Observation and quantification of protein production in single living cells

171
Observation and quantification of protein production in single living cells by Ibrahim Kays Integrated Program in Neuroscience McGill University, Montréal November, 2016 A thesis submitted to McGill University in partial fulfilment of the requirements of the degree of Doctor of Philosophy © Ibrahim Kays, 2016

Transcript of Observation and quantification of protein production in single living cells

Page 1: Observation and quantification of protein production in single living cells

Observation and quantification of protein production

in single living cells

by

Ibrahim Kays

Integrated Program in Neuroscience

McGill University, Montréal

November, 2016

A thesis submitted to McGill University in partial fulfilment of the requirements of

the degree of Doctor of Philosophy

© Ibrahim Kays, 2016

Page 2: Observation and quantification of protein production in single living cells

ii

Abstract

Accurate quantification of protein production is fundamental to understanding basic

molecular and cellular processes. Dysregulation of protein levels can harm cells and lead to

diseases such as cancer and neurodegenerative diseases. However, little is known about how

protein production in a cell changes over time and in response to external factors. The current

assays used to quantify protein production are invasive, time consuming, and have poor

resolution. As a result, researchers have turned to mRNA expression as a measure for protein

abundance, although this has been demonstrated to be inaccurate.

To address these issues, my thesis explores new tools and techniques I developed to

monitor protein production in single living cells. By simultaneously examining the levels of

mRNA and protein of a gene from a single cell, I describe a system used to determine how

individual cells vary in their transcriptional and translational landscapes, and demonstrate the

low predictive power of mRNA levels over protein abundance.

My second approach to understand protein production is aimed at directly observing

protein synthesis in living cells. I describe the generation of animals used for imaging protein

production in single cells in real time. I also describe a system that uses the reconstitution of split

GFP as a spatial and temporal quantitative marker of local protein synthesis. I used single-cell

quantitative imaging, electrophysiology and immunocytochemistry to demonstrate that proteins

produced with split GFP reporters function properly, and that their level of production correlates

with the intensity of the reconstituted GFP signal.

The experiments presented in this thesis demonstrate tools I developed to probe protein

production with high spatial and temporal resolution. The tools and reagents are accessible to a

wide range of researchers and the assays provide high accuracy and reliability. Protein analysis

in single cells can reveal unprecedented insight into the dynamics of the gene expression.

Page 3: Observation and quantification of protein production in single living cells

iii

Résumé

La quantification précise de la production de protéines est un outil fondamental pour

comprendre les processus cellulaires et moléculaires de base. La dérégulation des niveaux de

protéines peut endommager les cellules et mener à des maladies telles que le cancer et les

maladies neurodégénératives. Cependant nous en savons très peu sur la façon dont la production

de protéine change dans une cellule avec le temps ou en réponse à des facteurs externes. Les

méthodes actuelles de quantification de la production de protéines sont invasives, prennent

beaucoup de temps et ont une mauvaise résolution. En conséquence les chercheurs se sont

tournés vers l'expression de l'ARNm afin de mesurer l'abondance de protéines, bien qu'il ait été

démontré que cette méthode est imprécise.

En réponse à ce problème, mon travail de thèse explore de nouveaux outils et techniques

que j'ai développés afin de mesurer la production de protéines dans une cellule vivante. En

examinant simultanément les niveaux d'ARNm et de protéines pour un gène dans une cellule

unique, je décris un système qui peut être utilisé pour déterminer les variations de cellule à

cellule dans les paysages transcriptionel et traductionel, et je démontre la capacité faible du

niveau l'ARNm à prédire l'abondance de protéines.

Ma seconde approche visant à comprendre la production de protéines et dirigée

directement vers l'observation de la synthèse de protéines dans des cellules vivantes. Je décris

l’élaboration d'animaux utilisés pour capturer la production de protéines dans une cellule unique

en temps réel. Je décris aussi un système qui utilise la reconstitution d'une protéine fluorescente

verte (PFV) découpée comme marqueur quantitatif spatial et temporel de la synthèse locale de

protéines. J'ai utilisé de l'imagerie quantitative de cellules uniques, de l'électrophysiologie et de

l'immunocytochimie afin de démontrer que les protéines produites avec comme reporter la PFV

découpée fonctionnent correctement, et que leur niveau de production sont en corrélation avec

l'intensité du signal de la PFV reconstituée.

Les expériences présentées dans cette thèse démontrent les outils que j'ai développés afin

d'examiner la production de protéines avec une haute résolution spatiale et temporelle. Les outils

et réactifs chimiques sont accessibles à une grande variété de chercheurs, et cette méthode

fournit une haute précision et fiabilité. L'analyse de protéines dans des cellules uniques peut

révéler une connaissance approfondie sans précédent de la dynamique de l'expression génique.

Page 4: Observation and quantification of protein production in single living cells

iv

Acknowledgements

My experience in Dr. Brian Chen’s laboratory has been nothing short of transformative. I

was an undergraduate with no research experience when Brian provided me an opportunity to

join his team. He patiently guided me through this journey, helping develop my skills both in the

lab and outside along the way. My gratitude also goes to Drs. Don Van Meyel, David Stellwagen

and Keith Murai for invaluable advice and guidance, as well as to members of my advisory

committee Drs. Artur Kania and Hiroshi Tsuda.

I would like to thank all my colleagues at the Centre for Research in Neurosciences. A

special thank you to Drs. Tiago Ferreira, Emily Peco, Todd Farmer and Haider Al-Timimi for

hallway chats that taught me more lessons than seminars. To Sejal Davla for constantly sharing

her expertise in everything from fly food to Indian food. Many thanks to Chris Salmon, Benny

Kacerovsky and Dr. Gael Quesseveur for help with mouse work. I will always cherish the

memories of the breakfasts I had every Thursday with Hunter Shaw, charmer of the sixth floor

volunteers.

I am indebted to Dr. Chiu-An Lo, who watched and helped me grow since my undergraduate

years, and together with Tsung-Jung Lin taught me most of what I know about molecular

biology. A special thank you to Dr. Farida Emran, the jane of all trades who also coached and

helped me with every aspect of my graduate work.

Last but certainly not least, to my guide and co-conspirator Dr. Vedrana Cvetkovska, you

have helped realize this work in more ways than you know. I dedicate this thesis to my parents

Dima and Anwar, my sister Yasmina and my pug Winston, for your unconditional love, support

and prodding.

Page 5: Observation and quantification of protein production in single living cells

v

Contribution of authors

This thesis is presented in thesis-based format in accordance with McGill University

Graduate and Postdoctoral Studies guidelines. It comprises original work from one published

manuscript, one submitted manuscript and one currently in preparation.

The work in this thesis is based on the Protein Quantification Ratioing (PQR) technique

that I co-developed with Dr. Chiu-An Lo, published as:

Lo C*, Kays I*, Emran F, Lin T-J, Cvetkovska V, Chen BE. Quantification of Protein Levels in

Single Living Cells. Cell Reports. 2015;13(11):2634-2644. doi:10.1016/j.celrep.2015.11.048.

Brian E. Chen designed the experiments and supervised the project. Chiu-An Lo, Ibrahim Kays,

Farida Emran, Tsung-Jung Lin, Vedrana Cvetkovska and Brian E. Chen performed experiments

and analyzed the data. Chiu-An Lo, Ibrahim Kays, and Brian E. Chen wrote the manuscript.

The published technique, to which I contributed 4 years of my graduate work, constitutes Dr.

Lo’s PhD thesis work, obtained under Dr. Brian Chen in 2016, and its development and

validation are outside of the scope of my thesis. In this thesis I use our published technique,

described throughout the thesis and in detail in Chapter 1, as a stepping stone for the

development of novel systems and techniques. As per the McGill University Graduate and

Postdoctoral Studies guidelines, I have obtained written consent from Dr. Lo to describe the

technique and include it as a resource for my work.

A modified version of Chapter 2 has been submitted for publication as:

Kays I, and Chen BE. Protein and RNA quantification in single cells, submitted

Page 6: Observation and quantification of protein production in single living cells

vi

Ibrahim Kays and Brian Chen designed the experiments. Ibrahim Kays collected and analyzed all

the data. Ibrahim Kays wrote the manuscript. Brian Chen supervised the study.

A modified version of Chapter 3 is in preparation for publication as:

Kays I, and Chen BE. Direct observation of local protein synthesis in vivo, submitted

Ibrahim Kays and Brian Chen designed the experiments. Ibrahim Kays collected and analyzed all

the data. Ibrahim Kays drafted the manuscript. Brian Chen supervised the study.

Page 7: Observation and quantification of protein production in single living cells

vii

Table of contents

Abstract .......................................................................................................................................... ii

Résumé .......................................................................................................................................... iii

Acknowledgements ...................................................................................................................... iv

Contribution of authors ................................................................................................................ v

Table of contents ......................................................................................................................... vii

List of figures ................................................................................................................................. x

List of abbreviations .................................................................................................................... xi

Chapter I - Current state of quantification of protein production .......................................... 1

1.1 Introduction ........................................................................................................................... 2

1.2 Protein structure .................................................................................................................... 3

1.2.1 The primary structure of a protein is its linear chain ...................................................... 4

1.2.2 Secondary structures interconnect and stabilize protein residues .................................. 5

1.2.3 Protein folding and maturation are prerequisite to function ........................................... 6

1.3 Protein function ................................................................................................................... 11

1.3.1 Regulation of protein function ...................................................................................... 11

1.3.2 Regulated protein synthesis and degradation ............................................................... 12

1.3.3 Protein phosphorylation ................................................................................................ 13

1.3.4 Regulated translation of localized mRNAs .................................................................. 14

1.3.5 Local translation of mRNA shapes development ......................................................... 16

1.3.6 Local translation of mRNA in neurons ......................................................................... 18

1.4 Quantification of gene expression and protein levels ......................................................... 20

1.4.1 Quantification of mRNA levels .................................................................................... 22

1.4.2 Quantification of protein levels .................................................................................... 23

1.4.3 mRNA levels as proxy for protein abundance .............................................................. 25

1.5 Fluorescence-based single cell resolution protein quantification ....................................... 26

1.5.1 Quantification of protein levels in single living cells ................................................... 27

1.5.2 Protein production reporters must be carefully chosen ................................................ 29

1.5.3 Approaches to visualizing locally translated proteins .................................................. 30

1.5.4 Requirements for a local protein synthesis reporter ..................................................... 34

1.5.5 Split GFPs are indicators of protein interaction ........................................................... 35

1.6 Figures ................................................................................................................................. 37

Page 8: Observation and quantification of protein production in single living cells

viii

1.7 Thesis introduction .............................................................................................................. 40

Chapter II - Quantification of mRNA and protein levels in single cells ................................ 41

2.1 Relation to overall thesis ..................................................................................................... 41

2.2 Introduction ......................................................................................................................... 42

2.3 Experimental design and detailed protocol ......................................................................... 44

2.3.1 Materials ....................................................................................................................... 45

2.3.2 Gene editing using CRISPR-Cas9 ................................................................................ 47

2.3.3 Single-cell protein level quantification ......................................................................... 49

2.3.4 Total RNA extraction ................................................................................................... 49

2.3.5 Reverse-transcription .................................................................................................... 50

2.3.6 Real-time polymerase chain reaction ........................................................................... 51

2.3.7 Calculation of absolute mRNA transcript number ....................................................... 52

2.3.8 Readout of amplification .............................................................................................. 53

2.3.9 Assay controls............................................................................................................... 54

2.3.10 Image acquisition and analysis ................................................................................... 55

2.4 Results ................................................................................................................................. 56

2.5 Discussion ........................................................................................................................... 59

2.6 Conclusion ........................................................................................................................... 62

2.7 Figures ................................................................................................................................. 64

Chapter III - A system for direct observation of subcellular protein translation in single

living cells. .................................................................................................................................... 79

3.1 Relation to overall project ................................................................................................... 79

3.2 Introduction ......................................................................................................................... 80

3.3 Materials and Methods ........................................................................................................ 83

3.3.1 Protein Quantification Reporter constructs .................................................................. 83

3.3.2 Split GFP DNA constructs ........................................................................................... 83

3.3.3 GFP1-10 protein production and extraction ................................................................. 84

3.3.4 GFP 11 peptides............................................................................................................ 85

3.3.5 Cell culture ................................................................................................................... 85

3.3.6 In vitro protein reconstitution ....................................................................................... 86

3.3.7 Endoplasmic reticulum and ribosome staining ............................................................. 86

3.3.8 Electrophysiology ......................................................................................................... 87

3.3.9 Image acquisition and analysis ..................................................................................... 87

Page 9: Observation and quantification of protein production in single living cells

ix

3.3.10 Statistical analysis....................................................................................................... 88

3.4 Results ................................................................................................................................. 89

3.4.1 GFP11 and GFP1-10 reconstitute spontaneously in vitro ............................................ 89

3.4.2 GFP reconstitution in vitro occurs at millisecond timescales ....................................... 90

3.4.3 GFP11 detects GFP1-10 in living cells ........................................................................ 91

3.4.4 GFP reconstitution can report sites of protein translation ............................................ 92

3.4.5 Proteins co-translated with GFP11 reporters function properly ................................... 94

3.4.6 GFP reconstitution can quantitatively readout protein translation ............................... 95

3.5 Discussion and conclusions ................................................................................................. 96

3.5.1 GFP1-10 fluorophore maturation ................................................................................. 97

3.6 Figures ............................................................................................................................... 100

Chapter IV - Applications and future directions of protein quantification using PQR ..... 109

4.1 Relevance to overall project .............................................................................................. 109

4.2 Applications of optical protein quantification using PQR ................................................ 110

4.2.1 Dynamic observation of protein synthesis in vivo ..................................................... 110

4.2.2 Optical normalization of protein production in vivo .................................................. 112

4.3 Split GFP as a quantitative marker of local protein synthesis in vivo .............................. 114

4.3.1 Generation of animals constitutively expressing GFP1-10 ............................................ 115

4.3.2 Local translation of Gurken protein in Drosophila oocytes ........................................... 117

4.3.3 Detection of local protein translation in living neurons ................................................. 121

4.4 Detection of local protein synthesis using PQR photoconvertible reporters. ................... 123

4.5 Figures ............................................................................................................................... 128

Chapter V - Thesis directions and conclusions ...................................................................... 137

References .................................................................................................................................. 142

Page 10: Observation and quantification of protein production in single living cells

x

List of figures

Figure 1.1 Brightfield image of dissected whole ovarioles ......................................................... 37

Figure 1.2 PQR reporters allow quantification of protein production from cells. ....................... 39

Figure 2.1 Workflow of protein and mRNA measurement from the same cell. .......................... 64

Figure 2.2 Protein and mRNA measurement for multiple genes in a single cell. ........................ 66

Figure 2.3 Insertion of PQR-XFP reporters into the endogenous genomic loci of IgK and Rpl13a

using CRISPRs.............................................................................................................................. 68

Figure 2.4 Validation of CRISPR-mediated insertion of PQR-GFP in the endogenous IgK locus.

....................................................................................................................................................... 69

Figure 2.5 Illustration of the important steps and typical equipment used in the protocol. ......... 70

Figure 2.6 Titration of starting input cDNA volume. .................................................................. 72

Figure 2.7 Standard curve of serially diluted known amount of Rpl13a target. .......................... 73

Figure 2.8 Contamination of RNA sample quantification by genomic DNA can be assessed

using no-RT control reaction. ....................................................................................................... 74

Figure 2.9 Endogenous RNA and protein quantification from single cells. ................................ 75

Figure 2.10 Protein and mRNA relationships between multiple genes in single cells. ............... 76

Table 2.1 Sequences of primers and probes used in this protocol. .............................................. 78

Figure 3.1 Stoichiometric production of GFP11 reporters using PQR. ..................................... 100

Figure 3.2 In vitro characterization of the split GFP reconstitution reaction. ........................... 101

Figure 3.3 Reconstitution of split GFP occurs on the order of milliseconds in vitro. ............... 102

Figure 3.4 Split GFP reporters can be expressed using PQRs and the reconstitution of GFP

marks the presence GFP1-10 protein. ......................................................................................... 103

Figure 3.5 Split GFP reconstitution occurs at sites of active protein translation. ...................... 106

Figure 3.6 Co-translation of GFP11 reporters using PQR preserves the protein of interest’s

localization and function. ............................................................................................................ 107

Figure 4.1 PQR reporters are inserted in-frame into endogenous genes. .................................. 128

Figure 4.2 PQR constructs injected into mouse embryos result in red fluorescent pronuclei. .. 129

Figure 4.3 SplitGFP as a protein translation reporter. ............................................................... 131

Figure 4.4 GFP1-10 is expressed at high levels in transgenic animals. ..................................... 132

Figure 4.5 GFP11 can detect Gurken local translation in oocytes. ............................................ 134

Figure 4.6 Novel split fluorescent reporters exhibit more efficient reconstitution. ................... 136

Page 11: Observation and quantification of protein production in single living cells

xi

List of abbreviations

22c10 IgK-secreting mouse hybridoma cell line

2A Self-cleaving 2A peptide

3D Three dimensional

a.u. Arbitrary units

ActB Beta-actin

AMPA α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid

Arc Activity-regulated cytoskeleton-associated protein

ATP Adenosine triphosphate

BDNF Brain-derived neurotrophic factor

BFP blue fluorescent protein

bp Basepair

BSA Bovine serum albumin

cAMP Cyclic adenosine monophosphate

CCD Charge-coupled device

cDNA Complementary deoxyribonucleic acid

CHYSEL Cis-acting Hydrolase Element

CO2 Carbon dioxide

CRISPR Clustered regularly interspersed palindromic repeats

Ct Cycle threshold

ddH2O Double distilled water

DNA Deoxyribonucleic acid

DSB Double-strand break

E. coli Escherichia coli

ELISA Enzyme-linked immunosorbent assay

ER Endoplasmic reticulum

FACS Fluorescence-assisted cell sorting

FMRP Fragile X mental retardation protein

FP Fluorescent protein

GFP (sfGFP) Green fluorescent protein (superfolder GFP)

Gria1/GluR1 Glutamate ionotropic receptor AMPA type subunit 1

Grk Gurken

gRNA (sgRNA) Guide RNA

HEK293/T Human embryonic kidney 293 (Transformed)

hnRNP-R Heterogeneous nuclear ribonucleoprotein R

Ig Immunoglobulin

IgK Immunoglobulin light chain kappa

IHC Immunohistochemistry

IRES Internal ribosomal entry site

I-V Current-voltage

kDa Kilodalton

LTP Long-term potentiation

MAP1b Microtubule-associated protein 1b

MAP2 Microtubule-associated protein 2

mGluR1 Metabotropic glutamate receptor 1

mGRASP Mammalian GFP reconstitution across synaptic partners

Page 12: Observation and quantification of protein production in single living cells

xii

mRNA Messenger ribonucleic acid

NGF Nerve growth factor

nls Nuclear localization signal

nols Nucleolar localization signal

OMP Orotidine 5’phosphate decarboxylase

OPT GFP11-OPT, optimized

pCAG Promoter from Cytomegalovirus, beta-Actin and beta Globin genes

PCR Polymerase chain reaction

PEST Sequences containing Proline, Glutamic acid, Serine and Threonine

pJFRC Janelia Farms Research Center promoter

PQR Protein quantification reporter

PSD95 Postsynaptic density protein 95

qPCR Quantitative real time PCR

R2 Pearson’s linear regression correlation coefficient of determination

RFP Red fluorescent protein

RhoA Ras homolog family member A

RNA Ribonucleic acid

RNase Ribonuclease

ROI Region of interest

RPL13A Human ribosomal protein L13A

Rpl13a Mouse ribosomal protein L13a

RT Reverse-transcription

RT-qPCR Reverse-transcription real-time polymerase chain reaction

SDS-PAGE Sodium dodecyl sulfate-polyacrylamide gel electrophoresis

sec Second(s)

SMN1 Survival of motor neuron 1

UAS Upstream activator sequence

UTR Untranslated region

VegT VegT protein

XFP Generic fluorescent protein (any color)

αCaMKII Alpha Ca2+/Calmodulin-dependent protein kinase II

Page 13: Observation and quantification of protein production in single living cells

1

Chapter I - Current state of quantification of protein

production

Page 14: Observation and quantification of protein production in single living cells

2

1.1 Introduction

Proteins execute almost all fundamental cellular functions. They are the products of

genetic information and the molecular actors that affect change within a cell. To begin and

continue, life requires information and the ability to control this information. A hallmark of

organisms is the exquisite control that has evolved to ensure that biological processes, including

protein production, only occur when and where they are meant to.

Most chemical reactions in living systems are catalyzed by proteins. Proteins lower the

activation energy of cellular reactions, allowing them to proceed at faster rates. A classic and

illustrative example is the Orotidine 5’ phosphate (OMP) decarboxylase, which has been shown

to accelerate uncatalyzed reactions by a factor of 1017 (Radzicka & Wolfenden, 1995). To put

this into perspective, OMP decarboxylase does in 18 milliseconds what would otherwise require

78 million years to spontaneously occur (Callahan & Miller, 2007). Since all cellular function

involves defined chemical reactions, all cellular functions directly involve or require proteins.

Many proteins can be grouped into a few broad functional categories. Structural proteins give

cells shape and form highways within cells that direct the intracellular transport and localization

of various molecules and organelles. Regulatory proteins act as sensors or signals that orchestrate

and fine-tune cellular processes. Signaling proteins relay information from one cellular location

to another. Membrane transport proteins permit the flow of ions and molecules across cellular

membranes. Finally, protein enzymes form and break covalent bonds, allowing chemical

reactions to occur (Alberts et al., 2002; Lodish et al., 2008).

Proteins are the result of translation of messenger RNA by ribosomes, which are large

and complex molecular machines that link amino acids to form peptides and proteins according

to genomic instructions encoded by the mRNA sequence. The rate of synthesis of a protein is

Page 15: Observation and quantification of protein production in single living cells

3

dependent on the rate of transcription of the gene into mRNA, the steady-state abundance of the

mRNA within the cell, and the rate of translation or rate at which the mRNA is converted into

protein (Lodish et al., 2008). However, despite the widespread use of mRNA as a proxy measure

for protein abundance, tools that instead directly assay protein production unarguably offer more

insight into the regulation of the proteome, and its consequences.

In this thesis, I describe my efforts to understand when and where proteins are produced

by developing tools to monitor protein production in single cells. In Chapter 2, using a novel

technique I present results that demonstrate that mRNA levels do not necessarily correlate with

protein production, particularly in single cells. Therefore, Chapters 3 and 4 describe tools I have

developed to observe protein production in single living cells with high spatial and temporal

resolution. The work presented in this thesis provides a framework and resources that will allow

for monitoring of global and local protein synthesis in vivo.

1.2 Protein structure

Louis Sullivan is perhaps best known among scientists as the architect who proposed the

dicta “form follows function” (L. H. Sullivan, 1896), which is true for most man made structures.

However, in evolutionary and protein science, the reverse is true (Pauwels & Tompa, 2016). Key

to the elucidation of how proteins execute specific functions is the concept that protein function

is derived from three-dimensional (3D) structure. Therefore, a prerequisite to understanding and

manipulating protein function is understanding the structure of proteins. One of the best

examples illustrating this approach is with the green fluorescent protein (GFP) (Prasher et al.,

1992; Ward et al., 1980) .

GFP is arguably the most widely studied and known fluorescent protein. The fact that it is

genetically encoded and can autocatalytically develop the green fluorescent signal make it a tool

Page 16: Observation and quantification of protein production in single living cells

4

that has helped advance our understanding of several areas of biology in a revolutionary way,

allowing us to literally observe the interior of living cells and visualize cellular processes as they

occur. Since its discovery and cloning from the jellyfish Aequoria victoria, GFP has been the

subject of many experiments aimed at improving many of its properties including stability and

fluorescence emission.

The common variants of GFP used today offer a number of significant improvements

over the original wild type version, which was dim in brightness, unstable at 37°C, and had two

excitation peaks with the dominant one in the ultra-violet range (Prasher et al., 1992; Tsien,

1998). Over the years many mutations have been introduced and hundreds of variants of GFP are

now available for a wide range of applications (Kent et al., 2008; Tsien, 1998; Zimmer, 2002).

Understanding the structure of GFP was crucial to enable the rational design of mutations that

predictably altered its properties. For example, the elucidation of the GFP crystal structure in

1996 allowed researchers to design new versions of the protein in a wider range of colors,

brightness and stability (Remington, 2011; Tsien, 1998). In this thesis, I developed tools to

monitor protein translation in single living cells using fluorescent reporters, such as GFP. The

well-known history and properties of GFP allow us to exploit different characteristics to develop

new assay-tailored tools using GFP. In the following section I will describe aspects of protein

structure in the context of GFP.

1.2.1 The primary structure of a protein is its linear chain

The primary structure of a protein is the simple linear arrangement or sequence of its

amino acid constituents. At this fundamental level, proteins are constructed by the

polymerization of 20 different types of amino acid building blocks. Amino acids have an amine

(N) group and a carboxyl (C-O) group on either end, and peptide bond formation between the

Page 17: Observation and quantification of protein production in single living cells

5

amine group of one amino acid and the carboxyl group of another amino acid results in the net

release of one water molecule and their linkage via a peptide bond.

The average molecular weight of an amino acid in a protein, given average relative

abundance, is ~ 113 Da (Daltons). Therefore, the number of residues in a protein can be

estimated from the molecular weight of the protein, and vice versa. The Dalton mass unit is used

to report the size of a protein, with 1 Dalton = 1 atomic mass unit (or molecular weight) (Alberts

et al., 2002; Lodish et al., 2008). For example, GFP is a relatively small 238 amino acid protein

that has a molecular mass of 26.9 kDa, in contrast to large immunoglobulin antibody molecules

which are on average 150 kDa.

1.2.2 Secondary structures interconnect and stabilize protein residues

Protein secondary structures are stable spatial arrangements of segments or domains of a

protein chain. They are formed and linked together by non-covalent (mainly hydrogen) bonds

between backbone oxygen and hydrogen atoms. Repeating secondary structures are often found

along the protein chain, and a single protein may contain several types of secondary structures in

several parts of the polypeptide chain, depending on the amino acid sequence in that region. The

main types of secondary structures are the alpha helix, the beta sheet and the beta turn.

In an average protein, roughly 60% of the chain forms alpha helices and beta sheets.

Within an alpha helix, the tight spiralling of residues serves to hold portions of the protein

backbone in a rigid, rod-like cylindrical structure. Alpha helices are usually abundant in proteins

present in cell membranes, where their transmembrane domain is largely a straight alpha helix

that traverses the plasma membrane.

Beta sheets are composed of laterally packed beta strands, which in turn are composed of

5- to 8-residue stretches of fully extended polypeptide segments. The core of many proteins is

Page 18: Observation and quantification of protein production in single living cells

6

composed of regions rich in beta sheets, and these regions can form either between adjacent

domains along the polypeptide chain or between domains far apart within the sequence that can

fold back on themselves and form dense beta sheets. Both types are held together by extensive

hydrogen bonding and this results in highly rigid structures.

The GFP molecule is barrel shaped, consisting of 11 beta strands and a central coaxial

helix, with the chromophore forming from the central helix (Ormö et al., 1996). The beta sheets

shield the chromophore from the outer environment and this is important as removal of

individual beta sheets results in loss of fluorescence emission (Chapters 3 and 4). In addition, the

abnormal oligomerization of protein domains into beta sheet-rich structures is associated with

several pathological states, as epitomized by the implication of oligomerized amyloid beta

protein as one cause of Alzheimer’s disease (Nelson et al., 2005).

The transition of proteins from their linear string of residues into the three-dimensional

world of molecules enables proteins to have functions. The addition of salt bridges, disulfide and

hydrogen bonds, as well as the tight packing of side chains lock protein domains into place,

giving soluble proteins such as antibodies and GFP a compact globular three-dimensional

structure. The folding of a straight protein chain into the tertiary and quaternary structures

observed in living cells is a prerequisite and crucial step to “switch on” the function of proteins.

1.2.3 Protein folding and maturation are prerequisite to function

Thousands of types of proteins exist in every organism (Dobson, 2001). Following

synthesis by the ribosome, each protein must fold into the particular conformational shape

dictated by its sequence, to carry out its function. The Levinthal paradox is a thought experiment

that proposed it would take longer than the age of the universe for even a short polypeptide to go

Page 19: Observation and quantification of protein production in single living cells

7

through all possible conformations before arriving at the correct (lowest energy) structure

(Levinthal, 1969). Yet in the cell, protein folding takes place in seconds (Piana et al., 2013).

In Chapters 2 and 3, I use GFP as a reporter of antibody production by co-expressing it

with the kappa light chain of mouse antibodies. In the next sections I describe folding and

maturation of GFP and mammalian immunoglobulin antibodies.

1.2.3a Green fluorescent protein

Almost all cellular proteins require some folding before they can mature and begin to

function. GFP is nonfluorescent as it leaves the ribosome and only begins to emit fluorescence

when proper 3D structure is achieved, allowing the chromophore to form. The chromophore, p-

hydroxybenzylideneimidazolinone, is the source of fluorescence emission and is formed from the

spontaneous cyclization and oxidation of residues serine 56 (S65), tyrosine 66 (Y66) and glycine

67 (G67), by essentially a nucleophilic attack at the carbonyl carbon of S65 by the amide

nitrogen of G67 followed by dehydration. In the presence of molecular oxygen, Y66 and the new

imidazolinone group conjugate, forming the mature excitable chromophore (Cody et al., 1993;

Prasher et al., 1992). GFP exhibits green (509 nm) fluorescence when excited with blue (488 nm)

light (Prasher et al., 1992). At a rate of 6 amino acids/second (Ingolia et al., 2011), it takes

roughly 40 seconds for one molecule of GFP to be translated in generic mammalian cytoplasm

and several minutes for folding to be completed (Shaner et al., 2008). Therefore, crucial to

chromophore maturation and fluorescence emission is the proper folding of the protein to near

native conformation. Protein maturation always proceeds protein folding and it is important to

distinguish these two processes. It is also important to note, unlike most proteins that require

molecular chaperones to fold, no enzymes or cofactors are needed for GFP maturation to occur,

except molecular oxygen. Dependence of maturation on oxygen was established based on the

Page 20: Observation and quantification of protein production in single living cells

8

simple finding that GFP fluorescence does not develop in the absence of atmospheric oxygen

(Cubitt et al., 1995; Heim et al., 1995; Inouye & Tsuji, 1994). GFP that has been produced in

anaerobic conditions is non-fluorescent. However, the protein emits fluorescence that develops

exponentially after air is introduced. This process was measured to take 95 minutes and found to

be unaffected by either the concentration of starting GFP or the presence or absence of various

tested cellular factors (Heim et al., 1995).

1.2.3b Immunoglobulin antibodies

In the immune systems of vertebrates, four polypeptide molecules, two identical heavy

chains and two light chains constitute the basic structural unit of an antibody molecule. Five

classes of mammalian antibodies exist, each with its own class of heavy chain and either of two

classes of light chains: kappa (κ) and lambda (λ). Examination kappa chain amino acid

sequences has shown the C-terminal region consists of nearly identical residues, while the

variable N-terminal half consists of a relatively constant framework region and three small

hypervariable loops that provide the structural basis for the diversity of antigen-binding sites. A

pair of 7-9 antiparallel beta strands forms an 80 amino acid barrel-like structure termed the

immunoglobulin (Ig) domain, which forms the basis for protein-protein or protein-ligand

interactions (Janeway et al., 2001; Lodish et al., 2008).

The diversity in antigen binding sites is the result of extensive genomic rearrangement

that occurs at immunoglobulin gene loci. In mice, the total number of immunoglobulin kappa

(IgK) genes is 180 (174 variable genes, 5 joining genes and 1 constant gene). Recombinase

enzymes recognize recombination signal sequences which choose and join one variable gene in

frame with one joining gene and the common constant exon. Such mechanisms which are

essential to generate large diversities in antibody specificity, necessarily increase the probability

Page 21: Observation and quantification of protein production in single living cells

9

of producing proteins that fold improperly or cannot assemble. Instability in protein structure can

hinder the transport and secretion of antibodies in addition to restricting the interaction of

antibodies with signaling molecules, all of which compromise immune responses. It is therefore

not surprising that some of the earliest endoplasmic reticulum folding enzyme substrates were

immunoglobulin molecules (Haas & Wabl, 1983; Rao et al., 1976). Moreover, many protein

quality control mechanisms that ensure only correctly assembled proteins are retained, were

originally identified by their association to Ig antibody chain production. (Feige et al., 2010).

Using molecular dynamics simulations, it is now generally agreed that small biases

towards native-like states during protein folding lead to stable intermediary transition states that

reduce the conformational search to obtain realistic folding times (Martinez et al., 1998; D. C.

Sullivan & Kuntz, 2002). Following folding, quality control mechanisms ensure that only

properly folded and assembled proteins remain, and misfolded proteins are sent for degradation

and recycling in the proteasome (Lodish et al., 2008).

1.2.3c Measurement of protein folding

It is not trivial to measure the time it takes for a protein to fold and mature within a cell,

primarily because it is near impossible to establish a “time zero” moment when the protein is

translated. A complex orchestration of modifications and interactions follow the synthesis of

proteins in cells: for example, many proteins begin folding as the nascent peptide chain exits the

ribosome tunnel (Holtkamp et al., 2015). Many other proteins, such as proinsulin, are produced

in an inactive form and require posttranslational cleavage by a protease to generate the active

form of the protein (Orci et al., 1986). However, certain aspects of folding and maturation can be

utilized to untangle the two processes. Characteristic protein functions and properties such as the

emission of fluorescence in fluorescent proteins can be used, with caution, as reporters of protein

Page 22: Observation and quantification of protein production in single living cells

10

maturation (Craggs, 2009). For example, since the rate limiting step of GFP maturation is the

final oxidation step, one way maturation rates of different GFPs were determined in vitro and in

vivo was by manipulating the levels of available molecular oxygen in the assay, either by using

anaerobic in vitro protein synthesis (Iizuka et al., 2011) or by growing bacteria in anaerobic

conditions (Hansen et al., 2016; K. P. Scott et al., 1998). During synthesis the protein begins to

fold, and the maturation process begins with cyclization and dehydration, but halts at the rate

limiting oxidation step. Measuring the fluorescence recovery after oxygen admission effectively

measures the rate of the last and longest step of chromophore maturation (Iizuka et al., 2011).

GFP folding rates are usually measured by first completely denaturing the protein

rendering it unfolded using denaturants such as urea, then measuring the time it takes for

fluorescence to develop after washing the urea off (Pédelacq et al., 2006). It should be kept in

mind that prior to protein unfolding the chromophore is already in its mature form, and when the

protein is unfolded the chromophore remains in a cyclized state (Waldo et al., 1999; Ward et al.,

1980). Therefore, the maturation of the GFP chromophore is a permanent indicator that it had

once folded properly to its native state (Reid & Flynn, 1997). This eliminates the need for the

unfolded protein to go through the maturation step during subsequent refolding, therefore

measuring the fluorescence recovery effectively measures the rate of the folding step, unaffected

by maturation (Waldo et al., 1999).

There are several reasons why researchers require fluorescent proteins to have stable and

characterized folding and maturation times. One of the earliest and most common uses of GFP in

cells is its fusion to a protein to investigate that protein’s localization (Htun et al., 1996; Marshall

et al., 1995; Rizzuto et al., 1995). Early unstable versions of GFP posed several problems in

fusion protein experiments as improperly folded proteins aggregate into inclusion bodies. Protein

Page 23: Observation and quantification of protein production in single living cells

11

aggregates can be toxic to cellular health, and render the spatial visualization of proteins or

quantification of fluorescence intensity ambiguous (Kopito, 2000). Moreover, assays requiring

tagging of low abundance proteins, or those with short half-lives, require reporters that have high

spatiotemporal resolution. Therefore, reporters that present any time delay between their

production and the emission of fluorescence can preclude accuracy in determining when and

where the protein is produced.

1.3 Protein function

The biological properties of a protein determine the type of interaction or binding a

protein undergoes. The regulation and modification of such properties by cells is a direct way to

regulate protein function. All proteins interact with other molecules, be it small molecules,

nucleic acids or other proteins. For example, antibodies bind to antigens and mark them for

phagocytosis and destruction, cell surface receptors bind other protein ligands and transduce

signals, and ion channels bind ions and metals. The ability for a protein to bind its ligand with

high selectivity and affinity depends on the degree of formation of weak non-covalent bonds

such as hydrogen bonds, van der Waals attraction and ionic bonds, which are in turn dictated by

how closely the ligand and protein surfaces fit together, analogous to a lock and key concept

(Alberts et al., 2002; Lodish et al., 2008).

1.3.1 Regulation of protein function

Many proteins have critical cellular functions such as the catalysis of master metabolic

reactions that have widespread downstream effects, or the critical regulation of cell division to

prevent abnormal cell proliferation as seen in many cancers. Therefore, the existence of different

regulatory mechanisms that coordinate protein activity is key to proper cell function. As such,

Page 24: Observation and quantification of protein production in single living cells

12

the regulation of the function or activity of a protein is achieved at several levels during a

protein’s life span within the cell. In general, there are at least three ways in which cells can

regulate the activity of proteins (Lodish et al., 2008). One of the first levels of control over

protein activity is at the level of regulation of gene expression. Upregulation and downregulation

of mRNA expression and translation, in addition to protein degradation alter the steady-state

level of the protein, and by proxy the level of activity that results in the cell. A second level of

regulation of protein activity is using molecules that are either irreversibly added to the structure

of the protein, or that bind and dissociate acting as molecular on and off switches that control

protein activity. A third way by which cells can regulate protein activity is by localizing and

concentrating proteins to subcellular compartments to spatially restrict their activity. Similarly,

cells can produce point sources or gradients of another molecule or cofactor required for the

activity of the protein, thereby resulting in a spatial gradient of activation or suppression of

protein activity (Alberts et al., 2002; Lodish et al., 2008).

1.3.2 Regulated protein synthesis and degradation

Intracellular and extracellular cues often work in concert to regulate gene expression. For

example, increased neuronal firing stimulates the expression of immediate-early gene mRNAs

such as c-fos and Arc and their subsequent conversion to protein within minutes (Gissel et al.,

1997; Na et al., 2016). Such markers of external stimuli are labile, and must be recycled quickly

to prevent the constitutive activation of signaling and response mechanisms.

The stabilities and half-lives of cellular proteins vary widely. Proteins are constantly

degraded and replaced with newly synthesized copies, and this turnover process ensures a

constant supply of new functional protein to replace non-functional and damaged (sometimes

toxic) species. Turnover studies across organisms have shown that protein half-life can vary

Page 25: Observation and quantification of protein production in single living cells

13

many orders of magnitude within cells (Toyama & Hetzer, 2013). For example, examination of

turnover rates in budding yeast (cell cycle of 1.5 hours) revealed that mean protein half-life

under normal conditions is ~43 minutes (Belle et al., 2006), and in dividing human cells (cell

cycle of 24 hours), this figure increases to 0.5-35 hours (Cambridge et al., 2011). Although

protein half-life within the cell can vary from minutes to days, the turnover of proteins seems to

correlate with their function or subcellular localization (Toyama & Hetzer, 2013). For example,

endoplasmic reticulum and mitochondrial proteins on average have longer half-lives than other

cellular proteins (Price et al., 2010). In contrast, eye lens, tooth enamel and tooth dentine protein

half-lives are on the order of decades (Helfman & Bada, 1975, 1976; Masters et al., 1977).

Accelerated turnover of proteins is often the result of signals or modifications that are

added to the protein to increase its rate of degradation. For example, sequences rich in proline

(P), glutamic acid (E), serine (S) and threonine (T) (PEST) are associated with proteins that have

a short intracellular half-life such as the neuronal activity-regulated protein Arc (t1/2 ~ 37 mins),

whose expression marks short-term increased neuronal activity. In other cases, phosphorylation

can mark a protein for rapid turnover and degradation. For example, the production and

degradation of circadian clock protein such as period and timeless, are tightly regulated via

phosphorylation, and the resulting accelerated turnover ensures constant levels of circadian

proteins are never achieved, and instead a cycle is preserved (Kwon et al., 2006). Prior to

degradation, many proteins require the addition of ubiquitin groups, which is catalyzed by a

number of ubiquitin modification enzymes. The number of ubiquitin tags a protein contains

usually correlates with the level of recognition and degradation by the 26S proteasome (Smalle

& Vierstra, 2004).

1.3.3 Protein phosphorylation

Page 26: Observation and quantification of protein production in single living cells

14

The phosphorylation of proteins imparts negative charges, which result in subtle

conformational changes in protein shape that affect ligand binding or other functions of the

protein. Protein kinases catalyze phosphorylation of specific residues (tyrosine or serine and

threonine kinases) and this reaction is reversed by protein phosphatases, which hydrolyze and

release phosphate groups from modified residues. The combined action of protein kinases and

phosphatases is an important regulatory mechanism that can act as a reversible switch to increase

or decrease the activity of proteins, as such it is highly conserved across prokaryotes and

eukaryotes (Alberts et al., 2002). All classes of proteins including structural proteins, signaling

proteins, enzymes, membrane channels and scaffolds are known to be regulated by

phosphorylation. At any given time, 1/3 of the proteome is thought to be phosphorylated,

resulting in many proteins having more than one phosphate group (Lodish et al., 2008).

Through similar effects, reversible protein phosphorylation is known to control the

structure, localization and therefore function of a host of eukaryotic proteins. The simple

addition and removal of phosphate groups in response to signals is an elegant mechanism used to

ensure adequate protein activity, relay of signals between organelles or as a check-point for

molecular processes (Johnson, 2009).

1.3.4 Regulated translation of localized mRNAs

The localization and regulated translation of mRNA is an example of a cellular regulatory

mechanism used to spatially and temporally restrict gene expression to discrete cellular sites

(Martin & Ephrussi, 2010). Some of the best studied examples of locally translated mRNAs

include those whose protein products must be spatially sequestered in order to play specialized

roles within defined subcellular compartments (Ables, 2015; Martin & Ephrussi, 2010). For

example, the induction of mesodermal and endodermal cell fates in the Xenopus oocyte vegetal

Page 27: Observation and quantification of protein production in single living cells

15

pole is regulated by the local translation of VegT mRNA (King et al., 2005). Similarly, the

specification of asymmetric cell division in solely daughter cells of budding yeast is a result of

localized translation of ASH1 transcriptional repressor mRNA in the daughter cell (Paquin &

Chartrand, 2008). In addition, large-scale analyses have revealed several other genes whose

mRNAs are spatially sequestered and locally translated. For example, high-throughput in situ

hybridization of 3,000 mRNA transcripts in Drosophila has found that more than 70% of the

assayed transcripts were present in spatially distinct patterns (Lecuyer et al., 2007), which

indicates that the localization of mRNAs to subcellular compartments is more prevalent than

previously thought (S. Kim et al., 2010).

There are many advantages of regulating gene expression via local mRNA translation.

First, it is a mechanism that allows spatial restriction of gene expression within a cell’s

cytoplasm. Second, it provides cells with fine temporal resolution to control translation of

already spatially restricted transcripts, using local cues that can stimulate initiation of translation

on site. Third, almost every step involved in the production and transport of a protein to sites

where it is needed, requires expensive cellular reagents. Therefore, it is much more economical

for cells to locally produce many copies of a protein from the same few mRNA molecules, rather

than shuttle individual proteins from the cell soma to where they are required (Martin &

Ephrussi, 2010)

The targeting of mRNAs to their cellular destination involves cis-acting elements in the

RNA sequence. Most often, these localization elements are found in the 3’UTR, although some

transcripts such as the sensorin gene mRNA contain localization elements in their 5’UTR that

allow them to be targeted to dendrites and synapses (Dan Ohtan Wang et al., 2009). In addition,

some mRNA localization elements are found within the coding sequence, such as in the gurken

Page 28: Observation and quantification of protein production in single living cells

16

mRNA in Drosophila (Lasko, 2012). Several studies have shown that the processing of pre-

mRNA in the nucleus is required for its cytoplasmic localization by interacting with RNA

binding proteins that recognize localization and translational regulation signals within the

sequence (Giorgi & Moore, 2007). The association of RNAs with RNA binding proteins forms

ribonucleoprotein complexes which in many cases are part of a bigger RNA transport granule, a

structure transported by motor proteins that use the cell cytoskeleton to deliver the RNA to its

final destination (Martin & Ephrussi, 2010). During mRNA transport, mechanisms are in place to

repress the translation of RNA until the granule is anchored at its destination, where additional

mechanisms ensure it is translated at the right time (Besse & Ephrussi, 2008).

1.3.5 Local translation of mRNA shapes development

The local translation of mRNAs in the Drosophila oocyte is one of the earliest and best

studied examples of how the localization of mRNA is used as a mechanism for translational

regulation. During egg development, germ cell specification and embryonic axis patterning are

established via molecular asymmetries created by position-dependent regulation of the

translation of mRNAs deposited maternally into the oocyte (Richter & Lasko, 2011). The

location-dependent activation of translation is also coupled with mechanisms that localize and

concentrate specific mRNA transcripts in areas where the corresponding protein will be

produced. Such mechanisms ensure that proteins are present in their highest concentrations

where they are required, as opposed to where their presence would be deleterious.

The Drosophila ovary consists of ovarioles that harbor oocytes and provide the

microenvironment necessary for the development of the oocyte and the synthesis of maternally

deposited DNAs and RNAs that will be required post-fertilization for the proper development of

the egg axis and shell (Ables, 2015). At the anterior-most end of each ovariole is the germarium,

Page 29: Observation and quantification of protein production in single living cells

17

a structure which houses the primordial germ cells from which oocytes develop (Figure 1.1).

These germline stem cells undergo several rounds of asymmetric cell division to produce the

germline cyst, a cluster of 16 germ cells interconnected by intracellular ring canal bridges.

Fifteen of these cells will be destined to become nurse cells and the remaining cell becomes the

oocyte (Figure 1.1). Nurse cells produce several mRNAs, such as the gurken (grk) mRNA which

is transported into the oocyte where it colocalizes with the oocyte nucleus. The 1.7 kb grk

mRNA encodes the Gurken protein which is a ligand for Torpedo/EGF receptor, a receptor

located on the inner surface of follicle cells that envelop the oocyte. The analysis of grk mRNA

transcripts has revealed a conserved RNA stem loop element within the grk coding region that

forms the signal for dynein-dependent grk mRNA transport and localization to the oocyte

nucleus (Van De Bor et al., 2005).

During the early stages of oogenesis (stage 6 onwards) the nucleus is located at the posterior

end of the oocyte and grk mRNA accumulates there, however, during later stages of

development (stages 9,10 and 10B) the nucleus moves to the anterodorsal corner of the oocyte,

where the grk mRNA follows to create the characteristic crescent pattern of localization between

the apical surface of the nucleus and the surrounding cortex (Richter & Lasko, 2011). Localized

translation of Gurken protein at the anterodorsal corner creates a local source and molecular

gradient of Gurken signalling such that the highest EGF-R/Torpedo signalling occurs in

neighbouring anterodorsal follicle cells which initiates cell fate specification locally (Gavis,

1995). The anterior-posterior specification of polarity in the oocyte arises from the movement of

the oocyte to the posterior of the egg chamber prior to stage 6. This precedes the dorsoventral

specification of the oocyte axis at stage 8 which is mediated by the movement of the grk mRNA

from the anterior cortex of the oocyte to the anterodorsal corner.

Page 30: Observation and quantification of protein production in single living cells

18

The local translation of Gurken in Drosophila oocytes is therefore a defined system in which

the spatial and temporal characteristics of Gurken mRNA translation are known and invariable.

1.3.6 Local translation of mRNA in neurons

Animal development and behavior are shaped by wiring and activity of neural circuits.

During the formation and development of circuits, neurons elaborate processes that can extend

great distances before forming synaptic contacts with their partners. Axonal and dendritic

subcellular compartments must integrate a wide array of molecular cues whose correct spatial

and temporal processing is critical for correct patterning and circuit formation. Stimulus-induced

changes in the structure and function of these compartments are vital to the formation and

plasticity of neural circuits (Kandel, 2001). The ever-changing demands of growing axons and

dendrites raise the question of how gene expression can be spatially restricted within a neuron.

The localized translation of mRNA provides one solution, and translation of localized transcripts

within different subcellular neuronal compartments has been observed (Jung et al., 2012; Dan

Ohtan Wang et al., 2010). For example, local translation of proteins can occur in axonal growth

cones during axon guidance and circuit formation (Lin & Holt, 2008), in dendritic spines during

learning and memory formation (Dan Ohtan Wang et al., 2009) and in axons during injury-

related regeneration (Willis & Twiss, 2006). In addition, local protein translation rates have been

shown to vary throughout development. For example, in rat hippocampal neurons, radiolabeled

amino acid incorporation analysis showed that rates of local protein synthesis in dendrites rise

early in development and peak during synaptogenesis, but decline by adult stages (Steward et al.,

1998).

Studies of activity-regulated genes were among the first to show that synaptic activity

regulates dendritic mRNA localization, and in a transcript-specific manner. For example,

Page 31: Observation and quantification of protein production in single living cells

19

induction of LTP in rat hippocampus has been shown to result in increased localization of

αCaMKII and MAP2 mRNA to granule cell dendrites in vivo (Roberts et al., 1998). Similarly,

newly synthesized Arc mRNA specifically localizes to and accumulates in activated synapses of

the rat dentate gyrus (Steward et al., 1998).

Other types of cues may also signal mRNA transport and local protein synthesis. For

example, the regulation of axon growth cone architecture requires the local translation of

cytoskeletal proteins such as beta-actin and cofilin, in addition to signaling proteins such as

RhoA and MAP1b, demonstrating an important role for axonal protein synthesis in the

development and maintenance of neuronal function (Hengst & Jaffrey, 2007). During

development of mouse dorsal root ganglion neurons, axonal mRNA translation is responsible for

retrograde signaling that regulates transcription of genes in the nucleus. Specifically, cAMP

response binding element (CREB) local translation in axons via NGF signaling is required for

somatic CRE-dependent transcription and subsequent neuronal survival mediated by nerve

growth factor (NGF) (Cox et al., 2008; Sharon A Swanger & Bassell, 2011). This indicates that

the distal production of proteins has important implications on the expression of genes in the

nucleus and thus can result in cell-wide changes. Moreover, the involvement of neuronal survival

factors such as NGF, cAMP and BDNF suggests implications of these processes in neuronal and

developmental diseases.

Indeed, dysregulation in mRNA localization and local protein synthesis in neurons has

been observed in several neurological diseases. For example, spinal muscular atrophy (SMA) is a

degenerative disease that results in motor neuron death and muscle atrophy. It is caused by

mutations in survival of motor neuron1 protein (SMN1). The observation of SMN1 localization

to RNA granules in axons suggested a possible role in mRNA transport. In fact, SMN has been

Page 32: Observation and quantification of protein production in single living cells

20

shown to interact with axonal RNA binding proteins, such as hnRNP-R (Mourelatos et al., 2001).

In addition, SMN knockdown in cultured mouse motor neurons has been shown to result in a

reduction of axonal beta-actin mRNA levels, which was correlated with stunted axonal growth

and decreased size of growth cones, recapitulating SMN1-deficiency phenotypes (Glinka et al.,

2010).

Loss of fragile X mental retardation protein (FMRP) causes fragile X syndrome, a

developmental disorder associated with intellectual, behavioral and physical delays, and the most

common inherited form of cognitive deficiency (Bagni et al., 2012). FMRP is known to have

dual functions in regulating dendritic mRNA transport and local protein translation in an

activity-dependent manner (Besse & Ephrussi, 2008). Many FMRP targets are mRNAs encoding

synaptic proteins that include receptors, signaling molecules and cytoskeletal proteins such as

MAP1b and PSD95, in addition to its own mRNA (Bassell & Warren, 2008; Martin & Ephrussi,

2010; S. A. Swanger & Bassell, 2013). Metabotrobic glutamate receptor (mGluR) signaling is a

major regulator of FMRP-mediated local protein synthesis, and mGluR activation induces the

transport of FMRP into dendrites, where it associates with polyribosomes and inhibits protein

translation (Martin & Ephrussi, 2010).

The local translation of mRNAs in neurons is therefore a fundamental mechanism by

which gene expression can be spatially and temporally regulated. Detecting the location and time

of protein synthesis with high resolution can provide insight into the correlation between

localized protein production and cellular changes.

1.4 Quantification of gene expression and protein levels

The plethora of cellular mechanisms regulating mRNA and protein levels presents

unambiguous evidence of the importance of the careful balance of protein activity. The central

Page 33: Observation and quantification of protein production in single living cells

21

dogma of biology states that proteins are translated from mRNA molecules that get transcribed

from DNA. Biological identity and function are the result of complex interplay between the four

fundamental processes involved in gene expression: transcription, mRNA degradation, protein

translation and degradation, and their independent interactions with the environment. Each step

is governed by distinct regulatory mechanisms that ensure a careful balance of mRNA and

protein levels is present to properly meet constantly changing cellular needs. For example,

dendritic arbor complexity of Drosophila larval body wall neurons is dependent on the careful

regulation of the level of the transcription factor Cut, that activates genes involved in the growth

and stabilization of neuronal branches. The more Cut protein a particular neuron expresses, the

more complex its dendritic arbor will be (Grueber et al., 2003; Lo, Kays et al., 2015). Similarly,

cyclic changes in the concentration of mRNA and proteins of genes such as period, clock and tim

in the Drosophila lateral neurons drive circadian rhythms required for essential behaviors such as

locomotor activity and eclosion during metamorphosis (Benito et al., 2007). In human disease,

the quantification of mRNA and protein levels has helped develop clinical and diagnostic

markers that routinely aid the discovery of early stages of diseases such as in cancer (Kishikawa

et al., 2015; Krishna Prasad et al., 2013). For example, many human cancers are characterized by

the upregulation in expression of known oncogenes such as myc and akt2 (Prelich, 2012) and

downregulation of tumour suppressor genes such as p53 and PTEN in tumour cells (Shain &

Pollack, 2013). Dysregulation of protein levels is also a hallmark feature of many

neurodegenerative diseases as described in the previous section.

Copy numbers of mRNAs and proteins normally fluctuate over time and vary from cell to

cell, mostly as a response to various environmental and intracellular cues, and in part due to

stochastic molecular events during gene expression. In addition, post-translational mechanisms

Page 34: Observation and quantification of protein production in single living cells

22

can upregulate or downregulate protein activity, which in turn can alter the expression of many

other genes. Therefore, to understand the influence of genes on cellular phenotypes, it is

important to understand how different levels of gene products affect normal and abnormal

cellular states, which requires the development of specific tools for quantifying gene expression.

These gene expression assay tools can be broken down into two main categories: those that

quantify changes in mRNA levels and those that quantify protein levels.

1.4.1 Quantification of mRNA levels

Until recently, it was common practice to use mRNA levels as proxy measurements for

protein levels, primarily because of difficulties in estimating protein abundances on a large scale

(Greenbaum et al., 2003; Maier et al., 2009). Classical hybridization-based methods such as the

Northern blot, which use RNA probes complementary to target sequences, allowed the steady

state detection and quantification of select mRNA transcripts. More recent approaches use arrays

of thousands of RNA or cDNA probes complimentary to target mRNAs, which allows the

unbiased detection of mRNAs present in the sample. However, probe-based hybridization assays

require the a priori requirement to know the sequence of the targeted mRNAs to design the

hybridization probes, which hinders the detection of new transcripts. In addition, probe-based

assayed suffer from cross-hybridization artifacts that result from non-specific sequence

homology between unrelated transcripts, which can result in detection inaccuracies. More recent

RNA sequencing technologies assay all RNA species in the sample in an unbiased approach,

enabling the detection and quantification of previously unidentified transcripts.

With the advent of PCR in the early 1980s came unprecedented sensitivity and efficiency

in the exponential amplification of targets (Mullis et al., 1986; Saiki et al., 1985). The

development of real-time PCR (qPCR) in 1993 represented a milestone in PCR becoming a

Page 35: Observation and quantification of protein production in single living cells

23

quantitative assay (Higuchi et al., 1993). The addition of a reverse-transcription (RT) step prior

to qPCR produced the RT-qPCR assay which provided the ability to detect and amplify mRNA

transcripts. The amplification step allowed detection of relatively low-abundance mRNA

transcripts, which overcame sensitivity issues faced with traditional hybridization-based methods

such as the Northern blot (Alwine et al., 1977). RT-qPCR now represents a workhorse of mRNA

expression studies due to its high sensitivity, high reproducibility, broad dynamic range and ease

of use (S. A. Bustin, 2000; Ding & Cantor, 2004).

Measuring mRNA levels has come a long way (Kavanagh & Baker, 2009; Ozsolak &

Milos, 2011; Sage et al., 2015; Vogel & Marcotte, 2012). However, it should be kept in mind

that mRNA levels only reflect which genes have been transcribed, and give no insight into the

dynamics of the proteome. Therefore, the interpretation of mRNA levels as a proxy measure for

protein levels is purely correlative. This suggests that directly measuring protein levels instead is

an approach that is likely to yield data that can be interpreted as causative of observed cellular

changes.

1.4.2 Quantification of protein levels

Numerous analytical assays have been developed to measure protein levels. They can be

broken down into a few main categories: dye-based and colorimetric absorbance measurements

for the quantification of bulk levels of protein within a solution, antibody-based approaches for

the detection of single proteins and mass spectrometry approaches for the unbiased detection and

quantification of any protein species within a sample.

Bulk protein measurements using colorimetric assays are very sensitive assays that can

detect down to 20 nanograms of protein (Fowler, 1996). However, they are extremely sensitive

to reaction-quenching substances that commonly contaminate protein sample preparations, which

Page 36: Observation and quantification of protein production in single living cells

24

can result in erroneous quantification of protein levels. In bulk assays the protein source is often

the total protein fraction from populations of cells or dissected tissue, which inevitably includes

unwanted and unpredictable cell types, resulting in heterogeneity and poor cellular resolution.

The detection of proteins using antibodies is perhaps the most common approach used by

research and clinical labs. Techniques such as the quantitative immunoblot and the enzyme-

linked immunosorbent assay require the use of an antibody specific to the protein of interest,

which is then detected using chemical reactions or fluorescence to provide a “snapshot in time”

of the level of the protein of interest. Crucial to the success of these techniques is the availability

and quality of the antibody used, which often determines how consistent and accurate the

observed results are. A major disadvantage that prevents the use of these tools in vivo is that they

are inherently invasive, requiring the lysis of cells to release their protein content. Moreover,

assays requiring multi-antibody incubation, or high resolution separation of proteins, are time-

consuming, and can often require up to three days from start to finish (Baker, 2015). These

issues are complicated when using polyclonal antibodies, which display high variability in their

binding, and poor signal to noise ratios during quantification (Marx, 2013). Furthermore, the

phosphorylation state of the target protein, access of the antibody to the protein epitope as well

as inherent variability in the affinity and avidity of antibodies, often contribute to issues of poor

reproducibility (Baker, 2015).

In this thesis, I use genetically encoded fluorescent proteins to track protein production in

living cells. Similarly, genetically encoded fluorescent protein fusions have enabled the direct

observation of intracellular proteins, circumventing the need for secondary treatments or agents.

The fusion of a fluorescent protein to a protein of interest is routinely used to track the

production, trafficking and localization of the fused protein of interest. As discussed previously,

Page 37: Observation and quantification of protein production in single living cells

25

critical to the function of any protein is its 3D conformation which is dependent on the careful

folding of the amino acid sequence. Therefore, the fusion of fluorescent proteins, even via

peptide linkers, can be expected to interfere to some extent with protein function. For example,

fused fluorescent proteins have been shown to interfere with protein function via steric

hindrance, or by changing the fused protein’s half-life (Snapp, 2005). In addition, tracking

secreted proteins, low abundance proteins or proteins with punctate distributions can be difficult,

as non-geometric and open cellular spaces require relatively large local concentrations of

fluorescent proteins for them to be observed with standard fluorescence microscopy methods. A

typical excited FP can emit hundreds to thousands of photons before entering its dark state

(Kubitscheck et al., 2000). However this transition is a random event with a fixed half-life,

indicating that some fluorophores will enter the dark state before enough photons are captured by

the detector. Therefore, assuming the brightness of the fluorophore is constant, reliable detection

above noise levels requires either a larger concentration of single fluorescent proteins or more

sensitive detectors in order to overcome shot noise of photon movement and electronic noise in

the detectors when operating at low detection ranges (Bagshaw & Cherny, 2006). Nevertheless,

it is noteworthy that advances in microscopy and detector technologies have enabled the

detection of single fluorescent molecules (G.-W. Li & Xie, 2011; Pinaud & Dahan, 2011;

Tatavarty et al., 2012), however access to such equipment by common research labs is yet to

become mainstream.

1.4.3 mRNA levels as proxy for protein abundance

Despite significant improvements in technologies used to quantify proteins, detection and

measurement of protein production from cells is still laborious and invasive (Walker, 2009). As a

result, researchers have turned to correlations between mRNA and protein levels to estimate

Page 38: Observation and quantification of protein production in single living cells

26

protein abundances from quantitative mRNA data, which are easier to collect (Greenbaum et al.,

2003). However, combined analysis of proteomic and transcriptomic data from populations of

cells has recently revealed an unexpectedly poor correlation between mRNA and protein levels,

hovering around 40% explanatory power (Greenbaum et al., 2003; Maier et al., 2009;

Schwanhäusser et al., 2011).

Quantitative mRNA analyses have and continue to demonstrate their value and

usefulness in biological discovery. However, the poor correlation between mRNA and protein

levels suggests that understanding cellular phenotypes using transcript levels as a proxy for

protein abundances is inaccurate, particularly at the single cell level.

1.5 Fluorescence-based single cell resolution protein quantification

Most of the common tools used to quantify the levels of mRNA and proteins lack single

cell resolution. High detection limits, limited sensitivity and low signal-to-noise ratios force

researchers to increase the amount of starting material used in any assay. To overcome these

problems, populations of cells, tissues or whole animals are routinely used to prepare mRNA and

protein samples and consequently the collected data are averaged across thousands of individual

cells.

Individual cells are inherently heterogeneous entities; for example, neighboring neurons

and glia in the brain have been shown to express as little as 65% of the same genes (Schubert,

2011). Moreover, immune cells that are commonly grouped by surface expression of markers

such as CD1 and CD4, have also been found to express entirely different sets of genes and can

have widely varying responses to vaccines and therapy (Flatz et al., 2011). This means assays

performed on large numbers of cells result in population-averaged measurements that miss

important individual cellular differences. Therefore results obtained from a population sample

Page 39: Observation and quantification of protein production in single living cells

27

cannot be translated into the single-cell scale (Li & Xie, 2011). In addition, the fact that proteins

are what exert change within a cell, suggests we should be directly examining the production of

proteins and their interactions to understand the molecular processes that govern cellular

function. Assaying individual cells overcomes issues of cellular heterogeneity, and directly

assaying protein levels overcomes the significant weakness in mRNA abundance prediction

power. It is therefore intuitive and logical to conclude that the tools most likely to yield the most

informative results would be tools that directly assay protein production from single cells.

1.5.1 Quantification of protein levels in single living cells

I have recently contributed to the development of a technique to quantify protein

production from single cells in vivo (Lo, Kays et al., 2015). This Protein Quantification Ratioing

(PQR) technique uses a genetic tag that produces a stoichiometric ratio of a fluorescent protein

reporter and the protein of interest during protein translation (Figure 1.2a). The fluorescence

intensity is proportional to the number of molecules of protein of interest produced and is thus

used to determine the level of protein production within the cell (Figure 1.2b). Some RNA

viruses have sequences encoding for short polypeptides ~20 residues in length called CHYSEL

polypeptides (also known as “2A” and “2A-like” peptides, collectively) (de Felipe et al., 2006).

During protein translation, the interaction of the nascent CHYSEL peptide with residues in the

ribosome exit tunnel causes a conformational change which restricts peptide bond formation

between the last two residues of the CHYSEL peptide. This does not terminate translation, but

instead causes the ribosome to autonomously skip and continue translation starting from the last

CHYSEL residue, effectively resulting in the production of two separate functional proteins

(Figure 1.2a).

Page 40: Observation and quantification of protein production in single living cells

28

This unique mechanism can be exploited to allow multiple proteins to be produced from a

single polycistronic RNA strand to produce stoichiometric amounts of upstream and downstream

proteins, and this has been demonstrated in vitro and in vivo using high-throughput analyses (J.

H. Kim et al., 2011; Radcliffe & Mitrophanous, 2004; Szymczak et al., 2004; Szymczak-

Workman et al., 2012). We modified and tested different CHYSEL peptides for efficient and

stoichiometric separation of the upstream and downstream proteins and identified different

optimized sequences for use in mammalian and insect systems; we have collectively called these

DNA constructs Protein Quantification Reporters (PQRs). To demonstrate that PQRs can be

used to accurately and reliably quantify protein production, we validated the stoichiometric ratio

and linear relationship between different genes at the single cell level using quantitative imaging

and electrophysiology and found that PQRs allow for equimolar separation between different

proteins and this can be used to correlate changes in protein production with cellular phenotypes

in living cells (Figure 1.2b).

The bicistronic co-expression of a fluorescent reporter with a protein of interest using

PQR for tracking protein production has enormous advantages. Compared to internal ribosomal

entry sites (Pelletier & Sonenberg, 1988), the small size (20-30 amino acids) and low complexity

of the PQR peptide sequence allow it to be easily cloned anywhere along a DNA construct, and

this minimizes any extraneous amino acids that are fused to the proteins of interest. Its self-

processing mechanism is autonomous and insensitive to variabilities of cofactors, enzymes or

global differences in cellular states (J. H. Kim et al., 2011; Radcliffe & Mitrophanous, 2004).

Perhaps most importantly is the fact that the well-characterized and stoichiometric expression of

proteins separated by PQR-type linkers allows consistent and high level expression of proteins

from multicistronic constructs. In addition, since the signal is genetically encoded and is

Page 41: Observation and quantification of protein production in single living cells

29

fluorescence-based, quantification of protein production can be done non-invasively over time,

in a rapid and straightforward manner. Moreover, the fluorescence intensity of the reporter is

directly proportional to the level of production of the protein of interest over a wide dynamic

range of quantification and thus the fluorescence output of a cell is a measure of the production

of the protein of interest in that cell (Figure 1.2b). The fluorescence imaging of fluorescent

reporters such as GFP can be done at single cell resolution with very high sensitivity, using

standard fluorescence microscopy techniques which allows the detection of low abundance

proteins and small differences in cellular protein production rates with high spatial and temporal

resolution. Quantification of protein production using PQR reporters is also an approach that is

robust and insensitive to agents that routinely interfere with common protein assays and this

facilitates the observation of protein production over time in single living cells.

1.5.2 Protein production reporters must be carefully chosen

Wild-type GFP tends to misfold and aggregate when expressed in cells, which limits

many of its applications including its use as a reporter of gene expression. To address these

issues, GFP has been subjected to a series of mutations aimed at targeting residues that directly

result in changes in folding and maturation properties. The mutation of a phenylalanine at

residue 64 to a leucine (F64L) introduced EGFP, a more stable protein at 37°C which tends to

aggregate less. A serine to threonine mutation at residue 65 shifted the spectrum of GFP and

resulted in a single excitation peak at 484nm and an emission peak at 507nm (Heim et al., 1995).

This mutation made the use of GFP in living systems more amenable as blue light wavelengths

instead of the more damaging ultraviolet wavelengths could now be used to visualize the protein.

Early efforts by the lab of Willem Stemmer attempted to improve the brightness of native GFP

expressed in bacteria using DNA shuffling, a technique that utilized recombination of in vitro

Page 42: Observation and quantification of protein production in single living cells

30

homologous sequences, which they pioneered for the screening and propagation of beneficial

mutations (Crameri et al., 1996). Three cycles of DNA shuffling resulted in a variant known as

“GFP cycle 3” which had a 45-fold increase in fluorescence compared to the standard

commercial GFP at the time (Crameri et al., 1996). This improvement was thought to be most

likely due to a reduction in protein aggregation and improved solubility as a result of several

hydrophobic residues being mutated to hydrophilic residues in GFP cycle 3. However, the

folding and maturation rates were not changed (~95 minutes in bacteria). Using the same

technique, a fast folding GFP bearing six additional mutations was identified and termed

superfolder GFP. Superfolder GFP is one of the most commonly used variants due to its high

folding efficiency, fast folding rate and high denaturant resistance (Pédelacq et al., 2006).

In contrast, many efforts attempted to destabilize GFP, and instead accelerate its turnover

rate to produce variants with half-lives as short as 2 hours (Li et al., 1998). The rapid turnover of

destabilized GFPs has allowed its use in studies requiring reporters with especially short half-

lives, such as in circadian rhythm (Hastings, 2005) or gene expression studies (Li et al., 1998). In

addition, the rapid turnover results in less accumulation of GFP in cells which leads to lower

toxicity at high levels (Li et al., 1998), and decreased fluorescence intensity saturation during

fluorescence imaging enabling longer cellular expression times and longer imaging durations.

1.5.3 Approaches to visualizing locally translated proteins

The detection of new protein synthesis in localized cellular compartments provides

powerful spatial and temporal resolution that allows the correlation of protein production and

development of cellular phenotypes. Genetically encoded fluorescent proteins such as GFP can

be used in fusion constructs, similar to Figure 1.2a, and the development of green fluorescence in

subcellular compartments is used as a marker of new protein translation. This seemingly

Page 43: Observation and quantification of protein production in single living cells

31

unambiguous approach, however, raises a critical concern for the need to verify that the observed

new green signal is truly made locally instead of being produced at the soma, and then

transported to the local site. The majority of fluorescent proteins require on average several tens

of minutes from the moment they are translated for proper folding, maturation and for the

eventual fluorescence signal to develop (Shaner et al., 2005). During this time, the non-

fluorescent protein molecule can freely diffuse or be transported within the cell. In addition,

increased ribosomal translation rates at perinuclear regions can result in saturation of the

detection range from somatically translated GFP signals, impeding the detection of locally

synthesized signals (S. Kim et al., 2010).

In this thesis, I am interested in developing tools that define the location, time and rate of

translation of proteins, with the goal to examine local synthesis of proteins in neuronal

subcellular compartments. In the following sections I describe how the discovery of mRNAs in

neuronal subcellular compartments led to efforts to visualize local protein synthesis in neurons.

1.5.3a Identification of localized mRNAs in neurons

The effort to develop tools to detect newly synthesized proteins in neurons started with

the discovery that mRNAs were found to localize to discrete dendritic and axonal compartments

(Ochs et al., 1969; Tobias & Koenig, 1975). Such findings provided an attractive model to

answer how long lasting changes in structure and function of subcellular compartments such as

growth cones and dendrites can persist and be spatially restricted in highly polarized cells such

as neurons. However, this also posed a problem as it raised the need to restrict the study of

mRNA localization and protein translation specifically in those compartments.

Early experiments aimed at separating neuronal processes from cell somas, either

biochemically using fractionation (Krichevsky & Kosik, 2001) or by a variety of physical ways

Page 44: Observation and quantification of protein production in single living cells

32

to isolate processes and somas (Kim et al., 2010) and then profiling the RNA species present in

the sample. In such assays, the purity of the sample is critical as the amount of mRNA transcript

present in neuronal processes is orders of magnitude less than in cell bodies. Any somatic

contamination can therefore overwhelm the ability to detect process-localized transcripts. Once

separated from lysate, the RNA is profiled either by T7 RNA amplification (J Eberwine et al.,

2001; Miyashiro et al., 1994), hybridization to known probes (Eberwine et al., 2001) or by

quantitative real time PCR amplification of transcripts (Zheng et al., 2001). Such approaches

have led to the identification of over 400 dendritically localized transcripts and over 150

synaptically-enriched transcripts such as MAP2 and αCaMKII (Eberwine et al., 2001; Tian et al.,

1999). Similar methods were used to isolate axonally-localized mRNAs, such as CREB and

RhoA mRNA from rat sensory neurons (Cox et al., 2008; Zheng et al., 2001). Process-localized

transcripts can then be used as starting material for the generation of cDNA libraries of locally

translated genes (Moccia et al., 2003). These assays approach the problem in an unbiased

manner, however they required rigorous controls to rule out the contamination of somatically-

derived transcripts. For example, in a study examining the contribution of contaminating somatic

mRNAs in process-localized mRNA preparations, the presence of false-positive hits due to

somatic contamination was found to be minimal (S. Kim et al., 2010; Poon et al., 2006).

However, surprisingly little overlap was seen when comparing dendritically localized rodent

hippocampal mRNAs identified using independent approaches. Such results raise a crucial need

to independently confirm the localization of identified transcripts using different approaches (S.

Kim et al., 2010). One way to confirm the localization of mRNA transcripts identified using

amplification is in situ hybridization (ISH) which uses labeled RNA or DNA probes that

hybridize to specific mRNAs, which can then be imaged directly in the case of fluorescent ISH

Page 45: Observation and quantification of protein production in single living cells

33

(FISH) or revealed using secondary agents such as enzymes (van de Corput et al., 1998) or

antibodies (Poon et al., 2006). FISH approaches have been used to identify the localization of

transcripts with great success in a variety of tissues such as cultured hippocampal neurons (Lyles

et al., 2006), whole-mount Aplysia CNS (Dan Ohtan Wang et al., 2009) and sections of Aplysia

ganglia (Poon et al., 2006). However, ISH only provides a snapshot in time of mRNA

localization and thus cannot provide much information into the dynamics of mRNA localization

or translation of new protein. Therefore, new methods for the dynamic imaging of mRNA

translation are needed.

One important limitation of visualization methods is that they provide information that is

correlative in nature. Although these methods fill a critical gap in our understanding of activity-

dependent and neuromodulatory control of local translation, they do not reveal whether the

presence of mRNA transcripts contributes to active local protein translation. The presence of an

mRNA at dendrites does not necessarily indicate it is actively being translated (Na et al., 2016).

However, the presence of an mRNA at distal sites does suggest it is a process-localized mRNA

used to provide a rapid local supply of its protein product. Assays that instead directly monitor

protein production from localized mRNAs can determine the moment a protein is made with

higher spatial and temporal resolution than visualization methods. This enables correlating local

cellular changes with the moment a protein is produced locally. Below, I describe alternative

approaches that have been applied specifically with this goal in mind.

1.5.3b Direct detection of local protein synthesis using photoconvertible reporters

A significant portion of our understanding of local protein synthesis has been the result of

using genetically encoded fluorescent protein reporters such as GFP, RFP and Dendra2 (Kim et

al., 2010; Martin & Ephrussi, 2010). Genetically encoded fluorescent proteins fused to proteins

Page 46: Observation and quantification of protein production in single living cells

34

of interest, or placed between the 5’ and 3’ untranslated regions of a target mRNA result in the

production of new fluorescent signals each time the mRNA of interest is translated. Genetically

encoded fluorescent proteins eliminate the need to use exogenous detection agents, are

minimally toxic to cellular health and emit fluorescence with high signal to noise ratios. This

enables the non-invasive long term imaging of living tissues. In neurons, the most successful

approaches took advantage of an interesting class of fluorescent proteins known as

photoconvertible fluorescent proteins.

Kaede and Dendra2 are examples of photoconvertible fluorescent proteins that normally

emit green fluorescence, but can be permanently photoconverted to emit red fluorescence by

ultra-violet range (350-400nm) illumination. The photoconversion property of Kaede originates

from the Histidine 62-Tyrosine 63-Glycine 64 tripeptide motif that forms a green fluorescence

emitting chromophore analogous to that in GFP. Upon exposure to UV radiation, a cleavage of

the amide bond at the histidine 62 residue results in the formation of a new spectrally distinct red

chromophore (Ando et al., 2002; Dittrich et al., 2005).

The permanent photoconversion of fluorescent proteins is an interesting property that can

be exploited to examine the production of new proteins by simply photoconverting all

preexisting protein from green to red and then monitoring new translation as the development of

new green signal. While this approach provides several advantages over using standard

fluorescent proteins, the issue of whether the protein is truly locally synthesized or has diffused

into the site in its immature nonfluorescent state, is not addressed with photoconvertible

fluorescent proteins. These drawbacks must be considered and accounted for when using

genetically encoded fluorescent proteins for examining local protein translation events.

1.5.4 Requirements for a local protein synthesis reporter

Page 47: Observation and quantification of protein production in single living cells

35

An ideal local protein synthesis reporter must meet certain performance criteria for it to

be able to accurately and reliably detect local protein translation events in vivo. For instance,

introducing the reporter into the animal, and the collection of data should be as minimally

invasive as possible. Non-invasive approaches enable long-term and repeated tracking of

changes in a single cell within a living animal, effectively allowing the examination of cellular

function in its natural, dynamic and unperturbed context. Local protein synthesis reporters must

also have high temporal resolution with a signal that develops immediately after the moment of

protein translation to overcome issues of maturation time and diffusion, faced by most standard

fluorescent proteins. This signal will ideally also be quantifiable with high spatial resolution, to

enable the quantitative comparison of protein production levels in different subcellular

compartments within the same cell or between different cells.

1.5.5 Split GFPs are indicators of protein interaction

Bimolecular fluorescence complementation using GFP has emerged as a simple solution

and has been repeatedly used in protein and other macromolecular interaction studies (Kerppola,

2006). Split GFP is a GFP molecule that has been split into two nonfluorescent polypeptides that

compose the entire sequence of GFP. By fusing both halves of GFP to interacting proteins, the

close proximity and ensuing physical interaction of the partners results in the spontaneous

reassembly and emission of fluorescence from the reconstituted GFP molecule. However, despite

its widespread use for in vivo imaging, little is known about the mechanism of reconstitution

(Kent et al., 2008).

The most popular split GFP partners are GFP1-10 and GFP11, resulting from a separation

of the peptide chain between beta barrels 10 and 11. The GFP reconstitution across synaptic

partners (GRASP) techniques uses the fusion of the split GFP fragments to synaptic membranes

Page 48: Observation and quantification of protein production in single living cells

36

in distinct neuronal populations (Kim et al., 2011). If two neurons from these populations, each

expressing one of the fragments, interact and form a synapse, then the GFP molecule can

reconstitute and emit fluorescence. The green fluorescent signal thus marks the location synapses

along the cell within tissues. GRASP has been used to map cell contacts and neuronal

connections in Drosophila (Gordon & Scott, 2009) and C. elegans (Feinberg et al., 2008).

Page 49: Observation and quantification of protein production in single living cells

37

1.6 Figures

Figure 1.1 Brightfield image of dissected whole ovarioles

Representative image of dissected ovarioles containing stage 5-6 and 9-10B oocytes. During

oogenesis, a cluster of 16 germ cells interconnected by intracellular ring canal bridges, form the

germline cyst. Fifteen of these cells will become nurse cells and the remaining germ cell

becomes the oocyte (outlined). After fertilization, the oocyte chamber (outlined) grows in size to

support the development of the embryo.

Page 50: Observation and quantification of protein production in single living cells

38

(Lo, Kays et al., 2015)

Page 51: Observation and quantification of protein production in single living cells

39

Figure 1.2 PQR reporters allow quantification of protein production from cells.

(a) Insertion of a Protein Quantification Reporter (PQR) between a fluorescent reporter (GFP)

and a gene of interest creates a polycistronic mRNA for co-transcription and co-translation of

GFP and the gene of interest. The PQR construct allows for one molecule of GFP to be

synthesized for every one protein of interest synthesized. Because of the fluorescence output of

GFP is directly proportional to the concentration of GFP, then the fluorescence intensity of the

cell is used to quantify the level of production of the protein of interest. (b) Because the

fluorescence output of GFP is directly proportional to its concentrations (top panel), then by

using a protein Quantification Reporter sequence to produce a stoichiometric ratio between GFP

and the protein of interest (middle panel), the fluorescence intensity of GFP can be used as a

measure of the production of the protein of interest (bottom panel) (Lo, Kays et al., 2015).

Page 52: Observation and quantification of protein production in single living cells

40

1.7 Thesis introduction

The aim of this dissertation is to develop techniques and reagents that will be used to

explore the process of protein translation in vivo. Since proteins are the last molecular effectors

of change in a cell, and are responsible for the multitude of phenotypes we observe in cells, I am

particularly interested in knowing when, where and how much protein is being produced in the

context of a living animal. Expanding the molecular toolbox by which we can observe cellular

processes is key for our understanding of basic cellular processes. In Chapter 2, I demonstrate

combined protein and RNA measurement from single cells, in a system I have developed to

examine gene expression changes and how they relate to protein levels. In Chapter 3, I discuss

the development and validation of a novel technique to mark protein translation events at

subcellular resolution in living cells. Chapter 4 explores the applications of our technique to

quantifying protein translation events, and where I describe key proof of principle experiments

and reagents required for imaging protein translation in single cells in vivo. The results of this

work contribute to the tools we currently use examine protein production, with specific

advantages in spatial and temporal resolution and quantification accuracy.

Page 53: Observation and quantification of protein production in single living cells

41

Chapter II - Quantification of mRNA and protein

levels in single cells

2.1 Relation to overall thesis

Cells are inherently heterogeneous, and the current tools we use to quantify protein levels

miss important individual cellular differences that are key to understanding cell and molecular

biology. In addition, mRNA is routinely used as a proxy measurement for protein levels,

although this has repeatedly been demonstrated to be inaccurate. Assaying mRNA and protein

expression from single cells can therefore bring a new layer of information on the regulation of

protein expression. To address these issues, I have developed a system to combine mRNA and

protein quantification from the same cell. Single cells edited to co-produce PQR reporters each

time an endogenous protein of interest is translated are first imaged to determine the level of

translation of the protein of interest. The same cell is then lysed and the absolute mRNA levels

encoding the endogenous protein of interest are quantified. In this approach, mRNA and protein

production can be quantified simultaneously for multiple genes, providing valuable metrics that

can be used to: 1) Understand the regulation in expression of either the mRNA or protein of a

particular gene. 2) Correlate changes in mRNA levels with changes in protein levels of a gene. 3)

Reveal insight into the coregulation of multiple genes within the same cell, which can be used to

establish links between the transcriptional and translational regulation of gene products. This

chapter presents an extension of the PQR technique as part of my overall goal to develop

techniques to understand protein production at single cell resolution, by looking at concurrent

changes in mRNA production.

Page 54: Observation and quantification of protein production in single living cells

42

2.2 Introduction

Understanding the relationship between genes and phenotypes is a central component of

biology. Profiling the expression of genes with respect to messenger RNA (mRNA) or protein

products has advanced our fundamental understanding of cell biology, such as the maintenance

of cell structure (Gronowicz et al., 1992), progression of cell cycle (Ly et al., 2014), neural

development (Cáceres & Nilson, 2005), as well as how abnormalities in gene expression can lead

to disease states (Wong et al., 2001). Several analytic methods have been developed to profile

gene expression at the level of mRNA or protein. For example, microarrays and deep sequencing

technologies can identify mRNA from tens of thousands of different genes, whereas the

expression and abundance of proteins can be assayed using spectrometry, chromatography and

immunoassays (Greenbaum et al., 2003; Maier et al., 2009).

The current tools to assay mRNA and protein expression are generally performed on

large pools of cells or tissues (Wu & Singh, 2012a). Large numbers of cells increase the

reliability of results and overcome issues of low sensitivity and low signal to noise ratio of

common assays. However, it is becoming increasingly clear that even neighboring cells within

tissues or pools of clonal cells are not homogenous (Schubert, 2011). Cellular dynamics, such as

differing transcriptional and translational states, cell signaling, cell cycle, development, and other

molecular processes in addition to stochasticity together produce cellular heterogeneity, even in

immortalized clonal cell lines (Stockholm et al., 2007). Therefore, assays performed on large

numbers of cells cannot be translated into the single-cell scale (Schubert, 2011; Wu & Singh,

2012b). As described in Chapter 1, individual cells are inherently heterogeneous, therefore

analysis of single cells is required to identify molecular mechanisms that are masked by the

population average.

Page 55: Observation and quantification of protein production in single living cells

43

The traditional and more convenient quantification of mRNA levels as a proxy for

protein abundance implicitly assumes that changes in mRNA levels directly correspond to

changes in protein levels, and thus has biological relevance. This has repeatedly been shown to

be inaccurate, as more global analyses of transcriptomics and proteomics are revealing

surprisingly poor correlations in mRNA-protein level concordance (Greenbaum et al., 2003;

Gunawardana et al., 2015; Maier et al., 2009; Schwanhäusser et al., 2011). These findings create

a concern when making inferences only from mRNA data. The utility of mRNA expression

studies is clear, particularly for understanding mechanisms of transcription and in the continued

development of sensitive preclinical disease markers (Abrahamsen et al., 2004; Mikaelian et al.,

2013). However, the interpretation of quantitative mRNA results with respect to cellular

phenotypes is merely correlative. Therefore, protein expression directly needs to be examined to

understand causative change within cells. Taken together, this suggests that assaying mRNA and

protein expression together from a single cell overcomes cellular heterogeneity issues, and

protein data can be used to validate mRNA findings. In addition, the availability of mRNA data

from the same cell adds another level at which the regulation and correlation of mRNAs and

proteins can be examined. Examining these dynamics allows understanding of how changes in

mRNA levels contribute to the level of protein that ultimately generates cellular phenotypes.

We have previously developed a technique that allows for quantification of protein levels

in single cells in vivo (Lo, Kays et al., 2015). This Protein Quantification Ratioing (PQR)

technique uses a genetic tag that produces a stoichiometric ratio of a fluorescent protein reporter

and the protein of interest during protein translation. The fluorescence intensity is proportional

to the number of molecules of protein of interest produced and is used to determine the relative

protein amount within the cell (Figure 1.1a). Using genome editing tools, PQR constructs can be

Page 56: Observation and quantification of protein production in single living cells

44

inserted into any endogenous genomic locus to quantify endogenous protein levels in single

living cells (Lo, Kays et al., 2015).

This chapter describes a fast and accessible way for researchers to quantify the

expression of a gene of interest by measuring both its absolute mRNA transcript numbers and

protein production levels in the same cell (Figure 2.1). This tool can be used by any researcher

with access to a standard fluorescence microscope and a real-time PCR thermocycler and can

easily be applied to assay single cells obtained from various sources such as dissociated animal

and tissue samples, as well as FACS sorted and patient-derived stem cells. In addition, this

protocol is amenable to various cell types such as neurons, immune cells, immortalized cell lines

and dissociated animal or plant cells, if they are amenable to common genetic and biochemical

assays. By first measuring the level of protein translation using a PQR fluorescent reporter and

then quantifying the absolute mRNA copy number of that gene using RT-qPCR (Nolan et al.,

2006), the transcriptional and translational expression of a gene can be accurately determined at

the single cell level (Figure 2.1). This approach can be applied to examine the expression of

multiple genes at once, which results in an overarching view of the regulation of transcript and

protein expression of multiple genes in the same cell (Figure 2.2).

2.3 Experimental design and detailed protocol

This protocol describes a system for quantification of gene expression at the mRNA and

protein level from a single cell. We chose the mouse monoclonal antibody cell line 22c10 as a

source of single cells (Fujita et al., 1982). 22c10 hybridoma cells are large, easy to manipulate

and secrete high levels of antibody containing kappa-isotype immunoglobulin light chains (IgK).

Antibodies are composed of two heavy chains and two light chains. The genomic locus encoding

the kappa light chain contains a constant region that is not affected by genomic rearrangements

Page 57: Observation and quantification of protein production in single living cells

45

that occur following immune responses, and is thus common to all light chains produced from

the IgK locus (Janeway et al., 2001; Lodish et al., 2008). The insertion of a PQR-RFP reporter

within the constant region of the kappa light chain ensures the production of a molecule of RFP

with each kappa light chain translated, at high levels (Figure 2.3a). In contrast, we chose the

Rpl13a gene which encodes for Ribosomal Protein L13A and is stably expressed in all

eukaryotic cells at consistent levels. Due to its consistent expression across tissues, Rpl13a is

routinely used as a reference “housekeeping” gene for normalization of quantitative mRNA and

protein measurements (Curtis et al., 2010; Gubern et al., 2009; Mane et al., 2008). The

quantification of Rpl13a, by insertion of a PQR-GFPnols reporter at the end of its coding

sequence (Figure 2.3b), can therefore be used as a measure of a cell’s individual translational

status (Lo, Kays et al., 2015). In the following section, I present the protocol we use to take

advantage of the differences between IgK and Rpl13a expression, to highlight how this technique

can be used to profile the transcriptional and translational landscape of multiple genes

simultaneously at single cell resolution (Figure 2.2).

2.3.1 Materials

Reagents

Caution: Trizol reagent and chloroform are toxic and cause harm to skin and airways. Protect

skin and eyes and handle stock bottles in a fume hood.

• Serum free culture medium, e.g. H-Cell (Wisent, cat. no. 000-035-CL)

• Nuclease free water (Wisent, cat. no. 809-115-CL)

• Nuclease-free Glycoblue (Thermo-Fisher, cat. no. AM9515)

• RNase OUT (40 U/µl) (Thermo-Fisher, cat. no. 10777019)

• Superscript IV Reverse Transcriptase 200 U/µl (Thermo-Fisher, cat. no. 18090050)

• Taqman Fast Advanced 2X Mastermix (Thermo-Fisher, cat. no. 4444558)

Page 58: Observation and quantification of protein production in single living cells

46

• Trizol reagent (Thermo-Fisher, cat. no.15596018)

• Chloroform (Sigma, cat. no. 288306)

• Isopropanol (Sigma, cat. no. I9516)

• 75% ethanol prepared with nuclease-free water

• Gene-specific reverse transcription primers (Integrated DNA Technologies)

• Real-time polymerase primers and probes (Integrated DNA Technologies)

Equipment

Caution: Pipetting by mouth can be hazardous. Appropriate tubing filters must be used to

protect against accidental liquid or particulate inhalation. However, in our experience the

microliter volumes handled using a mouth pipette in this protocol are too small to pose any risk

of crossing the filter or the length of tubing, and the aspirated solutions are not generally

harmful. Nevertheless, a 1 ml syringe can be used instead of the mouth pipette, but much less

control over pipetting is achieved.

• Aspirator assembly unit for mouth pipette (Sigma, cat. no. A5177)

• 0.22 µm tubing filter (Millipore, cat. no. SLGP033NB)

• Teflon-coated patterned glass slides, 10-15 mm diameter spots (EMS, cat. no. 63426-06)

• 96-well qPCR microplates (Thermo-Fisher, cat. no.4346907)

• High optical clarity sealing film (Sarstedt, cat. no. 95.1999)

• 8-strip PCR tubes (Diamed, cat. no. DIATEC420-1402)

• Nuclease-free 1.5 mL microfuge tubes (Thermo-Fisher, cat. no.AM12400)

• 1.5 x 1.1 mm O.D/I.D borosilicate glass capillaries without filament (Sutter Instruments, cat.

no. B150-110-10)

• Glass micropipette puller e.g. P80/PC (Sutter Instruments)

• Microscope with excitation source and appropriate fluorescence emission filters (For example

(excitation/emission): GFP (488nm/515nm), RFP(543nm/594nm)

• Imaging camera e.g. Axiocam MRm CCD (Zeiss)

• Real-Time PCR System, e.g. StepOnePlus (Applied Biosciences)

Page 59: Observation and quantification of protein production in single living cells

47

2.3.2 Gene editing using CRISPR-Cas9

In order to target fluorescent PQR reporter constructs into endogenous genomic loci for

protein quantification, we used Clustered Regularly Interspaced Short Palindromic Repeats-Cas9

(CRISPR-Cas9) genome editing, which uses an RNA-guided Cas9 endonuclease that can be

targeted to induce double-strand breaks (DSB) in any known DNA sequence with pinpoint

sequence accuracy (Jinek et al., 2012). DNA DSBs can be repaired by the cellular machinery

using the homologous recombination pathway, which can be exploited to insert exogenous DNA

sequences into genomic loci (Mao et al., 2008).

To insert PQR-XFP reporters into the endogenous IgK or Rpl13a locus, we designed and

generated several CRISPR/Cas9 targeting vectors that cut the endogenous loci directly upstream

of the stop codon (Figure 2.3a). Co-transfection of IgK or RPL13A-specific repair plasmids into

22c10 cells results in a DSB, which stimulates the DNA repair machinery to recombine the PQR-

XFP fragment in-frame at the end of the coding sequence of the endogenous locus. When

successful integration of the PQR-XFP construct occurs, mRNAs transcribed from the

endogenous RPL13A or IgK locus will be slightly longer in size (~850 extra nucleotides), but

will preserve the 5’ and 3’ untranslated regions (UTRs) and native coding sequence, which

contain elements required for proper mRNA export, localization, stability and translation (Figure

2.3b). For every one molecule of endogenous RPL13A or IgK protein produced, there will be

one corresponding molecule of fluorescent reporter co-translated. Since the reporter and the

protein of interest are translated from one open reading frame, the fluorescence intensity of the

reporter (i.e. brightness of the cell) can be used as a measure of how much protein is being

produced (Figure 2.1) (Lo, Kays et al., 2015).

Page 60: Observation and quantification of protein production in single living cells

48

The repair templates used to repair the genomic DSB consist of a PQR-XFP construct

placed between two 1.0 kB-long homology arms specific to genomic IgK or RPL13A. The

homology arms lack any promoters or transcriptional activators, which prevents expression of

the PQR-XFP until in-frame genomic integration within an active coding gene. The arms align

the repair plasmid to the endogenous locus via molecular homology, and the DNA repair

machinery recombines the exogenous PQR-XFP into the break region, directly upstream of the

endogenous stop codon (Figure 2.3b).

We screened several CRISPR targeting vectors and repair templates for optimal nuclease

and recombination activity. Validation of the correct integration of PQR-GFP into the IgK locus

was achieved by genomic DNA extraction six days post-transfection and genotyping using

primers that lie outside and within the homology arms of the repair template (Figure 2.4a). The

5’ and 3’ ends of the endogenous IgK locus are probed using PCR with these two sets of primers

(Figure 2.4c). The PCR products are then digested using restriction enzymes specific for sites

within the PQR construct. In parallel, the edited genomic locus is sequenced to identify the PQR

and genomic junctions. The same procedure is repeated to validate the insertion of PQR-XFP

into a second locus (Lo, Kays et al., 2015).

To generate 22c10 cells carrying PQR reporters at both the endogenous Rpl13a and IgK

loci, we first established a stable line with a PQR-GFPnols at the Rpl13a locus, and then targeted

the IgK locus with a PQR-RFP reporter. Our preliminary experiments using an IgK-PQR-GFP

edited line (Figure 2.4) showed that the fluorescence intensity of a RFP-based red Rpl13a

reporter was too dim to be observed without a camera. High levels of GFP (from IgK), in

addition to overlap in the spectra of GFP and RFP led to high background fluorescence levels in

the red channel, posing a burden on the manual isolation of fluorescent cells and quantification

Page 61: Observation and quantification of protein production in single living cells

49

of fluorescence intensities. We therefore swapped the reporters and chose to include a nucleolar

(nols) targeting sequence to the GFP in order to sequester the GFP to nucleoli in Rpl13a-edited

cells, a cellular organelle 100 times smaller than the cytoplasm in volume (M. S. Scott et al.,

2010). Sequestering fluorescent proteins into small cellular organelles increases their local

concentration which results in a brighter signal, facilitating fluorescence quantification and cell

manipulation. Indeed, we found that using a green reporter for Rpl13a and a red reporter for IgK

produced cells that displayed clear green and red fluorescent signals (Figure 2.1b).

2.3.3 Single-cell protein level quantification

To determine the level of production of IgK or Rpl13a protein within a cell, genome

edited 22c10 cells carrying red and green fluorescent PQR reporters, respectively, were used

(Figure 2.3). Immediately prior to the experiment, a concentrated single cell suspension (> 5x105

cells/ml) was prepared and placed on ice to delay any changes in cell metabolism or gene

expression. Using freshly pulled wide-bore micropipettes (Figure 2.5a), a single red and green

fluorescent cell was pipetted (Figure 2.5c) into a droplet of culture medium on Teflon-coated

glass slide (Figure 2.5d). A high magnification objective was dipped into the drop of culture

medium, forming a meniscus around the lens of the microscope (Figure 2.5d), and the cell was

brought into focus to image the fluorescence intensity of the cell (Figure 2.5e). It is important to

use excitation and acquisition settings that have been previously determined by imaging control

cells to establish the range of minimum and maximum fluorescence levels that can be reasonably

expected from the cell line used to minimize imaging cells that are too dim or those that saturate

acquisition settings.

2.3.4 Total RNA extraction

Page 62: Observation and quantification of protein production in single living cells

50

After quantitative fluorescence imaging, the cell was immediately transferred using a new

clean micropipette to a microfuge tube prefilled with 200 µL of the phenolic reagent, Trizol

(Figure 2.5f). It is important to change the micropipette after each sample to avoid cross

contamination by nucleic acids or cells that may have adhered to the inside surface of the

micropipette glass. The cell is immediately lysed once in the phenolic solution and its nucleic

acid and protein content are separated into a top aqueous layer containing nucleic acids and a

bottom organic layer containing protein and cellular debris. One volume of chloroform was then

added to further eliminate contaminating phenol via a second phase separation and dilution of the

phenol. The total RNA dissolved in the aqueous layer was then recovered by precipitation with

isopropanol and subsequent centrifugation at ≥ 12,000g. Glycogen is an inert polysaccharide that

is often used as a carrier to increase RNA recovery by co-precipitating with nucleic acids,

effectively trapping precipitated RNA molecules in large sugar precipitate complexes. Glycoblue

is glycogen that has been conjugated to a blue dye, and addition of 5 µg of Glycoblue prior to

isopropanol improved RNA recovery without compromising subsequent transcription and

amplification steps. The RNA pellet obtained after centrifugation was washed twice with 70%

ethanol that has been prepared with RNase-free water, and then dried at room temperature for ~5

minutes, or until most of the ethanol was evaporated. The pellet was resuspended in 6 µL of

RNase-free water and this total volume was used as input RNA for the subsequent reverse-

transcription (RT) reaction.

2.3.5 Reverse-transcription

Superscript IV was chosen as a reverse transcriptase. It is a modified Moloney murine

leukemia virus (M-MLV) reverse transcriptase, a variant of thermostable reverse transcriptases

(Das et al., 2004). Thermostability allows RT reactions to be carried out at high temperatures at

Page 63: Observation and quantification of protein production in single living cells

51

which RNA is linear and secondary structures cannot form. Superscript IV has also been shown

to outperform other commercial enzymes in terms of reproducibility in the presence of inhibitors

commonly present in RNA samples such as carryover phenol, isothiocyanates and ethanol

(Suslov & Steindler, 2005), which pose a concern particularly in this protocol. To produce

complementary DNA (cDNA) from mRNA isolated from single cells, a mixture of gene-specific

primers (final concentration 0.1 µM) was used to specifically reverse transcribe edited IgK and

Rpl13a mRNA transcripts in a 15 µL RT reaction containing 30 units of RNase inhibitor and 150

units of Superscript IV enzyme. Gene-specific RT primers have been shown to be the most

sensitive and specific approach to convert target mRNA into cDNA, and produce less variability

compared to oligo-dT or random hexamer primers (Stephen A Bustin & Nolan, 2004). The

reaction was incubated at 55°C for 50 minutes to allow the RT reaction to proceed.

2.3.6 Real-time polymerase chain reaction

By titrating the amount of input cDNA comparing 1 µL, 3 µL and 6 µL, we found that

using 6 µL of cDNA from the RT reaction is optimal for obtaining high quality and reproducible

amplification without significant interference from the RT reaction mixture (Figure 2.6a). The

rationale was to determine the minimum amount of cDNA that could be used in the qPCR

reaction and the advantages are twofold: First, to reduce the total amount of glycerol, a common

component in RT and qPCR enzyme mixtures that is inhibitory beyond 20-30% v/v (Yasukawa

et al., 2010). Second, to spare as much RT mixture as possible in order to repeat failed

amplifications or to run replicate samples during optimization and control reactions. The

TaqMan Fast Advanced mastermix (ThermoFisher), which contains the AmpliTaq fast DNA

polymerase, was used to amplify the IgK and Rpl13a cDNA molecules. 20 µL reactions were

prepared and loaded into a 96-well optical plate sealed with high-optical clarity sealing film; the

Page 64: Observation and quantification of protein production in single living cells

52

plate was briefly centrifuged and loaded into the thermal cycler which has been preprogrammed

to define the cycling parameters and well assignments. Thresholding of amplification and

calculation of cycle threshold (Ct) values were performed automatically using the StepOnePlus

software. Ct values are inversely proportional to starting target concentration, with low Cts being

indicative of high concentrations of input DNA.

2.3.7 Calculation of absolute mRNA transcript number

qPCR standard curves are plots of obtained Ct values versus series of diluted known

standards that are used to convert Ct data into quantity values. Analysis of standard curves

provides a substantial amount of information about the qPCR assay. In addition to providing a

calibrator sample of known amounts for the absolute or relative quantification of mRNA

concentrations from unknown test samples, variability in the reverse-transcription reaction and

general interexperiment variability resulting from running multiple samples on different plates

can be accounted for and corrected (Nolan et al., 2006; Stahlberg & Bengtsson, 2010). The

inclusion of a standard curve with each experiment is an important control to normalize

quantitative measurements and to ensure the accurate comparison of data points.

The standard curve was generated by plotting Ct (y-axis) against the logarithm of

template quantity (x-axis) (Figure 2.7). Ct values obtained from single cell experiments could

then be compared to this curve to determine the quantity of starting nucleic acid material. A

linear dynamic range was normally seen over at least 6 logs of dilutions, which essentially

defines the working range of accurate quantification of the assay (Larionov et al., 2005).

In this single cell assay, standard amounts of DNA covering a range of approximately 10

to 107 copies were prepared to minimize any extrapolation outside of the predicted dynamic

Page 65: Observation and quantification of protein production in single living cells

53

range (Figure 2.7) (Ståhlberg & Kubista, 2014; Svec et al., 2013). It is ideal to amplify the serial

dilutions of standards in duplicate to determine the reproducibility of the assay. In addition, the

template used to generate the standard curve should accurately reflect the complexity of

experimental samples, ideally by using a total RNA preparation from the investigated cells or

using a known amount of pure cDNA, plasmid DNA or synthetic oligonucleotides containing the

target of interest (Dhanasekaran et al., 2010). Synthetic oligonucleotides containing the assayed

target region are easy, fast and inexpensive to commercially synthesize. In addition, commercial

synthetic oligonucleotides are usually highly purified samples of known concentration.

2.3.8 Readout of amplification

The qPCR assay uses the TaqMan chemistry principle for the detection of amplified

transcripts. The fluorescent signal originates from a 20bp DNA probe that contains a conjugated

fluorophore and a chemical quencher that hybridizes to the amplified target region (Holland et

al., 1991). Exonuclease activity of the passing DNA polymerase releases the quencher from the

probe. The otherwise quenched fluorophore can now change conformation and emit

fluorescence, which is then detected by the real-time PCR thermal cycler detectors. The

fluorescence intensity of the reaction during each cycle of the reaction is measured by the

machine and an automatic thresholding algorithm detects the cycle threshold (Ct), the cycle at

which the fluorescence intensity and its rate of increase reach a significant level above the noise

(S. Zhao & Fernald, 2005). Cts are inversely proportional to starting DNA concentration (Figure

2.6, Figure 2.7) and thus to convert a Ct value into number of mRNA molecules, the use of a

standard curve is required (Nolan et al., 2006).

Page 66: Observation and quantification of protein production in single living cells

54

In addition to the commonly used positive and negative qPCR amplification controls, it is

important to regularly include additional controls to monitor any drift in technical or

experimental variability.

2.3.9 Assay controls

The qPCR assay used must be optimized to produce the most sensitive and reliable

amplification (Tichopad et al., 2009). For the PCR reaction to produce twice as much product

during each cycle, the amplification of product must be as close to 100% efficient as possible, as

this results in the most sensitive and reliable quantification (Taylor et al., 2010). The efficiency

of the assay is most easily determined by using the equation efficiency = 10(-1/slope). The slope is

obtained from standard curves generated by amplifying serial dilutions of target template under

different primer and probe concentrations. The slope of our standard curve was calculated to be -

3.7, resulting in 86% efficiency (Figure 2.7).

Variability in the instrument and detectors, reaction setup, and stability of reagents can

contribute to erroneous quantification. The slope and y-axis of the standard curve can be used as

metrics to monitor and adjust for variability. Therefore, for the highest quantification accuracy, a

standard curve should be prepared and included on the same plate each time experimental

samples are assayed. Moreover, it is important to use the minimum amount of cDNA that

produces the most sensitive and highest quality amplification, to reduce any inhibitory effects

caused by components in the RT reaction mixture (Opel et al., 2010).

The reproducibility of a qPCR assay is reflected by the variability between technical

replicate reactions, as determined by the R2 of the best fit line of standard curve data. R2 values

of 0.98 or higher are indicative of a stable and reproducible assay (Nolan et al., 2006). Our

standard curve resulted in R2=0.99 (p<0.05) (Figure 2.7).

Page 67: Observation and quantification of protein production in single living cells

55

This protocol uses phenol-based purification of RNA which biochemically separates

away genomic DNA, however in some cases, genomic DNA may contaminate the obtained RNA

sample. Genomic DNA controls are thus crucial to determine the contribution of contaminating

genomic DNA to the observed amplification (Stephen A. Bustin et al., 2009). Although in the

case of single cells contamination by genomic DNA has minimal effect (2 extra copies of the

target in a diploid cell), some sensitive applications require maximum quantification accuracy.

Since this protocol requires that the entire RNA content of the cell be used in the RT reaction, it

is useful to randomly allocate some samples for genomic DNA contamination controls. The

contribution of contaminating genomic DNA to the observed single cell Ct can be determined by

performing a no-RT control reaction, where the reaction is assembled but the RT enzyme is

omitted (Figure 2.8). Any observed amplification can then be attributed to genomic DNA that

has been carried over during RNA purification. The contribution of genomic DNA from single

cells using this protocol was found to be negligible (in > 90% of randomly tested samples)

demonstrating that genomic DNA can be reliably separated from the aqueous RNA during Trizol

purification (Figures 2.8a-b). It is however useful when first performing this protocol or when

using different cell types to dedicate some samples for performing no-RT controls to determine

the contribution of contaminating genomic DNA. qPCR results are generally acceptable if the

cycle thresholds of genomic DNA control samples are at least 5 Ct larger than those of test

samples (S. Bustin & Nolan, 2004; Nolan et al., 2006). However, we found that most samples

contain no detectable genomic DNA, even when the PCR was programmed to run for 50 cycles

(data not shown).

2.3.10 Image acquisition and analysis

Page 68: Observation and quantification of protein production in single living cells

56

Fluorescence and brightfield microscopy was performed using a Zeiss AxioScope A1. All

images were acquired at 1388 x 1040 pixels using a 40× water objective, N.A. 1.0

(epifluorescence) (Figure 2.5d). Fluorescence emission was detected using a charge-coupled

device (CCD) camera (MRm). All image acquisition parameters were fixed for each imaging

channel for exposure time, excitation intensity and gain. Cells that were dimmer or brighter than

the fixed initial acquisition dynamic range were not included for analysis.

Images were selected for analysis based on identification of single healthy

(morphologically) cells and low background. Fluorescence pixel intensities were measured in

several random regions of interest (ROIs) within the cell cytoplasm (IgK) or nucleolus (Rpl13a)

using ImageJ. Average pixel intensities were calculated from five ROIs of 10x10 pixels for

measurements within the cytoplasm and 5x5 pixels for measurements within nucleoli (Figure

2.1b). All signal intensities were background subtracted from the average of three 10x10 pixel

ROIs surrounding the cell.

2.4 Results

The PQR fluorescent reporters are translated stoichiometrically each time Rpl13a or IgK

protein molecules are produced. The fluorescence intensity of the cell (in arbitrary units) is

proportional to the concentration of fluorescent reporter which is proportional to the production

rate of the protein of interest. Therefore, the fluorescence intensity or brightness of the cell can

be used as a measure of the amount of production of the protein of interest (Lo, Kays et al.,

2015). Similarly, the cycle threshold of the real time amplification reaction is proportional to the

concentration of input nucleic acid, and this is used to quantify the relative or absolute levels of

the target mRNA (Nolan et al., 2006). Measuring the fluorescence intensity of the cell and Ct

values from the qPCR amplification can therefore be used to assay gene expression of any

Page 69: Observation and quantification of protein production in single living cells

57

gene(s) from the same cell and determine the variation at both the transcriptional and

translational levels.

To determine the levels of IgK and Rpl13a protein in single 22c10 cells, we used

CRISPR-Cas9 to generate a cell line carrying a PQR-RFP and PQR-GFPnols insertion at the

endogenous IgK and Rpl13a genomic loci, respectively. Single cells fluorescent in both the red

and green channels were chosen and isolated to derive single cell clones. The relative levels of

IgK and Rpl13a protein production were determined by quantifying the nucleolar GFP

fluorescence and cytoplasmic RFP fluorescence intensities from a single cell, respectively. The

same cell was then lysed and the absolute abundance of IgK and Rpl13a mRNA transcripts were

determined using quantitative real time PCR (Figure 2.1).

Using quantitative imaging, we found that Rpl13a protein levels, as measured by

nucleolar GFP fluorescence intensity, ranged from 91 to 842 a.u. (arbitrary units) with an

average of 364 ± 205 (mean ± SD; n=26). The frequency distribution of Rpl13a protein levels

showed a peak at 150 a.u, and was broad and asymmetrical (Figure 2.9a). Specifically, over 60%

of 22c10 cells had GFP intensities of less than 500 a.u., and the remaining cells distributed in a

short tail of higher protein levels. There appeared to be only one peak in the distribution of

Rpl13a protein levels. In contrast, cytoplasmic RFP fluorescence intensities ranged from 936 to

3,263 a.u, with an average of 2,004 ± 619 a.u (n=24). The peak value was around 2,000 a.u., and

the overall shape of the distribution was more normal than that of Rpl13a (Figure 2.9b).

Using a standard curve of serially diluted known amounts of IgK and Rpl13a synthetic

oligonucleotide DNA templates, we determined the absolute number of IgK and Rpl13a mRNA

molecules (Figure 2.9c-d). We found that the absolute number of Rpl13a mRNA transcripts in

single 22c10 cells ranged from 287 to 3,023 transcripts, with an average of 1,156 mRNA

Page 70: Observation and quantification of protein production in single living cells

58

transcripts, giving the distribution of measured Rpl13a mRNA numbers a sharp peak at 960

transcripts (n=17) (Figure 2.9c). Similarly, absolute IgK mRNA numbers were found to range

from 2,420 to 159,000 transcripts, with an average of 63,835 IgK mRNA transcripts per cell

(n=25) (Figure 2.9d). This is expected as 22c10 cells are designed and selected to express high

levels of kappa light chain (Li et al., 2010; Lo & Gillies, 1991), while Rpl13a is a non-essential

housekeeping gene (Zhou et al., 2015).

To determine the relationship between gene expression at the transcriptional and

translational levels, we analyzed the linear correlations between mRNA and protein levels of IgK

and Rpl13a and found interesting differences that shed light on the variety of mRNA to protein

relationships that exists for different genes. Pearson’s linear correlations were calculated by

fitting the data to a simple linear regression model, with the coefficient of determination, R2. We

tested the null hypothesis that the variables were independent of each other and that the true R2

value was 0. Rpl13a mRNA levels did not correlate with Rpl13a protein levels, with a Pearson’s

correlation coefficient of 0.1 (p>0.05, n=17) (Figure 2.10a and e). In contrast, IgK mRNA levels

were a better predictor of IgK protein levels, with a correlation coefficient of 0.44 (p<0.05,

n=24) (Figure 2.10b and e), consistent with the generally reported power of predicting protein

abundances from mRNA levels (Greenbaum et al., 2003; Maier et al., 2009; Schwanhäusser et

al., 2011).

To determine whether changes in Rpl13a protein can predict changes in IgK protein, the

red and green fluorescence intensities in single 22c10 cells were compared. We found no

significant linear correlation (R2=0.14, p>0.05, n=24) which indicates that the levels of Rpl13a,

do not correlate with changes in the levels of IgK protein (Figure 2.10c), consistent with

Rpl13a’s role as a housekeeping protein. In contrast, to determine whether the expression of IgK

Page 71: Observation and quantification of protein production in single living cells

59

and Rpl13a mRNA might be linked, we compared absolute IgK and Rpl13a transcript numbers.

We found that Rpl13a transcript numbers were a better predictor of IgK mRNA levels (R2=0.3,

p<0.05, n=17) (Figure 2.10d), compared to protein levels. Taken together, these results

demonstrate that Rpl13a mRNA is likely to vary more in a single cell with other mRNAs under

normal conditions, suggesting that Rpl13a is more stable as a housekeeping normalization

control at the protein level.

In order to determine whether the regulation of Rpl13a mRNA and protein levels is

preserved across species, we generated human embryonic kidney 293 (HEK293) cells carrying a

PQR-GFP insertion at the endogenous human RPL13Aa (hRPL13A) locus. Using the same

approach, we found that hRPL13A protein levels ranged from 60 to 240 a.u. (compared to 91 to

842 a.u for mouse Rpl13a), and absolute hRPL13A mRNA transcript numbers ranged from 11 to

192 (compared to 287 to 3,023 mouse Rpl13a transcripts), resulting in a similarly poor inverse

correlation between hRpl13a mRNA and protein levels (R2=0.12, p>0.05, n=25) (Figure 2.10f).

These results demonstrate similar mRNA-protein relationships for RPL13a between the human

and mouse genomes.

2.5 Discussion

The genome-wide correlation between mRNA and protein levels in cell populations is

recognized to be poor, with mRNA predicting protein levels with only 40% power

(Schwanhäusser et al., 2011). This is typically attributed to regulatory mechanisms that can

independently affect each step of gene expression. Measuring and correlating the levels of IgK

and Rpl13a mRNA and protein from single 22c10 cells shows exactly how different genes can

have different mRNA-protein relationship dynamics: IgK mRNA level was a better predictor of

IgK protein levels (R2=0.4, p<0.05, n=24), compared to Rpl13a mRNA to protein levels

Page 72: Observation and quantification of protein production in single living cells

60

(R2=0.1, p>0.05, n=17) (Figure 2.10e). In addition, Rpl13a mRNA was more likely to co-vary

with other mRNAs, compared to Rpl13a protein variation (Figures 2.10c and e).

This protocol also allows the examination of relationships between the expression of

different genes and their products. Since this technique yields two metrics per gene, assaying

multiple genes allows the co-correlation of mRNA and protein levels across different genes. This

can be used to study the regulation and differential expression of mRNAs and proteins of

multiple genes at the single cell level. One analogous example is the use of housekeeping genes

for the normalization of quantitative mRNA and protein measurements. In this case changes in

IgK mRNA or protein levels may be normalized to the mRNA and proteins levels of Rpl13a, a

commonly used housekeeping gene for normalization of expression. For example, this knockin

22c10 cell line can be treated with drugs and then screened for cells that exclusively upregulate

or downregulate the expression of IgK antibodies, without affecting normal cellular

housekeeping functions, as would be assayed using Rpl13a. Similarly, this technique can be used

to examine whether different genes are co-regulated within a single cell, under different

treatment or disease conditions. Interestingly, our results revealed a large variability in Rpl13a

transcript numbers (mean ± S.D = 1156 ± 655), making it a poor normalization standard for

mRNA quantitation in 22c10 cells. While Rpl13a has been shown to be more stable than other

commonly used reference genes such as Gapdh or Hprt (Ling et al., 2011), our results show that

such universal conclusions cannot be drawn across all cell types and studies performed at the

single cell level.

The protocol is limited by the time required by the experimenter to collect and assay the

individual cells. The protocol can be improved by faster collection and imaging of cells using

automated or microfluidic chambers, and using non-precipitation approaches to isolate the total

Page 73: Observation and quantification of protein production in single living cells

61

RNA. However, in its present form the protocol requires 5 days to perform CRISPR experiments

and initiate clonal selection of cells carrying PQR reporters at any endogenous genomic locus

and ~4 h from start of imaging to qPCR amplification. In addition, the protocol does not require

uncommon reagents or equipment and is accessible to a wide range of researchers. Because of

the small quantity of starting template, certain low copy-number genes may fall below the

detection range of the reverse-transcriptase or PCR polymerase enzymes, and the success and

quality of the amplification depends on the optimization of the primer and probe conditions, in

addition to the quality of the template. With 85% efficiency primers, our results show that the

average amplification failure rate is around 2-3 in 10 cells, and is target-dependent (9 failed

Rpl13a reactions vs 0 failed IgK reactions). Such results can be explained by variability in the

RT step between IgK and Rpl13a RT primers and by the substantially lower number of Rpl13a

transcripts (mean= 1,156 transcripts) as compared to IgK (mean= 61,282 transcripts). The failure

rate may be improved with more efficient primers, and this further increases the reproducibility

of the assay.

The PQR technique is a translational reporter that marks when proteins are produced

within cells, however the differences in maturation and turnover rates between the PQR reporters

and the assayed genes may in some cases hinder accurate protein quantification. While we used

our most optimized PQR sequences that produce equimolar products and result in virtually no

fusion product (Lo, Kays et al., 2015), it cannot be guaranteed that within the cell the levels of

the reporter and protein of interest are equimolar at all times. Differences in folding, maturation

time and the inherent turnover rate of proteins can act differently on the reporter and protein of

interest. This intuitively suggests that protein quantification using PQR provides a relative

measure of protein production at the time of imaging, and it should be noted that PQR does not

Page 74: Observation and quantification of protein production in single living cells

62

measure absolute protein levels (Lo, Kays et al., 2015). However, several recent quantitative

protein studies indicate that effects such as protein stability, half-life and degradation account for

a little over 5% of the variation in protein abundance (Li et al., 2014; Schwanhäusser et al.,

2011). Using pulse-chase labelling of newly synthesized proteins, the abundance of a protein in a

given cell was determined to be mainly dictated by the abundance of its mRNA and rate of its

translation, accounting for 40% and 50 % of the variance, respectively (Schwanhäusser et al.,

2011). Since the protein of interest and the reporter are produced from the same mRNA molecule

by the same ribosome, the main source of variability between their levels is the difference in

their turnover rates. This means that the fluorescence intensity of the reporter can be reliably

used to readout the production of the protein of interest with an acceptable error margin.

Differences in turnover rate and stability merely change the slope of the linear relationship

between fluorescence intensity and protein production, as we have extensively demonstrated

using experiments and simulations in Lo, Kays et al., 2015.

2.6 Conclusion

Analysis of gene expression at the transcriptional and translational level in single cells

provides maximal resolution into the dynamics of gene expression. Combining novel techniques

with recent advances in genome editing has opened the door for the interrogation of the

endogenous genome, transcriptome and proteome. Tools such as this combined RNA and protein

measurement approach will reveal previously hidden molecular mechanisms that govern the

expression of proteins and phenotypes. For example, insertion of spectrally different PQR

reporters into different alleles of the same gene can be used to dissect out the allelic contribution

and balance in the expression of a gene (unpublished). In addition, insertion of PQR reporters

into endogenous disease loci can result in a reporter that can be used to readout mRNA and

Page 75: Observation and quantification of protein production in single living cells

63

protein responses to drugs or therapies (unpublished). Finally, this protocol can easily be applied

by any researcher who wishes to examine how well they can use the mRNA levels of their gene

of interest as a proxy for protein abundance. Differential mRNA expression studies are still faster

and more economical to perform than protein expression studies, and have a higher multiplex

potential, and with this technique our current arsenal of mRNA markers can be reduced to the

most informative and predictive transcripts.

Page 76: Observation and quantification of protein production in single living cells

64

2.7 Figures

Figure 2.1 Workflow of protein and mRNA measurement from the same cell.

Page 77: Observation and quantification of protein production in single living cells

65

(a) Insertion of a PQR reporter between a fluorescent reporter (RFP) and a gene of interest

results in the stoichiometric cotranslation of a fluorescent reporter which can be used to quantify

endogenous protein production levels. PQR constructs allow for one molecule of RFP to be

produced for each molecule of protein of interest translated. Since the fluorescence output of

GFP is directly proportional to its concentration, the fluorescence intensity of the cell can be

used to determine the level of production of the protein of interest. (b) Quantitative fluorescence

intensity measurement from regions of interest within the cell can be used to quantify the

production level of the upstream protein. Following imaging, the same cell was lysed for total

RNA extraction and absolute mRNA abundance quantification using single cell qPCR.

Page 78: Observation and quantification of protein production in single living cells

66

Figure 2.2 Protein and mRNA measurement for multiple genes in a single cell.

PQR constructs carrying an RFP and a GFPnols reporter were inserted into the endogenous loci of

IgK and Rpl13a, respectively, in 22c10 cells. The transcript numbers and protein levels of both

genes from the same cell were obtained using single cell qPCR and quantitative imaging.

Page 79: Observation and quantification of protein production in single living cells

67

a

Page 80: Observation and quantification of protein production in single living cells

68

b

Figure 2.3 Insertion of PQR-XFP reporters into the endogenous genomic loci of IgK

and Rpl13a using CRISPRs.

(a) Workflow and timeline of a typical CRISPR targeting experiment to insert PQR-XFP

reporters into any endogenous locus. (b) The fluorescent reporters (approximately 800 base pairs

in size) are inserted in-frame immediately upstream of the stop codon, preserving the

endogenous 5’ and 3’ untranslated regions and the native coding sequence.

Page 81: Observation and quantification of protein production in single living cells

69

Figure 2.4 Validation of CRISPR-mediated insertion of PQR-GFP in the

endogenous IgK locus.

(a-b) Insertion of a PQR-GFP reporter into the endogenous IgK locus in 22c10 cells results in

green fluorescent hybridomas. (c) Successful genomic integration is verified with PCR primers

that lie within and outside the PQR insertion (Primer sets A and B for 5’ and 3’ end verification,

respectively).

a

b c

Page 82: Observation and quantification of protein production in single living cells

70

Figure 2.5 Illustration of the important steps and typical equipment used in the

protocol.

(a) An aspirator assembly is used to manipulate cells into individual 100uL drops of culture

medium on a Teflon coated, spotted glass slide. DNA and RNA-free freshly pulled micropipettes

are used for transferring cells. The glass micropipette tip may be broken slightly on the bottom of

the slide in order to create a tip diameter wide enough for unperturbed cell flow (inset). (b) PQR-

edited 22c10 express fluorescent reporters each time a protein of interest is translated. The cell

suspension is prepared at a low enough density to permit easy manual manipulation of single

cells. (c) Representative examples of single 22c10 cells near the broken micropipette tip. Note

the large diameter of the micropipette tip (2-3 times the size of one cell). Arrowheads point to

single cells. (d) The microscope objective is dipped into the drop of culture medium and a liquid

Page 83: Observation and quantification of protein production in single living cells

71

meniscus is formed around the diameter of the lens. (e) A single fluorescent 22c10 cell is

brought into focus and its fluorescence intensity is measured. (f) Following imaging, the cell is

immediately transferred using a new micropipette directly into Trizol solution. The micropipette

tip may be broken inside the tube to fully expel the contents of the pipette into the solution.

Page 84: Observation and quantification of protein production in single living cells

72

Sam

ple 1

Sam

ple 2

Sam

ple 3

26

28

30

32

34 1 l

3 l

6 l

cDNA input

Cycle

th

resh

old

Figure 2.6 Titration of starting input cDNA volume.

The volume of input cDNA is titrated to find the optimal volume that produces the most sensitive

amplification without exceeding 30% of the total reaction volume. Cycle threshold is inversely

proportional to starting template amount and thus in all three samples above, 6uL of starting

cDNA results in high amplification sensitivity without obvious inhibitory effects. n=3 cells.

Page 85: Observation and quantification of protein production in single living cells

73

Figure 2.7 Standard curve of serially diluted known amounts of Rpl13a target.

Absolute mRNA abundances are determined by using standard curves which are generated by

running a serial dilution of a known quantity of DNA that contains the qPCR assay target. The

logarithm of DNA quantity is plotted against cycle threshold (Ct) values and the slope and y-

intercept are used to convert Ct values into defined quantity values (number of

molecules/quantity). Ct values obtained from single cell experiments can then be compared to

this curve, in order to interpolate and determine the absolute abundance of mRNA from single

cells. Standard curves are prepared and quantified with each experiment. The slope of the

standard curve above is -3.7, producing an assay with 86% efficiency, and the R2 of the linear

regression was 0.99 (p<0.05).

0 2 4 6 8 100

10

20

30

40

50

Log DNA quantity (ng)

Ct

Page 86: Observation and quantification of protein production in single living cells

74

Amplification plot

Figure 2.8 Contamination of RNA sample quantification by genomic DNA can be

assessed using no-RT control reaction.

Reverse transcription reactions in which the reverse transcriptase enzyme has been omitted can

be used as template for qPCR to quantify the contamination of the RNA sample by carryover

genomic DNA. (a) Raw amplification plot from no-RT control reactions. No above-threshold

amplification can be observed. Using Trizol purification, genomic DNA is efficiently removed

and no contamination is observed when no-RT control reactions are amplified using 1, 3 or 6 µl

starting volume. (b) Quantification of data from (a), n=3 cells.

Sam

ple 1

Sam

ple 2

Sam

ple 3

-1

0

11 l

3 l

6 l

cDNA Input

Am

plificati

on

b

a

Page 87: Observation and quantification of protein production in single living cells

75

010

020

030

040

050

060

070

080

090

0

0

2

4

6

F(GFP) a.u.

Fre

qu

en

cy

700

900

1100

1300

1500

1700

1900

2100

2300

2500

2700

2900

3100

3300

0

2

4

6

F(RFP) a.u.

Fre

qu

en

cy

200

400

600

800

1000

1200

1400

1600

1800

0

1

2

3

4

5

RPL13a number of transcripts

Fre

qu

en

cy

Figure 2.9 Endogenous RNA and protein quantification from single cells.

(a) GFP fluorescence intensity (in arbitrary units a.u.) from single genome-edited 22c10 cells

shows a moderate distribution, n=26 cells. (b) RFP fluorescence intensities were higher and

clustered more, indicating consistently high level IgK production, n=24 cells. (c) Frequency

distributions of absolute Rpl13a mRNA numbers shows moderate expression of the Rpl13a gene,

n=17 cells. (d) Absolute IgK mRNA levels were relatively higher, indicating high expression of

the IgK gene compared to Rpl13a, n=24 cells.

a b

c d

5000

2500

0

4500

0

6500

0

8500

0

1050

00

1250

00

1450

00

1650

00

0

2

4

6

IgK number of transcripts

Fre

qu

en

cy

Page 88: Observation and quantification of protein production in single living cells

76

Figure 2.10 Protein and mRNA relationships between multiple genes in single cells.

(a) Rpl13a mRNA levels did not correlate with green fluorescence intensities indicating Rpl13a

mRNA is a weak predictor of Rpl13a protein production (R2=0.0, p>0.05, n=17). (b) IgK

mRNA levels had a moderate correlation with red fluorescence intensities (R2=0.44, p<0.05,

Page 89: Observation and quantification of protein production in single living cells

77

n=24). (c-d) IgK and Rpl13a exhibit nonsignificant correlations at both the protein level

(R2=0.14, p>0.05, n=24), and at the mRNA level (R2=0.06, p>0.05, n=17). (e) Examination of

mRNA and protein levels of multiple genes simultaneously from the same cell allows the co-

correlation of mRNA and protein levels. Black lines connect data obtained from the same cell

(n=17). (f) Human RPL13A (hRPL13A) mRNA levels were also a poor predictor of red

fluorescent intensities (R2=0.12, p>0.05, n=25), indicating a similar kind of relationship

between RPL13A mRNA and protein across the mouse and human genomes.

Page 90: Observation and quantification of protein production in single living cells

78

Forward Primer Reverse Primer RT Primer qPCR Probe

Mouse Rpl13a TCCCTCCACCC

TATGACAAG

GTCACTGCCT

GGTACTTCC

GCAGCCCTGCT

ACTCATTTTC

AGACTAAAA

TTCGTCGCTC

CGCTTCC

Human RPL13A TGTTTGACGGC

ATCCCAC

CTGTCACTGC

CTGGTACTTC

CTGCTGGCCAC

ATTTTATGTC

CTTCAGACGC

ACGACCTTG

AGGG

IgK AGTGGAAGATT

GATGGCAGTG

CTGTCTTTGCT

GTCCTGATCA

GGTGGATTTCA

GGGCAACTA

ACAAAATGG

CGTCCTGAAC

AGTTGG

Mouse Rpl13a

outside of homology

arms

CGGGTTGCTAA

CCTGGAATA

CAGTCTCCAT

CAAGGGGAAA

IgK outside of

homology arms

GGGGGAAAGG

CTGCTCATAA

TAACTGGGGG

AAGGGACACT

Table 2.1 Sequences of primers and probes used in this protocol.

Page 91: Observation and quantification of protein production in single living cells

79

Chapter III - A system for direct observation of

subcellular protein translation in single living cells.

3.1 Relation to overall project

The localization of mRNA and its local translation is an elegant regulatory mechanism to

restrict the expression of genes to subcellular compartments. While local mRNAs in distal

cellular sites can be visualized, the current tools to detect local protein synthesis events are

invasive and lacking in spatial and temporal resolution. To address this issue, I modified our

original PQR technique (Lo, Kays et al., 2015) to develop a fluorescence-based technique to

directly observe local protein synthesis events in single living cells. This chapter describes the

technique and validation experiments used to demonstrate its use as a quantitative marker of

local protein translation events.

Page 92: Observation and quantification of protein production in single living cells

80

3.2 Introduction

Understanding protein dynamics in vivo requires accurate and sensitive tools that can

detect spatiotemporal changes in protein translation at subcellular resolution. The local

translation of new proteins is a widely used cellular mechanism to restrict gene products to

specific regions of a cell or animal (Besse & Ephrussi, 2008; S. Kim et al., 2010; Lin & Holt,

2008; Martin & Ephrussi, 2010; Sharon A Swanger & Bassell, 2011; Dan Ohtan Wang et al.,

2010). mRNAs of locally translated genes are transported from the cell soma to discrete distal

cellular sites, and their localized translation only at those sites ensures local delivery of cargo,

cell fate specification and local signal responses. For example, the local synthesis of proteins

during development plays a crucial role in the establishment of cell fate (Cáceres & Nilson,

2005). In the nervous system, the dynamics of the local protein translation of the immediate early

gene Arc have been imaged in real time in dendrites of neurons expressing a luciferase-based

reporter. Local translation of Arc in neurons was induced by glutamate and could be detected

within 15 seconds in dendrites but not spines, supporting a model where stalled ribosomes at

dendrites are reactivated following glutamate stimulation to rapidly produce Arc protein (Na et

al., 2016). While the luciferase reporter that was used in the study provides high temporal

resolution, a drawback of this approach was the requirement for a substrate for the luciferase

enzyme to emit light. The concentration of this substrate, provided in the extracellular medium,

must be maintained throughout the experiment in order to produce accurate and consistent

results, and several instability issues have been reported (Craig et al., 1991; Morse & Tannous,

2012). Therefore, maintaining controlled levels of substrates and ensuring its constant access to

the enzyme are unpredictable which can present variability that biases results.

Page 93: Observation and quantification of protein production in single living cells

81

Genetically encoded fluorescent protein fusions have enabled the direct observation of

intracellular proteins, however protein fusions have been shown to affect protein properties and

localization (Palmer & Freeman, 2004) and folding and function of both the protein of interest

and the reporter (H. L. Zhao et al., 2008). In addition, the constitutive fluorescence of the

reporters may saturate measurement ranges over time, making the detection of small changes in

signal intensity difficult. Internal ribosomal entry sites (IRES) are sequences that initiate

translation of mRNA in a 5’ cap-independent process (Pelletier & Sonenberg, 1988). Bicistronic

expression of using IRES placed between a gene of interest and a fluorescent reporter can

address some of the issues encountered with fluorescent protein fusions. However, IRES

sequences can be 500 bp long and IRES-based translation is not stoichiometric and has been

shown to strongly favor the translation of the upstream gene (Houdebine & Attal, 1999;

Mizuguchi et al., 2000) making them unamenable for quantitative coexpression in vivo.

Our protein quantification technique uses the stoichiometric separation and production of

a fluorescent reporter molecules each time a molecule of protein of interest is translated (Lo,

Kays et al., 2015). Therefore, the fluorescence intensity or the brightness of the cell can be used

as a proxy measure for how much protein of interest is being translated. This technique is an in

vivo protein translational reporter technique that takes advantage of a viral genetic tag that

triggers a ribosomal skipping mechanism during translation, resulting in the stoichiometric

production of two proteins. One drawback of using standard PQR reporters for the detection of

the small and transient local protein synthesis events is the fact that standard fluorescent

reporters such as GFP, RFP and YFP require on the order of tens of minutes to properly fold,

mature and emit fluorescence signals (Snapp, 2005). Molecular diffusion forces can move

Page 94: Observation and quantification of protein production in single living cells

82

proteins great distances in the cytoplasm prior to the emission of fluorescence which precludes

any accurate spatial or temporal determination of protein synthesis (S. Kim et al., 2010)

In an effort to enhance PQR reporters to be able to detect and quantify local protein

synthesis events, I have investigated whether we could use the split GFP reconstitution system

for the detection of local translation at single cell resolution, in vivo. The split GFP system is

based on bimolecular fluorescence complementation of proteins, which takes advantage of the

fact that the GFP protein can be split into two nonfluorescent complementary parts that can

spontaneously and non-covalently reassemble to form the full fluorescent protein (Kerppola,

2006). Splitting the GFP protein at the 213th residue between the 10th and 11th barrels to make

GFP1-10 and a 15 amino acid GFP11 peptide results in split GFP partners that do not require any

cofactors or enzymes for reconstitution (Cabantous, Terwilliger, et al., 2005). Placing the

sequence of a GFP11 peptide downstream of our PQR reporter would result in the translation of

a GFP11 peptide for every molecule of protein of interest produced. In the presence of GFP1-10,

the splitGFP parts can recombine and emit green fluorescence, which marks the time and

location of the translation of the protein of interest (Figure 3.1). Key to this reconstitution is

timing, the complementation and fluorescence signal must develop before the GFP protein can

diffuse too far away from the site where GFP11 is translated. The signal must develop at the

moment and location of translation for this system to offer advantages over using standard

fluorescent protein fusion or coexpression.

Using the split GFP approach to monitor protein translation events offers a number of

advantages over probe or antibody-based approaches. For example, the reporters are genetically

encoded and the signal is fluorescence based which minimizes the invasiveness associated with

detecting protein translation. In addition, the fast reconstitution of the reporter immediately after

Page 95: Observation and quantification of protein production in single living cells

83

translation would provide exceptional spatial and temporal resolution that can be used to localize

protein translation evens in subcellular compartments such as the endoplasmic reticulum or

neuronal dendrites. Such an approach would open the door to the direct observation of local

protein synthesis in neurons, which is key to understanding the distal processes that both

maintain cellular homeostasis and mediate plasticity.

3.3 Materials and Methods

3.3.1 Protein Quantification Reporter constructs

For work in mammalian systems, the DNA sequence chosen for the PQR peptide was

GGAAGCGGAGCGACGAATTTTAGTCTACTGAAACAAGCGGGAGACGTGGAGGAAA

ACCCTGGACCT. For work in Drosophila, the DNA sequence for the PQR peptide was

GGAAGCGGAGAAGGTCGTGGTAGTCTACTAACGTGTGGTGACGTCGAGGAAAATCC

TGGACCT (Lo, Kays et al., 2015). Sequences were generated using gene synthesis (BioBasic)

and cloned into pCAG or a modified version of pCFD3 for mammalian or Drosophila work,

respectively. Plasmids containing mammalian and insect PQR reporters are now available from

Addgene. GFP, RFP, and BFP constructs were based on superfolderGFP, TagRFP-T, and

mTagBFP2, respectively. SuperfolderGFP and TagRFP-T were chosen for their relatively fast

maturation times, 6 min and 100 min, and photostability, respectively (Pédelacq et al., 2006;

Shaner et al., 2008). ShakerGFP cDNA (R. Blunck, Université de Montréal) was a kind gift, and

ShakerRFP was generated by swapping out the GFP fluorophore in the original construct for a

Tag-RFP-T fluorophore. All other plasmids were obtained through Addgene (Cambridge, USA).

3.3.2 Split GFP DNA constructs

Page 96: Observation and quantification of protein production in single living cells

84

Residues 1-213 of GFP, corresponding to the first 10 beta barrels of GFP (GFP1-10),

were amplified and cloned from evolved superfolderGFP, according to published split GFP data

(Cabantous, Terwilliger, et al., 2005; Kim et al., 2011). GFP11 was generated by amplifying and

cloning the last 15 residues of GFP into pCAG. To stoichiometrically co-express GFP1-10 or

GFP11 with other proteins of interest, a PQR peptide was added in-frame upstream or

downstream of the GFP1-10 or GFP 11 sequence depending on the desired orientation to

produce reporter constructs that stoichiometrically express either GFP1-10 or GFP11

downstream of a gene of interest. For extracellular membrane-bound expression of GFP1-10, the

complete Neuroligin-1 signal sequence, in addition to portions of the Neuroligin-1 extracellular,

transmembrane, and intracellular anchoring domains were fused to the N-terminus of GFP1-10

(J. Kim et al., 2011). For electrophysiology experiments, a PQR-GFP11 reporter was placed

downstream of the ShakerRFP coding sequence to generate ShakerRFP-PQR-GFP11, and

cloned into pCAG.

3.3.3 GFP1-10 protein production and extraction

DNA encoding GFP1-10 protein was transformed and expressed under the control of an

arabinose inducible promoter in Escherichia coli strain BL21(DE3) (New England BioLabs).

Cells were grown in LB medium to an initial O.D of 0.2, at which point induction of protein

production was initiated with 0.2% L-Arabinose and cells were further grown at 37°C for an

additional 16 hours with shaking at 225 rpm to encourage inclusion body formation. Cultures

were harvested using centrifugation and GFP1-10 was purified from inclusion bodies by

resuspension with TNG buffer [100 mM Tris-HCl (pH 7.4), 150 mM NaCl, 10% glycerol

vol/vol] containing 0.5 mg/ml lysozyme, 50 units of DNase I. The lysate was then incubated at

37°C for 25 min. Crude lysates containing GFP1-10-rich inclusion bodies were separated using

Page 97: Observation and quantification of protein production in single living cells

85

centrifugation at 16,000g at 4°C. Inclusion bodies were lysed using B-Per (Thermo-Fisher) and

sonication (as above) and GFP1-10 protein was collected and filtered using a 0.22 µm filter

before concentration with 10,000 molecular weight cutoff columns.

3.3.4 GFP 11 peptides

Variants of the GFP11 peptide were chemically synthesized with >75% purity

(Genscript). The amino acid sequences of the GFP11 peptides were: GFP11v1:

RDHMVLHEYVNAAGIT, GFP11v2: RDHMVLLEFVTAAGIT and GFP11v3:

RDHMVLHEFVTAAGIT (see Table 1 for full sequence list). Lyophilized peptides were

resuspended in water to > 10 mg/ml and frozen at -20°C. For extracellular GFP reconstitution in

HEK293 cells, GFP11 peptide was dissolved into the culture medium at a final concentration of

50 µM and cells were returned to a 37°C incubator for 2 hours before live imaging.

3.3.5 Cell culture

HEK293 cells were cultured at 37°C under 5% CO2 in Dulbecco's Modified Eagle

Medium, supplemented with 10% fetal bovine serum (Wisent), or for Drosophila melanogaster

S2 cells, at 25°C in Ex-Cell 420 Medium (Sigma-Aldrich). Media were supplemented with 100

units/mL penicillin (Thermo-Fisher) and 100 μg/mL streptomycin (Thermo-Fisher). Mammalian

cells were transfected with 3.5 µg of plasmid DNA in 35 mm dishes using Lipofectamine 3000

(Thermo-Fisher), or Transit-Insect (Mirus) for Drosophila cells. For extracellular GFP

fluorescence reconstitution, HEK293 cells were transfected with constructs expressing GFP1-10

tagged to the transmembrane and extracellular domains of the cell surface molecule Neuroligin-1

and incubated for 24-36 hours. Cells displaying GFP1-10 on the extracellular side of the cell

membrane were incubated in culture medium containing 50 µM GFP11 peptide for 3 hours at

Page 98: Observation and quantification of protein production in single living cells

86

37°C before live imaging. GFP1-10 protein was allowed to accumulate for 24 hours before

transfection of GFP11 constructs.

3.3.6 In vitro protein reconstitution

In vitro fluorescence complementation was performed by mixing purified GFP1-10

protein and chemically synthesized GFP11 peptides and the fluorescence intensity of the reaction

was collected with a StepOnePlus real-time thermal cycler (Thermo-Fisher). Briefly, 3 mM

GFP1-10 in TNG buffer or PBS (varied pH) was added to wells of a 96-well microplate coated

with 1 mM bovine serum albumin (BSA) and allowed to equilibrate for 60 seconds. GFP11

peptide was added according to different final peptide concentrations and the microplate was

immediately loaded into the fluorescence reader. The fluorescence intensity was measured every

10 seconds for 45 minutes at 32°C or 37°C with excitation at 495 nm and emission set at 520

nm. The fluorescence intensity was normalized to the initial fluorescence intensity to express

relative fluorescence increase upon fluorescence reconstitution. Standard curves for GFP

fluorescence measurements were generated by either using reconstituted GFP or GFP purified

from E.coli using GFP-specific chromatography columns (Bio-Rad). GFP protein concentration

was determined using the Bradford assay and absorbance readings at 280nm with a NanoDrop

2000 (Thermo-Fisher). Samples were serially diluted (1:10 or 1:5) and 10 µl samples were

imaged to reduce any non-linear fluorescence excitation effects.

3.3.7 Endoplasmic reticulum and ribosome staining

To visualize endoplasmic reticula (ER), HEK293 cells were transfected with split GFP

PQR reporter constructs and stained (live or fixed) with the ER and ribosome-specific stain

Cytopainter (Abcam). Stained cells were imaged in the green and red channels to examine the

Page 99: Observation and quantification of protein production in single living cells

87

co-localization of reconstituted GFP and red ER signals. Colocalization of green and red signals

was determined by calculating the Pearson’s and Mander’s correlation coefficients for

overlapping green and red pixel intensities. Individual z-planes were background subtracted and

thresholded to remove the lowest and highest pixel intensities. Ten ROIs comprising cells and

excluding background and nuclear regions were used for analysis. Both Pearson and Mander’s

colocolization coefficients were independently obtained and cross-validated using Coloc2

(ImageJ) and BioImageXD (Kankaanpää et al., 2012).

3.3.8 Electrophysiology

Standard whole cell voltage clamp was used to record potassium currents from HEK293

cells. Cells were maintained at 25°C in extracellular solution containing 140 mM NaCl, 10 mM

CaCl2, 7.5 mM KCl, 10 mM HEPES, and 10 mM glucose at pH 7.4, 319 mOsm during

recordings. Patch electrodes were pulled from standard wall borosilicate glass (BF150-86-10,

Sutter instruments) with 3–5 MΩ resistances. The intracellular pipette solution was 120 mM

KCl, 2 mM MgCl2, 1 mM CaCl2, 2 mM EGTA, 20 mM HEPES, and 20 mM sucrose at pH 7.23,

326 mOsm. Whole cell currents were low pass filtered at 10 kHz and measured using an

Axopatch 200B amplifier (Axon instruments), and recorded using a DigiData 1200 with

pClamp9 software (Molecular Devices). Cells were held at -80 mV and then given +20 mV steps

of 45 ms. To accurately compare I-V curves and current data across cells and experiments, the

steady-state current was divided by the membrane capacitance (mean Cm=15 pF, n=3), and

current density (pA/pF) was used for comparisons. Consistent cell capacitance, and membrane

and access resistances were verified before and after recordings.

3.3.9 Image acquisition and analysis

Page 100: Observation and quantification of protein production in single living cells

88

Fluorescence and brightfield microscopy was performed using a Zeiss AxioScope A1. All

images were acquired at 1388 x 1040 pixels using a 40× water objective, N.A. 1.0

(epifluorescence). Fluorescence emission was detected using a charge-coupled device (CCD)

camera (MRm). 488 nm blue light was used to excite GFP, and 515 nm emission light was

collected. Similarly, 543 nm light was used to excite TagRFP-T and 594nm emission light was

collected. All image acquisition parameters were fixed for each imaging channel for exposure

time, excitation intensity and gain. Cells that were dimmer or brighter than the fixed initial

acquisition dynamic range were not included for analysis. Time-series images were collected

using an open-shutter video configuration in ZenLite (Zeiss). Images were acquired every 167

milliseconds with exposure times of 260 milliseconds.

Images were selected for analysis based on identification of healthy cells and low background.

Fluorescence pixel intensities were measured in several random regions of interest (ROIs) within

the target cellular region using a custom written program in MatLab (MathWorks) or ImageJ.

Average pixel intensities were calculated from five ROIs of 7x7 pixels for measurements within

the cytoplasm, perinucleus, and nucleus, and 3x3 pixels for measurements within the plasma

membrane. All signal intensities were background subtracted from the average of three ROIs

immediately surrounding the cell. For time-series image analysis, background was considered as

the region immediately adjacent (<15 µm) to the perinucleus, or cytoplasm. Kinetic increases in

fluorescence from timelapse or video data were plotted was Ft/F0.

3.3.10 Statistical analysis

Pearson’s linear correlations were calculated by fitting the data to a simple linear

regression model, with the coefficient of determination, R2. Kinetic reconstitution traces were fit

with a simple one-site binding model with the coefficient of determination R2 using Prism

Page 101: Observation and quantification of protein production in single living cells

89

(GraphPad). We tested the null hypothesis that the variables were independent of each other and

that the true R2 value was 0 for both linear and nonlinear models.

3.4 Results

3.4.1 GFP11 and GFP1-10 reconstitute spontaneously in vitro

To initially characterize the reconstitution of GFP1-10 and GFP11, we performed in vitro

fluorescence reconstitution with diluted GFP11 peptides. Diluted GFP11 peptides prepared over

a 100-fold range of molar ratios covering 0.1 pmol to 50 nmol were added to 3 mM of a crude

bacterial inclusion body lysate containing overexpressed GFP1-10 (Cabantous & Waldo, 2006;

Huang & Bystroff, 2009). The resulting fluorescence intensity of the reaction was collected over

20 minutes at 37°C (Figure 3.2a). The relative fluorescence intensity of the reaction was

normalized to the initial fluorescence intensity to express relative fluorescence increase upon

complementation. We found that the fluorescence intensity of the reaction was directly

dependent on the input amount of the split GFP components (Figure 3.2b), and the optimal ratio

of GFP1-10 to GFP11 that produced the most efficient reconstitution was 2:1 (Figures 3.2a-b).

The relative fluorescence increase was plotted against GFP11 peptide concentrations and fit to a

simple one-site binding model to determine the dissociation constant (kd) of the GFP11 peptide.

Consistent with previous results, we found that kd was = 0.48 ± 0.12 nmol (R2=0.95, p<0.05)

(Figure 3.2b) (Do & Boxer, 2011; Huang & Bystroff, 2009).Therefore, GFP1-10 must exist in

molar excess of GFP11 for reconstitution to be efficient.

To investigate the dependence of GFP reconstitution on temperature and pH, we

performed in vitro complementation under different temperature and pH conditions. GFP11

peptides (2 µmol) were mixed with an excess of GFP1-10 (1 mg/ml) in a 100 µL reaction

volume and the reconstitution was incubated at 32°C or 37°C in TNG or PBS buffered to pH 4,

Page 102: Observation and quantification of protein production in single living cells

90

pH 7 or pH 8.5 for 55 minutes. The resulting green fluorescence intensity was monitored every 5

minutes for 1 hour (Figures 3.2 c-d). Consistent with previous results, the complementation of

GFP was found to be severely reduced in acidic conditions (pH< 7) and lower temperatures and

is more efficient at pH >7 and 37°C (Patterson et al., 1997). These results demonstrate that the

reconstitution of GFP occurs efficiently at physiological conditions. Next, we sought to examine

the kinetics with which GFP reconstitution occurs in vitro.

3.4.2 GFP reconstitution in vitro occurs at millisecond timescales

For split GFP to be useful as a fast and reliable marker of local protein translation events,

the GFP parts must reconstitute immediately after translation. To characterize the kinetics of in

vitro complementation between GFP1-10 and GFP11 at a higher temporal resolution, we mixed

freshly purified GFP1-10 and GFP11 protein at 2:1 ratio and incubated the mixture in a

fluorescence reader at 37°C. The fluorescence intensity of the reaction was measured every 10

seconds for 20 minutes. The fluorescence intensity of the reaction rose logarithmically at a sharp

rate in the first 200 seconds and began to saturate after 5 minutes (Figure 3.3a), compared to the

GFP1-10 alone control. The relative fluorescence increase against time was fit to a simple one-

phase association model to determine the rate constant (k) of GFP11 peptide binding, resulting in

a value of ~ 7 ± 0.0001 x 10-3 s-1 and a half time of 95 seconds (R2=0.99, p<0.001) (Figure 3.3a),

consistent with previous results (Huang & Bystroff, 2009).

It is interesting to note that the initial fluorescence intensity of the most efficient

complementation reaction at t=0 is already almost four times that of the GFP1-10 alone control

(3.8 vs 0.7 x105 a.u.) (Figure 3.3b), indicating that fully formed and fluorescent GFP is present in

the reaction tube by the time it is loaded into the machine and before measurements are begun. It

takes roughly 30-45 seconds to pipet the GFP11 peptide into the reaction plate and load it into

Page 103: Observation and quantification of protein production in single living cells

91

the machine and in that time, at room temperature the complementation reaction is already taking

place and GFP fluorescence can be observed. Therefore, to determine the speed of the

complementation reaction with the available equipment, we sought to determine how many

molecules of GFP are being reconstituted per unit time. Using a serial dilution of known

quantities of GFP protein, we measured the fluorescence intensity to generate a standard curve of

fluorescence intensities versus amount of protein, covering a six-fold range of fluorescence

levels and GFP protein amounts (Figure 3.3c). Fitting the standard curve data to a linear fit

model results in an equation which essentially assigns a fluorescence value for any given

quantity of GFP protein (R2=0.99, p<0.0001, n=6). Given that 1 ng of GFP contains 2x1010

molecules, the rate of fluorescence increase per unit time obtained from the reaction

(fluorescence increase (a.u) / millisecond = 8.616) can thus be converted to molecules per unit

time, to obtain ~ 71x106 reconstituted molecules per millisecond. This result demonstrates the

fast reconstitution of split GFP and suggests that following protein translation, newly produced

GFP11 peptides can reconstitute with GFP1-10 on the order of milliseconds.

3.4.3 GFP11 detects GFP1-10 in living cells

To determine whether GFP11 can detect and recombine with GFP1-10 in living cells, we

expressed GFP1-10 on the extracellular surface of HEK293 cells. Portions of the synaptic protein

Neuroligin-1 were used to target GFP1-10 and anchor it on the extracellular surface of the cell

membrane (J. Kim et al., 2011). Cells expressing extracellular GFP1-10 were incubated with 50

µM synthetic GFP11 peptide for 2 hours and then imaged. Fluorescence live imaging shows

clear and distinct green fluorescent HEK293 cells after addition of GFP11 peptide (Figure 3.4b).

GFP was specifically detected at the membranes of HEK293 cell (Figure 3.4b, arrows). The

emission of fluorescence was solely triggered by the addition of GFP11 peptides, and no

Page 104: Observation and quantification of protein production in single living cells

92

fluorescence was observed from control untransfected cells, or when no GFP11 peptide was

added (Figure 3.4a).

We tested the intracellular complementation of split GFP by transiently transfecting

HEK293 cells with separate plasmids encoding cytoplasmic GFP1-10 and GFP11. We observed

green fluorescence as early as 7 hours after transfection and cells were homogenously filled with

fluorescent GFP 24-48 post transfection (Figure 3.4c). To test the effect of placing split GFP

reporters upstream or downstream of PQR constructs, we expressed GFP1-10 and GFP11

reporters separated by PQRs in different orders. Fluorescence imaging of live cells shows

homogenous green fluorescence throughout the cytoplasm, irrespective of whether the split GFP

reporter was placed before or after the PQR (Figure 3.4d). These results indicate that split GFP

reporters can be coexpressed with any protein of interest using PQR, and this does not affect the

reconstitution of fluorescence.

3.4.4 GFP reconstitution can report sites of protein translation

To better understand the sources of green fluorescence signals within the cell, we

expressed RFPnols-PQR-GFP11 in HEK293 cells (Figures 3.5b-c). The nucleolar localization of

RFP is a useful marker to highlight the size and boundary of the nucleus, enabling easy

identification of perinuclear regions where most protein translation occurs (M. S. Scott et al.,

2010). In the absence of GFP1-10, cells were nonfluorescent in the green channel but showed

clear red fluorescent nucleoli (not shown). In the presence of GFP1-10, the green fluorescence

signal was concentrated and clustered at perinuclear regions, often forming spots and halos

immediately adjacent to the nuclear boundary (Figures 3.5b-f). Similar patterns were observed

when a DNA construct encoding ShakerRFP-PQR-GFP11 was co-transfected with GFP1-10

(Figure 3.6). ShakerRFP is a membrane-bound potassium channel that is translated and

Page 105: Observation and quantification of protein production in single living cells

93

processed in the ER and Golgi apparatus before being transported to the plasma membrane.

Transfection of either the nucleolar RFP or membrane-bound ShakerRFP constructs showed the

correct localization of the upstream protein, indicating the co-expression of GFP11 reporter does

not perturb the translation or localization of upstream proteins (Figures 3.5 and 3.6).

To confirm that the local concentration of green fluorescence might correspond to sites of

active protein translation, we stained cells expressing GFP1-10 and GFP11 PQR constructs with

a red stain specific for endoplasmic reticula and ribosomes. High-resolution imaging of live and

fixed cells revealed high spatial overlap of the green and red fluorescence signals (Figures 3.5d-

f), indicating GFP fluorescence overlaps with endoplasmic reticulum and ribosomal signals

(Figures 3.5d-g). Mander and Pearson’s correlation coefficients were calculated for individual z-

slices to validate our results. Green and red pixel intensities were confirmed to colocalize;

specifically 71% of above-threshold GFP signals colocalized with 82% of above-threshold red

ER and ribosome stain signals (thresholded Mander’s (tM) coefficients tM1=0.71, tM2=0.82,

Pearson’s R2=0.84, Costes p-value=1 (i=100), n=10 ROIs) (Figure 3.5g). These results suggest

that the reconstitution of GFP is taking place at sites of active protein translation.

We further examined the dynamics of the fluorescent signal originating at perinuclear

sites, compared to regions within the cytoplasm to determine whether we could detect new GFP

synthesis. HEK293 cells transfected with split GFP components were imaged and the

fluorescence intensity within several regions of interest was quantified over time. We chose to

begin our imaging sessions 18 hours post transfections to coincide with high rates of translation

of transfected plasmids. Using quantitative imaging we found that the intensity of fluorescent

signals originating from perinuclear sites rose while cytoplasmic fluorescent signals did not over

the span of ~ 2 minutes (Figures 3.5g-h). Fluorescent signals from both perinuclear and

Page 106: Observation and quantification of protein production in single living cells

94

cytoplasmic regions showed an initial small decrease in fluorescence intensity, which we

attributed to photobleaching. The subsequent fluorescence increase at perinuclear sites was

consistently larger compared to cytoplasmic regions, even in cells that displayed more initial

photobleaching at perinuclear regions, indicating new GFP protein synthesis and reconstitution

(n=6) (Figure 3.5i). These results demonstrate that splitGFP can be used to observe the localized

synthesis of proteins and confirm that regions stained with the endoplasmic reticulum and

ribosome marker correspond to sites of new protein production.

3.4.5 Proteins co-translated with GFP11 reporters function properly

To verify that proteins produced with a PQR and GFP11 reporter can function properly,

we expressed the ShakerRFP ion channel separated from a GFP11 reporter by a PQR peptide

(ShakerRFP-PQR-GFP11) in HEK293 cells (Figure 3.6a). ShakerRFP is a large tetrameric

potassium channel that has a RFP molecule embedded in its N-terminal domain. The RFP

fluorophore acts as a tethered ball and chain which prevents the channel from inactivating,

allowing it to pass current as long as depolarization is maintained. As with most ion channels,

upon translation they are processed in the secretory pathway and packaged in vesicles to be sent

to the membrane, where they are inserted for periods of hours to days (Narahashi, 1988).

Fluorescence imaging of live HEK293 cells expressing ShakerRFP-PQR-GFP11 alone showed

distinct red fluorescent cell membranes, indicating the membranous distribution of the channel

(Figure 3.6a). Using whole-cell patch clamp, +20 mV voltage steps were applied and the steady

state current passed by the channel at each voltage step was used to generate an current-voltage

(I-V) relationship curve (Figure 3.6b). I-V curves are useful metrics that relate the current passed

across a membrane, when a voltage is applied across it. Since the current is determined by the

conductances present in the membrane, I-V curves serve as a diagnostic for proper channel

Page 107: Observation and quantification of protein production in single living cells

95

expression and function within a membrane (Tester, 1997). To normalize for variability in

transfection and uptake of plasmids by HEK293, current is divided by cell capacitance to give

current density (pA/pF), which is more appropriate when comparing I-V curves (Figure 3.6b). In

HEK293 cells in which ShakerRFP has been transiently overexpressed at very high levels, the

main conductance in the membrane is the Shaker potassium channel. Endogenous voltage-gated

calcium and potassium channels have been reported in the non-neuronal HEK293 cell line, but

the peak current is less than 300pA (He & Soderlund, 2010; Yu & Kerchner, 1998). Thus, the

contribution of endogenous HEK293 currents to the nanoampere range of currents observed in

cells with overexpressed channels can be considered negligible. The addition of a PQR-GFP11

reporter downstream of the coding sequence of ShakerRFP did not appear to affect the

translation or function of ShakerRFP, as determined by live imaging and electrophysiology.

HEK293 cells transiently transfected with ShakerRFP-PQR-GFP11 DNA showed the same

membrane pattern of red fluorescence as cells expressing ShakerRFP alone (Figure 3.6a). I-V

curves generated from cells transfected with ShakerRFP-PQR-GFP11 showed no significant

difference from those transfected with ShakerRFP alone, in terms of slope or reversal potential

(p<0.05 for both). In addition, the measured reversal potential (~ -73 mV) (Figure 3.6b),

corresponded to the potassium ion reversal potential predicted using the Nernst equation to

within 10% error. These results indicate that the use of GFP11 as a reporter of protein does not

perturb protein localization, function, or the general health of the cell in which they are

expressed, which suggests that it is amenable for expressing upstream proteins in vivo.

3.4.6 GFP reconstitution can quantitatively readout protein translation

Protein synthesis of a PQR-GFP11 reporter requires approximately 2 seconds (6 amino

acids/second) (Kramer et al., 2009; Ross & Orlowski, 1982). In the presence of GFP1-10, the

Page 108: Observation and quantification of protein production in single living cells

96

split proteins recombine in milliseconds to emit a green fluorescence signal to indicate the time

and location of protein translation within the cell. To determine whether the green signal

produced by GFP reconstitution can quantitatively readout the levels of translation of the

upstream protein, we took advantage of the RFP fluorophore that is embedded in the ShakerRFP

molecule. We co-expressed ShakerRFP-PQR-GFP11 and GFP1-10 in HEK293 cells and

verified that the fluorescent signals observed in the green and red channels are linearly

proportional in intensity (Figures 3.6a,c). Quantitative fluorescence imaging revealed a linear

correlation of the green and red fluorescence intensities (R2=0.71, p<0.001, n=35), indicating

the translation and reconstitution of the GFP11 can quantitatively readout the translation of an

upstream protein (Figure 3.6c). Therefore, the fluorescence intensity of the signal is directly

proportional to the level of translation of GFP11, which in turn is proportional to the translation

of the upstream protein of interest (Figure 3.1).

3.5 Discussion and conclusions

The reconstitution of GFP occurs on the order of milliseconds and split GFP partners can

be stoichiometrically co-expressed with proteins of interest using PQR. This is exploited as a fast

marker to read out protein translation events by using the GFP signal emitted after reconstitution

to mark the time, location and level of protein translation within a living cell. The rationale

behind this approach is to take advantage of the fast translation and reconstitution of the GFP11

peptide. The 15-amino acid GFP11 peptide requires just over 3 seconds to be translated. This is

crucial as once it is produced, it requires no further processing or folding and is free to

immediately recombine with GFP1-10 that has been previously expressed, allowed to fold and is

primed for reconstitution. The concentration of GFP1-10 is rate limiting factor in this approach,

Page 109: Observation and quantification of protein production in single living cells

97

and so GFP1-10 must be present in high concentrations in the cell as a critical prerequisite for

the accuracy and reliability of fast protein production quantification.

The GFP11 peptide requires ~ 3 seconds between the initiation of GFP11 translation and

the reconstitution and emission of fluorescence, indicating the moment of protein translation

within milliseconds for GFP11 and with a 3 second delay from the moment of translation of the

upstream protein. Whether or not the upstream protein diffuses away from the site of translation,

translation of GFP11 will mark the original site of mRNA translation, unless the RNA-bound

ribosome diffuses away. Ribosome diffusion in cytoplasm is 0.04 µm2/sec (Bakshi et al., 2012),

and thus the 3 second delays in detecting the initial protein translation even will produce a spatial

error of ~ 350 nm. Our results suggest that the kinetics and efficiency of the reconstitution

reaction can be improved, as different variants of the GFP11 peptide produced different rates of

reconstitution (Figure 3.2c). Minor differences in the solubility, charge and size of the GFP11

peptide can affect the rate and efficiency of reconstitution, and ultimately the properties of the

reconstituted protein. Therefore, screening for GFP11 peptides that result in the most sensitive

reconstitution will certainly improve this technique in the context of monitoring local protein

translation events. Split GFP components that fail to reconstitute, or reconstitute but fail to

fluoresce can affect the spatial, temporal and quantitative accuracy of the GFP11 reporter. While

we did not observe such problems in our experiments, their occurrence is impossible to predict or

estimate in vivo. However, the finding that different variants of GFP11 reconstitute differently

suggests that undiscovered GFP11 peptide variants may possibly outperform currently available

ones. Nonetheless, I am constantly screening for new GFP11 variants that reconstitute faster and

more efficiently, and this is discussed in the next chapter.

3.5.1 GFP1-10 fluorophore maturation

Page 110: Observation and quantification of protein production in single living cells

98

It is not known whether the fluorophore is GFP1-10 can mature when GFP1-10 is

produced separately from GFP11 (Cabantous, Pédelacq, et al., 2005; Chudakov et al., 2010;

Huang & Bystroff, 2009; Kaddoum et al., 2010; Kerppola, 2006). Unfolding and refolding

experiments have shown that unfolded GFP that was once native in structure preserved the

maturation of the fluorophore, which explains the fast recovery of fluorescence upon refolding of

the protein. However, in this case the incomplete structure of GFP (GFP1-10) is produced

directly from the gene, and it is not clear whether this precludes the maturation of the

fluorophore. The folding and maturation of the fluorophore occurs at residues 65, 66 and 67

which are over 150 residues away from the site where GFP11 is split, suggesting at least that

splitting the protein does not physically interfere with these residues. However, if the

fluorophore could mature without GFP11, then it would yet GFP1-10 is nonfluorescent, and

remains so until it interacts with GFP11. Residue Glu222, located in the C-terminal domain of

GFP and part of the GFP11 fragment is known to be required for the switching of the

fluorophore between protonated forms, which affects its stability (Kent et al., 2008).

Furthermore, Ser205, Thr203 and His148 are required for fluorophore stabilization (Heim et al.,

1995), but the question of whether a non-stabilized form of a mature fluorophore can form has

not been answered. In addition, the presence of water molecules inside the barrel structure leads

to quenching of fluorescence (Ormö et al., 1996). This raises the idea that an uncomplimented

GFP1-10 could have a mature yet unstable and quenched fluorophore, which quickly switches

conformations and emits fluorescence upon reconstitution with GFP11. This model is supported

by my data, as fluorescence emission began within seconds after introducing GFP11 in vitro.

Given that the maturation of the GFP fluorophore is known to require on average tens of minutes

(Iizuka et al., 2011; Shaner et al., 2008), this suggests that the GFP1-10 fluorophore is in some

Page 111: Observation and quantification of protein production in single living cells

99

transition state between folding and fluorophore maturation. Therefore, reconstitution with

GFP11 quickly allows the new molecular conditions to favor the full maturation and emission of

fluorescence from the chromophore.

Finally, the goal in developing this technique is to observe local protein synthesis in vivo,

particularly in neuronal cells. In the next chapter I describe the experimental design of the

strategy to apply the technique in the context of a living animal. The reagents that are currently

being developed are described and some preliminary results are presented.

Page 112: Observation and quantification of protein production in single living cells

100

3.6 Figures

Figure 3.1 Stoichiometric production of GFP11 reporters using PQR.

Insertion of a GFP11 reporter downstream of a PQR tag results in slightly longer mRNAs being

transcribed (extra 100 bases). During translation, a ribosomal skipping mechanism results in the

stoichiometric production of a GFP11 molecule each time a molecule of protein of interest is

translated. In the presence of GFP1-10, the splitGFP parts can recombine and emit green

fluorescence, which marks the time and location of the translation of the protein of interest.

Page 113: Observation and quantification of protein production in single living cells

101

Figure 3.2 In vitro characterization of the split GFP reconstitution reaction.

(a) Fluorescence reconstitution kinetic traces for amounts of GFP11 that were mixed with an

excess of GFP1-10 in vitro, and incubated at 37°C. (b) Sensitivity of split GFP reconstitution in

vitro. The fluorescence intensity of the reconstitution reaction is linearly dependent on input

GFP11 concentration. (c) Fluorescence reconstitution of split GFP occurs more efficiently at

37°C and different variants of GFP11 result in different reconstitution rates and efficiencies.

20pmol of GFP11 peptide were mixed with an excess of crude GFP1-10 protein and the resulting

fluorescence intensity was collected at regular intervals for 55 minutes. (d) In vitro reconstitution

of GFP was tested under different buffer and pH conditions at 37°C and was most efficient in

TNG buffer at pH 8.5 (n=4).

Page 114: Observation and quantification of protein production in single living cells

102

Figure 3.3 Reconstitution of split GFP occurs on the order of milliseconds in vitro.

(a) Kinetic trace of split GFP reconstitution. GFP1-10 and GFP11 were mixed at a 2:1 molar

ratio in TNG buffer and the fluorescence intensity was recorded over time. The fluorescence

intensity of the reaction rose logarithmically and began to saturate after 5 minutes. The

fluorescence increase against time was fit to a one-site model and the rate constant k was

determined to be 0.007 ± 0.0001 s-1, with a halftime of 95 seconds (R2=0.99, p<0.001). (b) The

first 50 seconds from (a) are shown to demonstrate that a portion of splitGFP was already

reconstituted before the earliest timepoint could be acquired. (c) Standard curve of serially

diluted GFP showing the linear dependence of fluorescence intensity on GFP amount (R2=0.997,

p<0.05).

Page 115: Observation and quantification of protein production in single living cells

103

Figure 3.4 Split GFP reporters can be expressed using PQRs and the reconstitution

of GFP marks the presence GFP1-10 protein.

(a-b) Extracellular membrane-tagged GFP1-10 was expressed in HEK293 cells. Cells display a

green membranous fluorescent signal upon addition of GFP11 peptide into the culture medium.

(c) Split GFP components can be expressed cytosolically and recombine upon translation to

produce fluorescent GFP. (d) SplitGFP reporters efficiently recombine to produce GFP when

Page 116: Observation and quantification of protein production in single living cells

104

expressed using PQRs placed upstream or downstream. Scale bars are 30 µm in (b) and 20 µm in

(c) and (d).

Page 117: Observation and quantification of protein production in single living cells

105

Page 118: Observation and quantification of protein production in single living cells

106

Figure 3.5 Split GFP reconstitution occurs at sites of active protein translation.

(a-c) SplitGFP reconstitution occurs at perinuclear sites. RFPnols was included in (b-c) to

highlight the nuclear boundary. (d-f) Cells expressing splitGFP constructs were fixed and stained

with a red marker for endoplasmic reticula and ribosomes (Cytopainter). High overlap between

green and red fluorescent signals was observed, particularly at perinuclear regions. (g)

Quantification of colocalization of green and red pixel intensities from ten ROIs confirmed that

green and red signals colocalized above chance (p<0.05). (h) Time-series normalized

fluorescence intensity analysis at perinuclear and cytoplasmic regions of interest in a cell

expressing splitGFP constructs. Perinuclear regions showed increasing levels of fluorescence

over time, compared to cytoplasmic levels. (i) Traces of fluorescence intensity over time from

perinuclear (top panel, n=6) or cytoplasmic (bottom panel, n=5) regions of interest from several

cells. Scale bars are 50 µm in (a-c) and 10 µm in (d-f).

Page 119: Observation and quantification of protein production in single living cells

107

Figure 3.6 Co-translation of GFP11 reporters using PQR preserves the protein of

interest’s localization and function.

Page 120: Observation and quantification of protein production in single living cells

108

(a) Co-expression of a GFP11 reporter does not affect the membrane localization of ShakerRFP

in HEK293 cells. (b) Representative I-V curve and voltage step protocol used to characterize the

potassium conductance in cells transfected with ShakerRFP-PQR-GFP11. (c) In the presence of

GFP1-10, GFP11 reporters co-produced with ShakerRFP reconstitute and form fluorescent GFP

(image). The green and red fluorescence intensities with linearly correlated, indicating the level

of GFP11 production, and thus reconstitution, is proportional to the level of production of

ShakerRFP (Pearson’s R2=0.72, p<0.05, n=35). Scale bars are 20 µm in (a) and 10 µm in (c).

Page 121: Observation and quantification of protein production in single living cells

109

Chapter IV - Applications and future directions of

protein quantification using PQR

4.1 Relevance to overall project

My goal in this part of the thesis is to develop proof of principle experiments to

demonstrate ways to monitor and quantify the translation of proteins in the living animal. The

specific questions to be addressed with these experiments are: 1) how does endogenous protein

production dynamically change in single cells in vivo? and 2) can I use the reconstitution of GFP

as a reliable and quantitative detector of subcellular local protein translation in single cells in

vivo? In this chapter I describe strategies and reagents that are being developed to visualize

protein translation in vivo, at both cellular and subcellular resolution. This chapter presents DNA

constructs, mice and flies currently being generated to demonstrate the applicability of the

techniques in vivo and provide useful resources for the research community.

Page 122: Observation and quantification of protein production in single living cells

110

4.2 Applications of optical protein quantification using PQR

4.2.1 Dynamic observation of protein synthesis in vivo

Antibodies are produced and secreted by the immune system in response to pathogen and

toxin exposure as part of the body’s immune response. In vertebrate immune systems, two

identical heavy chains and two light chains constitute the basic structural unit of an

immunoglobulin antibody molecule. There exist five classes of antibodies, each with its own

class of heavy chain and either class of light chains: kappa (κ) and lambda (λ). The detection and

measurement of antibody production has been crucial for the diagnosis and tracking of immune

responses as well as the progression or regression of immune-mediated diseases. Testing

therapeutic agents, particularly those of protein origin, can elicit harmful antibody responses in

the host against the agent, which are disadvantageous (De Groot & Scott, 2007). Therefore, the

immunogenicity of therapeutic proteins and agents remains a concern and monitoring levels of

antibody production is an important indicator in treatment of disease.

Currently, the detection of proteins is mostly performed using antibody-based approaches

such as the ELISA assay or the immunoblot (western blot). To develop a system to optically

monitor dynamic endogenous protein translation in single cells in vivo (question 1), we will

create animals using genome editing that express fluorescent PQR reporters as real time

indicators of protein translation. The first animal that we are generating carries a PQR-GFP

insertion at the endogenous IgK locus. Its main purpose is to allow for in vivo examination of

IgK antibody production in real time, at single cell resolution. B cells throughout the mouse body

will co-produce a molecule of GFP for every molecule of IgK light chain (Figure 4.1). B cells

from different immune regions in the body such as the spleen, lymph nodes, and bone marrow

differ in their reactivity to different stresses and thus they also differ in the amount of antibody

Page 123: Observation and quantification of protein production in single living cells

111

they produce (Leandro et al., 2013; Nutt et al., 2015). This mouse will serve as a platform to 1)

optically localize and quantify local immune responses in real time in vivo, 2) easily screen

agents or therapies that increase or decrease the levels of antibody production by simply

monitoring changes in fluorescence intensity and 3) allow the fast and easy identification,

extraction and enrichment of B cells that express varying levels of kappa isotype antibodies from

anywhere in the mouse body.

In order to generate mice carrying PQR reporters at the endogenous IgK locus, validated

CRISPR sgRNAs, Cas9 protein and PQR repair templates from Chapter 2 were injected into the

pronuclei of mouse embryos. We used a complex of Cas9 protein and synthetic RNA CRISPR

instead of using DNA or mRNA encoding the Cas9 and sgRNA, as this has been shown to

significantly decrease DNA toxicity, low expression levels, and issues of nucleic degradation by

nucleases (Liang et al., 2015).

Since immunoglobulin loci can undergo several forms of genomic rearrangements that

render the architecture of the genomic locus unpredictable, inserting the PQR reporters into the

end of the constant region of the light chain locus will ensure that the variety of kappa-type

antibodies, which all share a common constant region, are stoichiometrically co-produced with a

fluorescent reporter (Figure 4.1b). In addition, this approach preserves the endogenous IgK 5’

and 3’UTR, which may contain critical regulatory elements that influence the transcription,

splicing, translation and localization of IgK mRNA.

The resulting founder mice are bred and checked for the integration of the PQR reporter

into the germline by extracting genomic DNA from tail or ear clips followed by genotyping

using primers that lie within and outside the recombined fragment and by sequencing and

Page 124: Observation and quantification of protein production in single living cells

112

digestion-verification of the product. Positive founders are then raised and backcrossed to

determine germline transmission and eliminate any potential off-target CRISPR effects.

4.2.2 Optical normalization of protein production in vivo

The expression of nearly half of genes in the mouse genome, including circadian genes

has been shown to occur in an oscillatory manner in many types of tissues throughout the body

(Zhang et al., 2014). In contrast, stochastic “rush hours” of gene expression are frequently being

reported in gene expression atlases (Lein et al., 2007; J. Z. Li et al., 2013; Panda et al., 2002;

Zhang et al., 2014). Thus, expression of a gene can either be stochastic occurring in bursts or

rhythmic, exhibiting patterns.

In any quantitative measurement of change, particularly in vivo, it is crucial to ascertain

whether the observed changes are due to a specific effect, or a non-specific. This also applies to

the quantification of mRNA and protein levels. To address this issue, reference, or

“housekeeping”, gene expression is often used in quantitative mRNA and protein experiments to

normalize for observed changes across conditions and experiments. For example, PQRs inserted

at endogenous housekeeping loci such as RPL13A allow for the fluorescence from those loci to

be used as a normalization signal for measurements taken across cells and experiments. In vivo,

the complex crowding and depth of tissues introduce optical aberrations such as scattering of

excitation and emission light as well as an exponential drop in light penetration power with

increased tissue depth. Insertion of PQR reporters into the endogenous RPL13A locus in a mouse

will result in the production of a fluorescence-based normalization signal in all cells of all tissues

in that mouse. This enables the normalization of any protein measurements imaged in a second

channel (i.e., a second gene of interest), to changes in the levels of RPL13A in that same cell or

tissue. This permits the comparison of normalized data within and between experiments. In

Page 125: Observation and quantification of protein production in single living cells

113

addition, this mouse will allow for in vivo dynamic monitoring of Rpl13a production levels at

single cell resolution, which is particularly informative when examining global cellular states in

normal and disease conditions (Poddar et al., 2013). For example, quantifying the Rpl13a

fluorescence in a second channel can be used as a measure of an individual cell’s transcriptional

and translational levels (Chapter 2). Thus, any change in IgK protein production can be

normalized to the cell’s global protein production status, reflecting the true net increase in IgK

production.

The same approach and techniques described in previous chapters were used to insert a

PQR-RFPnols construct into the endogenous Rpl13a locus in mouse embryos. First, a double-

strand break (DSB) at the end of the Rpl13a coding sequence was induced by a guided Cas9

nuclease. PQR-RFPnols repair templates with arms homologous to Rpl13a were used as a

template for homologous recombination resulting in an edited locus of the form: Rpl13a-PQR-

RFPnols (Figure 4.2).

Preliminary results from the injection of a PQR-RFP reporter and Rpl13a-specific

CRISPR/Cas9 into mouse pronuclei showed that the validated CRISPRs from Chapter 2 can

efficiently and repeatedly produce red-fluorescent 2-cell, 4-cell and blastocyst-stage embryos

(Figure 4.2). The implantation of injected embryos into carrier mothers is ongoing and correct

genomic editing of the Rpl13a locus is verified by PCR genotyping and sequencing of the locus

as previously described.

Crossing the IgK knock-in mice with the Rpl13a knock-in mice will produce an F1

generation with PQR-GFP and PQR-RFPnols insertions at the endogenous Igk and Rpl13a loci,

respectively. Cells from this animal will stoichiometrically produce GFP and RFP under the

control of the endogenous genomic IgK and Rpl13A loci, respectively. Every cell in this animal

Page 126: Observation and quantification of protein production in single living cells

114

will have red fluorescent nucleoli, with immune cells expressing IgK being green fluorescent.

Using the green to red ratio of fluorescence, the relative protein levels of IgK can be determined

by normalizing to the levels of Rpl13a protein, and this can be used to account for optical

artifacts inherent to in vivo imaging and differences in global transcriptional and translational

states of the cell (Figure 4.2b).

A common limitation of using CRISPRs for genome editing is their tendency to induce

off-target effects that can introduce variability and confound experimental outcomes (Cho et al.,

2014). CRISPRs have been shown to induce DSBs and subsequently mutations at loci with more

than 5-nucleotide differences in the sequence of the target (Fu et al., 2013). Excessive off-target

effects can induce large chromosomal rearrangements that can be toxic to cells, in addition to the

misregulation and potential activation of deleterious genes such as oncogenes (Cho et al., 2014).

Homologous integration of the repair template at non-homologous sites is unlikely to occur due

to extremely low recombination efficiency (10-4 to 10-6 events per basepair per generation)

(Brown et al., 2011). Genome edited mice can easily be backcrossed to the parental strain to

remove almost all the genetic background except the PQR insertion. Backcrossing mice to

parental background at least twice ensures more than 75% of the genetic background has been

isogenized to the parental strain. Therefore, off-target effects pose a bigger problem for in vitro

work.

4.3 Split GFP as a quantitative marker of local protein synthesis in vivo

In Chapter 3, I describe a strategy we developed to use the reconstitution of GFP as a

reliable and quantitative indicator of protein translation in single living cells (Figure 4.3). For

this part of the thesis I aim to create and test different DNA constructs that will be used to

Page 127: Observation and quantification of protein production in single living cells

115

generate animals that express split GFP components for the quantitative imaging of local protein

synthesis events in vivo.

The local synthesis of proteins occurs as transient small magnitude events in spatially

restricted compartments far from the cell soma. We can take advantage of the small size and

rapid translation time of GFP11 peptides and place those downstream of PQR sequences. This

allows the generation of a protein synthesis reporter that requires ~ 3 seconds to be translated,

does not require any folding or post translational modification and is thus free to reconstitute

with the larger GFP1-10 fragment and emit green fluorescence. As demonstrated in Chapter 3,

the reconstitution of GFP occurs at extremely high temporal resolution, and this can be taken

advantage of to use GFP11 as a PQR reporter of local protein translation. By generating animals

that express separate split GFP reporters under control of a gene of interest, green fluorescent

signals observed in the F1 generation would be indicative of GFP reconstitution and thus a

protein translation event.

The concentration of GFP1-10 in this assay is a rate-limiting step. To fully exploit that

GFP11 can reconstitute with GFP1-10 in milliseconds, high levels of GFP1-10 are required

throughout the cell, particularly in small distal sites such as neuronal dendrites and axonal tips.

In the next section, I discuss our strategy to generate animals that stably express high levels of

GFP1-10 ubiquitously, and how they will be used as a reagent in our approach to visualize local

protein synthesis events.

4.3.1 Generation of animals constitutively expressing GFP1-10

To produce a stable and high level source of GFP1-10 expressing cells in vivo, we

generated transgenic animals that expresses GFP1-10 under the control of the ubiquitous actin

promoter. This transgenic animal has many uses: First, cell from any tissue can be used as a

Page 128: Observation and quantification of protein production in single living cells

116

source of high levels of GFP1-10 protein. Second, primary cell cultures where homogenous and

endogenous GFP1-10 expression is required can easily be generated from any tissue of the

animal. Third, by expressing GFP1-10 in one animal and simply crossing it to another expressing

a GFP11 reporter, fluorescence reconstitution is allowed to occur. This means these animals can

be repeatedly used in future local translation experiments using split GFP, without having to

modify their genotype. In other words, to screen for locally translated candidate mRNAs,

“Protein of interest-PQR-GFP11” animals can be crossed to the same GFP1-10 animal. This

makes the GFP1-10 animal a versatile and practical reagent for my work as well as the research

community.

There typically exist just over 1 million copies of Actin monomers per mammalian cell

typically at concentrations up to 95 µM in the cytoplasm (Luby-Phelps, 2000). This constitutes

roughly 1% of the total proteome, making Actin one of the most abundant proteins in any cell

containing a cytoskeleton (Lodish et al., 2008). In addition, the high cellular levels of Actin are

directly dependent on the long half-life of the actin mRNA (Dormoy-Raclet et al., 2007)

Therefore, driving the expression of GFP1-10 expression using the actin promoter in vivo can

reasonably be predicted to produce high levels of GFP1-10 high levels in ubiquitous cell types

throughout the life of the animal (Qin et al., 2010; Quitschke et al., 1989). To achieve a two-fold

increase in fluorescence intensity over cellular background autofluorescence, it was shown that

EGFP must be expressed at 200 nM, which translates to roughly 10,000 diffuse molecules in a

typical cytoplasm (Patterson et al., 1997; Snapp, 2005). However, fewer molecules may be

detected if they are spatially localized into a small organelle such as into the nucleolus (See

Chapter 2).

Page 129: Observation and quantification of protein production in single living cells

117

Using standard transgenic practices, we generated a fly that expresses GFP1-10 driven by

the fly actin (ActB) promoter. To verify the expression of GFP1-10 protein in this fly, we first

performed an immunoblot using adult fly head extracts. Using an anti-GFP antibody, we found

that the predicted 24 kDa GFP1-10 protein is correctly translated and runs at the expected size on

an SDS-PAGE gel (Figure 4.4a). To examine whether GFP1-10 is expressed at high levels in this

fly, we performed immunohistochemistry using the same anti-GFP antibody on larval brain

samples. We confirmed that nonfluorescent GFP1-10 is expressed and present at high levels as

shown by its accumulation in both neuronal and glial cell bodies and neurite projections of the

adult fly brain (Figure 4.4b) and larval ventral nerve cord (Figure 4.4c). From this evidence, we

conclude that the ActB promoter can drive the high-level expression of nonfluorescent GFP1-10

in flies, and this is particularly obvious in the somas and projections of cells in the fly brain.

To enable the translation of such experiments to mammalian systems, we are generating

transgenic mice that express GFP1-10 under the control of the beta actin promoter, a constitutive

and high expression promoter (Gunning et al., 1987; Qin et al., 2010). Mouse pups from the first

round of injection are now born and the correct insertion of GFP1-10 has been verified by

genotyping and Sanger sequencing of PCR products from genomic DNA.

4.3.2 Local translation of Gurken protein in Drosophila oocytes

The first step we took to demonstrate the applicability of the technique in vivo is to

identify a model or system best suited for the unambiguous identification of local translation.

The rationale is to take advantage of a known and characterized system in which the local

translation of a specific gene is known to occur. The local translation of mRNAs in the

Drosophila oocyte is one of the earliest and best studied examples of how the localization of

mRNA is used as a mechanism for translational regulation (Cáceres & Nilson, 2005; Driever &

Page 130: Observation and quantification of protein production in single living cells

118

Nüsslein-Volhard, 1988; Martin & Ephrussi, 2010). During egg development, germ cell

specification and embryonic axis patterning are established via molecular asymmetries created

by position-dependent regulation of the translation of mRNAs deposited maternally into the

oocyte (Richter & Lasko, 2011)

The 1.7 kb grk mRNA encodes the Gurken protein, a torpedo/EGF receptor ligand, a

receptor located on the inner surface of follicle cells that envelop the oocyte. A conserved RNA

stem loop element within the grk coding region forms the signal for dynein-dependent grk

mRNA transport and localization (Van De Bor et al., 2005). Additional elements in the 5’ and

3’UTRs ensure proper translational regulation (Saunders & Cohen, 1999). Gurken local

translation at the anterodorsal corner establishes a molecular EGF signaling gradient such that

the highest signalling occurs in neighbouring anterodorsal follicle, initiating cell fate

specification (Figure 4.5a) (Nilson & Schüpbach, 1999).

To observe the local translation of gurken transcripts within the Drosophila oocyte, we

generated a transgenic fly that expresses Grk-PQR-GFP11 under the control of the

transcriptional activator, Gal4 (Figure 4.5a). The Gal4 activator binds to elements in the

upstream activating sequence (UAS), a sequence placed upstream of the construct which allows

the recruitment of the transcription machinery and the subsequent initiation of transcription. The

spatially restricted expression of Gal4 allows the transcription of UAS transgenes only in those

tissues (Brand & Perrimon, 1993; Duffy, 2002). Widely available tissue-specific Gal4 lines such

as nanos-Gal4 allow the spatial restriction of Gal4 expression to germline cells (Rørth, 1998),

which results in the restricted expression of Grk-PQR-GFP11 specifically in those Gal4-

expressing cells. To generate the DNA construct for microinjection into embryos, the entire

1.7kb Grk cDNA including the coding sequence and the 5’ and 3’UTRs were used to flank a

Page 131: Observation and quantification of protein production in single living cells

119

Drosophila codon-optimized PQR-GFP11 reporter (Figure 4.5a). Our reason for including the

complete transcript sequence was to preserve all potential regulatory elements of the native Grk

mRNA, ensuring the correct processing, localization, and translational regulation of transcripts

generated from the integrated transgene.

The fundamental rationale behind the experiment is to cross the GFP1-10 and Grk-PQR-

GFP11 transgenic flies and observe Gurken local translation using GFP reconstitution in the F1

generation. F1 females bear the genotype ;GFP1-10/UAS-Grk-PQR-GFP11; nosGal4/+ and this

means that oocytes produced from these females contain high levels of GFP1-10 and nosGal4,

which will drive the expression of the Grk-PQR-GFP11 transgene in the oocyte. Fluorescence

imaging of Stage 9 and 10 oocytes showed a crescent shaped green signal that always localized

to the corner of the oocyte chamber where the nucleus was located (Figures 4.5b-c, n=6). The

observed Gurken translation pattern was consistent with the characteristic Grk signal observed

using immunofluorescence or in situ hybridization studies (Cáceres & Nilson, 2005; Jaramillo et

al., 2008; MacDougall et al., 2003). Imaging of control genotypes in which a component of the

split GFP system is omitted, or if the expression of UAS-Grk-PQR-GFP11 is repressed by

removing Gal4, showed complete absence of a green signal that localizes to the anterodorsal

corner in all assayed control oocytes (Figures 4.5d-e, n>25). This experiment demonstrates that

GFP11 reporters can be translated and can reconstitute with GFP1-10 with high spatial

resolution, and this can be used to monitor the local synthesis of proteins in vivo.

A major problem encountered in fluorescence imaging of Drosophila oocytes is the high

level of autofluorescence seen in several emission channels (Figures 4.4 and 4.5) (Boulina et al.,

2013; Parton et al., 2010). The level of autofluorescence in these tissues can easily obscure the

observation of true fluorescence signals, if the signal intensity is low (Mavrakis et al., 2008). For

Page 132: Observation and quantification of protein production in single living cells

120

example, the oocyte chamber contains a yolk which harbours factors necessary to sustain the

development of the oocyte and embryo. The concentration of macromolecular factors in the

cytoplasm can therefore reach concentrations as high as 400 mg/ml (Guigas et al., 2007), which

in addition to dense packing of tissue has been long known to cause nonspecific autofluorescence

emission (Mavrakis et al., 2008; H.-W. Wang et al., 1998). The high autofluorescence seen when

imaging Drosophila oocytes may explain the perceived low fluorescence intensity of the

observed GFP signals in the Gurken local translation experiments. Another possible explanation

is inefficient split GFP translation or reconstitution in an in vivo context. While it is difficult to

directly and quantitatively test reconstitution efficiency in vivo, the reconstitution and translation

of split GFP components can be examined in vitro. For example, in our constant effort to develop

split reporters can reconstitute faster and more efficiently, we recently identified a novel GFP11

variant, GFP11-OPT, that seems to exhibit superior reconstitution efficiency when expressed in

cultured Drosophila S2 cells (Figure 4.6a). The reconstitution efficiency is assessed by

examining how early the fluorescence can be seen from cells, and the characteristics of the signal

including relative brightness and spatial distribution within the cell. Once an adequate candidate

is identified, such as GFP11-OPT, the peptide is commercially synthesized and the reconstitution

kinetics are measured and quantitatively compared to the currently available GFP11 variants.

It is therefore reasonable to consider that using GFP11-OPT instead of the current

reporter in the Grk-PQR-GFP11 transgenic fly and future animals, might produce more reliable

fluorescence signals that will enable unambiguous identification of local protein translation

events. Weakly expressed transcripts, or imaging sites with high autofluorescence can preclude

the high contrast detection of signals. Therefore, it is beneficial to have faster and more efficient

reconstitution reporters.

Page 133: Observation and quantification of protein production in single living cells

121

4.3.3 Detection of local protein translation in living neurons

My ultimate goal in developing this technique is to develop tools to visualize and

quantify local protein translation events in living neuronal subcellular compartments. Ionotropic

and metabotropic neurotransmitter receptors are some of the most abundant proteins present at

synapses and they mediate most of neuronal communication in the brain. Glutamate is the most

abundant excitatory neurotransmitter in the vertebrate nervous system, and glutamate receptor

signaling is the brain’s main excitatory signaling mechanism. The AMPA (α-amino-3-hydroxy-

5-methyl-4-isoxazolepropionic acid) receptor is an ionotropic transmembrane receptor for

glutamate that in addition to mediating synaptic transmission, has been implicated as a major

player in forms of synaptic plasticity that underlie learning and memory, cognition, long term

potentiation and synaptic scaling (Henley & Wilkinson, 2013; Ibata et al., 2008; Willard &

Koochekpour, 2013). AMPA receptors are composed of highly conserved subunits encoded by

genes termed GluR1-4 (Gria1-4), that combine to form tetramers that are trafficked into and out

of synapses (Willard & Koochekpour, 2013). AMPA receptor trafficking is mediated by the

interaction of subunit-specific proteins in addition to various post-translational modifications that

occur at their C-termini, and is known to be highly dynamic (Anggono & Huganir, 2012). The

number of AMPA receptors present at synapses and spines is directly dependent on the rates of

endocytosis and exocytosis at the post-synaptic membrane. Increases in synaptic activity result in

enhanced GluR1-containing AMPA receptor exocytosis and insertion, while increased receptor

endocytosis rates have been associated with synaptic long-term depression (Kessels & Malinow,

2009).

As mentioned in Chapter 1, some of the earliest dendrite-targeted mRNAs that were

identified included BDNF, Arc and αCaMKII. AMPA receptor subunit mRNAs have also been

Page 134: Observation and quantification of protein production in single living cells

122

found to localize to dendrites and spines, which suggests that AMPA receptor local translation

regulates the local abundance and composition of the synaptic receptor pool (Grooms et al.,

2006; Ju et al., 2004). Endogenous GluR1 mRNAs have been reported to be localized to

proximal and distal dendrites of rat hippocampal neurons (Grooms et al., 2006). In addition,

transfected tagged GluR1 and GluR2 subunits have been observed to be synthesized in dendritic

compartments, independently of protein synthesis in the cell soma (Ju et al., 2004).

The local translation of AMPA receptors in neurons is therefore an attractive system to

test our local protein synthesis reporters in vivo. Specifically, we can take advantage of the

characterized expression patterns of GluR1 AMPA receptor subunits and the extensive literature

covering the localization of GluR1 mRNA and its regulation by neuronal activity. This enables

the design proof of principle experiments that would allow the unambiguous detection of GluR1

local protein synthesis in living neurons.

The design and rationale of the experiment to visualize GluR1 local translation in

neurons is similar to that used to observe Gurken local translation in the fly oocyte. A genome

edited animal expressing GluR1-PQR-GFP11 is crossed to an animal expressing GFP1-10 and

fluorescence imaging of GFP in neurons is used as a spatial and temporal marker of local GluR1

production. To generate an animal that expresses a GFP11 reporter each time a GluR1 subunit is

translated, we used CRISPR genome editing to insert a PQR-GFP11 reporter at the end of the

coding sequence of the endogenous GluR1 locus (Gria1) in mouse pronuclei. Every cell that

expresses GluR1 subunits will proportionally co-produce GFP11 reporters. In the presence of

GFP1-10 protein, the split GFP components can reconstitute and emit fluorescence. The

fluorescent signal is thus used to mark the time and location of protein translation and the

fluorescence intensity can be used as a quantitative measure of the level of GluR1 production.

Page 135: Observation and quantification of protein production in single living cells

123

Injections of Cas9 protein, repair oligonucleotides and gRNAs targeted against the GluR1

locus have been performed and the resulting mice are currently being tested for the correct

insertion of PQR-GFP11 into the GluR1 locus. A simple validation experiment would entail

crossing the transgenic GFP1-10 and genome edited GluR1-PQR-GFP11 mice and imaging in

the F1 generation brain for the development of green fluorescence signals in distal neurites. To

validate any observed signals and to determine that they in fact reflect known GluR1 dynamics,

we can take advantage of characterized scenarios in which GluR1 is known to be locally

produced. For example, it has been shown that potentiation of hippocampal synaptic currents

using pharmacological agents, or by dopamine receptor activation increases the surface

expression of GluR1 subunits in mice (Doyle & Kiebler, 2011; Ju et al., 2004; Smith et al.,

2005). GluR1 mRNA has also been shown to localize to distal dendrites of rodent cortical and

hippocampal neurons (Chen, Onisko, & Napoli, 2008; Doyle & Kiebler, 2011; Grooms et al.,

2006; Muddashetty et al., 2007).

4.4 Detection of local protein synthesis using PQR photoconvertible reporters.

The use of fluorescence proteins to detect local protein synthesis events has been

hindered by the long folding and maturation times of common fluorescent proteins (S. Kim et al.,

2010; Shaner et al., 2005). Immature fluorescent proteins can diffuse up to 50 µm2/ sec in their

non-fluorescent state producing a high temporal and spatial discrepancy between the original site

of protein synthesis, and the site where the fluorescent signal is observed. To address this issue,

photoconvertible fluorescent proteins have been used to track local protein translation in cells

(Kim et al., 2010; Wang et al., 2009).

Kaede is a 28kDa photoconvertible fluorescent protein that exists as a 116 kDa

homotetramer in vivo (Ando et al., 2002). Kaede normally emits green (~518 nm) fluorescence

Page 136: Observation and quantification of protein production in single living cells

124

but can be permanently photoconverted to emit red (~580 nm) fluorescence. Along with Dendra2

and mEos2, it is one of the commonly used photoactivatable fluorescent proteins to study local

protein translation, particularly in neurons (S. Kim et al., 2010). By photoconverting areas of the

cell from green to red, the development of new green fluorescence is associated to the translation

of new protein. However, a simple back-of-the-envelope calculation of the diffusion time and

distance of homotetrameric Kaede in typical mammalian cytoplasm shows that the protein can

diffuse upwards of 60 microns in just 1 minute. Although the maturation time of Kaede has not

yet been accurately measured, we can assume for the sake of argument that Kaede has a

maturation time identical to one of the fastest measured fluorescence protein Venus, with a

measured maturation time of 2 minutes (Nagai et al., 2002; Tatavarty et al., 2012). In 2 minutes

nonfluorescent Kaede can diffuse 90 microns; easily traversing the length of an average

mammalian neuronal soma (20 microns (García-López et al., 2006)). In addition, some of the

largest dendritic spines have volumes of 1 µm3 (Nimchinsky et al., 2002) and are spaced 1 µm

apart in more extensively branched neurons. Using slow-maturing fluorescent proteins therefore

poses spatial and temporal constraints and the examples above serve to show that in such

experiments, where Kaede is seen is likely not exactly where it was made.

Using split-photoconvertible fluorescent proteins is the most suitable approach to detect

local protein translation repeatedly, and if optimized can result in highly quantitative spatial and

temporal measurements of protein production over time. A splitKaede or splitDendra2 system

would work similar to splitGFP reconstitution, where the Dendra2 protein would be split into

two non-fluorescent parts of unequal sizes (Figure 4.6b). The larger, slower folding portion part

would be expressed at high levels in all parts of the cell, similar to GFP1-10, and the remaining

polypeptide would be placed downstream of a gene of interest, separated by a PQR. When the

Page 137: Observation and quantification of protein production in single living cells

125

protein of interest is translated, Dendra2 reconstitution would occur in milliseconds and mark the

exact time, location and rate of protein translation. The photoconvertible property therefore acts

a reset switch, overcoming signal saturation issues and would be used to define a time zero for

monitoring new translation events, as seen by development of new green fluorescence. Such as

system would combine the numerous advantages of co-expression of reporters using PQR, the

fast reconstitution of a fluorescent protein as a marker of translation, and an arbitrary time zero at

which measurements can be started.

Splitting a fluorescent protein into two nonfluorescent parts that can spontaneously

reassemble with no co-factors or enzymes presents several problems. In order for a split FP

system to work properly in the context of local protein synthesis detection, it should satisfy some

basic criteria. First and most important, each protein half must not exhibit any activity on its

own. Second, the affinities of the split parts to each other must be sufficiently high to ensure

proper interaction, and third the reconstituted FP must have an easily quantifiable readout. While

most fluorescent proteins in theory can be split into interacting parts, finding regions along the

protein structure that result in partners that conserve the above criteria is less straightforward. If

split FP parts that exist in near native conformation can be generated, then their binding affinity

could be made significantly higher (Huang & Bystroff, 2009). This is counterproductive for

traditional protein interaction studies using split fluorescent proteins such as mGRASP, since the

reporter could reconstitute without the interaction of the assayed proteins. However, in the

context of developing a fast protein translation sensor, increased binding affinity means reduced

delay in reconstitution and fluorescence emission after translation of the reporter. The original

split GFP developed by Cabantous et al. relied solely on the binding of the 11th beta barrel for

fluorescence reconstitution to occur, however, the slow maturation of the chromophore led to the

Page 138: Observation and quantification of protein production in single living cells

126

resulting fluorescence requiring several hours to reach a plateau (Cabantous, Terwilliger, et al.,

2005). This problem stimulated efforts to find variants with shorter chromophore maturation

times, and using DNA shuffling (Stemmer, 1994), the Waldo group developed GFP1-10

optimum (GFP1-10 OPT) which had 80-fold better reconstitution efficiency compared to the

original split GFP (Cabantous, Pédelacq, et al., 2005).

Experiments to split Dendra2 into non-fluorescent and spontaneously interacting partners

that can be used as local protein translation reporters are in progress and we have successfully

split the Dendra2 molecule into two nonfluorescent parts, which reconstitute to form a green

fluorescent protein that can be permanently photoconverted to emit red fluorescence using

ultraviolet light (Figure 4.6). High homology between the Dendra2 and GFP 3D crystal

structures allowed us to take advantage of the wealth of information we have on splitting and

folding GFP, and applied it to Dendra2. Similar to GFP, Dendra2 is composed of beta turn

secondary structures and beta sheets that envelop the chromophore in a barrel-like structure. This

allowed the comparison of the Dendra2 and GFP individual protein domains in parallel, which

offered locations along the sequence where the protein can in theory be split to generate

reconstitution partner candidates. Specifically, we split Dendra2 between the 190th and 191st

residue (Figure 4.6b) and verified that they could spontaneously reconstitute and emit green (~

507 nm) fluorescence by expressing the larger and smaller Dendra2 fragments in HEK293 cells

(Figure 4.6c).

Upon exposure to a flash of ultraviolet (~350 nm) light, we found that reconstituted

Dendra2 could be permanently photoconverted to emit red (~573 nm) fluorescence (Figure 4.6c).

By quantifying the levels of green and red fluorescence over time, we observed that immediately

following photoconversion, red fluorescence intensities rose and steadily declined over time,

Page 139: Observation and quantification of protein production in single living cells

127

while green fluorescence intensities dropped and then steadily rose over tens of minutes (Figure

4.6c). These results indicate that reconstituted green Dendra2 was permanently photoconverted,

and rising green fluorescence intensities within minutes post photoconversion suggest that new

Dendra 11 synthesis and reconstitution events could be observed immediately after

photoconversion. Importantly, no increases in fluorescence intensity were observed in the red

channel post photoconversion (Figure 4.6c), which suggests that unreconstituted Dendra1-10

does not photoconvert in the absence of Dendra11, and still emits green fluorescence when

reconstituted. This allows split Dendra2 to be used repetitively as a reset switch to constantly

allow observation of new protein synthesis events in the green channel.

Page 140: Observation and quantification of protein production in single living cells

128

4.5 Figures

a

b

Figure 4.1: PQR reporters are inserted in-frame into endogenous genes.

(a) An edited IgK locus showing the location of the insertion of a PQR-GFP reporter. The kappa

immunoglobulin locus is composed of 174 variable genes, 5 joining genes and 1 constant gene

that rearrange to form the mature light chain. One variable gene is joined to one joining gene and

both are joined to the constant gene. (b) The reporter is inserted at the end of the coding

sequence of the constant region, to avoid disruption of the open reading frame and result in

stoichiometric coexpression of GFP with each molecule of IgK.

Page 141: Observation and quantification of protein production in single living cells

129

a

b

Figure 4.2: PQR constructs injected into mouse embryos result in red fluorescent

pronuclei.

(a) Preliminary results showing mouse pronuclei that were co-injected with Rpl13a-specific

CRISPR and repair plasmids. Insertion of PQR-RFPnols reporters into the endogenous RPL13A

locus was achieved and results in fluorescent embryos. Injections and implantation of red-

fluorescent embryos are ongoing in collaboration with Mitra Cowan and the Goodman Cancer

research centre. Scale bar is 100 µm. (b) Hypothetical results obtained from infection of a

Rpl13a-PQR-RFPnols/IgK-PQR-GFP double knock-in mouse. Green indicates antibody response,

red indicate housekeeping protein production. Changes in IgK production between cells or

Page 142: Observation and quantification of protein production in single living cells

130

experiments can be normalized to Rpl13a production levels in those same cells, to reflect the net

increase in IgK protein production.

Page 143: Observation and quantification of protein production in single living cells

131

Figure 4.3: SplitGFP as a protein translation reporter.

Insertion of a PQR sequence between a gene of interest and a GFP11 reporter results in slightly

longer mRNAs being transcribed (< 100 bases), and stoichiometric translation of GFP11. Each

time a molecule of protein of interest is translated, a molecule of GFP11 is co-produced. In the

presence of GFP1-10, the split GFP parts reconstitute to form fluorescent GFP.

Page 144: Observation and quantification of protein production in single living cells

132

Figure 4.4: GFP1-10 is expressed at high levels in transgenic animals.

(a) Immunoblot performed using α-GFP antibody on whole fly and brain and head extracts. The

24 kDa predicted GFP1-10 protein is properly expressed in flies and runs at the expected size.

(b-c) Immunohistochemistry done in collaboration with Sejal Davla, showing ubiquitous high

level expression of GFP1-10 protein in the fly adult brain (b) and larval ventral cord (c). GFP1-

10 can clearly be seen in cell somas (arrows) and projections (arrowheads). Anti-GFP antibody

(cyan) was used to probe for GFP1-10 (cyan), RFP marks projections, nc82 is a marker of

synapses. Scale bar is 50 µm in (b) and 75 µm in (c).

Page 145: Observation and quantification of protein production in single living cells

133

Page 146: Observation and quantification of protein production in single living cells

134

Figure 4.5: GFP11 can detect Gurken local translation in oocytes.

Representative images of local Gurken translation in Drosophila oocytes. (a) The grk transcript

localizes to the anterior dorsal corner of the oocyte near the nucleus, where its translation is

initiated. (b-c) Translation of Grk-PQR-GFP11 shows green fluorescence (arrows) that is always

associated with the anterodorsal nucleus (star) in stage 9 (b) and stage 10 (c). (d-e) Control

animals in which GFP1-10 (d) or Gal4 (e) are not expressed show complete lack of the

characteristic signal that associates with the oocyte nucleus (n>25). Anterior is left, dorsal is up.

Scale bar is 10 µm.

Page 147: Observation and quantification of protein production in single living cells

135

Page 148: Observation and quantification of protein production in single living cells

136

Figure 4.6: Novel split fluorescent reporters exhibit more efficient reconstitution.

(a) GFP11 reporters can be improved to produce more efficient GFP reconstitution in cultured

S2 cells. (b) Illustration showing an example of splitting Dendra2 into two non-fluorescent

reconstitution partners. The protein is split into a larger, slower folding non-fluorescent part and

the remaining smaller polypeptide. The two are separately nonfluorescent but spontaneously

reconstitute to form the native fluorescent Dendra2 structure and emit fluorescence. In the

example above, Dendra2 is split between the 190th and 191st to generate the first split

photoconvertible Dendra2 reconstitution partners.

Page 149: Observation and quantification of protein production in single living cells

137

Chapter V - Thesis directions and conclusions

Page 150: Observation and quantification of protein production in single living cells

138

Proteins execute essential functions in cells. Dysregulation in protein production is a

hallmark feature of many diseases. Tremendous progress has been made towards understanding

when and where gene products are produced, yet we still know little about the complete picture

that is how the expression of genes, and their regulation contribute to the development of cellular

phenotypes. In this thesis, I have expanded the capabilities of our PQR technique by developing

alternative and new ways to examine the expression of genes and subsequent production of

proteins. To pave the way for understanding how proteins execute their functions in cells, I have

taken a two-pronged approach to examine the process of gene expression. By simultaneously

examining the levels of RNA and protein of a gene from a single cell, I have developed a system

to determine how individual cells vary in their transcriptional and translational landscapes within

a population.

My second approach to understand protein production aimed at directly observing the

process of new protein creation. The animals that will be generated to demonstrate endogenous

protein production measurement in vivo, the IgK and Rpl13a knockin mice, are useful not only as

proof of principle experiments in my work, but will also be of use to the wider research

community. Rpl13a is a well-established reference gene whose expression is used to reflect both

the transcriptional and translational status of cells. Therefore, the Rpl13a knockin mouse

provides researchers quantifying the expression of their gene of interest, from isolated cells or in

vivo, to normalize any changes observed to the general transcriptional and translational status of

the cell. The mouse represents a resource that complements, instead of replacing current

approaches used to study the production of proteins. The mouse is amenable to both antibody-

and fluorescence-based measurement and normalization of protein levels, which opens the door

Page 151: Observation and quantification of protein production in single living cells

139

for longer in vivo imaging of the production of proteins in addition to independent means of

quantifying and validating collected data.

The IgK mouse is a translational reporter of immune function. An increasing number of studies

now track immunological responses at the whole-body level in living mice, and this has revealed

a number of previously unknown mechanisms underlying variability in adaptive immune

responses (Nair-Gill et al., 2008), tissue transplantation (Donahue et al., 2015), infection (Prado

et al., 2015; Santangelo et al., 2015) and cancer growth (Edinger et al., 2003). Expression of

genetically encoded fluorescent reporters from endogenous antibody loci allows the optical

tracking of native immune responses at the microscopic and macroscopic levels, without the

need to laboriously label cells ex vivo and reintroduce them into the animal, one of the common

ways of examining immune responses in a living animal (Germain et al., 2006). Moreover,

genetically tagging antibody chains with fluorescent reporters opens the door for multicolor

imaging of immune responses, which allows the easy identification, screening and isolation of B

cells that express varying levels of antibodies of specific types produced against specific

responses. Ultimately, this mouse is most useful for the screening and development of

therapeutics that modulate immune responses. For example, the development of an immune

response against a pathogen in the IgK mouse can be optically monitored over time by observing

changes in GFP fluorescence intensity from B cells in the lymphatic system (Figure 4.2b).

Similarly, the progression of the infection or disease and efficacy of therapy can be optically

monitored in that same mouse, providing unprecedented resolution into the mechanisms that

mediate heterogeneous individual immune responses (Satija et al., 2014).

The concept of local protein translation in neurons has been examined over decades, and

only now are we beginning to unravel its mechanisms and understand its impact. My strategy to

Page 152: Observation and quantification of protein production in single living cells

140

use split fluorescent proteins as reporters of small protein translation events, is particularly well

suited to examine the occurrence of these events in neuronal cells. The fast reconstitution time,

linear dependence of the brightness of the signal on the concentration of reporter and

stoichiometric production of the genetically encoded reporter makes the technique amenable to

long-term in vivo imaging of protein production in neuronal subcellular compartments, which

will reveal unprecedented insight into the dynamics of gene expression response far from the cell

nucleus.

In continuing our progression of developing the PQR technique, we have found that local

protein translation reporters could be enhanced by optimizing their sequences for organism-

tailored split reporters. These reporters are more efficient at reconstitution in terms of speed,

brightness and efficiency (Figure 4.6a). Finally, our efforts to create split photoconvertible

fluorescent proteins may offer a complete shift in experiments examining protein production in

living cells. By photoconverting cells or subcellular compartments to a different spectrum, the

dynamics of protein production can be assayed from the same cell over the life of the animal,

with no issues of signal saturation or small dynamic range, in multiple channels. This type of

extended information obtained from single cells is extremely valuable. We know that cellular

heterogeneity could explain a substantial portion of the variability seen in assays conducted at

the tissue or population level (J Eberwine et al., 2001; Schubert, 2011). Upgraded reporters bring

increased spatial and temporal accuracy in determining protein levels, and so by ever-improving

reporters we can improve the quality and value of data that is collected. Similar to the efforts that

led to brighter and more stable GFPs, or the constant adaptation and expansion of optogenetic

techniques (Deisseroth, 2015; Remington, 2011; Tsien, 1998), I believe that the story of our PQR

Page 153: Observation and quantification of protein production in single living cells

141

technique has just begun. Protein analysis in single cells is certain to keep improving and open

new ways to understand cell biology.

Page 154: Observation and quantification of protein production in single living cells

142

References

Ables, E. T. (2015). Drosophila Oocytes as a Model for Understanding Meiosis: An Educational

Primer to Accompany “Corolla Is a Novel Protein That Contributes to the Architecture of

the Synaptonemal Complex of Drosophila.” Genetics, 199(1).

Abrahamsen, H. N., Nexo, E., Steiniche, T., Hamilton-Dutoit, S. J., & Sorensen, B. S. (2004).

Quantification of Melanoma mRNA Markers in Sentinel Nodes. The Journal of Molecular

Diagnostics, 6(3), 253–259. https://doi.org/10.1016/S1525-1578(10)60518-1

Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., & Walter, P. (2002). Molecular

Biology of the Cell, Fourth Edition. Molecular Biology. https://doi.org/citeulike-article-

id:691434

Alwine, J. C., Kemp, D. J., & Stark, G. R. (1977). Method for detection of specific RNAs in

agarose gels by transfer to diazobenzyloxymethyl-paper and hybridization with DNA

probes. Proceedings of the National Academy of Sciences of the United States of America,

74(12), 5350–4. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/414220

Ando, R., Hama, H., Yamamoto-Hino, M., Mizuno, H., & Miyawaki, A. (2002). An optical

marker based on the UV-induced green-to-red photoconversion of a fluorescent protein.

Proceedings of the National Academy of Sciences of the United States of America, 99(20),

12651–6. https://doi.org/10.1073/pnas.202320599

Anggono, V., & Huganir, R. L. (2012). Regulation of AMPA receptor trafficking and synaptic

plasticity. Current Opinion in Neurobiology, 22(3), 461–9.

https://doi.org/10.1016/j.conb.2011.12.006

Bagni, C., Tassone, F., Neri, G., Hagerman, R., Martin, J., Bell, J., … Narayanan, U. (2012).

Fragile X syndrome: causes, diagnosis, mechanisms, and therapeutics. The Journal of

Clinical Investigation, 122(12), 4314–22. https://doi.org/10.1172/JCI63141

Bagshaw, C. R., & Cherny, D. (2006). Blinking fluorophores: what do they tell us about protein

dynamics? Biochemical Society Transactions, 34(Pt 5), 979–982.

https://doi.org/10.1042/BST0340979

Baker, M. (2015). Reproducibility crisis: Blame it on the antibodies. Nature, 521(7552), 274–

276. https://doi.org/10.1038/521274a

Bakshi, S., Siryaporn, A., Goulian, M., & Weisshaar, J. C. (2012). Superresolution imaging of

ribosomes and RNA polymerase in live Escherichia coli cells. Molecular Microbiology,

85(1), 21–38. https://doi.org/10.1111/j.1365-2958.2012.08081.x

Bassell, G. J., & Warren, S. T. (2008). Fragile X Syndrome: Loss of Local mRNA Regulation

Alters Synaptic Development and Function. Neuron.

https://doi.org/10.1016/j.neuron.2008.10.004

Belle, A., Tanay, A., Bitincka, L., Shamir, R., & O’Shea, E. K. (2006). Quantification of protein

half-lives in the budding yeast proteome. Proceedings of the National Academy of Sciences

of the United States of America, 103(35), 13004–9.

https://doi.org/10.1073/pnas.0605420103

Benito, J., Zheng, H., Ng, F. S., & Hardin, P. E. (2007). Transcriptional feedback loop

regulation, function, and ontogeny in Drosophila. Cold Spring Harbor Symposia on

Quantitative Biology, 72, 437–44. https://doi.org/10.1101/sqb.2007.72.009

Besse, F., & Ephrussi, A. (2008). Translational control of localized mRNAs: restricting protein

synthesis in space and time. Nature Reviews Molecular Cell Biology, 9(12), 971–980.

https://doi.org/10.1038/nrm2548

Boulina, M., Samarajeewa, H., Baker, J. D., Kim, M. D., & Chiba, A. (2013). Live imaging of

Page 155: Observation and quantification of protein production in single living cells

143

multicolor-labeled cells in Drosophila. Development, 140(7).

Brand, A. H., & Perrimon, N. (1993). Targeted gene expression as a means of altering cell fates

and generating dominant phenotypes. Development (Cambridge, England), 118(2), 401–15.

Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/8223268

Brown, A. D., Claybon, A. B., & Bishop, A. J. R. (2011). A conditional mouse model for

measuring the frequency of homologous recombination events in vivo in the absence of

essential genes. Molecular and Cellular Biology, 31(17), 3593–3602.

https://doi.org/10.1128/MCB.00848-10

Bustin, S. A. (2000). Absolute quantification of mrna using real-time reverse transcription

polymerase chain reaction assays. Journal of Molecular Endocrinology.

https://doi.org/JME00927 [pii]

Bustin, S. A., Benes, V., Garson, J. A., Hellemans, J., Huggett, J., Kubista, M., … Wittwer, C. T.

(2009). The MIQE guidelines:Minimum Information for publication of quantitative real-

time PCR experiments. Clinical Chemistry, 55(4), 611–622.

https://doi.org/10.1373/clinchem.2008.112797

Bustin, S. A., & Nolan, T. (2004). Pitfalls of quantitative real-time reverse-transcription

polymerase chain reaction. Journal of Biomolecular Techniques : JBT, 15(3), 155–66.

https://doi.org/15/3/155 [pii]

Bustin, S., & Nolan, T. (2004, September). Pitfalls of quantitative real- time reverse-transcription

polymerase chain reaction. Journal of Biomolecular Techniques. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/15331581

Cabantous, S., Pédelacq, J. D., Mark, B. L., Naranjo, C., Terwilliger, T. C., & Waldo, G. S.

(2005). Recent advances in GFP folding reporter and split-GFP solubility reporter

technologies. Application to improving the folding and solubility of recalcitrant proteins

from Mycobacterium tuberculosis. Journal of Structural and Functional Genomics, 6(2–3),

113–119. https://doi.org/10.1007/s10969-005-5247-5

Cabantous, S., Terwilliger, T. C., & Waldo, G. S. (2005). Protein tagging and detection with

engineered self-assembling fragments of green fluorescent protein. Nature Biotechnology,

23(1), 102–7. https://doi.org/10.1038/nbt1044

Cabantous, S., & Waldo, G. S. (2006). In vivo and in vitro protein solubility assays using split

GFP. Nature Methods, 3(10), 845–854. https://doi.org/10.1038/nmeth932

Cáceres, L., & Nilson, L. A. (2005). Production of gurken in the nurse cells is sufficient for axis

determination in the Drosophila oocyte. Development (Cambridge, England), 132(10),

2345–53. https://doi.org/10.1242/dev.01820

Callahan, B. P., & Miller, B. G. (2007). OMP decarboxylase—An enigma persists. Bioorganic

Chemistry, 35(6), 465–469. https://doi.org/10.1016/j.bioorg.2007.07.004

Cambridge, S. B., Gnad, F., Nguyen, C., Bermejo, J. L., Krüger, M., & Mann, M. (2011).

Systems-wide proteomic analysis in mammalian cells reveals conserved, functional protein

turnover. Journal of Proteome Research, 10(12), 5275–84.

https://doi.org/10.1021/pr101183k

Chen, N., Onisko, B., & Napoli, J. L. (2008). The nuclear transcription factor RARalpha

associates with neuronal RNA granules and suppresses translation. The Journal of

Biological Chemistry, 283(30), 20841–7. https://doi.org/10.1074/jbc.M802314200

Cho, S. W., Kim, S., Kim, Y., Kweon, J., Kim, H. S., Bae, S., & Kim, J. S. (2014). Analysis of

off-target effects of CRISPR/Cas-derived RNA-guided endonucleases and nickases.

Genome Research, 24(1), 132–141. https://doi.org/10.1101/gr.162339.113

Page 156: Observation and quantification of protein production in single living cells

144

Chudakov, D. M., Matz, M. V., Lukyanov, S., & Lukyanov, K. A. (2010). Fluorescent Proteins

and Their Applications in Imaging Living Cells and Tissues. Physiological Reviews, 90(3).

Cody, C. W., Prasher, D. C., Westler, W. M., Prendergast, F. G., & Ward, W. W. (1993).

Chemical structure of the hexapeptide chromophore of the Aequorea green-fluorescent

protein. Biochemistry, 32(1979), 1212–1218. https://doi.org/10.1021/bi00056a003

Cox, L. J., Hengst, U., Gurskaya, N. G., Lukyanov, K. A., & Jaffrey, S. R. (2008). Intra-axonal

translation and retrograde trafficking of CREB promotes neuronal survival. Nature Cell

Biology, 10(2), 149–59. https://doi.org/10.1038/ncb1677

Craggs, T. D. (2009). Green fluorescent protein: structure, folding and chromophore maturation.

Chemical Society Reviews, 38(10), 2865–75. https://doi.org/10.1039/b903641p

Craig, F. F., Simmonds, A. C., Watmore, D., McCapra, F., & White, M. R. (1991). Membrane-

permeable luciferin esters for assay of firefly luciferase in live intact cells. The Biochemical

Journal, 276 ( Pt 3, 637–41. Retrieved from

http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=1151052&tool=pmcentrez&ren

dertype=abstract

Crameri, A., Whitehorn, E. A., Tate, E., & Stemmer, W. P. C. (1996). Improved green

fluorescent protein by molecular evolution using DNA shuffling. Nature Biotechnology,

14(3), 315–319. https://doi.org/10.1038/nbt0396-315

Cubitt, a B., Heim, R., Adams, S. R., Boyd, a E., Gross, L. a, & Tsien, R. Y. (1995).

Understanding, improving and using green fluorescent proteins. Trends in Biochemical

Sciences, 20(11), 448–55. https://doi.org/10.1016/S0968-0004(00)89099-4

Curtis, K. M., Gomez, L. A., Rios, C., Garbayo, E., Raval, A. P., Perez-Pinzon, M. A., …

Horwitz, E. (2010). EF1alpha and RPL13a represent normalization genes suitable for RT-

qPCR analysis of bone marrow derived mesenchymal stem cells. BMC Molecular Biology,

11(1), 61. https://doi.org/10.1186/1471-2199-11-61

Das, D., Georgiadis, M. M., Avidan, O., Loya, S., Tonjes, R. ., Sevilya, Z., … Steitz, T. . (2004).

The crystal structure of the monomeric reverse transcriptase from Moloney murine

leukemia virus. Structure (London, England : 1993), 12(5), 819–29.

https://doi.org/10.1016/j.str.2004.02.032

de Felipe, P., Luke, G. a, Hughes, L. E., Gani, D., Halpin, C., & Ryan, M. D. (2006). E unum

pluribus: multiple proteins from a self-processing polyprotein. Trends in Biotechnology,

24(2), 68–75. https://doi.org/10.1016/j.tibtech.2005.12.006

De Groot, A. S., & Scott, D. W. (2007). Immunogenicity of protein therapeutics. Trends in

Immunology. https://doi.org/10.1016/j.it.2007.07.011

Deisseroth, K. (2015). Optogenetics: 10 years of microbial opsins in neuroscience. Nature

Neuroscience, 18(9), 1213–1225. https://doi.org/10.1038/nn.4091

Dhanasekaran, S., Doherty, T. M., Kenneth, J., & TB Trials Study Group. (2010). Comparison of

different standards for real-time PCR-based absolute quantification. Journal of

Immunological Methods, 354(1–2), 34–9. https://doi.org/10.1016/j.jim.2010.01.004

Ding, C., & Cantor, C. R. (2004). Quantitative analysis of nucleic acids--the last few years of

progress. Journal of Biochemistry and Molecular Biology, 37(1), 1–10.

https://doi.org/<ARTICLE_ID IdType=doi> [pii]

Dittrich, P. S., Schäfer, S. P., & Schwille, P. (2005). Characterization of the photoconversion on

reaction of the fluorescent protein Kaede on the single-molecule level. Biophysical Journal,

89(5), 3446–55. https://doi.org/10.1529/biophysj.105.061713

Do, K., & Boxer, S. G. (2011). Thermodynamics, kinetics, and photochemistry of β-strand

Page 157: Observation and quantification of protein production in single living cells

145

association and dissociation in a split-GFP system. Journal of the American Chemical

Society, 133(45), 18078–81. https://doi.org/10.1021/ja207985w

Dobson, C. M. (2001). The structural basis of protein folding and its links with human disease.

Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences,

356(1406), 133–45. https://doi.org/10.1098/rstb.2000.0758

Donahue, R. E., Srinivasula, S., Uchida, N., Kim, I., St. Claire, A., Duralde, G., … Di Mascio,

M. (2015). Whole-body imaging of the immune system reveals discordance in CD4+

lymphoid tissue recovery after transplantation in rhesus macaques. Blood.

Dormoy-Raclet, V., Ménard, I., Clair, E., Kurban, G., Mazroui, R., Di Marco, S., … Gallouzi, I.-

E. (2007). The RNA-binding protein HuR promotes cell migration and cell invasion by

stabilizing the beta-actin mRNA in a U-rich-element-dependent manner. Molecular and

Cellular Biology, 27(15), 5365–5380. https://doi.org/10.1128/MCB.00113-07

Doyle, M., & Kiebler, M. A. (2011). Mechanisms of dendritic mRNA transport and its role in

synaptic tagging. The EMBO Journal, 30(17), 3540–52.

https://doi.org/10.1038/emboj.2011.278

Driever, W., & Nüsslein-Volhard, C. (1988). A gradient of bicoid protein in Drosophila

embryos. Cell, 54(1), 83–93. https://doi.org/10.1016/0092-8674(88)90182-1

Duffy, J. B. (2002). GAL4 system in Drosophila: a fly geneticist’s Swiss army knife. Genesis

(New York, N.Y. : 2000), 34(1–2), 1–15. https://doi.org/10.1002/gene.10150

Eberwine, J., Kacharmina, E. J., Andrews, C., Miyashiro, K., McIntosh, T., Becker, K., …

Marciano, P. (2001). mRNA Expression Analysis of Tissue Sections and Single Cells. The

Journal of Neuroscience, 21(21), 8310–8314.

Eberwine, J., Miyashiro, K., Kacharmina, J. E., & Job, C. (2001). Local translation of classes of

mRNAs that are targeted to neuronal dendrites. Proceedings of the National Academy of

Sciences of the United States of America, 98(13), 7080–5.

https://doi.org/10.1073/pnas.121146698

Edinger, M., Cao, Y.-A., Verneris, M. R., Bachmann, M. H., Contag, C. H., & Negrin, R. S.

(2003). Revealing lymphoma growth and the efficacy of immune cell therapies using in

vivo bioluminescence imaging. Blood, 101(2).

Feige, M. J., Hendershot, L. M., & Buchner, J. (2010). How antibodies fold. Trends in

Biochemical Sciences. https://doi.org/10.1016/j.tibs.2009.11.005

Feinberg, E. H., VanHoven, M. K., Bendesky, A., Wang, G., Fetter, R. D., Shen, K., &

Bargmann, C. I. (2008). GFP Reconstitution Across Synaptic Partners (GRASP) Defines

Cell Contacts and Synapses in Living Nervous Systems. Neuron, 57(3), 353–363.

https://doi.org/10.1016/j.neuron.2007.11.030

Flatz, L., Roychoudhuri, R., Honda, M., Filali-Mouhim, A., Goulet, J. P., Kettaf, N., … Nabel,

G. J. (2011). Single-cell gene-expression profiling reveals qualitatively distinct CD8 T cells

elicited by different gene-based vaccines. Proc Natl Acad Sci U S A, 108(14), 5724–5729.

https://doi.org/10.1073/pnas.1013084108

Fowler, S. J. (1996). Protein Staining and Immunodetection Using Colloidal Gold. In The

Protein Protocols Handbook (pp. 275–287). Totowa, NJ: Humana Press.

https://doi.org/10.1007/978-1-60327-259-9_44

Fu, Y., Foden, J. a, Khayter, C., Maeder, M. L., Reyon, D., Joung, J. K., & Sander, J. D. (2013).

High-frequency off-target mutagenesis induced by CRISPR-Cas nucleases in human cells.

Nature Biotechnology, 31(9), 822–6. https://doi.org/10.1038/nbt.2623

Fujita, S. C., Zipursky, S. L., Benzer, S., Ferrús, A., & Shotwell, S. L. (1982). Monoclonal

Page 158: Observation and quantification of protein production in single living cells

146

antibodies against the Drosophila nervous system. Proceedings of the National Academy of

Sciences of the United States of America, 79(24), 7929–33. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/6818557

García-López, P., García-Marín, V., & Freire, M. (2006). Three-dimensional reconstruction and

quantitative study of a pyramidal cell of a Cajal histological preparation. Journal of

Neuroscience, 26(44), 11249–11252. https://doi.org/10.1523/JNEUROSCI.3543-06.2006

Gavis, E. R. (1995). Pattern formation. Gurken meets torpedo for the first time. Current

Biology : CB, 5(11), 1252–4. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/8574581

Germain, R. N., Miller, M. J., Dustin, M. L., & Nussenzweig, M. C. (2006). Dynamic imaging of

the immune system: progress, pitfalls and promise. Nat Rev Immunol, 6(7), 497–507.

https://doi.org/10.1038/nri1884

Giorgi, C., & Moore, M. J. (2007). The nuclear nurture and cytoplasmic nature of localized

mRNPs. Seminars in Cell and Developmental Biology.

https://doi.org/10.1016/j.semcdb.2007.01.002

Gissel, C., Doutheil, J., & Paschen, W. (1997). Temporal analysis of changes in neuronal c-fos

mRNA levels induced by depletion of endoplasmic reticulum calcium stores: effect of

clamping cytoplasmic calcium activity at resting levels. Journal of Neurochemistry, 69(6),

2538–45. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9375687

Glinka, M., Herrmann, T., Funk, N., Havlicek, S., Rossoll, W., Winkler, C., & Sendtner, M.

(2010). The heterogeneous nuclear ribonucleoprotein-R is necessary for axonal β-actin

mRNA translocation in spinal motor neurons. Human Molecular Genetics, 19(10), 1951–

1966. https://doi.org/10.1093/hmg/ddq073

Gordon, M. D., & Scott, K. (2009). Motor Control in a Drosophila Taste Circuit. Neuron, 61(3),

373–384. https://doi.org/10.1016/j.neuron.2008.12.033

Greenbaum, D., Colangelo, C., Williams, K., & Gerstein, M. (2003). Comparing protein

abundance and mRNA expression levels on a genomic scale. Genome Biology, 4(9), 117.

Gronowicz, G., Hadjimichael, J., Richards, D., Cerami, A., & Rossomando, E. F. (1992).

Correlation between tumor necrosis factor-alpha (TNF-alpha)-induced cytoskeletal changes

and human collagenase gene induction. Journal of Periodontal Research, 27(6), 562–8.

Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/1334144

Grooms, S. Y., Noh, K.-M., Regis, R., Bassell, G. J., Bryan, M. K., Carroll, R. C., & Zukin, R. S.

(2006). Activity bidirectionally regulates AMPA receptor mRNA abundance in dendrites of

hippocampal neurons. The Journal of Neuroscience : The Official Journal of the Society for

Neuroscience, 26(32), 8339–51. https://doi.org/10.1523/JNEUROSCI.0472-06.2006

Grueber, W. B., Jan, L. Y., Jan, Y. N., & Francisco, S. (2003). Different Levels of the

Homeodomain Protein Cut Regulate Distinct Dendrite Branching Patterns of Drosophila

Multidendritic Neurons, 112, 805–818.

Gubern, C., Hurtado, O., Rodríguez, R., Morales, J. R., Romera, V. G., Moro, M. A., …

Davalos, A. (2009). Validation of housekeeping genes for quantitative real-time PCR in in-

vivo and in-vitro models of cerebral ischaemia. BMC Molecular Biology, 10(1), 57.

https://doi.org/10.1186/1471-2199-10-57

Guigas, G., Kalla, C., & Weiss, M. (2007). Probing the nanoscale viscoelasticity of intracellular

fluids in living cells. Biophysical Journal, 93(1), 316–23.

https://doi.org/10.1529/biophysj.106.099267

Gunawardana, Y., Fujiwara, S., Takeda, A., Woo, J., Woelk, C., & Niranjan, M. (2015). Outlier

Page 159: Observation and quantification of protein production in single living cells

147

detection at the transcriptome-proteome interface. Bioinformatics (Oxford, England),

31(15), 2530–6. https://doi.org/10.1093/bioinformatics/btv182

Gunning, P., Leavittt, J., Muscat, G., Ngt, S.-Y., & Kedes, L. (1987). A human f3-actin

expression vector system directs high-level accumulation of antisense transcripts.

Biochemistry, 84, 4831–4835.

Haas, I. G., & Wabl, M. (1983). Immunoglobulin heavy chain binding protein. Nature,

306(5941), 387–9. https://doi.org/10.1038/306387a0

Hansen, M. C., Palmer, R. J., Udsen, C., White, D. C., Molin, S., & Hansen, M. C. (2016).

Assessment of GFP fluorescence in cells of Streptococcus gordonii under conditions of low

pH and low oxygen concentration. Microbiology, 2337(147), 23–1383.

Hastings, M. H. (2005). Circadian biology: Fibroblast clocks keep ticking. Current Biology.

https://doi.org/10.1016/j.cub.2004.12.012

He, B., & Soderlund, D. M. (2010). Human embryonic kidney (HEK293) cells express

endogenous voltage-gated sodium currents and Nav1.7 sodium channels. Neuroscience

Letters, 469(2), 268–272. https://doi.org/10.1016/j.neulet.2009.12.012

Heim, R., Cubitt, a B., & Tsien, R. Y. (1995). Improved green fluorescence. Nature.

https://doi.org/10.1038/373663b0

Helfman, P. M., & Bada, J. L. (1975). Aspartic acid racemization in tooth enamel from living

humans. Proceedings of the National Academy of Sciences of the United States of America,

72(8), 2891–4. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/1059082

Helfman, P. M., & Bada, J. L. (1976). Aspartic acid racemisation in dentine as a measure of

ageing. Nature, 262(5566), 279–281. https://doi.org/10.1038/262279b0

Hengst, U., & Jaffrey, S. R. (2007). Function and translational regulation of mRNA in

developing axons. Seminars in Cell and Developmental Biology.

https://doi.org/10.1016/j.semcdb.2007.01.003

Henley, J. M., & Wilkinson, K. A. (2013). AMPA receptor trafficking and the mechanisms

underlying synaptic plasticity and cognitive aging. Dialogues in Clinical Neuroscience,

15(1), 11–27. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/23576886

Higuchi, R., Fockler, C., Dollinger, G., & Watson, R. (1993). Kinetic PCR analysis: real-time

monitoring of DNA amplification reactions. Bio/technology.

https://doi.org/10.1038/nbt0993-1026

Holland, P. M., Abramson, R. D., Watson, R., & Gelfand, D. H. (1991). Detection of specific

polymerase chain reaction product by utilizing the 5’----3’ exonuclease activity of Thermus

aquaticus DNA polymerase. Proceedings of the National Academy of Sciences of the United

States of America, 88(16), 7276–80. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/1871133

Holtkamp, W., Kokic, G., Jäger, M., Mittelstaet, J., Komar, A. A., & Rodnina, M. V. (2015).

Cotranslational protein folding on the ribosome monitored in real time. Science (New York,

N.Y.), 350(6264), 1104–7. https://doi.org/10.1126/science.aad0344

Houdebine, L. M., & Attal, J. (1999). Internal ribosome entry sites (IRESs): reality and use.

Transgenic Research, 8(3), 157–77. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/10478488

Htun, H., Barsony, J., Renyi, I., Gould, D. L., & Hager, G. L. (1996). Visualization of

glucocorticoid receptor translocation and intranuclear organization in living cells with a

green fluorescent protein chimera. Proceedings of the National Academy of Sciences of the

United States of America, 93(10), 4845–50. Retrieved from

Page 160: Observation and quantification of protein production in single living cells

148

http://www.ncbi.nlm.nih.gov/pubmed/8643491

Huang, Y., & Bystroff, C. (2009). Complementation and Reconstitution of Fluorescence from

Circularly Permuted and Truncated Green Fluorescent Protein †. Biochemistry, 48(5), 929–

940. https://doi.org/10.1021/bi802027g

Ibata, K., Sun, Q., & Turrigiano, G. G. (2008). Rapid Synaptic Scaling Induced by Changes in

Postsynaptic Firing. Neuron, 57(6), 819–826. https://doi.org/10.1016/j.neuron.2008.02.031

Iizuka, R., Yamagishi-Shirasaki, M., & Funatsu, T. (2011). Kinetic study of de novo

chromophore maturation of fluorescent proteins. Analytical Biochemistry, 414(2), 173–178.

https://doi.org/10.1016/j.ab.2011.03.036

Ingolia, N. T., Lareau, L. F., & Weissman, J. S. (2011). Ribosome profiling of mouse embryonic

stem cells reveals the complexity and dynamics of mammalian proteomes. Cell, 147(4),

789–802. https://doi.org/10.1016/j.cell.2011.10.002

Inouye, S., & Tsuji, F. I. (1994). Evidence for redox forms of the Aequorea green fluorescent

protein. FEBS Letters, 351(2), 211–214. https://doi.org/10.1016/0014-5793(94)00859-0

Janeway, C. A., Travers, P., Walport, M., & Shlomchik, M. (2001). Immunobiology: The

Immune System In Health And Disease. Immuno Biology 5. https://doi.org/10.1111/j.1467-

2494.1995.tb00120.x

Jaramillo, A. M., Weil, T. T., Goodhouse, J., Gavis, E. R., & Schupbach, T. (2008). The

dynamics of fluorescently labeled endogenous gurken mRNA in Drosophila. Journal of Cell

Science, 121(Pt 6), 887–94. https://doi.org/10.1242/jcs.019091

Jinek, M., Chylinski, K., Fonfara, I., Hauer, M., Doudna, J. A., & Charpentier, E. (2012). A

Programmable Dual-RNA – Guided DNA Endonuclease in Adaptive Bacterial Immunity.

Science, 337(August), 816–822. https://doi.org/10.1126/science.1225829

Johnson, L. N. (2009). The regulation of protein phosphorylation. Biochemical Society

Transactions, 37(Pt 4), 627–41. https://doi.org/10.1042/BST0370627

Ju, W., Morishita, W., Tsui, J., Gaietta, G., Deerinck, T. J., Adams, S. R., … Malenka, R. C.

(2004). Activity-dependent regulation of dendritic synthesis and trafficking of AMPA

receptors. Nature Neuroscience, 7(3), 244–53. https://doi.org/10.1038/nn1189

Jung, H., Yoon, B. C., & Holt, C. E. (2012). Axonal mRNA localization and local protein

synthesis in nervous system assembly, maintenance and repair. Nature Reviews.

Neuroscience, 13(5), 308–24. https://doi.org/10.1038/nrn3210

Kaddoum, L., Magdeleine, E., Waldo, G. S., Joly, E., & Cabantous, S. (2010). One-step split

GFP staining for sensitive protein detection and localization in mammalian cells.

BioTechniques, 49(4), 727–736. https://doi.org/10.2144/000113512

Kandel, E. R. (2001). The Molecular Biology of Memory Storage: A Dialogue between Genes

and Synapses. Science, 294(5544), 1030–1038. https://doi.org/10.1126/science.1067020

Kankaanpää, P., Paavolainen, L., Tiitta, S., Karjalainen, M., Päivärinne, J., Nieminen, J., …

White, D. J. (2012). BioImageXD: an open, general-purpose and high-throughput image-

processing platform. Nature Methods, 9(7), 683–689. https://doi.org/10.1038/nmeth.2047

Kavanagh, I. C., & Baker, S. C. (2009). Advances in Nucleic Acid Detection and Quantification.

Biochemical Society Transactions, 37(2).

Kent, K. P., Childs, W., & Boxer, S. G. (2008). Deconstructing Green Fluorescent Protein.

Journal of the American Chemical Society, 130(30), 9664–9665.

https://doi.org/10.1021/ja803782x

Kerppola, T. K. (2006). Visualization of molecular interactions by fluorescence

complementation. Nature Reviews Molecular Cell Biology, 7(6), 449–456.

Page 161: Observation and quantification of protein production in single living cells

149

https://doi.org/10.1038/nrm1929

Kessels, H. W., & Malinow, R. (2009). Synaptic AMPA receptor plasticity and behavior.

Neuron, 61(3), 340–50. https://doi.org/10.1016/j.neuron.2009.01.015

Kim, J. H., Lee, S.-R., Li, L.-H., Park, H.-J., Park, J.-H., Lee, K. Y., … Choi, S.-Y. (2011). High

cleavage efficiency of a 2A peptide derived from porcine teschovirus-1 in human cell lines,

zebrafish and mice. PloS One, 6(4), e18556. https://doi.org/10.1371/journal.pone.0018556

Kim, J., Zhao, T., Petralia, R. S., Yu, Y., Peng, H., Myers, E., & Magee, J. C. (2011). mGRASP

enables mapping mammalian synaptic connectivity with light microscopy. Nature Methods,

9(1), 96–102. https://doi.org/10.1038/nmeth.1784

Kim, S., Chen, P., & Martin, K. (2010). Visualizing mRNA Trafficking and Local Translation

Within Individual Neurons. Society for Neuroscience, 35. Retrieved from

http://www.sfn.org/siteobjects/published/0000BDF20016F63800FD712C30FA42DD/64CF

FFC781B594C25468FF82134E35AD/file/Short Course Book II.pdf#page=37

King, M. Lou, Messitt, T. J., & Mowry, K. L. (2005). Putting RNAs in the right place at the right

time: RNA localization in the frog oocyte. Biology of the Cell / under the Auspices of the

European Cell Biology Organization, 97, 19–33. https://doi.org/10.1042/BC20040067

Kishikawa, T., Otsuka, M., Ohno, M., Yoshikawa, T., Takata, A., & Koike, K. (2015).

Circulating RNAs as new biomarkers for detecting pancreatic cancer. World Journal of

Gastroenterology, 21(28), 8527–40. https://doi.org/10.3748/wjg.v21.i28.8527

Kopito, R. R. (2000). Aggresomes, inclusion bodies and protein aggregation. Trends in Cell

Biology. https://doi.org/10.1016/S0962-8924(00)01852-3

Kramer, G., Boehringer, D., Ban, N., & Bukau, B. (2009). The ribosome as a platform for co-

translational processing, folding and targeting of newly synthesized proteins. Nature

Structural & Molecular Biology, 16(6), 589–597. https://doi.org/10.1038/nsmb.1614

Krichevsky, A. M., & Kosik, K. S. (2001). Neuronal RNA Granules: A Link between RNA

Localization and Stimulation-Dependent Translation. Neuron, 32(4), 683–696.

https://doi.org/10.1016/S0896-6273(01)00508-6

Krishna Prasad, R. B., Sharma, A., & Babu, H. M. (2013). An insight into salivary markers in

oral cancer. Dental Research Journal, 10(3), 287–95. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/24019794

Kubitscheck, U., Kückmann, O., Kues, T., & Peters, R. (2000). Imaging and Tracking of Single

GFP Molecules in Solution. Biophysical Journal, 78(4), 2170–2179.

https://doi.org/10.1016/S0006-3495(00)76764-6

Kwon, I., Lee, J., Chang, S. H., Jung, N. C., Lee, B. J., Son, G. H., … Lee, K. H. (2006).

BMAL1 shuttling controls transactivation and degradation of the CLOCK/BMAL1

heterodimer. Molecular and Cellular Biology, 26(19), 7318–30.

https://doi.org/10.1128/MCB.00337-06

Larionov, A., Krause, A., & Miller, W. (2005). A standard curve based method for relative real

time PCR data processing. BMC Bioinformatics, 6, 62. https://doi.org/10.1186/1471-2105-

6-62

Lasko, P. (2012). mRNA localization and translational control in Drosophila oogenesis. Cold

Spring Harbor Perspectives in Biology, 4(10). https://doi.org/10.1101/cshperspect.a012294

Leandro, M. J., Weill, J., Weller, S., Reynaud, C., Lammers, A., Porto, A. de, … Nicoletti, A.

(2013). B-cell subpopulations in humans and their differential susceptibility to depletion

with anti-CD20 monoclonal antibodies. Arthritis Research & Therapy, 15(Suppl 1), S3.

https://doi.org/10.1186/ar3908

Page 162: Observation and quantification of protein production in single living cells

150

Lecuyer, E., Yoshida, H., Parthasarathy, N., Alm, C., Babak, T., Cerovina, T., … Krause, H. M.

(2007). Global Analysis of mRNA Localization Reveals a Prominent Role in Organizing

Cellular Architecture and Function. Cell, 131(1), 174–187.

https://doi.org/10.1016/j.cell.2007.08.003

Lein, E. S., Hawrylycz, M. J., Ao, N., Ayres, M., Bensinger, A., Bernard, A., … Jones, A. R.

(2007). Genome-wide atlas of gene expression in the adult mouse brain. Nature, 445(7124),

168–176. https://doi.org/10.1038/nature05453

Levinthal, C. (1969). How to fold graciously. Mössbauer Spectroscopy in Biological Systems

Proceedings, 24(41), 22–24. https://doi.org/citeulike-article-id:380320

Li, F., Vijayasankaran, N., Shen, A. (Yijuan), Kiss, R., & Amanullah, A. (2010). Cell culture

processes for monoclonal antibody production. mAbs, 2(5), 466–479.

https://doi.org/10.4161/mabs.2.5.12720

Li, G.-W., & Xie, X. S. (2011). Central dogma at the single-molecule level in living cells.

Nature, 475(7356), 308–15. https://doi.org/10.1038/nature10315

Li, J. J., Bickel, P. J., & Biggin, M. D. (2014). System wide analyses have underestimated

protein abundances and the importance of transcription in mammals. PeerJ, 2, e270.

https://doi.org/10.7717/peerj.270

Li, J. Z., Bunney, B. G., Meng, F., Hagenauer, M. H., Walsh, D. M., Vawter, M. P., … Bunney,

W. E. (2013). Circadian patterns of gene expression in the human brain and disruption in

major depressive disorder. Proceedings of the National Academy of Sciences of the United

States of America, 110(24), 9950–5. https://doi.org/10.1073/pnas.1305814110

Li, X., Zhao, X., Fang, Y., Jiang, X., Duong, T., Fan, C., … Kain, S. R. (1998). Generation of

destabilized green fluorescent protein as a transcription reporter. Journal of Biological

Chemistry, 273(52), 34970–34975. https://doi.org/10.1074/jbc.273.52.34970

Liang, X., Potter, J., Kumar, S., Zou, Y., Quintanilla, R., Sridharan, M., … Chesnut, J. D. (2015).

Rapid and highly efficient mammalian cell engineering via Cas9 protein transfection.

Journal of Biotechnology, 208, 44–53. https://doi.org/10.1016/j.jbiotec.2015.04.024

Lin, A. C., & Holt, C. E. (2008). Function and regulation of local axonal translation. Current

Opinion in Neurobiology. https://doi.org/10.1016/j.conb.2008.05.004

Ling, D., Salvaterra, P. M., White, K., Salvaterra, P., & Lewis, J. (2011). Robust RT-qPCR Data

Normalization: Validation and Selection of Internal Reference Genes during Post-

Experimental Data Analysis. PLoS ONE, 6(3), e17762.

https://doi.org/10.1371/journal.pone.0017762

Lo, C.-A., Kays, I., Emran, F., Lin, T.-J., Cvetkovska, V., & Chen, B. E. (2015). Quantification

of Protein Levels in Single Living Cells. Cell Reports, 13(11), 2634–2644.

https://doi.org/10.1016/j.celrep.2015.11.048

Lo, K. M., & Gillies, S. D. (1991). High level expression of human proteins in murine

hybridoma cells: induction by methotrexate in the absence of gene amplification.

Biochimica et Biophysica Acta, 1088(2), 217–24. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/1672095

Lodish, H., Berk, A., Kaiser, C. A., Matsudaira, P., Baltimore, D., Bretscher, A., … Matsudaira,

P. (2008). Molecular Cell Biology. Book (Vol. 5). https://doi.org/10.1016/S1470-

8175(01)00023-6

Luby-Phelps, K. (2000). Cytoarchitecture and physical properties of cytoplasm: volume,

viscosity, diffusion, intracellular surface area. International Review of Cytology, 192, 189–

221. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10553280

Page 163: Observation and quantification of protein production in single living cells

151

Ly, T., Ahmad, Y., Shlien, A., Soroka, D., Mills, A., Emanuele, M. J., … Lamond, A. I. (2014).

A proteomic chronology of gene expression through the cell cycle in human myeloid

leukemia cells. eLife, 3, e01630. https://doi.org/10.7554/elife.01630

Lyles, V., Zhao, Y., & Martin, K. C. (2006). Synapse formation and mRNA localization in

cultured Aplysia neurons. Neuron, 49(3), 349–356.

https://doi.org/10.1016/j.neuron.2005.12.029

MacDougall, N., Clark, A., MacDougall, E., & Davis, I. (2003). Drosophila gurken (TGFα)

mRNA localizes as particles that move within the oocyte in two dynein-dependent steps.

Developmental Cell, 4(3), 307–319. https://doi.org/10.1016/S1534-5807(03)00058-3

Maier, T., Güell, M., & Serrano, L. (2009). Correlation of mRNA and protein in complex

biological samples. FEBS Letters, 583(24), 3966–73.

https://doi.org/10.1016/j.febslet.2009.10.036

Mane, V. P., Heuer, M. A., Hillyer, P., Navarro, M. B., & Rabin, R. L. (2008). Systematic

method for determining an ideal housekeeping gene for real-time PCR analysis. Journal of

Biomolecular Techniques : JBT, 19(5), 342–7.

Mao, Z., Bozzella, M., Seluanov, A., & Gorbunova, V. (2008). DNA repair by nonhomologous

end joining and homologous recombination during cell cycle in human cells. Cell Cycle

(Georgetown, Tex.), 7(18), 2902–6. https://doi.org/10.2964/jsik.kuni0223

Marshall, J., Molloy, R., Moss, G. W., Howe, J. R., & Hughes, T. E. (1995). The jellyfish green

fluorescent protein: a new tool for studying ion channel expression and function. Neuron,

14(2), 211–5. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/7531985

Martin, K. C., & Ephrussi, A. (2010). mRNA Localization : Gene Expression in the Spatial

Dimension. Cell, 136(4), 1–21. https://doi.org/10.1016/j.cell.2009.01.044.mRNA

Martinez, J. C., Pisabarro, M. T., & Serrano, L. (1998). Obligatory steps in protein folding and

the conformational diversity of the transition state. Nature Structural Biology, 5(8), 721–

729. https://doi.org/10.1038/1418

Marx, V. (2013). Finding the right antibody for the job. Nature Methods, 10(8), 703–707.

https://doi.org/10.1038/nmeth.2570

Masters, P. M., Bada, J. L., & Zigler Jr., J. S. (1977). Aspartic acid racemisation in the human

lens during ageing and in cataract formation. Nature, 268(5615), 71–73.

https://doi.org/10.1038/268071a0

Mavrakis, M., Rikhy, R., Lilly, M., Lippincott-Schwartz, J., Mavrakis, M., Rikhy, R., …

Lippincott‐ Schwartz, J. (2008). Fluorescence Imaging Techniques for Studying

Drosophila Embryo Development. In Current Protocols in Cell Biology (p. 4.18.1-4.18.43).

Hoboken, NJ, USA: John Wiley & Sons, Inc.

https://doi.org/10.1002/0471143030.cb0418s39

Mikaelian, I., Scicchitano, M., Mendes, O., Thomas, R. A., & Leroy, B. E. (2013). Frontiers in

preclinical safety biomarkers: microRNAs and messenger RNAs. Toxicologic Pathology,

41(1), 18–31. https://doi.org/10.1177/0192623312448939

Miyashiro, K., Dichter, M., & Eberwine, J. (1994). On the nature and differential distribution of

mRNAs in hippocampal neurites: implications for neuronal functioning. Proceedings of the

National Academy of Sciences of the United States of America, 91(23), 10800–4.

https://doi.org/10.1073/pnas.91.23.10800

Mizuguchi, H., Xu, Z., Ishii-Watabe, A., Uchida, E., & Hayakawa, T. (2000). IRES-dependent

second gene expression is significantly lower than cap-dependent first gene expression in a

bicistronic vector. Molecular Therapy : The Journal of the American Society of Gene

Page 164: Observation and quantification of protein production in single living cells

152

Therapy, 1(4), 376–82. https://doi.org/10.1006/mthe.2000.0050

Moccia, R., Chen, D., Lyles, V., Kapuya, E., E, Y., Kalachikov, S., … Martin, K. C. (2003). An

unbiased cDNA library prepared from isolated Aplysia sensory neuron processes is

enriched for cytoskeletal and translational mRNAs. The Journal of Neuroscience : The

Official Journal of the Society for Neuroscience, 23(28), 9409–9417.

https://doi.org/23/28/9409 [pii]

Morse, D., & Tannous, B. A. (2012). A water-soluble coelenterazine for sensitive in vivo

imaging of coelenterate luciferases. Molecular Therapy : The Journal of the American

Society of Gene Therapy, 20(4), 692–3. https://doi.org/10.1038/mt.2012.38

Mourelatos, Z., Abel, L., Yong, J., Kataoka, N., & Dreyfuss, G. (2001). SMN interacts with a

novel family of hnRNP and spliceosomal proteins. The EMBO Journal, 20(19), 5443–52.

https://doi.org/10.1093/emboj/20.19.5443

Muddashetty, R. S., Kelić, S., Gross, C., Xu, M., & Bassell, G. J. (2007). Dysregulated

metabotropic glutamate receptor-dependent translation of AMPA receptor and postsynaptic

density-95 mRNAs at synapses in a mouse model of fragile X syndrome. The Journal of

Neuroscience : The Official Journal of the Society for Neuroscience, 27(20), 5338–5348.

https://doi.org/10.1523/JNEUROSCI.0937-07.2007

Mullis, K., Faloona, F., Scharf, S., Saiki, R., Horn, G., & Erlich, H. (1986). Specific enzymatic

amplification of DNA in vitro: The polymerase chain reaction. Cold Spring Harbor

Symposia on Quantitative Biology, 51(1), 263–273.

https://doi.org/10.1101/SQB.1986.051.01.032

Na, Y., Park, S., Lee, C., Kim, D., Park, J. M., Sockanathan, S., … Worley, P. F. (2016). Real-

Time Imaging Reveals Properties of Glutamate-Induced Arc/Arg 3.1 Translation in

Neuronal Dendrites. Neuron, 91(3), 561–573. https://doi.org/10.1016/j.neuron.2016.06.017

Nagai, T., Ibata, K., Park, E. S., Kubota, M., Mikoshiba, K., & Miyawaki, A. (2002). A variant

of yellow fluorescent protein with fast and efficient maturation for cell-biological

applications. Nature Biotechnology, 20(1), 87–90. https://doi.org/10.1038/nbt0102-87

Nair-Gill, E. D., Shu, C. J., Radu, C. G., & Witte, O. N. (2008). Non-invasive imaging of

adaptive immunity using positron emission tomography. Immunological Reviews, 221, 214–

28. https://doi.org/10.1111/j.1600-065X.2008.00585.x

Narahashi, T. (1988). Ion Channels : Volume 1. Springer US.

Nelson, R., Sawaya, M. R., Balbirnie, M., Madsen, A. Ø., Riekel, C., Grothe, R., & Eisenberg,

D. (2005). Structure of the cross-β spine of amyloid-like fibrils. Nature Cell Biology,

435(7043), 773–778. https://doi.org/10.1038/nature03680

Nilson, L. A., & Schüpbach, T. (1999). EGF receptor signaling in Drosophila oogenesis. Current

Topics in Developmental Biology, 44, 203–43. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/9891881

Nimchinsky, E. A., Sabatini, B. L., & Svoboda, K. (2002). Structure and function of dendritic

spines. Annual Review of Physiology, 64, 313–53.

https://doi.org/10.1146/annurev.physiol.64.081501.160008

Nolan, T., Hands, R. E., & Bustin, S. A. (2006). Quantification of mRNA using real-time RT-

PCR. Nature Protocols, 1(3), 1559–82. https://doi.org/10.1038/nprot.2006.236

Nutt, S. L., Hodgkin, P. D., Tarlinton, D. M., & Corcoran, L. M. (2015). The generation of

antibody-secreting plasma cells. Nature Reviews. Immunology, 15(3), 160–171.

https://doi.org/10.1038/nri3795

Ochs, S., Sabri, M. I., & Johnson, J. (1969). Fast transport system of materials in mammalian

Page 165: Observation and quantification of protein production in single living cells

153

nerve fibers. Science (New York, N.Y.), 163(3868), 686–7. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/5762934

Opel, K. L., Chung, D., & McCord, B. R. (2010). A Study of PCR Inhibition Mechanisms Using

Real Time PCR. Journal of Forensic Sciences, 55(1), 25–33. https://doi.org/10.1111/j.1556-

4029.2009.01245.x

Orci, L., Ravazzola, M., Amherdt, M., Madsen, O., Perrelet, A., Vassalli, J. D., & Anderson, R.

G. (1986). Conversion of proinsulin to insulin occurs coordinately with acidification of

maturing secretory vesicles. The Journal of Cell Biology, 103(6 Pt 1), 2273–81.

https://doi.org/10.1083/JCB.103.6.2273

Ormö, M., Cubitt, A. B., Kallio, K., Gross, L. A., Tsien, R. Y., Remington, S. J., … YANG, F.

(1996). Crystal structure of the Aequorea victoria green fluorescent protein. Science (New

York, N.Y.), 273(5280), 1392–5. https://doi.org/10.1126/science.273.5280.1392

Ozsolak, F., & Milos, P. M. (2011). RNA sequencing: advances, challenges and opportunities.

Nature Reviews. Genetics, 12(2), 87–98. https://doi.org/10.1038/nrg2934

Palmer, E., & Freeman, T. (2004). Investigation into the use of C- and N-terminal GFP fusion

proteins for subcellular localization studies using reverse transfection microarrays.

Comparative and Functional Genomics, 5(4), 342–353. https://doi.org/10.1002/cfg.405

Panda, S., Antoch, M. P., Miller, B. H., Su, A. I., Schook, A. B., Straume, M., … Hogenesch, J.

B. (2002). Coordinated transcription of key pathways in the mouse by the circadian clock.

Cell, 109(3), 307–20. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/12015981

Paquin, N., & Chartrand, P. (2008). Local regulation of mRNA translation: new insights from the

bud. Trends in Cell Biology. https://doi.org/10.1016/j.tcb.2007.12.004

Parton, R. M., Vallés, A. M., Dobbie, I. M., & Davis, I. (2010). Collection and mounting of

Drosophila embryos for imaging. Cold Spring Harbor Protocols, 2010(4), pdb.prot5403.

https://doi.org/10.1101/pdb.prot5403

Patterson, G. H., Knobel, S. M., Sharif, W. D., Kain, S. R., & Piston, D. W. (1997). Use of the

green fluorescent protein and its mutants in quantitative fluorescence microscopy.

Biophysical Journal, 73(5), 2782–90. https://doi.org/10.1016/S0006-3495(97)78307-3

Pauwels, K., & Tompa, P. (2016). Editorial: Function and Flexibility: Friend or Foe? Frontiers

in Molecular Biosciences, 3. https://doi.org/10.3389/fmolb.2016.00031

Pédelacq, J.-D., Cabantous, S., Tran, T., Terwilliger, T. C., & Waldo, G. S. (2006). Engineering

and characterization of a superfolder green fluorescent protein. Nature Biotechnology,

24(1), 79–88. https://doi.org/10.1038/nbt1172

Pelletier, J., & Sonenberg, N. (1988). Internal initiation of translation of eukaryotic mRNA

directed by a sequence derived from poliovirus RNA. Nature, 334(6180), 320–325.

https://doi.org/10.1038/334320a0

Piana, S., Lindorff-Larsen, K., & Shaw, D. E. (2013). Atomic-level description of ubiquitin

folding. Proceedings of the National Academy of Sciences of the United States of America,

110(15), 5915–20. https://doi.org/10.1073/pnas.1218321110

Pinaud, F., & Dahan, M. (2011). Targeting and imaging single biomolecules in living cells by

complementation-activated light microscopy with split-fluorescent proteins. Proc Natl Acad

Sci U S A, 108(24), E201-10. https://doi.org/10.1073/pnas.1101929108

Poddar, D., Basu, A., Baldwin, W. M., Kondratov, R. V, Barik, S., & Mazumder, B. (2013). An

extraribosomal function of ribosomal protein L13a in macrophages resolves inflammation.

Journal of Immunology (Baltimore, Md. : 1950), 190(7), 3600–12.

https://doi.org/10.4049/jimmunol.1201933

Page 166: Observation and quantification of protein production in single living cells

154

Poon, M. M., Choi, S.-H., Jamieson, C. a M., Geschwind, D. H., & Martin, K. C. (2006).

Identification of process-localized mRNAs from cultured rodent hippocampal neurons. The

Journal of Neuroscience : The Official Journal of the Society for Neuroscience, 26(51),

13390–13399. https://doi.org/10.1523/JNEUROSCI.3432-06.2006

Prado, M., Eickel, N., De Niz, M., Heitmann, A., Agop-Nersesian, C., Wacker, R., … Heussler,

V. T. (2015). Long-term live imaging reveals cytosolic immune responses of host

hepatocytes against Plasmodium infection and parasite escape mechanisms. Autophagy,

11(9), 1561–79. https://doi.org/10.1080/15548627.2015.1067361

Prasher, D. C., Eckenrode, V. K., Ward, W. W., Prendergast, F. G., & Cormier, M. J. (1992).

Primary structure of the Aequorea victoria green-fluorescent protein. Gene, 111(2), 229–

233. https://doi.org/10.1016/0378-1119(92)90691-H

Prelich, G. (2012). Gene overexpression: uses, mechanisms, and interpretation. Genetics, 190(3),

841–54. https://doi.org/10.1534/genetics.111.136911

Price, J. C., Guan, S., Burlingame, A., Prusiner, S. B., & Ghaemmaghami, S. (2010). Analysis of

proteome dynamics in the mouse brain. Proceedings of the National Academy of Sciences of

the United States of America, 107(32), 14508–13. https://doi.org/10.1073/pnas.1006551107

Qin, J. Y., Zhang, L., Clift, K. L., Hulur, I., Xiang, A. P., Ren, B.-Z., & Lahn, B. T. (2010).

Systematic Comparison of Constitutive Promoters and the Doxycycline-Inducible Promoter.

PLoS ONE, 5(5), e10611. https://doi.org/10.1371/journal.pone.0010611

Quitschke, W. W., Lin, Z. Y., DePonti-Zilli, L., & Paterson, B. M. (1989). The beta actin

promoter. High levels of transcription depend upon a CCAAT binding factor. The Journal

of Biological Chemistry, 264(16), 9539–46. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/2722849

Radcliffe, P. A., & Mitrophanous, K. A. (2004). Multiple gene products from a single vector:

“self-cleaving” 2A peptides. Gene Therapy, 11(23), 1673–1674.

https://doi.org/10.1038/sj.gt.3302361

Radzicka, A., & Wolfenden, R. (1995). A proficient enzyme. Science, 267(5194).

Rao, A. K., Garver, F., & Mendicino, J. (1976). Biosynthesis of the carbohydrate units of

immunoglobulins. 1. Purification and properties of galactosyltransferases from swine

mesentary lymph nodes. Biochemistry, 15(23), 5001–5009.

Reid, B. G., & Flynn, G. C. (1997). Chromophore formation in green fluorescent protein.

Biochemistry, 36(22), 6786–6791. https://doi.org/10.1021/bi970281w

Remington, S. J. (2011). Green fluorescent protein: a perspective. Protein Science : A

Publication of the Protein Society, 20(9), 1509–19. https://doi.org/10.1002/pro.684

Richter, J. D., & Lasko, P. (2011). Translational control in oocyte development. Cold Spring

Harbor Perspectives in Biology. https://doi.org/10.1101/cshperspect.a002758

Rizzuto, R., Brini, M., Pizzo, P., Murgia, M., & Pozzan, T. (1995). Chimeric green fluorescent

protein as a tool for visualizing subcellular organelles in living cells. Current Biology : CB,

5(6), 635–42. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/7552174

Roberts, L. A., Large, C. H., Higgins, M. J., Stone, T. W., O’Shaughnessy, C. T., & Morris, B. J.

(1998). Increased expression of dendritic mRNA following the induction of long-term

potentiation. Molecular Brain Research, 56(1–2), 38–44. https://doi.org/10.1016/S0169-

328X(98)00026-6

Rørth, P. (1998). Gal4 in the Drosophila female germline. Mechanisms of Development, 78(1–2),

113–118. https://doi.org/10.1016/S0925-4773(98)00157-9

Ross, J. F., & Orlowski, M. (1982). Growth-rate-dependent adjustment of ribosome function in

Page 167: Observation and quantification of protein production in single living cells

155

chemostat-grown cells of the fungus Mucor racemosus. Journal of Bacteriology, 149(2),

650–3. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/6799491

Sage, A. T., Besant, J. D., Mahmoudian, L., Poudineh, M., Bai, X., Zamel, R., … Kelley, S. O.

(2015). Fractal circuit sensors enable rapid quantification of biomarkers for donor lung

assessment for transplantation. Science Advances, 1(7).

Saiki, R. K., Scharf, S., Faloona, F., Mullis, K. B., Horn, G. T., Erlich, H. a, & Arnheim, N.

(1985). Enzymatic amplification of beta-globin genomic sequences and restriction site

analysis for diagnosis of sickle cell anemia. Science (New York, N.Y.), 230(4732), 1350–

1354. https://doi.org/10.1126/science.2999980

Santangelo, P. J., Rogers, K. a, Zurla, C., Blanchard, E. L., Gumber, S., Strait, K., … Villinger,

F. (2015). Whole-body immunoPET reveals active SIV dynamics in viremic and

antiretroviral therapy-treated macaques. Nature Methods, 12(5), 427–32.

https://doi.org/10.1038/nmeth.3320

Satija, R., Shalek, A. K., Raj, A., al., et, Germain, R. N., Littman, D. R., … al., et. (2014).

Heterogeneity in immune responses: from populations to single cells. Trends in

Immunology, 35(5), 219–29. https://doi.org/10.1016/j.it.2014.03.004

Saunders, C., & Cohen, R. S. (1999). The role of oocyte transcription, the 5’UTR, and translation

repression and derepression in Drosophila gurken mRNA and protein localization.

Molecular Cell, 3(1), 43–54. https://doi.org/10.1016/S1097-2765(00)80173-2

Schubert, C. (2011). Single-cell analysis: The deepest differences. Nature, 480(7375), 133–137.

https://doi.org/10.1038/480133a

Schwanhäusser, B., Busse, D., Li, N., Dittmar, G., Schuchhardt, J., Wolf, J., … Selbach, M.

(2011). Global quantification of mammalian gene expression control. Nature, 473(7347),

337–342. https://doi.org/10.1038/nature10098

Scott, K. P., Mercer, D. K., Glover, L. A., & Flint, H. J. (1998). The green fluorescent protein as

a visible marker for lactic acid bacteria in complex ecosystems. FEMS Microbiology

Ecology, 26(3), 219–230. https://doi.org/10.1016/S0168-6496(98)00037-3

Scott, M. S., Boisvert, F. M., McDowall, M. D., Lamond, A. I., & Barton, G. J. (2010).

Characterization and prediction of protein nucleolar localization sequences. Nucleic Acids

Research, 38(21), 7388–7399. https://doi.org/10.1093/nar/gkq653

Shain, A. H., & Pollack, J. R. (2013). The spectrum of SWI/SNF mutations, ubiquitous in human

cancers. PloS One, 8(1), e55119. https://doi.org/10.1371/journal.pone.0055119

Shaner, N. C., Lin, M. Z., McKeown, M. R., Steinbach, P. A., Hazelwood, K. L., Davidson, M.

W., & Tsien, R. Y. (2008). Improving the photostability of bright monomeric orange and

red fluorescent proteins. Nature Methods, 5(6), 545–51. https://doi.org/10.1038/nmeth.1209

Shaner, N. C., Steinbach, P. A., & Tsien, R. Y. (2005). A guide to choosing fluorescent proteins.

Nature Methods, 2(12), 905–909. https://doi.org/10.1038/nmeth819

Smalle, J., & Vierstra, R. D. (2004). The ubiquitin 26S proteasome proteolytic pathway. Annual

Review of Plant Biology, 55, 555–90.

https://doi.org/10.1146/annurev.arplant.55.031903.141801

Smith, W. B., Starck, S. R., Roberts, R. W., & Schuman, E. M. (2005). Dopaminergic

stimulation of local protein synthesis enhances surface expression of GluR1 and synaptic

transmission in hippocampal neurons. Neuron, 45(5), 765–79.

https://doi.org/10.1016/j.neuron.2005.01.015

Snapp, E. (2005). Design and use of fluorescent fusion proteins in cell biology. Current

Protocols in Cell Biology / Editorial Board, Juan S. Bonifacino ... [et Al.], Chapter 21, Unit

Page 168: Observation and quantification of protein production in single living cells

156

21.4. https://doi.org/10.1002/0471143030.cb2104s27

Stahlberg, A., & Bengtsson, M. (2010). Single-cell gene expression profiling using reverse

transcription quantitative real-time PCR. Methods, 50(4), 282–288.

https://doi.org/10.1016/j.ymeth.2010.01.002

Ståhlberg, A., & Kubista, M. (2014). The workflow of single-cell expression profiling using

quantitative real-time PCR. Expert Review of Molecular Diagnostics, 14(3), 323–331.

https://doi.org/10.1586/14737159.2014.901154

Stemmer, W. P. (1994). DNA shuffling by random fragmentation and reassembly: in vitro

recombination for molecular evolution. Proceedings of the National Academy of Sciences of

the United States of America, 91(22), 10747–10751.

https://doi.org/10.1073/pnas.91.22.10747

Steward, O., Wallace, C. S., Lyford, G. L., & Worley, P. F. (1998). Synaptic activation causes

the mRNA for the leg Arc to localize selectively near activated postsynaptic sites on

dendrites. Neuron, 21(4), 741–751. https://doi.org/10.1016/S0896-6273(00)80591-7

Stockholm, D., Benchaouir, R., Picot, J., Rameau, P., Neildez, T. M. A., Landini, G., … Paldi,

A. (2007). The Origin of Phenotypic Heterogeneity in a Clonal Cell Population In Vitro.

PLoS ONE, 2(4), e394. https://doi.org/10.1371/journal.pone.0000394

Sullivan, D. C., & Kuntz, I. D. (2002). Protein Folding as Biased Conformational Diffusion. The

Journal of Physical Chemistry B, 106(12), 3255–3262. https://doi.org/10.1021/jp012911g

Sullivan, L. H. (1896). The tall office building artistically considered. Lippincott’s Magazine.

Retrieved from http://ocw.mit.edu/courses/architecture/4-205-analysis-of-contemporary-

architecture-fall-2009/readings/MIT4_205F09_Sullivan.pdf

Suslov, O., & Steindler, D. A. (2005). PCR inhibition by reverse transcriptase leads to an

overestimation of amplification efficiency. Nucleic Acids Research, 33(20), e181.

https://doi.org/10.1093/nar/gni176

Svec, D., Andersson, D., Pekny, M., Sjöback, R., Kubista, M., & Ståhlberg, A. (2013). Direct

cell lysis for single-cell gene expression profiling. Frontiers in Oncology, 3(November),

274. https://doi.org/10.3389/fonc.2013.00274

Swanger, S. A., & Bassell, G. J. (2011). Making and breaking synapses through local mRNA

regulation. Current Opinion in Genetics & Development, 21(4), 414–21.

https://doi.org/10.1016/j.gde.2011.04.002

Swanger, S. A., & Bassell, G. J. (2013). Dendritic protein synthesis in the normal and diseased

brain. Neuroscience, 232(404), 106–127.

https://doi.org/10.1016/j.neuroscience.2012.12.003

Szymczak, A. L., Workman, C. J., Wang, Y., Vignali, K. M., Dilioglou, S., Vanin, E. F., &

Vignali, D. a a. (2004). Correction of multi-gene deficiency in vivo using a single “self-

cleaving” 2A peptide-based retroviral vector. Nature Biotechnology, 22(5), 589–94.

https://doi.org/10.1038/nbt957

Szymczak-Workman, A. L., Vignali, K. M., & Vignali, D. A. A. (2012). Design and

Construction of 2A Peptide-Linked Multicistronic Vectors. Cold Spring Harbor Protocols,

2012(2), pdb.ip067876-ip067876. https://doi.org/10.1101/pdb.ip067876

Tatavarty, V., Ifrim, M. F., Levin, M., Korza, G., Barbarese, E., Yu, J., & Carson, J. H. (2012).

Single-molecule imaging of translational output from individual RNA granules in neurons.

Molecular Biology of the Cell, 23(5), 918–29. https://doi.org/10.1091/mbc.E11-07-0622

Taylor, S., Wakem, M., Dijkman, G., Alsarraj, M., & Nguyen, M. (2010). A practical approach

to RT-qPCR—Publishing data that conform to the MIQE guidelines. Methods, 50(4), S1–

Page 169: Observation and quantification of protein production in single living cells

157

S5. https://doi.org/10.1016/j.ymeth.2010.01.005

Tester, M. (1997). Techniques for studying ion channels: an introduction. Journal of

Experimental Botany, 48 Spec No, 353–9. https://doi.org/10.1093/jxb/48.Special_Issue.353

Tian, Q. B., Nakayama, K., Okano, A., & Suzuki, T. (1999). Identification of mRNAs localizing

in the postsynaptic region. Molecular Brain Research, 72(2), 147–157.

https://doi.org/10.1016/S0169-328X(99)00214-4

Tichopad, A., Kitchen, R., Riedmaier, I., Becker, C., Ståhlberg, A., & Kubista, M. (2009).

Design and Optimization of Reverse-Transcription Quantitative PCR Experiments. Clinical

Chemistry, 55(10).

Tobias, G. S., & Koenig, E. (1975). Influence of nerve cell body and neurolemma cell on local

axonal protein synthesis following neurotomy. Experimental Neurology, 49(1 Pt 1), 235–45.

Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/52557

Toyama, B. H., & Hetzer, M. W. (2013). Protein homeostasis: live long, won’t prosper. Nature

Reviews. Molecular Cell Biology, 14(1), 55–61. https://doi.org/10.1038/nrm3496

Tsien, R. Y. (1998). The green fluorescent protein. Annual Review of Biochemistry, 67, 509–44.

https://doi.org/10.1146/annurev.biochem.67.1.509

Van De Bor, V., Hartswood, E., Jones, C., Finnegan, D., & Davis, I. (2005). gurken and the I

factor retrotransposon RNAs share common localization signals and machinery.

Developmental Cell, 9(1), 51–62. https://doi.org/10.1016/j.devcel.2005.04.012

van de Corput, M. P., Dirks, R. W., van Gijlswijk, R. P., van de Rijke, F. M., & Raap, A. K.

(1998). Fluorescence in situ hybridization using horseradish peroxidase-labeled

oligodeoxynucleotides and tyramide signal amplification for sensitive DNA and mRNA

detection. Histochemistry and Cell Biology, 110(4), 431–7. Retrieved from

http://www.ncbi.nlm.nih.gov/pubmed/9792422

Vogel, C., & Marcotte, E. M. (2012). Insights into the regulation of protein abundance from

proteomic and transcriptomic analyses. Nature Reviews. Genetics, 13(4), 227–32.

https://doi.org/10.1038/nrg3185

Waldo, G. S., Standish, B. M., Berendzen, J., & Terwilliger, T. C. (1999). Rapid protein-folding

assay using green fluorescent protein. Nature Biotechnology, 17(7), 691–695.

https://doi.org/10.1038/10904

Walker, J. M. (2009). The Protein Protocols Handbook. Humana Press.

https://doi.org/10.1007/978-1-59745-198-7_26

Wang, D. O., Kim, S. M., Zhao, Y., Hwang, H., Miura, S. K., Sossin, W. S., & Martin, K. C.

(2009). Synapse- and Stimulus-Specific Local Translation During Long-Term Neuronal

Plasticity. Science, 324(5934), 1536–1540. https://doi.org/10.1126/science.1173205

Wang, D. O., Kim, S. M., Zhao, Y., Hwang, H., Miura, S. K., Sossin, W. S., & Martin, K. C.

(2009). Synapse- and stimulus-specific local translation during long-term neuronal

plasticity. Science (New York, N.Y.), 324(5934), 1536–1540.

https://doi.org/10.1126/science.1173205

Wang, D. O., Martin, K. C., & Zukin, R. S. (2010). Spatially restricting gene expression by local

translation at synapses. Trends in Neurosciences, 33(4), 173–182.

https://doi.org/10.1016/j.tins.2010.01.005

Wang, H.-W., Willis, J., Sivak, M. V., & Izatt, J. A. (1998). Correlation of autofluorescence to

proliferation and inflammatory indices in human premalignant colonic tissues. In

Biomedical Optical Spectroscopy and Diagnostics / Therapeutic Laser Applications (p.

BTuA4). Washington, D.C.: OSA. https://doi.org/10.1364/BOSD.1998.BTuA4

Page 170: Observation and quantification of protein production in single living cells

158

Ward, W. W., Cody, C. W., Hart, R. C., & Cormier, M. J. (1980). Spectrophotometric Identity of

the Energy Transfer Chromophores in Renilla and Aequorea Green-Fluorescent Proteins.

Photochemistry and Photobiology, 31(6), 611–615. https://doi.org/10.1111/j.1751-

1097.1980.tb03755.x

Willard, S. S., & Koochekpour, S. (2013). Glutamate, glutamate receptors, and downstream

signaling pathways. International Journal of Biological Sciences, 9(9), 948–59.

https://doi.org/10.7150/ijbs.6426

Willis, D. E., & Twiss, J. L. (2006). The evolving roles of axonally synthesized proteins in

regeneration. Current Opinion in Neurobiology. https://doi.org/10.1016/j.conb.2006.01.002

Wong, I. H., Yeo, W., Chan, A. T., & Johnson, P. J. (2001). Quantitative correlation of

cytokeratin 19 mRNA level in peripheral blood with disease stage and metastasis in breast

cancer patients: potential prognostic implications. International Journal of Oncology, 18(3),

633–8. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/11179498

Wu, M., & Singh, A. K. (2012a). Single-cell protein analysis. Current Opinion in Biotechnology,

23(1), 83–88. https://doi.org/10.1016/j.copbio.2011.11.023

Wu, M., & Singh, A. K. (2012b). Single-cell protein analysis. Current Opinion in Biotechnology,

23(1), 83–8. https://doi.org/10.1016/j.copbio.2011.11.023

Yasukawa, K., Konishi, A., & Inouye, K. (2010). Effects of organic solvents on the reverse

transcription reaction catalyzed by reverse transcriptases from avian myeloblastosis virus

and Moloney murine leukemia virus. Bioscience, Biotechnology, and Biochemistry, 74(9),

1925–30. https://doi.org/10.1271/bbb.100337

Yu, S. P., & Kerchner, G. A. (1998). Endogenous voltage-gated potassium channels in human

embryonic kidney (HEK293) cells. Journal of Neuroscience Research, 52(5), 612–7.

https://doi.org/10.1002/(SICI)1097-4547(19980601)52:5<612::AID-JNR13>3.0.CO;2-3

[pii]

Zhang, R., Lahens, N. F., Ballance, H. I., Hughes, M. E., & Hogenesch, J. B. (2014). A circadian

gene expression atlas in mammals: implications for biology and medicine. Proceedings of

the National Academy of Sciences of the United States of America, 111(45), 16219–24.

https://doi.org/10.1073/pnas.1408886111

Zhao, H. L., Yao, X. Q., Xue, C., Wang, Y., Xiong, X. H., & Liu, Z. M. (2008). Increasing the

homogeneity, stability and activity of human serum albumin and interferon-alpha2b fusion

protein by linker engineering. Protein Expression and Purification, 61(1), 73–7.

https://doi.org/10.1016/j.pep.2008.04.013

Zhao, S., & Fernald, R. D. (2005). Comprehensive algorithm for quantitative real-time

polymerase chain reaction. Journal of Computational Biology : A Journal of Computational

Molecular Cell Biology, 12(8), 1047–64. https://doi.org/10.1089/cmb.2005.12.1047

Zheng, J. Q., Kelly, T. K., Chang, B., Ryazantsev, S., Rajasekaran, a K., Martin, K. C., & Twiss,

J. L. (2001). A functional role for intra-axonal protein synthesis during axonal regeneration

from adult sensory neurons. The Journal of Neuroscience : The Official Journal of the

Society for Neuroscience, 21(23), 9291–9303. https://doi.org/21/23/9291 [pii]

Zhou, X., Liao, W. J., Liao, J. M., Liao, P., & Lu, H. (2015). Ribosomal proteins: Functions

beyond the ribosome. Journal of Molecular Cell Biology, 7(2), 92–104.

https://doi.org/10.1093/jmcb/mjv014

Zimmer, M. (2002). Green Fluorescent Protein (GFP): Applications, Structure, and Related

Photophysical Behavior. Chemical Reviews, 102(3), 759–782.

https://doi.org/10.1021/cr010142r

Page 171: Observation and quantification of protein production in single living cells

159

Fin