1: Size and distribution of NDs separated by centrifugation determined using DLS

and TEM analysis

DLS Analysis TEM Analysis

Median (nm) Range (nm) Median (nm) Range (nm)

Starting Material 18.2 11.7 - 58.8 9.7 2.2 - 31.8

15,000 x g for 1 hour 11.17 7.5 - 37.8 7.7 1.9 -15.9

15,000 x g for 2 hours 10.1 6.5 - 32.7 6.1 2.4 - 13.7

15,000 x g for 3 hours 8.72 6.5 - 32.7 6.4 2.3 - 13.6

100,000 x g for 30 minutes 6.5 4.85 - 21 - -

100,000 x g for one 1 hour 4.85 3.12 - 15.7 2.4 1.8 - 4.3

Increasing centrifugation time and speed yields NDs with smaller median size and more

narrow size distribution as observed from the TEM images and in the DLS measurements

(Figure 4-4 and Table 4-1). The TEM images also illustrate the shape irregularity of the NDs,

highlighting the lower accuracy of DLS size measurements that use spherical shape assumptions.

While the size distribution of the starting material measured from DLS is 11.7 - 58.8 nm, the

TEM images report the distribution to be 2.2 - 31.8 nm, which is much closer to the

manufacturer's reported distribution of 0 - 30 nm. Nevertheless, both DLS measurements and

TEM images exhibit the general trend of decreasing size with greater centrifugation.

Both DLS and TEM measurements also illustrate that shorter periods of

ultracentrifugation were more effective than longer periods of normal centrifugation for isolating

smaller NDs (Figure 4-4 and Table 4-1). Increasing centrifugation at 40,000 x g from two to

three hours did not yield as substantial of change in median size and size distribution as observed

in the first or second hour of centrifugation. However, thirty minutes of ultracentrifugation at

100,000 x g isolated NDs with smaller size and size distribution than those obtained using

normal centrifugation. Further exploration of centrifugation speed and time could produce more

ND solutions with more varied median size and size distributions.

Page 133: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

122

The surface chemistry of the unoxidized NDs may explain the difficulty applying

biochemical size separation methods successfully. These NDs are covered in graphitic material

and are stabilized in solution through weak electrostatic interactions. In comparison, the surface

of most biomolecules is highly hydrophilic, making these proteins soluble in aqueous solutions.

As a result, the interaction of NDs with separation materials developed for proteins could be

quite different. To date, there are no other reports of HPHT ND size selection through methods

other than centrifugation, which is consistent with the difficult experienced here testing other

methods.

Effective separation of the NDs to smaller sizes required increasingly higher

centrifugation speed rather than time. This is likely due to the need for greater forces to

overcome electrostatic stabilization and pull down the NDs from suspension. Similar results

were observed using oxidized NDs, with the application of gradient during centrifugation

improving the separation resolution by slowing the progress of pulled down NDs. 19,51

All in all,

centrifugation is a fast and straightforward method to isolate small NDs from size disperse

solutions for biological applications.

4.4.2 Nanodiamond Surface Modification

Following methods previously published in the literature, two surface functionalization

schemes based on carboxylic acid modification were attempted for oxidized 0 - 30 nm NDs

(Figure 4-5).35,36,52

Rather than directly conjugate TMP to NDs in a single step, two multistep

synthetic routes were developed that incorporate the production of functionalized ND

intermediates that can be modified with robust and well-established chemistry. The synthesis of

these ND intermediates allows conjugation to other biomolecules other than just TMP,

broadening the application of NDs for other biological tagging applications. Even more

Page 134: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

123

importantly, these intermediates include a limited number of chemical groups that can be

detected using FT-IR for straightforward confirmation of covalent bond formation. The first

method covalently conjugates propargylamine to ND carboxylic acids using EDC coupling,

forming an amide bond whose presence can be verified by FT-IR (Figure 4-5a). The

incorporated alkyne can also be detected through FT-IR to evaluate successful coupling and later

used to react with azide-TMP or another azide-containing biomolecules. The second method

reduces ND surface carboxylic acids to hydroxyls using borane reduction (Figure 4-5b). These

hydroxy groups are coupled to a aminofunctionalized silane reagent, such as APTES ((3-

aminopropyl)triethoxysilane) (Figure 4-5c), whose successful conjugation can be determined by

the presence of silicon-oxygen vibrational stretches in FT-IR. The incorporated amine is also

detectable using FT-IR and allows later EDC coupling with carboxylic acid-TMP or other

carboxylic acid derivatives.

Page 135: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

124

Figure 4-5: Schemes for nanodiamond functionalization

Synthetic schemes for ND functionalization based on carboxylic acid modification. a) EDC coupling of

propargylamine with ND carboxylic acids yield FT-IR detectable amide and alkyne functional groups. The alkyne

can be modified with any number of azide compounds. b) Boron reduction of ND carboxylic acids yield

hydroxylated NDs. c) Hydroxylated NDs are silanized with (3-aminopropyl)triethoxysilane, producing FT-IR

detectable silicon-oxygen bonds and amine groups. The amine functionalized NDs can be coupled to any number of

carboxylic acid compounds.

The starting material of oxidized NDs was evaluated for the presence of carboxylic acid

groups upon which the proposed surface modification schemes are dependent. 0-30 nm NDs

were oxidized for two hours at 575°C in a tube furnace to remove graphitic material and

subjected to acid-base-acid washing that has previously been reported to solubilize and

carboxylate the NDs surface.36,53

The FT-IR spectrum of this material reveals few features that

indicate the presence of carboxylic acids (Figure 4-6a). Most notably, the characteristic

carboxylic acid ketone and hydroxyl peaks at 1600-1700 cm-1

and 3500 cm-1

, respectively, are

not present. Instead, there is broad and weak absorption at 1410 cm-1

that could possibly be

attributed to C-H stretches of hydrogenated surface carbons. The lack of carboxylic acids is

further supported by the results of attempted EDC coupling with propargylamine (Figure 4-6b).

Although additional peaks are present in the FT-IR spectra after the undergoing the reaction, the

Page 136: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

125

characteristic peaks for amide bond formation are missing. In addition, no alkyne signal in the

2200 cm-1

region is observed. Taken together, the lack of distinctive features in the FT-IR

spectra of these NDs indicates that carboxylation and functionalization by EDC coupling was

either unsuccessful or proceeded to an extent that was undetectable by FT-IR.

Figure 4-6: FT-IR of nanodiamond after surface modification treatments

Diffused Reflectance Infrared Fourier Transform (DRIFT) of NDs in KBr and schematics of their expected

products. a) NDs oxidized at 575°C for 2 hours used as starting material for all reactions, b) NDs after coupling of

propargylamine with ND-carboxylic acids activated with EDC in water, c) NDs after boron reduction in THF, and d)

Boron reduced NDs after aminosilanization with APTES in MeOH.

Despite the lack of confirmed carboxylic acids on the initial starting material, boron

reduction of these NDs appear to produce oxygen-containing hydroxy and ketone groups (Figure

Page 137: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

126

4-6c). The broad peak between 3600-3100 cm-1

that suggests the presence of hydroxyl groups. In

addition, there is also a strong peak at 1600 cm-1

that likely corresponds to a ketone. The origin

of these absorbances is unclear if in fact carboxylic acids are not present on the ND starting

material. One possibility is that hydroxy groups are produced from boron reduction of surface

alkenes that lack characteristic C=C vibrational stretches in the starting material due to their

confinement on the surface. More likely, is that the hydroxyl absorbance originates from

adsorbed water to the ND surface rather than surface functional hydroxy groups.

Silane coupling to hydroxylated NDs would help confirm the presence of surface

hydroxyl groups. If hydroxylated NDs had undergone successful reaction with an aminosilane,

the product is expected to produce characteristic FT-IR peaks in the 3500-3100 cm-1

range and at

1100 cm-1

, corresponding to stretches from the primary amine and silicon-oxygen, respectively.

Instead, its spectrum does exhibit peaks at 1730 cm-1

, 1590 cm-1

, and broad absorbances between

3500-2700 cm-1

and 1500-1000 cm-1

(Figure 4-6d) The broad peak at 3500-2700 cm-1

is likely

attributed to adsorbed water on the surface and may obscure any peaks from amine groups.

While the broad 1500-1000 cm-1

peak could be due to silica, previous reports of silanizations

exhibit well defined 1100 cm-1

absorbances.39,40,54,55

Without the characteristic peaks from the

aminosilane on the FT-IR, the success of silanization and previous hydroxylation cannot be

concluded.

Overall, the FT-IR data on the ND chemical surface modifications is not convincing of

the covalent bond formation to the ND surface. Given the possibility that there are not carboxylic

acids in the starting material, it is in retrospect not surprising that the reactions did not yield

conclusive ND functionalization. The reliance on carboxylic acids for ND surface chemistry

limits the ability to functionalize the surface, especially if those functional groups are not present

Page 138: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

127

or present in very low numbers on the ND surface. Indeed, the lack of carboxylic acids on the

oxidized ND surface is consistent with findings from other groups regarding HPHT NDs.41

In

addition, reports of successful carboxylic acid derivatization with NDs has been limited to a few

research groups, so the presence of these functional groups may not as abundant as initially

reported.35,37,45

The use of HPHT NDs from different manufacturers may also create

discrepancies in ND surface properties. Differing milling techniques that produce NDs with

more or fewer edges that are hypothesized to host functional groups. While HPHT NDs between

different manufacturers has not yet been compared, doing so could reveal the chemical basis for

these different reactivities.

While changes in the FT-IR spectra are observed between reactions, these absorbances

could be due to adsorption of reagents to the ND surface rather than covalent modification.

Although the NDs were thoroughly washed, non-specific adhesion could have withstood

repeated washings, especially if contained within larger ND aggregates. Improved ND dispersion

through ultrasonication during reactions may reduce aggregation and may improve overall ND

reactivity by providing greater exposure to limited surface functional groups. At the same time, if

the strength of these nonspecific interactions is high enough to resistant washing and other

methods of dispersion, then using nonspecific adsorption of targeting molecules may be explored

as an alternative method to associate NDs with specific biomolecules.

4.4.3 Nanodiamond Cell Labeling and Imaging

Despite the uncertainty of surface functionalization, the synthesis of TMP conjugated

NDs was attempted for live cell labeling with the TMP-tag. 0-30 nm unoxidized NDs underwent

boron reduction and silanization with APTES (Figure 4-5b,c). TMP-NDs were then prepared by

EDC coupling to a carboxylic acid TMP (Figure 4-7). Comparison of the FT-IR between

Page 139: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

128

conjugated TMP-NDs and a mixture of TMP and NDs reveals several additional absorbance with

the TMP-NDs that are suggestive of covalent bond formation between the TMP moiety and the

ND (Figure 4-7).

Figure 4-7: FT-IR of TMP-functionalized nanodiamonds

Diffused Reflectance Infrared Fourier Transform (DRIFT) of NDs in KBr and schematic of their expected products.

Top (Purple): Mixture of NDs with TMP. Middle (Green): TMP-conjugated NDs. Bottom (Red): NDs alone. Figure

courtesy of Yongjun Li.

The TMP functionalized NDs were then evaluated for their ability to label extracellular

proteins in live cells using the TMP-tag. Successful labeling would not only demonstrate the

ability of the TMP-tag to label proteins with nanomaterials, but also confirm conjugation of TMP

to the ND surface. TMP-NDs were incubated with live HEK293T cells expressing extracellular

eDHFR-GFP fusion for colocalization of ND fluorescence with GFP fluorescence (Figure 4-8b).

Unfortunately, ND fluorescence is not colocalized to GFP fluorescence in cells indicating

unsuccessful ND labeling. But, attempts to label the same construct with the red organic

Page 140: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

129

fluorophore-ligand, TMP-Alexa647, also failed to produce co-localized fluorescence (Figure 4-

8a). The unsuccessful co-labeling of cell impermeable TMP-Alexa647 indicates that the eDHFR-

GFP fusion lacks TMP-binding capabilities or is not correctly expressed on the outer membrane

of the cell. Therefore, the lack of ND co-labeling with the TMP-tag may or may not reflect

successful TMP-ND conjugation.

Figure 4-8: Extracellular protein labeling with NDs using the TMP-tag in live cells

Live HEK293T cells expressing an extracellular eDHFR-GFP fusion were labeled with a) TMP-Alexa647 and b)

TMP-NDs. Scale bar = 20 m. Images from left to right: (1) green fluorescence with 488nm excitation, (2) red

fluorescence with 639nm excitation, (3) transmitted light, (4) merged image.

Overall cell health was very poor during the ND imaging experiments, complicating

analysis of TMP-ND binding behavior. While NDs have been demonstrated to be biologically

inert to cells, the uptake of highly functionalized NDs may have still adverse effects on cells.56

Although the TMP antibiotic has very high affinity for the eDHFR protein over endogenous

mammalian DHFR, a TMP-ND with high concentrations of the TMP could still bind to

mammalian DHFR and inhibit cell growth. However, rounded and floating dead cells are visible

Page 141: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

130

in cultures that were incubated with and without NDs (Figures 4-8, 4-9). The consistency

between these cultures suggests that the NDs are not necessarily cause of poor cell health, but the

impact of high TMP concentrations from ND should be more thoroughly investigated. Future

experiments should improve and optimize protocols for cell culturing and imaging to preserve

cell health during extracellular protein expression and labeling results.

The high level of TMP-ND aggregation when incubated with cells suggests incomplete

surface modification. The highly soluble TMP should stabilize individual NDs in solution if

there is sufficient surface coverage. Given that size of individual NDs are beyond the diffraction

limit of light, individual NDs in solution should be not visible from transmitted light images and

fluorescence should only be visible as diffraction limited spots. However, large micron sized

aggregates of TMP-NDs are observed (Figure 4-9). These aggregates would be completely

unsuitable for cell imaging due to their bulky size. Further investigation of surface chemistry to

reduce aggregation is worthwhile to develop, keeping ND and ND clusters as small as possible to

minimize perturbations to the biological systems.

Figure 4-9: HEK293T cells with TMP-ND aggregates

Merged transmitted light and red fluorescence (excitation 639nm) of live HEK293T cells incubated with TMP-NDs

for one hour without washing.

Page 142: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

131

The TMP-NDs exhibit very low fluorescence when imaged with confocal microscopy

(Figures 4-8, 4-9). Because the TMP-NDs are unoxidized, the NDs contain a significant graphitic

material on the surface that lowers fluorescence. In addition, the NDs used did not have

implanted NV- centers. so fluorescence originates from naturally occurring NV

- centers. While

small clusters of TMP-NDs are visible when they were incubated with cells, the NDs lacked

strong red fluorescence (Figure 4-8). When cells were incubated with TMP-NDs without

washing excess TMP-NDs from culture, the presence and fluorescence from large ND

aggregates is more plainly visible (Figure 4-9). These images highlight the importance of using

oxidized NDs with implanted NV- centers for bright fluorescence imaging.

4.5 Conclusions and Outlook

FNDs have tremendous potential to revolutionize SM imaging and biosensing in live

cells. However, the barrier to the widespread application of FNDs in biological studies the lack

of reliable surface modification protocols to functionalize FNDs for targeted labeling. As

demonstrated in this chapter, replicating and developing surface modification for FNDs is

challenging due to differences in the synthetic origin of NDs. Despite these difficulties,

advancements in isolating small NDs may improve the surface chemistry reactivity by providing

greater surface area relative to diamond mass. With continued development, small and

functionalized ND fluorophores can be used with chemical tags for selective labeling of

biomolecules.

The applications of FNDs for live cell imaging go beyond conventional fluorescence.

FND biocompatibility and photophysical properties are unmatched in comparison to organic

dyes and other fluorescent nanomaterials. The ability to image individual FNDs over long

periods of time, as well as perform optical tracking, magnetic sensing, and orientation

Page 143: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

132

measurements opens unprecedented opportunities for probing biomolecular mechanism. Creative

applications of FNDs using multi-color imaging based on resonant microwave wavelengths or

coupling to magnetic iron oxide particles for distance measurements can be explored. Along with

simply being brighter and more photostable fluorophores, FNDs are an exciting material to use

for live cell SM imaging

4.6 References

(1) Rabeau, J. R.; Reichart, P.; Tamanyan, G.; Jamieson, D. N.; Prawer, S.; Jelezko,

F.; Gaebel, T.; Popa, I.; Domhan, M.; Wrachtrup, J. Applied Physics Letters 2006, 88.

(2) Schirhagl, R.; Chang, K.; Loretz, M.; Degen, C. L. Annual Review of Physical

Chemistry 2014, 65, 83.

(3) Boudou, J. P.; Curmi, P. A.; Jelezko, F.; Wrachtrup, J.; Aubert, P.; Sennour, M.;

Balasubramanian, G.; Reuter, R.; Thorel, A.; Gaffet, E. Nanotechnology 2009, 20, 235602.

(4) Chang, Y.-R.; Lee, H.-Y.; Chen, K.; Chang, C.-C.; Tsai, D.-S.; Fu, C.-C.; Lim,

T.-S.; Tzeng, Y.-K.; Fang, C.-Y.; Han, C.-C.; Chang, H.-C.; Fann, W. Nat Nano 2008, 3, 284.

(5) Su, L. J.; Fang, C. Y.; Chang, Y. T.; Chen, K. M.; Yu, Y. C.; Hsu, J. H.; Chang,

H. C. Nanotechnology 2013, 24, 315702.

(6) Vlasov, I. I.; Shiryaev, A. A.; Rendler, T.; Steinert, S.; Lee, S.-Y.; Antonov, D.;

Voros, M.; Jelezko, F.; Fisenko, A. V.; Semjonova, L. F.; Biskupek, J.; Kaiser, U.; Lebedev, O.

I.; Sildos, I.; Hemmer, P. R.; Konov, V. I.; Gali, A.; Wrachtrup, J. Nat Nano 2014, 9, 54.

(7) Boudou, J.-P.; Tisler, J.; Reuter, R.; Thorel, A.; Curmi, P. A.; Jelezko, F.;

Wrachtrup, J. Diamond and Related Materials 2013, 37, 80.

(8) Krueger, A.; Lang, D. Advanced Functional Materials 2012, 22, 890.

(9) Dolmatov, V. Y. In Ultrananocrystalline diamond; Gruen, O. A. S. M., Ed.;

William Andrew Publishing: Norwich, NY, 2006, p 347.

(10) Danilenko, V. V. Phys. Solid State 2004, 46, 595.

(11) Greiner, N. R.; Phillips, D. S.; Johnson, J. D.; Volk, F. Nature 1988, 333, 440.

(12) Dolmatov, V. Y. Russian Chemical Reviews 2001, 70, 607.

Page 144: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

133

(13) Ozawa, M.; Inagu a, M.; Takahashi, M.; ataoka, F.; r ger, A.; Ōsawa, E.

Advanced Materials 2007, 19, 1201.

(14) Pentecost, A.; Gour, S.; Mochalin, V.; Knoke, I.; Gogotsi, Y. ACS Applied

Materials & Interfaces 2010, 2, 3289.

(15) Neitzel, I.; Mochalin, V.; Knoke, I.; Palmese, G. R.; Gogotsi, Y. Composites

Science and Technology 2011, 71, 710.

(16) Ivanov, M. G.; Pavlyshko, S. V.; Ivanov, D. M.; Petrov, I.; Shenderova, O.

Journal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures 2010,

28, 869.

(17) Zhu, Y.; Li, J.; Li, W.; Zhang, Y.; Yang, X.; Chen, N.; Sun, Y.; Zhao, Y.; Fan, C.;

Huang, Q. Theranostics 2012, 2, 302.

(18) Liang, Z. Z.; Jia, X.; Ma, H. A.; Zang, C. Y.; Zhu, P. W.; Guan, Q. F.; Kanda, H.

Diamond and Related Materials 2005, 14, 1932.

(19) Morita, Y.; Takimoto, T.; Yamanaka, H.; Kumekawa, K.; Morino, S.; Aonuma,

S.; Kimura, T.; Komatsu, N. Small 2008, 4, 2154.

(20) Trusheim, M. E.; Li, L.; Laraoui, A.; Chen, E. H.; Bakhru, H.; Schröder, T.;

Gaathon, O.; Meriles, C. A.; Englund, D. Nano Letters 2013, 14, 32.

(21) Rabeau, J. R.; Stacey, A.; Rabeau, A.; Prawer, S.; Jelezko, F.; Mirza, I.;

Wrachtrup, J. Nano Letters 2007, 7, 3433.

(22) Fu, C.-C.; Lee, H.-Y.; Chen, K.; Lim, T.-S.; Wu, H.-Y.; Lin, P.-K.; Wei, P.-K.;

Tsao, P.-H.; Chang, H.-C.; Fann, W. Proceedings of the National Academy of Sciences 2007,

104, 727.

(23) Balasubramanian, G.; Lazariev, A.; Arumugam, S. R.; Duan, D.-w. Current

Opinion in Chemical Biology 2014, 20, 69.

(24) Doherty, M. W.; Manson, N. B.; Delaney, P.; Jelezko, F.; Wrachtrup, J.;

Hollenberg, L. C. L. Physics Reports 2013, 528, 1.

(25) Hegyi, A.; Yablonovitch, E. Nano Letters 2013, 13, 1173.

(26) McGuinness, L. P.; YanY; StaceyA; Simpson, D. A.; Hall, L. T.; MaclaurinD;

PrawerS; MulvaneyP; WrachtrupJ; CarusoF; Scholten, R. E.; Hollenberg, L. C. L. Nat Nano

2011, 6, 358.

Page 145: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

134

(27) Gruber, A.; Dräbenstedt, A.; Tietz, C.; Fleury, L.; Wrachtrup, J.; Borczyskowski,

C. v. Science 1997, 276, 2012.

(28) Rittweger, E.; Han, K. Y.; Irvine, S. E.; Eggeling, C.; Hell, S. W. Nat Photon

2009, 3, 144.

(29) Wildanger, D.; Patton, B. R.; Schill, H.; Marseglia, L.; Hadden, J. P.; Knauer, S.;

Schönle, A.; Rarity, J. G.; O'Brien, J. L.; Hell, S. W.; Smith, J. M. Advanced Materials 2012, 24,

OP309.

(30) Arroyo-Camejo, S.; Adam, M.-P.; Besbes, M.; Hugonin, J.-P.; Jacques, V.;

Greffet, J.-J.; Roch, J.-F.; Hell, S. W.; Treussart, F. ACS nano 2013, 7, 10912.

(31) Faklaris, O.; Garrot, D.; Joshi, V.; Druon, F.; Boudou, J.-P.; Sauvage, T.;

Georges, P.; Curmi, P. A.; Treussart, F. Small 2008, 4, 2236.

(32) BradacC; GaebelT; NaidooN; Sellars, M. J.; TwamleyJ; Brown, L. J.; Barnard, A.

S.; PlakhotnikT; Zvyagin, A. V.; Rabeau, J. R. Nat Nano 2010, 5, 345.

(33) Cheng, C. L.; Chen, C. F.; Shaio, W. C.; Tsai, D. S.; Chen, K. H. Diamond and

Related Materials 2005, 14, 1455.

(34) Yin, L. W.; Li, M. S.; Cui, J. J.; Song, Y. J.; Li, F. Z.; Hao, Z. Y. Applied Physics

A Materials Science & Processing 2001, 73, 653.

(35) Liu, Y. L.; Sun, K. W. Applied Physics Letters 2011, 98.

(36) Osswald, S.; Yushin, G.; Mochalin, V.; Kucheyev, S. O.; Gogotsi, Y. Journal of

the American Chemical Society 2006, 128, 11635.

(37) Mochalin, V. N.; Gogotsi, Y. Journal of the American Chemical Society 2009,

131, 4594.

(38) Krueger, A.; Stegk, J.; Liang, Y.; Lu, L.; Jarre, G. Langmuir 2008, 24, 4200.

(39) Rehor, I.; Slegerova, J.; Kucka, J.; Proks, V.; Petrakova, V.; Adam, M. P.;

Treussart, F.; Turner, S.; Bals, S.; Sacha, P.; Ledvina, M.; Wen, A. M.; Steinmetz, N. F.; Cigler,

P. Small 2014, 10, 1106.

(40) Liu, Y.; Khabashesku, V. N.; Halas, N. J. Journal of the American Chemical

Society 2005, 127, 3712.

(41) Wolcott, A.; Schiros, T.; Trusheim, M. E.; Chen, E. H.; Nordlund, D.; Diaz, R. E.;

Gaathon, O.; Englund, D.; Owen, J. S. The Journal of Physical Chemistry C 2014, 118, 26695.

Page 146: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

135

(42) Perevedentseva, E.; Cheng, C.-Y.; Chung, P.-H.; Tu, J.-S.; Hsieh, Y.-H.; Cheng,

C.-L. Nanotechnology 2007, 18, 315102.

(43) Cheng, C. Y.; Perevedentseva, E.; Tu, J. S.; Chung, P. H.; Cheng, C. L.; Liu, K.

K.; Chao, J. I.; Chen, P. H.; Chang, C. C. Applied Physics Letters 2007, 90, 163903.

(44) Weng, M.-F.; Chiang, S.-Y.; Wang, N.-S.; Niu, H. Diamond and Related

Materials 2009, 18, 587.

(45) Chang, B.-M.; Lin, H.-H.; Su, L.-J.; Lin, W.-D.; Lin, R.-J.; Tzeng, Y.-K.; Lee, R.

T.; Lee, Y. C.; Yu, A. L.; Chang, H.-C. Advanced Functional Materials 2013, 23, 5737.

(46) Liu, D. S.; Phipps, W. S.; Loh, K. H.; Howarth, M.; Ting, A. Y. ACS nano 2012,

6, 11080.

(47) Bonasio, R.; Carman, C. V.; Kim, E.; Sage, P. T.; Love, K. R.; Mempel, T. R.;

Springer, T. A.; von Andrian, U. H. Proceedings of the National Academy of Sciences of the

United States of America 2007, 104, 14753.

(48) Long, M. J.; Pan, Y.; Lin, H. C.; Hedstrom, L.; Xu, B. Journal of the American

Chemical Society 2011, 133, 10006.

(49) Huynh, V. T.; Pearson, S.; Noy, J.-M.; Abboud, A.; Utama, R. H.; Lu, H.;

Stenzel, M. H. ACS Macro Letters 2013, 2, 246.

(50) Gibson, D. G.; Young, L.; Chuang, R. Y.; Venter, J. C.; Hutchison, C. A., 3rd;

Smith, H. O. Nature methods 2009, 6, 343.

(51) Mahfouz, R.; Floyd, D. L.; Peng, W.; Choy, J. T.; Loncar, M.; Bakr, O. M.

Nanoscale 2013, 5, 11776.

(52) Börsch, M.; Reuter, R.; Balasubramanian, G.; Erdmann, R.; Jelezko, F.;

Wrachtrup, J. 2009; Vol. 7183, p 71832N.

(53) Habuchi, S.; Ando, R.; Dedecker, P.; Verheijen, W.; Mizuno, H.; Miyawaki, A.;

Hofkens, J. Proceedings of the National Academy of Sciences of the United States of America

2005, 102, 9511.

(54) Bumb, A.; Sarkar, S. K.; Billington, N.; Brechbiel, M. W.; Neuman, K. C.

Journal of the American Chemical Society 2013, 135, 7815.

(55) Vial, S.; Mansu  , C.; Sagan, S.; Irinopoulou, T.; Burlina, F.; Boudou, J.-P.;

Chassaing, G.; Lavielle, S. Chembiochem : a European journal of chemical biology 2008, 9,

2113.

Page 147: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

136

(56) Schrand, A. M.; Huang, H.; Carlson, C.; Schlager, J. J.; Ōsawa, E.; Hussain, S.

M.; Dai, L. The Journal of Physical Chemistry B 2006, 111, 2.

Page 148: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

137

Chapter 5

Development of Imaging Ion Channel Trafficking Assays

Page 149: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

138

5.1 Chapter Overview

The full benefit of the chemical tags is realized when they are used to tag proteins that

cannot be labeled with FPs. Ion channels are known for their poor stability when fused to GFP,

especially during investigations of mutant or misfolded ion channels that can cause rare diseases.

The inability to label these proteins with GFP inhibits the study their trafficking and cellular

localization using high throughput fluorescence imaging. Chemical tags offer an alternative

potential strategy for reliable fluorescence labeling of ion channels for high resolution imaging

assays. The smaller size and better folding properties of the chemical tags compared to FPs can

be used to label ion channels with fewer perturbations. By visualizing ion channels localization

with fluorescence imaging, the chemical tags can greatly facilitate the identification of

therapeutics that can rescue protein trafficking through the cell and be used for drug

development.

In this chapter, we detail the development of a fluorescence imaging assay to identify

compounds that pharmacologically rescue misfolded and mistrafficked ion channels. We explore

the use of the TMP-tag to fluorescently label the hERG potassium ion channel and its mutants

without interfering with protein trafficking. We compare differences between fluorescent

labeling with GFP and the TMP-tag for imaging hERG ion channel localization. I am the main

contributors of this work. Inspiration for this work came from Dr. Stuart Licht. I created the cell

constructs, cultured the cells, and performed imaging with assistance from Mia Shandell.

5.2 Introduction

The mistrafficking of proteins is responsible for many different and rare genetic diseases,

including cystic fibrosis, familial hypercholesterolemia, diabetes mellitus, oestogenesis

imperfecta, and retinitis pigmentosa.1 Certain proteins, such as receptors and ion channels, must

Page 150: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

139

be expressed on the cell membrane for correct cell function. Properly trafficked proteins are first

synthesized in the endoplasmic reticulum by ribosomes, progress through the Golgi apparatus

where they may undergo post-translation modifications, and then placed in the cell membrane.

Mutations, however, can inhibit normal trafficking due to protein misfolding. Instead of reaching

the cell membrane, misfolded proteins are redirected to lysosomes for degradation or accumulate

in the endoplasmic reticulum, causing cellular stress.2,3

The lack of protein activity on the cell

membrane is the pathogenesis of these mistrafficking diseases. One strategy to treat these

mistrafficking diseases is to use pharmacological chaperones that assist in protein folding to

restore conformation, trafficking, and functioning.

Chaperones help misfolded or partially folded proteins assume their correct

conformation. There are three classes of chaperones; molecular, chemical, and pharmacological.

Molecular chaperones, such as heat shock proteins and GroEL, are endogenously expressed by

cells and either actively support protein folding or simply prevent aggregation during folding.4

Chemical chaperones are small molecule osmolytes, such as glycerol and DMSO, which

nonspecifically assist in protein folding. While mechanisms for each osmolyte are slightly

different, chemical chaperones generally improve the solubility and thermodynamic stability of

the folded state.5 Since chemical chaperones are non-specific and require high concentrations,

they are unsuitable as therapeutics. One the other hand, pharmacological chaperones are most

relevant for drug development because they are small molecules that interact and assist in the

folding of a specific protein. A pharmacological chaperone assists with folding by stabilizing the

folded conformation, facilitating folding from an intermediate or unfolded state, or possibly both

(Figure 5-1).6 Pharmacological chaperones that encourage proper protein folding while

maintaining function can rescue protein trafficking and be used as therapeutics.

Page 151: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

140

Figure 5-1: Mechanism of pharmacological chaperone assisted protein folding.

Without the pharmacological chaperone (red), the unfolded state is favored (F to U). The pharmacological

chaperone can either facilitate folding (Uc to Fc), stabilize the folded state (F to Fc) or both. Figure courtesy of Dr.

Stuart Licht

Pharmacological chaperones have already been developed as cystic fibrosis (CF)

therapeutics. Previous treatments for CF were aimed at symptom management without targeting

the disease cause. CF is caused by mutations to the cystic fibrosis transmembrane conductance

regulator (CFTR), which encodes for a chloride ion channel. One of more common mutations to

the CFTR is the F508del mutation, which causes improper protein folding and reduced ion

channel gating activity. Ultimately, the F085del mutation prevents the expression of properly

functioning CFTR chloride ion channels on the cell membrane.7 To develop targeted therapeutics

for the F508del CF genotype, pharmacological chaperones that rescue trafficking and gating

activity of the ion channel were identified. In a study using high throughput voltage clamping,

scientists at Vertex screened over 164,000 small molecules for their ability to restore CFTR

expression. The presence of voltage gated activity identified pharmacological chaperones that

both correctly trafficked ion channels to the cell membrane and restored ion channel function.8

Page 152: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

141

One of the identified molecules, VX-809, has recently been submitted for FDA approval in

2014.9 The success of developing targeted CF therapeutics demonstrates the potential of using

pharmacological chaperones as therapeutics for other ion channel mistrafficking diseases.

5.2.1 Cellular Trafficking of hERG Ion Channels

The success of pharmacological chaperones in CF lends hope to the development of such

therapeutics for genetic long QT type 2 syndrome (LQT2). LQT2 is one of seven types of long

QT syndrome that are all characterized by a prolonged QT interval on cardiovascular EKG,

corresponding to delayed repolarization of cardiac tissue. The prolonged QT interval can cause

arrhythmias that lead to sudden decreases in blood pressure, resulting in fainting, seizures, and

even death. LQT2 is the second most common type of long QT syndrome, affecting 25-30% of

all patients, and is caused by mutations to the human ether related a go-go gene (hERG).10

Over

70 different mutations have so far been identified in LQT2 patients, with the majority of these

mutations causing hERG ion channel misfolding and mistrafficking.11

Like CF, there are

currently no targeted therapeutics for LQT2, raising the potential of pharmacological chaperones

for drug development.

Some compounds have been identified that can rescue hERG ion channel trafficking.

Because properly trafficked hERG undergoes 20 kDa of post-translational glycosylation in the

Golgi apparatus, successful hERG expression on the cell membrane can be evaluated by the

presence of both pre- and post-glycosylated hERG using Western Blot analysis.12

The

compounds astemizole, cisapride, terfenadine, E-4031 and fexofenadine have thus far been

identified as chaperones that rescued trafficking of both N470D and G601S hERG mutants.

However, later electrophysiological evaluations revealed that only fexofenadine was capable of

rescuing both trafficking and ion gated voltage activity.13-15

The other compounds managed to

Page 153: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

142

promote proper folding of mutant hERG ion channels for membrane trafficking, but bind to the

ion channel pore and inhibit activity.16

With only fexofenadine identified as a suitable candidate

for drug development so far, there is an obvious need to discover additional compounds that

rescue both trafficking and activity of the wide array of hERG mutations that affect patients with

LQT2.

In addition, there is another distinct need to identify therapeutics that interfere with wild

type hERG ion channel trafficking. The hERG ion channel has a large pore that is known to bind

to many small molecule drugs, destabilizing the channel and causing mistrafficking.17,18

This

condition induces long QT syndrome in patients and is called acquired long QT (acqLQT). Not

limited to a certain class or family of drugs, acqLQT has been associated with the anti-

protozoical pentamidine, the antidepressant fluoxetine (trade name: Prozac), and the cholesterol-

lowering drug probucol, among others.19

Currently, small molecules are not routinely screened

for acqLQT activity during drug development, despite the side effects of acqLQT being just as

dangerous as those of genetic LQT. The development of a convenient and cheap assay to

evaluate hERG trafficking and localization could significantly improve the ability to identify

either pharmacological chaperones for mutant hERG channels or therapeutics for other diseases

that do not induce acqLQT.

Current methods to evaluate hERG localization and trafficking are not amenable to high

throughput screening, hampering the ability to swiftly screen small molecules as

pharmacological chaperone drug leads. Previous work with CFTR chaperones for CF

therapeutics evaluated a library of over 164,000 compounds.8 This work was accomplished using

high throughput electrophysiology, which is still time consuming and technically challenging to

perform in comparison to other high throughput screening methods.7 With hERG,

Page 154: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

143

pharmacological chaperones were previously evaluated used Western Blots, which require large

cell culture preparations and are also difficult to perform as a high throughput method.20

The

holistic cellular activity of ion channel folding and trafficking requires the development of a cell-

based assay that can be adapted for rapid screening. Because LQT2 is an orphan disease with

multiple mutant phenotypes, screens for hERG pharmacological chaperones would also benefit

from being both cost and time effective. Such an assay could also be applied to the development

of therapeutics for other orphan protein trafficking diseases. Fluorescence imaging is an

appealing option for high throughput screens of ion channel trafficking in live cells because it

can be performed quickly using small volumes. However, labeling hERG and other ion channels

for fluorescence imaging presents a substantial challenge.

5.2.2 Visualizing hERG Ion Channel Trafficking with the TMP-tag

There are limited methods for reliably and conveniently tagging ion channels with

fluorescent labels for imaging. Immunofluorescence labeling of ion channels in fixed cells can be

used, but they have limited applicability in live cells and require expensive antibodies.

Genetically encoded GFP fusions are among the most straightforward for protein labeling in

living cells, but ion channel GFP fusions are known for their poor stability. When fused to

misfolded ion channel mutants, GFP itself can become improperly folded and have inhibited

chromophore formation.21

Consequently, GFP is a poor fluorescent label for investigating

mistrafficking diseases caused by misfolded ion channels. The incorporation of fluorescent

unnatural amino acid is another technique that is widely being explored for direct labeling of ion

channels with minimal perturbations.22

However, the use of fluorescent unnatural amino acid is

very technically challenging to carry out, requiring highly engineered t-RNA synthetases and

many controls to reduce nonspecific incorporation.23

Given these limitations, GFP fusions are

Page 155: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

144

cautiously employed for fluorescence imaging experiments, and subjected to careful evaluations

to ensure preservation of native activity before use.

In the case of the hERG ion channel, various GFP fusions have been created and assessed

for their impact on function and trafficking. N-terminal GFP fusions to wild type hERG have

been demonstrated to inhibit activity and membrane trafficking based on electrophysiology and

fluorescence imaging analysis.24,25

However, N-terminal GFP fusions to the G601S hERG

mutant ion channel showed no impact on trafficking.26

Also, C-terminal GFP fusions shows little

interference with trafficking and activity for both the wild type, N598Q, and N629Q hERG

glycosylation mutant ion channels.25,27

While certain GFP fusions have little impact on hERG

ion channel activity, the difference in activity between wild type and mutant hERG ion channel

using N-terminal GFP fusions raises concerns for mutant dependent fusion behavior.22

The use

of a smaller and more robust fluorescent label enable more convenient hERG ion channel

imaging by eliminating mutant and fusion dependent effects on trafficking and function.

The TMP-tag has the potential to overcome the challenges of fluorescent labeling hERG

and other ion channels. With the TMP-tag, the ion channel would be fused to an E. coli

dihydrofolate reductase (eDHFR) protein domain. The introduction of a cell permeable

fluorescent trimethoprim (TMP) ligand would label the ion channel based on the high affinity

and selective interaction between eDHFR and TMP (Figure 5-2). The eDHFR protein domain

used with the TMP-tag is two-thirds the size of GFP, and its smaller size is less likely to interfere

with ion channel function and trafficking. The eDHFR protein domain is also well behaved with

stable folding, allowing the tag to retain its binding and labeling abilities even when fused to

mutant misfolded ion channels. In addition, the TMP-tag can be used with virtually any

fluorophore that can be chemically attached to TMP without compromising the TMP-eDHFR

Page 156: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

145

affinity.28

The demonstrated live cell labeling and imaging capabilities of the TMP-tag further

demonstrate its compatibility for imaging hERG ion channel trafficking in mammalian cells.28-30

With the TMP-tag, localization and trafficking of fluorescent labeled hERG ion channels can be

straightforwardly analyzed with fluorescence microscopy.

Figure 5-2: Imaging hERG trafficking using the TMP-tag

The TMP-tag can label the hERG potassium ion channel to image cellular trafficking. a) A hERG-eDHFR fusion

protein is labeled by binding of the eDHFR domain to a cell permeable fluorescent TMP ligand. b) Localization of

the hERG-eDHFR is fluorescently imaged in cells. Traffic-competent or rescued hERG will exhibit fluorescence at

the cell membrane, while traffic-deficient hERG will have fluorescence localized only on the endoplasmic

reticulum.

Fluorescence microscopy of TMP-tagged mutant hERG ion channels can be used to

quickly screen for pharmacological chaperones that rescue trafficking to the cell membrane. The

localization and trafficking response of fluorescently labeled hERG ion channels to potential

drugs can be visually assessed, identifying lead compounds for LQT2 drug discovery (Figure 5-

2). While fluorescence imaging is unable to directly probe for the rescue of hERG ion channel

Page 157: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

146

activity, this microscopy-based assay can be used as a preliminary screen that identifies a smaller

subset of compounds that are later evaluated using more demanding high throughput

electrophysiology. With future developments in voltage sensitive fluorophores that can be used

with the TMP-tag, simultaneous analysis of hERG trafficking and activity based on fluorescence

based readout activity could also be accomplished. By allowing fluorescent readout of hERG ion

channel trafficking, the TMP-tag has the potential to significantly speed the discovery of

pharmacological chaperones as therapeutics for LQT2 and other protein mistrafficking diseases.

5.3 Materials and Methods

A vector containing the hERG gene (pcDNA3-hERG) was graciously given by the Kass

laboratory. Constructs for hERG gene expression as fusion proteins were created using Gibson

assembly. A hERG fusion with GFP was created, as well as a fusion with GFP containing a

SGLRST linker. The same was created with eDHFR. The hERG gene was amplified by PCR

using the primers in table for different constructs. The backbone of the published vector

pLM1208 containing eDHFR was amplified using the primers 5'-

CACGCCTACCGCCCATTTGCG - 3' and 5' - ATCAGTCTGATTGCGGCGTTAG - 3'(Miller).

The backbone of the published vector pH2B-GFP (Addgene Cat#11680) containing GFP was

amplified using the primers 5' - CACGCCTACCGCCCATTTGCG - 3' and 5' -

ATGGTGAGCAAGGGCGAGGA - 3'. hERG fusion constructs were assembled using Gibson

Assembly of the PCR amplified hERG gene and the GFP or eDHFR containing vector

backbones.31

All constructs were verified by sequencing.

Page 158: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

147

Table 5-1 : Primers for cloning the hERG gene for Gibson Assembly

Construct hERG gene primers

pcDNA-hERG-GFP 5' - CGTAACAACTCCGCCCCATTGA - 3'

5' - GCACCACCCCGGTGAACAGCTCCTCGCCCTTGCTCACCAT

ACTGCCCGGGTCCGAGCCGT - 3'

pcDNA-hERG-eDHFR 5' - CGTAACAACTCCGCCCCATTGA - 3'

5' - CGATAACGCGATCTACCGCTAACGCCGCAATCAGACTGA

TACTGCCCGGGTCCGAGCCGT - 3'

pcDNA-hERG-linker-

GFP

5' - CGTAACAACTCCGCCCCATTGA - 3'

5' - GCTCCTCGCCCTTGCTCACCAT TGTACTTCGCAATCCACT

ACTGCCCGGGTCCGAGCCGT - 3'

pcDNA-hERG-linker-

eDHFR

5' - CGTAACAACTCCGCCCCATTGA - 3'

5' - CTAACGCCGCAATCAGACTGATTGTACTTCGCAATCCACT

ACTGCCCGGGTCCGAGCCGT - 3'

Vectors for hERG mutants were constructed using Gibson Assembly of PCR amplified

pcDNA3-hERG (see Table 5-2 for primers). Fusions of mutant hERG with GFP and eDHFR

were created using the same protocol and primers above.

Table 5-2 : Primers for mutagenesis of the hERG gene

Construct Primers

pcDNA-hERG:N470D 5' - GACTTCCGCACCACCTACGTC - 3'

5' - GACGTAGGTGGTGCGGAAGTCGATGAGGATGTCCACAATG- 3'

pcDNA-hERG:G601S 5' - AGCCTGGGCGGCCCCTCCATC - 3'

5' - GATGGAGGGGCCGCCCAGGCTGCTGCTGTTGTAGGGTTTG - 3'

pcDNA-hERG:R752W 5' - TGGGCCCTGGCCATGAAGTTC - 3'

5' - GAACTTCATGGCCAGGGCCCAAAGGCAGCCCTTGGTGGCC - 3'

pcDNA-hERG:F805C 5' - TGGGGAGCCTCTGAACCTGT - 3'

5' - ACAGGTTCAGAGGCTCCCCACAGATGTCATTCTTCCCCAG - 3'

HEK 293T cells were cultured in Dulbecco's Modified Eagle Medium (DMEM)

w/ glutamine (Gibco #11995) with 10% v/v fetal bovine serum (FBS) and 1% v/v PenStrep

plated in 6 well plates at 500,000 cells per well 12 hours before transfection. Cells were

transfected with the hERG expression plas id (2 μg per well) using Xtre e Gene HP (Roche).

Page 159: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

148

12 hours after transfection, cells were replated in a 1:2 ratio onto 35 mm Fluorodish cell culture

dishes. All cells were maintained under 5% CO2 at 37°C. For cells expressing TMP-tagged

proteins, cells were incubated with 1 M of TMP-fluorescein diacetate (Active Motif) for 1 hour,

followed by two washes with fresh media before imaging.

For immunofluorescence imaging, cells washed three times with sterile PBS and fixed

using 4% formaldehyde in PBS for 15 minutes at room temperature. Cells were washed three

times with PBS and blocked with 10% BSA, 0.02% Triton X-100, and 0.05% azide in PBS for 1

hour at room temperature. Cells were washed three times with PBS before incubation with rabbit

anti-hERG antibody (Sigma) 1:40 overnight at 4C. Cells were washed three times with PBS

before imaging.

All confocal images were obtained using an Zeiss LSM 700 confocal microscope. Images

were processed by Zeiss Zen image processing software.

5.4 Results and Discussion

The application of the TMP-tag for evaluating trafficking of hERG ion channels requires

demonstrating the TMP-tag is convenient to use, reduces impact on ion channel trafficking,

improves image quality, and labels channels similar selectivity compared to using GFP or

immunofluorescence. Consequently, immunofluorescence images of HEK293T cells expressing

wild type hERG ion channels were first evaluated for labeling specificity and image resolution.

Cells were fixed and immunostained with a fluorescent anti-hERG antibody (Figure 5-3). Both

transfected cells expressing hERG ion channels and non-transfected cells exhibited fluorescence

throughout cell and membrane. Since the hERG ion channel is not endogenously expressed by

HEK293T cells, fluorescence in non-transfected cells is likely due to nonspecific labeling. The

non-transfected cells are less fluorescent and have lower labeling than the hERG transfected

Page 160: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

149

cells, indicating that the antibody does exhibit some preference for binding to the hERG ion

channel. However, the degree of nonspecific labeling observed in this investigation could make

membrane-localized fluorescence from successfully trafficked hERG ion channels difficult to

distinguish in future investigations. Overall, immunofluorescence is not reliable for selectively

labeling and visualizing hERG ion channels localization.

Figure 5-3: Immunofluorescence labeling of hERG ion channels

Confocal microscopy images of fixed and permeabilized HEK293T cells with a) no transfection and b) wild type

hERG ion channel expression are labeled with fluorescent anti-hERG antibody. Scale bar = 20 m. Images from left

to right: (1) green fluorescence (488nm excitation), (2) transmitted light, (3) merged image.

The nonspecific labeling observed with immunofluorescence highlights some of the

challenges of using antibody labeling for evaluating ion channel trafficking.

Immunofluorescence has been widely used for hERG ion channel imaging, but those reports

used in-house produced anti-hERG antibodies that are not commercially available.12,32

These

antibodies as well as other commercially available antibodies not employed in this work could

exhibit higher specificity for hERG and lower nonspecific interactions. In addition, recent reports

Page 161: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

150

indicate that nonspecific antibody labeling is likely not attributed to technical errors performed

during cell fixation or blocking steps, further confirming that the observed labeling is due to poor

antibody selectivity.33

The potentially differing performance of antibodies raises additional

challenges in using immunofluorescence for imaging of hERG ion channel trafficking in cells.

Next, we examined whether hERG ion channels labeled with the TMP-tag would produce

higher resolution cell images and exhibit less trafficking interference compared to hERG ion

channels labeled with GFP. HEK293T cells were transiently transfected with a hERG-GFP

fusion or a hERG-eDHFR fusion, which was labeled with TMP-fluorescein before imaging

(Figure 5-4). Both hERG ion channel fusions have similar fluorescence distributions throughout

the cell and on the membrane. Interestingly, thin cell membrane protrusions are visualized in

cells with TMP-tagged hERG ion channels and not observed in cells with the hERG-GFP fusion.

These thin areas of the membranes could be more easily detected and imaged because the TMP-

tag uses organic fluorophores that are brighter than GFP. As a result, the TMP-tag could more

easily identify successfully trafficking hERG ion channel based on the presence of these cell

features. Alternatively, the smaller size of the TMP-tag may reduce cell stress with trafficking

hERG fusions, allowing the membrane to form more points of cell adhesion. Further study into

differences between cell health and image quality of cells labeled with the TMP-tag and GFP can

be used to more completely determine the advantages of the TMP-tag for visualize hERG ion

channel trafficking.

Page 162: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

151

Figure 5-4: Live cell images of hERG fusion proteins

Confocal microscopy images of live HEK293T cells with a) no transfection, b) wild type hERG expression, c)

hERG-eDHFR expression labeled with TMP-fluorescein, and d) hERG-GFP fusion expression. Scale bar = 20 m.

Images from left to right: (1) green fluorescence (488nm excitation), (2) transmitted light, (3) merged image.

The overall similar localization of hERG ion channel fusions with eDHFR and GFP

illustrates that the TMP-tag is, at the very least, a viable substitute for GFP as a fluorescent label.

The hERG ion channel fusions examined here were both C-terminal fusions, with the C-terminal

GFP fusion having previously been shown to not inhibit hERG trafficking and funcion.25

Like

GFP, the TMP-tag based on a hERG ion channel C-terminal eDHFR fusion also does not

interfere with normal hERG trafficking. Since hERG ion channels with N-terminal GFP fusions

have impacted membrane trafficking, a demonstration of proper localization of hERG ion

channel with N-terminal eDHFR fusion would confirm the benefit of the TMP-tag over GFP.25

The eDHFR domain is two-thirds the size of GFP, so it remains uncertain whether the size

reduction would be sufficient to eliminate impacts on cellular trafficking. More likely, the well

behaved folding of the eDHFR domain in comparison to GFP could reduce destabilizing effects

Page 163: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

152

of the fusion. Additional studies using the TMP-tag in place of GFP fusion-induced trafficking

defects could clearly establish the benefit of the TMP-tag over GFP for ion channel labeling and

imaging.

Finally, the effect of the TMP-tag on labeling and imaging trafficking-deficient hERG

ion channel mutants was examined. Four hERG ion channel misfolding and trafficking mutants

with known mechanisms of rescue, N470D, G601S, R752W, and F805C, were evaluated for

eDHFR and GFP induced impacts on trafficking. HEK293T cells were transiently transfected

with mutant hERG-eDHFR and mutant hERG-GFP fusions and imaged (Figure 5-5 and Figure

5-6). In general, less membrane fluorescence was observed in the hERG trafficking mutants in

both GFP and eDHFR fusions. In addition, cells expressing the hERG trafficking mutants

exhibited rounded cell morphologies that are characteristic of unhealthy or dying cells. In

comparison to cells expressing wild type hERG ion channel fusions, cells expressing mutant

hERG ion channels are considerably less healthy. This is consistent with the known cellular

distress caused by accumulation of hERG trafficking mutants in the endoplasmic reticulum.1

Unfortunately, the rounded shape of unhealthy cells increases the difficulty of differentiating

hERG ion channel localization between the cell membrane and endoplasmic reticulum. The

rescue of cell morphology itself could be used as a indirect indicator of proper hERG localization

that could be further confirmed using fluorescence imaging. Overall, these results indicate that

the TMP-tag has similar labeling capabilities and impacts on ion channel trafficking and cell

health as GFP when used with the hERG ion channel.

Page 164: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

153

Figure 5-5: Live cell images of hERG-eDHFR mutants

Confocal microscopy images of live HEK293T cells with a) no transfection, b) hERG-eDHFR expression, c) hERG-

eDHFR: N470D expression, d) hERG-eDHFR:G601S expression, and e) hERG-eDHFR:F805C expression. All cells

were labeled with TMP-fluorescein. Scale bar = 20 m. Images from left to right: (1) green fluorescence (488nm

excitation), (2) transmitted light, (3) merged image.

Page 165: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

154

Figure 5-6: Live cell images of hERG-GFP mutants

Confocal microscopy images of live HEK293T cells with a) no transfection, b) hERG-GFP expression, c) hERG-

GFP: N470D expression, d) hERG-GFP:G601S expression, e) hERG-GFP:R752W expression, and e) hERG-

GFP:F805C expression. Scale bar = 20 m. Images from left to right: (1) green fluorescence (488nm excitation), (2)

transmitted light, (3) merged image.

Further evaluations are needed to see whether the TMP-tag can better visualize

membrane fluorescence from rescued mutant hERG ion channel trafficking in response to

pharmacological chaperones. At the same time, the impact of the TMP-tag itself also needs to be

assessed for any potential effects on hERG ion channel pharmacological rescue. Because the

eDHFR domain is very stable and well-folded, the TMP-tag may actually act as a chaperone for

mutant hERG ion channel ions. Because this behavior has been previously observed with

eDHFR fusions, the TMP-tag may overselect for small molecules that only partial or minimal

pharmacological chaperoning activity and rescuing capability.34

As a result, the impact of both

the GFP and TMP-tag on folding stability of wild-type and mutant hERG ion channel should be

further studied to determine which label has the least impact.

Page 166: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

155

5.5 Conclusion and Outlook

In this study, labeling hERG potassium ion channels with the TMP-tag did not impact

protein trafficking any differently than using GFP labeling. The ability to label hERG ion

channels with organic fluorophores enabled improved resolution of fluorescent cell membrane

protrusions that could used to rapidly identify properly trafficked hERG to the cell membrane.

Unfortunately, the benefit of small size and stable folding of the TMP-tag for labeling hERG ion

channels and identifying pharmacological chaperones has yet to be determined. Overall, the

ability to image hERG ion channel in live cell with the TMP-tag does present the advantage of

direct observation of trafficking and mistrafficking. With future development, the TMP-tag could

be used as a robust assay to identify pharmacological chaperones of hERG ion channels and

other ion channels responsible for diseases.

The prospect of using the TMP-tag for SM imaging of ion channel in live cells is very

appealing. Protein trafficking is a whole cell phenomena, requiring compartmentalization of

organelles that is simply not feasible to reconstitute in vitro. Live cell labeling and imaging

techniques can provide mechanistic insights into trafficking. In particular, SM imaging of

channels starting from synthesis to membrane localization could offer a more comprehensive

understanding of this complex cellular activity. With the heterogeneity of other biological

systems revealed by SM studies, live cell SM imaging of ion channels would undoubtedly

provide remarkable insights into protein trafficking.

5.6 References

(1) Hebert, D. N.; Molinari, M. Physiological reviews 2007, 87, 1377.

(2) Sitia, R.; Braakman, I. Nature 2003, 426, 891.

Page 167: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

156

(3) Selkoe, D. J. Nature 2003, 426, 900.

(4) Hartl, F. U. Nature 1996, 381, 571.

(5) Perlmutter, D. H. Pediatr Res 2002, 52, 832.

(6) Arakawa, T.; Ejima, D.; Kita, Y.; Tsumoto, K. Biochimica et Biophysica Acta

(BBA) - Proteins and Proteomics 2006, 1764, 1677.

(7) Cai, Z. W.; Liu, J.; Li, H. Y.; Sheppard, D. N. Acta pharmacologica Sinica 2011,

32, 693.

(8) Van Goor, F.; Straley, K. S.; Cao, D.; Gonzalez, J.; Hadida, S.; Hazlewood, A.;

Joubran, J.; Knapp, T.; Makings, L. R.; Miller, M.; Neuberger, T.; Olson, E.; Panchenko, V.;

Rader, J.; Singh, A.; Stack, J. H.; Tung, R.; Grootenhuis, P. D.; Negulescu, P. American journal

of physiology. Lung cellular and molecular physiology 2006, 290, L1117.

(9) Van Goor, F.; Hadida, S.; Grootenhuis, P. D.; Burton, B.; Stack, J. H.; Straley, K.

S.; Decker, C. J.; Miller, M.; McCartney, J.; Olson, E. R.; Wine, J. J.; Frizzell, R. A.; Ashlock,

M.; Negulescu, P. A. Proceedings of the National Academy of Sciences of the United States of

America 2011, 108, 18843.

(10) Wu, C. T.; Nattel, S. The Journal of Physiology 2012, 590, 1311.

(11) Anderson, C. L.; Delisle, B. P.; Anson, B. D.; Kilby, J. A.; Will, M. L.; Tester, D.

J.; Gong, Q.; Zhou, Z.; Ackerman, M. J.; January, C. T. Circulation 2006, 113, 365.

(12) Zhou, Z.; Gong, Q.; Epstein, M. L.; January, C. T. Journal of Biological

Chemistry 1998, 273, 21061.

(13) Zhou, Z.; Gong, Q.; January, C. T. Journal of Biological Chemistry 1999, 274,

31123.

(14) Ficker, E.; Obejero-Paz, C. A.; Zhao, S.; Brown, A. M. The Journal of biological

chemistry 2002, 277, 4989.

(15) Rajamani, S. Circulation 2002, 105, 2830.

(16) Shah, R. R. British Journal of Clinical Pharmacology 2002, 54, 188.

(17) Recanatini, M.; Poluzzi, E.; Masetti, M.; Cavalli, A.; De Ponti, F. Medicinal

Research Reviews 2005, 25, 133.

(18) Finlayson, K.; Witchel, H. J.; McCulloch, J.; Sharkey, J. European Journal of

Pharmacology 2004, 500, 129.

Page 168: Development of Trimethoprim Chemical Tags for Single Molecule Imaging in Live Cells Tracy Y

157

(19) Dennis, A.; Wang, L.; Wan, X.; Ficker, E. Biochemical Society transactions 2007,

35, 1060.

(20) Zhou, Z.; Gong, Q.; Ye, B.; Fan, Z.; Makielski, J. C.; Robertson, G. A.; January,

C. T. Biophysical Journal, 74, 230.

(21) Waldo, G. S.; Standish, B. M.; Berendzen, J.; Terwilliger, T. C. Nat Biotech 1999,

17, 691.

(22) Beene, D. L.; Dougherty, D. A.; Lester, H. A. Current opinion in neurobiology

2003, 13, 264.

(23) Dougherty, D. A. Current Opinion in Chemical Biology 2000, 4, 645.

(24) Huang, N.; Lian, J. F.; Huo, J. H.; Liu, L. Y.; Ni, L.; Yang, X.; Zhou, J. Q.; Li, Z.

F.; Song, T. S.; Huang, C. Cell biology international 2011, 35, 193.

(25) Claassen, S.; Schwarzer, S.; Ludwig, J.; Zünkler, B. J Membrane Biol 2008, 222,

31.

(26) Furutani, M.; Trudeau, M. C.; Hagiwara, N.; Seki, A.; Gong, Q.; Zhou, Z.;

Imamura, S.-i.; Nagashima, H.; Kasanuki, H.; Takao, A.; Momma, K.; January, C. T.; Robertson,

G. A.; Matsuoka, R. Circulation 1999, 99, 2290.

(27) Gong, Q.; Jones, M. A.; Zhou, Z. The Journal of biological chemistry 2006, 281,

4069.

(28) Miller, L. W.; Cai, Y.; Sheetz, M. P.; Cornish, V. W. Nat Meth 2005, 2, 255.

(29) Chen, Z.; Jing, C.; Gallagher, S. S.; Sheetz, M. P.; Cornish, V. W. Journal of the

American Chemical Society 2012, 134, 13692.

(30) Jing, C.; Cornish, V. W. ACS chemical biology 2013, 8, 1704.

(31) Gibson, D. G.; Young, L.; Chuang, R. Y.; Venter, J. C.; Hutchison, C. A., 3rd;

Smith, H. O. Nature methods 2009, 6, 343.

(32) Delisle, B. P.; Anderson, C. L.; Balijepalli, R. C.; Anson, B. D.; Kamp, T. J.;

January, C. T. Journal of Biological Chemistry 2003, 278, 35749.

(33) Buchwalow, I.; Samoilova, V.; Boecker, W.; Tiemann, M. Sci. Rep. 2011, 1.

(34) Hausler, T.; Stierhof, Y. D.; Wirtz, E.; Clayton, C. J Cell Biol 1996, 132, 311.


Recommended