isolation and characterization of exosomes

180
ISOLATION AND CHARACTERIZATION OF EXOSOMES BY BOW JESSE TAURO SUBMITTED IN TOTAL FULFILMENT OF THE REQUIREMENTS OF THE DEGREE OF DOCTOR OF PHILOSOPHY AUGUST 2013 DEPARTMENT OF BIOCHEMISTRY AND MOLECULAR BIOLOGY THE UNIVERSITY OF MELBOURNE LUDWIG INSTITUTE FOR CANCER RESEARCH LTD. PARKVILLE BRANCH & LA TROBE INSTITUTE FOR MOLECULAR SCIENCE LA TROBE UNIVERSITY

Upload: others

Post on 13-Mar-2022

14 views

Category:

Documents


0 download

TRANSCRIPT

ISOLATION AND

CHARACTERIZATION OF EXOSOMES

BY

BOW JESSE TAURO

SUBMITTED IN TOTAL FULFILMENT OF THE REQUIREMENTS OF THE DEGREE

OF DOCTOR OF PHILOSOPHY

AUGUST 2013

DEPARTMENT OF BIOCHEMISTRY AND MOLECULAR BIOLOGY

THE UNIVERSITY OF MELBOURNE

LUDWIG INSTITUTE FOR CANCER RESEARCH LTD.

PARKVILLE BRANCH

&

LA TROBE INSTITUTE FOR MOLECULAR SCIENCE

LA TROBE UNIVERSITY

iii

“It is the theory which decides what we can observe.”

Albert Einstein

“At the heart of science is an essential balance between two seemingly contradictory

attitudes - an openness to new ideas, no matter how bizarre or counterintuitive they may

be, and the most ruthless skeptical scrutiny of all ideas, old and new. This is how deep

truths are winnowed from deep nonsense.”

Carl Sagan

iv

Abstract

Cell-cell communication is an integral physiological process that relies on the sending and

receiving of signals. Communication may involve direct contact between adjoining cells, or

require the release of secreted molecules to facilitate the interaction. Recently, extracellular

vesicles (EVs) secreted from cells have been recognized to be involved in cell-cell

communication. EVs comprise shed microvesicles (sMVs), apoptotic bodies and exosomes

which differ based on their mechanism of biogenesis and size; of these, exosomes have

been most widely studied. Exosomes are ~40-100 nm EVs released from a multitude of cell

types that perform pleiotropic extracellular functions within the cellular microenvironment.

These functions include autocrine and paracrine signaling, immunological modulation, and

horizontal transfer of proteins, lipids and genetic material (miRNA/mRNA) to recipient

cells. When investigating exosome functionality, sample homogeneity is of critical

importance, a caveat which, thus far, has been ostensibly overlooked. Although numerous

functions have been ascribed to exosomes, the interpretation of the majority of these studies

has been confounded by sample heterogeneity with other secreted vesicles.

This thesis aims to address the concern of exosome purity by evaluating widely-used

exosome isolation procedures, using the human colorectal cancer (CRC) cell line LIM1863

as a model. The information gained from this comparative purification study will be used to

provide insights into the isolation and characterization of two exosome sub-populations

released from this highly-polarized CRC cell line. Additionally, using an epithelial-

mesenchymal transition (EMT) cell model, exosomes will be isolated and characterized for

the purpose of identifying their contribution to the EMT process.

The first experimental chapter compared and evaluated three strategies for exosome

isolation: differential ultracentrifugation, density gradient centrifugation using iodixanol

(OptiPrep™) and EpCAM immunoaffinity capture. All exosome preparations contained 40-

100 nm diameter vesicles based on electron microscopy, and were positive for stereotypical

exosome markers Alix, TSG101, HSP70 using Western blotting. Using MS–based

proteomic profiling, the protein composition of exosomes isolated from each of these

v

procedures was investigated. The effectiveness of each method was assessed by

quantitating the number of MS/MS spectra identified for exosome markers and proteins

associated with key exosome processes including their biogenesis, intracellular trafficking,

and exocytic release. This study revealed that proteins in all aforementioned categories

were significantly enriched (at least 2-fold) in immunoaffinity isolated exosomes when

compared to density gradient- or differential ultracentrifugation-derived exosomes. Overall,

this study concluded that immunoaffinity capture was the most effective method for

isolating exosomes.

The second experimental chapter investigated sub-populations of exosomes from highly-

polarized LIM1863 cells. In this study, a sequential immunoaffinity capture strategy was

developed using anti-A33- and anti-EpCAM-coupled magnetic beads to isolate A33- and

EpCAM-Exos. A key finding of this study was the exclusive identification in EpCAM-

Exos of the classical intracellular apical trafficking molecules CD63, mucin 13, and apical

intestinal enzyme sucrase isomaltase. The increased expression of dipeptidyl peptidase IV

and the apically-restricted pentaspan membrane glycoprotein prominin 1 was also

observed. In comparison, A33-Exos were enriched with intracellular basolateral trafficking

molecules including early endosome antigen 1 (EEA1), the Golgi membrane protein ADP-

ribosylation factor (ARF1) and clathrin. Collectively, these data are consistent with

EpCAM- and A33-Exos being released from the apical and basolateral surfaces,

respectively. Intriguingly, several members of the MHC class I family of antigen

presentation molecules were exclusively observed in A33-Exos. Additionally, EpCAM-

Exos contained molecules known to complex together to promote tumor progression

including, EpCAM, claudin-7 and CD44. This study is the first robust characterization of

two distinct populations of exosomes from highly-polarized epithelial cells.

A proteome analysis of LIM1863-derived sMVs (also referred to as plasma membrane

blebs, microvesicles and oncosomes) was also performed to reveal that sMVs are clearly

distinguishable from A33- and EpCAM-Exos. Interestingly, sMVs were shown to be

significantly enriched with members of the human ATP-binding cassette (ABC) transporter

superfamily, which act to shuttle substrates (e.g., drugs) across cellular membranes.

vi

The third experimental chapter analyzed exosomes released from cells following EMT.

EMT is a highly conserved morphogenic process defined by the loss of epithelial

characteristics and the acquisition of a mesenchymal phenotype. This process is associated

with increased aggressiveness, invasiveness, and metastatic potential in carcinoma cells.

Using an EMT cell model consisting of Madin-Darby canine kidney cells (MDCK) and its

oncogenic H-Ras-induced variant (21D1) established by Dr. Zhu (The University of

Melbourne), MDCK- and 21D1- exosomes (MDCK- and 21D1-Exos) were isolated using

density gradient centrifugation (OptiPrep™). Proteomic profiling revealed that typical

cellular EMT marker proteins are present in MDCK- and 21D1-Exos. These include

reduction of a characteristic inhibitor of angiogenesis (e.g., thrombospondin-1) and

epithelial markers (e.g., E-cadherin and EpCAM), with a concomitant upregulation of a

mesenchymal marker (e.g., vimentin). Further, this study revealed that 21D1-Exos were

enriched with several proteases and integrins that have been recently implicated in

regulating the tumor microenvironment to promote metastatic progression. A salient

finding, however, was the unique enrichment of key transcriptional regulators (e.g., the

master transcriptional regulator YBX1) and core splicing complex components in

mesenchymal 21D1 cell-derived exosomes. Overall, these findings demonstrate that

exosomes from Ras-transformed MDCK cells are reprogrammed with factors which may

be capable of inducing EMT in recipient cells.

In summary, this thesis evaluated three strategies for purifying exosomes. The findings of

this study revealed that immunoaffinity capture was significantly more effective at

enriching exosomes from cellular media than density gradient centrifugation or differential

centrifugation. This purification knowledge was used to isolate and characterize two

populations of exosomes secreted from the colon carcinoma cell line LIM1863 as well as to

evaluate the contribution of exosomes in the epithelial-mesenchymal transition process.

vii

Declaration

This is to certify that:

(i) this thesis comprises only my original work towards the PhD except where

indicated in the Preface.

(ii) due acknowledgement has been made in the text to all other material used.

(iii) the thesis is less than 100,000 words in length, exclusive of tables, maps,

bibliographies, appendices and footnotes.

Signed: _________________________

Bow Tauro

viii

Preface

Publications during candidature not included in this thesis:

1. Mathivanan, S., Lim, J. W. E., Tauro, B. J., Ji, H., Moritz, R. L., and Simpson, R. J.

(2010) Proteomics analysis of A33 immunoaffinity-purified exosomes released from the

human colon tumor cell line LIM1215 reveals a tissue-specific protein signature. Molecular

& Cellular Proteomics 9 (2), 197-208. PMID: 19837982.

2. Mathivanan, S., Ji, H., Tauro, B. J., Chen, Y. S., and Simpson, R. J. (2012) Identifying

mutated proteins secreted by colon cancer cell lines using mass spectrometry. Journal of

Proteomics 5 (76), 141-149. PMID: 22796352.

3. Ji, H., Greening, D. W., Barnes, T. W., Lim, J. W. E., Tauro, B. J., Adda, C.,

Mathivanan, S., Zhao, W., Xue, Y., Xu, T., and Simpson, R. J. (2013) Proteome profiling

of exosomes derived from human primary and metastatic colorectal cells reveal differential

expression of key metastatic factors and signal transduction components. Proteomics 13

(10-11), 1672-1686. PMID: 23585443.

ix

Publications during candidature included in this thesis:

1. Tauro, B. J., Greening, D. W., Mathias, R. A., Ji, H., Mathivanan, S., Scott, A. M., and

Simpson, R. J. (2012) Comparison of ultracentrifugation, density gradient separation, and

immunoaffinity capture methods for isolating human colon cancer cell line LIM1863-

derived exosomes. Methods 56, 293-304. PMID: 22285593. Chapter 2 of this thesis.

2. Tauro, B. J., Greening, D. W., Mathias, R. A., Mathivanan, S., Ji, H., and Simpson, R. J.

(2012) Two distinct populations of exosomes are released from LIM1863 colon carcinoma

cell-derived organoids. Molecular & Cellular Proteomics 12, 587-598 PMID: 23230278.

Chapter 3 of this thesis.

3. Tauro, B. J., Mathias, R. A., Greening, D. W., Gopal, S. K., Ji, H, Kapp, E. A.,

Coleman, B. M., Hill, A. F., Kusebauch, U., Hallows, J. L., Shteynberg, D., Moritz, R. L.,

Zhu, H. J., and Simpson R. J. (2013) Oncogenic H-Ras reprograms Madin-Darby canine

kidney (MDCK) cell-derived exosomal proteins during epithelial-mesenchymal transition.

Molecular & Cellular Proteomics 12, 2148-2159. PMID: 23645497. Chapter 4 of this

thesis.

x

The work described in the manuscripts included this thesis was performed entirely by

myself except for the following collaborations: Chapters 2 and 3, Dr. David Greening and

Dr. Rommel Mathias (assistance with manuscript editing), Dr. Suresh Mathivanan

(performed initial bioinformatic analysis), Chapter 4, Dr. Rommel Mathias (assistance with

manuscript writing), Dr. David Greening (assistance with manuscript editing), Mr. Shashi

Kumar Gopal Krishnan (assistance with cell culture for Figure 2), Mr. Eugene Kapp

(performed initial bioinformatic analysis), Dr. Bradley Coleman (cryo-transmission

electron microscopy imaging), A/Prof. Robert Moritz (MS analysis of a second (replicate)

dataset). I estimate that my contribution to Chapter 2 and 3 to be greater than 90% and my

contribution to Chapter 4 to be greater than 80% therefore achieving an overall contribution

of my work in the thesis to be greater than 85%.

xi

Acknowledgements

This thesis is an account of several years of devoted work while enrolled through the

Department of Biochemistry and Molecular Biology at The University of Melbourne.

Experimental work was performed at the Ludwig Institute for Cancer Research (LICR) and

the La Trobe Institute for Molecular Science (LIMS).

First, I would like to express my deep appreciation to my primary supervisor, Prof. Richard

Simpson, for allowing me to undertake my PhD. He is incredibly accomplished within the

scientific field and his level of success is something to aspire to. His ability to continually

enthuse and inspire me, even in the toughest of times, has been amazing. I have genuinely

enjoyed spending these years with him and value the friendship that has evolved further

than a supervisor/student relationship. Attending conferences together, particularly the

International Society for Extracellular Vesicles meeting in Sweden, have been a great

highlight with many fond memories.

Next I would like to thank my co-supervisors, Dr. Hong Ji and Dr. Hong-Jian Zhu. From

the first day I stepped into the laboratory, a friendship with Ji Hong grew like the orchids

she would nurture; we were instantly a great team. Her skills within the laboratory are

second to none. Her experimental design and execution are impeccable; I can only hope

that my skills will continue to grow to one day reach her level. Hong-Jian’s guidance as an

external supervisor has been amazing; his knowledge regarding key concepts including

epithelial-mesenchymal transition has been invaluable.

I would also like to thank the members of my PhD Committee, Prof. Andrew Hill and

A/Prof. Matthew Perugini, who both provided critical review and guidance during my PhD

research. Additionally, I would also like to acknowledge Prof. Antony Burgess and Prof.

Nick Hoogenraad who were the directors of LICR and LIMS, respectively, during my PhD

candidature.

xii

I would also like to thank staff and students, past and present, who have given me unending

support. In particular, Dr. David Greening and Dr. Rommel Mathias whose clear thinking

and focused attitudes I admire greatly. With these qualities they assisted me to distill my

vast amounts of experimental data into a publishable form for which I am deeply greatful.

Additionally, I would like to thank Dr. Thomas Barnes, Mr. Robert Goode, Dr. Yuan-Shou

Chen, Dr. Justin Lim, Dr. Nicole Harris, Mr. Robert Speirs, Mr. Shashi Kumar Gopal

Krishnan, Mr. Alin Rai, and Mr. Rong Xu. I thank them all for the support, laughter and the

great times we shared together throughout my PhD.

Thanks must also go to A/Prof. Robert Moritz, Dr. Suresh Mathivanan and Dr. Eugene

Kapp. Their collective knowledge of proteomics and mass spectrometry has been vital to

my learning and the completion of this thesis.

I would also like to give thanks to my parents, Geraldine and Peter, my sister Eve, and

brothers Rory and Tristan, I love and thank them all. My family have given me unending

support in all of my endeavors and helped me to grow to become who I am today. I would

also like to acknowledge my influential and motivating chemistry teacher, Mr. Robert

Timoney, who taught me many key scientific concepts that I still reflect on to this day.

Finally, I would like to thank my fiancée, Jess Williamson. Literally since day one, Jess has

been by my side, encouraging and supporting me along on the way, I could not have done it

without her. My love continues to grow for this inspiring person and I look forward to

tackling the next chapter with her in Ghana.

xiii

Table of Contents

Abstract .................................................................................................................................... iv

Declaration............................................................................................................................. vii

Preface .................................................................................................................................. viii

Acknowledgements ................................................................................................................. xi

List of Tables ......................................................................................................................... xvi

List of Figures and Illustrations ........................................................................................... xvii

Abbreviations ........................................................................................................................ xix

Chapter 1: Background and literature review ........................................................................... 1

1.1 Preface ................................................................................................................... 1

1.2 Secretome .............................................................................................................. 2

1.3 Extracellular vesicles ............................................................................................. 2

1.3.1 Shed microvesicles ................................................................................................ 5

1.3.2 Apoptotic blebs ...................................................................................................... 6

1.4 Exosomes ............................................................................................................... 7

1.4.1 Exosome biogenesis and cargo selection ............................................................... 8

1.4.2 Endosome motility and exosome release ............................................................. 11

1.4.2.1 Exosome biogenesis and release from polarized cells ......................................... 12

1.4.3 Exosome sizing and morphology ........................................................................ 13

1.4.4 Exosomal proteins ............................................................................................... 14

1.4.5 Exosomal lipids ................................................................................................... 16

1.4.6 Exosomal RNA .................................................................................................... 18

1.4.7 Exosome targeting and uptake ............................................................................. 20

1.4.8 Exosome function in the tumor microenvironment and cancer progression ....... 23

xiv

1.4.9 Exosome therapeutics .......................................................................................... 27

1.4.10 Exosome and microvesicle diagnostics ............................................................... 32

1.4.11 Exosome isolation strategies................................................................................ 32

1.5 Research models used in this thesis ..................................................................... 34

1.5.1 Colorectal cancer model ...................................................................................... 34

1.5.1.1 Colorectal cancer ................................................................................................. 34

1.5.1.2 CRC statistics and subtypes ................................................................................. 35

1.5.1.3 Colorectal cancer cell line LIM1863 ................................................................... 37

1.5.2 Epithelial-mesenchymal transition (EMT) model ............................................... 38

1.5.2.1 EMT during cancer progression .......................................................................... 38

1.5.2.2 Molecular markers of EMT ................................................................................. 41

1.5.2.3 Signaling pathways and inducers of EMT ........................................................... 42

1.5.2.4 Madin-Darby canine kidney (MDCK) and 21D1 cell lines................................. 43

1.5.2.5 Proteomic analysis of the secretome from MDCK cells following EMT ........... 43

1.5.2.6 Exosomes and EMT ............................................................................................. 45

1.6 Thesis aims .......................................................................................................... 45

Chapter 2: Comparison of methods for isolating exosomes ................................................... 49

2.1 Overview.............................................................................................................. 49

Chapter 3: Use of sequential immunoaffinity capture to isolate and characterize two distinct

populations of exosomes released from LIM1863 colon carcinoma cells ............................. 65

3.1 Overview.............................................................................................................. 65

Chapter 4: Characterization of Madin-Darby canine kidney (MDCK) cell-derived exosomal

proteins following oncogenic H-Ras induced epithelial-mesenchymal transition ................. 81

xv

4.1 Overview.............................................................................................................. 81

Chapter 5: Summary and future directions ............................................................................. 95

5.1 Evaluation of exosome isolation methods and the importance of purity............. 95

5.2 Analysis of two exosome sub-populations released from LIM1863 cells ........... 96

5.3 Analysis of exosomes following epithelial-mesenchymal transition (EMT) ...... 98

5.4 Conclusion ......................................................................................................... 100

References ............................................................................................................................ 102

Appendix .............................................................................................................................. 139

Appendix A-1 Isolation and characterization of cell line-derived exosomes ............... 140

Appendix A-2 Isolation and characterization of body fluid-derived exosomes............ 146

Appendix A-3 Human colon cancer LIM1863 cells ..................................................... 149

Appendix A-4 Distinguishing features of EMT ............................................................ 150

Appendix A-5 Secreted extracellular modulators of EMT ........................................... 151

Appendix A-6 Supplemental information for manuscripts included in this thesis ....... 152

Appendix A-7 Distribution of proteins identified in UC-, DG-, and IAC-Exos ........... 154

Appendix A-8 Plasma membrane proteins identified in the exosome datasets ............ 155

Appendix A-9 21D1 cells require their own culture medium to maintain a

mesenchymal phenotype ....................................................................... 156

Appendix A-10 Exosome characterization following OptiPrep™ density gradient

separation .............................................................................................. 157

Appendix A-11 Experimental procedures for replicate GeLC-MS/MS analysis of

MDCK- and 21D1-Exos ........................................................................ 158

xvi

List of Tables

Chapter 1

Table 1-1 Properties of extracellular vesicles ................................................................... 4

Table 1-2 A summary of in vitro studies of miRNAs identified in cancer cell

exosomes ........................................................................................................ 19

Table 1-3 Examples of EV-mediated transfer of oncogenic/transforming molecules .... 25

Table 1-4 Therapeutic applications of exosomes ............................................................ 31

Chapter 2

Table 2-1 Relative quantitation of selected exosome proteins by label-free spectral

counting .......................................................................................................... 61

Chapter 3

Table 3-1 Representative list of proteins either uniquely localized or significantly

enriched in EpCAM-Exos .............................................................................. 74

Table 3-2 Representative list of proteins either uniquely localized or significantly

enriched in A33-Exos ..................................................................................... 75

Table 3-3 List of colon cancer-related and normal colon tissue specific membrane

proteins in LIM1863-derived exosomes ......................................................... 76

Chapter 4

Table 4-1 EMT hallmark proteins identified in MDCK- and 21D1-Exos .......................... 88

Table 4-2 Exosomal factors involved in metastatic niche formation and metastasis .......... 89

Table 4-3 Splicing factors and transcription factors enriched in 21D1-Exos ...................... 90

xvii

List of Figures and Illustrations

Chapter 1

Figure 1-1 Extracellular membranous vesicles released from cells ................................... 3

Figure 1-2 Exosomal cargo selection involving ESCRT complexes.................................. 9

Figure 1-3 Exosome targeting and uptake ........................................................................ 21

Figure 1-4 Extracellular vesicles (EVs) are components of the intra- and intercellular

oncogenic signaling circuitry ......................................................................... 24

Figure 1-5 Functional roles and therapeutic targeting of extracellular vesicles (EVs) .... 28

Figure 1-6 Advantages and drawbacks of siRNA delivery by exosomes, viruses and

lipid nanoparticles .......................................................................................... 29

Figure 1-7 Model of intestinal cancer initiation and progression ..................................... 36

Figure 1-8 Epithelial-mesenchymal transition.................................................................. 39

Figure 1-9 EMT during tumor progression and metastasis .............................................. 40

Figure 1-10 MDCK and Ras-transformed MDCK cells (21D1) morphology .................... 44

Chapter 2

Figure 2-1 Exosome isolation .......................................................................................... 56

Figure 2-2 Characterization of exosomes ........................................................................ 57

Figure 2-3 Semi-quantitative normalized spectral count ratios of selected exosome

proteins ........................................................................................................... 58

Chapter 3

Figure 3-1 Isolation of exosomes and shed microvesicles from the colon carcinoma cell

line LIM1863 .................................................................................................. 72

Figure 3-2 Morphological characterization and proteome analysis of LIM1863 cell-

derived A33- and EpCAM-Exos .................................................................... 73

Figure 3-3 A confocal optical section though a preparation of LIM1863 organoids ...... 77

Figure 3-4 A model depicting the molecular structure of two discreet populations of

LIM1863-derived exosomes ........................................................................... 70

xviii

Chapter 4

Figure 4-1 Isolation of exosomes from MDCK and 21D1 cells ...................................... 86

Figure 4-2 Characterization of MDCK- and 21D1-Exos................................................. 87

Figure 4-3 Proteomic analysis of exosomes .................................................................... 87

xix

Abbreviations

ADAM a disintegrin and metalloproteinase

A33-Exos exosomes isolated using anti-A33 immunoaffinity beads

BMDC bone marrow-derived hematopoietic progenitor cells

BMP bone morphogenic protein

CCM concentrated culture medium

CM culture medium

CRC colorectal cancer

DMEM Dulbecco’s modified eagle medium

EGF epidermal growth factor

EGFR epidermal growth factor receptor

EM electron microscopy

EMMPRIN extracellular matrix metalloproteinase inducer

EMT epithelial-mesenchymal transition

eMVs extracellular microvesicles

EpCAM epithelial cell adhesion molecule

EpCAM-Exos exosomes isolated using anti-EpCAM immunoaffinity beads

ESCRT endosomal sorting complex required for transport

FACS fluorescence-activated cell sorting

FCS fetal calf serum

FGF fibroblast growth factor

FITC fluorescein isothiocyanate

FOXC2 forkhead box protein C2

GI gastrointestinal

GSK-3β glycogen synthase kinase 3 β

HMGB high mobility group box

HPLC high-performance liquid chromatography

HSP heat shock protein

HUVEC human umbilical vein endothelial cells

ILV intraluminal vesicle

ITS insulin-transferrin-selenium

LDH lactate dehydrogenase

xx

LEF lymphoid enhancer factor

MACS magnetic activated cell sorting

MAPK mitogen-activated protein kinase

MDCK Madin-Darby canine kidney

MMP matrix metalloproteinase

MTT 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazoliumbromide

MVB multivesicular body

MVs microvesicles

NSF N-Ethylmaleimide-sensitive fusion protein

PDCD6IP/Alix programmed cell death 6 interacting protein

PI(3)K phosphatidylinositol 3-kinase

PM plasma membrane

Rsc relative spectral count fold change ratios

SIP smad interacting protein

SMART simple modular architecture research tool

sMVs shed microvesicles

SNARE soluble N-ethylmaleimide-sensitive factor attachment protein receptor

SSM solid support magnet

TEM tetraspanin enriched microdomains

TGN trans-Golgi network

TMHMM transmembrane hidden Markov model

TF tissue factor

TGFβ transforming growth factor β

TNFα tumor necrosis factor α

TPP trans-proteomic pipeline

TNTs tunneling nanotubes

TSG101 tumor susceptibility gene 101

VEGF-A vascular endothelial growth factor A

VEGFR-1 vascular endothelial growth factor receptor 1

VAMP vesicle-associated membrane protein

WCL whole cell lysate

YBX1 Y-box binding protein 1

ZEB1 zinc finger E-box binding homeobox 1

21D1 Ras-transformed MDCK cells

1

Chapter 1: Background and literature review

1.1 Preface

Exosomes are 40-100 nm diameter membraneous vesicles released from most cell types.

They function in extracellular communication by transferring a variety of bioactive

molecules (e.g., proteins, lipids, RNA species such as mRNA/miRNAs, and possibly DNA)

between cells [1, 2]. Over the past decade they have gained much attention for their role in

cancer progression [3]. Although exosomes were first described almost thirty years ago [4-

6], their purification to homogeneity has proven a difficult task. This is due to their inherent

biophysical properties (e.g., size and buoyant density) being similar to extracellular vesicles

(EVs) (e.g., shed microvesicles and apoptotic blebs), the problem of co-purification with

protein oligomers (e.g., proteasome complexes), and the possible contamination with

cellular pathogens (e.g., adenoviruses and prions) [7]. Obtaining homogeneous exosomes is

a pre-requisite for in-depth biophysical and functional analyses.

Using semi-quantitative MS-based proteomic techniques, such as label-free spectral

counting [8], this thesis first compares widely used exosome isolation strategies to assess

their capability for enriching several classes of proteins involved in exosome biogenesis,

trafficking, release and uptake. Secondly, a sequential immunoaffinity capture technique

was developed to isolate exosome sub-populations released from apical and basolateral

surfaces of highly-polarized cells. Finally, in the absence of an immunoaffinity capture

strategy, density gradient centrifugation (OptiPrep™) was used to isolate and characterize

exosomes, for the first time, from an epithelial-mesenchymal transition (EMT) cell model.

2

1.2 Secretome

The term ‘secretome’ refers to all soluble proteins and extracellular membranous vesicles

that are secreted or shed into the extracellular space [9-15]. The secretome is a tightly

regulated and sensitive feature required for cell-cell communication and normal

physiological function [9]. Secreted proteins can constitute structural components of the

extracellular matrix (ECM), such as collagens and laminins, while others are able to trigger

intercellular responses in target cells [16, 17]. Secreted proteins are of particular importance

as aberrant protein secretion is associated with pathological events within diseases such as

cancer [13, 14], including pre-metastatic niche formation and metastasis [18-21]. In

addition to soluble-secreted proteins, the secretome contains several types of EVs including

shed microvesicles (sMVs), apoptotic blebs and exosomes (Figure 1-1, Table 1-1) [1]. It is

now recognized that these vesicles are also important regulators of cell-cell communication,

particularly within the tumor microenvironment [19, 22].

1.3 Extracellular vesicles

The release of EVs was first described in 1967 when Wolf and colleagues recognized pro-

coagulant particulate matter around activated platelets (‘platelet dust’) [23]. Platelets were

later identified to release microvesicles and exosomes [24]. In addition to normal cellular

functioning, EV release can be caused by diverse biological mechanisms that are triggered

during oncogenic transformation, cellular activation by stimuli in the microenvironment,

stress or apoptosis [25]. The formation of vesicles may occur as budding events at the

plasma membrane (e.g., sMVs) or within late endosomal compartments such as

multivesicular bodies (e.g., exosomes), or as a consequence of apoptosis (e.g., apoptotic

blebs). Following release, different EVs may have distinct functions. It is understood that

isolating EVs to homogeneity is the first step towards obtaining a clear understanding of

their biophysical properties and biological functions. However, such an undertaking has

been impeded by the inherent technical difficulties – due largely to the similarity of their

physiochemical characteristics (Table 1-1) [26].

3

Figure 1-1. Extracellular membranous vesicles released from cells

Shed microvesicles (sMVs) range in size between 100-1000 nm in diameter and are formed

by direct outward budding of the plasma membrane (PM). During endocytosis, a receptor is

internalized and sorted in early endosomes, where cargo may be recycled to the PM via

recycling endosomes [27, 28]. Intraluminal vesicles (ILVs) are generated by inward

budding from the limiting membrane of late endosomes, forming multivesicular bodies

(MVBs). MVBs may be (i) sorted to the lysosome and become degradative or (ii) fuse with

the PM to release 40-100 nm ILVs, now termed exosomes [29]. Exosomes display the same

surface topology as the cell with extracellular domains of proteins at the surface, and

enclosing cytoplasmic contents (cargo) in the lumen. Apoptotic blebs (50-5000 nm in

diameter) are formed from cells undergoing programmed cell death.

4

Table 1-1. Properties of extracellular vesicles [1, 7, 26, 30-36]

Exosomes Shed Microvesicles

(sMVs) Apoptotic blebs

Size 40-100 nm 100-1000 nm 50-5000 nm

Mechanism

of

biogenesis Exocytosis of MVBs

Outward budding of

the plasma membrane Release of fragments/blebs from

cells undergoing apoptosis

Isolation

techniques

Differential centrifugation at 100,000 g [35]

Sucrose and OptiPrep™ density gradient

ultracentrifugation (1.08-1.22 g/mL) [35]

Immunoaffinity capture: HER2 positive exosomes from breast cell lines [37] Rab-5b positive exosomes from melanoma cell lines

[38] CD45 positive exosomes from Jurkat T cells [39] MHC class II positive exosomes from antigen

presenting cells [40] CD63 positive exosome-like vesicles from blood [41] EpCAM positive exosomes from human serum of

lung and ovarian cancer [42, 43] A33 positive exosomes from colorectal cancer cell

line [44] Size-exclusion HPLC [45-47]

Differential

centrifugation at

10,000 g [24]

Sucrose density

gradient

ultracentrifugation

(1.16 g/mL) [48]

Fluorescence-activated

cell sorting (FACS) 1-

10 µm size gate [49]

Induction of apoptosis in cells,

isolate apoptotic bodies at 50,000

g [50, 51]

Sucrose density gradient

ultracentrifugation (1.16-1.28 g/mL) [32]

Markers Alix, TSG101, CD9, CD63 Annexin V binding,

ARF6, ABCG2 Annexin V binding, DNA content,

phosphatidylserine

5

A recent initiative to assist in the cataloguing of EVs is the database Vesiclepedia

(http://www.microvesicles.org/) [52]. Vesiclepedia is a manually-curated compendium of

molecular information (protein, lipid and RNA) identified in different classes of EVs.

Vesiclepedia comprises 35,264 protein, 18,718 mRNA, 1,772 miRNA and 342 lipid entries

comprised from 341 independent studies. A summary of EVs (sMVs, apoptotic blebs and

exosomes) and their biophysical properties is outlined below.

1.3.1 Shed microvesicles

Shed microvesicles (sMVs) range in diameter between 100-1000 nm and are formed by

direct budding of cytoplasmic protrusions from the cell membrane upon activation of

various internal and external stimuli (Figure 1-1) [7, 26]. Nomenclature of these vesicles

has been defined depending on their parent cell; for example, microvesicles released by

platelets are referred to as ‘microparticles’ while microvesicles released by leukocytes are

called ‘ectosomes’ [31]. Prostate cancer xenograft released vesicles are termed

‘oncosomes’ [49]. They are collectively referred to as sMVs in this thesis. The biogenesis

of sMVs is reported to involve phospholipid redistribution caused by flippases and

floppases translocating phospholipids from the outer-to-inner and inner-to-outer membrane

leaflets, respectively [36]. Following this rearrangement, contraction of cytoskeletal

structures by actin-myosin interactions, caused by ADP-ribosylation factor 6 (ARF6),

initiates a signaling cascade that culminates in the phosphorylation and activation of

myosin light chain, facilitating membrane budding and sMV release [36]. Shedding of

vesicles occurs in resting cells, however, Ca2+

is known to increase the rate of the process

dramatically [53]. Interestingly, despite being generated by the analogous process of

budding from the cell surface, sMVs isolated at rest and following stimulation can be

molecularly different. For example, it was shown that vesicles shed from stimulated

neutrophils express increased levels of integrin αMβ2 [54].

sMVs are recognized to participate in important biological events including coagulation by

mediating the contribution of platelets, macrophages and neutrophils, and horizontal

6

trafficking of protein and RNA between cells [26] (for an excellent review see [55]). It is

well recognized that various classes of sMVs are rich in ATP-Binding Cassette (ABC)

transporters which form a defense network against a range of chemotherapeutics [56]. This

network can result in multidrug resistance, a major impediment to curative cancer

chemotherapy [57-60]. Following release into the extracellular space, sMVs may be

degraded locally or move by diffusion to appear in biological fluids including cerebrospinal

fluid, blood and urine [26]. When investigating sMVs released from tumor cells, it was

demonstrated that they may rapidly break down to release cargo. This may include

extracellular matrix metalloproteinase inducer (EMMPRIN), a transmembrane glycoprotein

expressed at high levels by tumor cells that can stimulate matrix metalloproteinase

expression in fibroblasts to facilitate tumor invasion and metastasis [61]. Additionally,

sMVs (oncosomes) isolated by FACS from prostate cancer xenograft mouse blood were

shown to stimulate the migration of normal endothelial cells indicating their ability to

modify the microvasculature [49]. Methods used to isolate sMVs from culture medium or

platelet-poor plasma include density gradient centrifugation (buoyant density of 1.16 g/mL)

on sucrose gradients [24, 48, 49, 62]. sMVs can also be isolated by FACS using size-gating

determined by polymeric bead standards [49]. Although separation by size-exclusion

chromatography has been performed to isolate sMVs of varying size from urine, this

process resulted in contamination with other EVs and proteins [63]. To date, little is known

about sMVs structure and specific information about their biogenesis is limited.

1.3.2 Apoptotic blebs

Apoptotic blebs (and larger ‘apoptotic bodies’) are heterogeneous fragments of the cell (50-

5000 nm) formed during programmed cell death (Figure 1-1) [36]. The occurrence of cell

shrinkage and condensation of nuclear chromatin, hallmarks of apoptosis, leads to the

generation of apoptotic blebs. These blebs contain remnants of the cell degradation

processes including cytoplasmic components and DNA [64]. It is understood that during

apoptosis, phosphatidylserine (PS) is relocalized from the inner leaflet of the PM to the

outer leaflet where it triggers recognition for phagocytosis [65]. In addition to PS, other

7

phagocytic surface signals including integrins αVβ3/αVβ5 and CD36 are displayed [33, 34,

66]. Interestingly, from a functional aspect, phagocytosis of apoptotic blebs by dendritic

cells results in efficient processing and presentation of apoptotic cell antigens on CD4+ and

CD8+ T lymphocytes [67]. Further, apoptotic bodies derived from H-RasV12 or human c-

myc transfected cells were up taken by murine fibroblasts resulting in loss of contact

inhibition in vitro and a tumorigenic phenotype in vivo [68]. Protocols for the biogenesis

and isolation of apoptotic blebs involve induction of cell apoptosis, commonly by UV

irradiation or ethanol induction, and subsequent cell culture for up to 24 h. At different

times after apoptotic induction, cells are collected and stained with FITC-labeled annexin

V, an early marker of apoptosis. Blebs are then isolated by centrifugation at 50,000 g [50,

51] or by sucrose density gradients at a buoyant density of 1.16-1.28 g/mL [32]. In contrast

to sMVs, which are formed and released at the cell membrane, and apoptotic blebs which

are formed by cell shrinkage during programmed cell death, exosomes are formed by

inward budding of late endosomes. These MVBs subsequently traffic to and fuse with the

plasma membrane to release exosomes into the extracellular environment.

1.4 Exosomes

Exosomes were first described in the 1980s by Johnstone and colleagues while culturing

maturing reticulocytes and were originally thought to be a mechanism for removal of

obsolete cellular proteins [5, 6]. Using transmission electron microscopy they observed the

internalization and release of immunogold labeled transferrin receptors associated with

small (50 nm diameter) vesicles [4]. Exosomes are now recognized to be released from

most cell types including B lymphocytes [69], dendritic cells [70], neurons [71], intestinal

epithelial cells [72] and tumor cell lines [37, 44, 73]. A current definition describes

exosomes as a discrete population of small 40-100 nm diameter membranous vesicles that

are formed within endosomes and released into the extracellular space following fusion of

MVBs with the plasma membrane (Figure 1-1) [74, 75]. This mechanism of biogenesis

differs from that of sMVs and apoptotic blebs and led to the hypothesis that exosomes

might have specialized functions that differ from other EVs [76].

8

1.4.1 Exosome biogenesis and cargo selection

The biogenesis of exosomes occurs within the endosomal network, a membranous

compartment used to sort various intraluminal vesicles and proteins within the cell [77].

Endosomes are separated into three distinct compartments including early, late and

recycling endosomes, each of which are defined by the association of specific proteins on

their cytosolic surface [27, 78]. Early endosomes are characterized by the presence of Rab5

together with its effector VPS34/p150, a phosphatidylinositol 3-kinase (PI(3)K) complex

[78]. These endosomes patrol the peripheral cytoplasm close to the plasma membrane by

movement along microtubules [79] and are the major entry site for endocytosed material

that arrives by clathrin-mediated-, caveolar- and ARF6-dependent pathways [80]. It is

reported that endosomes accept content for approximately 10 min and cargo destined for

recycling is sorted, via syntenin and syndecan heparin sulphate proteoglycans, into

recycling endosomes, which fuse with the plasma membrane [27, 28]. The remaining

material accumulates and is retained in the early endosomes which then undergo maturation

into late endosomes [27].

Maturation of early to late endosomes occurs to reveal a series of transformations including

increased size, a decrease in pH from (6.8-5.9) to (6.0-4.9) and a gradual change in markers

including the loss of Rab5 and acquisition of Rab7 [81, 82]. Following this transformation,

inward budding occurs to form intraluminal vesicles (ILVs/exosomes) in the late endosome

lumen (also termed MVB). This process is governed by endosomal sorting complex

required for transport complexes (ESCRT-0, -I, -II, -III) which are also reported to be

involved in cargo selection [83-85]. The molecular mechanisms regulating cargo selection

into exosomes remain to be fully characterized, however, mono-ubiquitination of the

cytoplasmic tail of membrane proteins has been reported as one of the methods for cargo

selection and sorting [84, 85] (Figure. 1-2). This process requires at least 18 conserved

ESCRT complex proteins that were originally identified in yeast Saccharomyces cerevisiae

[85-87]. ESCRT components are individually recruited from the cytosol to the surface of

endosomes to segregate cargo into domains within the limiting membrane. Specifically,

ESCRT-0, -I, and -II complexes function in sorting of ubiquitin-tagged proteins into the

9

Figure 1-2. Exosomal-cargo selection involving ESCRT complexes

Plasma membrane proteins (e.g., Integrin, E-cadherin) are selected as cargo following

ubiquitination (Ub). The ubiquitin- and lipid-binding domains are indicated (FYVE, UIM,

UEV, GLUE). The ESCRT machinery components form ESCRT complexes -0, -I, -II, -III,

which are recruited in a sequential manner to the limiting membrane of the multivesicular

endosome (MVE). Following inward budding, ATPase VPS4 mediates the disassembly of

ESCRT-III oligomers. Yeast nomenclature is used in this figure, with the exception of

HRS, STAM, and TSG101 (mammalian). UIM, ubiquitin interacting motif; UEV, ubiquitin

E2 variant; GLUE, GRAM-like ubiquitin-binding in Eap45; PI(3)P, phosphatidylinositol-3-

phosphate; PI(3,5)P2, phosphatidylinositol-3,5-bisphosphate; DUBs, de-ubiquitinating

enzymes. Adapted from [88].

10

endosomal delimiting membrane [84, 89]. After sorting, the ESCRT-III complex

sequentially associates and recruits the de-ubiquitinating enzyme Doa4 to remove the

ubiquitin tag from the protein cargo, prior to incorporation into ILVs [90]. The ESCRT-III

complex is also responsible for membrane budding and scission in a ‘purse string’ model

[84, 89, 91]. It is reported that depletion of ESCRT machinery components results in fewer

ILVs, and accumulation of cargo in endosomes with abnormal morphology. This indicates

their importance in this process [92].

Conversely, it has been reported that cargo incorporation into exosomes does not depend on

ubiquitination and ESCRT machinery [1]. This is exemplified by the trafficking of pre-

melanosome protein 17 (Pmel17), which through association with CD63, relies on a

luminal domain enabling its sorting and inclusion into exosomes [93, 94]. Exosome

biogenesis also appears to be dependent on cytoplasmic adaptors syntenin and syndecan

[95] and sphingolipid ceramide [96]. In a study by Trajkovic and colleagues, ceramide was

shown to be enriched in exosomes; they demonstrated that the release of exosomes was

reduced following inhibition of neutral sphingomyelinase (nSMase), suggesting the lateral

segregation of lipids contributes to exosome formation and subsequent release, a process

devoid of ESCRT machinery [96].

Cargo selection is also dependent on adaptor proteins. For example, phosphatase and tensin

homolog deleted on chromosome 10 (PTEN), a tumor suppressor protein localized to the

cytoplasm and nucleus, has been demonstrated to be secreted in exosomes [97]. It was

found that PTEN secretion in exosomes requires Ndfip1, an adaptor protein for members of

the Nedd4 family of E3 ubiquity ligases [97]. To clearly define the exosome formation

pathway, further examination into ESCRT-dependent and ESCRT-independent formation

mechanisms is required.

11

1.4.2 Endosome motility and exosome release

Following vesicle formation, motility of endosomes is dependent on dynein and kinesin

motors that migrate endosomes in opposite directions along microtubules [98]. Rab

GTPases are also integral in ensuring endosome cargo is delivered to the correct

destination. Through sorting adaptors and tethering factors, Rab GTPases control

membrane identity and vesicle budding, motility and fusion. Following endosome

maturation (see section 1.4.1), late endosomes containing ILVs may be targeted to

lysosomes for degradation via the Rab7 effector Rab-interacting lysosomal protein (RILP),

which mediates minus-end-directed trafficking [99]. Late endosomes move to the

perinuclear area of the cell, fuse together to form larger bodies, undergo transient ‘kiss-and-

run’ interactions and eventually full fusions with lysosomes [100].

Alternative to the lysosome pathway, late endosomes may traffic to, and fuse with, the

plasma membrane to release ILVs; ILVs released into the extracellular space are referred to

as exosomes. This process requires various proteins involved in docking, tethering and

fusion. Previous studies have reported the involvement of Rab35 in regulating exosome

secretion by assisting docking and tethering of MVBs to the plasma membrane [101].

Additionally, Rab27a/b and their effectors Slp4 and Slac2b are reported to be involved

exosome release [102, 103]. Soluble N-Ethylmaleimide-sensitive factor attachment protein

receptor (SNARE) molecules are also involved in recognition and fusion of endosomal

membranes to the plasma membrane by forming tetrahelical bundles on opposing

membranes [104]. In a recent study, it was demonstrated that the SNARE Ykt6 was integral

in the secretion of exosomes carrying the morphogen Wnt [105]. Further, vesicle-associated

membrane proteins (VAMPs or synaptobrevins) are anchored in the vesicular membrane to

mediate intracellular vesicle fusion [106, 107]. Finally, VAMP7 and ATPase NSF, a

protein required for SNARE disassembly, are shown to mediate the fusion between MVBs

with the plasma membrane to release exosomes into the extracellular space [108].

12

1.4.2.1 Exosome biogenesis and release from polarized cells

Investigation into exosome biogenesis and release in polarized cell systems has received

little attention to date. In polarized cells, proteins are sorted into different sub-populations

of carrier vesicles, a process mediated by sorting signals (e.g., tyrosine motifs, GPI anchors

and N-glycans) and signal-decoding machinery (e.g., clathrin adaptors and lipid rafts)

[109]. In these cells, endosomes are strategically located between the biosynthetic and

endocytic routes. This allows newly synthesized and endocytosed proteins to be directed to

the appropriate membrane for interaction with different extracellular environments at their

apical and basolateral surfaces [110]. When considering the biogenesis, trafficking and

release of exosomes in polarized cells, it is interesting to note that apical and basolateral

endocytic routes have been shown to converge at the level of late endosomes in Madin-

Darby canine kidney (MDCK) cells [111]. This has led to a hypothesis that exosomes

destined for apical or basolateral membrane release might be generated within a common

late endosomal pool [110]. However, it has also been suggested that epithelial cells could

have distinct MVBs into which apical and basolateral proteins are differentially sorted and

exosomes are formed [110].

Studies investigating these hypotheses have been quite limited. To date, only a single

experimental study has investigated exosomes released from apical and basolateral

surfaces. It was observed that polarized intestinal epithelial cells (T84), grown as a

monolayer on microporous filters, allowed for collection of apical or basolateral exosomes.

These exosomes contained proteins localized to corresponding plasma membrane surfaces

including dipeptidyl peptidase IV on apical exosomes and glycoprotein A33 on basolateral

exosomes [72]. Interestingly, using gel filtration, another study identified two populations

of exosomes in human saliva, however it cannot be claimed that these are apical/basolateral

exosomes due to the potential of multiple cell types contributing to the saliva sample [47].

Further investigation is required to isolate and characterize the contents of apical and

basolateral exosomes. This may lead to an understanding of their formation and trafficking

within polarized cells and potentially identify their varying functions following release.

13

1.4.3 Exosome sizing and morphology

Exosome are characterized by their size of 40-100 nm in diameter [74, 75], however some

studies have reported the identification of larger vesicles up to 150 nm [112]. Various

methods have been used to define vesicle size including transmission electron microscopy

(TEM) which originally identified exosomes to have ‘cup shaped’ morphology [35]. It was

later realized that this phenotype may be a consequence of fixation and dehydration during

sample preparation [1]. In contrast, cryo-TEM preparations, where samples are snap frozen

in liquid ethane, revealed spherical-shaped exosomes that visibly contained ultra-structural

features [113, 114].

Another method used to characterize exosomes is atomic force microscopy (AFM). AFM

provides topographical imaging of exosomes by scanning the surface of deposited samples

with a tipped cantilever and translating tip deflection into a three-dimensional (lateral and

vertical) image of the surface [115]. This method can reach sub-nanometer resolution of

polydisperse samples. However, surface binding may affect the morphology of vesicles and

can hamper the determination of their true diameter [116].

Size-based analysis may also be accomplished by examining the light-scattering properties

of vesicles within a fluid medium that is based upon Brownian motion [117]. Dynamic light

scattering (DLS) is a fast and accurate measurement tool for size distribution and

electrokinetic potential (zeta potential) analysis of monosized particles in suspension.

However, it does have limitations when analyzing polydisperse samples where large

fluctuations in intensity of scattered light are measured [118, 119]. This limitation occurs as

the size distribution analysis is highly influenced by the presence of small amounts of large

particles which scatter more light than small vesicles, causing the distribution to be

weighted towards larger vesicles [116]. Alternatively, it is reported that nanoparticle

tracking analysis (NTA) can analyze polydisperse samples of vesicles between 50-1000 nm

by measuring the absolute size distribution of particles in a mixture [116]. Particles in a

fluid are illuminated by a laser and viewed as small bright spots through a conventional

optical microscope. The Brownian motion of individual particles is recorded through live

14

imaging video and the mean velocity of each particle is calculated with image analysis

software [116]. Although individual particles may be tracked, some limitations of NTA in

comparison to DLS include a limited concentration range (107-10

9 vs. 10

8-10

12

particles/mL) and increased acquisition time (2-5 min vs. 5-60 min). Additionally in NTA,

large aggregates have been reported to mask smaller aggregates, which may enter and leave

the viewing area, making size measurement impossible [120].

Finally, a novel high resolution flow cytometry technique has been developed to quantitate

immunolabeled nano-sized vesicles. Using a new custom-made Influx™ flow cytometer,

small-particle detection has been improved by a wider forward scatter angler and a

photomultiplier tube to allow detection below the previously limiting 300 nm level [121].

This technological advance in the physical characterization of exosomes and other released

vesicles awaits commercialization before becoming widely applicable.

1.4.4 Exosomal proteins

Exosomes have been extensively characterized with respect to their proteome content [35,

52, 122, 123]. Proteomics describes the unbiased and global analysis of all proteins in a

given system [124]. Needless to say, for an accurate protein representation, a homogeneous

exosome sample is essential. Several proteomic techniques, including 1-D sodium dodecyl

sulfate polyacrylamide gel electrophoresis (SDS-PAGE) and 2-D differential in gel

electrophoresis (DIGE) coupled with mass spectrometry (MS) have been used to identify

between ~30 to ~900 proteins from exosome samples [114, 125]. In these workflows, SDS-

PAGE bands/spots are cut, reduced, alkylated and trypsinised. Generated tryptic peptides

are extracted and separated by reversed-phase high-performance liquid chromatography

(HPLC) that is coupled online to a mass spectrometer for analysis [126]. MS tryptic peptide

spectra are then searched against databases using programs including SEQUEST [127],

PROWL [128], and MASCOT [129] to identify proteins. Additionally, Trans-Proteomic

Pipeline (TPP) is an integrated software platform that encompasses most of the steps in a

proteomic data analysis workflow [130]. In addition to identifying proteins, techniques can

15

be used to define the abundance of proteins between samples. Stable isotope labeling by

amino acids in cell culture (SILAC) experimentally measures relative-abundance ratio of

peptides by comparing heavy/light peptide pairs [131]. Other examples of chemical

derivatization techniques for quantitative proteomics include isotope-coded affinity tags

(ICAT), used for the labeling of free cysteine, and isobaric tags for relative and absolute

quantification (iTRAQ), used for the labeling of free primary amine groups [132]. Label-

free quantitation includes either spectral counting or peptide signal intensity to estimate

abundance of proteins [133]. Additionally, semi-quantitative spectral counting using

significant spectral count fold change ratios (RSC) can also be used to quantitate protein

levels between samples [8, 134]. Alternatively, absolute quantitation using selected reaction

monitoring (SRM) [135], also referred to as multiple reaction monitoring (MRM), is a

targeted approach requiring individually designed assays. This involves the selection of a

‘quantotypic’ peptide that is unique to the protein of interest and directly proportional to the

amount of protein present [136]. These isotopically labeled peptides become the internal

calibrant and therefore the unknown abundance of a specific protein can be determined by

comparing its peptide signal intensity with the known calibrant standards [135, 137].

Proteomic studies have been performed on exosomes derived from multiple cell lines and

body fluid sources [122]. Exosomes derived from hematopoietic cells including mast [138],

B and T cells [139, 140], and dendritic cells [141]. Tumor cell lines derived from breast

adenocarcinoma [37], melanoma [142], mesothelioma [143], medulloblastoma [144],

bladder [145], and colorectal cancer [114] are used to characterize cancer-related exosomes

which may assist in defining cell specific markers for disease. Proteomic studies have also

been performed on exosomes derived from body fluids including breast milk [146],

malignant pleural effusions [147], semen [148], urine [149], saliva [47] and blood [41]. An

extensive review of proteomic studies of exosomes isolated from cell lines and body fluids

has been included in Appendix A1 and A2. For a comprehensive account of exosome

studies, including unpublished datasets, see ExoCarta (http://www.exocarta.org/) [122, 123]

and Vesiclepedia (http://www.microvesicles.org/) [52].

16

A conserved set of proteins are observed in exosomes including those involved in their

biogenesis, these include, Alix, TSG101 and other ESCRT complex proteins. Tetraspanin

superfamily proteins CD9, CD63, CD81 and CD82 are commonly found in exosomes; they

are reported to contribute to the protein sorting pathway [74, 150, 151]. Cytosolic proteins

actin and tubulin, and chaperone proteins HSP70 and HSP90 are also often identified in

exosomes. Additionally, exosomes are enriched in proteins that associate with lipid rafts,

including glycosylphosphatidylinositol (GPI) -anchored proteins and flotillin [139] and Rab

GTPases involved in trafficking and endosome fusion events [152]. Exosomes also contain

lipid and nucleic acid (mRNA/miRNA) species [138, 153] (see section 1.4.5 and 1.4.6).

In addition to a conserved set of proteins generally identified in exosomes, a seminal study

comparing proteomic analyses of exosomes revealed that proteins harbored by exosomes

may be cell or tissue specific. This was demonstrated by a comparison of urine, mast cell

and colorectal carcinoma-derived exosomes by Mathivanan and colleagues [44]. A

proteomic analysis was performed on exosomes isolated from the LIM1215 colorectal

carcinoma cell model using immunoaffinity targeting anti-A33, a marker of the basolateral

surface in intestinal epithelial cells [154-156]. This study identified 394 exosomal proteins.

A comparative protein profiling analysis of this study with murine mast cell and human

urine-derived exosomes revealed a common set of 31 proteins including Alix, Rab5 and

annexins A6 and A11. Interestingly, a protein signature reflecting the colorectal cell origin

(LIM1215) was identified. This signature included A33, cadherin-17, carcinoembryonic

antigen (CEA) and epithelial cell adhesion molecule (EpCAM).

1.4.5 Exosomal lipids

In addition to the protein content of exosomes, the lipid components of exosomes have also

been investigated. Characterizing the lipid content of exosomes may assist in our

understanding of their biogenesis and cell-mediated uptake, and in the longer term, the

development of exosome (liposome) encapsulated drugs [153, 157].

17

The process of inward budding of MVBs and exosome formation requires critical lipid

sorting [153]. During inward budding, as two thirds of the lipids required for ILV

formation are located within the inner leaflet of the endosomal membrane, careful

rearrangement and equilibration of lipids from inner and outer leaflets is required [153]. It

is reported that lysobisphosphatidic acid (LBPA) accumulates in the MVB membrane and

interacts strongly with Alix during exosome biogenesis [153]. Interestingly, addition of

LBPA to a lipid composition similar to that of MVBs caused internal budding of small

vesicles within liposomes when adjusted to pH 5.5 (pH of MVBs) [158]. A lipidomic study

of exosomes in comparison to parental cells revealed enrichment in sphingomyelins by 1.3

times in erythrocytes and 2.3 times in B lymphocytes. It was also observed that cholesterol

was enriched in B-lymphocyte-derived exosomes while phosphatidylserine was enriched in

dendritic cell-derived exosomes [153]. Further, it was recognized that phosphatidylcholines

and phosphatidylethanolamines present in exosomes are enriched in saturated fatty acids

comparatively to parental cells [153]. Interestingly, bioactive lipids such as prostaglandins,

and lipid mediators that display multiple biological effects related to inflammation, are

sorted into exosomes [159]. Lipid raft-like domains have also been reported in exosome

membranes. These domains consist of glycerophospholipids with long saturated fatty acyl

chains, cholesterol, flotillin, stomatin, high amounts of sphingolipids and low levels of

phosphatidylcholine [160].

A recent comprehensive lipid profiling study of isogenic primary and metastatic colon cell

lines (SW480 and SW620) revealed differential expression of lipids identified as being

associated with cancer progression [161]. These included increased plasmanylcholine and

triglyceride lipid levels, decreased plasmenylethanolamine and ceramide lipid levels, and a

significant increase of total cholesterol ester and triglyceride lipids in the SW620 cells

compared to those in SW480 cells [161]. It is interesting to hypothesize if similar lipid

contributions would be observed in exosomes. Understanding the lipid content of

exosomes, through comprehensive lipidomic studies of exosomes from various sources,

may shed light on their fate and physiological function.

18

1.4.6 Exosomal RNA

A major breakthrough in characterizing exosomes was the finding of functional mRNAs

and miRNAs, small 22-25 nucleotide non-coding RNAs that suppress the translation of

target mRNAs [162]. This finding highlighted exosomes as prospective vehicles for the

horizontal transfer of biologically important information [163, 164]. Valadi and colleagues

demonstrated that mast cells secreted exosomes containing mRNAs from approximately

1300 genes and more than 100 different miRNAs [138]. Further, an in vitro translational

assay demonstrated that exosomal mRNAs from mouse mast cell line MC/9 are functional

in human mast cell line HMC-1 cells [138]. In a separate study, glioblastoma cell-derived

exosomes containing a reporter gene were incubated with recipient human brain

microvascular endothelial cells (HBMVECs) to reveal mRNA was delivered and translated

[165]. A summary of miRNAs identified from in vitro studies of cancer cell exosomes is

outlined (Table 1-2).

In-depth RNA analysis has been performed to characterize multiple RNA species contained

in exosomes. For example, a small RNA deep sequencing study revealed RNA profiles in

exosomes released from prion-infected neuronal cells [166]. Bellingham and colleagues

demonstrated that neuronal exosomes contain a diverse range of RNA species including

retroviral RNA repeat regions, messenger RNA fragments, transfer RNA fragments, non-

coding RNA, small nuclear RNA, small nucleolar RNA, small cytoplasmic RNA, silencing

RNA as well as known and novel candidate miRNAs. The authors concluded that

exosomes released from prion-infected neuronal cells have a distinct mRNA/miRNA

signature that can be utilized for diagnosis of prion disease [166]. Further work is required

to establish a link between these exosomal RNA species and prion disease pathogenesis.

In another study, deep sequencing technology was used to characterize the RNA content of

exosomes isolated from breast milk [167]. This study revealed 602 unique miRNAs. Of

these, 59 were well characterized immune-relate miRNAs [167]. The longevity of miRNAs

contained within exosomes was increased when compared with circulating RNAs. It was

speculated that due to the protection and durability of exosomal miRNAs, they may assist

19

Table 1-2. A summary of in vitro studies of miRNAs identified in cancer cell exosomes. Modified from [168].

Cell line model Major findings Predominant

microRNAs Target genes or pathways

Ref.

Human and mouse

mast cells

Identified small RNAs, including 121 microRNAs and

1,300 specific mRNAs. Detected mouse exosomal RNA

and new mouse proteins in human mast cells after

treatment with mouse mast cell exosomes. Coined the

term ‘exosomal shuttle RNA (esRNA)’

let-7, miR-1,

miR-15, miR-

16, miR-181,

miR-375

None tested [138]

Metastatic gastric

cell line

Profiled microRNA expression by microarray in

exosomes isolated from gastric cancer cells. let-7

microRNA family was enriched in exosomes

let-7 family None tested [169]

Co-culture of

IL-4-activated

macrophages and

breast cancer cells

miRNAs can be transferred from macrophages to breast

cancer cells. miR-223 released by macrophages was

found in MCF7 and MDA-MB-231 cells and promoted

invasion

miR-223 Mef2c-β-catenin pathway [170]

Mouse dendritic

cells

Exosomal microRNA from dendritic cells can be

transferred to a recipient dendritic cell and repress

microRNA target mRNAs in the acceptor cell

miR-148a,

miR-451 Luciferase reporter containing

tandem microRNA target sequences [171]

Leukemia cells and

endothelial cells

Leukemia cells released microRNAs from the miR-17-92

cluster and were taken up by human umbilical vein

endothelial cells (HUVECs) and repressed a target

mRNA. Did not affect the growth of HUVEC cells, but

did enhance cell migration and tube formation

miR-92a Integrin α5 [172]

Metastatic rat

adenocarcinoma

cells

Exosomes were preferentially taken up by lymph node

stroma cells and lung fibroblasts. The transferred

microRNAs affected mRNA translation of many genes

miR-494, miR-

542-3p

Cadherin-17, proteases, adhesion

molecules, chemokine ligands, cell

cycle- and angiogenesis-promoting

and oxidative stress response genes.

[173]

Renal cancer

stem cells

Microvesicles were secreted from human renal cell

carcinoma that could trigger angiogenesis and pre-

metastatic niche formation in a severe combined

immunodeficient (SCID) mouse model

miR-92, miR-

141, miR-29a,

miR-650, miR-

151, miR-19b,

miR-29c

Increase in VEGFR1 and MMP-9

expression [174]

20

the development of the infant immune system [167]. Further studies are required to

determine whether specific intracellular miRNAs are enriched in exosomes. Recent

evidence supporting specific miRNA loading of exosomes was the identification that

different miRNA content depended on the maturation stage of the DCs [171], and that cis-

acting elements may target specific RNAs into exosomes [175]. Interestingly, the

identification of a ‘zipcode’ contained within the 3’ region of mRNA has been identified

that leads to incorporation of specific mRNAs into EVs [176]. In-depth characterization of

mRNA and miRNA signatures within exosomes from cancer cells may lead to their use as

clinical biomarkers and give insight into their contribution in disease progression.

1.4.7 Exosome targeting and uptake

It is understood that exosomes are key mediators of intercellular communication and,

therefore, the molecular mechanisms governing exosome recognition and uptake by

recipient cells have gained much attention. An overview of possible modes of exosome

interaction and uptake by recipient cells is given in (Figure 1-3) [32].

Differences in exosomal tetraspanin complexes appear to influence target cell interaction in

vitro and in vivo, possibly by modulating the functions of associated integrin adhesion

molecules [177]. For example, exosome capture by dendritic cells was reduced by 5-30%

by co-incubation with blocking antibodies specific for various integrins, adhesion

molecules or tetraspanins [178]. Other membrane proteins also reported to be important in

targeting exosomes to recipient cells include intercellular adhesion molecule 1 (ICAM-1)

and milk fat globule-epidermal growth factor VIII protein (MFGE-8) [179, 180]. In

addition to membrane proteins, the delivery efficiency of exosomes to cells is reported to

be directly related to rigidity of exosomes that contain lipids including sphingomyelin and

N-acetylneuraminyl-galactosylglucosylceramide (GM3) [181].

In addition to receptor-mediated interactions leading to activation of cell signaling

pathways [25], exosomes may also be internalized. Escrevente and colleagues

demonstrated that fluorescently-labeled ovarian cancer cell-derived exosomes were

21

Figure 1-3. Exosome targeting and uptake

Following release by parent cells, exosomes are shown to communicate and elicit an effect

with recipient cells by (i) receptor-mediated interactions involving tetraspanins and

integrins [177], (ii) membrane fusion [181], and (iii) endocytic uptake [182] by

phagocytosis [183] and macropinocytosis [184].

22

internalized by clathrin-mediated endocytosis in ovarian cancer cells [182], while Feng and

colleagues reported exosomes being internalized via phagocytosis in monocyte-derived

macrophages [183]. Fitzner and co-workers also showed that selective transfer of exosomes

from oligodendrocytes to microglia could occur by macropinocytosis [184], a process

whereby bending of single surface lamellipodia gives rise to circular curved ruffles which

are finally sealed to form discrete endocytic vacuoles [184]. Exosome uptake has been

reported to be a temperature dependent process. For example, exosomes from rat

pheochromocytoma (PC12 cells) were internalized by resting PC12 cells by actin-

dependent endocytosis [185]. Exosomes were shown to be internalized by lipid raft

mediated endocytosis in HUVECs – this process was subsequently negatively regulated by

caveolin-1 [186, 187].

In another study, exosomes were shown to fuse with the plasma membrane in a pH

dependent manner [181]. For example, exosomes purified from metastatic melanoma cell

culture medium were labeled with a self-quenching lipid fluorescent probe, R18, and shown

to fuse at the plasma membrane of recipient melanoma cells [181]. Interestingly, fusion

efficiency was higher with exosomes released from metastatic cells compared to those

derived from primary melanoma cell tumors or normal peripheral blood mononuclear cells

[181]. A mechanism of exosome uptake involving a combination of fusion and endocytosis

has also been reported [171]. Montecalvo and colleagues demonstrated that dendritic cell-

derived exosomes were shown to fuse in ‘2-step event’ consisting of exosome hemifusion

with the cell membrane, followed by endocytosis and complete fusion of the exosomes with

the limiting membrane of the phagosome [171].

Although there are numerous studies demonstrating exosome uptake by recipient cells, very

little is known about the specific target recognition motifs in exosomes. This highlights the

need for in-depth characterization of exosome membranes to identify barcode signatures

that allow and signal for their recognition and uptake by specific recipient cells. Following

exosome uptake there is limited data showing how exosomes traffic intracellularly to exert

their biological effect.

23

1.4.8 Exosome function in the tumor microenvironment and cancer progression

In addition to cancer cells, the tumor microenvironment comprises endothelial cells,

supporting pericytes, fibroblasts, and both innate and adaptive infiltrating immune cells

[188]. Interestingly, as cancer cells accumulate genetic mutations, they can secrete

molecules (and vesicles) that can reprogram normal stromal cells to serve the budding

neoplasm; this process ultimately enables cancer cells to invade adjacent tissues and

disseminate to distant loci [189]. Recently, there has been significant interest in

understanding the involvement of exosomes in the tumor microenvironment and cancer

progression [190]. For example, exosomes have been reported to be involved in multiple

aspects of cancer progression including causing immune suppression [191], stimulating

angiogenesis [192], modulating stromal cells [193], promoting metastatic ability [194], and

establishing the pre-metastatic niche [195] through the transfer of oncogenic material in an

intra- and inter-cellular circuitry [196] (Figure 1-4, Table 1-3).

It has been reported that tumor-derived exosomes bear immunosuppressive molecules and

are implicated in immune suppression [191]. Further, it was demonstrated that exosomes

can inactivate T lymphocytes, or inhibit the differentiation of precursors to mature antigen

presenting cells to suppress immune surveillance, thereby assisting cancer progression

[191]. Angiogenesis can also be influenced by exosomes [192]. For example, melanoma

exosomes were observed to rapidly stimulate the production of endothelial spheroids and

sprouts in a dose-dependent manner [192]. These exosomes elicited paracrine endothelial

signaling by regulation of inflammatory cytokines, including IL-1α, TNFα and VEGF,

suggesting that exosomes can promote endothelial angiogenic responses and contribute to

tumor metastatic potential [192]. Tumor-derived exosomes were also shown to express

membrane TGF-β and activate the SMAD-dependent signaling pathway in stromal

fibroblasts, resulting in myofibroblast differentiation [193]. Myofibroblasts are enriched in

solid tumors and represent an altered stroma that usually supports tumor growth,

vascularization, and metastasis [193].

24

Figure 1-4. Extracellular vesicles (EVs) are components of the intra- and intercellular

oncogenic signaling circuitry

Oncogenic mutations (onc) and the resulting synthesis of altered mRNA and oncoproteins

(ONC) result in perturbations of the intracellular signaling circuitry (SIGNAL) in a primary

tumor. EVs containing onc are released into the extracellular space, and sometimes further

into circulation, enabling exposure of stromal or indolent cells to transforming onc

molecules. Upon uptake of tumor-derived EVs, oncogenic mutations may affect cellular

signaling which can contribute to increased proliferation [197] and invasiveness [198].

Within the circulation, EVs can educate bone marrow-derived cells (BMDC onc) resulting

in mobilization of BMDC onc to lymph nodes to assist pre-metastatic niche formation

[195]. EVs alone can also home to the lymph node for pre-metastatic niche formation

[119]. Figure adapted from [196] using Sevier Medical Art.

25

Table 1-3. Examples of EV-mediated transfer of oncogenic/transforming molecules. Adapted and updated from [196].

Transforming

molecule Vesicle Observations Ref.

Protein/EGFRvIII EVs Glioblastoma EVs transfer EGFRvIII protein to indolent cancer cells causing oncogenic activity

including activation of transforming signaling pathways MAPK and Akt [199]

mRNA/EGFRvIII EVs Glioblastoma EVs containing EGFRvIII oncogene transfer functional mRNA causing stimulation

of tubule formation by endothelial cells [165]

Protein/EGFR EVs Cancer EVs transfer oncogenic EGFR to endothelial cells causing autocrine reprogramming of

VEGF [200]

Protein/MyrAkt1 EVs Cancer EVs transfer oncogenic Akt to recipient cells causing rapid phosphor-tyrosine and Akt

pathway signaling stimulating proliferation and migration [62]

Protein/tTG and FNa EVs Cancer EVs transfer tissue transglutaminase and fibronectin to NIH3T3 fibroblasts causing

tumorigenic conversion [201]

DNA/retrotransposona EVs Cancer cell-derived EVs transfer retrotransposons to endothelial cells [202]

Protein/LMP1 Exosomes Cancer exosomes transfer Epstein-Barr virus latent membrane protein 1 causing activation of

ERK and AKT signaling pathways in recipient epithelial cells [203]

Mutant KRAS Exosomes Exosomes containing mutant KRAS can cause enhanced three-dimensional growth of wild-type

KRAS-expressing non-transformed cells [197]

Wnt11 Exosomes Fibroblast-derived exosomes mobilize autocrine Wnt-PCP signaling in breast cancer cells to

stimulate invasive behavior and metastasis in animal models [198]

Wnt3a Exosomes Exosomes containing Wnt3a can induce Wnt signaling activity in target cells [105]

Tyrosine kinase MET Exosomes Exosomes expressing tyrosine kinase MET increased the metastatic capability of bone marrow

cells [195]

a Oncogenic activity not directly demonstrated

26

The transfer of metastatic activity has also been attributed to exosomes. Melanoma-derived

exosomes were shown to promote epigenetic transfer of metastatic activity in murine

models. Highly metastatic BL6-10-derived exosomes were transferred to poorly metastatic

melanoma F1 cells to express Met 72 tumor antigen [194]. The subsequent Met 72-positive

F1 tumor cells showed increased metastatic activity by growing numerous lung tumor

colonies following their intravenous injection into mice [194]. In addition to exosomes,

microvesicles (MVs) derived from platelets are also capable of enhancing the invasive

potential of breast cancer and induce angiogenesis and metastasis in lung cancer [204, 205].

In a further study, transfer of oncogenic potential to a recipient cell through activation of

MAPK and Akt signaling pathways highlights new mechanisms of intercellular

communication of oncogenic material via MVs in the tumor microenvironment [165, 200].

It is well recognized that tumors preferentially home to specific metastatic sites. For

example, metastasis of melanoma often homes to the lung and brain [206], while colorectal

cancer to the liver [207]. Interestingly, it has recently been shown that malignant cancers

use secreted signals to effectively prime a specific site for metastasis in a ‘seed-soil’

manner by preparing a fertile environment [19, 208]. Melanoma-derived exosomes injected

into the foot pads of mice preferentially colonized and conditioned sentinel lymph nodes by

the introduction of genetic molecular signals for cell recruitment, extracellular matrix

deposition, and vascular proliferation [119]. Further, studies by Peinado and colleagues

have demonstrated that tumor exosomes can horizontally transfer genetic material to bone

marrow-derived cells (BMDCs) resulting in upregulation of the MET oncoprotein. This led

to mobilization of BMDCs to home to the lung and lymph node to support tumor

vasculogenesis, invasion and metastasis [195] (Figure 1-4).

Interestingly, the ability to interfere with molecular pathways responsible for exosome

formation and release may in turn lessen their contribution to disease progression [209].

This could potentially involve the modulation of a key protein required for their formation

(i.e., Rab27a) [103, 210]. Additionally, it has previously been demonstrated that regulation

of Rab GTPase-activating proteins (GAPs) can alter Rab35 activity which leads to the

accumulation of endosomal vesicles and impairment of exosome secretion [101].

27

Attenuation of exosome release from cancer cells can also be accomplished by

administration of neutral sphingomyelinase inhibitors [96], or treatment with the blood-

pressure-lowering drug amiloride [211]. Alternatively, the inhibition of EV uptake by

recipient cells can be altered using diannexin to block phosphatidylserine, which is an

important molecule required for cell adhesion [200, 212]. This experimental process was

shown to reduce the growth rate of human glioma xenografts in mice [200].

1.4.9 Exosome therapeutics

The ability of exosomes to transport bioactive molecules has driven research to improve

their large-scale isolation and characterization for use in translational nano-medicine [213]

(Figure 1-5). Current drug delivery vehicles, including viral and bacterial vectors, have

limited efficacy following repeat administration. These vectors may also elicit cellular

toxicity by causing insertional mutagenesis in host cells [214] (Figure 1-6). Alternatively,

exosomes may have advantageous properties over other drug delivery vehicles as they have

limited immunogenicity when derived from ‘self’, increased stability and efficient delivery

of cargo delivery abilities.

A groundbreaking discovery in exosome therapeutics was the ability of exosomes to

function as natural cell-derived nano-carriers that can cross biological barriers [215, 216].

For example, exosomes may be used to encapsulate and protect exogenous RNA

interference (RNAi) and miRNA regulatory molecules in addition to other single stranded

oligonucleotides [216]. Exciting new work by Alvarez-Erviti and colleagues have

demonstrated that following systemic injections, neuron-specific peptide targeting allowed

effective delivery of exosomes loaded with siRNA into the mouse brain [216]. As a result

of siRNA contained within exosomes, they observed the knock-down of BACE-1, a

therapeutic target in Alzheimer’s disease. Additionally, injections of exosomes (150 µg)

appear to be well-tolerated with little or no immunogenicity or toxicity, even following

repeat (n=3) administration [210, 216].

28

Figure 1-5. Functional roles and therapeutic targeting of extracellular vesicles (EVs)

EVs containing protein and RNA may be exploited for functional roles in therapeutics. The

presence of MHC molecules on exosomes enables efficient antigen presentation to suppress

tumor growth [218]. Further, EVs can cause immune modulation by protecting T cells from

activation-induced cell death [219], or cause immune suppression by EV/FASL-mediated T

cell apoptosis [220]. Regenerative medicine applications of EVs have also been

demonstrated. For example, EVs derived from mesenchymal stem cells are shown to reduce

myocardial ischemia/reperfusion injury by mediating cardioprotection and reducing infarct

size in mouse models [46, 221]. EVs may also be endogenously loaded using recombinant

expression of targeting ligands or overexpressing therapeutic RNAs [210, 216].

Alternatively, EVs may be exogenously loaded with siRNA by electroporation prior to

being used in therapeutic treatments [222]. Adapted from [210].

29

Figure 1-6. Advantages and drawbacks of siRNA delivery by exosomes, viruses and

lipid nanoparticles

Major advantages are indicated in green, disadvantages in red. (a) Dendritic cell-derived

exosomes can directly fuse with the plasma membrane through CD9 tetraspanin

interactions with surface glycoproteins on the target cell to ensure direct cytosolic delivery

of siRNA [223]. (b) Viruses mediate direct delivery of vectors encoding short hairpin (sh)

RNA to the cytosol. However, viruses are prone to clearance by opsonins, complement,

coagulation factors and virus-specific antibodies in the systemic circulation. Although

genomic insertion of the shRNA-encoding vectors allows long-term RNAi, a potential risk

of this strategy is genetic deregulation and oncogene activation by insertional mutagenesis

[214]. (c) Internalization of lipid nanoparticles occurs by endocytosis. Endosomal escape is

required for cytoplasmic delivery by fusion with, or rupture of, the endosomal membrane

using viral or bacterial proteins. This escape is often hampered by the man-made nature of

lipid nanoparticles The direct delivery of RNA into endosomes can lead to unwanted

TLR7/8 activation [224]. Adapted from [225].

30

The ability of exosomes to activate T cells has prompted their potential use as vaccines

[217]. Chaput and colleagues demonstrated that peptide loaded dendritic cell-derived

exosomes could mediate tumor rejection in mice [217]. Further, exosomes secreted from

Gm-CSF/IL-4 bone marrow dendritic cells were shown to eradicate established murine

tumors in mammary TSA carcinoma, P815 mastocytoma and MCA205 sarcoma models

[217]. These immunogenic properties have allowed the first phase I clinical trial of

dendritic cell-derived exosomes to treat metastatic-tumor-bearing patients [226]. Good

manufacturing practice (GMP) was used to isolate large-scale pharmaceutical-grade

exosomes and directly load peptides on functional MHC complexes. Autologous exosomes

were inserted into patients to observe an increase in NK cells, thus highlighting their

potential use as a vaccine [226]. It was also observed that increased levels of hsa-let-7b and

hsa-let-7g miRNAs in mesenchymal stem cell-derived exosomes were able to treat

myocardial infarction [227, 228] and provide protection against myocardial ischemia [46].

A list of other therapeutic applications of exosomes is outlined in (Table 1-4).

An alternative to tailored biological exosomes are artificial exosome ‘mimetics’, which are

being developed for clinical immunology [229]. It is hypothesized that the incorporation of

only key exosomal lipids, protein and nucleic acids will lead to the creation of a viable

therapeutic vehicle [229]. Using nanotechnology-based approaches, targeted and traceable

in vivo super paramagnetic artificial exosomes (termed liposomes) have been developed

[230, 231]. Liposomes coated with a MHC class I/peptide complexes and a range of ligands

for adhesion were tracked in vivo using fluorescence and magnetic resonance imaging and

shown to activate and expand functional T cells [232]. Interestingly, Delcayre and

colleagues have coined ‘exosome display technology’ which utilizes the fusion of antigens

onto the C1C2 domain of lactadherin [233], allowing for presentation of antigens to the

immune system. Further, this technology could also be incorporated into liposomes.

Inclusion of tetraspanin proteins in exosome mimetics has been proposed due to their

ability to functionally mediate fusion and cell-cell adhesion [234]. Challenges associated

with liposome creation and usability include their sensitiveness to clearance by opsonins,

complement and coagulation factors in the blood [225] (Figure 1-6). Additionally,

although lipid assembly is understood, incorporation of functional membrane proteins into

31

Table 1-4. Therapeutic applications of exosomes

Application Therapeutic observation Ref.

Antigen

presentation

Dendritic cells pulsed with tumor-related peptides are release on exosomes to suppress tumor

growth [218]

Immune

modulation

Exosomes cause immune activation by protecting T cells from activation-induced cell death [219]

Immune suppression by exosome/FASL-mediated T cell apoptosis [220]

Tissue repair Manufacturing of therapeutic exosomes following immortalizing of mesenchymal stem cells [236]

Exosomes secreted by mesenchymal stem cells reduces myocardial ischemia/reperfusion injury [46, 221]

Gene therapy

Transfection of miR-150 in monocytes released miR-150 containing functional exosomes [237]

siRNA loaded into exosomes by electroporation is shown to be functional by knocking down

BACE1 in mouse brain, a therapeutic target in Alzheimer’s disease

[210,

216]

Human plasma-derived exosomes were electroporated with siRNA to deliver a functional RNAi

response in human monocytes and lymphocytes in vitro [222]

Adendo-associated viruses (AAVs) displaying viral caspid proteins contained within ‘vexosomes’

can deliver cargo to recipient cells [238]

Tissue targeting

Plasmid expressing exosomal proteins fused to brain specific peptide (rabies virus glycoprotein,

RVG) is displayed on exosomes and delivered across blood brain barrier [216]

Exosomes containing the transmembrane domain of platelet-derived growth factor receptor fused

to the GE11 peptide homed to EGFR-expressing tumor cells to deliver let-7a anti-tumor miRNA [239]

32

liposomes has proven to be a more difficult process [235]. Overall, it is hypothesized that

these challenges can be overcome, and that liposomes may be produced on a large scale to

act as pharmaceutical-grade delivery vehicles that carry similar therapeutic properties to

exosomes [229].

1.4.10 Exosome and microvesicle diagnostics

The capacity of exosomes and other EVs to protect genomic material makes them a

promising source of potential biomarkers. One assay reaching clinical utility is the use of

mRNA profiling of serum-derived EVs to identify glioblastoma-bearing patients [240]. It

has also been reported by Mitchell and colleagues that miR-141 levels from serum,

proposed to be contained within exosomes, can robustly distinguish prostate cancer patients

from healthy controls [241]. Additionally, there are several companies that have begun

development of exosome based diagnostics. New York based biotechnology company,

Exosome Diagnostics (http://www.exosomedx.com/), is performing the characterization of

mRNA and miRNA contained within exosomes (and other EVs) to develop biofluid-based

molecular diagnostic tests for use in personalized medicine. Caris Life Sciences

(http://www.carislifesciences.com/) is also developing methods for isolating and

characterizing blood based circulating microvesicles. Finally, HansaBioMed

(http://www.hansabiomed.eu/) has developed an ELISA based platform called Exotest

which utilizes CD63, CD9 and CD81 antibodies immobilized on 96-well plates in an effort

to capture, quantify and characterize exosomes. It is evident that there is significant interest

in this field of research, with several companies having initiated programs that exploit the

diagnostic potential of exosomes.

1.4.11 Exosome isolation strategies

A major challenge in the field is to improve and standardize methods for exosome isolation

and characterization. Current strategies for isolating exosomes differ significantly based

33

upon their various physical attributes. These strategies include ultracentrifugation [242],

sucrose or OptiPrep™ gradients relying on vesicle density [35, 243] and immunoaffinity

approaches relying on a specific exosome surface protein target (Table 1-1). This diversity

in purification methodology has led to variations in sample homogeneity and confusion

with respect to nomenclature within the field (e.g., exosomes vs. exosome-like vesicles). As

different isolation strategies may have varying fractionation capabilities, it is critical to

evaluate these techniques and their abilities to isolate a homogeneous population exosomes

These isolation strategies are described in turn:

Differential ultracentrifugation involves removal of intact cells and bulky cell debris by low

speed centrifugation (e.g., 500 g, 2,000 g) followed by high-speed centrifugation (e.g.,

100,000 g) to sediment exosomes [242]. This widely-used methodology, however, can

potentially result in co-sedimentation of protein aggregates and/or other membranous

vesicles which may be pelleted during ultracentrifugation [7].

Ultracentrifugation using a linear sucrose gradient has been developed to isolate exosomes

based on their density of 1.08-1.22 g/cm3 [35, 242]. For clinical purposes, the preparation

of GMP-grade exosomes can be prepared using a combination of ultrafiltration,

ultracentrifugation and a 30% sucrose/deuterium (D2O) (98%) cushion (1.21 g/cm3) [244].

Interestingly, sucrose gradients have been inefficient in separating exosomes from HIV-1

particles due to similarities in their size and buoyant density [245]. To overcome this

problem, Cantin and colleagues describe the use of iodixanol (OptiPrep™) 6-18% gradients

to separate HIV-1 particles and apoptotic vesicles from exosomes [243]. Iodixanol is a non-

ionic, water-soluble radiographic contrast medium [246] that may be beneficial for future

clinical use with exosomes due to its physical tolerance and low incidence of adverse

effects when injected and used as an x-ray contrast medium [247].

Immunoaffinity capture (IAC) is another widely used method for exosome isolation. This

strategy is based on the use of specific exosome-surface protein directed antibodies coupled

to magnetic beads (Table 1-1). IAC can be a powerful tool for exosome isolation, provided

a specific exosomal cell surface protein can be identified that discriminates the exosome of

34

interest from other EVs present in the biological matrix. Immuno-isolation of exosomes has

been performed to purifying antigen presenting cell-derived exosomes [40], CD45-

exosomes from Jurkat T cells [39], HER2-positive exosomes from breast adenocarcinoma

cell lines and ovarian cancer patient-derived ascites [37], A33-positive exosomes released

from colon carcinoma cancer cells [44], and EpCAM-positive exosomes from the sera of

lung cancer [42] and ovarian cancer [43] patients. This technique is limited by the number

of known cell, and thus, specific exosome membrane proteins. A final strategy, size-

exclusion HPLC, has also been successfully employed for the large-scale production of

microparticles from human mesenchymal stem cells [45, 46], and human saliva [47].

1.5 Research models used in this thesis

Cell lines that are readily cultured in vitro are often used as surrogates for tumors as they

can reflect their tissue of origin [248]. Cell lines have an advantage over cancer biopsy

tissue in that they are homogeneous as compared to heterogenic solid tumors which, in

addition to cancer cells, contain multiple other cell types [249]. Cell lines also provide an

unlimited auto-replicative source that can be easily handled, may contain specific known

mutations, or can be genetically manipulated [250]. In this thesis, two different cell line

models were utilized: the colorectal cancer (CRC) cell line LIM1863 [251] and an EMT

cell model created from epithelial Madin-Darby canine kidney (MDCK) cells [250].

1.5.1 Colorectal cancer model

1.5.1.1 Colorectal cancer

The structure of the adult colon, or large intestine, is composed of three concentric tissue

layers. The outer layer consists of several sheets of smooth muscle that execute the

rhythmic peristaltic movements of the intestine [252]. The space between the outer muscle

and the inner epithelial layer is filled by connective stromal tissue that contains numerous

35

blood and lymph vessels, nerve fibers, and various immune cells. The inner luminal surface

consists of a single-cell layer comprising differentiated epithelial cells that mediate

functions of the intestinal epithelium [252]. Finally, finger like invaginations of this

epithelial sheet into the underlying connective tissue form ‘colonic crypts’, the basic

functional unit of the colon [253]. CRC is defined as the development of malignant tissue

in the wall of the colon or rectum and is caused by a series of stepwise genetic changes

(Figure 1-7) [254].

1.5.1.2 CRC statistics and subtypes

About 608,000 deaths from colorectal cancer are estimated to occur worldwide each year.

This accounts for 8% of all cancer deaths, making it the fourth most common cause of

cancer-related death [255]. According to the Australian Institute of Health and Welfare,

CRC was the most commonly registered form of cancer in Australia in 2009 with 14,410

new cases and 3,982 deaths, making up 9.3% of all cancer deaths in Australia [256]. The

Dukes’ staging system (A-D) is commonly employed to stage progression and it has been

shown that CRC can be effectively treated if detected early [257, 258]. Based on 5-year

survival rates, patients presenting with Dukes’ stage A CRC have >90% chance of survival,

while more commonly, patients presenting with late stage Dukes’ stage D CRC have 5-year

survival rates <10% [259]. A lack of early detection occurs due to low patient compliance

with current testing modalities coupled with the limited sensitivity of current detection

techniques [260].

Typically there are two forms of CRC, termed sporadic and hereditary. Sporadic CRC

refers to random occurrences in which family history is absent [261]. This accounts for

approximately 70% of all CRCs [262]. Sporadic CRC arises from the accumulation of

multiple genetic mutations in somatic cells over an extended period of time and therefore a

late age onset is commonly observed [263]. Hereditary CRC is divided into two classes,

familial adenomatous polyposis (FAP) and hereditary non-polyposis colorectal cancer

36

Figure 1-7. Model of intestinal cancer initiation and progression

Crypt base columnar cells (green) and adenoma cell (red) are located at the base of the

crypt. Loss of adenomatous polyposis coli (APC) is known to be an initiating event behind

early adenoma formation (red-hued) where cells begin to migrate up the transit amplifying

compartment (blue). The accumulation of additional oncogenic mutations (shown in italics)

drives tumor progression from a benign intermediate and late adenoma to an invasive

carcinoma. Other cell types include entero-endocrine cell (yellow), goblet cell (orange), and

tuft cell (purple). Adapted from [254].

37

(HNPCC), which together make up ~10% of all colon cancers [262]. FAP is characterized

by germline mutations in the adenomatous polyposis coli (APC) gene, resulting in the

presence of numerous adenomatous polyps [264]. The APC gene encodes a ~300 kDa

protein which selectively localizes to the nucleus and cytoplasm [265]. Under normal

circumstances, APC functions as a tumor suppressor by mediating the downregulation of β-

catenin. When APC is mutated, β-catenin accumulates in the nucleus and binds to

transcription factors to promote cellular proliferation [266]. HNPCC is caused by mutations

in one or more DNA mismatch repair genes, most commonly hMSH2 and hMLH1 [267].

Other less frequently mutated genes are hMSH6 and hPMS2 [267]. These mutations are

autosomal dominant, resulting in aggressive tumor growth in the right side colon in 70% of

cases [268].

1.5.1.3 Colorectal cancer cell line LIM1863

The human LIM1863 (Ludwig Institute Melbourne) cell line was derived from a tumor

sample taken from the ileocecal valve, between the ileum of the small intestine and the

caecum of the large intestine, from a 74-year-old Caucasian female [251]. This sporadic

carcinoma extended through the full thickness of the muscle wall and mutational analysis

revealed truncated APC (stop at aa 1506) and mutated p53 (aa 234 Y to H) genes. In vitro,

the LIM1863 cell line grows as a suspension culture of floating organoids, clusters of cells

resembling crypt like structures, which are morphologically and functionally organized.

LIM1863 organoids comprise two polarized cell types, columnar and goblet cells

surrounding a central lumen (Appendix A-3). Along with continuous cell proliferation and

organoid formation, dead cells are continuously shed. This organization is very similar to

that of in vivo, where intestinal epithelial cells form intestinal crypts [251]. Significantly,

single cell cloning was used to isolate 29 subclones of the LIM1863 cell line, all of which

grew to display the same phenotype and percentage of morphological cell types as the

parent line, thus indicating ‘stemness’ like properties of the cells [269].

38

Isolating exosomes from CRC cell lines may allow insights into their contribution to the

disease. As the LIM1863 cell line grows as floating organoids in suspension, it can be

cultured on a large scale, this is beneficial for characterization studies as large amounts of

exosomes can be isolated. Further, as LIM1863 cells are polarized, the cell model can be

used to understand exosomes in the context of apical and basolateral trafficking and release.

1.5.2 Epithelial-mesenchymal transition (EMT) model

Epithelial-mesenchymal transition is a process by which epithelial cells lose polarity and

cell adhesion characteristics and acquire migratory and invasive properties of mesenchymal

cells [270, 271]. For many decades, EMT has been recognized as a cell phenotypic switch

in developmental biology, embryogenesis, tissue remodeling and wound healing [272,

273]. During these processes, a series of molecular and cellular changes occur allowing

epithelial cells to escape from their rigid structural constraints to accommodate a phenotype

more amenable to cell migration and invasion (Figure 1-8) [274]. These changes include

reduced cell-cell adhesion, loss of cell polarity, elongated spindle-shaped morphology, an

invasive phenotype, induction of mesenchymal-specific proteins, ECM alterations, and

increased migratory capacity (Appendix A-4) [275]. The activation of an EMT program

has also been proposed as the critical mechanism for the acquisition of malignant

phenotypes by epithelial cancer cells [189, 276, 277] (Figure 1-9).

1.5.2.1 EMT during cancer progression

Early growth of primary epithelial cancers is characterized by excessive cell proliferation

and angiogenesis [189]. Following these changes, local invasiveness, initially through the

basement membrane, a dense specialized ECM barrier between the epithelial cell layer and

underlying stroma, is thought to begin the multi-step process that leads to metastatic

dissemination [270]. The ability of cells to penetrate through the basement membrane may

be attributed to the increased expression of secreted proteases [278].

39

Figure 1-8. Epithelial-mesenchymal transition

Epithelial-mesenchymal transition (EMT) is a morphogenetic process whereby immotile

polarized epithelial cells become migratory mesenchymal cells. This highly conserved

cellular process involves dissolution of cell-cell junctions, i.e., tight junctions (dark blue),

adherens junctions (light blue) and desmosomes (green), and loss of apical-basolateral

polarity. Cells will then acquire a mesenchymal phenotype, this is characterized by actin

reorganization, stress fiber formation (red), a ‘leading-edge’ polarity, migration and

invasion phenotypes. Modified from [274].

40

Figure 1-9. EMT during tumor progression and metastasis

Normal epithelia proliferate to form a local adenoma. Genetic alterations and epigenetic

events produce a carcinoma in situ. Loss of cell-cell adhesion and increased invasiveness,

possibly by EMT, drives cells through a fragmented basement membrane where they can

intravasate into lymph or blood vessels. It is proposed that the invasive tumor cells are

passaged to distant sites where they repopulate and form secondary metastases. Modified

from [276].

41

These proteases include matrix metalloproteinases (MMPs) and a disintegrin and

metalloproteinases (ADAMs) that function at cells invasive leading edge to proteolyse

collagens and laminins to assist cell migration [278, 279]. Cells expressing a mesenchymal

phenotype are typically at the invasive tumor front and are considered to ultimately follow

an invasion-metastasis cascade [270]. Interestingly, at distant sites colonized mesenchymal

cells are reported to revert back and resemble their originating tumor phenotype. This

process is known as mesenchymal-epithelial transition (MET), whereby mesenchymal

characteristics are lost in favor of epithelial traits [270]. It is proposed that this may occur

as signals originally required for driving and maintaining the mesenchymal phenotype in

the primary tumor environment are no longer apparent [270, 280].

1.5.2.2 Molecular markers of EMT

The architecture of epithelial cells is controlled by the expression of various proteins

including E-Cadherin, which are associated with cell-cell contact [281]. E-cadherin is a

transmembrane protein required for the establishment and maintenance of adherens

junctions and is fundamental to the epithelial phenotype [282-284]. The suppression of E-

cadherin diminishes epithelial cell adhesion, alters polarity and is associated with increased

cell invasiveness during cancer metastasis [285]. Following E-cadherin suppression, the

mesenchymal cell-cell adhesion molecule N-cadherin becomes highly expressed, a process

also known as the ‘cadherin switch’ [284].

In addition to E-cadherin, the modulation of many other components has been reported

during EMT [286]. These alterations include downregulation of epithelial adhesion

molecules (e.g., claudins, occludin and the ZO-family) and upregulation of mesenchymal

molecules (e.g., vimentin, fibronectin and α5β1 integrin) [287]. Additionally, during EMT

cytokeratin networks are replaced by vimentin-rich intermediate filaments, leading to

enhanced cell migration and invasiveness [288].

42

1.5.2.3 Signaling pathways and inducers of EMT

The induction of EMT can occur via a complex array of extracellular signaling

components, cytoplasmic molecules and nuclear effectors of transcription [289]. The EMT

process may be initiated by growth factors such as fibroblast growth factor (FGF),

epidermal growth factor (EGF), TGF-β, TNF-α, and bone morphogenic protein (BMP)

[270, 290, 291]. These initiating events can activate signaling pathways of cytoplasmic

effectors including SMADs, glycogen synthase kinase 3 β (GSK-3β), mitogen-activated

protein kinase (MAPK), and reactive oxygen species leading to activation of transcriptional

regulators [289]. It is understood that once expressed and activated, EMT-inducing

transcription factors, notably, smad interacting protein (SIP-1), Lymphoid enhancer factor

(LEF), Snail, Slug, Zinc finger E-box binding homeobox 1 (ZEB1), Twist, Goosecoid, and

Forkhead box protein C2 (FOXC2) can act pleiotropically to choreograph the complex

EMT program [276, 289, 292].

In addition to extracellular signaling components, the transfection of oncogenes into

epithelial cells can also induce EMT [293]. Ras is a small GTPase that is involved in many

signaling pathways that regulate cellular processes including cell proliferation,

differentiation, and cell motility [294]. Active Ras-GTP initiates a MAPK phosphorylation

cascade that facilitates the altered expression of transcription factors including Jun, Fos and

Slug, which can induce EMT [282]. Additionally, it was found that oncogenic Ras

signaling downregulated Rac activity and upregulated Rho activity, leading to

rearrangement of the actin cytoskeleton during EMT [295]. Further, the Ras effector

pathway, Raf/MAPK, was shown to transform Madin-Darby canine kidney (MDCK) cells

towards a mesenchymal phenotype [296]. Examination of MDCK cells following

transformation with oncogenic H-RasV12 demonstrated a loss of polarity, detachment from

the substratum, and multilayer growth [297].

43

1.5.2.4 Madin-Darby canine kidney (MDCK) and 21D1 cell lines

The MDCK cell line was derived from the kidney of a normal male cocker spaniel in 1958

[298] and is one of the most-widely used and best characterized epithelial cell lines

available [299]. MDCK cells appear round and resemble cobble-stone like structures. They

grow as a polarized monolayer, with strong cell-cell contact and show expression of

epithelial cell junction markers E-cadherin and ZO-1 [250].

MDCK cells and oncogenic H-Ras-transformed MDCK cells, termed 21D1 cells, are

together recognized as an epithelial-mesenchymal cell model [250]. 21D1 cells were

generated by transfecting pcDNA3 containing v-Ha-Ras under the control of a CMV

promoter and a SV40-driven neomycin resistance plasmid gene into MDCK cells [250]. In

this study, 21D1 cells were shown to have reduced cell-cell contact and display spindle

shaped morphology following transformation [250] (Figure 1-10). Confocal microscopy

images of mesenchymal 21D1 cells showed reduced expression of E-cadherin and ZO-1

between cell junctions, and increased expression of the mesenchymal marker vimentin.

Wound healing and migration assays also revealed that 21D1 cells have enhanced motility

and invasive capabilities when compared to MDCK cells [300].

1.5.2.5 Proteomic analysis of the secretome from MDCK cells following EMT

Using the EMT model previously described, the secretome of MDCK cells following Ras-

induced EMT was analyzed in the Simpson laboratory to identify differentially expressed

secreted proteins which may influence tumor cell state and invasive potential [250, 300].

Using a semi-quantitative label-free spectral counting method, common constituents of the

basement membrane including laminins and collagens were downregulated, indicating

ECM remodeling following EMT. Several proteases such as MMP-1 and kallikrein 6 were

identified to be highly upregulated after EMT (Appendix A5). The precise function of

MMP-1 in EMT was further studied by an MMP-1 siRNA knock-down assay in 21D1 cells

which led to reduced cell migration. Collectively, it is hypothesized that secreted effectors

44

Figure 1-10. MDCK and H-Ras-transformed MDCK cells (21D1) morphology

(A) MDCK cells grow as round cobblestone-like sheets. (B) 21D1 cells exhibit elongated

and spindle shaped cell morphology.

A B

45

might co-ordinate cell migration and EMT in a hierarchical manner [300].

1.5.2.6 Exosomes and EMT

While the direct involvement of exosomes in the process of EMT has not yet been studied,

pioneering work investigating exosome-like vesicles following EMT has been undertaken

[301]. It was found that A431 epithelial cancer cells adopt mesenchymal-like features upon

activation of epidermal growth factor receptor (EGFR) and repression of E-cadherin.

Exosomes were isolated from these mesenchymal-like cells and shown to contain increased

levels of tissue factor (TF) that caused pro-coagulant conversion of endothelial cells. This

indicated that cells with mesenchymal-like features release vesicles capable of altering cell

interactions within the vasculature [301]. Further, there are a significant number of studies

identifying that exosomes are able to transfer ‘oncogenic’ material to assist cancer

progression (see section 1.4.8) [196]. Although exosomes are identified to be released from

mesenchymal cells [301], a detailed proteomic characterization has not yet been performed.

It is therefore hypothesized that an in-depth proteomic analysis of exosomes isolated from

MDCK and 21D1 cell lines may define extracellular modulators contained in exosomes and

lead to a better understanding of their funtional roles.

1.6 Thesis aims

Exosomes are 40-100 nm diameter membrane-derived nanovesicles of endocytic origin that

are released by most cell types. They are reported to have pleiotropic biological roles

including involvement in cell-cell communication. The majority of functional studies

performed on exosomes to date have involved heterogeneous or crude vesicle populations.

It is anticipated that by developing robust exosome purification strategies, and studying

precise protein profiles in homogeneous exosome populations, valuable biological insights

will be gained.

46

The overarching aim of this thesis is to undertake in-depth proteomic analyses on highly-

purified exosomes in order to gain insights into the role of exosomes in cancer biology and

the epithelial-mesenchymal transition process.

The specific aims of this thesis are:

1. To evaluate exosome purification strategies

Our capacity to gain meaningful biochemical insights from exosomes is impeded by the

fundamental difficulty in isolating exosomes to homogeneity. Chapter 2 describes a

comparative evaluation of three widely-used methods for isolating exosomes: differential

ultracentrifugation, OptiPrep™ density gradient centrifugation and immunoaffinity capture

using EpCAM monoclonal antibody bound to magnetic beads. In this study the human

colorectal cancer cell line LIM1863 [251] was used as a model. In order to compare

proteome profiles between the three isolation methods, GeLC-MS/MS and label-free MS

spectral counting was employed. Following isolation of exosomse, sample homogeneity

was based on the expression of defined stereotypical exosomal proteins reported in the

literature.

This aim is addressed in Chapter 2 and incorporates - Tauro, B. J., Greening, D. W.,

Mathias, R. A., Ji, H., Mathivanan, S., Scott, A. M., and Simpson, R. J. (2012) Comparison

of ultracentrifugation, density gradient separation, and immunoaffinity capture methods for

isolating human colon cancer cell line LIM1863-derived exosomes. Methods 56, 293-304.

2. To analyze exosome sub-populations released from LIM1863 cells

The colon cancer cell line LIM1863 [251] grows as crypt-like ‘organoid’ structures with

highly-polarized columnar and goblet cells organized around a central lumen. Chapter 3

describes a strategy that involves the sequential use of two immunoaffinity target reagents

47

(anti-A33 antibody and anti-EpCAM antibody coated magnetic beads) for isolating

different sub-populations of exosomes. This study was directed at characterizing protein

cargo contained in apical- (e.g., positive for epithelial cell marker EpCAM) and basolateral-

(e.g., positive for basolateral marker A33) derived exosome populations. A proteomic

analysis of sMVs was also undertaken to assess whether these vesicles differ from

exosomes.

This aim is addressed in Chapter 3 and incorporates - Tauro, B. J., Greening, D. W.,

Mathias, R. A., Mathivanan, S., Ji, H., and Simpson, R. J. (2013) Two distinct populations

of exosomes are released from LIM1863 colon carcinoma cell-derived organoids.

Molecular & Cellular Proteomics 12, 587-598.

3. To analyze exosomes following epithelial-mesenchymal transition

While it is becoming increasingly evident that epithelial-mesenchymal transition (EMT)

plays a critical role in cancer cell plasticity [277], it is not known whether exosomes

contribute to the EMT process. Chapter 4 describes the in-depth proteome analysis of

exosomes released from Madin-Darby canine kidney (MDCK) cells and upon oncogenic H-

Ras induced EMT (21D1 cells). This study identified exosomal proteins that were

significantly dysregulated following oncogenic H-Ras induced EMT. This finding indicates

potential functions of exosomes associated with the EMT process.

This aim is addressed in Chapter 4 and incorporates - Tauro, B. J., Mathias, R. A.,

Greening, D. W., Gopal, S. K., Ji, H, Kapp, E. A., Coleman, B. M., Hill, A. F., Kusebauch,

U., Hallows, J. L., Shteynberg, D., Moritz, R. L., Zhu, H. J., and Simpson R. J. (2013)

Oncogenic H-Ras reprograms Madin-Darby canine kidney (MDCK) cell-derived exosomal

proteins during epithelial-mesenchymal transition. Molecular & Cellular Proteomics 12,

2148-2159.

48

49

Chapter 2: Comparison of methods for isolating

exosomes

2.1 Overview

The secretome, a rich complex subset of macromolecules secreted by a cell, is a vital aspect

of cell-cell communication in eukaryotes [9]. There has been much interest in the secretome

from cancer cells due to the recognition that secreted molecules could spawn new therapies

[9, 13, 14, 302]. In addition to soluble proteins, the secretome also contains a multitude of

extracellular nano-membraneous vesicles (EVs) [303] (see section 1.2); of these, 40-100

nm diameter exosomes have been the most widely studied from a functional perspective.

However, one of the setbacks in the exosome field is that most functional studies have been

conducted using impure preparations, typically containing a mixture of EVs. Since this

situation confounds the interpretation of such studies there is a fundamental need to purify

exosomes to homogeneity. For these reasons, I have performed, and describe in this

chapter, a detailed comparison of typical exosome isolation methods currently used in the

field. Due to the scope of this chapter and instrument limitations, we were unable to include

HPLC purification of microparticles in this comparison. Size-exclusion HPLC has been

successfully employed for the large-scale production of cardioprotective microparticles

from human fetal mesenchymal stem cells for clinical purposes [45, 46], and for isolating

two distinct exosome populations from human saliva [47].

Commonly used exosome isolation methods include: (i) differential ultracentrifugation to

‘pellet’ exosomes, (ii) density-gradient centrifugation to fractionate exosomes from other

EVs by exploiting their buoyant density, and (iii) immunoaffinity capture on magnetic

beads using antibodies directed towards known exosomal membrane proteins (see section

1.4.11) [35].

50

Differential centrifugation utilizes multiple centrifugation and filtration stages to separate

organelles [304], intracellular vesicles [305], and exosomes [50] based on differences in

density. Typically, this method involves serial low speed centrifugation (500 g, 2000 g)

to remove cells and cellular debris, followed by 0.1 µm [306] or 0.22 µm [50] membrane

filtration, and ultracentrifugation between 60-100,000 g to pellet exosomes [307]. Due to

its limited resolving ability, the differential ultracentrifugation procedure can result in

exosome preparations being significantly contaminated by other classes of EVs, protein

oligomers and pathogens (e.g., adenoviruses, prions etc) [7]. For example, it has been

shown that shed microvesicles (sMVs, also referred to as oncosomes) and exosomes are

morphologically distinct, the former being heterogeneous in size (100 nm-1000 nm) and

sedimenting at lower speeds relative to exosomes [308]. Thus it is likely that EVs that

sediment at 100,000 g may contain a mixture of sMVs and exosomes.

The biogenesis of sMVs and exosomes are quite distinct. sMVs are formed by outward

budding of the plasma membrane, however, it should be noted that not all plasma

membrane proteins are incorporated into shed microvesicles, yet the topology of the

membrane remains intact. Oncogenic membrane proteins and other growth factor receptor

proteins, integrin receptors and soluble proteins such as proteases, cytokines and nucleic

acids are found in microvesicles, however, it is not clear how specific cargo proteins are

trafficked to microvesicles [49]. In contrast, exosomes are formed by endocytic

invagination of late endosomes to form multivesicular bodies (MVBs) which then traffic to,

and fuse with, the plasma membrane, releasing their cargo of exosomes (see section 1.4.1 –

1.4.2). Like sMVs, exosomes contain bioactive molecules such as oncogenic proteins,

mRNAs and miRNAs that can be horizontally transferred to recipient cells and are thought

to be implicated in tumor invasion and metastasis [165]. Whether biological differences

exist between sMVs and exosomes remains to be determined. For these reasons it is

important to purify exosomes to homogeneity (free of sMVs) in order to understand their

biological function [49, 308].

Isolation of exosomes by density gradient centrifugation involves their fractionation across

a continuous or discontinuous density gradient, i.e., sucrose or iodixanol (OptiPrep).

51

Exosomes can be purified by floatation on linear sucrose gradients (2.0-0.25 M) at densities

of 1.08-1.22 g/cm3 [35, 242]. However, sucrose gradients have been shown to be inefficient

in separating exosomes and HIV-1 particles due to their similarity in size and buoyant

densities [245]. Recently, OptiPrep™ gradients have been shown to be an alternative to

sucrose gradients due to their ability to separate exosomes from HIV-1 particles [243].

Interestingly, iodixanol (sold under the FDA approved Visipaque™ UCM198414) is

commonly used as a radiopaque contrast agent for intravascular use to diagnose coronary

angiopathy, brain, blood vessel and kidney disorders. For this reason, OptiPrep prepared

exosomes are compatible with in vivo functional studies.

Immunoaffinity capture can be used to isolate exosomes from a variety of cellular sources

(Table 1-1), however, this method is dependent on the availability of monoclonal

antibodies directed towards tissue-specific exosomal membrane markers [44]. For example,

immunoaffinity capture has been used to isolate exosomes containing A33 from epithelial

colorectal cancer cells [44], HER2-tumor exosomes from breast adenocarcinoma cell lines

[37] and ascites of ovarian cancer patients [37], Rab5b-positive exosomes from melanoma

cell lines [38], CD45-positive exosomes from Jurkat T cells [39], MHC class II-exosomes

from antigen presenting cells [40], CD63-positive exosome-like vesicles from blood [41],

and EpCAM-positive exosomes from human sera of lung cancer [42] and ovarian cancer

patients [43].

In this chapter, label-free MS spectral counting [8, 134] was used to quantitate the relative

levels of proteins in exosomes prepared by the three commonly-used exosomal isolation

strategies described above. Using the LIM1863 colorectal carcinoma cell line as a model,

exosomes were isolated from concentrated culture medium by differential

ultracentrifugation (UC-Exos), OptiPrep density gradient centrifugation (DG-Exos), and

EpCAM-magnetic microbead immunoaffinity capture (IAC-Exos). This study revealed that

stereotypical exosome markers Alix, TSG101, and HSP70 were significantly enriched in

the IAC-Exos dataset when compared with differential centrifugation and density gradient

centrifugation (OptiPrep) prepared samples. Detailed examination of the IAC-Exos

dataset also revealed the enrichment of other ESCRT complex subunit proteins involved in

52

exosome biogenesis, and Rab proteins involved in exosome trafficking and release. Taken

together, the findings from this study indicate that EpCAM-based immunoaffinity capture

is more effective than density gradient centrifugation, which in turn is more effective than

differential centrifugation, for enriching exosomes.

For IAC to be more generally applicable there is a need, first, to identify tissue- and cell-

type specific membrane proteins and, secondly, to generate mAbs directed to such proteins

that allow for the selective capture of exosomes using mAb-loaded magnetic beads. In

addition, there is a need to fully evaluate the effectiveness of different immunoaffinity

capture beads, since in my studies (used in Chapter 3), the background contaminants

associated with Protein G Dynabeads® and magnetic assisted cell sorting (MACS)

microbeads varied significantly. Further studies need to be performed to discriminate

whether this is due to mAb isotype or the chemistry of the magnetic beads employed.

Advances in the identification of tissue-specific exosome surface proteins, the development

of corresponding mAbs directed toward these proteins, and improvement of magnetic bead

chemistry will lead to better understanding of exosome biochemistry and may also allow

them to be exploited for clinical use (e.g., biomarker diagnostic), vaccines [309], and

delivery vehicles for therapeutic cargo [215].

Following is the manuscript: Comparison of ultracentrifugation, density gradient

separation, and immunoaffinity capture methods for isolating human colon cancer cell line

LIM1863-derived exosomes. Tauro, B. J., Greening, D. W., Mathias, R. A., Ji, H.,

Mathivanan, S., Scott, A. M., and Simpson, R. J. (2012) Methods 56, 293-304.

Supplemental Information located in Appendix A-6 and online at

doi:10.1016/j.ymeth.2012.01.002.

53

SEE CHAPTER 2 MANUSCRIPT - PAGE 1

54

SEE CHAPTER 2 MANUSCRIPT - PAGE 2

55

SEE CHAPTER 2 MANUSCRIPT - PAGE 3

56

SEE CHAPTER 2 MANUSCRIPT - PAGE 4

57

SEE CHAPTER 2 MANUSCRIPT - PAGE 5

58

SEE CHAPTER 2 MANUSCRIPT - PAGE 6

59

SEE CHAPTER 2 MANUSCRIPT - PAGE 7

60

SEE CHAPTER 2 MANUSCRIPT - PAGE 8

61

SEE CHAPTER 2 MANUSCRIPT - PAGE 9

62

SEE CHAPTER 2 MANUSCRIPT - PAGE 10

63

SEE CHAPTER 2 MANUSCRIPT - PAGE 11

64

SEE CHAPTER 2 MANUSCRIPT - PAGE 12

65

Chapter 3: Use of sequential immunoaffinity capture to

isolate and characterize two distinct populations of

exosomes released from LIM1863 colon carcinoma cells

3.1 Overview

To date, the content and function of exosomes released from highly-polarized epithelial

cells has received limited attention. In polarized cells, the apical and basolateral surfaces

are separated by adherens and tight junctions [310], and endosomes are used to recycle

apical and basolateral proteins to their respective plasma membrane surfaces in order to

maintain cell polarity. It has been reported that apical and basolateral exosomes, each

containing different cargos, may traffic to and be released from apical and basolateral

plasma membranes, respectively [110]. Van Neil and colleagues have hypothesized that

basolateral exosomes and their cargo may be able to cross the basement membrane and

convey immune information to non-contiguous immune cells [72]. To test this hypothesis,

polarized T84 colorectal cancer cells were grown as a monolayer on microporous filters to

allow for apical and basolateral exosome collection. This group subsequently performed a

limited proteomic analysis of exosomes to identify 31 proteins. A key finding was the

differential protein expression of apical trafficking molecule syntaxin 3 in apical exosomes

and basolateral transmembrane protein A33 in basolateral exosomes [72].

To further investigate apical and basolateral trafficking of exosomes in the context of colon

cancer biology, I performed an in-depth proteomic analysis of these exosome sub-

populations from a polarized colon cancer cell line. Having established in Chapter 2 that

immunoaffinity capture (IAC) is the preferred method to enrich exosomes from cell culture

medium, I used this approach to isolate highly-purified apical and basolateral exosomes

using the human colon carcinoma cell line LIM1863 as an experimental model (see section

1.5.1.3). I designed a sequential immunoaffinity method utilizing magnetic beads coated

66

with mAbs directed towards A33 antigen, a definitive marker of basolateral surfaces in

intestinal epithelial cells [154-156], followed by anti-EpCAM loaded magnetic beads, a

ubiquitously expressed epithelial cancer marker [311]. This strategy allowed for the

isolation of two sub-populations of exosomes, A33-Exos and EpCAM-Exos. Proteomic-

profiling revealed the unique identification in EpCAM-Exos of classical apical trafficking

molecules, apical intestinal enzymes and the apically-restricted membrane proteins. By

comparison, A33-Exos were selectively enriched with basolateral trafficking molecules,

such as early endosome antigen 1 (EEA1), ADP-ribosylation factor (ARF1) and clathrin.

Intriguingly, several members of the MHC class I family of antigen presentation molecules

were exclusively observed in A33-Exos, while neither MHC class I or II molecules were

observed in EpCAM-Exos. Findings from this study also showed the co-localization of

EpCAM, claudin-7 and CD44 in EpCAM-Exos. When complexed together, but not alone,

these proteins promote tumor progression [312]. Further investigation is required to

ascertain whether EpCAM/claudin-7/CD44 do in fact complex together in this study, and

whether this complex confers tumor promoting activity in recipient cells.

The findings in this chapter also shed light on the membrane composition of A33- and

EpCAM-Exos. For example, a total of 119 transmembrane-containing proteins were

identified, of which 13 have been previously implicated in colon cancer (e.g., ADAM10

and tumor-associated calcium signal transducer 2). An interesting finding was the

expression of three cell surface proteins specifically localized on normal colon epithelium,

MUC13, TSPAN8 and A33, as indicated following examination of the Human Protein

Atlas (http://www.proteinatlas.org/). Although these three membrane proteins are

specifically expressed in normal colon epithelium, they are over-expressed in many other

cancer tissues. Ongoing studies are required to identify specific membrane proteins in colon

cancer cell-derived exosomes that will morphologically distinguish these exosomes in

blood and other body fluids from exosomes generated from other tissue/cell types. In

addition, two other integral membrane proteins (CD46 and CD59) identified in EpCAM-

Exos, but not A33-Exos, may have the potential to increase the half-life of these exosomes

by protecting them from complement mediated lysis. CD59 prevents activation of the

membrane attack complex by inhibition of C9 incorporation in C5b-9 [313], whereas CD46

67

has cofactor activity for inactivation of complement components C3b and C46 by serum

factor 1, which protects cells from damage by the complement system [314]. Further,

EpCAM-Exos also uniquely expressed high-mobility group box proteins HMGB2 and

HMGB3 which act as sensors to distinct types of nucleic acid and activate innate immune

responses [315]. Studies aimed at determining the functional significance of apical and

basolateral exosomes in colon cancer biology were limited due to time constraints.

In order to eliminate the possibility that A33- and EpCAM-Exos were actually shed

microvesicles (sMVs) (see section 1.3.1), I isolated sMVs and examined their protein

content. This analysis identified 1392 proteins, of which 426 were unique to sMVs. This

showed that A33- and EpCAM-Exos were clearly distinguishable from sMVs. A key

finding was the enrichment of ATP binding cassette transporter proteins, which suggests

that sMVs may be implicated in multidrug resistance [316, 317]. Ongoing studies are

underway in the Simpson laboratory to establish whether LIM1863-derived sMVs are

capable of transferring resistance to drug sensitive colon cancer cells.

Following is the manuscript: Tauro, B. J., Greening, D. W., Mathias, R. A., Mathivanan,

S., Ji, H., and Simpson, R. J. (2013) Two distinct populations of exosomes are released

from LIM1863 colon carcinoma cell-derived organoids. Molecular & Cellular Proteomics

12, 587-598. Supplemental Information located in Appendix A-6 and online at

http://www.mcponline.org/content/suppl/2013/07/11/M112.021303.DC1.

68

69

SEE CHAPTER 3 MANUSCRIPT - PAGE 1

70

SEE CHAPTER 3 MANUSCRIPT - PAGE 2

71

SEE CHAPTER 3 MANUSCRIPT - PAGE 3

72

SEE CHAPTER 3 MANUSCRIPT - PAGE 4

73

SEE CHAPTER 3 MANUSCRIPT - PAGE 5

74

SEE CHAPTER 3 MANUSCRIPT - PAGE 6

75

SEE CHAPTER 3 MANUSCRIPT - PAGE 7

76

SEE CHAPTER 3 MANUSCRIPT - PAGE 8

77

SEE CHAPTER 3 MANUSCRIPT - PAGE 9

78

SEE CHAPTER 3 MANUSCRIPT - PAGE 10

79

SEE CHAPTER 3 MANUSCRIPT - PAGE 11

80

SEE CHAPTER 3 MANUSCRIPT - PAGE 12

81

Chapter 4: Characterization of Madin-Darby canine

kidney (MDCK) cell-derived exosomal proteins following

oncogenic H-Ras induced epithelial-mesenchymal

transition

4.1 Overview

Epithelial-mesenchymal transition (EMT) occurs during embryonic morphogenesis and has

recently been implicated in cancer invasion and metastasis [277, 318-320]. In previous

studies, Mathias and colleagues in the Simpson laboratory examined the components of the

secretome in EMT using oncogenic H-Ras transformed Madin-Darby canine kidney

(MDCK) cells as a model (see sections 1.5.2.4 and 1.5.2.5). These studies, which focused

on the proteomic profiles of soluble-secreted proteins, revealed that following EMT,

secreted components mediating cell-cell and cell-matrix adhesion (e.g., collagen IV, XVII,

and laminin 5) were decreased in abundance, with the concordant upregulation of proteases

and ECM constituents promoting cell motility and invasion (e.g., MMP-1, TIMP-1

kallikrein-6, -7) [250, 300].

In this chapter I extended the secretome studies of Mathias and colleagues [250, 300] by

examining the contribution of exosomes in oncogenic H-Ras-induced EMT in MDCK cells.

As this was the first study of the role of exosomes in EMT using the MDCK model, and

given that there are no known specific MDCK-derived exosomal membrane marker

proteins, immunoaffinity capture isolation was not a viable purification option. Instead I

employed density gradient centrifugation using iodixanol (OptiPrep™) described in

Chapter 2. Proteomic profiling of MDCK-derived exosomes (MDCK-Exos) and oncogenic

H-Ras transformed MDCK-derived exosomes (21D1-Exos) was performed using GeLC-

MS/MS as described in Chapter 2 and 3.

Findings from this study showed that purified MDCK- and 21D1-Exos were

morphologically similar by cryo-electron microscopy and contained stereotypical exosome

82

marker proteins such as TSG101, Alix and CD63. Typical cellular EMT hallmark protein

changes are also mirrored in MDCK and 21D1-Exos, including decreased expression of E-

Cadherin and increased expression of vimentin. This study also revealed that 21D1-Exos

are enriched with several metalloproteinases (e.g., MMP-1, -14, -19, ADAM10,

ADAMTS1), and integrins (e.g., ITGB1, ITGA3, and ITGA6) that have been recently

implicated in regulating the tumor microenvironment to promote metastatic progression

[321-323]. A significant finding of this study was the observation that upon EMT,

mesenchymal 21D1-Exos are extensively reprogrammed to be enriched with key

transcriptional regulators (e.g., master transcriptional regulator YBX1) and core splicing

complex components (e.g., SF3B1, SF3B3, and SFRS1). These findings suggest that Ras-

transformed MDCK cell-derived exosomes may be reprogrammed with factors capable of

inducing EMT in recipient cells. In this regard, preliminary experiments showed that 21D1-

Exos were capable of maintaining the mesenchymal morphology of 21D1 cells. Further

experiments are required to reveal whether 21D1-Exos are capable of inducing EMT in

other recipient cell types, and whether over-expression of YBX1 alone in epithelial cells, or

in conjunction with soluble secreted proteins, is capable of inducing tumorigenesis in

‘normal’ epithelial cells.

Following is the manuscript: Tauro, B. J., Mathias, R. A., Greening, D. W., Gopal, S. K.,

Ji, H, Kapp, E. A., Coleman, B. M., Hill, A. F., Kusebauch, U., Hallows, J. L., Shteynberg,

D., Moritz, R. L., Zhu, H. J., and Simpson R. J. (2013) Oncogenic H-Ras reprograms

Madin-Darby canine kidney (MDCK) cell-derived exosomal proteins during epithelial-

mesenchymal transition. Molecular & Cellular Proteomics 12, 2148-2159. Supplemental

Information located in Appendix A-6 and online at

http://www.mcponline.org/content/early/2013/05/03/mcp.M112.027086/suppl/DC1.

83

SEE CHAPTER 4 MANUSCRIPT - PAGE 1

84

SEE CHAPTER 4 MANUSCRIPT - PAGE 2

85

SEE CHAPTER 4 MANUSCRIPT - PAGE 3

86

SEE CHAPTER 4 MANUSCRIPT - PAGE 4

87

SEE CHAPTER 4 MANUSCRIPT - PAGE 5

88

SEE CHAPTER 4 MANUSCRIPT - PAGE 6

89

SEE CHAPTER 4 MANUSCRIPT - PAGE 7

90

SEE CHAPTER 4 MANUSCRIPT - PAGE 8

91

SEE CHAPTER 4 MANUSCRIPT - PAGE 9

92

SEE CHAPTER 4 MANUSCRIPT - PAGE 10

93

SEE CHAPTER 4 MANUSCRIPT - PAGE 11

94

SEE CHAPTER 4 MANUSCRIPT - PAGE 12

95

Chapter 5: Summary and future directions

5.1 Evaluation of exosome isolation methods and the importance of purity

As microvesicles, 100-1000 nm in diameter, and exosomes, 40-100 nm in diameter, are

morphologically distinct and may have different biological functions, there is a pressing

need to evaluate methods for obtaining homogeneous vesicle preparations in order to

characterise their bioactive content (e.g., proteins, lipids and RNA species). This is

particularly important when considering exosomes for use as clinical agents in therapeutics

(see section 1.4.9). A clear understanding of all exosome components may also lead to the

design of exosome mimetics for personalized medicine [229]. Recently, as circulating

exosomes have been identified to contain disease-specific miRNA signatures, there is also

evidence for the use of exosomes as diagnostic biomarkers [241]. Exosomes also display

the capabilities to be therapeutic agents, these include immune modulation, tissue repair,

and drug delivery due to their biocompatibility, immunological inertness, and ability to

cross the blood brain barrier [209, 210]. Before the full potential of exosomes use in drug

delivery can be realized, large-scale production protocols must be standardized [209].

Currently, detailed biochemical and functional analyses of exosomes are confounded by

technical difficulties in their isolation from cell culture medium and body fluids. Without

stringent purification, exosome samples are typically contaminated with other membranous

vesicles including shed plasma membrane vesicles (e.g., microparticles and oncosomes)

and apoptotic blebs. As these other secreted vesicles share similar physiochemical

characteristics (size, buoyant density) to exosomes, Chapter 2 compared the effectiveness

of three widely-used isolation strategies (ultracentrifugation, density gradient separation

and immunoaffinity capture) to obtain exosomes from colorectal cancer cells. Further, an

in-depth proteomic analysis was performed to establish the purity of secreted exosomes.

The findings from this chapter showed that immunoaffinity capture (IAC) using anti-

96

EpCAM antibody was the optimum method for purifying exosomes. In this study the

assessment of exosome sample homogeneity was based upon selective enrichment of

proteins involved in exosome biogenesis, trafficking, and release. To further this research,

investigation into the isolation of exosomes from other cell-derived sources and biofluids

(e.g., blood, urine, and malignant ascites) needs to be performed to validate the

compatability of the IAC approach.

Further, it is noted that the potential for exosomes to act as a blood-based biomarker

diagnostic source (e.g., using specific disease miRNA signatures) is confounded by the

challenge in isolating exosomes that are characteristic of specific disease cell types, such as

cancer. Such cancer-derived exosomes may be in low abundance when compared to the

general population of exosomes and other microvesicles in circulation. Additionally,

because these disease specific miRNA signatures may be masked by analyzing low

abundance disease-derived exosomes in blood, there is a need to isolate highly pure

exosomes (e.g., using IAC) for diagnostic purposes. In future studies it is therefore essential

to identify tissue-specific membrane protein markers for IAC isolation of disease-related

exosomes.

It remains to be seen whether a large-scale IAC approach will emerge as an alternative to

the filtration-based/size-exclusion HPLC approach currently used for preparing clinical

grade exosomes for regenerative medical purposes [45, 221]. Needless to say, such

evaluation of exosome purification methods is required, as pre-clinical and clinical trials

(http://www.australianclinicaltrials.gov.au/) require large-scale cost-effective standardized

protocols for the production of GMP exosomes.

5.2 Analysis of two exosome sub-populations released from LIM1863 cells

In Chapter 3, a sequential immunoaffinity capture strategy using magnetic beads coupled

with antibodies towards A33 and EpCAM was developed to isolate exosome sub-

populations released from apical and basolateral surfaces of polarized human colon

97

carcinoma cells. An in-depth proteomic characterization of exosome sub-populations

revealed distinct apical and basolateral trafficking proteins, consistent with release of

exosomes from the respective surfaces of polarized epithelial cells. Proteomic analysis

revealed 684 proteins were common to both exosomal datasets, while 340 and 214 proteins

were unique to A33- and EpCAM-Exos, respectively. Interestingly, the key components of

a metastasis-promoting complex, EpCAM/claudin-7/CD44, were uniquely identified in

apically-derived exosomes, while MHC-I components were unique to basolaterally-derived

exosomes.

The finding of selectively enriched proteins in apical and basolateral exosomes raises

questions into the functional roles of these two distinct exosome populations, especially in

the context of tumor biology. To examine their biological function, studies investigating the

abilities of the metastasis-promoting complex, EpCAM/claudin-7/CD44, may be

performed. Although all three components were uniquely identified in EpCAM-Exos, it

could not be claimed that they form a functional complex within exosomes. Further studies

are required to determine whether exosomes contain active signaling complexes involving

oncogenic proteins and, if so, whether such complexes may remain intact upon exosome

uptake in recipient cells and contribute to functional signaling pathways in target cells.

Ongoing efforts using Blue-Native PAGE and co-immunoprecipitation are underway in the

Simpson laboratory to assess the presence of this protein complex and functional validation

of its constituents. Following identification of this complex, experiments involving its

modification (e.g., the knock-down of one or more of the components from the originating

cell line) could be performed. To assess the ability of the complex to act as a metastasis

promoter, experiments involving the application of exosomes, with or without the intact

complex, to recipient cells may be performed. Functional readouts assessing cell

proliferation, migration, invasion or resistance to apoptosis may indicate the hypothesized

metastatic effect of exosomes on recipient cells.

Interestingly, the presence of MHC-I components were only observed in A33-Exos. Further

experiments are required to establish whether these components also form a functional

complex in exosomes. It would be exciting to determine whether the MHC complexes carry

98

peptides, or are capable of carrying peptides, that can elicit a stimulatory or inhibitory

immune cell response.

To further differentiate apical and basolateral exosomes and understand their biological

functions, characterization of their RNA content (miRNA and mRNA) may be investigated.

The first report of exosome-mediated transfer of exogenous nucleic acids was published in

2010, when it was shown that monocytic miR-150 was functionally delivered to endothelial

cells [237]. Since then, numerous other studies demonstrating the RNA transporting

capabilities of exosomes have been reported [171, 173, 324, 325]. Although at an early

stage, these findings collectively illustrate the potential of exosomes for therapeutic

delivery of nucleic acids – for reviews see El-Andaloussi et al., Kooijmans et al., and Rak

et al. [196, 210, 229].

In addition to the characterization of apical and basolateral-derived exosomes, shed

microvesicles (sMVs) were also isolated using differential centrifugation (10,000 g) to

differentiate these vesicles from exosomes. A proteomic characterization revealed 1392

proteins in sMVs with 462 proteins unique to these vesicles. This characterization revealed

another distinct vesicle population and therefore, the potential of different functional

capabilities to that of exosomes. One intriguing finding was the enrichment of ATP binding

cassette transporter proteins, which suggests that sMVs may be involved in multidrug

resistance [316, 317]. Functional studies are planned to establish whether LIM1863-derived

sMVs are capable of transferring resistance to drug sensitive colon cancer cells. Further

investigation into the isolation of sMVs is also being performed in the Simpson laboratory

to identify if sup-populations of sMVs exist within their large size range of 100-1000 nm

diameter.

5.3 Analysis of exosomes following epithelial-mesenchymal transition (EMT)

In Chapter 4 of this thesis, I investigated whether exosomes are modified during the EMT

process. I studied the EMT process using the epithelial MDCK cell line and its oncogenic

99

H-Ras-transformed mesenchymal variant (21D1) as an experimental model. Exosomes

released into the culture media from both cell lines were isolated by density gradient

centrifugation (as described in Chapter 2) and their protein content characterized by MS-

based profiling, ultimately to gain insight into any functional roles of exosomes in EMT.

Proteomic data revealed profound changes in 21D1 cell-derived exosomes. Foremost, these

studies revealed that cellular hallmarks typically identified in the epithelial-mesenchymal

transition (e.g., downregulation of E-cadherin and upregulation of vimentin in

mesenchymal cells) were mirrored in exosomes. These studies also showed that oncogenic

H-Ras reprograms MDCK cell-derived exosomal proteins. For example, mesenchymal

exosomes were enriched with several proteases (e.g., MMP-1, -14, -19, ADAM10 and

ADAMTS1) and integrins (e.g., ITGB1, ITGA3 and ITGA6) which are implicated in

promoting metastasis. A key finding in mesenchymal exosomes was the unique expression

of numerous splicing factors (e.g., SF3B1, SF3B3, and SFRS1) and master transcriptional

regulator, YBX1. Whether these proteins are functional in recipient cells remains to be

investigated.

It is interesting to speculate that exosomes from 21D1 cells may induce a mesenchymal

phenotype in recipient cells. If so, it would be important to identify the specific cargo (e.g.,

proteins, transcription factors or RNAs) that can cause such a transformation. To assess

this, initial experiments involving the addition of exosomes to recipient cells are planned to

be performed to assess changes in proliferation, invasion and migration abilities. Following

an observed morphological change, identification of the specific factors contributing to

these effects could be undertaken. For example, using an RNAi based system to create

stable knock-down cell lines; the expression of selected target genes that govern key

cellular processes may be altered in 21D1 cells. For these studies, target gene selection will

be based on proteomic data identified in 21D1-Exos. For example, highly upregulated

protein and inducer of EMT (e.g., transcription factor YBX1), or characteristic proteins of

the mesenchymal phenotype (e.g., MMP-1) will be selected first for generating knock-

down lines from 21D1 cells. Following gene knock-down, exosomes will be isolated and

characterized to assess that the corresponding protein expression is also reduced. Exosomes

isolated from knock-down and wild-type cell lines will then be applied to recipient cells to

100

identify the contribution of these modified constituents (i.e., YBX1 and MMP-1) in

exosome functionality and recipient cell modulation.

Additionally, the characterization of exosomal miRNA may also allow functional insights

into epithelial and mesenchymal cell-derived exosomes. Rana and colleagues have recently

demonstrated that miRNA contained within exosomes from metastatic rat adenocarcinoma

cells can alter mRNA transcription of proteases, adhesion molecules, chemokine ligands,

cell cycle- and angiogenesis-promoting genes, and genes engaged in oxidative stress

responses [173]. It is hypothesized that similarly to the proteomic analysis of exosomes,

miRNA content may also be dysregulated following EMT and therefore cause different

mRNA transcription in target cells which may result in altered biological functions. To

further support this theory, it has been demonstrated that exosomes can also deliver

functional miRNA to dysregulate mRNA in dendritic cells [171, 324]. Additionally it is

reported that mRNA and miRNA within melanoma exosomes may be actively transported

into normal melanocytes enabling increased invasive properties of these cells [325]. Based

on these recent observations, it would be exciting to examine the RNA species that are

contained within exosomes derived from MDCK to 21D1 cells. This may allow the

identification of bioactive molecules (RNAs) that have functional capabilities associated

with the EMT process.

5.4 Conclusion

Exosomes are important extracellular conveyors of information between cells. This may

occur through the transmission of bioactive molecules such as oncogenic proteins, tumor

suppressors, transcription factors, bioactive lipids and genetic information such as

mRNAs/miRNAs that alter the phenotypic and biological function of target cells. One of

the key obstacles in ascribing any biochemical or functional characteristics to exosomes is

that of sample homogeneity. To date, the majority of functional exosome studies reported

in the literature have been performed on impure samples that are significantly contaminated

with other EV types such as microvesicles and possibly apoptotic bodies.

101

In this thesis I carefully evaluated three commonly-used exosome purification protocols

(differential ultracentrifugation, density gradient centrifugation and immunoaffinity

capture) to establish that immunoaffinity capture was the superior method for the

enrichment of exosomes. With this knowledge, I then developed a sequential

immunoaffinity capture method to isolate and characterize, for the first time, two distinct

populations of exosomes from LIM1863 organoids. In a further application, I used density

gradient centrifugation (OptiPrep™) to investigate, for the first time, exosomes in

oncogenic H-Ras MDCK induced EMT. This study showed that exosomes were

significantly reprogrammed during the EMT process, being selectively enriched with the

master transcriptional regulator, transcription factor YBX1, and components of protein

splicing machinery.

In summary, the findings of this thesis are the first step towards standardizing protocols for

the purification of exosomes. Already, the applications of these purification strategies have

paved the way for a better understanding of the roles of exosomes in colon cancer biology

and their contribution to the oncogenic H-Ras induced EMT process.

102

References

1. Raposo, G. and W. Stoorvogel. (2013) Extracellular vesicles: Exosomes,

microvesicles, and friends. J Cell Biol. 200(4). 373-383

2. Thery, C. (2011) Exosomes: secreted vesicles and intercellular communications.

F1000 Biol Rep. 3. 15

3. Kahlert, C. and R. Kalluri. (2013) Exosomes in tumor microenvironment influence

cancer progression and metastasis. J Mol Med (Berl). 91(4). 431-437

4. Pan, B.T., K. Teng, C. Wu, M. Adam, and R.M. Johnstone. (1985) Electron

microscopic evidence for externalization of the transferrin receptor in vesicular

form in sheep reticulocytes. J Cell Biol. 101(3). 942-948

5. Pan, B.T. and R.M. Johnstone. (1983) Fate of the transferrin receptor during

maturation of sheep reticulocytes in vitro: selective externalization of the receptor.

Cell. 33(3). 967-978

6. Johnstone, R.M., M. Adam, J.R. Hammond, L. Orr, and C. Turbide. (1987) Vesicle

formation during reticulocyte maturation. Association of plasma membrane

activities with released vesicles (exosomes). J Biol Chem. 262(19). 9412-9420

7. Gyorgy, B., T.G. Szabo, M. Pasztoi, Z. Pal, P. Misjak, B. Aradi, V. Laszlo, E.

Pallinger, E. Pap, A. Kittel, G. Nagy, A. Falus, and E.I. Buzas. (2011) Membrane

vesicles, current state-of-the-art: emerging role of extracellular vesicles. Cell Mol

Life Sci. 68(16). 2667-2688

8. Vogel, C. and E.M. Marcotte. (2012) Label-free protein quantitation using weighted

spectral counting. Methods Mol Biol. 893. 321-341

9. Paltridge, J.L., L. Belle, and Y. Khew-Goodall. (2013) The secretome in cancer

progression. Biochim Biophys Acta. Epub ahead of print

10. Greening, D.W., E.A. Kapp, H. Ji, T.P. Speed, and R.J. Simpson. (2013) Colon

tumour secretopeptidome: Insights into endogenous proteolytic cleavage events in

the colon tumour microenvironment. Biochim Biophys Acta. Epub ahead of print

11. Greening, D. and R. Simpson. (2013) An updated secretome. Biochim Biophys Acta,

12. Makridakis, M. and A. Vlahou. (2010) Secretome proteomics for discovery of

cancer biomarkers. J Proteomics. 73(12). 2291-2305

103

13. Pavlou, M.P. and E.P. Diamandis. (2010) The cancer cell secretome: a good source

for discovering biomarkers? J Proteomics. 73(10). 1896-1906

14. Karagiannis, G.S., M.P. Pavlou, and E.P. Diamandis. (2010) Cancer secretomics

reveal pathophysiological pathways in cancer molecular oncology. Mol Oncol. 4(6).

496-510

15. Ji, H., R.J. Goode, F. Vaillant, S. Mathivanan, E.A. Kapp, R.A. Mathias, G.J.

Lindeman, J.E. Visvader, and R.J. Simpson. (2011) Proteomic profiling of

secretome and adherent plasma membranes from distinct mammary epithelial cell

subpopulations. Proteomics. 11(20). 4029-4039

16. Dumortier, J., D.N. Streblow, A.V. Moses, J.M. Jacobs, C.N. Kreklywich, D.

Camp, R.D. Smith, S.L. Orloff, and J.A. Nelson. (2008) Human cytomegalovirus

secretome contains factors that induce angiogenesis and wound healing. J Virol.

82(13). 6524-6535

17. Formolo, C.A., R. Williams, H. Gordish-Dressman, T.J. MacDonald, N.H. Lee, and

Y. Hathout. (2011) Secretome signature of invasive glioblastoma multiforme. J

Proteome Res. 10(7). 3149-3159

18. Hathout, Y. (2007) Approaches to the study of the cell secretome. Expert Rev

Proteomics. 4(2). 239-248

19. Peinado, H., S. Lavotshkin, and D. Lyden. (2011) The secreted factors responsible

for pre-metastatic niche formation: old sayings and new thoughts. Semin Cancer

Biol. 21(2). 139-146

20. Gadea, B.B. and J.A. Joyce. (2006) Tumour-host interactions: implications for

developing anti-cancer therapies. Expert Rev Mol Med. 8(30). 1-32

21. Joyce, J.A. (2005) Therapeutic targeting of the tumor microenvironment. Cancer

Cell. 7(6). 513-520

22. Al-Nedawi, K., B. Meehan, and J. Rak. (2009) Microvesicles: messengers and

mediators of tumor progression. Cell Cycle. 8(13). 2014-2018

23. Wolf, P. (1967) The nature and significance of platelet products in human plasma.

Br J Haematol. 13(3). 269-288

104

24. Heijnen, H.F., A.E. Schiel, R. Fijnheer, H.J. Geuze, and J.J. Sixma. (1999)

Activated platelets release two types of membrane vesicles: microvesicles by

surface shedding and exosomes derived from exocytosis of multivesicular bodies

and alpha-granules. Blood. 94(11). 3791-3799

25. Lee, T.H., E. D'Asti, N. Magnus, K. Al-Nedawi, B. Meehan, and J. Rak. (2011)

Microvesicles as mediators of intercellular communication in cancer--the emerging

science of cellular 'debris'. Semin Immunopathol. 33(5). 455-467

26. Cocucci, E., G. Racchetti, and J. Meldolesi. (2009) Shedding microvesicles:

artefacts no more. Trends Cell Biol. 19(2). 43-51

27. Huotari, J. and A. Helenius. (2011) Endosome maturation. Embo J. 30(17). 3481-

3500

28. Beekman, J.M. and P.J. Coffer. (2008) The ins and outs of syntenin, a

multifunctional intracellular adaptor protein. J Cell Sci. 121(Pt 9). 1349-1355

29. Camussi, G., M.C. Deregibus, S. Bruno, V. Cantaluppi, and L. Biancone. (2010)

Exosomes/microvesicles as a mechanism of cell-to-cell communication. Kidney Int.

78(9). 838-848

30. Breakefield, X.O., R.M. Frederickson, and R.J. Simpson. (2011) Gesicles:

Microvesicle "Cookies" for Transient Information Transfer Between Cells. Mol

Ther. 19(9). 1574-1576

31. Gasser, O., C. Hess, S. Miot, C. Deon, J.C. Sanchez, and J.A. Schifferli. (2003)

Characterisation and properties of ectosomes released by human polymorphonuclear

neutrophils. Exp Cell Res. 285(2). 243-257

32. Thery, C., M. Ostrowski, and E. Segura. (2009) Membrane vesicles as conveyors of

immune responses. Nat Rev Immunol. 9(8). 581-593

33. Beyer, C. and D.S. Pisetsky. (2010) The role of microparticles in the pathogenesis

of rheumatic diseases. Nat Rev Rheumatol. 6(1). 21-29

34. Gregory, C.D. and A. Devitt. (2004) The macrophage and the apoptotic cell: an

innate immune interaction viewed simplistically? Immunology. 113(1). 1-14

35. Thery, C., S. Amigorena, G. Raposo, and A. Clayton. (2006) Isolation and

characterization of exosomes from cell culture supernatants and biological fluids.

Curr Protoc Cell Biol. Chapter 3. Unit 3 22

105

36. Akers, J.C., D. Gonda, R. Kim, B.S. Carter, and C.C. Chen. (2013) Biogenesis of

extracellular vesicles (EV): exosomes, microvesicles, retrovirus-like vesicles, and

apoptotic bodies. J Neurooncol. 113(1). 1-11

37. Koga, K., K. Matsumoto, T. Akiyoshi, M. Kubo, N. Yamanaka, A. Tasaki, H.

Nakashima, M. Nakamura, S. Kuroki, M. Tanaka, and M. Katano. (2005)

Purification, characterization and biological significance of tumor-derived

exosomes. Anticancer Res. 25(6A). 3703-3707

38. Logozzi, M., A. De Milito, L. Lugini, M. Borghi, L. Calabro, M. Spada, M.

Perdicchio, M.L. Marino, C. Federici, E. Iessi, D. Brambilla, G. Venturi, F.

Lozupone, M. Santinami, V. Huber, M. Maio, L. Rivoltini, and S. Fais. (2009) High

levels of exosomes expressing CD63 and caveolin-1 in plasma of melanoma

patients. PLoS One. 4(4). e5219

39. Coren, L.V., T. Shatzer, and D.E. Ott. (2008) CD45 immunoaffinity depletion of

vesicles from Jurkat T cells demonstrates that exosomes contain CD45: no evidence

for a distinct exosome/HIV-1 budding pathway. Retrovirology. 5. 64

40. Clayton, A., J. Court, H. Navabi, M. Adams, M.D. Mason, J.A. Hobot, G.R.

Newman, and B. Jasani. (2001) Analysis of antigen presenting cell derived

exosomes, based on immuno-magnetic isolation and flow cytometry. J Immunol

Methods. 247(1-2). 163-174

41. Caby, M.P., D. Lankar, C. Vincendeau-Scherrer, G. Raposo, and C. Bonnerot.

(2005) Exosomal-like vesicles are present in human blood plasma. Int Immunol.

17(7). 879-887

42. Rabinowits, G., C. Gercel-Taylor, J.M. Day, D.D. Taylor, and G.H. Kloecker.

(2009) Exosomal microRNA: a diagnostic marker for lung cancer. Clin Lung

Cancer. 10(1). 42-46

43. Taylor, D.D. and C. Gercel-Taylor. (2008) MicroRNA signatures of tumor-derived

exosomes as diagnostic biomarkers of ovarian cancer. Gynecol Oncol. 110(1). 13-

21

106

44. Mathivanan, S., J.W. Lim, B.J. Tauro, H. Ji, R.L. Moritz, and R.J. Simpson. (2010)

Proteomics analysis of A33 immunoaffinity-purified exosomes released from the

human colon tumor cell line LIM1215 reveals a tissue-specific protein signature.

Mol Cell Proteomics. 9(2). 197-208

45. Lai, R.C., F. Arslan, S.S. Tan, B. Tan, A. Choo, M.M. Lee, T.S. Chen, B.J. Teh,

J.K. Eng, H. Sidik, V. Tanavde, W.S. Hwang, C.N. Lee, R.M. El Oakley, G.

Pasterkamp, D.P. de Kleijn, K.H. Tan, and S.K. Lim. (2010) Derivation and

characterization of human fetal MSCs: an alternative cell source for large-scale

production of cardioprotective microparticles. J Mol Cell Cardiol. 48(6). 1215-1224

46. Lai, R.C., F. Arslan, M.M. Lee, N.S. Sze, A. Choo, T.S. Chen, M. Salto-Tellez, L.

Timmers, C.N. Lee, R.M. El Oakley, G. Pasterkamp, D.P. de Kleijn, and S.K. Lim.

(2010) Exosome secreted by MSC reduces myocardial ischemia/reperfusion injury.

Stem Cell Res. 4(3). 214-222

47. Ogawa, Y., Y. Miura, A. Harazono, M. Kanai-Azuma, Y. Akimoto, H. Kawakami,

T. Yamaguchi, T. Toda, T. Endo, M. Tsubuki, and R. Yanoshita. (2011) Proteomic

analysis of two types of exosomes in human whole saliva. Biol Pharm Bull. 34(1).

13-23

48. Choi, D.S., J.M. Lee, G.W. Park, H.W. Lim, J.Y. Bang, Y.K. Kim, K.H. Kwon, H.J.

Kwon, K.P. Kim, and Y.S. Gho. (2007) Proteomic analysis of microvesicles derived

from human colorectal cancer cells. J Proteome Res. 6(12). 4646-4655

49. Di Vizio, D., M. Morello, A.C. Dudley, P.W. Schow, R.M. Adam, S. Morley, D.

Mulholland, M. Rotinen, M.H. Hager, L. Insabato, M.A. Moses, F. Demichelis,

M.P. Lisanti, H. Wu, M. Klagsbrun, N.A. Bhowmick, M.A. Rubin, C. D'Souza-

Schorey, and M.R. Freeman. (2012) Large oncosomes in human prostate cancer

tissues and in the circulation of mice with metastatic disease. Am J Pathol. 181(5).

1573-1584

50. Thery, C., M. Boussac, P. Veron, P. Ricciardi-Castagnoli, G. Raposo, J. Garin, and

S. Amigorena. (2001) Proteomic analysis of dendritic cell-derived exosomes: a

secreted subcellular compartment distinct from apoptotic vesicles. J Immunol.

166(12). 7309-7318

107

51. Zhou, Z., X. Sun, and Y.J. Kang. (2001) Ethanol-induced apoptosis in mouse liver:

Fas- and cytochrome c-mediated caspase-3 activation pathway. Am J Pathol.

159(1). 329-338

52. Kalra, H., R.J. Simpson, H. Ji, E. Aikawa, P. Altevogt, P. Askenase, V.C. Bond,

F.E. Borras, X. Breakefield, V. Budnik, E. Buzas, G. Camussi, A. Clayton, E.

Cocucci, J.M. Falcon-Perez, S. Gabrielsson, Y.S. Gho, D. Gupta, H.C. Harsha, A.

Hendrix, A.F. Hill, J.M. Inal, G. Jenster, E.M. Kramer-Albers, S.K. Lim, A.

Llorente, J. Lotvall, A. Marcilla, L. Mincheva-Nilsson, I. Nazarenko, R. Nieuwland,

E.N. Nolte-'t Hoen, A. Pandey, T. Patel, M.G. Piper, S. Pluchino, T.S. Prasad, L.

Rajendran, G. Raposo, M. Record, G.E. Reid, F. Sanchez-Madrid, R.M. Schiffelers,

P. Siljander, A. Stensballe, W. Stoorvogel, D. Taylor, C. Thery, H. Valadi, B.W.

van Balkom, J. Vazquez, M. Vidal, M.H. Wauben, M. Yanez-Mo, M. Zoeller, and

S. Mathivanan. (2012) Vesiclepedia: a compendium for extracellular vesicles with

continuous community annotation. PLoS Biol. 10(12). e1001450

53. Gonzalez, L.J., E. Gibbons, R.W. Bailey, J. Fairbourn, T. Nguyen, S.K. Smith, K.B.

Best, J. Nelson, A.M. Judd, and J.D. Bell. (2009) The influence of membrane

physical properties on microvesicle release in human erythrocytes. PMC Biophys.

2(1). 7

54. Pluskota, E., N.M. Woody, D. Szpak, C.M. Ballantyne, D.A. Soloviev, D.I. Simon,

and E.F. Plow. (2008) Expression, activation, and function of integrin alphaMbeta2

(Mac-1) on neutrophil-derived microparticles. Blood. 112(6). 2327-2335

55. Muralidharan-Chari, V., J.W. Clancy, A. Sedgwick, and C. D'Souza-Schorey.

(2010) Microvesicles: mediators of extracellular communication during cancer

progression. J Cell Sci. 123(Pt 10). 1603-1611

56. Goler-Baron, V. and Y.G. Assaraf. (2011) Structure and function of ABCG2-rich

extracellular vesicles mediating multidrug resistance. PLoS One. 6(1). e16007

57. Han, B. and J.T. Zhang. (2004) Multidrug resistance in cancer chemotherapy and

xenobiotic protection mediated by the half ATP-binding cassette transporter

ABCG2. Curr Med Chem Anticancer Agents. 4(1). 31-42

58. Ejendal, K.F. and C.A. Hrycyna. (2002) Multidrug resistance and cancer: the role of

the human ABC transporter ABCG2. Curr Protein Pept Sci. 3(5). 503-511

108

59. Robey, R.W., O. Polgar, J. Deeken, K.W. To, and S.E. Bates. (2007) ABCG2:

determining its relevance in clinical drug resistance. Cancer Metastasis Rev. 26(1).

39-57

60. Mealey, K.L. (2012) ABCG2 transporter: therapeutic and physiologic implications

in veterinary species. J Vet Pharmacol Ther. 35(2). 105-112

61. Sidhu, S.S., A.T. Mengistab, A.N. Tauscher, J. LaVail, and C. Basbaum. (2004)

The microvesicle as a vehicle for EMMPRIN in tumor-stromal interactions.

Oncogene. 23(4). 956-963

62. Di Vizio, D., J. Kim, M.H. Hager, M. Morello, W. Yang, C.J. Lafargue, L.D. True,

M.A. Rubin, R.M. Adam, R. Beroukhim, F. Demichelis, and M.R. Freeman. (2009)

Oncosome formation in prostate cancer: association with a region of frequent

chromosomal deletion in metastatic disease. Cancer Res. 69(13). 5601-5609

63. Rood, I.M., J.K. Deegens, M.L. Merchant, W.P. Tamboer, D.W. Wilkey, J.F.

Wetzels, and J.B. Klein. (2010) Comparison of three methods for isolation of

urinary microvesicles to identify biomarkers of nephrotic syndrome. Kidney Int.

78(8). 810-816

64. Elmore, S. (2007) Apoptosis: a review of programmed cell death. Toxicol Pathol.

35(4). 495-516

65. Bratton, D.L., V.A. Fadok, D.A. Richter, J.M. Kailey, S.C. Frasch, T. Nakamura,

and P.M. Henson. (1999) Polyamine regulation of plasma membrane phospholipid

flip-flop during apoptosis. J Biol Chem. 274(40). 28113-28120

66. Albert, M.L., S.F. Pearce, L.M. Francisco, B. Sauter, P. Roy, R.L. Silverstein, and

N. Bhardwaj. (1998) Immature dendritic cells phagocytose apoptotic cells via

alphavbeta5 and CD36, and cross-present antigens to cytotoxic T lymphocytes. J

Exp Med. 188(7). 1359-1368

67. Albert, M.L., B. Sauter, and N. Bhardwaj. (1998) Dendritic cells acquire antigen

from apoptotic cells and induce class I-restricted CTLs. Nature. 392(6671). 86-89

68. Bergsmedh, A., A. Szeles, M. Henriksson, A. Bratt, M.J. Folkman, A.L. Spetz, and

L. Holmgren. (2001) Horizontal transfer of oncogenes by uptake of apoptotic

bodies. Proc Natl Acad Sci U S A. 98(11). 6407-6411

109

69. McLellan, A.D. (2009) Exosome release by primary B cells. Crit Rev Immunol.

29(3). 203-217

70. Buschow, S.I., E.N. Nolte-'t Hoen, G. van Niel, M.S. Pols, T. ten Broeke, M.

Lauwen, F. Ossendorp, C.J. Melief, G. Raposo, R. Wubbolts, M.H. Wauben, and

W. Stoorvogel. (2009) MHC II in dendritic cells is targeted to lysosomes or T cell-

induced exosomes via distinct multivesicular body pathways. Traffic. 10(10). 1528-

1542

71. Ghidoni, R., A. Paterlini, V. Albertini, M. Glionna, E. Monti, L. Schiaffonati, L.

Benussi, E. Levy, and G. Binetti. (2009) Cystatin C is released in association with

exosomes: A new tool of neuronal communication which is unbalanced in

Alzheimer's disease. Neurobiol Aging,

72. van Niel, G., G. Raposo, C. Candalh, M. Boussac, R. Hershberg, N. Cerf-

Bensussan, and M. Heyman. (2001) Intestinal epithelial cells secrete exosome-like

vesicles. Gastroenterology. 121(2). 337-349

73. Wolfers, J., A. Lozier, G. Raposo, A. Regnault, C. Thery, C. Masurier, C. Flament,

S. Pouzieux, F. Faure, T. Tursz, E. Angevin, S. Amigorena, and L. Zitvogel. (2001)

Tumor-derived exosomes are a source of shared tumor rejection antigens for CTL

cross-priming. Nat Med. 7(3). 297-303

74. Simpson, R.J., J.W. Lim, R.L. Moritz, and S. Mathivanan. (2009) Exosomes:

proteomic insights and diagnostic potential. Expert Rev Proteomics. 6(3). 267-283

75. Bobrie, A., M. Colombo, G. Raposo, and C. Thery. (2011) Exosome secretion:

molecular mechanisms and roles in immune responses. Traffic. 12(12). 1659-1668

76. van Niel, G., I. Porto-Carreiro, S. Simoes, and G. Raposo. (2006) Exosomes: a

common pathway for a specialized function. J Biochem. 140(1). 13-21

77. Simons, M. and G. Raposo. (2009) Exosomes--vesicular carriers for intercellular

communication. Curr Opin Cell Biol. 21(4). 575-581

78. Christoforidis, S., M. Miaczynska, K. Ashman, M. Wilm, L. Zhao, S.C. Yip, M.D.

Waterfield, J.M. Backer, and M. Zerial. (1999) Phosphatidylinositol-3-OH kinases

are Rab5 effectors. Nat Cell Biol. 1(4). 249-252

79. Nielsen, E., F. Severin, J.M. Backer, A.A. Hyman, and M. Zerial. (1999) Rab5

regulates motility of early endosomes on microtubules. Nat Cell Biol. 1(6). 376-382

110

80. Mayor, S. and R.E. Pagano. (2007) Pathways of clathrin-independent endocytosis.

Nat Rev Mol Cell Biol. 8(8). 603-612

81. Maxfield, F.R. and D.J. Yamashiro. (1987) Endosome acidification and the

pathways of receptor-mediated endocytosis. Adv Exp Med Biol. 225. 189-198

82. Poteryaev, D., S. Datta, K. Ackema, M. Zerial, and A. Spang. (2010) Identification

of the switch in early-to-late endosome transition. Cell. 141(3). 497-508

83. Hurley, J.H. and G. Odorizzi. (2012) Get on the exosome bus with ALIX. Nat Cell

Biol. 14(7). 654-655

84. Raiborg, C. and H. Stenmark. (2009) The ESCRT machinery in endosomal sorting

of ubiquitylated membrane proteins. Nature. 458(7237). 445-452

85. Babst, M. (2005) A protein's final ESCRT. Traffic. 6(1). 2-9

86. Raiborg, C., T.E. Rusten, and H. Stenmark. (2003) Protein sorting into

multivesicular endosomes. Curr Opin Cell Biol. 15(4). 446-455

87. Katzmann, D.J., G. Odorizzi, and S.D. Emr. (2002) Receptor downregulation and

multivesicular-body sorting. Nat Rev Mol Cell Biol. 3(12). 893-905

88. Lobert, V.H. and H. Stenmark. (2011) Cell polarity and migration: emerging role

for the endosomal sorting machinery. Physiology (Bethesda). 26(3). 171-180

89. Hurley, J.H. (2010) The ESCRT complexes. Crit Rev Biochem Mol Biol. 45(6).

463-487

90. Amerik, A.Y., J. Nowak, S. Swaminathan, and M. Hochstrasser. (2000) The Doa4

deubiquitinating enzyme is functionally linked to the vacuolar protein-sorting and

endocytic pathways. Mol Biol Cell. 11(10). 3365-3380

91. Saksena, S., J. Wahlman, D. Teis, A.E. Johnson, and S.D. Emr. (2009) Functional

reconstitution of ESCRT-III assembly and disassembly. Cell. 136(1). 97-109

92. Doyotte, A., M.R. Russell, C.R. Hopkins, and P.G. Woodman. (2005) Depletion of

TSG101 forms a mammalian "Class E" compartment: a multicisternal early

endosome with multiple sorting defects. J Cell Sci. 118(Pt 14). 3003-3017

93. Theos, A.C., S.T. Truschel, D. Tenza, I. Hurbain, D.C. Harper, J.F. Berson, P.C.

Thomas, G. Raposo, and M.S. Marks. (2006) A lumenal domain-dependent

pathway for sorting to intralumenal vesicles of multivesicular endosomes involved

in organelle morphogenesis. Dev Cell. 10(3). 343-354

111

94. van Niel, G., S. Charrin, S. Simoes, M. Romao, L. Rochin, P. Saftig, M.S. Marks,

E. Rubinstein, and G. Raposo. (2011) The tetraspanin CD63 regulates ESCRT-

independent and -dependent endosomal sorting during melanogenesis. Dev Cell.

21(4). 708-721

95. Baietti, M.F., Z. Zhang, E. Mortier, A. Melchior, G. Degeest, A. Geeraerts, Y.

Ivarsson, F. Depoortere, C. Coomans, E. Vermeiren, P. Zimmermann, and G.

David. (2012) Syndecan-syntenin-ALIX regulates the biogenesis of exosomes. Nat

Cell Biol. 14(7). 677-685

96. Trajkovic, K., C. Hsu, S. Chiantia, L. Rajendran, D. Wenzel, F. Wieland, P.

Schwille, B. Brugger, and M. Simons. (2008) Ceramide triggers budding of

exosome vesicles into multivesicular endosomes. Science. 319(5867). 1244-1247

97. Putz, U., J. Howitt, A. Doan, C.P. Goh, L.H. Low, J. Silke, and S.S. Tan. (2012)

The tumor suppressor PTEN is exported in exosomes and has phosphatase activity

in recipient cells. Sci Signal. 5(243). ra70

98. Loubery, S., C. Wilhelm, I. Hurbain, S. Neveu, D. Louvard, and E. Coudrier. (2008)

Different microtubule motors move early and late endocytic compartments. Traffic.

9(4). 492-509

99. Jordens, I., M. Fernandez-Borja, M. Marsman, S. Dusseljee, L. Janssen, J. Calafat,

H. Janssen, R. Wubbolts, and J. Neefjes. (2001) The Rab7 effector protein RILP

controls lysosomal transport by inducing the recruitment of dynein-dynactin motors.

Curr Biol. 11(21). 1680-1685

100. Luzio, J.P., P.R. Pryor, and N.A. Bright. (2007) Lysosomes: fusion and function.

Nat Rev Mol Cell Biol. 8(8). 622-632

101. Hsu, C., Y. Morohashi, S. Yoshimura, N. Manrique-Hoyos, S. Jung, M.A.

Lauterbach, M. Bakhti, M. Gronborg, W. Mobius, J. Rhee, F.A. Barr, and M.

Simons. (2010) Regulation of exosome secretion by Rab35 and its GTPase-

activating proteins TBC1D10A-C. J Cell Biol. 189(2). 223-232

102. Savina, A., M. Vidal, and M.I. Colombo. (2002) The exosome pathway in K562

cells is regulated by Rab11. J Cell Sci. 115(Pt 12). 2505-2515

112

103. Ostrowski, M., N.B. Carmo, S. Krumeich, I. Fanget, G. Raposo, A. Savina, C.F.

Moita, K. Schauer, A.N. Hume, R.P. Freitas, B. Goud, P. Benaroch, N. Hacohen,

M. Fukuda, C. Desnos, M.C. Seabra, F. Darchen, S. Amigorena, L.F. Moita, and C.

Thery. (2010) Rab27a and Rab27b control different steps of the exosome secretion

pathway. Nat Cell Biol. 12(1). 19-30; sup pp 11-13

104. Hong, W. (2005) SNAREs and traffic. Biochim Biophys Acta. 1744(3). 493-517

105. Gross, J.C., V. Chaudhary, K. Bartscherer, and M. Boutros. (2012) Active Wnt

proteins are secreted on exosomes. Nat Cell Biol. 14(10). 1036-1045

106. Hasan, N. and C. Hu. (2010) Vesicle-associated membrane protein 2 mediates

trafficking of alpha5beta1 integrin to the plasma membrane. Exp Cell Res. 316(1).

12-23

107. Jahn, R. and R.H. Scheller. (2006) SNAREs--engines for membrane fusion. Nat Rev

Mol Cell Biol. 7(9). 631-643

108. Fader, C.M., D.G. Sanchez, M.B. Mestre, and M.I. Colombo. (2009) TI-

VAMP/VAMP7 and VAMP3/cellubrevin: two v-SNARE proteins involved in

specific steps of the autophagy/multivesicular body pathways. Biochim Biophys

Acta. 1793(12). 1901-1916

109. Rodriguez-Boulan, E., G. Kreitzer, and A. Musch. (2005) Organization of vesicular

trafficking in epithelia. Nat Rev Mol Cell Biol. 6(3). 233-247

110. Lakkaraju, A. and E. Rodriguez-Boulan. (2008) Itinerant exosomes: emerging roles

in cell and tissue polarity. Trends Cell Biol. 18(5). 199-209

111. Parton, R.G., K. Prydz, M. Bomsel, K. Simons, and G. Griffiths. (1989) Meeting of

the apical and basolateral endocytic pathways of the Madin-Darby canine kidney

cell in late endosomes. J Cell Biol. 109(6 Pt 2). 3259-3272

112. Ng, Y.H., S. Rome, A. Jalabert, A. Forterre, H. Singh, C.L. Hincks, and L.A.

Salamonsen. (2013) Endometrial exosomes/microvesicles in the uterine

microenvironment: a new paradigm for embryo-endometrial cross talk at

implantation. PLoS One. 8(3). e58502

113. Coleman, B.M., E. Hanssen, V.A. Lawson, and A.F. Hill. (2012) Prion-infected

cells regulate the release of exosomes with distinct ultrastructural features. Faseb J.

26(10). 4160-4173

113

114. Ji, H., D.W. Greening, T.W. Barnes, J.W. Lim, B.J. Tauro, A. Rai, R. Xu, C. Adda,

S. Mathivanan, W. Zhao, Y. Xue, T. Xu, H.J. Zhu, and R.J. Simpson. (2013)

Proteome profiling of exosomes derived from human primary and metastatic

colorectal cells reveal differential expression of key metastatic factors and signal

transduction components. Proteomics. 10-11. 1672-1686

115. Binnig, G., C.F. Quate, and C. Gerber. (1986) Atomic force microscope. Phys Rev

Lett. 56(9). 930-933

116. van der Pol, E., A.G. Hoekstra, A. Sturk, C. Otto, T.G. van Leeuwen, and R.

Nieuwland. (2010) Optical and non-optical methods for detection and

characterization of microparticles and exosomes. J Thromb Haemost. 8(12). 2596-

2607

117. D'Souza-Schorey, C. and J.W. Clancy. (2012) Tumor-derived microvesicles:

shedding light on novel microenvironment modulators and prospective cancer

biomarkers. Genes Dev. 26(12). 1287-1299

118. Lawrie, A.S., A. Albanyan, R.A. Cardigan, I.J. Mackie, and P. Harrison. (2009)

Microparticle sizing by dynamic light scattering in fresh-frozen plasma. Vox Sang.

96(3). 206-212

119. Hood, J.L., R.S. San, and S.A. Wickline. (2011) Exosomes released by melanoma

cells prepare sentinel lymph nodes for tumor metastasis. Cancer Res. 71(11). 3792-

3801

120. Filipe, V., A. Hawe, and W. Jiskoot. (2010) Critical evaluation of Nanoparticle

Tracking Analysis (NTA) by NanoSight for the measurement of nanoparticles and

protein aggregates. Pharm Res. 27(5). 796-810

121. Nolte-'t Hoen, E.N., E.J. van der Vlist, M. Aalberts, H.C. Mertens, B.J. Bosch, W.

Bartelink, E. Mastrobattista, E.V. van Gaal, W. Stoorvogel, G.J. Arkesteijn, and

M.H. Wauben. (2012) Quantitative and qualitative flow cytometric analysis of

nanosized cell-derived membrane vesicles. Nanomedicine. 8(5). 712-720

122. Mathivanan, S., C.J. Fahner, G.E. Reid, and R.J. Simpson. (2012) ExoCarta 2012:

database of exosomal proteins, RNA and lipids. Nucleic Acids Res. 40(1). D1241-

1244

114

123. Mathivanan, S. and R.J. Simpson. (2009) ExoCarta: A compendium of exosomal

proteins and RNA. Proteomics. 9(21). 4997-5000

124. Anderson, N.L. and N.G. Anderson. (1998) Proteome and proteomics: new

technologies, new concepts, and new words. Electrophoresis. 19(11). 1853-1861

125. Graner, M.W., O. Alzate, A.M. Dechkovskaia, J.D. Keene, J.H. Sampson, D.A.

Mitchell, and D.D. Bigner. (2008) Proteomic and immunologic analyses of brain

tumor exosomes. Faseb J. 5. 1541-1557

126. Raimondo, F., L. Morosi, C. Chinello, F. Magni, and M. Pitto. (2011) Advances in

membranous vesicle and exosome proteomics improving biological understanding

and biomarker discovery. Proteomics. 11(4). 709-720

127. Ducret, A., I. Van Oostveen, J.K. Eng, J.R. Yates, 3rd, and R. Aebersold. (1998)

High throughput protein characterization by automated reverse-phase

chromatography/electrospray tandem mass spectrometry. Protein Sci. 7(3). 706-719

128. Qin, J., D. Fenyo, Y. Zhao, W.W. Hall, D.M. Chao, C.J. Wilson, R.A. Young, and

B.T. Chait. (1997) A strategy for rapid, high-confidence protein identification. Anal

Chem. 69(19). 3995-4001

129. Perkins, D.N., D.J. Pappin, D.M. Creasy, and J.S. Cottrell. (1999) Probability-based

protein identification by searching sequence databases using mass spectrometry

data. Electrophoresis. 20(18). 3551-3567

130. Deutsch, E.W., L. Mendoza, D. Shteynberg, T. Farrah, H. Lam, N. Tasman, Z. Sun,

E. Nilsson, B. Pratt, B. Prazen, J.K. Eng, D.B. Martin, A.I. Nesvizhskii, and R.

Aebersold. (2010) A guided tour of the Trans-Proteomic Pipeline. Proteomics.

10(6). 1150-1159

131. Ong, S.E., B. Blagoev, I. Kratchmarova, D.B. Kristensen, H. Steen, A. Pandey, and

M. Mann. (2002) Stable isotope labeling by amino acids in cell culture, SILAC, as a

simple and accurate approach to expression proteomics. Mol Cell Proteomics. 1(5).

376-386

132. Chong, P.K., C.S. Gan, T.K. Pham, and P.C. Wright. (2006) Isobaric tags for

relative and absolute quantitation (iTRAQ) reproducibility: Implication of multiple

injections. J Proteome Res. 5(5). 1232-1240

115

133. Domon, B. and R. Aebersold. (2006) Mass spectrometry and protein analysis.

Science. 312(5771). 212-217

134. Beissbarth, T., L. Hyde, G.K. Smyth, C. Job, W.M. Boon, S.S. Tan, H.S. Scott, and

T.P. Speed. (2004) Statistical modeling of sequencing errors in SAGE libraries.

Bioinformatics. 20 Suppl 1. i31-39

135. Holman, S.W., P.F. Sims, and C.E. Eyers. (2012) The use of selected reaction

monitoring in quantitative proteomics. Bioanalysis. 4(14). 1763-1786

136. Brownridge, P., S.W. Holman, S.J. Gaskell, C.M. Grant, V.M. Harman, S.J.

Hubbard, K. Lanthaler, C. Lawless, R. O'Cualain, P. Sims, R. Watkins, and R.J.

Beynon. (2011) Global absolute quantification of a proteome: Challenges in the

deployment of a QconCAT strategy. Proteomics. 11(15). 2957-2970

137. Schmidt, C. and H. Urlaub. (2012) Absolute quantification of proteins using

standard peptides and multiple reaction monitoring. Methods Mol Biol. 893. 249-

265

138. Valadi, H., K. Ekstrom, A. Bossios, M. Sjostrand, J.J. Lee, and J.O. Lotvall. (2007)

Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of

genetic exchange between cells. Nat Cell Biol. 9(6). 654-659

139. Wubbolts, R., R.S. Leckie, P.T. Veenhuizen, G. Schwarzmann, W. Mobius, J.

Hoernschemeyer, J.W. Slot, H.J. Geuze, and W. Stoorvogel. (2003) Proteomic and

biochemical analyses of human B cell-derived exosomes. Potential implications for

their function and multivesicular body formation. J Biol Chem. 278(13). 10963-

10972

140. Blanchard, N., D. Lankar, F. Faure, A. Regnault, C. Dumont, G. Raposo, and C.

Hivroz. (2002) TCR activation of human T cells induces the production of

exosomes bearing the TCR/CD3/zeta complex. J Immunol. 168(7). 3235-3241

141. Segura, E., C. Nicco, B. Lombard, P. Veron, G. Raposo, F. Batteux, S. Amigorena,

and C. Thery. (2005) ICAM-1 on exosomes from mature dendritic cells is critical

for efficient naive T-cell priming. Blood. 106(1). 216-223

116

142. Mears, R., R.A. Craven, S. Hanrahan, N. Totty, C. Upton, S.L. Young, P. Patel, P.J.

Selby, and R.E. Banks. (2004) Proteomic analysis of melanoma-derived exosomes

by two-dimensional polyacrylamide gel electrophoresis and mass spectrometry.

Proteomics. 4(12). 4019-4031

143. Hegmans, J.P., M.P. Bard, A. Hemmes, T.M. Luider, M.J. Kleijmeer, J.B. Prins, L.

Zitvogel, S.A. Burgers, H.C. Hoogsteden, and B.N. Lambrecht. (2004) Proteomic

analysis of exosomes secreted by human mesothelioma cells. Am J Pathol. 164(5).

1807-1815

144. Epple, L.M., S.G. Griffiths, A.M. Dechkovskaia, N.L. Dusto, J. White, R.J.

Ouellette, T.J. Anchordoquy, L.T. Bemis, and M.W. Graner. (2012)

Medulloblastoma exosome proteomics yield functional roles for extracellular

vesicles. PLoS One. 7(7). e42064

145. Welton, J.L., S. Khanna, P.J. Giles, P. Brennan, I.A. Brewis, J. Staffurth, M.D.

Mason, and A. Clayton. (2010) Proteomics analysis of bladder cancer exosomes.

Mol Cell Proteomics. 9(6). 1324-1338

146. Admyre, C., S.M. Johansson, K.R. Qazi, J.J. Filen, R. Lahesmaa, M. Norman, E.P.

Neve, A. Scheynius, and S. Gabrielsson. (2007) Exosomes with immune

modulatory features are present in human breast milk. J Immunol. 179(3). 1969-

1978

147. Bard, M.P., J.P. Hegmans, A. Hemmes, T.M. Luider, R. Willemsen, L.A.

Severijnen, J.P. van Meerbeeck, S.A. Burgers, H.C. Hoogsteden, and B.N.

Lambrecht. (2004) Proteomic analysis of exosomes isolated from human malignant

pleural effusions. Am J Respir Cell Mol Biol. 31(1). 114-121

148. Poliakov, A., M. Spilman, T. Dokland, C.L. Amling, and J.A. Mobley. (2009)

Structural heterogeneity and protein composition of exosome-like vesicles

(prostasomes) in human semen. Prostate. 69(2). 159-167

149. Raimondo, F., S. Corbetta, L. Morosi, C. Chinello, E. Gianazza, G. Castoldi, C. Di

Gioia, C. Bombardi, A. Stella, C. Battaglia, C. Bianchi, F. Magni, and M. Pitto.

(2013) Urinary exosomes and diabetic nephropathy: a proteomic approach. Mol

Biosyst. 9(6). 1139-1146

117

150. Hurley, J.H. (2008) ESCRT complexes and the biogenesis of multivesicular bodies.

Curr Opin Cell Biol. 20(1). 4-11

151. Fehon, R.G., A.I. McClatchey, and A. Bretscher. (2010) Organizing the cell cortex:

the role of ERM proteins. Nat Rev Mol Cell Biol. 11(4). 276-287

152. Savina, A., C.M. Fader, M.T. Damiani, and M.I. Colombo. (2005) Rab11 promotes

docking and fusion of multivesicular bodies in a calcium-dependent manner.

Traffic. 6(2). 131-143

153. Subra, C., K. Laulagnier, B. Perret, and M. Record. (2007) Exosome lipidomics

unravels lipid sorting at the level of multivesicular bodies. Biochimie. 89(2). 205-

212

154. Heath, J.K., S.J. White, C.N. Johnstone, B. Catimel, R.J. Simpson, R.L. Moritz,

G.F. Tu, H. Ji, R.H. Whitehead, L.C. Groenen, A.M. Scott, G. Ritter, L. Cohen, S.

Welt, L.J. Old, E.C. Nice, and A.W. Burgess. (1997) The human A33 antigen is a

transmembrane glycoprotein and a novel member of the immunoglobulin

superfamily. Proc Natl Acad Sci U S A. 94(2). 469-474

155. Ji, H., R.L. Moritz, G.E. Reid, G. Ritter, B. Catimel, E. Nice, J.K. Heath, S.J.

White, S. Welt, L.J. Old, A.W. Burgess, and R.J. Simpson. (1997) Electrophoretic

analysis of the novel antigen for the gastrointestinal-specific monoclonal antibody,

A33. Electrophoresis. 18(3-4). 614-621

156. Ritter, G., L.S. Cohen, E.C. Nice, B. Catimel, A.W. Burgess, R.L. Moritz, H. Ji,

J.K. Heath, S.J. White, S. Welt, L.J. Old, and R.J. Simpson. (1997) Characterization

of posttranslational modifications of human A33 antigen, a novel palmitoylated

surface glycoprotein of human gastrointestinal epithelium. Biochem Biophys Res

Commun. 236(3). 682-686

157. Record, M., C. Subra, S. Silvente-Poirot, and M. Poirot. (2011) Exosomes as

intercellular signalosomes and pharmacological effectors. Biochem Pharmacol.

81(10). 1171-1182

158. Matsuo, H., J. Chevallier, N. Mayran, I. Le Blanc, C. Ferguson, J. Faure, N.S.

Blanc, S. Matile, J. Dubochet, R. Sadoul, R.G. Parton, F. Vilbois, and J. Gruenberg.

(2004) Role of LBPA and Alix in multivesicular liposome formation and endosome

organization. Science. 303(5657). 531-534

118

159. Subra, C., D. Grand, K. Laulagnier, A. Stella, G. Lambeau, M. Paillasse, P. De

Medina, B. Monsarrat, B. Perret, S. Silvente-Poirot, M. Poirot, and M. Record.

(2010) Exosomes account for vesicle-mediated transcellular transport of activatable

phospholipases and prostaglandins. J Lipid Res. 51(8). 2105-2120

160. Laulagnier, K., C. Motta, S. Hamdi, S. Roy, F. Fauvelle, J.F. Pageaux, T.

Kobayashi, J.P. Salles, B. Perret, C. Bonnerot, and M. Record. (2004) Mast cell-

and dendritic cell-derived exosomes display a specific lipid composition and an

unusual membrane organization. Biochem J. 380(Pt 1). 161-171

161. Fhaner, C.J., S. Liu, H. Ji, R.J. Simpson, and G.E. Reid. (2012) Comprehensive

lipidome profiling of isogenic primary and metastatic colon adenocarcinoma cell

lines. Anal Chem. 84(21). 8917-8926

162. Calin, G.A. and C.M. Croce. (2006) MicroRNA signatures in human cancers. Nat

Rev Cancer. 6(11). 857-866

163. Baj-Krzyworzeka, M., R. Szatanek, K. Weglarczyk, J. Baran, B. Urbanowicz, P.

Branski, M.Z. Ratajczak, and M. Zembala. (2006) Tumour-derived microvesicles

carry several surface determinants and mRNA of tumour cells and transfer some of

these determinants to monocytes. Cancer Immunol Immunother. 55(7). 808-818

164. Ratajczak, J., K. Miekus, M. Kucia, J. Zhang, R. Reca, P. Dvorak, and M.Z.

Ratajczak. (2006) Embryonic stem cell-derived microvesicles reprogram

hematopoietic progenitors: evidence for horizontal transfer of mRNA and protein

delivery. Leukemia. 20(5). 847-856

165. Skog, J., T. Wurdinger, S. van Rijn, D.H. Meijer, L. Gainche, M. Sena-Esteves,

W.T. Curry, Jr., B.S. Carter, A.M. Krichevsky, and X.O. Breakefield. (2008)

Glioblastoma microvesicles transport RNA and proteins that promote tumour

growth and provide diagnostic biomarkers. Nat Cell Biol. 10(12). 1470-1476

166. Bellingham, S.A., B.M. Coleman, and A.F. Hill. (2012) Small RNA deep

sequencing reveals a distinct miRNA signature released in exosomes from prion-

infected neuronal cells. Nucleic Acids Res. 40(21). 10937-10949

167. Zhou, Q., M. Li, X. Wang, Q. Li, T. Wang, Q. Zhu, X. Zhou, X. Gao, and X. Li.

(2012) Immune-related microRNAs are abundant in breast milk exosomes. Int J

Biol Sci. 8(1). 118-123

119

168. Hannafon, B.N. and W.Q. Ding. (2013) Intercellular Communication by Exosome-

Derived microRNAs in Cancer. Int J Mol Sci. 14(7). 14240-14269

169. Ohshima, K., K. Inoue, A. Fujiwara, K. Hatakeyama, K. Kanto, Y. Watanabe, K.

Muramatsu, Y. Fukuda, S. Ogura, K. Yamaguchi, and T. Mochizuki. (2010) Let-7

microRNA family is selectively secreted into the extracellular environment via

exosomes in a metastatic gastric cancer cell line. PLoS One. 5(10). e13247

170. Yang, M., J. Chen, F. Su, B. Yu, L. Lin, Y. Liu, J.D. Huang, and E. Song. (2011)

Microvesicles secreted by macrophages shuttle invasion-potentiating microRNAs

into breast cancer cells. Mol Cancer. 10. 117

171. Montecalvo, A., A.T. Larregina, W.J. Shufesky, D.B. Stolz, M.L. Sullivan, J.M.

Karlsson, C.J. Baty, G.A. Gibson, G. Erdos, Z. Wang, J. Milosevic, O.A. Tkacheva,

S.J. Divito, R. Jordan, J. Lyons-Weiler, S.C. Watkins, and A.E. Morelli. (2012)

Mechanism of transfer of functional microRNAs between mouse dendritic cells via

exosomes. Blood. 119(3). 756-766

172. Umezu, T., K. Ohyashiki, M. Kuroda, and J.H. Ohyashiki. (2013) Leukemia cell to

endothelial cell communication via exosomal miRNAs. Oncogene. 32(22). 2747-

2755

173. Rana, S., K. Malinowska, and M. Zoller. (2013) Exosomal tumor microRNA

modulates premetastatic organ cells. Neoplasia. 15(3). 281-295

174. Grange, C., M. Tapparo, F. Collino, L. Vitillo, C. Damasco, M.C. Deregibus, C.

Tetta, B. Bussolati, and G. Camussi. (2011) Microvesicles released from human

renal cancer stem cells stimulate angiogenesis and formation of lung premetastatic

niche. Cancer Res. 71(15). 5346-5356

175. Batagov, A.O., V.A. Kuznetsov, and I.V. Kurochkin. (2011) Identification of

nucleotide patterns enriched in secreted RNAs as putative cis-acting elements

targeting them to exosome nano-vesicles. BMC Genomics. 12 Suppl 3. S18

176. Bolukbasi, M.F., A. Mizrak, G.B. Ozdener, S. Madlener, T. Strobel, E.P. Erkan,

J.B. Fan, X.O. Breakefield, and O. Saydam. (2012) miR-1289 and "Zipcode"-like

Sequence Enrich mRNAs in Microvesicles. Mol Ther Nucleic Acids. 1. e10

120

177. Rana, S., S. Yue, D. Stadel, and M. Zoller. (2012) Toward tailored exosomes: the

exosomal tetraspanin web contributes to target cell selection. Int J Biochem Cell

Biol. 44(9). 1574-1584

178. Morelli, A.E., A.T. Larregina, W.J. Shufesky, M.L. Sullivan, D.B. Stolz, G.D.

Papworth, A.F. Zahorchak, A.J. Logar, Z. Wang, S.C. Watkins, L.D. Falo, Jr., and

A.W. Thomson. (2004) Endocytosis, intracellular sorting, and processing of

exosomes by dendritic cells. Blood. 104(10). 3257-3266

179. Segura, E., C. Guerin, N. Hogg, S. Amigorena, and C. Thery. (2007) CD8+

dendritic cells use LFA-1 to capture MHC-peptide complexes from exosomes in

vivo. J Immunol. 179(3). 1489-1496

180. Hanayama, R., M. Tanaka, K. Miwa, A. Shinohara, A. Iwamatsu, and S. Nagata.

(2002) Identification of a factor that links apoptotic cells to phagocytes. Nature.

417(6885). 182-187

181. Parolini, I., C. Federici, C. Raggi, L. Lugini, S. Palleschi, A. De Milito, C. Coscia,

E. Iessi, M. Logozzi, A. Molinari, M. Colone, M. Tatti, M. Sargiacomo, and S. Fais.

(2009) Microenvironmental pH is a key factor for exosome traffic in tumor cells. J

Biol Chem. 284(49). 34211-34222

182. Escrevente, C., S. Keller, P. Altevogt, and J. Costa. (2011) Interaction and uptake of

exosomes by ovarian cancer cells. BMC Cancer. 11. 108

183. Feng, D., W.L. Zhao, Y.Y. Ye, X.C. Bai, R.Q. Liu, L.F. Chang, Q. Zhou, and S.F.

Sui. (2010) Cellular internalization of exosomes occurs through phagocytosis.

Traffic. 11(5). 675-687

184. Fitzner, D., M. Schnaars, D. van Rossum, G. Krishnamoorthy, P. Dibaj, M. Bakhti,

T. Regen, U.K. Hanisch, and M. Simons. (2011) Selective transfer of exosomes

from oligodendrocytes to microglia by macropinocytosis. J Cell Sci. 124(Pt 3). 447-

458

185. Tian, T., Y. Wang, H. Wang, Z. Zhu, and Z. Xiao. (2010) Visualizing of the cellular

uptake and intracellular trafficking of exosomes by live-cell microscopy. J Cell

Biochem. 111(2). 488-496

186. Ewers, H. and A. Helenius. (2011) Lipid-mediated endocytosis. Cold Spring Harb

Perspect Biol. 3(8). a004721

121

187. Svensson, K.J., H.C. Christianson, A. Wittrup, E. Bourseau-Guilmain, E. Lindqvist,

L.M. Svensson, M. Morgelin, and M. Belting. (2013) Exosome uptake depends on

ERK1/2-heat shock protein 27 signalling and lipid raft-mediated endocytosis

negatively regulated by caveolin-1. J Biol Chem,

188. Kucerova, L. and S. Skolekova. (2013) Tumor microenvironment and the role of

mesenchymal stromal cells. Neoplasma. 60(1). 1-10

189. Hanahan, D. and R.A. Weinberg. (2011) Hallmarks of cancer: the next generation.

Cell. 144(5). 646-674

190. Ge, R., E. Tan, S. Sharghi-Namini, and H.H. Asada. (2012) Exosomes in Cancer

Microenvironment and Beyond: have we Overlooked these Extracellular

Messengers? Cancer Microenviron. 5(3). 323-332

191. Zhang, H.G. and W.E. Grizzle. (2011) Exosomes and cancer: a newly described

pathway of immune suppression. Clin Cancer Res. 17(5). 959-964

192. Hood, J.L., H. Pan, G.M. Lanza, and S.A. Wickline. (2009) Paracrine induction of

endothelium by tumor exosomes. Lab Invest. 89(11). 1317-1328

193. Webber, J., R. Steadman, M.D. Mason, Z. Tabi, and A. Clayton. (2010) Cancer

exosomes trigger fibroblast to myofibroblast differentiation. Cancer Res. 70(23).

9621-9630

194. Hao, S., Z. Ye, F. Li, Q. Meng, M. Qureshi, J. Yang, and J. Xiang. (2006)

Epigenetic transfer of metastatic activity by uptake of highly metastatic B16

melanoma cell-released exosomes. Exp Oncol. 28(2). 126-131

195. Peinado, H., M. Aleckovic, S. Lavotshkin, I. Matei, B. Costa-Silva, G. Moreno-

Bueno, M. Hergueta-Redondo, C. Williams, G. Garcia-Santos, C.M. Ghajar, A.

Nitadori-Hoshino, C. Hoffman, K. Badal, B.A. Garcia, M.K. Callahan, J. Yuan,

V.R. Martins, J. Skog, R.N. Kaplan, M.S. Brady, J.D. Wolchok, P.B. Chapman, Y.

Kang, J. Bromberg, and D. Lyden. (2012) Melanoma exosomes educate bone

marrow progenitor cells toward a pro-metastatic phenotype through MET. Nat Med.

18(6). 883-891

196. Rak, J. and A. Guha. (2012) Extracellular vesicles--vehicles that spread cancer

genes. Bioessays. 34(6). 489-497

122

197. Demory Beckler, M., J.N. Higginbotham, J.L. Franklin, A.J. Ham, P.J. Halvey, I.E.

Imasuen, C. Whitwell, M. Li, D.C. Liebler, and R.J. Coffey. (2013) Proteomic

analysis of exosomes from mutant KRAS colon cancer cells identifies intercellular

transfer of mutant KRAS. Mol Cell Proteomics. 12(2). 343-355

198. Luga, V., L. Zhang, A.M. Viloria-Petit, A.A. Ogunjimi, M.R. Inanlou, E. Chiu, M.

Buchanan, A.N. Hosein, M. Basik, and J.L. Wrana. (2012) Exosomes mediate

stromal mobilization of autocrine Wnt-PCP signaling in breast cancer cell

migration. Cell. 151(7). 1542-1556

199. Al-Nedawi, K., B. Meehan, J. Micallef, V. Lhotak, L. May, A. Guha, and J. Rak.

(2008) Intercellular transfer of the oncogenic receptor EGFRvIII by microvesicles

derived from tumour cells. Nat Cell Biol. 10(5). 619-624

200. Al-Nedawi, K., B. Meehan, R.S. Kerbel, A.C. Allison, and J. Rak. (2009)

Endothelial expression of autocrine VEGF upon the uptake of tumor-derived

microvesicles containing oncogenic EGFR. Proc Natl Acad Sci U S A. 106(10).

3794-3799

201. Antonyak, M.A., B. Li, L.K. Boroughs, J.L. Johnson, J.E. Druso, K.L. Bryant, D.A.

Holowka, and R.A. Cerione. (2011) Cancer cell-derived microvesicles induce

transformation by transferring tissue transglutaminase and fibronectin to recipient

cells. Proc Natl Acad Sci U S A. 108(12). 4852-4857

202. Balaj, L., R. Lessard, L. Dai, Y.J. Cho, S.L. Pomeroy, X.O. Breakefield, and J.

Skog. (2011) Tumour microvesicles contain retrotransposon elements and amplified

oncogene sequences. Nat Commun. 2. 180

203. Meckes, D.G., Jr., K.H. Shair, A.R. Marquitz, C.P. Kung, R.H. Edwards, and N.

Raab-Traub. (2010) Human tumor virus utilizes exosomes for intercellular

communication. Proc Natl Acad Sci U S A. 107(47). 20370-20375

204. Janowska-Wieczorek, A., L.A. Marquez-Curtis, M. Wysoczynski, and M.Z.

Ratajczak. (2006) Enhancing effect of platelet-derived microvesicles on the invasive

potential of breast cancer cells. Transfusion. 46(7). 1199-1209

123

205. Janowska-Wieczorek, A., M. Wysoczynski, J. Kijowski, L. Marquez-Curtis, B.

Machalinski, J. Ratajczak, and M.Z. Ratajczak. (2005) Microvesicles derived from

activated platelets induce metastasis and angiogenesis in lung cancer. Int J Cancer.

113(5). 752-760

206. Fidler, I.J., S. Yano, R.D. Zhang, T. Fujimaki, and C.D. Bucana. (2002) The seed

and soil hypothesis: vascularisation and brain metastases. Lancet Oncol. 3(1). 53-57

207. Schluter, K., P. Gassmann, A. Enns, T. Korb, A. Hemping-Bovenkerk, J. Holzen,

and J. Haier. (2006) Organ-specific metastatic tumor cell adhesion and

extravasation of colon carcinoma cells with different metastatic potential. Am J

Pathol. 169(3). 1064-1073

208. Kaplan, R.N., S. Rafii, and D. Lyden. (2006) Preparing the "soil": the premetastatic

niche. Cancer Res. 66(23). 11089-11093

209. S, E.L.A., I. Mager, X.O. Breakefield, and M.J. Wood. (2013) Extracellular

vesicles: biology and emerging therapeutic opportunities. Nat Rev Drug Discov.

12(5). 347-357

210. El-Andaloussi, S., Y. Lee, S. Lakhal-Littleton, J. Li, Y. Seow, C. Gardiner, L.

Alvarez-Erviti, I.L. Sargent, and M.J. Wood. (2012) Exosome-mediated delivery of

siRNA in vitro and in vivo. Nat Protoc. 7(12). 2112-2126

211. Chalmin, F., S. Ladoire, G. Mignot, J. Vincent, M. Bruchard, J.P. Remy-Martin, W.

Boireau, A. Rouleau, B. Simon, D. Lanneau, A. De Thonel, G. Multhoff, A.

Hamman, F. Martin, B. Chauffert, E. Solary, L. Zitvogel, C. Garrido, B. Ryffel, C.

Borg, L. Apetoh, C. Rebe, and F. Ghiringhelli. (2010) Membrane-associated Hsp72

from tumor-derived exosomes mediates STAT3-dependent immunosuppressive

function of mouse and human myeloid-derived suppressor cells. J Clin Invest.

120(2). 457-471

212. Lima, L.G., R. Chammas, R.Q. Monteiro, M.E. Moreira, and M.A. Barcinski.

(2009) Tumor-derived microvesicles modulate the establishment of metastatic

melanoma in a phosphatidylserine-dependent manner. Cancer Lett. 283(2). 168-175

213. Hood, J.L. and S.A. Wickline. (2012) A systematic approach to exosome-based

translational nanomedicine. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 4(4).

458-467

124

214. Seow, Y. and M.J. Wood. (2009) Biological gene delivery vehicles: beyond viral

vectors. Mol Ther. 17(5). 767-777

215. O'Loughlin, A.J., C.A. Woffindale, and M.J. Wood. (2012) Exosomes and the

emerging field of exosome-based gene therapy. Curr Gene Ther. 12(4). 262-274

216. Alvarez-Erviti, L., Y. Seow, H. Yin, C. Betts, S. Lakhal, and M.J. Wood. (2011)

Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes.

Nat Biotechnol. 29(4). 341-345

217. Chaput, N., J. Taieb, N. Schartz, C. Flament, S. Novault, F. Andre, and L. Zitvogel.

(2005) The potential of exosomes in immunotherapy of cancer. Blood Cells Mol

Dis. 35(2). 111-115

218. Zitvogel, L., A. Regnault, A. Lozier, J. Wolfers, C. Flament, D. Tenza, P. Ricciardi-

Castagnoli, G. Raposo, and S. Amigorena. (1998) Eradication of established murine

tumors using a novel cell-free vaccine: dendritic cell-derived exosomes. Nat Med.

4(5). 594-600

219. Monleon, I., M.J. Martinez-Lorenzo, L. Monteagudo, P. Lasierra, M. Taules, M.

Iturralde, A. Pineiro, L. Larrad, M.A. Alava, J. Naval, and A. Anel. (2001)

Differential secretion of Fas ligand- or APO2 ligand/TNF-related apoptosis-

inducing ligand-carrying microvesicles during activation-induced death of human T

cells. J Immunol. 167(12). 6736-6744

220. Kim, S.H., E.R. Lechman, N. Bianco, R. Menon, A. Keravala, J. Nash, Z. Mi, S.C.

Watkins, A. Gambotto, and P.D. Robbins. (2005) Exosomes derived from IL-10-

treated dendritic cells can suppress inflammation and collagen-induced arthritis. J

Immunol. 174(10). 6440-6448

221. Lai, R.C., R.W. Yeo, K.H. Tan, and S.K. Lim. (2013) Exosomes for drug delivery -

a novel application for the mesenchymal stem cell. Biotechnol Adv. 31(5). 543-551

222. Wahlgren, J., L.K.T. De, M. Brisslert, F. Vaziri Sani, E. Telemo, P. Sunnerhagen,

and H. Valadi. (2012) Plasma exosomes can deliver exogenous short interfering

RNA to monocytes and lymphocytes. Nucleic Acids Res. 40(17). e130

125

223. Thery, C., A. Regnault, J. Garin, J. Wolfers, L. Zitvogel, P. Ricciardi-Castagnoli, G.

Raposo, and S. Amigorena. (1999) Molecular characterization of dendritic cell-

derived exosomes. Selective accumulation of the heat shock protein hsc73. J Cell

Biol. 147(3). 599-610

224. Dobrovolskaia, M.A., P. Aggarwal, J.B. Hall, and S.E. McNeil. (2008) Preclinical

studies to understand nanoparticle interaction with the immune system and its

potential effects on nanoparticle biodistribution. Mol Pharm. 5(4). 487-495

225. van den Boorn, J.G., M. Schlee, C. Coch, and G. Hartmann. (2011) SiRNA delivery

with exosome nanoparticles. Nat Biotechnol. 29(4). 325-326

226. Escudier, B., T. Dorval, N. Chaput, F. Andre, M.P. Caby, S. Novault, C. Flament,

C. Leboulaire, C. Borg, S. Amigorena, C. Boccaccio, C. Bonnerot, O. Dhellin, M.

Movassagh, S. Piperno, C. Robert, V. Serra, N. Valente, J.B. Le Pecq, A. Spatz, O.

Lantz, T. Tursz, E. Angevin, and L. Zitvogel. (2005) Vaccination of metastatic

melanoma patients with autologous dendritic cell (DC) derived-exosomes: results of

the first phase I clinical trial. J Transl Med. 3(1). 10

227. Chen, T.S., R.C. Lai, M.M. Lee, A.B. Choo, C.N. Lee, and S.K. Lim. (2010)

Mesenchymal stem cell secretes microparticles enriched in pre-microRNAs. Nucleic

Acids Res. 38(1). 215-224

228. Lai, R.C., T.S. Chen, and S.K. Lim. (2011) Mesenchymal stem cell exosome: a

novel stem cell-based therapy for cardiovascular disease. Regen Med. 6(4). 481-492

229. Kooijmans, S.A., P. Vader, S.M. van Dommelen, W.W. van Solinge, and R.M.

Schiffelers. (2012) Exosome mimetics: a novel class of drug delivery systems. Int J

Nanomedicine. 7. 1525-1541

230. Puri, A., K. Loomis, B. Smith, J.H. Lee, A. Yavlovich, E. Heldman, and R.

Blumenthal. (2009) Lipid-based nanoparticles as pharmaceutical drug carriers: from

concepts to clinic. Crit Rev Ther Drug Carrier Syst. 26(6). 523-580

231. Fenske, D.B. and P.R. Cullis. (2008) Liposomal nanomedicines. Expert Opin Drug

Deliv. 5(1). 25-44

232. De La Pena, H., J.A. Madrigal, S. Rusakiewicz, M. Bencsik, G.W. Cave, A.

Selman, R.C. Rees, P.J. Travers, and I.A. Dodi. (2009) Artificial exosomes as tools

for basic and clinical immunology. J Immunol Methods. 344(2). 121-132

126

233. Delcayre, A., A. Estelles, J. Sperinde, T. Roulon, P. Paz, B. Aguilar, J. Villanueva,

S. Khine, and J.B. Le Pecq. (2005) Exosome Display technology: applications to the

development of new diagnostics and therapeutics. Blood Cells Mol Dis. 35(2). 158-

168

234. Hemler, M.E. (2005) Tetraspanin functions and associated microdomains. Nat Rev

Mol Cell Biol. 6(10). 801-811

235. Varnier, A., F. Kermarrec, I. Blesneac, C. Moreau, L. Liguori, J.L. Lenormand, and

N. Picollet-D'hahan. (2010) A simple method for the reconstitution of membrane

proteins into giant unilamellar vesicles. J Membr Biol. 233(1-3). 85-92

236. Chen, T.S., F. Arslan, Y. Yin, S.S. Tan, R.C. Lai, A.B. Choo, J. Padmanabhan, C.N.

Lee, D.P. de Kleijn, and S.K. Lim. (2011) Enabling a robust scalable manufacturing

process for therapeutic exosomes through oncogenic immortalization of human

ESC-derived MSCs. J Transl Med. 9. 47

237. Zhang, Y., D. Liu, X. Chen, J. Li, L. Li, Z. Bian, F. Sun, J. Lu, Y. Yin, X. Cai, Q.

Sun, K. Wang, Y. Ba, Q. Wang, D. Wang, J. Yang, P. Liu, T. Xu, Q. Yan, J. Zhang,

K. Zen, and C.Y. Zhang. (2010) Secreted monocytic miR-150 enhances targeted

endothelial cell migration. Mol Cell. 39(1). 133-144

238. Maguire, C.A., L. Balaj, S. Sivaraman, M.H. Crommentuijn, M. Ericsson, L.

Mincheva-Nilsson, V. Baranov, D. Gianni, B.A. Tannous, M. Sena-Esteves, X.O.

Breakefield, and J. Skog. (2012) Microvesicle-associated AAV vector as a novel

gene delivery system. Mol Ther. 20(5). 960-971

239. Ohno, S., M. Takanashi, K. Sudo, S. Ueda, A. Ishikawa, N. Matsuyama, K. Fujita,

T. Mizutani, T. Ohgi, T. Ochiya, N. Gotoh, and M. Kuroda. (2013) Systemically

injected exosomes targeted to EGFR deliver antitumor microRNA to breast cancer

cells. Mol Ther. 21(1). 185-191

240. Noerholm, M., L. Balaj, T. Limperg, A. Salehi, L.D. Zhu, F.H. Hochberg, X.O.

Breakefield, B.S. Carter, and J. Skog. (2012) RNA expression patterns in serum

microvesicles from patients with glioblastoma multiforme and controls. BMC

Cancer. 12(22). 1-11

127

241. Mitchell, P.S., R.K. Parkin, E.M. Kroh, B.R. Fritz, S.K. Wyman, E.L. Pogosova-

Agadjanyan, A. Peterson, J. Noteboom, K.C. O'Briant, A. Allen, D.W. Lin, N.

Urban, C.W. Drescher, B.S. Knudsen, D.L. Stirewalt, R. Gentleman, R.L. Vessella,

P.S. Nelson, D.B. Martin, and M. Tewari. (2008) Circulating microRNAs as stable

blood-based markers for cancer detection. Proc Natl Acad Sci U S A. 105(30).

10513-10518

242. Raposo, G., H.W. Nijman, W. Stoorvogel, R. Liejendekker, C.V. Harding, C.J.

Melief, and H.J. Geuze. (1996) B lymphocytes secrete antigen-presenting vesicles. J

Exp Med. 183(3). 1161-1172

243. Cantin, R., J. Diou, D. Belanger, A.M. Tremblay, and C. Gilbert. (2008)

Discrimination between exosomes and HIV-1: purification of both vesicles from

cell-free supernatants. J Immunol Methods. 338(1-2). 21-30

244. Lamparski, H.G., A. Metha-Damani, J.Y. Yao, S. Patel, D.H. Hsu, C. Ruegg, and

J.B. Le Pecq. (2002) Production and characterization of clinical grade exosomes

derived from dendritic cells. J Immunol Methods. 270(2). 211-226

245. Nguyen, D.G., A. Booth, S.J. Gould, and J.E. Hildreth. (2003) Evidence that HIV

budding in primary macrophages occurs through the exosome release pathway. J

Biol Chem. 278(52). 52347-52354

246. Andersen, R., J. Hosaka, T.S. Halstensen, A. Stordahl, A. Tverdal, L. Aabakken,

and F. Laerum. (1996) The X-ray contrast medium iodixanol detects increased

colonic permeability equally well as 51Cr-labeled ethylenediaminetetraacetic acid

in experimental colitis of rats. Scand J Gastroenterol. 31(2). 140-146

247. Spencer, C.M. and K.L. Goa. (1996) Iodixanol. A review of its pharmacodynamic

and pharmacokinetic properties and diagnostic use as an x-ray contrast medium.

Drugs. 52(6). 899-927

248. Borrell, B. (2010) How accurate are cancer cell lines? Nature. 463(7283). 858

249. Anglard, P., E. Trahan, S. Liu, F. Latif, M.J. Merino, M.I. Lerman, B. Zbar, and

W.M. Linehan. (1992) Molecular and cellular characterization of human renal cell

carcinoma cell lines. Cancer Res. 52(2). 348-356

128

250. Mathias, R.A., B. Wang, H. Ji, E.A. Kapp, R.L. Moritz, H.J. Zhu, and R.J. Simpson.

(2009) Secretome-based proteomic profiling of Ras-transformed MDCK cells

reveals extracellular modulators of epithelial-mesenchymal transition. J Proteome

Res. 8(6). 2827-2837

251. Whitehead, R.H., J.K. Jones, A. Gabriel, and R.E. Lukies. (1987) A new colon

carcinoma cell line (LIM1863) that grows as organoids with spontaneous

differentiation into crypt-like structures in vitro. Cancer Res. 47(10). 2683-2689

252. Sancho, E., E. Batlle, and H. Clevers. (2003) Live and let die in the intestinal

epithelium. Curr Opin Cell Biol. 15(6). 763-770

253. Humphries, A. and N.A. Wright. (2008) Colonic crypt organization and

tumorigenesis. Nat Rev Cancer. 8(6). 415-424

254. Rizk, P. and N. Barker. (2012) Gut stem cells in tissue renewal and disease:

methods, markers, and myths. Wiley Interdiscip Rev Syst Biol Med. 4(5). 475-496

255. Ferlay, J., H.R. Shin, F. Bray, D. Forman, C. Mathers, and D.M. Parkin. (2010)

Estimates of worldwide burden of cancer in 2008: GLOBOCAN 2008. Int J

Cancer. 127(12). 2893-2917

256. McDermid, I. (2005) Cancer incidence projections Australia 2002 to 2011.

Australian Institute of Health and Welfare. 1-166

257. Gomez-Dominguez, E., M. Trapero-Marugan, A.J. del Pozo, J. Cantero, J.P.

Gisbert, and J. Mate. (2006) The colorectal carcinoma prognosis factors.

Significance of diagnosis delay. Rev Esp Enferm Dig. 98(5). 322-329

258. Cserni, G. (2003) Nodal staging of colorectal carcinomas and sentinel nodes. J Clin

Pathol. 56(5). 327-335

259. Steele, G.D., Jr. (1994) The National Cancer Data Base report on colorectal cancer.

Cancer. 74(7). 1979-1989

260. Whitlock, E.P., J. Lin, E. Liles, T. Beil, R. Fu, E. O'Connor, R.N. Thompson, and T.

Cardenas, in Screening for Colorectal Cancer: An Updated Systematic Review2008:

Rockville (MD).

261. Hamilton, S.R. (1992) The adenoma-adenocarcinoma sequence in the large bowel:

variations on a theme. J Cell Biochem Suppl. 16G. 41-46

129

262. Calvert, P.M. and H. Frucht. (2002) The genetics of colorectal cancer. Ann Intern

Med. 137(7). 603-612

263. Fearon, E.R. and B. Vogelstein. (1990) A genetic model for colorectal

tumorigenesis. Cell. 61(5). 759-767

264. Debinski, H.S., S. Love, A.D. Spigelman, and R.K. Phillips. (1996) Colorectal

polyp counts and cancer risk in familial adenomatous polyposis. Gastroenterology.

110(4). 1028-1030

265. Akiyama, T. (2000) Wnt/beta-catenin signaling. Cytokine Growth Factor Rev.

11(4). 273-282

266. MacDonald, B.T., K. Tamai, and X. He. (2009) Wnt/beta-catenin signaling:

components, mechanisms, and diseases. Dev Cell. 17(1). 9-26

267. Fishel, R., M.K. Lescoe, M.R. Rao, N.G. Copeland, N.A. Jenkins, J. Garber, M.

Kane, and R. Kolodner. (1993) The human mutator gene homolog MSH2 and its

association with hereditary nonpolyposis colon cancer. Cell. 75(5). 1027-1038

268. Muller, A. and R. Fishel. (2002) Mismatch repair and the hereditary non-polyposis

colorectal cancer syndrome (HNPCC). Cancer Invest. 20(1). 102-109

269. Hayward, I.P. and R.H. Whitehead. (1992) Patterns of growth and differentiation in

the colon carcinoma cell line LIM 1863. Int J Cancer. 50(5). 752-759

270. Kalluri, R. and R.A. Weinberg. (2009) The basics of epithelial-mesenchymal

transition. J Clin Invest. 119(6). 1420-1428

271. Thiery, J.P. and J.P. Sleeman. (2006) Complex networks orchestrate epithelial-

mesenchymal transitions. Nat Rev Mol Cell Biol. 7(2). 131-142

272. Greenburg, G. and E.D. Hay. (1982) Epithelia suspended in collagen gels can lose

polarity and express characteristics of migrating mesenchymal cells. J Cell Biol.

95(1). 333-339

273. Thiery, J.P. (2003) Epithelial-mesenchymal transitions in development and

pathologies. Curr Opin Cell Biol. 15(6). 740-746

274. Xu, J., S. Lamouille, and R. Derynck. (2009) TGF-beta-induced epithelial to

mesenchymal transition. Cell Res. 19(2). 156-172

130

275. Grunert, S., M. Jechlinger, and H. Beug. (2003) Diverse cellular and molecular

mechanisms contribute to epithelial plasticity and metastasis. Nat Rev Mol Cell

Biol. 4(8). 657-665

276. Thiery, J.P. (2002) Epithelial-mesenchymal transitions in tumour progression. Nat

Rev Cancer. 2(6). 442-454

277. Thompson, E.W., D.F. Newgreen, and D. Tarin. (2005) Carcinoma invasion and

metastasis: a role for epithelial-mesenchymal transition? Cancer Res. 65(14). 5991-

5995

278. Egeblad, M. and Z. Werb. (2002) New functions for the matrix metalloproteinases

in cancer progression. Nat Rev Cancer. 2(3). 161-174

279. Chang, C. and Z. Werb. (2001) The many faces of metalloproteases: cell growth,

invasion, angiogenesis and metastasis. Trends Cell Biol. 11(11). S37-S43

280. Bissell, M.J., D.C. Radisky, A. Rizki, V.M. Weaver, and O.W. Petersen. (2002) The

organizing principle: microenvironmental influences in the normal and malignant

breast. Differentiation. 70(9-10). 537-546

281. Yang, J. and R.A. Weinberg. (2008) Epithelial-mesenchymal transition: at the

crossroads of development and tumor metastasis. Dev Cell. 14(6). 818-829

282. Guarino, M., B. Rubino, and G. Ballabio. (2007) The role of epithelial-

mesenchymal transition in cancer pathology. Pathology. 39(3). 305-318

283. Thiery, J.P. (2002) Epithelial-mesenchymal transitions in tumour progression. Nat

Rev Cancer. 2(6). 442-454

284. Yilmaz, M. and G. Christofori. (2009) EMT, the cytoskeleton, and cancer cell

invasion. Cancer Metastasis Rev. 28(1-2). 15-33

285. Canel, M., A. Serrels, M.C. Frame, and V.G. Brunton. (2013) E-cadherin-integrin

crosstalk in cancer invasion and metastasis. J Cell Sci. 126(Pt 2). 393-401

286. Savagner, P. (2010) The epithelial-mesenchymal transition (EMT) phenomenon.

Ann Oncol. 21 Suppl 7. vii89-92

287. Moustakas, A. and C.H. Heldin. (2007) Signaling networks guiding epithelial-

mesenchymal transitions during embryogenesis and cancer progression. Cancer Sci.

98(10). 1512-1520

131

288. Kokkinos, M.I., R. Wafai, M.K. Wong, D.F. Newgreen, E.W. Thompson, and M.

Waltham. (2007) Vimentin and epithelial-mesenchymal transition in human breast

cancer--observations in vitro and in vivo. Cells Tissues Organs. 185(1-3). 191-203

289. Radisky, D.C. (2005) Epithelial-mesenchymal transition. J Cell Sci. 118(Pt 19).

4325-4326

290. Brattain, M.G., G. Howell, L.Z. Sun, and J.K. Willson. (1994) Growth factor

balance and tumor progression. Curr Opin Oncol. 6(1). 77-81

291. Strutz, F., M. Zeisberg, F.N. Ziyadeh, C.Q. Yang, R. Kalluri, G.A. Muller, and E.G.

Neilson. (2002) Role of basic fibroblast growth factor-2 in epithelial-mesenchymal

transformation. Kidney Int. 61(5). 1714-1728

292. Niessen, K., Y. Fu, L. Chang, P.A. Hoodless, D. McFadden, and A. Karsan. (2008)

Slug is a direct Notch target required for initiation of cardiac cushion

cellularization. J Cell Biol. 182(2). 315-325

293. Pelaez, I.M., M. Kalogeropoulou, A. Ferraro, A. Voulgari, T. Pankotai, I. Boros,

and A. Pintzas. (2010) Oncogenic RAS alters the global and gene-specific histone

modification pattern during epithelial-mesenchymal transition in colorectal

carcinoma cells. Int J Biochem Cell Biol. 42(6). 911-920

294. Giehl, K. (2005) Oncogenic Ras in tumour progression and metastasis. Biol Chem.

386(3). 193-205

295. Zondag, G.C., E.E. Evers, J.P. ten Klooster, L. Janssen, R.A. van der Kammen, and

J.G. Collard. (2000) Oncogenic Ras downregulates Rac activity, which leads to

increased Rho activity and epithelial-mesenchymal transition. J Cell Biol. 149(4).

775-782

296. Schramek, H., E. Feifel, I. Marschitz, N. Golochtchapova, G. Gstraunthaler, and R.

Montesano. (2003) Loss of active MEK1-ERK1/2 restores epithelial phenotype and

morphogenesis in transdifferentiated MDCK cells. Am J Physiol Cell Physiol.

285(3). C652-661

297. Hogan, C., S. Dupre-Crochet, M. Norman, M. Kajita, C. Zimmermann, A.E.

Pelling, E. Piddini, L.A. Baena-Lopez, J.P. Vincent, Y. Itoh, H. Hosoya, F. Pichaud,

and Y. Fujita. (2009) Characterization of the interface between normal and

transformed epithelial cells. Nat Cell Biol. 11(4). 460-467

132

298. Madin, S.H., P.C. Andriese, and N.B. Darby. (1957) The in vitro cultivation of

tissues of domestic and laboratory animals. Am J Vet Res. 18(69). 932-941

299. Gaush, C.R., W.L. Hard, and T.F. Smith. (1966) Characterization of an established

line of canine kidney cells (MDCK). Proc Soc Exp Biol Med. 122(3). 931-935

300. Mathias, R.A., Y.S. Chen, B. Wang, H. Ji, E.A. Kapp, R.L. Moritz, H.J. Zhu, and

R.J. Simpson. (2010) Extracellular remodelling during oncogenic Ras-induced

epithelial-mesenchymal transition facilitates MDCK cell migration. J Proteome Res.

9(2). 1007-1019

301. Garnier, D., N. Magnus, T.H. Lee, V. Bentley, B. Meehan, C. Milsom, L.

Montermini, T. Kislinger, and J. Rak. (2012) Cancer cells induced to express

mesenchymal phenotype release exosome-like extracellular vesicles carrying tissue

factor. J Biol Chem. 287(52). 43565-43572

302. Clark, H.F., A.L. Gurney, E. Abaya, K. Baker, D. Baldwin, J. Brush, J. Chen, B.

Chow, C. Chui, C. Crowley, B. Currell, B. Deuel, P. Dowd, D. Eaton, J. Foster, C.

Grimaldi, Q. Gu, P.E. Hass, S. Heldens, A. Huang, H.S. Kim, L. Klimowski, Y. Jin,

S. Johnson, J. Lee, L. Lewis, D. Liao, M. Mark, E. Robbie, C. Sanchez, J.

Schoenfeld, S. Seshagiri, L. Simmons, J. Singh, V. Smith, J. Stinson, A. Vagts, R.

Vandlen, C. Watanabe, D. Wieand, K. Woods, M.H. Xie, D. Yansura, S. Yi, G. Yu,

J. Yuan, M. Zhang, Z. Zhang, A. Goddard, W.I. Wood, P. Godowski, and A. Gray.

(2003) The secreted protein discovery initiative (SPDI), a large-scale effort to

identify novel human secreted and transmembrane proteins: a bioinformatics

assessment. Genome Res. 13(10). 2265-2270

303. Mathivanan, S., H. Ji, and R.J. Simpson. (2010) Exosomes: extracellular organelles

important in intercellular communication. J Proteomics. 73(10). 1907-1920

304. Bronfman, M., G. Loyola, and C.S. Koenig. (1998) Isolation of intact organelles by

differential centrifugation of digitonin-treated hepatocytes using a table Eppendorf

centrifuge. Anal Biochem. 255(2). 252-256

305. Girard, M., P.D. Allaire, F. Blondeau, and P.S. McPherson. (2005) Isolation of

clathrin-coated vesicles by differential and density gradient centrifugation. Curr

Protoc Cell Biol. Chapter 3. Unit 3 13

133

306. Ji, H., N. Erfani, B.J. Tauro, E.A. Kapp, H.J. Zhu, R.L. Moritz, J.W. Lim, and R.J.

Simpson. (2008) Difference gel electrophoresis analysis of Ras-transformed

fibroblast cell-derived exosomes. Electrophoresis. 29(12). 2660-2671

307. Mathias, R.A., J.W. Lim, H. Ji, and R.J. Simpson. (2009) Isolation of extracellular

membranous vesicles for proteomic analysis. Methods Mol Biol. 528. 227-242

308. Muralidharan-Chari, V., J. Clancy, C. Plou, M. Romao, P. Chavrier, G. Raposo, and

C. D'Souza-Schorey. (2009) ARF6-regulated shedding of tumor cell-derived plasma

membrane microvesicles. Curr Biol. 19(22). 1875-1885

309. Tan, A., H. De La Pena, and A.M. Seifalian. (2010) The application of exosomes as

a nanoscale cancer vaccine. Int J Nanomedicine. 5. 889-900

310. Martin-Belmonte, F. and M. Perez-Moreno. (2012) Epithelial cell polarity, stem

cells and cancer. Nat Rev Cancer. 12(1). 23-38

311. Went, P., M. Vasei, L. Bubendorf, L. Terracciano, L. Tornillo, U. Riede, J.

Kononen, R. Simon, G. Sauter, and P.A. Baeuerle. (2006) Frequent high-level

expression of the immunotherapeutic target Ep-CAM in colon, stomach, prostate

and lung cancers. Br J Cancer. 94(1). 128-135

312. Kuhn, S., M. Koch, T. Nubel, M. Ladwein, D. Antolovic, P. Klingbeil, D.

Hildebrand, G. Moldenhauer, L. Langbein, W.W. Franke, J. Weitz, and M. Zoller.

(2007) A complex of EpCAM, claudin-7, CD44 variant isoforms, and tetraspanins

promotes colorectal cancer progression. Mol Cancer Res. 5(6). 553-567

313. Rollins, S.A. and P.J. Sims. (1990) The complement-inhibitory activity of CD59

resides in its capacity to block incorporation of C9 into membrane C5b-9. J

Immunol. 144(9). 3478-3483

314. Riley-Vargas, R.C., D.B. Gill, C. Kemper, M.K. Liszewski, and J.P. Atkinson.

(2004) CD46: expanding beyond complement regulation. Trends Immunol. 25(9).

496-503

315. Yanai, H., T. Ban, Z. Wang, M.K. Choi, T. Kawamura, H. Negishi, M. Nakasato, Y.

Lu, S. Hangai, R. Koshiba, D. Savitsky, L. Ronfani, S. Akira, M.E. Bianchi, K.

Honda, T. Tamura, T. Kodama, and T. Taniguchi. (2009) HMGB proteins function

as universal sentinels for nucleic-acid-mediated innate immune responses. Nature.

462(7269). 99-103

134

316. Jaiswal, R., F. Luk, P.V. Dalla, G.E. Grau, and M. Bebawy. (2013) Breast cancer-

derived microparticles display tissue selectivity in the transfer of resistance proteins

to cells. PLoS One. 8(4). e61515

317. Sarkadi, B., C. Ozvegy-Laczka, K. Nemet, and A. Varadi. (2004) ABCG2 -- a

transporter for all seasons. FEBS Lett. 567(1). 116-120

318. Huber, M.A., N. Kraut, and H. Beug. (2005) Molecular requirements for epithelial-

mesenchymal transition during tumor progression. Curr Opin Cell Biol. 17(5). 548-

558

319. Nistico, P., M.J. Bissell, and D.C. Radisky. (2012) Epithelial-mesenchymal

transition: general principles and pathological relevance with special emphasis on

the role of matrix metalloproteinases. Cold Spring Harb Perspect Biol. 4(2). pii:

a011908

320. De Craene, B. and G. Berx. (2013) Regulatory networks defining EMT during

cancer initiation and progression. Nat Rev Cancer. 13(2). 97-110

321. Pulukuri, S.M. and J.S. Rao. (2008) Matrix metalloproteinase-1 promotes prostate

tumor growth and metastasis. Int J Oncol. 32(4). 757-765

322. Guise, T.A. (2009) Breaking down bone: new insight into site-specific mechanisms

of breast cancer osteolysis mediated by metalloproteinases. Genes Dev. 23(18).

2117-2123

323. Guo, W. and F.G. Giancotti. (2004) Integrin signalling during tumour progression.

Nat Rev Mol Cell Biol. 5(10). 816-826

324. Pegtel, D.M., K. Cosmopoulos, D.A. Thorley-Lawson, M.A. van Eijndhoven, E.S.

Hopmans, J.L. Lindenberg, T.D. de Gruijl, T. Wurdinger, and J.M. Middeldorp.

(2010) Functional delivery of viral miRNAs via exosomes. Proc Natl Acad Sci U S

A. 107(14). 6328-6333

325. Xiao, D., J. Ohlendorf, Y. Chen, D.D. Taylor, S.N. Rai, S. Waigel, W. Zacharias, H.

Hao, and K.M. McMasters. (2012) Identifying mRNA, microRNA and protein

profiles of melanoma exosomes. PLoS One. 7(10). e46874

135

326. Escola, J.M., M.J. Kleijmeer, W. Stoorvogel, J.M. Griffith, O. Yoshie, and H.J.

Geuze. (1998) Selective enrichment of tetraspan proteins on the internal vesicles of

multivesicular endosomes and on exosomes secreted by human B-lymphocytes. J

Biol Chem. 273(32). 20121-20127

327. Chaput, N., C. Flament, S. Viaud, J. Taieb, S. Roux, A. Spatz, F. Andre, J.B.

LePecq, M. Boussac, J. Garin, S. Amigorena, C. Thery, and L. Zitvogel. (2006)

Dendritic cell derived-exosomes: biology and clinical implementations. J Leukoc

Biol. 80(3). 471-478

328. Skokos, D., S. Le Panse, I. Villa, J.C. Rousselle, R. Peronet, B. David, A. Namane,

and S. Mecheri. (2001) Mast cell-dependent B and T lymphocyte activation is

mediated by the secretion of immunologically active exosomes. J Immunol. 166(2).

868-876

329. Huber, V., S. Fais, M. Iero, L. Lugini, P. Canese, P. Squarcina, A. Zaccheddu, M.

Colone, G. Arancia, M. Gentile, E. Seregni, R. Valenti, G. Ballabio, F. Belli, E.

Leo, G. Parmiani, and L. Rivoltini. (2005) Human colorectal cancer cells induce T-

cell death through release of proapoptotic microvesicles: role in immune escape.

Gastroenterology. 128(7). 1796-1804

330. Lim, J.W., R.A. Mathias, E.A. Kapp, M.J. Layton, M.C. Faux, A.W. Burgess, H. Ji,

and R.J. Simpson. (2012) Restoration of full-length APC protein in SW480 colon

cancer cells induces exosome-mediated secretion of DKK-4. Electrophoresis.

33(12). 1873-1880

331. Malik, Z.A., K.S. Kott, A.J. Poe, T. Kuo, L. Chen, K.W. Ferrara, and A.A.

Knowlton. (2013) Cardiac myocyte exosomes: stability, HSP60, and proteomics.

Am J Physiol Heart Circ Physiol. 304(7). H954-965

332. Faure, J., G. Lachenal, M. Court, J. Hirrlinger, C. Chatellard-Causse, B. Blot, J.

Grange, G. Schoehn, Y. Goldberg, V. Boyer, F. Kirchhoff, G. Raposo, J. Garin, and

R. Sadoul. (2006) Exosomes are released by cultured cortical neurones. Mol Cell

Neurosci. 31(4). 642-648

136

333. Potolicchio, I., G.J. Carven, X. Xu, C. Stipp, R.J. Riese, L.J. Stern, and L.

Santambrogio. (2005) Proteomic analysis of microglia-derived exosomes: metabolic

role of the aminopeptidase CD13 in neuropeptide catabolism. J Immunol. 175(4).

2237-2243

334. Krämer-Albers, E.-M., N. Bretz, S. Tenzer, C. Winterstein, W. Möbius, H. Berger,

K.-A. Nave, H. Schild, and J. Trotter. (2007) Oligodendrocytes secrete exosomes

containing major myelin and stress-protective proteins: Trophic support for axons?

PROTEOMICS - CLINICAL APPLICATIONS. 1(11). 1446-1461

335. Kesimer, M., M. Scull, B. Brighton, G. Demaria, K. Burns, W. O'Neal, R.J. Pickles,

and J.K. Sheehan. (2009) Characterization of exosome-like vesicles released from

human tracheobronchial ciliated epithelium: a possible role in innate defense. Faseb

J. 23(6). 1858-1868

336. Conde-Vancells, J., E. Rodriguez-Suarez, N. Embade, D. Gil, R. Matthiesen, M.

Valle, F. Elortza, S.C. Lu, J.M. Mato, and J.M. Falcon-Perez. (2008)

Characterization and Comprehensive Proteome Profiling of Exosomes Secreted by

Hepatocytes. J Proteome Res. 7(12). 5157-5166

337. Chavez-Munoz, C., J. Morse, R. Kilani, and A. Ghahary. (2008) Primary human

keratinocytes externalize stratifin protein via exosomes. J Cell Biochem. 104(6).

2165-2173

338. Perez-Hernandez, D., C. Gutierrez-Vazquez, I. Jorge, S. Lopez-Martin, A. Ursa, F.

Sanchez-Madrid, J. Vazquez, and M. Yanez-Mo. (2013) The intracellular

interactome of tetraspanin-enriched microdomains reveals their function as sorting

machineries to exosomes. J Biol Chem. 288(17). 11649-11661

339. Fevrier, B., D. Vilette, F. Archer, D. Loew, W. Faigle, M. Vidal, H. Laude, and G.

Raposo. (2004) Cells release prions in association with exosomes. Proc Natl Acad

Sci U S A. 101(26). 9683-9688

340. Dukers, D.F., P. Meij, M.B. Vervoort, W. Vos, R.J. Scheper, C.J. Meijer, E.

Bloemena, and J.M. Middeldorp. (2000) Direct immunosuppressive effects of EBV-

encoded latent membrane protein 1. J Immunol. 165(2). 663-670

137

341. Admyre, C., J. Grunewald, J. Thyberg, S. Gripenback, G. Tornling, A. Eklund, A.

Scheynius, and S. Gabrielsson. (2003) Exosomes with major histocompatibility

complex class II and co-stimulatory molecules are present in human BAL fluid. Eur

Respir J. 22(4). 578-583

342. Andre, F., N.E. Schartz, M. Movassagh, C. Flament, P. Pautier, P. Morice, C.

Pomel, C. Lhomme, B. Escudier, T. Le Chevalier, T. Tursz, S. Amigorena, G.

Raposo, E. Angevin, and L. Zitvogel. (2002) Malignant effusions and immunogenic

tumour-derived exosomes. Lancet. 360(9329). 295-305

343. Ogawa, Y., M. Kanai-Azuma, Y. Akimoto, H. Kawakami, and R. Yanoshita. (2008)

Exosome-like vesicles with dipeptidyl peptidase IV in human saliva. Biol Pharm

Bull. 31(6). 1059-1062

344. Skriner, K., K. Adolph, P.R. Jungblut, and G.R. Burmester. (2006) Association of

citrullinated proteins with synovial exosomes. Arthritis Rheum. 54(12). 3809-3814

345. Keller, S., C. Rupp, A. Stoeck, S. Runz, M. Fogel, S. Lugert, H.D. Hager, M.S.

Abdel-Bakky, P. Gutwein, and P. Altevogt. (2007) CD24 is a marker of exosomes

secreted into urine and amniotic fluid. Kidney Int. 72(9). 1095-1102

346. Pisitkun, T., R.F. Shen, and M.A. Knepper. (2004) Identification and proteomic

profiling of exosomes in human urine. Proc Natl Acad Sci U S A. 101(36). 13368-

13373

347. Gonzales, P.A., T. Pisitkun, J.D. Hoffert, D. Tchapyjnikov, R.A. Star, R. Kleta,

N.S. Wang, and M.A. Knepper. (2009) Large-scale proteomics and

phosphoproteomics of urinary exosomes. J Am Soc Nephrol. 20(2). 363-379

348. Principe, S., E.E. Jones, Y. Kim, A. Sinha, J.O. Nyalwidhe, J. Brooks, O.J.

Semmes, D.A. Troyer, R.S. Lance, T. Kislinger, and R.R. Drake. (2013) In-depth

proteomic analyses of exosomes isolated from expressed prostatic secretions in

urine. Proteomics. 13(10-11). 1667-1671

349. Raimondo, F., L. Morosi, S. Corbetta, C. Chinello, P. Brambilla, P. Della Mina, A.

Villa, G. Albo, C. Battaglia, S. Bosari, F. Magni, and M. Pitto. (2013) Differential

protein profiling of renal cell carcinoma urinary exosomes. Mol Biosyst. 9(6). 1220-

1233

138

350. Wang, Z., S. Hill, J.M. Luther, D.L. Hachey, and K.L. Schey. (2012) Proteomic

analysis of urine exosomes by multidimensional protein identification technology

(MudPIT). Proteomics. 12(2). 329-338

351. Saleem, R.A., R.S. Rogers, A.V. Ratushny, D.J. Dilworth, P.T. Shannon, D.

Shteynberg, Y. Wan, R.L. Moritz, A.I. Nesvizhskii, R.A. Rachubinski, and J.D.

Aitchison. (2010) Integrated phosphoproteomics analysis of a signaling network

governing nutrient response and peroxisome induction. Mol Cell Proteomics. 9(9).

2076-2088

352. Deutsch, E.W., H. Lam, and R. Aebersold. (2008) PeptideAtlas: a resource for

target selection for emerging targeted proteomics workflows. EMBO Rep. 9(5). 429-

434

353. Desiere, F., E.W. Deutsch, N.L. King, A.I. Nesvizhskii, P. Mallick, J. Eng, S. Chen,

J. Eddes, S.N. Loevenich, and R. Aebersold. (2006) The PeptideAtlas project.

Nucleic Acids Res. 34(Database issue). D655-658

354. Desiere, F., E.W. Deutsch, A.I. Nesvizhskii, P. Mallick, N.L. King, J.K. Eng, A.

Aderem, R. Boyle, E. Brunner, S. Donohoe, N. Fausto, E. Hafen, L. Hood, M.G.

Katze, K.A. Kennedy, F. Kregenow, H. Lee, B. Lin, D. Martin, J.A. Ranish, D.J.

Rawlings, L.E. Samelson, Y. Shiio, J.D. Watts, B. Wollscheid, M.E. Wright, W.

Yan, L. Yang, E.C. Yi, H. Zhang, and R. Aebersold. (2005) Integration with the

human genome of peptide sequences obtained by high-throughput mass

spectrometry. Genome Biol. 6(1). R9

139

Appendix

140

Appendix A-1 Isolation and characterization of cell line-derived exosomes

Cell line

Isolation

strategiesa)

Validationb)

Proteomic

strategies Comments

c) Refs.

Haematopoietic cells

B cells (RN HLA-

DR15+)

DC, DG and

Immunobead

s (MHC-II)

WB (MHC-II)

1-DE, MALDI-

TOF MS and

nESI/MS/MS

21 proteins identified including MHC-I and

MHC-II, CD45, integrin α4, hsc70, hsp90,

Giα2, actin, tubulin, moesin, clathrin, GAPDH,

enolase and EF1α1

[139]

B cells (RN HLA-

DR15+)

DC, DG

IEM (MHC-II, CD53 and

CD82), WB (MHC-II, HLA-

DM, CD37, CD63, CD81,

CD82 and CD86)

-

Identified proteins CD37, CD53, CD63, CD81,

CD82 and CD86

[326]

Dendritic cells (D1) DC, F

EM, WB (MHC-II and

Lamp2), FACS (MHC- I and -

II, CD9, Mac-1 and CD86)

1-DE, MALDI-

TOF MS and

nESI/MS/MS

37 proteins identified including actin, tubulin,

cofilin, MFG-E8, annexins, rabs, CD9, hsp90β,

TSG101, syntenin, histones, Alix, 14-3-3

proteins, galectin-3, gag, reverse

transcriptase/pol and Mac-1 αβ

[50]

Dendritic cells (MD-

DC) DC, DG EM, WB

1-DE, MALDI-

TOF MS

35 proteins identified including Alix, annexins,

ICAM-1 and cofilin [327]

Dendritic cells (D1 and

BMDC) F, UC, DG

EM, WB (clathrin, hsc70,

annexin-II, CD9, flotillin 1,

ICAM-1, MHC I, MHC II,

TSG101, and MFG-E8),

FACS

1-DE, LC-

MS/MS

150 proteins identified including CD9, annexin

II, ICAM-1 and TSG101 [141]

Dendritic cells (D1 and

BMDC) DC, DG

EM, IEM (MHC-II), WB

(MHC-II)

1-DE, MALDI-

TOF MS

9 proteins identified including CD9, gag, Mac-

1, MFG-E8, hsc73 and annexin-II [223]

141

Mast cell (MC/9,

HMC-1, BMMC) DC, F, DG EM, FACS (CD63)

1-DE, LC-

MS/MS

271 proteins identified including mast

carboxypeptidase A, tubulins, TCP proteins,

ezrin, moesin, 40S ribosomal proteins, 14-3-3

proteins,CD43, CD63, CD97, annexins, MHC-I,

histones, hsc70 and integrin-α6

[138]

Mast cell (MC/9 and

BMMC) and

mastocytoma (P815)

DC EM, IEM

1-DE, MALDI-

TOF MS,

ELISA

7 proteins identified including MHC-II, CD40,

CD40L, CD86, LFA-1, ICAM-1, CD13,

annexin-VI, actins and CDC25

[328]

T cells (Jurkat cells, T

cell blasts, E+ cells and

MART-1+ T cell)

F, UC

EM, IEM (TCR β and CD3ε),

WB (TCR β and CD3ε),

FACS (CD63, TCR β, CD3ε,

MHC-I and -II)

WB, FACS

Proteins identified including TCR β, CD3ε and

ζ (phosphorylated ζ), MHC-I and -II, CD2,

CD18, chemokine receptor CXCR4, c-Cbl,

tyrosine kinases Fyn and Lck

[140]

Tumor cells

Bladder cancer cells

(HT1376) DC, SC

WB (TSG101, CD9 and

CD81)

LC-MALDI-

TOF/TOF MS

353 proteins identified including basigin,

galectin-3, trophoblast glycoprotein (5T4),

LAMP2, CD73 HLA-G and β-Catenin

[145]

Brain tumor

(EGFRvIII-transfected

SMA560)

DC, DG

(OptiPrep™)

EM, WB (Alix, GAPDH, α1-

anti-trypsin, CD9, PDI, CRT,

transferrin, GPNMB, TGF-β1

and EGFRvIII),

acetylcholinesterase assay

2-DE, MALDI-

TOF/TOF

36 proteins identified including HSP70, α-

enolase, MVP, ARRDC2, CENP-P, GAPDH,

syntenin, PCNA, CRMP-2 and sialic acid

synthase

[125]

Breast adenocarcinoma

(BT-474 and MDA-

MB-231)

DC, F, DG,

Immunobead

s (HER2)

EM, FACS (HER2), WB

(HER2 and actin) WB, FACS Protein HER2 identified [37]

Colorectal cancer

(HT29)

DC,

diafiltration

(100K), DG

EM, WB (CD63 and CD81) 1-DE, LC-

MS/MS

547 proteins identified including annexins,

ARFs, rabs, ADAM10, CD44, NG2, ephrin-B1,

MIF, β-catenin, junction plakoglobin, galectin-

4, RACK1 and tetraspanin-8

[48]

142

Colon carcinoma cell

lines (SW403, 1869col

and CRC28462)

DC

EM, IEM (CD63, FasL and

TRAIL), WB and FACS

(CD63, FasL, TRAIL, CEA

and MHC-I)

WB FasL and TRAIL identified [329]

Colorectal cancer

(LIM1215)

F,

diafiltration

(5K), UC

EM, IEM (A33), WB (CD9,

A33, TSG101, Rab7 and

hsc70)

1-DE, LC-

MS/MS

400 proteins identified including A33, CEA,

EGFR, ADAM10, dipeptidase 1, ephrin-B1,

hsc70, tetraspanins, ESCRT proteins, integrins,

annexins, rabs and GTPases

[44]

Colorectal carcinoma

cell lines SW480 and

SW620

DG

(OptiPrep™)

Cryo-EM, WB (Alix,

TSG101, CD9, HSP70,

Flotillin-1, and EGFR

1-DE, LC-

MS/MS

941 and 796 proteins identified in SW480 and

SW620 exosomes, respectively, including MET,

S100A8, S100A9, TNC, EFNB2, EGFR, JAG1,

SRC and TNIK

[114]

Colorectal carcinoma

cell lines SW480 and

SW480APC

F, DC EM, WB (Alix, TSG101, CD9

and HSP70)

1-DE, LC-

MS/MS

563 and 497 proteins were identified in SW480

and SW480APC exosomes respectively

including DKK4, Alix, TSG101, VPS28,

VPS25, CHMP2A and tetraspanins CD9, CD63,

CD81 and CD82

[330]

Mammary

adenocarcinoma

(TS/A, H-2d), P815

mastocytoma (H-2d),

melanoma (Fon and

Mel-888)

DC, DG EM, WB (hsc70 and MHC-I) WB, IEM Proteins MHC-I, hsp70, MART-1 and TRP

identified [73]

Melanoma (MeWo and

SK-MEL-28) F, UC

WB (MHC-I, MART-1, Mel-

CAM and annexin II), DG

2-DE, MALDI-

TOF/TOF MS

41 proteins identified including Alix, hsp70,

Giβ2, Giα, moesin, GAPDH, malate

dehydrogenase, p120 catenin, PGRL, syntaxin-

binding protein 1 and 2, septin-2 and WD

repeat-containing protein 1

[142]

143

Mesothelioma (PMR-

MM7 and 8) DC IEM (CD63)

1-DE, MALDI-

TOF MS

30 proteins identified including annexins,

actins, actinin-4 tubulins, hsc70, hsp90,

integrins, fibronectin, GAPDH, MHC-I,

PLVAP and DEL-1

[143]

Medulloblastoma cells

(D283MED) F, DG

NTA, WB (CD9 and HSP70)

acetylcholinesterase assay

1-DE, LC-

MS/MS

148 proteins proteins identified including PDI,

HPX, HNF4A and GPNMB [144]

Primary and normal cells

Cardiac myocytes

(Sprague-Dawly rats) DC, EQ EM

1-DE LC-

MS/MS

57 proteins identified including tropomyosin-

1α, myomesin 2 (M-band), α-crystallin B,

cardiac α-actin, GAPDH and long-chain

specific acyl-CoA dehydrogenase.

[331]

Cortical neurons (8-

day primary culture) DC, DG

EM, WB (Alix, TSG101 and

flotillin)

1-DE, LC-

MS/MS

19 proteins identified including GLAST1,

brain-specific ceruloplasmin, L1 cell adhesion

molecule, GPI-anchored prion protein and

GluR2/3

[332]

Intestinal epithelial

(HT29-19A and T84-

DRB1*0401/CIITA)

DC, DG

EM, IEM (CD26, CD63,

MHC-I and -IIα), WB

(MHC-I and -IIα, CD26,

CD63, TfR and Na+K

+-

ATPase)

1-DE, MALDI-

TOF MS

28 proteins identified including syntaxin-3,

syntaxin-binding protein 2, EPS8, microsomal

dipeptidase in AM and A33, epithelial cell

surface antigen and major vault protein

[72]

Microglia (N9 and

primary culture from

SJL/J mice)

DC, DG

EM, WB (CD9, CD63,

syntaxin-8, Rab7, Rab11,

clathrin, Lamp-1 and -2, Vti-

1A and -1B)

1-DE, LC-

MS/MS

59 proteins identified including cathepsin S,

aminopeptidase N (CD13), MCT-1, MHC class

II-associated chaperone Ii, CD14, NAP-22, FcR

for IgE and GP42

[333]

Oligodendrocytes

(primary culture and

Oli-neu)

DC, DG EM, WB (Alix, PLP, CNP

and TSG101) LC-MS/MS

143 proteins identified including CD81, 14-3-3

proteins, actins, tubulins, histones, EF1 and 2,

hsp90, hsc70, Na+K

+ATPase α chains, PLP,

CNP, MBP and MOG

[334]

144

Human

tracheobronchial

epithelial cell

DC, F, DG

EM, WB (Muc1, EBP50,

CD133, Annexin II, TSG101

and CD63)

1-DE, LC-

MS/MS

40 proteins identified including CD63, TSG101,

mucins, actins and tubulins [335]

Platelets DC, DG EM, IEM (CD63), WB

(CD63) WB, FACS CD63 identified [24]

Hepatocytes (MLP-29

and livers from 9-week

old Sprague-Dawley

rats)

DC, F, DG

EM, WB (TSG101, Alix,

integrin-β1, CD63, CD81,

ICAM-1 and lactadherin)

1-DE, LC-

MS/MS

251 proteins identified including tetraspanins,

ASGR, cytochromes P450, cytoskeletal

proteins, apolipoprotein-E and -AV,

paraoxonase-1 and -3, regucalcin, UDP-

glucuronosyltransferases

[336]

Keratinocytes

(2-day primary culture

from foreskin)

DC,

diafiltration

(100K), DG

EM, WB (Hsc70 and

LAMP2) WB 14-3-3σ (stratifin) identified [337]

Lymphoblasts F, DC EM, NTA LC-MS/MS

489 proteins identified including MHC

isoforms, CD4, CD8, ADAM10, CD44, CD98,

integrins, ERM proteins and syntenin

[338]

Virus-infected cells

Rov epithelial and

Mov neuroglial cells

infected with PrP

DC, DG

IEM (PrP, flotillin, TSG101

and TfR), WB (PrP, flotillin,

TfR, TSG101 and hsc70)

1-DE, LC-

MS/MS

93 proteins identified including 14-3-3 proteins,

annexins, hsc70, integrins, rabs, actin, tubulins,

MFG-E8, Gi2α, histones, PrP and PrPsc

[339]

EBV-positive

lymphoblastoid cells

(JY and B95-8)

DC WB (HLA-DR and CD86) WB LMP1 [340]

145

a) DC: Differential centrifugation; DG: Sucrose density gradient or otherwise stated; EQ: Exoquick; F: Filtration; GF: Gel filtration; SC: Sucrose

cushion; UC: Ultracentrifugation

b) CRT: Calreticulin; EM: Electron microscopy; EGFRvIII: Epidermal growth factor receptor variant III; FACS: Fluorescence–activated cell sorting;

GPNMB: Glycoprotein nonmetastatic B; IEM: Immunoelectron microscopy; NTA: Nanoparticle tracking analysis; PDI: Protein disulfide isomerase;

TGF-β1: Transforming growth factor beta 1; TfR: Transferrin receptor; WB: Western blot.

c) ADAM10: A Disintegrin and metalloproteinase domain-containing protein 10; AM: Apical membrane; ARF: ADP-ribosylation factor; ARRDC2:

Arrestin domain-containing protein 2; ASGR: Asiaglycoprotein receptor; BM: Basolateral membrane; CENP-P: Centromere protein P; CHMP2A:

Charged multivesicular body protein 2A; CNP: 2',3'-cyclic-nucleotide 3'-phosphodiesterase; CRMP-2: Collapsing response mediator protein 2; DEL-

1: developmental endothelial locus-1; DKK4: Dickkopf 4; Di EBV: Epstein-Barr virus; EFNB2: Ephrin-B2; EGFR: Epidermal growth factor

receptor; ERM proteins (ezrin, radixin, and moesin); ESCRT: Endosomal sorting complex required for transport; FasL: Fas Ligand; GAPDH:

glyceraldehyde-3-phosphate dehydrogenase; GPNMB: glycoprotein non-metastatic B; HNF4A: hepatocyte nuclear factor alpha; HPX: Hemopexin;

JAG1: Jagged 1; LAMP1: Lysosome associated membrane protein 1; LMP1: Latent membrane protein 1; MART-1:Melanoma antigen recognized by

T cells 1; MBP: Myelin basic protein; MET: Met proto-oncogene (hepatocyte growth factor receptor); MIF: Macrophage migration inhibitory factor;

MOG: Myelin oligodendrocyte glycoprotein; MHC: Major histocompatibility complex; MVP: Major vault protein; NG2: Chondroitin sulfate

proteoglycan; PDI: Protein disulfide isomerase; PLP: Myelin proteolipid protein; PrP: Prion protein; PrPsc: Prion protein scrapie; RACK1: Lung

cancer oncogene 7; S100A8: S100 calcium binding protein A8; S100A9: S100 calcium binding protein A9; SRC: v-src sarcoma; TNC: tenascin C;

TRAIL: TNF-related apoptosis-inducing ligand; TNIK; TRAF2 and NCK interacting kinase; TRP: 5,6-dihydroxyindole-2-carboxylic acid oxidase;

VPS proteins: Class E Vacuolar protein-sorting proteins. Reproduced and updated from [74].

146

Appendix A-2 Isolation and characterization of body fluid-derived exosomes

Body fluid Isolation

strategiesa)

Validationb)

Proteomic

strategies Comments

c) Refs.

Breast milk DC, F, DG

IEM (CD63 and HLA-DR), WB

(HLA-DR, CD81 and HSC70),

FACS (HLA-DR, CD63, CD81 and

mucin-1)

In-solution

trypsinization,

scx-LC-MS/MS

73 proteins identified including MFG-

E8, MUC1, hsp70, ARF-1, EH

domain-containing protein 1, CD36,

butyrophilin and polymeric-Ig

receptor

[146]

Bronchoalveolar

lavage fluid

DC, immuno-

beads (MHC-II)

IEM (HLA-DR and CD63), FACS

(HLA-DR, CD54, CD63 and CD86)

-

-

[341]

Malignant pleural

effusions DC, DG EM

1-DE, MALDI-

TOF MS

50 proteins identified including MHC-I,

actin, G protein, hsp90, BTG1, Bamacan,

PEDF, BTG-1, TSG14 and TSP2

[147]

Malignant pleural

effusions and

malignant ascites

DC, DG

EM, IEM (MHC-I and -II, TRP,

gp100 and CD81), WB (MHC-I and

-II ,MART-1, HER2 and hsc70)

WB Proteins including MART-1, TRP, gp100

and HER2 identified [342]

Saliva GF EM WB (Alix, TSG101, HSP70 and

CD63)

2-DE LC-

MS/MS

68 proteins were identified including

GW182, immunoglobulin A, polymeric

immunoglobulin receptor, Mucin-5B, Alpha

–amylase 1, Dipeptidyl peptidase 4 (CD26).

[47]

Saliva F, diafiltration

(100K), GF EM

N-terminal

amino acid

sequencing, WB

Dipeptidyl peptidase IV (CD26), polymeric-

Ig receptor, actin, galectin-3 and Ig chains [343]

Semen DG EM 1-DE, LC-

MS/MS 440 proteins identified [148]

Synovial fluid

(RA, OA and

reactive arthritis)

DC, DG EM, IEM (IgG)

2-DE, WB,

MALDI-TOF

MS

Citrullinated fibrin α-chain, CD5 antigen-

like, fibrinogen fragment D and β-chain [344]

147

Urine and

amniotic fluid DC, DG

EM, WB (hsp70, AQP2, annexin-1

and CD9) WB CD24 [345]

Urine UC

EM, IEM (APN, AQP2, CD9 and

NCC), WB (TSG101, Alix, CD9,

Rab-4, -5B and -11, SNX18)

1-DE, LC-

MS/MS

295 proteins identified including VPS

proteins, AQP2, polycystin-1, carbonic

anhydrase II and IV

[346]

Urine DC Not performed 1-DE, LC-

MS/MS

1132 proteins identified including AQP2,

vacuolar H+-ATPase subunits and ESCRTs.

14 phosphorylated proteins including NCC,

GPRC5B and GPRC5C

[347]

Urine

(Prostatic

secretions)

DC Not Performed 1-DE LC-

MS/MS

877 proteins identified including markers

CD9, CD63, CD81, Alix, TSG101, FLOT1,

FLOT2, ANXA2 and ANXA5. Prostate-

enriched proteins PSA, ACPP, TGM4 and

PSMA

[348]

Urine

(Renal cell

carcinoma, 9

patients)

DC, DG

(OptiPrep™)

EM, WB (Flotillin-1, TSG101 and

CD9)

1-DE LC-

MS/MS

186 proteins identified including matrix

metalloproteinase 9, ceruloplasmin,

podocalyxin , dickkopf related protein 4 and

carbonic anhydrase IX

[349]

Urine DC WB (Alix, TSG101 and AQP1)

In-solution

trypsinization

LC-MS/MS

280 proteins identified including major

urinary protein 1, CD10 [149]

Urine DC EM

1-DE LC-

MS/MS and In-

solution

trypsinization

LC-MS/MS

1025 proteins identified including 12 annexin

proteins, 32 RAB proteins, 5 RAP proteins,

aquaporin-1, aquaporin-2, aquaporin-7,

solute carrier family 12 member 1, Kidney

specific Na-K-Cl symporter and solute

carrier family 12 member 3

[350]

148

a) DC: Differential centrifugations; DG: Sucrose density gradient or otherwise stated; F: Filtration; GF: Gel filtration; UC: Ultracentrifugation.

b) APN: Aminopeptidase N; EM: Electron microscopy; FACS: Fluorescence–activated cell sorter; IEM: Immunoelectron microscopy; SNX18:

Sorting nexin-18; WB: Western blot.

c) ACPP: Prostatic Acid Phosphatase; ANXA2: Annexin A2; ANXA5: Annexin A5; AQP1: Aquaporin-1; AQP2: Aquaporin-2; ARF: ADP-

ribosylation factor; Bamacan: Basement membrane chrondroitin sulfate proteoglycan; BTG1: B-cell translocation gene 1; FLOT: Flotillin GP100:

Melanocyte protein Pmel 17; GPRC5B and GPRC5C: G-protein coupled receptor family C group 5 member B and C; MART-1: Melanoma antigen

recognized by T cells 1; NCC: Na-Cl Cotransporter; OA: osteoarthritis; PEDF: Pigment epithelium- derived factor; PSA: Prostate specific antigen;

PSMA: Prostate specific membrane antigen; RA: Rheumatoid arthritis; TGM4; Transglutaminase 4; TRP: 5,6-dihydroxyindole-2-carboxylic acid

oxidase; TSG14: Tumor necrosis factor-stimulated gene 14; TSP2: Thrombospondin-2; VPS proteins: Class E Vacuolar protein-sorting proteins.

Reproduced and updated from [74].

149

Appendix A-3 Human colon cancer LIM1863 cells

(A) Phase contrast microscopy of LIM1863 organoids Scale 20 μm. (B) A confocal

microscopy cross-section through a LIM1863 organoid highlights concentrated syntaxin 3

staining of the apical ring (red), and A33 staining at the basolateral cell periphery (green).

Scale 20 μm. (C) LIM1863 cell metabolic activity was measured using the MTT assay

following 24 h culture in RPMI containing either 5% FCS or 0.6% ITS. (D) Cell viability

of LIM1863 cells was measured using the lactate dehydrogenase (LDH) assay following 24

h culture in RPMI containing either 5% FCS or 0.6% ITS. Data represents the mean ± S.D

of three independent experiments performed in triplicate.

150

Appendix A-4 Distinguishing features of EMT

References [270, 286, 289].

Feature

observed Epithelial cells Mesenchymal cells

Morphology

• Round cobblestone growth in culture

• Strong cell-cell contact

• Elongated, spindle shaped, scattered

growth in culture

• Weak cell-cell contact

Polarity

• Well maintained apical basolateral

polarity

• Intact tight junctions (ZO-1), adherens

junctions (E-cadherin) and desmosomes

•Leading edge front-back polarity

• No tight junction, adherens junction or

desmosomes. Diminished E-cadherin and

ZO-1 expression

Motility • Little cell migration due to intact

epithelial sheets

• Increased cell migration of individual

cells

Invasiveness • Minimal, spherical cyst growth in

collagen gel

• Increased, penetrating growth into

collagen gel

Cytoskeleton

• Apical actin ring and actin stress fibers

• Cytokeratin intermediate filament

network inside cell border

• No vimentin intermediate filaments

• No apical actin ring and dissolved actin

stress fibers

• Loss of cytokeratin intermediate filaments

• Vimentin intermediate filaments across

cell body

151

Appendix A-5 Secreted extracellular modulators of EMT

Secretome analysis of MDCK cells reduced expression of cell-cell and cell-matrix

attachment proteins following oncogenic Ras-transformed EMT, possibly to enable greater

mobility. Basement membrane constituents were downregulated, and increased protease

expression was observed following transformation. Production of factors associated with

migration and invasion were also elevated in the 21D1 cell secretome. Adapted from [250].

152

Appendix A-6 Supplemental information for manuscripts included in this thesis

Chapter 2. Tauro, B. J., Greening, D. W., Mathias, R. A., Ji, H., Mathivanan, S., Scott, A.

M., and Simpson, R. J. (2012) Comparison of ultracentrifugation, density gradient

separation, and immunoaffinity capture methods for isolating human colon cancer cell line

LIM1863-derived exosomes. Methods 56, 293-304. PMID: 22285593.

Supplemental data associated with this article can be found at

doi:10.1016/j.ymeth.2012.01.002.

Supplemental Table 1 - Proteins identified in UC-Exos, DG-Exos, and IAC-Exos.

Supplemental Table 2 - Relative quantification of cancer related proteins using label-free-

based spectral counting.

Supplemental Figure 1 - Human colon cancer LIM1863 cells (Appendix A-3).

Supplemental Figure 2 - Distribution of proteins identified in UC-, DG-, and IAC-Exos

(Appendix A-7).

Supplemental Figure 3 - Plasma membrane proteins identified in the exosome datasets

(Appendix A-8).

Chapter 3 Tauro, B. J., Greening, D. W., Mathias, R. A., Mathivanan, S., Ji, H., and

Simpson, R. J. (2012) Two distinct populations of exosomes are released from LIM1863

colon carcinoma cell-derived organoids. Molecular & Cellular Proteomics 12, 587-598

PMID: 23230278.

Supplemental data associated with this article can be found at

http://www.mcponline.org/content/suppl/2013/07/11/M112.021303.DC1.

Supplemental Table 1 - A33- and EpCAM-Exos Protein Datasets.

Supplemental Table 2 - sMVs Protein Dataset.

153

Supplemental Table 3 - A33-Exos Peptide Identifications.

Supplemental Table 4 - EpCAM-Exos Peptide Identifications.

Supplemental Table 5 - sMVs Peptide Identifications.

Chapter 4 Tauro, B. J., Mathias, R. A., Greening, D. W., Gopal, S. K., Ji, H, Kapp, E. A.,

Coleman, B. M., Hill, A. F., Kusebauch, U., Hallows, J. L., Shteynberg, D., Moritz, R. L.,

Zhu, H. J., and Simpson R. J. (2013) Oncogenic H-Ras reprograms Madin-Darby canine

kidney (MDCK) cell-derived exosomal proteins during epithelial-mesenchymal transition.

Molecular & Cellular Proteomics 12, 2148-2159. PMID: 23645497.

Supplemental data associated with this article can be found at

http://www.mcponline.org/content/early/2013/05/03/mcp.M112.027086/suppl/DC1.

Supplemental Table 1- Proteins identified in exosome proteome profiling of H-Ras-induced

EMT (21D1) in MDCK cells.

Supplemental Table 2 - Peptides identified in MDCK-Exos.

Supplemental Table 3 - Peptides identified in 21D1-Exos.

Supplemental Table 4 - 139 proteins identified in MDCK- and 21D1-Exos not reported in

ExoCarta database.

Supplemental Figure 1 - 21D1 cells require their own culture medium to maintain a

mesenchymal phenotype (Appendix A-9).

Supplemental Figure 2 - Exosome characterization following OptiPrep™ density gradient

separation (Appendix A-10).

Supplemental Data - Experimental procedures for replicate GeLC-MS/MS analysis of

MDCK- and 21D1-Exos (Appendix A-11).

154

Appendix A-7 Distribution of proteins identified in UC-, DG-, and IAC-Exos

A three-way Venn diagram depicting number of proteins commonly observed in all three

datasets (265), and those that are unique to each purification strategy; ultracentrifugation

(UC-Exos), density gradient centrifugation (DG-Exos), and immunoaffinity capture (IAC-

Exos).

155

Appendix A-8 Plasma membrane proteins identified in the exosome datasets

Proteins were annotated as plasma membrane based on HPRD classification (GO:

0005886). Proteins containing one or more transmembrane domains were predicted using

TMHMM (v2.0). IAC-Exos contained the highest proportion of proteins annotated as

plasma membrane (32.1%), and predicted to contain at least one TM spanning domain

(19.8%).

156

Appendix A-9 21D1 cells require their own culture medium to maintain a

mesenchymal phenotype

21D1 cells were grown as described in Chapter 4 Experimental Procedures with growth

media changed daily and supplemented with either 10% (A) or without (B) their own

culture medium. After a period of 5 days, phase contrast images of 21D1 cells

supplemented with 10% culture medium display an elongated mesenchymal-like spindle

shape (A), while 21D1 cells not supplemented with 10% culture medium revert back to an

epithelial-like cobblestone morphology (B).

157

Appendix A-10 Exosome characterization following OptiPrep™ density gradient

separation

Concentrated culture medium from MDCK and 21D1 cells was isolated as described in

Chapter 4 Experimental Procedures (Cell culture and CCM preparation) and separated

using OptiPrep™ density gradient. Protein fractions (1.01-1.30 g/mL) from MDCK (A) and

21D1 (B) were stained with SYPRO® Ruby protein stain. Western blot analysis of MDCK

(C) and 21D1 (D) OptiPrep™ density gradient fractions (10 µg) revealed the exosomal

marker Alix predominantly detected in fraction 1.09 g/mL. This fraction was selected for

proteomic analysis.

158

Appendix A-11 Experimental procedures for replicate GeLC-MS/MS analysis of

MDCK- and 21D1-Exos

MDCK- and 21D1-Exos samples (20 μg) were lysed in SDS sample buffer, and proteins

separated by SDS-PAGE and visualized by Imperial™ Protein Stain (Thermo Fisher

Scientific), according to manufacturer’s instructions. Gel lanes were cut into 20 × 2 mm

bands using a GridCutter (The Gel Company, San Francisco, CA) and individual bands

subjected to automated in-gel reduction, alkylation and trypsinization using a Tecan EVO

liquid handler robotic system. Briefly, gel bands were reduced with 10 mM DTT

(Calbiochem, San Diego, USA) for 20 min, washed and alkylated for 30 min with 50 mM

iodoacetic acid (Fluka, St. Louis, USA), and digested with 100 ng trypsin (Promega

Sequencing Grade Product V5111X, Wisconsin, USA) for 8 h at 35 °C with automated

agitation. Tryptic peptides were extracted with 3 50 μL 50% (v/v) acetonitrile, 50 mM

ammonium bicarbonate, concentrated to ~10 μL by centrifugal lyophilisation, pooled (2

bands pooled as 1 fraction) and analyzed by LC-MS/MS. RP-HPLC was performed on a

Dr. Maisch 3 μm ReproSil-Pur C18 AQ (Dr Maisch GmbH, Germany) 150 × 0.075-mm-

internal diameter reversed phase column with an integrated KASIL Frit using an Agilent

1100 nanoHPLC coupled online to an LTQ-Orbitrap mass spectrometer equipped with a

nanoelectrospray ion source (Thermo Fisher Scientific, San Jose, USA). The column was

developed with a linear 60 min gradient with a flow rate of 0.3 μL/min from 0-60% solvent

B, 60-80% solvent B in 2 min and held for 10 min and then returned to 2% solvent B for

15 min equilibration where solvent A was 0.1% (v/v) aqueous formic acid and solvent B

was 0.1% (v/v) aqueous formic acid/100% acetonitrile. Survey MS scans were acquired

with the resolution set to a value of 30,000. Real-time recalibration was performed using a

background ion from ambient air in the C-trap. Up to 10 selected target ions were

fragmented and then were dynamically excluded from further analysis for 120 secs from

sample analysis [351]. Raw mass spectrometry data is deposited in the PeptideAtlas and

can be accessed at http://www.peptideatlas.org/PASS/PASS00225 [352-354].

159

Minerva Access is the Institutional Repository of The University of Melbourne

Author/s:

Tauro, Bow Jesse

Title:

The isolation and characterization of exosomes

Date:

2013

Citation:

Tauro, B. J. (2013). The isolation and characterization of exosomes. PhD thesis, Ludwig

Institute for Cancer Research Ltd. (Parkville Branch), La Trobe Institute for Molecular

Science, La Trobe University & Department of Biochemistry and Molecular Biology, The

University of Melbourne.

Persistent Link:

http://hdl.handle.net/11343/38509

File Description:

Thesis text