epigenetic reprogramming and dna demethylation · however reprogramming of imprinting is only...

158
Epigenetic Reprogramming and DNA Demethylation HAKAN BAGCI Imperial College London Faculty of Medicine MRC Clinical Sciences Centre Doctor of Philosophy 1

Upload: others

Post on 21-Aug-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Epigenetic Reprogramming and DNA Demethylation

HAKAN BAGCI

Imperial College London

Faculty of Medicine

MRC Clinical Sciences Centre

Doctor of Philosophy

1

Page 2: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

I, Hakan Bagci, hereby declare that this thesis is my own work and that work performed by others has been appropriately acknowledged and referenced.

2

Page 3: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

The copyright of this thesis rests with the author and is made available under a Creative Commons Attribution Non-Commercial No Derivatives licence. Researchers are free to copy, distribute or transmit the thesis on the condition that they attribute it, that they do not use it for commercial purposes and that they do not alter, transform or build upon it. For any reuse or redistribution, researchers must make clear to others the licence terms

of this work.

3

Page 4: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Abstract Embryonic development begins with fertilization of the egg, a progressive process

that gives rise to the zygote and subsequently to the formation of somatic tissues.

Normally once cells acquire a fate, it is stably maintained. Conversion back to a

multipotent state occurs rarely in-vivo, but can be achieved experimentally by inducing

‘reprogramming’. In this study I am looking at the epigenetic mechanisms that underlie

reprogramming and, in particular, DNA methylation and demethylation. To address this I

am taking advantage of the cellular fusion system. Fusion of pluripotent cells with

differentiated cells results in the formation of transient heterokaryon and hybrid cells,

where the somatic partner is efficiently reprogrammed. This gives me the opportunity to

monitor early and late events in pluripotent conversion, in which global remodelling of

chromatin and changes in DNA methylation occur.

Here, I examine changes in DNA methylation that are induced at imprinted loci and

pluripotency-associated genes when somatic cells are fused with either mouse embryonic

stem (ES) or embryonic germ (EG) cells. I focus on defining the factors and order of events

that accompany reprogramming. I show that acquisition of pluripotency is an early process

occurring at the heterokaryon stage, and is followed by imprint erasure later in hybrids.

However reprogramming of imprinting is only induced by EG, but not ES cells, and it

requires sequential steps of 5-methylcytosine oxidation mediated by Tet proteins and

nucleotide exchange upon several rounds of DNA synthesis. I provide evidence that Tet

proteins are dispensable for pluripotent reprogramming using CRISPR-Cas9 genome

editing to abrogate the expression of both Tet1 and Tet2. This result suggests that either

DNA demethylation can occur without TET activity (implying a redundancy with other

demethylating agents and routes), or that DNA demethylation is not required for inducing

pluripotency. Finally, I describe how CRISPR/Cas9 approaches were used to demonstrate

that non-canonical Wnt signalling components are downstream targets of JARID2 in ES

cells.

4

Page 5: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Acknowledgments

I would first like to express my gratitude to Mandy and Matthias for giving me the opportunity to be a part of the scientifically stimulating, socially interacting, internationally competent Lymphocyte Development Team at the CSC. I am especially grateful to Mandy, for teaching me how to be open-minded, patient and collaborative in science, how to ask the right questions and how to approach them. Thank you very much for encouraging curiosity and innovation, always being tolerant, thoughtful and supportive with me. I would also like to thank the Medical Research Council, for financially supporting me and my PhD work.

I would like to thank all the past and present members of the LympDeve, for collectively creating a joyful and friendly scientific environment, for help, advice, support and long discussions, and for making my four years unforgettable. I would like to say “air hair lair” to Allifia, thank you for not only being a valuable friend, but also for trusting me and for providing me shelter in the last months of my thesis. I cannot put into words how great it feels to know you will be there whenever I am in trouble. Thank you Amélie for our little conversations, for your help with experiments, and for bringing order to our laboratory; without you, we would all be lost. Andy, my desk buddy, who by now should have become a YoYo master, thank you for your friendship, for marvellous memories, your jokes and puns that will always make me giggle. Thank you for bringing colour and laughter to the PhD office (also thank you to your iPod!). Feng, thank you for your humour, I wish we made a list, but I believe it would be a bit inappropriate to be written in here. Thank you Irene for long discussions, and for never being tired of sharing your extensive knowledge with me (and for delicious tiramisus). Thank you Jorge for your scientific help, for sharing your ideas, for always being kind, supportive and positive. Lee the east Londoner, thank you for great house parties, for teaching me the contemporary usage of the English language and exposing me to the alternative London lifestyle. Lesly, thank you for setting limits to our bad jokes. Preksha, I have always enjoyed discussing with you, on any subject, and thank you for always being there since the first day of our PhD. Thais, thank you for your never-ending kindness, and for our dialogues while cycling to the west. Ziwei, it has always been a pleasure talking to you, and thank you for your help with experiments. I am thankful to the present members Anne-Céline, Grainne, Isabel, Kotryna, Liz, Ludovica, Matt, Sergi, Tom, Vlad, and to the past members Antoine, Bryony, Cynthia, Hegias, Luke, Rory, for making my time more enjoyable in the laboratory, you will never be forgotten. I would specially like to thank Francesco, David and Karen, for instructive discussions and for your help and contribution to this thesis. I would finally like to thank James in the FACS facility, Zoe in the Transgenics facility, Microscopy facility and Sequencing facility for their help in conducting my PhD work.

Special thanks to my great friends Matteo, Joana, Silvia and Joao. Thank you for delicious dinners, for pub nights, for party nights, for laughters, for concerts, for travels, for festivals starting from the first days of our PhD… Our bond will never be broken.

Thank you Mélanie for your love, for your patience with me, for everything we shared and for all of our cheerful memories. Je t’envoie pleins de bisous!

And I would finally like to thank my parents. Annem ve babama: her ne kosulda olursa olsun her zaman arkamda oldugunuz icin, beni her kararimda desteklediginiz ve buralara kadar gelmemde en buyuk rolu oynadiginiz icin cok tesekkur ederim.

5

Page 6: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Contents

Abstract........................................................................................................................... . 4

Acknowledgments.......................................................................................................... 5

Contents.......................................................................................................................... . 6

List of Figures & Tables ................................................................................................ 9

Abbreviations................................................................................................................. 11

Chapter 1. Introduction ............................................................................................... 13

1.1. Pluripotent embryonic stem cells .............................................................................. 13 1.1.1. Chromatin dynamics in embryonic stem cells ................................................................................... 15 1.1.2. Polycomb regulation and bivalency ................................................................................................... 16

1.2. DNA methylation dynamics in mammals ................................................................... 17 1.2.1. Maintenance and de novo establishment of DNA methylation ......................................................... 17 1.2.2. Roles of DNA methylation in gene regulation .................................................................................... 19 1.2.3. Genomic Imprinting ........................................................................................................................... 20 1.2.4. DNA demethylation ........................................................................................................................... 23

1.2.4.1. Passive demethylation ................................................................................................................ 23 1.2.4.2. Active demethylation ................................................................................................................. 25 1.2.4.3. TET protein mediated 5-mC oxidation in passive and active demethylation ............................. 27 1.2.4.4. TET-associated DNA demethylation dynamics in embryonic development and pluripotency .. 33

1.3. Reprogramming cell fate ........................................................................................... 38 1.3.1. Transdifferentiation ........................................................................................................................... 39 1.3.2. Pluripotent conversion of somatic cells ............................................................................................. 40

1.3.2.1. Nuclear transfer .......................................................................................................................... 40 1.3.2.2. Induced pluripotent stem cells ................................................................................................... 42 1.3.2.3. Cell fusion ................................................................................................................................... 43

1.4. Aims of this study ..................................................................................................... 47

Chapter 2. Materials and Methods .............................................................................. 48

2.1. Materials ................................................................................................................... 48 2.1.1. Cell lines ............................................................................................................................................. 48 2.1.2. Antibodies .......................................................................................................................................... 49

2.2. Methods.................................................................................................................... 49 2.2.1. Cell culture ......................................................................................................................................... 49 2.2.2. Cell fusion experiments ..................................................................................................................... 50 2.2.3. Fluorescence activated cell sorting (FACS) ........................................................................................ 51 2.2.4. Quantitative Reverse Transcription Polymerase Chain Reaction (qRT-PCR) Analysis ....................... 51

2.2.4.1. RNA extraction and reverse transcription .................................................................................. 51 2.2.4.2. Quantitative PCR ........................................................................................................................ 52

2.2.5. DNA methylation and hydroxymethylation analyses......................................................................... 52 2.2.5.1. Bisulfite sequencing analysis ...................................................................................................... 53 2.2.5.2. 5-hmC quantification by enzyme protection assay .................................................................... 53

6

Page 7: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

2.2.6. Imaging analysis ................................................................................................................................. 54 2.2.6.1. Immunofluorescence and confocal microscopy imaging ........................................................... 54 2.2.6.2. X-gal staining .............................................................................................................................. 54

2.2.7. Western Blot analysis ......................................................................................................................... 54 2.2.8. Chromatin Immunoprecipitation (ChIP) analysis ............................................................................... 55 2.2.9. Plasmid construction and delivery into ES cells ................................................................................. 56 2.2.10. CRISPR/Cas9 genome editing system .............................................................................................. 57

2.2.10.1. CRISPR/Cas9 plasmid construction ........................................................................................... 57 2.2.10.2. Surveyor and RFLP Assays ........................................................................................................ 58

Chapter 3. Pluripotency Gene Demethylation during Reprogramming ................... 59

3.1. Introduction ............................................................................................................... 59

3.2. Reprogramming of human B lymphocytes upon fusion with mouse embryonic stem cells ................................................................................................................................. 59

3.3. DNA methylation profiles of pluripotency associated genes in human B lymphocytes and human ES cells ......................................................................................................... 61

3.4. Changes in DNA methylation of OCT4 accompanies reprogramming but the extent is variable ............................................................................................................................ 62

3.5. Reprogramming human fibroblasts and OCT4 induction without detectable changes in DNA methylation ............................................................................................................. 64

3.6. No evidence of DNA methylation changes at site upstream of the OCT4 transcription start site. .......................................................................................................................... 66

3.7. DNA demethylation kinetics of somatic Oct4 transgene in reprogrammed mouse hybrids. ............................................................................................................................ 67

3.8. Summary and Discussion ......................................................................................... 68

Chapter 4. Mechanisms of Imprint Erasure in Somatic Cells mediated by Embryonic Germ Cell Fusion ...................................................................................... 72

4.1. Introduction ............................................................................................................... 72

4.2. Imprint erasure in somatic cells induced by embryonic germ cell fusion. ................... 72

4.3. Using dual reporter (2rB) somatic cells to assess the kinetics of imprint erasure during EG-reprogramming. ......................................................................................................... 73

4.4. EG cell capacity to induce demethylation is not restricted to imprinted genes. .......... 75

4.5. Imprint erasure is not seen in fusions with mouse ES cells or female ES cells that are globally hypomethylated. ................................................................................................. 76

4.6. Hydroxymethylation at imprinted loci upon fusion with mouse EG cells. ................... 78

4.7. Tet regulated 5-mC oxidation at imprinted loci upon fusion with mouse EG cells. ..... 79

4.8. Summary and Discussion ......................................................................................... 81

7

Page 8: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5. Analysis of TET Protein Requirement in Mouse Embryonic Stem Cell Induced Reprogramming of Human B Lymphocytes................................................ 84

5.1. Introduction ............................................................................................................... 84

5.2. Tet Knockdown in mouse ES cells and cell fusion .................................................... 84 5.2.1. Tet1 knockdown and fusion ............................................................................................................... 84 5.2.2. Tet2 knockdown and fusion ............................................................................................................... 87 5.2.3. Tet1 and Tet2 double knockdown and fusion .................................................................................... 89

5.3. CRISPR/Cas9 mediated Tet gene editing and cellular fusion ................................... 91 5.3.1. CRISPR/Cas9 system construction against Tet1 and Tet2 genes and delivery into mES cells ............ 91 5.3.2. Surveyor Assay for analysis of CRISPR/Cas efficiency ........................................................................ 94 5.3.3. Restriction Fragment Length Polymorphism screen on CRISPR/Cas9 targeted mES cells for Tet1 and Tet2 .............................................................................................................................................................. 95 5.3.4. Sequencing of Tet1&Tet2 CRISPR targeted ES clones ........................................................................ 97 5.3.5. Reprogramming capacity of CRISPR/Cas9 mediated Tet1 and Tet2 mutant ES cell clones upon cell fusion ........................................................................................................................................................... 99

5.4. Summary and Discussion ....................................................................................... 100

Chapter 6. CRISPR/Cas Editing of Jarid2 and Non-Canonical WNT Pathway Components…………………………………………………………………………………. 106

6.1. Introduction ............................................................................................................. 106

6.2. CRISPR/Cas9 editing of Jarid2 and Prickle1/Fzd2/Wnt9a in mouse embryonic stem cells ............................................................................................................................... 106

6.2.1. Guide RNA design and delivery into mouse ES cells ........................................................................ 107 6.2.2. Surveyor Assay for analysis of CRISPR/Cas9 efficiency .................................................................... 108 6.2.3. Clonal screens and sequencing for targeted Jarid2 locus in mouse ES cells .................................... 108 6.2.4. Clonal screens and sequencing for targeted Prickle1, Fzd2 and Wnt9a loci in mES cells. ............... 110

6.3. JARID2 deficiency in mouse ES cells can be phenocopied by Prickle1/Fzd2/Wnt targeting ........................................................................................................................ 113

6.4. Summary and Discussion ....................................................................................... 115

Chapter 7. General Discussion ................................................................................. 118

7.1. DNA methylation dynamics in reprogramming ........................................................ 118

7.2. Genome editing and the use of CRISPR/Cas9-based approaches ......................... 120

7.3. Future Studies ........................................................................................................ 123 7.3.1. Droplet-based microfluidics and cell fusion ..................................................................................... 124 7.3.2. Single-cell heterokaryon analysis ..................................................................................................... 126

Bibliography................................................................................................................ 128

Appendix...................................................................................................................... 155

Publications................................................................................................................. 158

8

Page 9: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

List of Figures & Tables

FIGURES

Figure 1.1. Maintenance/replication-coupled loss of DNA methylation

Figure 1.2. Mechanisms of dynamic modifications of cytosine

Figure 1.3. TET-induced demethylation in mouse embryonic development and ES cells

Figure 1.4. In-vitro strategies for nuclear reprogramming to pluripotency

Figure 2.1. Vectors used for delivery

Figure 2.2. px330 vector and the guide RNA sequence

Figure 3.1. Interspecies cell fusion and reprogramming of human B lymphocyte by mouse ES cells

Figure 3.2. Bisulfite sequencing of OCT4, NANOG and CRIPTO promoters in human B lymphocytes and human ES cells

Figure 3.3. Transcript analysis and bisulfite sequencing of human OCT4 in heterokaryons after 72 hours of fusion

Figure 3.4. Transcript analysis and bisulfite sequencing of human OCT4 in heterokaryons after 72 hours of fusion in four different experiments

Figure 3.5. Fusion of human fibroblasts with mouse ES cells and transcript and bisulfite sequencing analyses

Figure 3.6. Bisulfite sequencing of human OCT4 upstream region in human fibroblasts before and after fusion and in human ES cells

Figure 3.7. DNA demethylation kinetics upon reprogramming in mouse hybrids

Figure 4.1. CpG methylation analysis of imprinted H19 locus upon reprogramming mediated by mouse EG cells

Figure 4.2. CpG methylation analysis of imprinted Peg1 locus upon pluripotent reprogramming mediated by mouse EG cells

Figure 4.3. Functional resetting of somatic imprints mediated by mouse EG cells

Figure 4.4. CpG methylation analysis of LINE1 repeats upon pluripotent reprogramming mediated by mouse EG cells

Figure 4.5. CpG methylation analysis of imprinted H19, Peg3 and Gtl2/Dlk1 loci upon pluripotent reprogramming induced by mouse ES cells

Figure 4.6. CpG methylation analysis of imprinted H19, Peg3 and Gtl2/Dlk1 loci upon pluripotent reprogramming induced by Pgk12.1 female mouse ES cells

Figure 4.7. Acquisition of 5-hmC at human B lymphocyte ICRs upon fusion with mouse EG cells

Figure 4.8. Roles of TET proteins in the acquisition of 5-hmC at human B lymphocyte ICRs upon fusion with mouse EG cells.

Figure 5.1. Effect of Tet1 knockdown on reprogramming

Figure 5.2. Effect of Tet2 knockdown on reprogramming

9

Page 10: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Figure 5.3. Effect of Tet1/Tet2 double knockdown on reprogramming

Figure 5.4. Tet1 and Tet2 targeting by CRISPR/Cas

Figure 5.5. Workflow for CRISPR/Cas mediated gene targeting in mouse ES cells

Figure 5.6. Schematic Representation of Surveyor Assay.

Figure 5.7. Surveyor Assay on Tet1 and Tet2 in wild type and CRISPR/Cas targeted mES cells, treated with Puroymcin or mCherry sorted upon co-transfection

Figure 5.8. Schematic Representation of RFLP

Figure 5.9. RFLP Assay on WT and Tet1&Tet2 CRISPR co-targeted 32 clones

Figure 5.10. DNA sequencing results on Tet1&Tet2 CRISPR co-targeted loci

Figure 5.11. Effect of Tet1/Tet2 knockout on reprogramming

Figure 5.12. HP1α redistribution in mESxhF heterokaryons

Figure 6.1. Jarid2, Prickle1, Fzd2 and Wnt9a targeting by CRISPR/Cas

Figure 6.2. Surveyor Assay on Puromycin treated populations of Jarid2 single and Prickle1, Wnt9a, Fzd2 triple CRISPR/Cas targeted mES cells

Figure 6.3. Surveyor Assay on CRISPR/Cas targeted single mES cell clones for Jarid2

Figure 6.4. DNA sequencing results on Jarid2 CRISPR/Cas targeted locus

Figure 6.5. Western Blot detection of Jarid2 in wild type and CRISPR/Cas targeted clones

Figure 6.6. Surveyor Assay on CRISPR/Cas triple targeted single mES cell clones for Fzd2

Figure 6.7. Surveyor Assay on CRISPR/Cas triple targeted single mES cell clones for Prickle1 and Wnt9a

Figure 6.8. DNA sequencing results on Prickle1, Fzd2 and Wnt9a CRISPR/Cas co-targeted loci

Figure 6.9. mRNA levels of CRISPR/Cas targeted Prickle1, Fzd2 and Wnt9a in selected clones and wild type cells

Figure 6.10. mRNA levels of Prickle1, Fzd2 and Wnt9a in Jarid2 mutant lines and wild type cells

Figure 6.11. Flow Cytometry analysis of Nanog expression in mES cells Clones 12C and 2D

Figure 6.12. Off-target identification of CRISPR/Cas targets

Figure 7.1. Schematic representation of RNA-guided Cas9 targeting on DNA.

TABLES

Table 1 Primers for transcript analysis by quantitative RT-PCR

Table 2 Primers for bisulfite sequencing analysis

Table 3 Primers for enzyme protection assay

Table 4 Primers for ChIP assay

Table 5 Primers for genomic DNA amplification for Surveyor and RFLP Assays

10

Page 11: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Abbreviations

2i Small molecule inhibitors of MEK and GSK3β 3C Chromosome Conformation Capture 5-caC 5-carboxycytosine 5-fC 5-formylcytosine 5-hmC 5-hydroxymethylcytosine 5-hmU 5-hydroxymethyluracil 5-mC 5-methylcytosine β-gal β-galactosidase µ Micro Ac acetyl group BER Base excision repair bp base pair CAGE Cap analysis of gene expression Cas CRISPR-associated cDNA Complementary DNA CGI Cytosine Guanine dinucleotide islands CHD Chromodomain helicase DNA-binding ChIP Chromatin immunoprecipitation CpG Cytosine Guanine dinucleotide Ct Threshold Cycle CRISPR clustered regularly interspaced short palindromic repeats Ctrl Control DAPI 4,6-diaminido-2-phenylindole DMEM Dulbecco’s Modified Eagle DMR Differentially methylated region DNA Deoxyribonucleic acid DNMT DNA methyltransferase DSB Double strand break E Embryonic day EC Embryonic carcinoma EDTA Ethylene diamine tetraacetic acid EdU 5-ethynyl-2-deoxyuridine EG Embryonic germ ES Embryonic stem FACS Fluorescence activated cell sorting FBS Fetal Bovine Serum FGF Fibroblast Growth Factor FRAP Fluorescence recovery after photobleaching g Gram GFP Green Fluorescent Protein H Histone hB human B HCP High CpG density promoter HDAC Histone deacetylase HDR Homology-directed break HP1α Heterochromatin Protein 1α HR Homologous recombination ICM Inner cell mass ICP Intermediate CpG density promoter

11

Page 12: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

ICR Imprinting control region Indel Insertion / Deletion iPS Induced pluripotent stem kb Kilo base K Lysine KRAB Krüppel associated box KO Knock-out l Litre LCP Low CpG density promoter LINE Long interspersed element LIF Leukemia Inhibitory Factor m Milli M Molar MBD Methyl binding domain Me methyl group MET mesenchymal-to-epithelial NEAA Non-essential aminoacids n Nano Neo Neomycin NHEJ Non-homologous end joining ntES Nuclear-transfer-embryonic stem NURD Nucleosome Remodelling Deacetylase PAM Protospacer Adjacent Motif PBS Phosphate Buffered Saline PCR Polymerase Chain Reaction PEG Polyethylene Glycol PGC Primordial germ cell PN Pronuclear PRC Polycomb repressor complex RFLP Restriction Fragment Length Polymorphism RNA Ribonucleic acid RNAi RNA interference RT Reverse Transcription S Serine SAM S-adenosyl-l-methionine SCNT Somatic-cell nuclear transfer sgRNA single guide RNA shRNA short hairpin RNA SNP Single nuclear Polymorphism SWI/SNF SWItch/Sucrose NonFermentable TF Transcription Factor TKO Triple Knock-out tracrRNA trans-activating crRNA U Unit WT Wild type X X-chromosome

12

Page 13: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

Chapter 1. Introduction

The faithful transmittance of the genetic material from mother to daughter cells

ensures that every cell of a higher eukaryotic organism shares the same genomic

sequence, originated from a single fertilised egg. Yet, each cell type performs a discrete

function that depends on information decoded from the DNA. Cell fate is acquired in the

course of development, characterised by a distinct gene expression profile, and is

maintained by epigenetic mechanisms.

1.1. Pluripotent embryonic stem cells

Cells within the inner cell mass (ICM) of the embryo are pluripotent and possess

the ability to differentiate into all three germ layers. Although the ICM exists transiently in

the blastocysts during pre-implantation development, it is possible to derive cells from this

stage (Evans and Kaufman, 1981; Martin, 1981). These in-vitro counterparts are known

as embryonic stem (ES) cells, they are pluripotent and mouse ES cells can be indefinitely

propagated in culture on feeder embryonic fibroblasts or in the presence of leukaemia

inhibitory factor (LIF). Remarkably, once injected into the blastocysts, ES cells can give

rise to both adult somatic and germ cells following mouse chimera production. Although

human ES cells share main properties with mouse ES cells including self-renewal and

pluripotency, they are morphologically different, proliferate with slower kinetics, require

fibroblast growth factor 2 (FGF2) and Activin/Nodal pathway activity and proposed to

exhibit similar characteristics to mouse epiblast stem cells (Schnerch et al., 2010).

ES cell state is maintained by tight control over gene regulation with OCT4, SOX2

and NANOG providing the core transcription factors (TFs) that underwrite the pluripotency

network (Boyer et al., 2005). The equilibrium between pluripotency and differentiation is

established by the relative levels of TF expression. For example, a modest increase of

Oct4 expression triggers mesodermal and endodermal differentiation, while its decrease

directs differentiation towards a trophectoderm lineage (Niwa et al., 2000). On the other

hand, mono-allelic disruption of Oct4 results in increased genomic OCT4 protein

occupancy, maintenance of a stable pluripotent state and a delay in differentiation

(Karwacki-Neisius et al., 2013). Similarly, changes in SOX2 levels cause deregulation of

pluripotency (Kopp et al., 2008). Interestingly, ES cells regularly oscillate between high

and low Nanog expression levels, where the former state associates with efficient self-

13

Page 14: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

renewal, while the latter is linked to a propensity to differentiate (Chambers et al., 2007).

Similar fluctuations have been observed for additional pluripotency network components

including Essrb, Klf4, Tbx3, Rex1 and Stella (Cahan and Daley, 2013). Homogeneous

high expression of pluripotency genes can be achieved by culturing mouse ES cells in the

presence of two small-molecule inhibitors (PD0325901 and CHIR99021, denoted as 2i)

that selectively target mitogen-activated protein kinase kinase (MEK) and glycogen

synthase kinase-3 (GSK3) (Wray et al., 2010; Ying et al., 2008). Moreover, mouse ES

cells cultured under 2i condition exhibit reduced levels of lineage-specific genes, reduced

bivalent (H3K27me3/H3K4me3) marking (see below) and genome-wide DNA

hypomethylation (see section 1.2) (Leitch et al., 2013; Marks et al., 2012). On this basis,

ES cells maintained in 2i conditions are proposed to exist in a ground state of pluripotency

(Wray et al., 2010).

The OCT4 interactome is composed of over 160 proteins, including pluripotency-

associated transcription factors SALL4, ESRRB, DAX1 and TCFCP2I1 and chromatin

modifiers, such as NURD and SWI/SNF complex components (van den Berg et al., 2010).

Acute depletion of OCT4 results in reduced recruitment of transcription factors on their

genomic targets, suggesting that OCT4 provides an anchor point for the assembly and

co-localisation of associated factors. Core transcription factors share a significant number

of common targets and can regulate their own expression, which may be a key concept

in pluripotent homeostasis (Boyer et al., 2005). Interestingly, genes that show a high level

of transcription factor occupancy tend to be active, while those bound by only a few

transcription-factors tend to be silent, implying that gene regulation is linked to

combinatorial localisation of pluripotency network components (Kim et al., 2008).

Moreover, core transcription factors and their interactors show preferential binding at

enhancers and NANOG-OCT4-SOX2 binding can recruit p300, a histone

acetyltransferase to the enhancer regions (Chen et al., 2008). These ES-cell specific

enhanceosomes can interact with the target gene promoters to induce gene expression

(Chen et al., 2008), and this interaction may be facilitated by mediator and cohesin

complexes forming a DNA loop structure, as demonstrated by chromosome conformation

capture (3C) technique (Kagey et al., 2010). Inspection of several ES-specific gene

enhancers has revealed that they comprise clusters of constituent enhancers spanning

large distances (as much as 50 kb). These so-called “super-enhancers” are enriched for

OCT4, SOX2, NANOG, KLF4 and ESRRB binding, and contain high levels of the Mediator

coactivator complex that is shown to be associated with general transcription factors and

RNA polymerase II (Whyte et al., 2013).

14

Page 15: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

1.1.1. Chromatin dynamics in embryonic stem cells

The basic repeating unit of eukaryotic chromatin is the nucleosome, which consists

of a central histone octamer core (normally composed of two copies of each of the

histones H2A, H2B, H3 and H4) wrapped by a segment of ~147 base pairs of DNA and a

linker DNA associated with histone H1 (Luger et al., 1997). Gene expression activity

correlates with post-translational modifications of the histone residues that can impact

expression by two main mechanisms. The first one is via influencing the physical structure

of the chromatin, and the second by creating docking sites for the binding of effector

proteins (Bannister and Kouzarides, 2011). These modifications occur at the N-terminal

histone tails or within the globular domains on specific residues [such as lysine (K) or

serine (S)] and include acetylation, mono- di- and tri-methylation, phosphorylation and

ubiquitylation, that can crosstalk or fine-tune the transcriptional readout (Lee et al., 2010).

Originally termed by Emil Heitz in 1928 (Heitz, 1928) according to the staining profiles,

chromatin exists in two forms; densely packed, transcriptionally inactive heterochromatin

and de-condensed, transcription permissive euchromatin. Both structures can be

recognised by their signature histone marks; while H3K27me3 and H3K9me3 are

generally associated with heterochromatin, H3K4me3 and histone acetylation often

associate with euchromatin. Collectively, these histone modifications, and others, have

been postulated to establish a ‘histone code’ that forms the basis of epigenetic regulation

of gene expression (Jenuwein and Allis, 2001).

Embryonic stem cells have been reported to have a generalised ‘open chromatin

structure’, a feature that is also observed in the ICM, which gradually condenses in the

course of lineage commitment (Ahmed et al., 2010; Azuara et al., 2006; Meshorer et al.,

2006). Remarkably, core histones and structural proteins of the chromatin, including HP1,

are hyperdynamic in pluripotent cells, as demonstrated by sensitivity to salt extraction and

by fluorescence recovery after photobleaching (FRAP) (Bhattacharya et al., 2009;

Meshorer et al., 2006). Consistently, electron microscopy indicates that heterochromatic

regions are infrequent in ES cells, where H3K9me3 modification is underrepresented

(Efroni et al., 2008; Meshorer et al., 2006). In addition, histone H3K27me3, which is not

abundant in ES cells, is significantly increased upon ES cell differentiation (Hawkins et al.,

2010; Zhu et al., 2013), accompanied by reduced H3K9 acetylation levels (Krejci et al.,

2009). From a higher-order genome organization perspective, OCT4, SOX2 and NANOG

binding sites co-localise at three dimensional space establishing transcriptionally active

regions in the nucleus (Denholtz et al., 2013), while inactive regions form less long-range

15

Page 16: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

interactions in ES cells than differentiated cells (de Wit et al., 2013). In pluripotency, an

open chromatin landscape may actively be maintained by chromatin remodelling factors

including CHD (chromodomain helicase DNA-binding) family members that can slide or

eject nucleosomes in an ATP-dependent manner to promote transcription (Clapier and

Cairns, 2009). In this regard, Chd1 knockdown in ES cells leads to accumulation of

heterochromatin, and disrupted differentiation capacity (Gaspar-maia et al., 2009).

1.1.2. Polycomb regulation and bivalency

Although ES cell chromatin is open and dynamic, expression of lineage-specific

genes must be restricted for the maintenance of pluripotency. Silencing of developmental

regulators in ES cells is primarily mediated by polycomb repressor complex (PRC)

proteins which exist in two sub-complexes. PRC1 catalyses the mono-ubiquitinylation of

histone H2A on position 119 (H2AK119ub) and PRC2 is responsible for H3K27me3

deposition (Boyer et al., 2006). It has recently been demonstrated that CXXC domain

containing proteins are involved in targeting PRC components for gene repression.

Notably, KDM2B has been shown to recruit non-canonical PRC1 components on CpG

island promoters by associating with unmethylated DNA via its CXXC domain, leading to

targeted mono-ubiquitinylation of H2AK119 (Farcas et al., 2012; Wu et al., 2013a). This

results in recruitment of PRC2 that in turn generate H3K27me3 for repression of gene

expression (Blackledge et al., 2014). Consequently, a positive feedback loop is

established, in which H3K27me3 further promotes binding of the canonical PRC1 on the

chromatin to expand the PRC silenced domain (Kalb et al., 2014). Recently, a Jumonji

histone demethylase family member JARID2 has been identified to associate with PRC2

and is critical for mouse development and ES cell differentiation (Landeira and Fisher,

2011). Differentiation defects observed in Jarid2-/- ES cells could reflect the lack of serine

5 phosphorylated RNA polymerase II enrichment at the bivalent domains (Landeira et al.,

2010). More recent analyses have revealed that, although catalytically inactive (Klose et

al., 2006) JARID2 exhibits nucleosome-binding activity and promotes PRC2 recruitment

to nucleosomes (Son et al., 2013), which is in part mediated by long noncoding RNAs

(Kaneko et al., 2014).

In ES cells, while developmentally regulated genes are silent, they possess the

capacity to be rapidly activated upon external differentiation signals. These silenced

lineage-specific regulators possess both repressive H3K27me3 and active H3K4me3

histone marks at their promoters, a property known as bivalency, which is thought to be

16

Page 17: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

important for fast gene reactivation (Azuara et al., 2006; Bernstein et al., 2006). Of

importance, bivalent gene promoters are occupied by serine 5 phosphorylated RNA

polymerase II, which marks transcription initiation, but not elongation, indicating that the

transcription is ‘poised’ but ready to be initiated at these genes (Brookes et al., 2012; Stock

et al., 2007).

1.2. DNA methylation dynamics in mammals

DNA methylation mainly involves cytosine residues. Cytosine methylation on the

fifth carbon (abbreviated as 5-mC) is an essential epigenetic mark for mammalian

development and homeostasis, and is often found in the context of symmetrical CpG

dinucleotides (Bird, 2002). While non-clustered CpGs are often methylated, clustered

CpGs (so called CpG islands, CGIs, which are on average 1000 base pairs long with high

C+G density) exhibit hypomethylation (Deaton and Bird, 2011). It is important to note that

CpG dinucleotide sequences outside CGIs are evolutionarily underrepresented in the

animal genome due to spontaneous deamination of methylcytosine to thymidine and

accumulation of TpG dinucleotides because of inaccurate mismatch repair (Bird, 1980;

Lander et al., 2001). DNA methylation is involved in key developmental processes that

involve regulation of gene expression, repetitive element silencing, X-inactivation and

imprinting (Smith and Meissner, 2013).

1.2.1. Maintenance and de novo establishment of DNA methylation

Methylation of the cytosine residues is catalysed by a group of DNA

methyltransferase (DNMT) enzymes; DNMT1 is responsible for maintenance, and

DNMT3A and DNMT3B mediate establishment of de novo methylation patterns (Goll and

Bestor, 2005). DNMT2, despite possessing sequence and structural characteristics of

DNA methyltransferases, methylates transfer RNA, but not DNA (Goll et al., 2006). A last

member, DNMT3L, is catalytically inactive and lacks DNA affinity, but stimulates DNMT3A

and 3B activities by direct interaction (Suetake et al., 2004).

Maintenance of DNA methylation is based on the recognition of hemi-methylated

CpGs, sites that are generated by incorporation of the unmodified cytosine into the newly

synthesised DNA during semiconservative DNA replication. DNMT1 can then methylate

these sites to propagate the information at every cell cycle. DNMT1 was the first DNA

17

Page 18: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

methyltransferase to be cloned and sequenced (Bestor et al., 1988), and its depletion

results in embryonic lethality accompanied with significantly reduced 5-mC levels in the

embryo (Li et al., 1992). Similarly, DNMT1 deficient ES cells undergo rapid genome-wide

demethylation, yet are viable and retain low level but stable DNA methylation patterns (Lei

et al., 1996). DNMT1 is highly abundant in mitotic cells (Kishikawa et al., 2003) and is

recruited to the replication foci during DNA replication (Leonhardt et al., 1992).

Recruitment is mediated by its tethering factor UHRF1 (also known as NP95), which

specifically recognises and binds hemi-methylated CpG sites (Bostick et al., 2007; Sharif

et al., 2007). This results in the flipping of the 5-mC out of the DNA helix that is positioned

in the SRA domain of UHRF1, which then correctly orients unmethylated cytosine to

DNMT1 for methylation (Arita et al., 2008; Avvakumov et al., 2008; Hashimoto et al.,

2008). In addition, UHRF1 can recognise H3K9me2/me3 marks, which is proposed to

enhance DNMT1 recruitment and DNA methylation (Liu et al., 2013). Of importance, Uhrf1

deletion results in substantially decreased DNA methylation levels due to lack of

maintenance and Uhrf1 deficiency causes embryonic lethalily (Sharif et al., 2007).

DNMT3A and DNMT3B, the de novo methyltransferases, are required for

establishment of DNA methylation patterns in early mammalian development (Okano et

al., 1998, 1999). DNMT3B was first shown to specifically methylate pericentromeric

regions (Okano et al., 1999) and is expressed in the ICM, epiblast and the embryonic

ectoderm, while DNMT3A is not detected at these stages (Watanabe et al., 2002). Instead

Dnmt3a is expressed between E10.5-E14.5 embryos (Watanabe et al., 2002), partly

related to its role in the establishment of Imprint Control Regions (ICRs) (Kaneda et al.,

2004). Although both proteins share homology, deletion of Dnmt3b results in embryonic

death, while Dnmt3a-null mice die after birth (Okano et al., 1999). However, combined

deletion of both genes causes earlier embryonic lethality, revealing some level of partial

redundancy between the two enzymes and overlapping functions during embryogenesis

(Okano et al., 1999). In ES cells that lack both DNMT3A and 3B, loss of methylation is

gradual and requires long term passaging (Chen et al., 2003). In addition,

Dnmt1/Dnmt3a/Dnmt3b triple KO (TKO) ES cells exhibit extensive loss of CpG

methylation, but retain their self-renewal capacity and undifferentiated state, suggesting

that in ES cells other epigenetic mechanisms are sufficient to maintain stable

heterochromatin and chromosome structures (Tsumura et al., 2006). However, although

TKO cells can contribute to the blastocyst stage and differentiate into extraembryonic

tissues in vivo, they do not contribute to the embryonic lineages (Sakaue et al., 2010).

18

Page 19: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

1.2.2. Roles of DNA methylation in gene regulation

CpG islands constitute essential platforms for DNA methylation to exert its role as

transcriptional repressor, in combination with chromatin modifications (Deaton and Bird,

2011). Around 70% of human gene promoters are associated with CpG islands (Saxonov

et al., 2006). These high CpG density promoters (HCPs) are involved in regulation of

housekeeping as well as key developmental genes, and are generally free of DNA

methylation (Saxonov et al., 2006). This is evidenced by cap analysis of gene expression

(CAGE) technique that revealed the correlation between initiation of transcription and

presence of CGIs (Carninci et al., 2006). It is important to mention that CGI promoters can

be methylated and methylation provides long-term stabilization of silencing, including

imprinted and inactive X-chromosome genes. Two additional classes are intermediate and

low CpG density promoters (ICPs and LCPs, respectively), which exhibit more frequent

DNA methylation (Weber et al., 2007). Although LCPs are generally methylated,

methylation does not have an influence on gene expression, which remains active (Borgel

et al., 2010; Weber et al., 2007). On the other hand, ICP methylation results in gene

silencing, examples of which include repression of pluripotency-associated and germline

specific genes during differentiation (Borgel et al., 2010; Farthing et al., 2008; Meissner et

al., 2008).

CGI promoters are enriched for H3K4me3, a chromatin mark of actively transcribed

genes. However it is important to note that not all of the analysed promoters carrying this

mark are active (Guenther et al., 2007; Mikkelsen et al., 2007). Association of H3K4me3

with the CGIs is in part mediated by CFP1, which binds to the unmethylated CpGs upon

its interaction with SETD1, a H3K4 methyltransferase (Clouaire et al., 2012; Thomson et

al., 2010). H3K4me3 and histone variant H2A.Z, which also mark nucleosome depleted

transcription start sites, protect promoters from acquisition of de novo methylation (Ooi et

al., 2007; Zilberman et al., 2008). Similarly, KDM2A binding to unmodified CpGs in CGIs

results in depletion of di-methylated H3K36, necessary for a permissive environment for

transcriptional machinery (Blackledge et al., 2010). On the other hand, another H3K36

demethylase ,KDM2B, is required for recruiting PRC1 to the unmethylated CGIs and is

associated with silencing of genes involved in embryonic development and cellular

differentiation in ES cells (as described earlier, (Farcas et al., 2012)).

Gene repression mediated by CpG methylation can be attributed to the

combinatorial effects of inhibition of transcription factor binding and recruitment of

chromatin re-modellers by methyl-binding domain (MBD) proteins (Klose and Bird, 2006).

19

Page 20: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

However, it is speculated that gene silencing precedes DNA methylation, which then acts

as a lock to stabilise the repressed status (Jones, 2012). For example, silencing of Oct4

and Nanog genes during carcinoma cell differentiation is accompanied with chromatin

remodelling of regulatory sequences that are subsequently methylated by DNMT3A (You

et al., 2011). This is in agreement with the fact that G9A, a H3K9 di-methyltransferase,

promotes de novo methylation during differentiation, suggesting that DNA methylation

follows changes in chromatin structure and histone marks (Dong et al., 2008; Epsztejn-

Litman et al., 2008; Tachibana et al., 2008). For example, ectopic targeting of HP1α at the

Oct4 locus in ES cells results in H3K9 tri-methylation and gene silencing that are later

followed by DNA methylation (Hathaway et al., 2012). Furthermore, H3K9 tri-methylation

is an early event in X-inactivation leading to gene repression, which precedes CGI

methylation (Mermoud et al., 2002). Histone marks around methylated CpGs are also

important for mitotic inheritance of DNA methylation. This is exemplified by the

requirement of UHRF1 binding to the methylated H3K9 to enable the recruitment of DNA

methylation maintenance machinery (Rothbart et al., 2012). Therefore, DNA methylation

can be considered as a provider of high-fidelity epigenetic memory by stabilising gene

repression.

1.2.3. Genomic Imprinting

Genomic imprinting ensures mono-allelic gene expression that is dependent on the

parental origin. DNA methylation is the major epigenetic component of genomic imprinting

and marks differentially methylated regions (DMRs) in the genome. These DMRs are

established during gametogenesis (denoted as germline DMR) or after fertilization

(denoted as somatic DMRs) and can functionally control expression of imprinted genes in

cis within clusters, acting as imprinting control regions (ICRs) (Ferguson-Smith, 2011).

While the majority of methylated germline DMRs are maternally inherited, thus far only

four germline DMRs, H19-Igf2, Dlk1, Rasgrf1, and Zdbf2, are characterised as paternally

silenced (John and Lefebvre, 2011). Since the initial discovery of imprinted Igf2r, Igf2 and

H19 loci (Barlow et al., 1991; Bartolomei et al., 1991; DeChiara et al., 1991; Ferguson-

Smith and Cattanach, 1991) more than a hundred genes in mouse and human have been

identified as imprinted (Henckel and Arnaud, 2010). Genomic imprinting is involved in

many processes such as embryonic growth (Smith et al., 2006a), placental development

(Frost and Moore, 2010), metabolism (Radford et al., 2011) and behaviour (Wilkinson et

20

Page 21: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

al., 2007), and perturbations are linked to several human disorders including Prader-Willi

and Angelman syndromes (Butler, 2009).

For new gamete-specific methylation to be established in the germline, parental-

specific DNA methylation marks at the DMRs must first be erased. This occurs during

proliferation and migration of the primordial germ cells (PGCs) to the gonads and is

completed by embryonic day 13.5 in both male and female mouse embryos (Guibert et

al., 2012; Hajkova et al., 2002; Popp et al., 2010; Seisenberger et al., 2012). At that point

parent-of-origin-specific de novo germline DMR methylation starts and is completed during

the early neonatal period in females and late foetal development in males (Hajkova et al.,

2008; Lucifero et al., 2004). During this period, DNA methyltransferase DNMT3A is

recruited at DMRs via interaction with DNA methyltransferase-like protein DNMT3L, and

establishes de novo methylation of both maternal and paternal imprinted loci (Hata et al.,

2002; Kaneda et al., 2004). DMR methylation is then maintained through DNA replication

by DNMT1.

One important difference between maternal and paternal DMRs is their genomic

location; while maternal DMRs are intragenic and are mostly found at transcription start

sites of protein-coding or non-coding RNA genes, paternal DMRs are intergenic (Edwards

and Ferguson-Smith, 2007). In male foetal germ cells, promoter associated maternal

DMRs are transcriptionally active and are enriched for H3K4me3 (Henckel et al., 2012).

Mechanistically, interaction of DNMT3L with the DNA is strongly inhibited by this histone

modification thus preventing the recruitment of DNMT3A for de novo methylation (Ooi et

al., 2007). This correlates with the requirement of KDM1B, a H3K4 demethylase that is

highly expressed in oocytes, for proper establishment of DNA methylation at several

maternal DMRs (Ciccone et al., 2009). Many other histone marks and trans-acting

elements are collectively involved in the establishment of imprinting, and consequent

designation of actual parental-specific imprinting occurs during the global DNA

demethylation wave that occurs soon after fertilization (Kelsey and Feil, 2013).

Global erasure of DNA methylation happens shortly after fertilization in both

parental pronuclei. This global DNA demethylation is thought to be necessary for the

acquisition of totipotency and the establishment of the developmental programme. This

occurs asymmetrically; CpG methylation levels are rapidly reduced in the paternal

pronucleus at the one-cell stage, while the maternal pronucleus undergoes a gradual loss

of DNA methylation throughout pre-implantation development (Santos et al., 2002). At this

stage, imprinted genes must be selectively protected from the global wave of DNA

demethylation. It has recently demonstrated that paternal pronucleus undergoes

21

Page 22: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

extensive TET3-mediated CpG hydroxylation (Gu et al., 2011; Iqbal et al., 2011; Wossidlo

et al., 2011), while the maternal pronucleus remains protected from TET3 activity by a

maternal factor, STELLA (also known as PGC7 or DPPA3) (Nakamura et al., 2012). In

addition, STELLA also protects maternal germline DMRs at Peg1, Peg3, Peg10 loci as

well as paternal germline DMRs at H19 and Rasgrf1 loci from demethylation (Nakamura

et al., 2007). However, STELLA’s specificity for imprinted genes is questionable as its

protective function is genome-wide (Nakamura et al., 2012). ZFP57, a zinc finger protein

with a KRAB (Krüppel associated box) domain, stands as a strong candidate for

imprinting-associated maintenance and its deficiency in mouse embryos results in

hypomethylation of both parental germline DMRs (Li et al., 2008). ZFP57 binds methylated

alleles of germline DMRs in mouse ES cells, via sequence-specific recognition of a

methylated hexanucleotide motif (TGCCGC) found in all murine ICRs (Quenneville et al.,

2011). ZFP57 belongs to a family of proteins that interact with TRIM28 (also known as

KRAB-associated protein 1, KAP1) a component of a multifunctional repressive complex,

which in turn brings repressive histone marks and DNA methyltransferases

(DNMT1/3A/3B) on to the zinc finger bound DNA (Iyengar and Farnham, 2011). Indeed,

ChIP experiments in mouse ES cells revealed co-localisation of ZFP57, TRIM28 and

H3K9me3 at methylated alleles of ICRs (Quenneville et al., 2011). Moreover, maternal

loss of Trim28 leads to ICR hypomethylation and embryonic lethality (Messerschmidt et

al., 2012). Further investigation of DNA methylation on each individual cell of the 8-cell

blastomeres revealed that maternal TRIM28 deficiency causes mosaic demethylation at

the DMRs thus confirming its importance in pre-implantation development

(Lorthongpanich et al., 2013). Interaction of TRIM28 with ZFP57 leads to sequence-

specific recruitment of the DNA methylation maintenance machinery to the imprinted

genes (Zuo et al., 2012) and therefore ensures protection of methylation at imprinted loci

during pre-implantation embryo development (Messerschmidt, 2012).

Unmethylated alleles of DMRs must also be protected from de novo methylation.

One example is the association of zinc finger protein CTCF with H19 DMR. By specifically

binding onto the unmethylated CTCF binding sites in the maternal allele, CTCF brings the

H19 enhancer and promoter together (Murrell et al., 2004) to ensure transcriptional activity

that in turn prevents acquisition of DNA methylation (Engel et al., 2006). In addition, loop

formation at this region, stabilized by Cohesin (Nativio et al., 2009), prevents the

enhancers interacting with the promoter of Igf2 gene (which is therefore not activated on

the maternal allele). As CTCF does not bind to the methylated H19 DMR in the paternal

allele, Igf2 gene remains active via interaction between enhancers (Nativio et al., 2009).

22

Page 23: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

Early nuclear transplantation studies have demonstrated that mouse embryos that

contain biparental gynogenones (two female pronuclei) or biparental androgenones (two

male pronuclei) failed to undergo successful embryogenesis (McGrath and Solter, 1984;

Surani et al., 1984). This shows the importance and requirement of one paternal and one

maternal genome for normal development. Although the origins and evolution of genomic

imprinting are still being debated (Patten et al., 2014), it constitutes an ideal model system

to study epigenetic mechanisms and their interplay.

1.2.4. DNA demethylation

DNA methylation is a vital epigenetic modification in mammalian development.

Nonetheless, despite its chemical and hereditary stability, rapid loss of methylation is

observed in the zygote and during PGC development. Understanding the cause and

consequences of this vital switch is essential for elucidating the molecular details of life

cycle.

The sixth carbon of a cytosine nucleotide plays an important role in the chemistry

of methylation as it has electrophilic characteristics and it is therefore prone to be attacked

by nucleophilic cysteine thiolate, which is found on the catalytic site of DNA

methyltransferases. Enzyme binding on the sixth carbon of the cytosine leads to the

activation and increased nucleophilicity of the fifth carbon. Upon induction, the fifth carbon

receives an electrophilic attack mediated by the cofactor S-adenosyl-l-methionine (SAM),

which transfers a methyl group. This process is known as cytosine methylation, and its

stability arises from the established carbon-carbon bond (Smith et al., 1992). For this

reason, direct removal of the methyl group has been speculated to be energetically

unfavourable (Ooi and Bestor, 2008). Yet, especially in the past decade, several

mechanisms that mediate loss of DNA methylation have been reported, and are grouped

into two major pathways: (1) passive demethylation, via DNA replication-mediated dilution

of 5-methylcytosine; (2) active demethylation, via replication-independent enzymatic

activities. Interestingly, recent investigations have also revealed that both pathways can

co-exist and lead to loss of DNA methylation.

1.2.4.1. Passive demethylation

There are several ways for the passive DNA demethylation to occur: inhibition,

downregulation or exclusion of maintenance machinery, which is composed of DNMT1

23

Page 24: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

and UHRF1, or further modification of 5-mC that impairs its recognition (Figure 1.1). As a

consequence, upon DNA replication, the nascent cytosine that is incorporated into the

newly synthesized strand remains unmodified. One pharmacological implication of

passive demethylation is the usage of 5-azacytidine for treatment of diseases that stem

from hyper-methylation. 5-azacytidine is a cytosine derivative, where the fifth carbon atom

is substituted by a nitrogen atom, and can be incorporated into the newly synthesized

DNA. Although DNMT1 can recognise 5-azacytidine, it gets trapped on the DNA due to

chemically unresolved methylation reaction. This triggers DNA damage signalling

resulting in degradation of DNMT1. Therefore lack of nascent cytosine methylation leads

to gradual loss of DNA methylation via cell division (Stresemann and Lyko, 2008). Passive

demethylation has been described both in plants and animals. In plants, downregulation

of Dnmt1 homologue MET1 is essential for activation of imprinted genes during female

gametogenesis (Jullien et al., 2008). MET1 downregulation also in part explains activation

of transposable elements in the endosperm of Arabidopsis Thaliana (Slotkin et al., 2009).

Figure 1.1. Maintenance/replication-coupled loss of DNA methylation. (A) DNA replication at the methylated CpG sites results in hemimethylated DNA, which is restored to full methylation by maintenance machinery composed of UHRF1 and DNMT1. Downregulation, inhibition, or nuclear exclusion of UHRF1 and/or DNMT1 leads to replication-coupled dilution of 5-mC mark as incorporated cytosine to the newly synthesized DNA remains unmodified. (B) Iterative oxidation of the 5-mC mark prevents its recognition by the maintenance machinery, which similarly leads to progressive loss of methylation.

Similarly, nuclear exclusion and/or transcriptional repression of UHRF1 (which is the

tethering factor of DNMT1) during mouse germ cell development has been suggested to

24

Page 25: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

result in genome-wide DNA demethylation (Kagiwada et al., 2013; Seisenberger et al.,

2012). Moreover, nuclear exclusion of maintenance machinery causes progressive loss

of methylation levels in the mouse maternal DNA soon after fertilization (Cardoso and

Leonhardt, 1999). These examples indicate that passive demethylation is a common

mechanism to achieve relatively rapid and efficient resetting of global methylation

programme.

1.2.4.2. Active demethylation

Methylation-dependent binding protein MBD2 was initially proposed to be able to

directly remove the methyl group of the 5-methylcytosine in-vitro (Bhattacharya et al.,

1999; Ramchandani et al., 1999) and in-vivo (Cervoni and Szyf, 2001). Nevertheless, this

suggestion proved to be unreliable as the results were not reproduced in subsequent

studies, and mice lacking MBD2 showed normal methylation patterns (Hendrich et al.,

2001; Santos et al., 2002). Since then, although no protein with such enzymatic activity

has been identified, multi-layered enzymatic pathways that could lead to DNA

demethylation have been proposed, as outlined below (Ooi and Bestor, 2008).

Active demethylation is well characterized in plants. DNA glycosylases DME,

ROS1, DML2 and DML3 are involved in DNA demethylation of imprinting, silenced

transgenes as well as specific genomic loci. These enzymes can recognise and remove

5-mC from double stranded DNA irrespective of sequence context. They additionally

possess apyrimidinic lyase activity to cut off the abasic site that is left after 5-mC excision.

A nascent cytosine is then introduced to the excision site via DNA polymerase and DNA

ligase enzymes that are components of base excision repair (BER) pathway (Ikeda and

Kinoshita, 2009).

Mammalian orthologs of such DNA glycosylases have not yet been identified and

a universal pathway that is responsible for DNA demethylation has not been found. In

recent years, studies in many model systems have led to several different hypotheses with

contrasting mechanisms. BER pathway constitutes one of the proposed mechanisms for

active DNA demethylation in mammals (Wu and Zhang, 2010). For example, BER

components were shown to be recruited to the paternal pronucleus of the zygote where

DNA breaks occur after fertilization (Hajkova et al., 2010; Wossidlo et al., 2010). BER-

mediated DNA repair results in incorporation of unmodified cytosine, and the inhibition of

BER components such as PARP1 and APE1 was shown to interfere with loss of DNA

methylation (Hajkova et al., 2010). PARP1 was also reported to be part of active DNA

25

Page 26: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

demethylation in PGCs in DNA damage-dependent and –independent ways (Ciccarone

et al., 2012; Hajkova et al., 2010; Kawasaki et al., 2014). A screening in Xenopus

suggested that Gadd45a is a key regulator of active DNA demethylation in non-dividing

cells, by promoting DNA repair (Barreto et al., 2007). The link between BER and GADD45

was further elucidated in Danio rerio (Zebrafish) (Rai et al., 2008). GADD45 lacks

enzymatic activity, but in this study it was demonstrated to interact with MBD4, a thymine

glycosylase, and DNA deaminases AID (activation-induced deaminase) and APOBEC

(apolipoprotein B mRNA editing catalytic polypeptide) family members such as

APOBEC2A and APOBEC2B. According to the suggested model, 5-mC deamination

gives rise to a T:G mismatch, which is then repaired by MBD4-mediated removal of

thymine and incorporation of an unmodified cytosine (Rai et al., 2008). GADD45A was

also shown to interact with AID and thymine DNA glycosylase (TDG) that collectively

exhibit active DNA demethylation activity in a similar two-step model (Cortellino et al.,

2011). Interestingly, DNMT3A and 3B were reported to exhibit 5-mC deaminase activity

for the establishment of a dynamic TDG-coupled DNA demethylation/methylation process

upon external stimuli (Kangaspeska et al., 2008; Métivier et al., 2008). It should be noted

that although Mbd4 knockout mice are viable and fertile (Millar et al., 2002), Tdg deficiency

is associated with embryonic lethality (Cortellino et al., 2011). In the meantime, conflicting

results were reported regarding the involvement of GADD45A in active DNA

demethylation (Jin et al., 2008). Furthermore, knock-out mice for Gadd45a and its

homologous Gadd45b are viable and show no significant change in methylation levels

(Engel et al., 2009; Ma et al., 2009). However this does not rule out the possibility of

redundant roles of Gadd45 members (Niehrs and Schäfer, 2012). Although GADD45 is

not involved in global control of DNA demethylation, site-specific Gadd45-mediated DNA

demethylation may be induced upon various external stimuli, including neurogenesis in

adult mouse brain (Ma et al., 2009). Gene-specific loss of methylation mediated by

GADD45A were further demonstrated upon targeted recruitment by inhibitor of growth

protein 1 (ING) (Schäfer et al., 2013) and by long non-coding RNA TARID (Arab et al.,

2014).

The mechanism of AID-mediated deamination of 5-mC to thymine as an

intermediate for DNA demethylation was suggested to occur in PGCs, where AID

deficiency resulted in less efficient loss of DNA methylation compared to wild type mice

(Popp et al., 2010). The fact that AID knock-out mice are viable (Muramatsu et al., 2000)

suggests the presence of redundant mechanisms of DNA demethylation exerted by

APOBEC family members (Popp et al., 2010) or by other pathways (see below).

26

Page 27: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

Interestingly, AID was shown to be essential for experimental cell fusion-based

reprogramming, especially for demethylation and transcriptional reactivation of

pluripotency-associated genes OCT4 and NANOG (Bhutani et al., 2010). However, this

finding was later revoked by Foshay et. al., who demonstrated that Aid was not expressed

in the cell lines used in the former study. Furthermore, the authors reported that Aid over-

expression did not alter the reprogramming kinetics nor changed the DNA demethylation

kinetics after cell fusion (Foshay et al., 2012). Similar results have recently been shown

regarding the role of AID in induced reprogramming. While one report suggested that

acute loss of AID affects reprogramming efficiency (Bhutani et al., 2013), subsequent

reports could not reproduce this finding, and suggested that AID is dispensable for early

reprogramming of iPS cells (Habib et al., 2014; Shimamoto et al., 2014), but might be

required later for the stabilisation of the pluripotency programme (Kumar et al., 2013). It is

worth mentioning that biochemical analysis of AID/APOBEC activity revealed a

substantially lower propensity (~10-fold) for 5-mC deamination relative to their canonical

substrate, cytosine (Nabel et al., 2012). Moreover, AID/APOBEC requires single-stranded

DNA for efficient deamination that has not been found in the investigated model systems

(Bransteitter et al., 2003). Therefore, further investigation is necessary to clarify the

intermediary effect of AID/APOBEC mediated 5-mC deamination for active DNA

demethylation.

1.2.4.3. TET protein mediated 5-mC oxidation in passive and active demethylation

As described above, several mechanisms for DNA demethylation have been

postulated in the past, although a consensus mechanism has not been evident. This may

be because certain mechanisms are restricted to particular biological settings, or that

conflicting results are genuinely unreliable (Wu and Zhang, 2014, 2010). However, the

recent discovery of 5-hydroxymethylcytosine (5-hmC) as a proposed intermediate for DNA

demethylation, has captivated attention (Pastor et al., 2013).

In 2009, two laboratories independently demonstrated the presence of 5-hmC in

the mammalian genome (Kriaucionis and Heintz, 2009; Tahiliani et al., 2009). The

presence of an “unusual” nucleotide in mammals was first observed in mouse Purkinje

and granule cell nuclei by thin layer chromatography (found in ~0.6% and ~0.2% of all CG

dinucleotides, respectively), and was confirmed to be 5-hmC, the oxidative derivative of

5-mC, by mass spectrometry (Kriaucionis and Heintz, 2009). Nucleotide oxidation was

previously characterised in Trypanasome brucei, where J-binding proteins (JBPs)

27

Page 28: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

catalyse thymine oxidation into 5-hydroxymethyluracil (Borst and Sabatini, 2008). In a

computational screen for mammalian homologs of JBPs, Tahiliani et al. identified three

paralogous human proteins TET1, TET2 and TET3 (Tahiliani et al., 2009). The authors

further demonstrated that TET1 overexpression resulted in decreased genomic 5-mC

levels, and that this was due to TET-mediated 5-mC conversion into 5-hmC (Tahiliani et

al., 2009). Similarly, mouse TET proteins (TET1-3) were shown to catalyse 5-hmC

production (Ito et al., 2010). Following studies reported that TET proteins can further

oxidize 5-hmC to yield 5-formylcytosine (5-fC) and 5-carboxycytosine (5-caC) (~15 5-fC

and ~3 5-caC for every 103 5-hmC) (He et al., 2011; Ito et al., 2011; Pfaffeneder et al.,

2011). Structurally, TET proteins contain CXXC-type zinc finger domain that has affinity

for clustered CpG sites, and a carboxy-terminal catalytic domain that provides Fe(II) and

2-oxoglutarate-dependent dioxygenase activity (Iyer et al., 2009). In jawed vertebrates

TET2 underwent chromosomal gene inversion, which resulted in detachment of the exon

containing the CXXC domain that became an independent gene encoding IDAX (inhibition

of the Dvl and axin complex) protein (Iyer et al., 2009; Ko et al., 2013). Interestingly, TET

proteins can target methylated and hemimethylated DNA in a CpG or non-CpG context

for catalytic oxidation (Ficz et al., 2011; Tahiliani et al., 2009)

A broad analysis in mouse tissues revealed that 0.03% to 0.69% of cytosines are

hydroxymethylated, with highest levels in the brain tissues (Globisch et al., 2010).

Similarly, pluripotent embryonic stem cells contain significant levels of 5-hmC, attributed

to the presence of TET1 and TET2 (Ito et al., 2010; Koh et al., 2011; Tahiliani et al., 2009).

Upon retinoic acid-mediated ES cell differentiation, Tet1 and Tet2 genes are

downregulated, accompanied by the upregulation of Tet3, suggesting differential and

developmental regulation of Tet genes (Koh et al., 2011). Indeed, during pre-implantation

development while Tet3 expression is limited to the zygote, Tet1 and Tet2 expression

programme is initiated in two cell embryos (Wossidlo et al., 2011) and maintained in the

inner cell mass of the blastocysts (Ito et al., 2010). Moreover, Tet1 and Tet2 are expressed

in the PGCs, whereas Tet3 is found particularly in the somatic cells during PGC

development (Yamaguchi et al., 2012). This dynamic control can in part be explained by

the presence of large number of pluripotency-associated transcription factor binding sites

at the Tet promoter regions (Ficz et al., 2011) and can be exemplified by OCT4-SOX2

complex-mediated regulation of Tet1 and Tet2 expression in mouse ES cells (Koh et al.,

2011).

To elucidate the relationship between pluripotency and 5-mC oxidation, several

studies have focused on genome-wide mapping of 5-hmC and TET binding in mouse ES

28

Page 29: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

cells. Due to the presence of CXXC domain, TET1 preferentially associates with CpG

islands found at the promoter regions of actively transcribed genes, co-localising with

H3K4me3, yet no enrichment of 5-hmC is observed at these loci (Williams et al., 2011;

Wu et al., 2011; Xu et al., 2011). Confirming this, TET1 depletion results in 5-mC

accumulation at many TET1-associated regions (Wu et al., 2011; Xu et al., 2011).

Interestingly, both TET1 and 5-hmC co-localise at Polycomb repressed CpG-rich bivalent

gene promoters, and this is proposed to protect these regions from the acquisition of de

novo DNA methylation, providing an additional layer of epigenetic control over lineage

commitment (Pastor et al., 2011; Williams et al., 2011; Wu et al., 2011). It should be noted

that, TET1 depletion results in similar levels of upregulated and downregulated genes (Wu

et al., 2011). This suggests that in addition to its catalytic role, TET1-mediated gene

regulation may take place in a catalytic activity-independent manner. In this regard TET1

was demonstrated to recruit SIN3A (Deplus et al., 2013; Vella et al., 2013; Williams et al.,

2011) and MBD3-NURD repressor complexes (Yildirim et al., 2011). In addition, recent

studies have documented TET protein interaction with O-linked N-acetylglucosamine (O-

GlcNAc) transferase (OGT) enzyme, which catalyses GlcNAc addition to serine/threonine

residues of proteins including histones. It was also revealed that TET mediated OGT

recruitment is associated with positive regulation of gene expression (Balasubramani and

Rao, 2013; Chen et al., 2013b; Deplus et al., 2013; Vella et al., 2013). Furthermore, most

of the active promoters that contain high CpG density are devoid of 5-hmC, 5-fC and 5-

caC modifications (Shen et al., 2013; Wu et al., 2011; Yu et al., 2012). Enrichment of TET1

in these regions as well as in active distal enhancers suggests that TET1 might be acting

as a safe-guard machinery to remove any randomly occurring de novo methylation (Shen

et al., 2013). On the other hand, 5-hmC is mainly found at the promoters that contain lower

CpG densities, bearing lower transcriptional activity (Yu et al., 2012). Similarly, 5-hmC is

enriched at poised enhancers, which may require rapid DNA demethylation during lineage

specification (Shen et al., 2013). On the contrary, a strong link exists between gene body

enrichment of 5-hmC and active transcription. TET2 protein, which lacks the CXXC

domain, has been shown to be responsible for 5-hmC deposition at these regions of the

active genes (Huang et al., 2014), as confirmed by TET2 depletion (Chen et al., 2013b).

Oxidation of 5-mC has also been linked to regulation of gene expression,

independent of DNA demethylation. DNA methylation is recognised by methyl binding

domain proteins, which recruit chromatin remodellers to prevent gene expression (Klose

and Bird, 2006). Oxidation of 5-methylcytosine strongly inhibits association of MBDs

(including MBD1, MBD2 and MBD4) with the DNA that in turn may positively affect

29

Page 30: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

transcription (Jin et al., 2010; Valinluck et al., 2004). Interestingly, although methyl-CpG-

binding protein 2 (MeCP2) associates with 5-mC to exhibit repressive functions, it can

bind 5-hmC with similar high affinity and this association facilitates transcription in neurons

(Mellén et al., 2012). On the other hand, in-vitro binding assays revealed that MBD3 is

able to associate with 5-hmC but not 5-mC, which may lead to recruitment of repressive

NURD complex for negative regulation of transcription (Yildirim et al., 2011). However this

finding was not confirmed in a later report, where readers of 5-mC and its oxidized

derivatives were screened by unbiased quantitative mass spectroscopy (Spruijt et al.,

2013). It was further demonstrated that there is a limited overlap between readers (which

are expressed in a cell type specific manner) of each of the 5-mC, 5-hmC, 5-fC and 5-

caC, suggesting individual and cell specific roles of such epigenetic modifications (Spruijt

et al., 2013). Therefore, although a general mechanism has not yet been elucidated,

growing evidence indicates that relative stability of 5-hmC in the genome can influence

gene expression.

5-hmC and passive demethylation

The major excitement of 5-mC oxidation stems from the fact that 5-hmC can act as

an intermediate of DNA demethylation, involved in both DNA replication-dependent and

independent pathways (Figures 1.1 and 1.2). DNA methylation is stably inherited via

collaborative action of DNMT1 with its partner UHRF1 (as described earlier). However

modifications on the methylated cytosine may interfere with the maintenance machinery.

A previous study demonstrated that oxidation of 5-mC prevents methylation of the newly

synthesised DNA, resulting in heritable changes in DNA methylation patterns (Valinluck

and Sowers, 2007). However, whether this is due to the inability of UHRF1 in recognition

of the hemi-hyrdoxymethylated CpG sites has been controversial. While Frauer et al.

reported that UHRF1 can bind to hemi-methylated and hemi-hydroxymethylated DNA with

similar affinities, Hashimoto et al. demonstrated that the efficiency is tenfold less in the

latter case (Frauer et al., 2011; Hashimoto et al., 2012). Nevertheless, DNMT1 is

significantly inefficient in catalytic methylation of the unmodified cytosine in hemi-

hydroxymethlated DNA (60-fold less compared to hemimethylated DNA) (Hashimoto et

al., 2012; Otani et al., 2013). This shows that 5-mC oxidation coupled passive dilution can

occur even in the presence of maintenance machinery. It should be noted that in cells

where TET activity is high (such as neurons and ES cells), 5-mC and 5-hmC levels may

30

Page 31: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

be necessary to be stably maintained. In this regard, it has been shown that de novo

methyltransferases DNMT3A and 3B can efficiently methylate hemi-methylated and hemi-

hydroxymethylated CpG sites (Hashimoto et al., 2012; Otani et al., 2013).

5-hmC and active demethylation

Oxidative derivatives of 5-mC have recently been suggested to be the

intermediates of replication-independent DNA demethylation. Several enzymatic

pathways have been proposed for this process, with a particular emphasis on BER.

The first DNA-repair based pathway that has been proposed is the one involving

AID/APOBEC family members (Figure 1.2). Guo et al. demonstrated that 5-hmC levels

induced by TET1 in HEK293 cells were significantly reduced upon AID over-expression.

In addition, AID and some of the APOBEC family members were able to demethylate

reporter DNA in the presence of TET1. According to the authors, demethylation pathway

first involves TET mediated hydroxylation of 5-mC that is followed by deamination to yield

5-hydroxymethyluracil (5-hmU). The resulting 5-hmU:G mismatch is then repaired by BER

components and 5-hmU is replaced by unmodified cytosine (Guo et al., 2011). In line with

this report, Cortellino et al. demonstrated that SMUG1 (single-strand-selective

monofunctional uracil DNA glycosylase 1) and TDG glycosylases interact with AID and

perform in-vitro 5-hmU removal on double stranded DNA (Cortellino et al., 2011). However

subsequent studies showed conflicting evidence on AID/APOBEC-mediated 5-hmU

generation. Firstly, no enrichment of 5-hmU was detected in mammalian cells using

sensitive mass spectroscopy, suggesting that either this is a very short-lived nucleotide

base or deamination of 5-hmC does not occur (Globisch et al., 2010; Pfaffeneder et al.,

2011). Secondly, although AID/APOBEC members have significantly reduced activity

towards 5-mC, no activity for 5-hmC was detected neither in vitro nor upon overexpression

in cells (Nabel et al., 2012). Mechanistically, AID does not show affinity on 5-hmC due to

steric hindrance and increased electron cloud size (Rangam et al., 2012). Therefore,

involvement of AID/APOBEC deaminase activity in 5-hmC associated DNA demethylation

seems unlikely to occur.

The second DNA repair-mediated pathway involved in active DNA demethylation

is based on TET ability to perform iterative 5-mC oxidation that results in the formation of

5-fC and 5-caC (Figure 1.2) (He et al., 2011; Ito et al., 2011; Pfaffeneder et al., 2011). 5-

fC and 5-caC can be removed by TDG glycosylase activity and this will result in DNA

31

Page 32: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

demethylation upon incorporation of unmodified cytosine by the BER components (He et

al., 2011; Maiti and Drohat, 2011; Zhang et al., 2012). Compared to 5-hmC, TDG exhibits

great affinity for 5-fC and 5-caC (paired with guanine in double stranded DNA) (Ito et al.,

2011; Maiti and Drohat, 2011; Zhang et al., 2012). Supporting this, TDG deficiency causes

significantly increased levels of 5-caC in mouse ES cells (He et al., 2011; Shen et al.,

2013; Song et al., 2013). This collaborative action has been further demonstrated in germ

cell development (Nettersheim et al., 2013), in embryonic stem cells (Okashita et al., 2014)

and in somatic cell reprogramming (Hu et al., 2014). Therefore, TDG stands out as an

essential component in active DNA demethylation and epigenetics not only because it

interacts with numerous transcription factors and chromatin remodellers (Dalton and

Bellacosa, 2012) but also because its deletion in mice is embryonically lethal (Cortázar et

al., 2011; Cortellino et al., 2011). An alternative hypothesis is the direct removal of

carboxyl group from 5-caC. Although chemically feasible, such decarboxylase has not

been found yet (Schiesser et al., 2012).

Figure 1.2. Mechanisms of dynamic modifications of cytosine. 5-methylcytosine (5-mC) generated from cytosine by DNA methyltransferases (DNMTs), can further be oxidized by TET proteins to yield 5-hydroxymethylcytosine (5-hmC), 5-formylcytosine (5-fC) and 5-carboxylcytosine (5-caC). 5-fC and 5-caC can be excised by thymine DNA glycosylase (TDG) and replaced by an unmodified cytosine via base excision repair (BER) pathway. Alternatively, 5-mC and 5-hmC deamination by activation-induced deaminase (AID) and/or apolipoprotein B mRNA editing enzyme catalytic popylpeptide (APOBEC) proteins have been proposed yield thymine and 5-hydroxymethyluracil for DNA demethylation after excision by TDG and SMUG1 (single-strand-selective monofunctional uracil DNA glycosylase) followed by BER. However this model has been controversial (see the main text).

32

Page 33: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

1.2.4.4. TET-associated DNA demethylation dynamics in embryonic development and pluripotency

Following the discovery of TET-mediated 5-mC oxidation, the mechanisms of

genome-wide DNA demethylation occurring in pre-implantation embryos and during germ

cell development have been re-visited.

In the zygote

The first wave of demethylation occurs soon after fertilization allowing the

activation of the transcriptional programme in the totipotent zygote. Initial reports

demonstrated that the paternal pronucleus undergoes rapid loss of 5-mC before the onset

of DNA replication, whereas DNA methylation at the maternal pronucleus is gradually lost

during cleavage stages (Mayer et al., 2000; Oswald et al., 2000; Santos et al., 2002).

Notably, however, later analyses revealed that the observed paternal loss of methylation

signal was actually due to formation of 5-hmC as well as 5-fC and 5-caC, mediated by

TET3 protein (Gu et al., 2011; Inoue and Zhang, 2011; Iqbal et al., 2011; Wossidlo et al.,

2011). 5-hmC begins to be significantly accumulated in the paternal genome at around

pronuclear stage PN3 of the zygote, which coincides with the loss of cytosine methylation

signal, whereas the maternal genome is protected from 5-mC oxidation (Wossidlo et al.,

2011). At this stage, while Tet1 and Tet2 genes are silenced, Tet3, which is transcribed in

the oocyte, is present at high levels (Iqbal et al., 2011; Wossidlo et al., 2011). Importantly,

zygotic RNA interference-mediated silencing of Tet3 prevents 5-hmC generation

accompanied by increased 5-mC signal in the paternal pronuclei (Wossidlo et al., 2011).

Confirming this finding, generation of 5-hmC was not detected in the zygote upon maternal

deletion of Tet3 and blocked demethylation of paternal pluripotency-associated genes and

prevented subsequent gene activation (Gu et al., 2011). Furthermore, maternal Tet3

mutant embryos are severely degenerated and possess morphological abnormalities, and

homozygous Tet3 deletion results in prenatal lethality (Gu et al., 2011). It is important to

note that Tet3 is specifically associated with paternal pronuclei, indicating that it is actively

excluded from the maternal genome (Gu et al., 2011). Later it was shown that protection

of the maternal genome from oxidation is mediated by STELLA, which prevents TET3

binding by possibly altering chromatin configuration (Nakamura et al., 2012). STELLA

preferentially binds maternal chromatin due to its association with H3K9me3 that is

predominantly in the maternal pronucleus. From this point, replication-dependent

33

Page 34: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

demethylation of the paternal DNA occurs as evidenced by gradually decreased levels of

5-hmC during cleavage stages (Inoue and Zhang, 2011). This is further supported by the

limited affinity shown by DNMT1 for hemi-hydroxymethylated DNA and its exclusion of

maintenance machinery (Hashimoto et al., 2012; Hirasawa et al., 2008). Furthermore,

iterative oxidation products (5-fC and 5-caC) of 5-mC that are also present in the paternal

pronucleus can contribute to passive demethylation (Inoue et al., 2011). It is worth

mentioning that biological significance of genome-wide paternal 5-mC oxidation is not well

elucidated and TET3 deficiency does not affect global transcription levels in the embryo

(Inoue et al., 2012).

In the Primordial Germ Cells

The second wave of global DNA demethylation takes place during primordial germ

cell development and migration to the genital ridges, which is essential for the

establishment of totipotency in the next generation. Loss of DNA methylation in PGCs

follows a strict bi-phasic temporal order; demethylation first starts at the pluripotency

associated and germ cell specific genes, and only after involves ICRs, LINE repeats and

X-chromosome linked loci once PGCs settle in the gonads (Guibert et al., 2012;

Seisenberger et al., 2012). The bulk of the DNA demethylation occurs in the first phase

between E6.5 to E9.5, where overall CpG methylation level decreases from ~70% to 30%

(Seisenberger et al., 2012). It affects promoters of the genes that are expressed early in

PGC development, including transcription factors of pluripotency network genes such as

Nanog, while DMRs of the imprinted genes and X-linked CGIs maintain methylation at this

stage (Seisenberger et al., 2012). De novo methyltransferases are silenced in PGCs, while

DNMT1 is present and is localised in the nucleus (Hajkova et al., 2002; Kurimoto et al.,

2008). However DNMT1 tethering factor UHRF1 is significantly downregulated and the

remaining protein is cytoplasmic (Kurimoto et al., 2008; Seisenberger et al., 2012). This

suggests that genome-wide DNA demethylation may occur passively due to largely

impaired maintenance machinery, which may specifically protect some loci (including

DMRs of imprinted genes) from loss of methylation during this phase (Seisenberger et al.,

2012).

The second phase of DNA de-methylation further decreases overall CpG

methylation to 14% in male and 7% in female PGCs at E13.5 (Seisenberger et al., 2012)

and it seems to involve TET1 and TET2 mediated 5-mC oxidation (Hackett et al., 2013;

Yamaguchi et al., 2013a). Expression of Tet1 and Tet2 genes reaches maximum levels

34

Page 35: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

between E9.5 and E10.5, and is accompanied by substantial amount of 5-hmC generation

in the PGC nuclei and increases progressively until 5-mC is no more detectable at E11.5

(Hackett et al., 2013; Yamaguchi et al., 2012, 2013a). Genome-wide mapping

demonstrated that 5-hmC is accumulated at several late demethylated loci, including

DMRs of the imprinted genes (Hackett et al., 2013). Interestingly unlike ES and somatic

cells, 5-hmC is also enriched at heterochromatic satellite repeats in PGCs (Hackett et al.,

2013). From E11.5 onward, 5-hmC signal is progressively lost in the PGC nuclei as

evidenced by time-course immunostaining analysis. As PGCs undergo cell division at

every 12.6 hours at this stage (Kagiwada et al., 2013), this gradual loss is suggested to

occur by a replication-coupled mechanism where unmodified cytosine is incorporated

during DNA synthesis (Hackett et al., 2013). It is worth mentioning that, BER pathway

components including Tdg are expressed in PGCs, (but not Aid or Apobec1, calling into

question their contribution in loss of methylation as previously proposed (Popp et al.,

2010)), suggesting a potential role for these factors in DNA demethylation (Hajkova et al.,

2002; Kagiwada et al., 2013). The bi-phasic character of DNA demethylation and

contribution of 5-mC oxidation during germ cell development was further confirmed in-vitro

by ES cell into PGC differentiation (Vincent et al., 2013). This study revealed that TET

proteins are dispensable for the first phase, as evidenced by the presence of global

demethylation that occurred in TET1 and TET2 deficiency. On the other hand, the second

phase of locus-specific loss of methylation is TET-dependent (Vincent et al., 2013).

Tet1 depletion interferes with demethylation of meiosis related gene promoter and

transcriptional activation in female PGCs. This results in loss of oocytes, decreased

fertilization and smaller litters (Yamaguchi et al., 2012). Although male Tet1 deficiency

does not interfere with testes development, E13.5 PGCs exhibit locus specific

hypermethylation, mainly observed in late-demethylated group of imprinted genes

(Yamaguchi et al., 2013a). In the progeny derived from crossing Tet1 knockout males with

wild type females, aberrant hypermethylation is observed at the DMRs of imprinted genes

including Peg3 and Peg10. Dysregulation of imprinting is in part responsible for various

phenotypes observed in the new heterozygous generation, such as placental, foetal and

post-natal defects as well as embryonic lethality (Yamaguchi et al., 2013a). Tet1/Tet2

double knockout PGCs are completely devoid of 5-hmC and the progenies display

compromised imprinting, supporting the role of Tet activity in erasure of paternal imprints

(Dawlaty et al., 2013). The majority of Tet1/Tet2-null mice die within the first two days of

birth; however those who survive to adulthood are fertile, suggesting a compensatory role

35

Page 36: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

of TET3. Indeed, double knockout male sperm contains almost normal levels of 5-hmC

(Dawlaty et al., 2013).

In embryonic stem cells

Since the discovery of relatively high levels of 5-hmC in ES cells, studies have

focused on the roles of TET proteins in ES cell-specific properties that are self-renewal

and pluripotency. Initially, shRNA-mediated Tet1 knockdown in mouse ES cells was

shown to result in Nanog downregulation and defects in stem cell maintenance (Ito et al.,

2010). Furthermore, acute depletion of Tet1 resulted in impaired LIF/Stat3 signalling with

adverse effects on ES cell identity (Freudenberg et al., 2012). In contrast, several other

studies reported that Tet1 knockdown does not affect the expression of key pluripotency-

associated factors including Nanog. The discrepancy between different studies may stem

from culture conditions, use of different cell lines and RNA interference based off-target

effects (Ficz et al., 2011; Koh et al., 2011; Williams et al., 2011). More evidence was

provided upon genetic deletion of Tet1, where knockout ES cells exhibit normal

undifferentiated ESC morphology and express pluripotency markers (Dawlaty et al.,

2011). To rule out possible redundant functions and compensation by TET2 in chronic

Tet1 depletion, Dawlaty et al. further deleted Tet2 gene and generated double Tet1/Tet2

knockout ES cells (Dawlaty et al., 2013). This resulted in complete depletion of 5-hmC

signal, while the cells remained pluripotent, as evidenced by their capacity to generate

tissues from all embryonic layers by teratoma assay (Dawlaty et al., 2013). Interestingly,

teratomas formed from Tet1 single and Tet1/Tet2 double knockout ES cells are

haemorrhagic and enriched for throphoblast cells, suggesting a skewed differentiation

potential toward extraembryonic tissues. It should be noted that these cells can efficiently

contribute to the chimaeras upon blastocyst injection, indicating that differentiation

problems can be overcome during embryonic development (Dawlaty et al., 2013).

Similarly, ES cells deficient of all three members of TET proteins maintain normal

morphology and express pluripotency markers, however their differentiation potential is

severely restricted, as reported by lack of endodermal and mesodermal structures in

teratomas (Dawlaty et al., 2014). Moreover, triple knockout ES cells contribute poorly to

developing embryos upon blastocyst injection, probably due to hypermethylation of

promoter regions of several developmental genes. Confirming the importance of TET

proteins in this process, ectopic Tet1 expression in TET-null ES cells rescues their

differentiation defects and reinstates chimeric contribution. (Dawlaty et al., 2014). Overall

36

Page 37: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

this data demonstrate partial redundancy between TET proteins and their critical roles in

proper differentiation.

TET-assisted DNA demethylation has recently been shown to contribute to the

acquisition of an ICM-like naïve state in mouse ES cells. Traditionally mouse ES cells are

isolated and grown in serum supplemented with leukaemia inhibitory factor (LIF), however

they exhibit heterogeneity with two distinguishable populations. These are primed

epiblast-like and naïve ICM-like states, as exemplified by biphasic Nanog expression

(Chambers et al., 2007). Interestingly ICM-like ground state homogeneity can be

established by growing mouse ES cells in media supplemented with two kinase inhibitors

(2i) (Ying et al., 2008) and the epigenetic basis of this naïve status has recently begun to

be elucidated. For example, ES cells isolated in the presence of 2i display genome-wide

hypomethylation that span CGIs, LINE1 elements, as well as minor and major satellite

sequences, conversely to what observed in serum supplemented media (Leitch et al.,

2013). Similarly, 2i addition to serum grown ES cells results in a rapid shift of methylation

status, accompanied by downregulation of Dnmt3a, Dnmt3b and Dnmt3l gene expression

that is partly mediated by PRDM14 (Leitch et al., 2013). Consequently, genome-wide 5-

hmC accumulation catalysed by TET1/TET2, coupled with impaired methylation

maintenance and disabled de novo methyltransferase activity, results in global replication-

coupled passive loss of methylation in 2i condition (Ficz et al., 2013; Habibi et al., 2013).

In addition, global methylation data sets confirmed that ES cells grown in 2i show

similarities to ICM cells at E3.5 or migratory PGCs at E9.5 (Ficz et al., 2013; Habibi et al.,

2013), reminiscent of TET activity during embryonic development.

Interestingly, in 2i condition further demethylation can occur in the presence of a

rather unexpected supplement, vitamin C. Actually, vitamin C is a necessary ingredient of

human embryonic stem cell media and activates genes implicated in growth, proliferation

and pluripotency by regulating methylation dynamics at their promoters (Chung et al.,

2010). Interestingly, mouse embryonic fibroblasts supplemented with vitamin C undergo

significant 5-mC oxidation, while TET levels remain stable. This is due to enhanced

enzymatic TET activity upon vitamin C association with TET catalytic domains that results

in structural change and enhanced product recycling (Minor et al., 2013). In mouse ES

cells where Tet1 and Tet2 genes are active, vitamin C supplementation leads to 40%

decrease in genome-wide 5-mC levels due to 5-hmC generation, affecting previously

methylated gene promoters and results in upregulation of germline genes (Blaschke et al.,

2013; Yin et al., 2013). Yet, imprinted genes and intracisternal A particle retroelements

37

Page 38: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

resist demethylation induced by vitamin C, reminiscent of pre-implantation embryo

development (Blaschke et al., 2013).

Figure 1.3. TET-induced demethylation in mouse embryonic development and ES cells. TET proteins are involved in both genome-wide demethylation waves that occur in the pre-implantation embryo and in primordial germ cells (PGCs). TET-assisted demethylation is also observed in mouse ES cells cultured in 2i media and in the presence of Vitamin C. TET involvement in the transitions is highlighted as surf board and DNA methylation dynamics are indicated. The methylated cytosine is shown in red, unmodified cytosine in grey. Adapted and modified from (Bagci and Fisher, 2013).

1.3. Reprogramming cell fate

Early mammalian development involves sequential cell fate decisions

accompanied by specialization and progressive restriction of potential, giving rise to the

germ lineage and over two hundred different types of somatic cells. It has long been

thought that cell differentiation is an irreversible process, with a few exceptional

pathological cases such as metaplasia and malignancy. However, decades of research

have now challenged this view, unravelling the hidden potential of differentiated cells.

38

Page 39: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

1.3.1. Transdifferentiation Transdifferentiation is a process of direct switch of cell fate, which transforms one

cell type into another without passing through a pluripotent state. A very important

example of this switch was demonstrated by treating immortalized fibroblasts with 5-

azacytidine (inhibitor of DNA methylation), which resulted in their conversion into

adipocytes and chondrocytes (Taylor and Jones, 1979). This suggested that lineage-

specific transcription factors are tightly controlled by DNA methylation. It was later

demonstrated that, a single factor, MyoD, can convert fibroblasts into myocytes via ectopic

expression (Davis et al., 1987). The blood system has been an informative platform in

terms of cell conversion analyses. For example, high level expression of the erythroid-

megakaryocyte-affiliated transcription factor GATA1 in monocytes (macrophage

precursors) does not only downregulate monocyte-associated factors, but also activates

erythroid-megakaryocyte gene expression (Kulessa et al., 1995; Visvader et al., 1992).

Interestingly, conversion towards the opposite direction can also occur; expression of

PU.1 in erythroid-megakaryocytic cells promotes their switch to monocytes by repressing

GATA1 (Nerlov and Graf, 1998). Furthermore, the granulocyte/macrophage-specific

transcription factor C/EBPα can very efficiently convert primary B-cell progenitors into

macrophages, or to a lesser extent immunoglobulin-producing B lymphocytes (Xie et al.,

2004). Identification of factors that can modulate cellular epigenome for fate switching has

implications in regenerative therapy. For example, pancreatic β-cells are key targets for

treatment of diabetes, and a transcription factor screen in pancreatic tissue revealed three

important genes: Pdx1, Ngn3 and MafA for their normal function. In vivo adenoviral

delivery of these factors for transient expression in fully differentiated pancreatic exocrine

cells resulted in conversion into β-cells with an efficiency of 20% (Zhou et al., 2008). A

similar candidate-based approach was used to identify Gata4, Mef2c and Tbx5 genes that

can directly convert mouse cardiac and dermal fibroblasts into cardiomyocyte-like cells

(Ieda et al., 2010). Additional examples include Ascl1-mediated fibroblast-to-neuron

conversion (Vierbuchen et al., 2010) and fibroblast-to hepatocyte conversion by defined

factors (Huang et al., 2011; Sekiya and Suzuki, 2011).

These examples have demonstrated powerful trans-acting functions of

transcription factors. This is especially interesting considering the presence of higher order

chromatin organization that packs the DNA to repress gene expression and to prevent

accessibility of activating factors. Although cell division has been proposed to allow

transcription factor binding due to unwinding of the chromatin during DNA replication, in

39

Page 40: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

some cases cell fate can rapidly be converted in the absence of cell proliferation. This is

proposed to be accomplished by ‘pioneer transcription factors’ which can target repressive

lineage-specific regulatory sites, to create a permissive environment for accessibility and

binding of additional activators (Zaret and Carroll, 2011). It is important to note that

transdifferentiation efficiency depends on developmental proximity between two cell types

and degree of their lineage specificity. Cell fate conversion between different germ layers

has been more challenging due to genome-wide differences in epigenetic status.

Therefore, a distant switch may require additional transcription factors and epigenetic

modulators (Graf and Enver, 2009).

1.3.2. Pluripotent conversion of somatic cells

Cellular differentiation is accompanied by the acquisition of lineage-specific

transcription programme. Although mechanisms of epigenetic memory ensures

conservation of fate during proliferation, cells remain surprisingly plastic. This does not

only render the process of transdifferentiation possible as described earlier, but also

allows pluripotent conversion of differentiated cells. Once a pluripotent state is acquired,

it is possible to generate cells from different germ lines, which has significant therapeutic

implications. Cellular plasticity was demonstrated by early studies of nuclear transfer and

cell fusion, and recently by factor mediated induction of pluripotency. It is crucial to

elucidate molecular details of pluripotent conversion, to better understand epigenetic

mechanisms that govern cellular differentiation and establishment of pluripotency.

1.3.2.1. Nuclear transfer

Once transplanted into an enucleated oocyte, the nucleus of a differentiated cell

undergoes extensive nuclear reprogramming that can lead to the generation of an entire

organism, a process also known as cloning. Pioneering work was conducted in

amphibians, where blastocyst cell implantation into enucleated oocyte resulted in clones

of swimming tadpoles (Briggs and King, 1952). Although the efficiency was lower when

more differentiated intestinal epithelium cell nuclei were transferred (~1% compared with

~30% in the case of blastocyst cells), adult frog clones could still be obtained (Gurdon,

1962a, 1962b). These initial results proved the plasticity and reversibility of differentiated

nuclear state, and revealed that all the genetic information to generate an entire individual

40

Page 41: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

Figure1.4. In-vitro strategies for nuclear reprogramming to pluripotency. Nuclear transfer. This approach involves the injection of somatic cell nucleus into an enucleated oocyte. Factors in the oocyte cytoplasm act on the somatic nucleus and reprogram it to a pluripotent state. This oocyte can then be further cultured to obtain genetically matched nuclear transfer embryonic stem (ntES) cells or to obtain a clone animal upon implantation into a surrogate mother. Induction of pluripotency factors. Ectopic expression of pluripotency-associated genes in a somatic cell can initiate reprogramming, and resulting “induced pluripotent stem cells” (IPSCs), which have embryonic stem cell-like properties can be stably propagated in culture. Cell fusion. In this approach, a somatic cell can be reprogrammed to pluripotency after fusion with a pluripotent cell type such as embryonic stem or embryonic germ cell. The fusion first yields a heterokaryon where parental nuclei are separated but share the same cytoplasm. The merging of the nuclei results in the formation of the pluripotent hybrid, which, in the case of an intraspecies fusion, can stably be cultured.

is stably kept in the specialized cell nucleus. The same approach in mammalians gave

rise to the first cloned sheep (named after Dolly Parton) (Wilmut et al., 1997), and a year

later led to cloning of mice (Wakayama et al., 1998). Mice clones could further be

generated by somatic-cell nuclear transfer (SCNT), using mature B and T cells as well as

olfactory sensory neurons as donors (Eggan et al., 2004; Hochedlinger and Jaenisch,

2002). It was notably, however, that reprogramming of terminally differentiated cell was

41

Page 42: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

very inefficient and a two-step model was used. In this system, nuclear transplantation of

the somatic cell gives rise to the blastocysts, from which nuclear-transfer-embryonic stem

(ntES) cells are derived and retransplanted into enucleated oocytes. It should be noted

that cloned animals display common abnormalities that involve premature death (Thuan

et al., 2010). That is most probably due to a defect in proper erasure of epigenetic memory

that has been imposed during differentiation by DNA methylation and histone

modifications (Simonsson and Gurdon, 2004). Supporting this, inhibition of histone

deacetylases improved the efficiency of cloning, possibly by promoting histone acetylation

that leads to structural changes in the chromatin and enhanced DNA demethylation

(Kishigami et al., 2006). In the past decade, a wide range of species have been cloned

(Thuan et al., 2010), and ntES cells have been successfully derived from non-human

primates (Byrne et al., 2007). Although several attempts to isolate human ntES cells failed

in the past (Egli et al., 2011; Noggle et al., 2011), a recent study demonstrated their

isolation by using high quality enuclated human oocytes as recipients and human

fibroblast nuclei as donors (Tachibana et al., 2013). This has important implications in

derivation of patient-specific pluripotent cells for developing specific therapies.

1.3.2.2. Induced pluripotent stem cells

In 2006, Takahashi and Yamanaka reported the derivation of induced pluripotent

stem (iPS) cells from mouse embryonic fibroblasts (MEFs) and skin cells (Takahashi and

Yamanaka, 2006). They conducted a systematic candidate-based interrogation of ES cell-

associated genes for their potential to induce pluripotency in adult cells and identified four

factors: OCT4, SOX2, KLF4 and c-MYC (OSKM). This breakthrough in reprogramming

was quickly embraced and iPS cells were derived upon ectopic expression of OSKM from

human fibroblasts and from a plethora of differentiated cells in subsequent studies

(Yamanaka, 2009). iPS cells form compact ES cell-like colonies with distinct borders, can

propagate in culture indefinitely and possess the capacity to form each of the three

embryonic germ layers (Hanna et al., 2010). It is important to note that the differentiation

potential of iPS cells, which would be a prerequisite for therapy, is influenced by the

parental-origin. For example, monocyte-derived iPS cells can differentiate more efficiently

along the haematopoietic lineages compared to fibroblast-derived iPS cells. The fact that

their potential can be enhanced by HDAC inhibitors and 5-azacytidine indicates a defect

in erasure of epigenetic memory during reprogramming (Kim et al., 2010). Somatic cell

reprogramming is an inefficient and lengthy process (Hochedlinger and Plath, 2009), and

42

Page 43: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

has been suggested to be composed of three phases: initiation (marked by mesenchymal-

to-epithelial (MET) transition), maturation and stabilization (marked by expression of

genes associated with embryonic development and maintenance) (Samavarchi-Tehrani

et al., 2010). OSKM activation first induces stochastic regulation of genes involved in

pluripotency, in MET, in proliferation and in metabolism, and is possibly the rate-limiting

step of reprogramming. Those cells with the appropriate combination of gene expression

can then enter a hierarchical phase, in which pluripotency network is activated leading to

a decrease in intercellular variation (Buganim et al., 2012). Transcriptional changes that

are observed during pluripotent conversion are driven by epigenetic alterations in the

nucleus. These involve histone modifications, chromatin reorganization and DNA

demethylation. Remarkably, these events are ordered: although histones are immediately

modified after induction (Koche et al., 2011). DNA demethylation occurs late in the process

and coincides with the acquisition of a stably reprogrammed state (Polo et al., 2012).

OSKM can be in part, or fully replaced by other transcription factors or chemicals for the

induction of pluripotency, yet a wide discrepancy exists between studies, which stems

from the stochasticity and the inefficiency of the reprogramming process (Theunissen and

Jaenisch, 2014).

1.3.2.3. Cell fusion

Cell fusion experiments were initiated in 1960s. Since then this methodology has

provided valuable information on molecular details of nuclear-cytoplasmic interaction,

dominant effects of trans-acting factors, nuclear plasticity and regulatory mechanisms of

pluripotency. Fusion of two or more cell types creates a single cytoplasmic entity called a

heterokaryon, which harbours both parental nuclei. Heterokaryons are generally transient,

and upon further culture, the nuclei merge and give rise to hybrid cells which undergo cell

division. It should be noted that stability of hybrids depends on the genetic background

compatibility of the fusion partners. While fusion between same species can generate

stably proliferating hybrids, interspecies hybrids are susceptible to chromosomal

abnormalities and loss.

Based on the fact that certain animal viruses can induce the formation of

multinucleated cells, Harris and Watkins used Sendai virus to induce interspecies fusion

between human HeLa and mouse Ehrlich ascites tumour cells (Harris and Watkins, 1965).

Tritiated uridine and leucine labelling experiments demonstrated that both sets of nuclei

could produce RNA in the resulting heterokaryons that also contain normal protein levels.

43

Page 44: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

This suggested that fundamental cellular functions are not altered upon fusion.

Furthermore, although heterokaryons did not undergo cell division, tritiated thymidine

staining revealed the presence of partially synchronised DNA synthesis in unfused nuclei

(Harris and Watkins, 1965). This was further confirmed in a systematic approach, where

a parental cell line undergoing DNA replication was shown to induce premature DNA

synthesis on its fusion partner nucleus (Rao and Johnson, 1970). In the meantime, cell

fusion system provided evidence on regulation of gene expression by trans-acting

elements. This was demonstrated by ceased melanin and tyrosine amino-transferase

synthesis in hamster melanocytes or rat hepatocytes, respectively, after fusion with mouse

fibroblasts, revealing very important initial clues on dominant action of a cell type on its

fusion partner (Davidson et al., 1966; Weiss and Chaplain, 1971). Dominance was further

evaluated in fusion experiments conducted between malignant with non-malignant cells,

and tumorigenicity was found to be suppressed in hybrid cells by the possible trans-acting

functions of anti-oncogenes (Harris and Miller, 1969). Previously silenced genes can also

be activated upon cell fusion, as initially analysed in hybrid cells (Davidson, 1972;

Peterson and Weiss, 1972). One caveat of these early reports arises from the instability

of interspecies hybrids, where chromosomal loss or rearrangement may result in the

observed phenotypes. However, enhanced sensitivity of detection techniques facilitated

the analysis of early events, and led subsequent studies to focus on heterokaryons. For

example, heterokaryons formed between human amniocytes and mouse muscle cells

started to produce human specific muscle proteins that are normally repressed in

amniocytes, as early as 3 days of fusion (Blau et al., 1983). As parental nuclei remain

distinct and retain their chromosomes in heterokaryons, observed gene activation could

be directly attributed to the mouse muscle specific trans-acting factors in the cytoplasm.

Furthermore, muscle cells were able to alter the differentiated state of cells originated from

different embryonic lineages (mesoderm, ectoderm and endoderm), revealing the nuclear

plasticity of diverse cell types (Blau et al., 1985). It is important to note that frequency and

kinetics of reprogramming is a cell-type specific phenomenon, for example mesoderm-

derived human cells (which are more closely related to muscle cells) initiated muscle

specific gene expression sooner compared to ectoderm or endoderm derived cells (Blau

et al., 1985). Reprogramming of distant lineages by muscle cells (such as hepatocytes

derived from the endoderm) is dependent on the gene dosage, as lowest reprogramming

efficiency was observed in heterokaryons when hepatocyte nuclei proportion was higher

(Blau et al., 1985). In addition, HeLa cells that could not be reprogrammed by muscle cells,

were only reprogrammed following 5-azacytidine treatment, increasing their

44

Page 45: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

responsiveness for muscle specific trans-acting regulatory factors (Chiu and Blau, 1985).

These reports demonstrated that somatic cells can influence the transcription

programme of their fusion partners, by the action of cell specific master regulators. Early

studies also addressed whether pluripotency can be induced in differentiated cells. For

example, when mouse embryonic carcinoma (EC) cells were fused with mouse

thymocytes, not only the features of the somatic partner were extinguished, but also

pluripotent properties of the parental EC cells (Miller and Ruddle, 1976), including

reactivation of the silent X-chromosome (Takagi, 1993), were acquired in the generated

hybrids. Dominant pluripotent conversion capacity was later extended by Tada and

colleagues, by fusing mouse thymocytes with mouse EG (Tada et al., 1997) or mouse ES

cells (Tada et al., 2001) that resulted in pluripotent hybrids. It is worth mentioning that

these studies revealed an important finding; somatic originated imprinted genes remained

methylated when the fusion partner was ES cells, but were demethylated upon fusion with

EG cells. This resembles to the regulation of imprinted gene methylation in early embryo

and PGC development, from which ES and EG cells are derived (Tada et al., 1997, 2001).

Subsequent studies demonstrated that ES cells from human origin can successfully

reprogram fibroblasts, and resulting stable tetraploid hybrids possess the characteristics

of parental ES cells. They actively express OCT4 gene, whose promoter is

hypomethylated and can differentiate into each embryonic germ layer (Cowan et al.,

2005). Cell fusion-based reprogramming is a fast process as evidenced by rapid activation

of Oct4-GFP transgene embedded in the somatic genome (Do and Schöler, 2004; Han et

al., 2008; Silva et al., 2006; Wong et al., 2008). Remarkably, pluripotent-specific

transcription program is initiated in interspecies heterokaryons as early as 24 hours of

fusion, indicated by induction of OCT4, NANOG, CRIPTO, ESRRB, TLE1 and REX1

genes from the somatic nucleus (Bhutani et al., 2010; Pereira et al., 2008; Soza-Ried and

Fisher, 2012). Regulation of gene expression is accompanied by substantial changes in

the chromatin structure. For instance, somatic DNA in pluripotent heterokaryons and

hybrids acquire histone H3 and H4 acetylation as well as di- and tri-methylated H3K4me3,

all of which are marks of a pluripotent state (Kimura et al., 2004; Piccolo et al., 2011).

Although cell fusion based reprogramming kinetics are similar to nuclear transfer,

the latter method is technically challenging (Yamanaka and Blau, 2010). The speed and

extent of reprogramming by fusion are not surprising considering the effectiveness of the

pluripotency-associated network in maintaining pluripotent identity by regulating gene

expression (Ng and Surani, 2011). This already established powerful network can quickly

act on the somatic nucleus. It is noteworthy that Oct4, Klf4 and Sox2 have recently been

45

Page 46: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

identified as pioneering transcription factors, suggesting that their modulatory functions

are more powerful than previously anticipated (Soufi et al., 2012). This is exemplified by

rapid localisation of the OCT4 protein on the somatic DNA after fusion with ES cells and

reprogramming cannot be achieved in its absence (Pereira et al., 2008). In this regard,

cell fusion model establishes a platform to screen for important factors that drive

pluripotent conversion. For example, fusion of ES cells that overexpress Nanog with

neural stem cells yields at least 200-fold increased number of pluripotent hybrids (Silva et

al., 2006). Similarly Polycomb-group proteins are essential for successful reprogramming

of lymphocytes by ES cells (Pereira et al., 2010). AID has also been demonstrated to be

crucial in pluripotent conversion by actively demethylating OCT4 and NANOG gene

promoters in somatic DNA (Bhutani et al., 2010); however this observation has been

revoked by subsequent studies (Foshay et al., 2012).

46

Page 47: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 1

1.4. Aims of this study

Pluripotent conversion of differentiated cells involves the installation of a

pluripotency-associated transcription program which is mediated by nuclear

reorganisation, chromatin remodelling and resetting of DNA methylation patterns. The

overall aim of this thesis is to elucidate molecular mechanisms of DNA demethylation in

the course of initiation and propagation of successful reprogramming. In this regard, cell-

fusion based reprogramming provides a tractable experimental platform that helps

dissection of early events leading to pluripotent conversion. For this reason, I will generate

experimental heterokaryons and hybrids between pluripotent and somatic cells, and

analyse the epigenetic stages of reprogramming.

Initially, I will focus on the dynamics of DNA demethylation in heterokaryons

(generated between mouse embryonic stem cells and B lymphocytes or fibroblasts) to

understand its role in initiation of pluripotency-associated gene expression. Then I will

investigate the molecular details of mouse embryonic germ cell mediated imprint erasure

in somatic genome upon fusion. I will finally analyse the involvement of TET proteins at

the early stages of successful reprogramming, by using gene knockdown and knockout

approaches. For this, I will explore CRISPR/Cas9-induced gene editing system, which I

will also use as part of a technical chapter, to generate mouse embryonic stem cell lines

deficient of non-canonical Wnt pathway components.

47

Page 48: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

Chapter 2. Materials and Methods

2.1. Materials

2.1.1. Cell lines

Mouse embryonic stem cells

E14tg2A : Feeder-free, hypoxanthine guanine phophoribosyltransferase (HPRT) –deficient mouse

embryonic stem cell line (Hooper et al., 1987). The main mouse embryonic stem cell line used throughout this

thesis unless otherwise specified. Pgk12.1 : Feeder-free female embryonic stem cell line, kindly provided by Professor Neil Brockdorff

(Penny et al., 1996).

Mouse embryonic germ cells

58G : Mouse embryonic germ cell line derived from primordial germ cells of female E12.5 embryos,

kindly provided by Professor Takashi Tada (Tada et al., 1997).

Mouse B lymphocytes

mB : Mouse pre-B cell line derived from transgenic (GOF-18/ΔPE/GFP) mice (Yoshimizu et al.,

1999). These cells were Aberson transformed and contain a Puromycin resistance cassette. The main mouse B

lymphocyte line used throughout this thesis unless otherwise specified. 2rB : Mouse pre-B cell line derived from transgenic mouse generated by mating male GOF-

18/ΔPE/GFP (Yoshimizu et al., 1999) with female Peg1M-β-gal (Lefebvre et al., 1998) mice. These cells were

Aberson transformed and contain a Puromycin resistance cassette.

Human embryonic stem cells

H7 : Human embryonic stem cell line, derived from human blastocysts.

Human B lymphocytes

hB : Epstein-Barr Virus transformed human adult B cell line, kindly provided by Emily R. Eden (Eden

et al., 2002).

Human fibroblasts

IMR90 : Female human fibroblasts (Coriell), immortalised by telomerase reverse transcriptase (pBABE-

hTERT-blastocydin plasmid).

48

Page 49: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

2.1.2. Antibodies

mTet1 : Rabbit Polyclonal anti-Tet1 (Millipore, 09-872). WB at 1:1000

mTet2 : Rabbit Polyclonal anti-Tet2 (Santa Cruz Biotechnology, sc136926). WB at 1:250 mOct4 : Goat Polyclonal anti-Oct-3/4 (Santa Cruz Biotechnology, sc-8628). WB at 1:2000.

mNanog : Rabbit Polyclonal anti-Nanog (Cosmo Bio, REC-RCAB0002P-F). WB at 1:5000

mTubulin : Rabbit Monoclonal anti-α Tubulin (Sigma, T9026). WB at 1:10000

mJarid2 : Rabbit Polyclonal anti-Jarid2 (Abcam, ab48137). WB at 1:750

mLamin B : Goat Polyclonal anti-Lamin B (Santa Cruz Biotechnology, sc-6216). WB at 1:20000

H3 : Rabbit Polyclonal anti-histone 3 (Abcam, ab1791). ChIP, 2µg per IP

H3K4me3 : Rabbit Polyclonal anti-histone 3 trimethyl K4 (Millipore, 07-473). ChIP, 2µg per IP

2.2. Methods

2.2.1. Cell culture

All of the reagents used for tissue culture were supplied from Gibco (Life

Technologies), unless otherwise specified. All of the cells were maintained at 37oC and

5% (v/v) CO2, unless otherwise specified.

Mouse embryonic stem cells were cultured in Knockout Dulbecco’s Modified Eagle

Medium (KO-DMEM), supplemented with 10% (v/v) Fetal Bovine Serum (FBS), 1mM Non-

Essential Amino Acids (NEAA), 2mM L-Glutamine, antibiotics (100U/ml penicillin &

100µg/ml streptomycin), 50µM β-mercaptoethanol and 1000U/ml Leukemia Inhibitory

Factor (LIF) (ESGRO-LIF, Millipore). The cells were maintained on 0.1% gelatin (Sigma)

coated dishes. The cells were routinely passaged at a 1:8 dilution by treatment with 0.05%

Trypsin-EDTA.

Mouse embryonic germ cells were cultured in DMEM/F12, supplemented with 20%

(v/v) FBS, 1mM NEAA, 2mM L-Glutamine, antibiotics (100U/ml penicillin & 100µg/ml

streptomycin), 50µM β-mercaptoethanol, 1mM Sodium Pyruvate, 0.12% Sodium

Bicarbonate, 1% (v/v) Nucleosides (stock solution is prepared in distilled water: 0.08%

(m/v) Adenosine, 0.073% (m/v) Cytidine, 0.024% (m/v) Thymidine, 0.085% (m/v)

Guanosine, 0.073% (m/v) Uridine (Sigma)) and 1000U/ml Leukemia Inhibitory Factor (LIF)

(ESGRO-LIF, Millipore). The cells were maintained on γ-irradiated primary mouse

49

Page 50: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

embryonic fibroblast feeder layers and were routinely passaged at a 1:8 dilution by

treatment with 0.05% Trypsin-EDTA. Embryonic germ cells were treated with Plasmocin

(25µg/ml, Invivogen) for 1 week, and all the experiments were performed between

passages 24 and 28.

Mouse B lymphocytes were cultured in Roswell Park Memorial Institute-1640

(RPMI-1640), supplemented with heat inactivated 20% (v/v) FBS, 1mM NEAA, 2mM L-

Glutamine, antibiotics (100U/ml penicillin & 100µg/ml streptomycin) and 50µM β-

mercaptoethanol.

Human embryonic stem cells were cultured in mouse embryonic fibroblast

conditioned KO-DMEM, supplemented with 20% (v/v) Knockout Serum Replacement

(KSR), 1mM NEAA, 2mM L-Glutamine, antibiotics (100U/ml penicillin & 100µg/ml

streptomycin), 50µM β-mercaptoethanol, 1mM Sodium Pyruvate, 40ng/ml bFGF

(Peprotech) and 5ng/ml Activin A (Peprotech). The cells were maintained on dishes

overnight coated with 0.5mg/mL growth-factor-reduced matrigel (BD Biosciences) and

were routinely passaged by a 1:3 dilution by treatment with 0.02% EDTA.

Human B lymphocytes were cultured in RPMI-1640, supplemented with heat

inactivated 10% FBS, 1mM NEAA, 2mM L-Glutamine and antibiotics (100U/ml penicillin

& 100µg/ml streptomycin).

Human fibroblasts were cultured in DMEM, supplemented with supplemented with

10% FBS, 1mM NEAA, 2mM L-Glutamine and antibiotics (100U/ml penicillin & 100µg/ml

streptomycin). The cells were cultured in 3% (v/v) oxygen and were routinely passaged at

a 1:3 dilution by treatment with 0.05% Trypsin-EDTA.

Cells were frozen for later usage using 10% dimethyl sulfoxide (DMSO) in FBS.

2.2.2. Cell fusion experiments

Inter-species or intra-species cell fusion experiments were performed between

mouse ES or mouse EG cells and either human B, mouse B, 2rB lymphocytes or IMR90

fibroblasts as previously described (Pereira and Fisher, 2009), with minor modifications.

Briefly, cells were mixed in a conical 20-ml universal tube with a 1:1 ratio and washed

twice with KO-DMEM. After the last wash, the supernatant was completely removed, the

cell pellet was gently broken and 1mL of polyethylene glycol (PEG 1500, Sigma) was

added onto the cell pellet over a 60 second period. The PEG-cell mixture was incubated

at 37oC for 90 seconds with constant stirring. Then PEG was diluted by addition of 4mL of

50

Page 51: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

KO-DMEM over a 4 minute period, followed by additional 10 mL KO-DMEM, and

incubation at 37oC for 3 minutes. The cells were then centrifuged at 1350 rpm for 5 minutes

followed by the removal of the supernatant and the cells were let to swell in complete

mouse ES or mouse EG cell media for 3 minutes before resuspension. Cell mixtures were

plated under conditions that promote maintenance of pluripotency at 0.5x106cells/cm2

density. Non-fused mouse ES or mouse EG cells were eliminated by the addition of

puromycin (1µg/ml, Sigma), after 6-12 hours fusion and onwards. In the case of mouse

ES (H2BmCherry) with IMR90 (HP1αGFP) fusion, double positive heterokaryons were

isolated by Fluorescence Activated Cell Sorting (FACS, see section 2.1.3) after 48 hours

of fusion and plated for further culture. GFP positive reprogrammed inter-species hybrid

colonies generated between mouse ES or mouse EG cells and mouse B or 2rB

lymphocytes were picked (between 8 to 10 days of fusion) under fluorescence microscope

(Leica DM IRE2), dissociated with 0.05% Trypsin-EDTA, plated back and further cultured.

At earlier time points, GFP positive reprogrammed cells were isolated by FACS and further

processed for DNA isolation.

2.2.3. Fluorescence activated cell sorting (FACS)

Cells were resuspended in Phosphate Buffered Saline -/- (PBS, Gibco)

supplemented with 2% FBS and 2mM EDTA (FACS buffer) and run in the FACS

instrument (BD FACSAria III). Sorted cells were collected in FACS buffer, and plated in

appropriate culture conditions.

2.2.4. Quantitative Reverse Transcription Polymerase Chain Reaction (qRT-PCR) Analysis

2.2.4.1. RNA extraction and reverse transcription Cell lysates were homogenised with the QIAshredder columns (Qiagen), total RNA

was extracted using the RNeasy Mini kit (Qiagen) and residual DNA was eliminated with

the DNA-free kit (Ambion), following manufacturer’s instructions. RNA concentration was

measured using NanoDrop (Thermo). Reverse transcription of the total RNA was

performed using SuperScript III First-Strand Synthesis system (Invitrogen). Briefly, up to

51

Page 52: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

3µg of RNA was diluted in RNase free water containing 1µl of 10mM dNTP mix and 1µl of

oligo(dT)12-18 (Invitrogen) to a final volume of 11µl. The mixture was incubated at 65oC for

5 minutes and kept at 4oC for 2 minutes, while 4µl of 5X first strand buffer, 1µl of 0.1M

dithiothreitol (DTT), 1µl of RNase OUT and 1µl of 200U/µl of SuperScript III (Invitrogen)

were added. An additional mixture was also prepared without the reverse transcriptase as

a negative control. The mixture was then incubated at 25oC for 5 minutes, 50oC for 50

minutes and 75oC for 15 minutes to generate cDNA.

2.2.4.2. Quantitative PCR

The PCR mixture was prepared as follows: 2µl of cDNA, 300nM of forward and

reverse primers, 1X QuantiTect Sybr Green PCR mix (Qiagen) completed to 20µl with

DNase free water. An additional mixture was also prepared without cDNA as a negative

control for primer dimer formation. The cDNA in the mixture was amplified (in duplicates)

in Chromo 4 PCR instrument (Bio-Rad), using Opticon Monitor 3 software (MJ Research)

under the following cycling conditions: initial denaturation at 95oC for 15 minutes, 40 cycles

of denaturation at 94oC for 15 seconds, annealing at 60oC for 30 seconds and elongation

at 72oC for 30 seconds. Fluorescence quantification was performed at 72oC, 75oC, 78oC

and 83oC, and melting curve was determined from 70oC to 90oC, at 0.2oC intervals. The

qPCR data analysis was performed using the Opticon Monitor 3 software and the relative

cDNA abundance of the samples (to GAPDH and Ubc housekeeping genes) was

calculated with the obtained Ct values (threshold cycle number in which the fluorescence

resulting from the amplification becomes detectable above background) using the formula

2ΔΔCt. Primer sequences for qPCR analysis can be found in the Appendix Table 1.

For quantitative PCR analysis of ChIP samples similar settings were used except

for instead of the cDNA, 5µl of DNA was added in the reaction mixture. Primer sequenced

for ChIP analysis can be found in the Appendix Table 4.

2.2.5. DNA methylation and hydroxymethylation analyses

All the reagents described below were supplied from Sigma, unless otherwise

specified. For genomic DNA isolation, cells were first lysed by overnight incubation at 55oC

in 500µl of lysis buffer that contains 10mM NaCl, 10mM Tris-HCl pH 7.5, 10mM EDTA,

0.5% Sarcosyl and 200µg/ml Proteinase K. The genomic DNA containing aqueous

52

Page 53: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

solution was collected upon subsequent treatments with Phenol,

Phenol/Chloroform/Isoamyl Alcohol and Chloroform and the DNA was precipitated using

1/10 volume 3M Sodium Acetate pH 5.2 and 2.5X volume Ethanol, followed by

resuspension in Tris-EDTA Buffer (10mM Tris. 1mM EDTA). DNA concentration was

measured by NanoDrop (Thermo).

2.2.5.1. Bisulfite sequencing analysis

Up to 1.5µg of genomic DNA was used for bisulfite conversion reaction that was

performed using the EZ DNA Methylation Kit (Zymo Research), following manufacturer’s

instructions. Bisulfite converted DNA was recovered by columns provided in the kit in 10µl

elution buffer (provided in the kit). 2µl of bisulfite converted DNA was used for PCR

amplification, using the Taq PCR Kit (New England BioLabs) following manufacturer’s

instructions, with 10µM primers listed in Appendix Table 2. PCR primers were designed

to amplify bisulfite converted species-specific DNA and tested in-silico using the Bisearch

Web Server (http://bisearch.enzim.hu). Amplified PCR products were cloned into pGEM-

T easy vector (Promega), following manufacturer’s instructions, and competent DH5α

bacteria were transformed by heat shock. At least ten bacterial colonies were randomly

picked, incubated overnight in lysogeny broth at 37oC with agitation and plasmid DNA was

isolated using Wizard SV 96 plasmid purification system (Promega). DNA sequencing was

performed by the MRC Clinical Sciences Centre Sequencing Facility and obtained

sequences were analysed by CLC Main Workbench software (Qiagen).

2.2.5.2. 5-hmC quantification by enzyme protection assay

10µg of genomic DNA was either treated with T4 Phage β-glucosyltransferase (T4-

BGT, New England BioLabs) (T4+) or not (T4-) following manufacturer’s instructions.

Glucosylated and non-glucosylated genomic DNA were either treated with methylation

insensitive 100U of MspI restriction enzyme (New England BioLAbs), or not treated with

enzyme (mock digestion) at 37oC for 4 hours, followed by Proteinase K treatment for 30

minutes at 40oC. Primers that span the MspI restriction site were used to quantify by

quantitative PCR the T4+ and T4- samples, normalised individually to the amplification of

a control region that does not contain MspI site, and subtraction between the two

normalised levels translates into percentage 5-hmC levels. Primer sequences for enzyme

protection assay can be found in the Appendix Table 3.

53

Page 54: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

2.2.6. Imaging analysis

2.2.6.1. Immunofluorescence and confocal microscopy imaging

IMR90 (HP1αGFP) fibroblasts, mouse ES (H2BmCherry) cells and heterokaryons

were grown on glass coverslips coated with 0.1% gelatin. The coverslips were washed

with PBS and the cells were fixed in 2% paraformaldehyde (Sigma) for 15 minutes. Fixed

samples were washed three times in PBS for 5 minutes and the cells were permeabilized

with 0.4% Triton X-100 (Sigma) for 5 minutes. The samples were incubated in blocking

solution [2.5% bovine serum albumin (BSA, Sigma), 0.05% Tween 20 (Sigma), normal

goat serum (Sigma) in PBS], supplemented with Alexa Fluor 647 Phalloidin (1:50 dilution,

Invitrogen), for 1 hour in humid chamber. After three washes in washing buffer (0.2% BSA,

0.05% Tween 20, in PBS) for five minutes, coverslips were mounted using Vectashield

medium supplemented with DAPI (Vector Laboratories). Samples were visualised with a

SP5 Leica laser-scanning confocal microscope and processed using Leica Application

Suite Software and Adobe Photoshop CS5.

2.2.6.2. X-gal staining

For X-gal staining all reagents were supplied from Sigma. mEG/2rB hybrids were

fixed for 15 minutes in 0.1M PBS supplemented with 5mM ethylene glycol tetraacetic acid

(EGTA), 2mM MgCl2, 0.2% gluteraldehyde, followed by washing with 0.1M PBS

supplemented with 2mM MgCl2, and incubated in staining buffer (0.1M PBS supplemented

with 2mM MgCl2, 5mM potassium ferrocyanide, 5mM potassium ferricyanide and 1mg/ml

X-gal) for 16 hours at 37oC. After three washes, stained plates were analysed by inverted

microscope with 10X objective (Leica).

2.2.7. Western Blot analysis

All reagents were supplied from Sigma unless otherwise specified. Whole cell

protein extract was prepared by cell lysis in sample buffer (50mM Tris-HCl pH6.8, 1%SDS,

10% Glycerol) by incubation at 95oC and vortexing. After quantification (Pierce BCA

assay), Bromophenol Blue and 5% β-mercaptoethanol were added (0.001% and 5%,

respectively) into the samples. Bio-Rad minigel system was used to perform Sodium

54

Page 55: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

dodecyl sulphate – polyacrylamide gel electrophoresis (SDS-PAGE). 10 to 50µg of protein

was loaded on acrylamide (Bio-Rad) stacking gel (4% acrylamide, 0.125M Tris-HCl pH

6.8, 0.1% SDS, 0.1% ammonium persulphate and 0.1% tetramethylethylenediamine

(TEMED), and separated in a 10% acrylamide resolving gel (10% acrylamide, 0.4M Tris

pH8.8, 0.1%SDS, 0.1% ammonium persulphate and 0.1% TEMED), using Tris-glycine

buffer (25mM Tris, 192mM glycine, 0.1% SDS). Resolved proteins in acrylamide gels were

blotted to Protran nitrocellulose transfer membrane (Schleicher&Schuell Bioscience), in

transfer buffer (48mM Tris, 39mM glycine, 0.037% SDS and 20% methanol), using the

trans-blot semi-dry electrophoretic transfer instrument (Bio-Rad). The membranes were

incubated for 1 hour with blocking buffer (5% fat free milk powder, 20mM Tris pH 7.5,

150mM NaCl), followed by incubation with primary antibody diluted in blocking buffer for

1 hour at room temperature. After three washes in washing buffer (20mM Tris pH 7.5,

150mM NaCl, 0.1% Tween 20), membranes were incubated with horseradish peroxidase-

coupled secondary antibodies (anti-rabbit at 1:5000, anti-mouse at 1:2000 dilutions, both

from Amersham, anti-goat at 1:2000 dilution, from Santa Cruz), diluted in blocking buffer

for 1 hour. ECL Prime western blotting detection kit (Amersham) was used for signal

detection using Kodak Carestream photographic films and Kodak X-Omat Developer.

2.2.8. Chromatin Immunoprecipitation (ChIP) analysis

ChIP analysis was performed using the LowCell# ChIP Kit from Diagenode (all

reagents described below were supplied from Diagenode). Briefly, Anti-H3, Anti-H3K4me3

and IGG (as negative control) antibodies were bound to magnetic beads following

manufacturer’s instructions. Up to 100,000 cells were resuspended in PBS, fixed with

formaldehyde for 8 minutes before the solution was quenched with glycine, using the

concentrations advised in the kit. Fixed cells were washed twice in cold PBS and lysed

and sonicated (Bioruptor Plus, Diagenode) for 12 cycles of 30 seconds ON and 30

seconds pause. Sheared chromatin was overnight incubated with antibody coated

magnetic beads at 4oC. After the washes, magnetic bead-bound chromatin was eluted,

and the DNA was isolated using IPure kit (Diagenode). Isolated DNA was later processed

by quantitative PCR as described in section 2.2.4.2.

55

Page 56: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

2.2.9. Plasmid construction and delivery into ES cells

Short hairpin RNA (shRNA) plasmids were constructed using pSuper.GFP.Neo

backbone vector (Oligoengine, Figure 2.1.A). Synthetic primer pairs (sense and

antisense) that contain the shRNA targeting sequence were annealed in a reaction

mixture that contain 1X T4 ligase buffer (Invitrogen), by incubation at 90oC for 4 minutes,

and slowly cooling down to room temperature. The annealed oligo inserts were ligated

into BglII and HindIII digested pSuper.GFP.Neo vector by T4 ligation reaction (Invitrogen).

H2BmCherry sequence was kindly provided by Dr. Nobuaki Kudo. The sequence

was amplified with primers that contain XhoI and NotI sites and ligated into XhoI and NotI

digested pCAGIresPuro vector (Figure 2.1.B) (Niwa et al., 2002).

HP1αGFP sequence was kindly provided by Dr. Jesus Gil. The sequence was

amplified with primers that contain BglII and EcoRI and ligated into BglII and EcoRI

digested pMIP retroviral vector (kindly provided by Bradley Cobb, Figure 2.1.C).

All of the above mentioned ligation products were transformed into competent

DH5α bacteria by heat shock. Bacteria colonies were picked, incubated overnight in

lysogeny broth at 37oC with agitation and plasmid DNA was isolated by QIAprep Spin

Miniprep Kit (Qiagen). DNA sequencing was performed by the MRC Clinical Sciences

Centre Sequencing Facility and obtained sequences were analysed for correct insertion

by CLC Main Workbench software (Qiagen).

Figure 2.1. Vectors used for delivery. (A) short hairpin RNA sequence was cloned downstream of H1 RNA Polymerase III promoter. NeoR: Neomycin resistance Cassette, GFP: Green Fluorescence Protein. (B) H2BmCherry coding sequence was cloned downstream of Chicken Beta Actin (CBA) Promoter and upstream of IRES (Internal Ribosome Entry Site). PuroR: Puromycin resistance cassette. (C) Hp1αGFP coding sequence was cloned downstream of Murine Stem Cell Virus (MSCV) promoter.

shRNA constructs were delivered into mouse ES and mouse EG cells by

electroporation using Amaxa Nucleofector 2b system (VPH-1001, programme A-024).

PCAGIresPuro-H2BmCherry plasmid was similarly electroporated into mouse ES cells

56

Page 57: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

and Puromycin resistant mCherry positive single clones were picked and expanded. For

retroviral infection, pMSCVIresPuro-HP1αGFP plasmid was transfected into 293T cells

with helper p10A1 plasmid using calcium phosphate. Culture supernatants containing the

retrovirus were harvested post-transfection and added on IMR90 fibroblast culture

together with 2 µg/ml polybrene. Cells were treated with Puromycin (1.5µg/ml) and

expanded.

2.2.10. CRISPR/Cas9 genome editing system

2.2.10.1. CRISPR/Cas9 plasmid construction

All of the reagents used for plasmid construction were supplied from New England

BioLabs, unless otherwise specified. The plasmid used for cloning was supplied from

Addgene (px330, plasmid 42230, Figure 2.2) (Cong et al., 2013). The targeting guide RNA

sequence was designed using a bioinformatics software available online at

http://crispr.mit.edu/ as sense and antisense complementary oligonucleotides as

illustrated in Figure 2.2.

Figure 2.2. px330 vector and the guide RNA sequence. Annealed sense and antisense guide RNA sequences contain 5’ and 3’ overhangs for ligation into BbsI digested px330 plasmid. The ligation results in chimeric single guide RNA sequence, transcribed by the RNA U6 Polymerase III promoter. Chicken beta actin promoter drives the transcription of human codon optimised Streptococcus Pyogenes Cas9, which contains nuclear localisation signal (NLS) sequences on both ends.

The px330 plasmid was digested with BbsI restriction enzyme (Thermo, #ER1011)

at 37oC for 1 hour, heat inactivated, dephosphorylated by Antarctic Phosphatase, heat

inactivated and gel purified (QIAquick Gel Extraction Kit, Qiagen). In the meantime

complementary guide sequences were annealed in 1X T4 Ligase buffer (Invitrogen), by

incubating at 95oC for 3 minutes, and slowly cooling down to room temperature. After

phosphorylation by T4 PNK, annealed oligos were ligated to the linearized vector using

T4 DNA ligase (Invitrogen) and competent DH5α bacteria were transformed by heat

shock. Bacteria colonies were picked, incubated overnight in lysogeny broth at 37oC with

57

Page 58: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 2

agitation and plasmid DNA was isolated by QIAprep Spin Miniprep Kit (Qiagen). DNA

sequencing was performed by the MRC Clinical Sciences Centre Sequencing Facility and

obtained sequences were analysed for correct insertion by CLC Main Workbench software

(Qiagen).

2.2.10.2. Surveyor and RFLP Assays

The Surveyor assay, as described in Chapter 5.3.2, was used to detect the

presence of indel mutations at the CRISPR target locus. The target locus was PCR

amplified (primers are listed in Appendix Table 5) using Phusion High-Fidelity DNA

Polymerase (New England BioLabs) followed by PCR purification (QIAquick PCR

Purification Kit, Qiagen). 360ng of purified amplicons were mixed with 1X Taq Buffer (New

England BioLabs), denatured and reannealed [as described in (Ran et al., 2013a)] and

treated with Surveyor DNA endonuclease (Transgenomic), following manufacturer’s

instructions. Samples were run on 2% Agarose gel for detection.

The Restriction Fragment Length Polymorphism (RFLP) was used where

applicable, as described in Chapter 5.3.3. CRISPR target locus was amplified followed by

PCR purification as described above. PCR amplicons were digested by suitable restriction

enzymes and ran on 2% Agarose gel for visualisation.

58

Page 59: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

Chapter 3. Pluripotency Gene Demethylation during

Reprogramming

3.1. Introduction

The fate of pluripotent cells gradually becomes restricted as they acquire specific

properties and differentiate to become part of a particular tissue or organ. Under normal

conditions, once a cell is differentiated, it appears to be epigenetically protected in the

sense that its identity remains stable in-vivo and propagated to daughters upon cell

division. We know that although cells don’t generally change their identity, cells remain

flexible, and their epigenetic status can be reset (Reik, 2007). This can be achieved by

specific experimental approaches including transcription factor induced-reprogramming

and nuclear transfer. A third method is cell fusion, where a pluripotent cell can dominantly

alter the identity of the differentiated fusion partner towards pluripotency (Yamanaka and

Blau, 2010). Our laboratory and others have shown that pluripotent mouse embryonic

stem (mES) cells can efficiently induce reprogramming of somatic cells upon fusion, over

a short period of time (Pereira et al., 2008, 2010; Tada et al., 2001). Although it has

previously been reported that DNA demethylation is critical for the acquisition of

pluripotency (Mikkelsen et al., 2008; Simonsson and Gurdon, 2004), how DNA

demethylation is achieved early in reprogramming is not yet clear. In this Chapter, I use

cell fusion-based reprogramming to examine the changing status of DNA methylation at

pluripotency associated genes during pluripotent conversion of the somatic nucleus.

3.2. Reprogramming of human B lymphocytes upon fusion with mouse embryonic stem cells

Our laboratory has previously demonstrated that interspecies fusion between

mouse ES and human B lymphocytes results in efficient reprogramming of the somatic

nuclei, and that this happens at the early heterokaryon stage, in which discrete nuclei

originating from both fusion partners are apparent (Pereira et al., 2008, 2010) (Figure

3.1.A). This transient period lasts up to 72 hours after fusion and terminates when the two

nuclei merge to form a tetraploid hybrid. Because of genetic incompatibility, although

59

Page 60: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

interspecies hybrids can divide, they are not karyotypically stable and are not viable for

long-term. In addition, although cell division does not occur in heterokaryons, we have

recently shown that DNA synthesis is widespread at this stage, and that it has a positive

impact on reprogramming efficiency (Tsubouchi et al., 2013).

To begin to examine DNA methylation changes during cell fusion-based

reprogramming, I fused mouse ES cells (E14tg2A) and human B lymphocytes (Epstein -

Barr virus transformed adult B cells), using polyethylene glycol (PEG). The advantage of

performing interspecies fusion is that it allows us to analyse species-specific gene

expression. By using primers specific for human genes including pluripotency factors

(OCT4, NANOG, CRIPTO), I analysed the induction of human pluripotency gene

expression by performing qRT-PCR in human B lymphocytes (0h, before fusion) and in

heterokaryons (72h after fusion) (Figure 3.1.B).

Figure 3.1. Interspecies cell fusion and reprogramming of human B lymphocyte by mES cells. (A) Schematic representation of interspecies fusion between mES cells (light grey) and human B lymphocytes (dark grey). Nuclei originating from both cell types are separate at the heterokaryon stage (persisting for ~72 hours after fusion) and tetraploid hybrids are formed upon nuclei merging. (B) qRT-PCR analysis of pluripotency factor gene expression (OCT4, NANOG and CRIPTO) in hB lymphocytes before (0h) and at 72 hours after fusion with mES cells. Data were normalised to human GAPDH and represent mean and SEM of 3 different experiments.

Transcript analysis revealed that pluripotency network is silent in B lymphocytes,

as evidenced by the lack of expression of OCT4, NANOG, and CRIPTO before fusion (0h)

60

Page 61: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

(Figure 3.1.B). This profile was altered as early as 72 hours after fusion with mES cells,

where I observed significant upregulation of pluripotency-associated gene expression

originating from human nucleus. These results together with previous work (Pereira et al.,

2008, 2010) show that cell fusion mediated reprogramming provides an opportunity to

study the molecular details of early stages of pluripotent induction.

3.3. DNA methylation profiles of pluripotency associated genes in human B lymphocytes and human ES cells

Among many of the epigenetic regulators, DNA methylation is a stable modification

and is associated with inhibition of gene expression (Klose and Bird, 2006). To understand

the status of DNA methylation at the promoters of pluripotency-associated genes in

human B lymphocytes and human ES cells (H7), I conducted bisulfite sequencing analysis

(Figure 3.2). This method enables the assessment of individual 5-methylcytosine residues

in DNA strands. It is based on bisulfite-induced modification of genomic DNA, where

cytosine molecules are converted to uracil while 5-methylcytosine remains nonreactive

Figure 3.2. Bisulfite sequencing of OCT4, NANOG and CRIPTO promoters in human B lymphocytes and human ES cells. Genomic DNA from human B lymphocytes (top) and human ES cells (bottom) were bisulfite converted, and PCR amplicons for OCT4, NANOG and CRIPTO promoters were sequenced. Transcription start sites for each gene were taken as reference and relative positions of CpG sites were indicated. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

61

Page 62: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

that can be distinguished by sequencing of the PCR products (Frommer et al., 1992). H7

human ES cells express high levels of pluripotency-associated factors, including OCT4,

and have the capacity to differentiate into cell types from all the three germ layers

(Thomson, 1998; Xu et al., 2001). In differentiated cells (human B lymphocytes), the

promoters of pluripotency-associated OCT4, NANOG and CRIPTO genes were highly

methylated (Figure 3.2, top). Consistently, these genes lacked CpG methylation in

pluripotent human ES cells (Figure 3.2, bottom).

Overall, these data suggest that there is a correlation between pluripotency-

associated gene expression and lack of cytosine methylation.

3.4. Changes in DNA methylation of OCT4 accompanies reprogramming but the extent is variable

Upon fusion with mES cells, human B cell nucleus is rapidly reprogrammed and

starts expressing components of pluripotency network genes. To assess how DNA

methylation signatures were altered in the course of pluripotent reactivation, I conducted

genomic bisulfite sequencing using human specific primers for amplification and

sequencing of the OCT4 promoter. Figure 3.3 demonstrates a single cell fusion

experiment between mES cells and human B cells, where I analysed gene expression

levels and methylation status of OCT4 in heterokaryons after 72 hours of fusion.

Successful reprogramming was evident as shown by transcript analysis, and this was

accompanied by partial demethylation of human OCT4 promoter (Figure 3.3).

Figure 3.3. Transcript analysis and bisulfite sequencing of human OCT4 in heterokaryons after 72 hours of fusion. (Left) qRT-PCR analysis of pluripotency factor gene expression (OCT4, NANOG and CRIPTO) in hB lymphocytes before (0h) and at 72 hours after fusion with mES cells, data were normalised to human GAPDH. (Right) Bisulfite sequencing analysis on human OCT4 promoter in heterokaryons after 72 hours of fusion, where relative positions of CpG sites to transcription start site were indicated. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

62

Page 63: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

Even though this data demonstrated a relation between demethylation and

induction of gene expression in heterokaryons, a second experiment I conducted showed

no correspondence between OCT4 expression and cytosine demethylation at the OCT4

promoter (Figure 3.4.A). I performed this analysis in three more experiments, but again

saw no strong correlation between gene reactivation and DNA demethylation, at least at

the population level in heterokaryons (Figure 3.4.B).

Figure 3.4. Transcript analysis and bisulfite sequencing of human OCT4 in heterokaryons after 72 hours of fusion in four different experiments. (A) (Left) qRT-PCR analysis of pluripotency factor gene expression (OCT4, NANOG and CRIPTO) in hB lymphocytes before (0h) and at 72 hours after fusion with mES cells, data were normalised to human GAPDH. (Right) Bisulfite sequencing analysis on human OCT4 promoter in heterokaryons after 72 hours of fusion, where relative positions of CpG sites to transcription start site were indicated. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively. (B) Same analysis as in (A) in 3 more experiments.

63

Page 64: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

3.5. Reprogramming human fibroblasts and OCT4 induction without detectable changes in DNA methylation

In order to repeat previous investigations in a different cell type, I used human

fibroblasts as fusion partners of mouse ES cells. I generated human fibroblasts (IMR90)

and mouse ES cells constitutively expressing HP1α-GFP and H2B-mCherry respectively.

After two days of fusion, I isolated double positive heterokaryons by Fluorescence

Activated Cell Sorting (FACS) and plated them back for two more days of culture. Then I

re-sorted double positive heterokaryons, and used these to do transcript and bisulfite

sequencing analyses (Figure 3.5.A).

Gene expression analysis after 96 hours of fusion demonstrated upregulation of

pluripotency-associated genes OCT4, NANOG and CRIPTO originating from human

fibroblast (Figure 3.5.B). Despite this, bisulfite analysis showed that the human OCT4

promoter remained methylated at this stage as assessed in two different experiments

(Figure 3.5.C). Collectively these data suggest that demethylation of the human OCT4

promoter may not be essential for expression.

64

Page 65: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

Figure 3.5. Fusion of human fibroblasts with mouse ES cells and transcript and bisulfite sequencing analyses. (A) Schematic demonstration of the experimental approach where human fibroblasts (IMR90) expressing HP1α-GFP were fused with mouse ES cells expressing H2B-mCherry. Double positive cells (GFP+&mCherry+) were sorted after 2 days of fusion (Q2 population) and plated. 2 days later, same sorting settings were applied and the re-sorted population was analysed for gene expression and methylation profiles. (B) qRT-PCR analysis of pluripotency factor gene expression (OCT4, NANOG and CRIPTO) in hFibroblasts before (0h) and at 96 hours after fusion with mES cells. Data were normalised to human GAPDH and represent mean and SEM of 2 different experiments. (C) Bisulfite sequencing analysis on human OCT4 promoter in human fibroblasts before (0h) and at 96 hours after fusion (in 2 independent experiments), where relative positions of CpG sites to transcription start site were indicated. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

65

Page 66: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

3.6. No evidence of DNA methylation changes at site upstream of the OCT4 transcription start site.

Using similar system, settings and cell lines, Bhutani et al. have previously reported

DNA demethylation at the OCT4 upstream region upon fusion (Bhutani et al., 2010). The

analysed region in this report is localised between the OCT4 promoter and the proximal

enhancer sites (Nordhoff et al., 2001) (Figure 3.6.A). Therefore, I investigated the

methylation status of the same locus in human fibroblast DNA before and after fusion.

Human fibroblast DNA was fully methylated as expected, however I did not detect a loss

of DNA methylation after fusion with mES cells as demonstrated in two independent

experiments (Figure 3.6.B). Interestingly, this region (unlike the promoter), was mostly

methylated in H7 human ES cells (Figure 3.6.C). In addition, the same DNA methylation

pattern at the OCT4 upstream was previously reported in pluripotent human embryonal

carcinoma cells (Deb-Rinker et al., 2005), calling into question the published role of this

site in regulating OCT4 expression.

Figure 3.6. Bisulfite sequencing of human OCT4 upstream region in human fibroblasts before and after fusion and in human ES cells. (A) Schematic representation of human OCT4 locus. The arrow indicates transcription start site (TSS), and red lines indicate locations of the CpG sites relative to the TSS. Location of the five CpG sites analysed by Bhutani et al. and here is marked. Bisulfite sequencing in (B) human fibroblasts (IMR90) before (0h) and at 96 hours after fusion (in 2 independent experiments) and in (C) human embryonic stem cells (H7) where relative positions of CpG sites to transcription start site were indicated. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

66

Page 67: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

3.7. DNA demethylation kinetics of somatic Oct4 transgene in reprogrammed mouse hybrids.

To better understand the relationship between Oct4 re-expression and locus

demethylation during reprogramming, I made use of a mouse indicator B cell line in which

re-expression of Oct4 was evident by GFP expression. In order to track reprogrammed

cells, I used B cell lines isolated from transgenic mice bearing GFP coding sequence

under the control of 18 kb genomic mouse Oct4 fragment (GOF-18/ΔPE/GFP) (Yoshimizu

et al., 1999) (Figure 3.5.7, left). This fragment was previously shown to be sufficient to

recapitulate endogenous expression patterns of Oct4 in embryonic development, thus

GFP expression under its regulation can be considered as a direct sign for pluripotency

(Yeom et al., 1996). Upon fusion with mouse ES cells, the silent transgene carried by the

mouse B cells was reactivated, leading to a prominent GFP expression in hybrid colonies

(Figure 3.7.A, right). To assess the methylation status of the transgene in mouse B cells

and to track how it is altered during pluripotency induction, I conducted bisulfite

sequencing on the region which spans exogenous Oct4 promoter site and GFP coding

sequence. As expected, the transgene was fully methylated in mouse B lymphocytes

(Figure 3.7.B), in line with the lack of GFP transgene expression (data not shown). To

examine the kinetics of DNA demethylation of the transgene in the course of

reprogramming, I isolated GFP expressing hybrid cells by FACS at different time points

(3, 6 and 13 days after fusion). Then I conducted bisulfite conversion on each sample and

sequenced the exogenous Oct4-GFP genomic fragments (by using PCR primers that

specifically span the promoter/GFP region). Analysis of DNA methylation revealed that

among GFP positive cells isolated 3 days after fusion, partial demethylation was seen. By

6 days, DNA methylation was completely lost, and the region remained un-methylated at

least until 13 days of fusion. (Figure 3.7.C).

These data demonstrated that reprogramming was rapidly induced following intra-

species cell fusion and resulted in GFP expression under the control of the exogenous

Oct4 promoter. Interestingly, this was accompanied by partial demethylation (as observed

3 days after fusion), followed later by complete loss of DNA methylation.

67

Page 68: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

Figure 3.7. DNA demethylation kinetics upon reprogramming in mouse hybrids. (A) (Left) Schematic representation of interspecies fusion between mES cells and mouse B lymphocytes carrying an exogenous Oct4-GFP fragment. Nuclei originating from both cell types are separate at the heterokaryon stage (persisting for ~72 hours after fusion) and tetraploid hybrids are formed upon nuclei merging. (Right) Bright field and fluorescent microscopy images of a hybrid colony, cultured for 10 days after fusion. (B) Bisulfite sequencing analysis on Oct4-GFP transgene in mouse B lymphocytes and (C) in GFP expressing hybrids at indicated time points. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

3.8. Summary and Discussion

Cell fusion allows an assessment of the impact of one cell type on another. This is

achieved by trans-acting factor binding to DNA sites that can lead to reversing of cell fate.

Our laboratory and others have previously demonstrated that differentiated cells can

successfully acquire a pluripotent-like identity upon fusion with mouse embryonic stem

cells (Pereira et al., 2008, 2010; Tada et al., 2001). In this Chapter I have re-examined

these findings with a focus on DNA methylation dynamics that occur during

reprogramming.

The activation of silenced genes in interspecies heterokaryons was first

demonstrated by fusing mouse muscle cells with human amniotic cells, where initiation of

previously silent human muscle-specific gene expression was detected (Blau et al., 1983).

68

Page 69: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

Here I confirm that a silent pluripotent program can be reactivated in somatic human B

lymphocytes after fusion with mouse ES cells (Pereira et al., 2008). Pluripotency-related

gene activation is detected as early as 24 hours of fusion in heterokaryons.

How trans-acting factors remodel DNA elements and relieve repressive chromatin

structure is still largely unknown. DNA methylation is a key epigenetic mechanism involved

in stable transcriptional silencing of gene expression. Here I demonstrated that DNA

methylation was inversely correlated with pluripotency-associated gene expression in

human B lymphocytes. During reprogramming, it might be assumed that DNA

demethylation would occur in parallel to gene re-activation. Indeed, nuclear

transplantation experiments have revealed that DNA demethylation is necessary and

precedes Oct4 transcription during pluripotent conversion (Simonsson and Gurdon, 2004).

However, in my interspecies heterokaryon experiments I did not see a direct correlation

between promoter DNA demethylation and transcriptional activation of OCT4 gene.

Instead, I observed either lack of, or partial demethylation of OCT4, despite OCT4 gene

re-expression. Similar to these results, a study has demonstrated that when 293T cells

were treated by pluripotent embryonic carcinoma cell extract, OCT4 gene reactivation was

accompanied by a mosaic demethylation pattern after 4 weeks of culture (Freberg et al.,

2007). In a related study, Foshay et. al. analysed changes in gene expression in rat

fibroblasts after fusion with mouse ES cells, and concluded that reprogramming of cis-

silenced genes occurs with rather slow kinetics and requires DNA synthesis (Foshay et

al., 2012). These authors suggested that this was due to repressive histone marks, and

the presence of DNA methylation around genes such as Oct4 and Nanog. Moreover, the

rate of Oct4 promoter demethylation was slow when analysed at different times after

fusion (at population level), and was only complete in selected (and cultured) stable

hybrids (Foshay et al., 2012). One explanation for my observations might be that the

initiation of OCT4 gene expression may take place even in the presence of methylated

CpGs and be regulated by epigenetic features and mechanisms such as active histone

modifications (See Chapter 5.4 for discussion). A second rationale for the lack of

demethylation might be due to the fact that the analysis was conducted at a population

level; in a group of heterokaryons, it is possible that some fused cells may not undergo

reprogramming. This would obscure the bisulfite analysis, even though mRNA transcripts

for OCT4 can be detected in the population. To overcome this issue, it would be crucial to

conduct a single heterokaryon gene expression analysis coupled with single cell DNA

methylation profiling [which are now possible with the advent of microfluidic platforms

(Lorthongpanich et al., 2013)]. I am currently looking at the feasibility of this analysis.

69

Page 70: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

Previously, Bhutani et. al. showed demethylation of OCT4 upstream region in

human fibroblasts soon after fusion with mouse ES cells (Bhutani et al., 2010). The

authors claimed that this was mediated by AID, as Aid knockdown impaired

reprogramming and OCT4 demethylation. Using similar approach and cell lines, I did not

observe demethylation at the OCT4 transcription initiation site, nor in the upstream region

highlighted in the research article. Instead, I detected CpG methylation at this upstream

locus in human ES cells, which raises a major concern about the validity of this region in

regulating gene expression. Furthermore, the role of AID in DNA demethylation and

reprogramming has recently been re-evaluated. For example, in vitro examination

revealed a substantially reduced activity of AID on 5mC relative to cytosine, caused by

the adverse effect of steric hindrance on its function (Nabel et al., 2012). In addition, iPS

cells can successfully be generated from Aid deficient fibroblasts without any significant

changes in the DNA methylome (Habib et al., 2014; Shimamoto et al., 2014). Consistent

with these reports, I did not see any effect of Aid knockdown in reprogramming upon cell-

fusion and Aid mRNA levels in ES cells were already very low (data not shown), as also

observed in a similar study (Foshay et al., 2012).

I have demonstrated the kinetics of DNA demethylation of an exogenous Oct4-

GFP fragment integrated into mouse B cell genome, upon fusion with mouse ES cells.

With this system, successful reprogramming can be visualised using GFP marker as

previously demonstrated (Do and Schöler, 2004; Han et al., 2008; Silva et al., 2006; Wong

et al., 2008). Although the transgene was silent and fully methylated in mouse B cells, loss

of methylation was evident by 3 days of fusion, and was completed by 6 days. In these

experiments I sorted GFP-positive cells and therefore the bisulfite sequencing is biased

to successfully reprogrammed cells. Interestingly, at Day 3, there was a substantial

number of methylated clones, despite the expression of GFP. This can be attributed to a

couple of different possibilities. First, the transgene might have multiple copies in the

mouse B genome, and according to where the transgene was integrated, different kinetics

of demethylation would be encountered according to genomic context. Secondly, although

DNA methylation is symmetrical, in the course of loss of methylation one strand might be

demethylated while the complementary strand remains unmodified. This profile would be

expected to be seen during passive demethylation, and could be visualised using a hairpin

bisulfite sequencing strategy (Laird et al., 2004). Thirdly, the partial demethylation might

reflect ongoing conversion of 5-mC to 5-hmC and to subsequent derivatives either

actively, or through passive mechanisms.

70

Page 71: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 3

In conclusion my results demonstrated that reprogramming of somatic nuclei by

cellular fusion is a rapid process, evidenced by initiation of pluripotency associated gene

expression. Although this system allows us to study the early events that contribute to the

pluripotent conversion, exactly how gene reactivation is synchronised with DNA

demethylation is unclear due to heterogeneity in the heterokaryon population. One

contribution to this heterogeneity comes from the various cell cycle stages of the fusion

partners. Indeed we have shown that mouse ES cells in S/G2 phase are more efficient in

reprogramming human B lymphocytes upon fusion (Tsubouchi et al., 2013). This is due to

their capacity to induce DNA synthesis in the somatic nuclei. To better understand

sequential events taking place during early stages of reprogramming, it is important to

collect data from single heterokaryons in a high-throughput manner to collect transcript,

DNA methylation and reprogramming (imaging) analysis. The combination of those would

provide us a global understanding of progressive conversion towards pluripotency.

71

Page 72: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

Chapter 4. Mechanisms of Imprint Erasure in Somatic

Cells mediated by Embryonic Germ Cell Fusion

4.1. Introduction

It has previously been shown that despite their dominant pluripotent

reprogramming potential, mouse ES cells were not able to induce DNA demethylation of

imprinted regions in the somatic genome upon fusion (Tada et al., 2001). This is especially

interesting considering that some mouse embryonic germ (mEG) cells derived from

primordial germ cells, can on the other hand, re-set imprints (Tada et al., 1997). In this

Chapter, I analyse the kinetics of imprint erasure in somatic cells following fusion with

mouse EG cells, and show some of the molecular mechanisms likely to be involved in this

process.

4.2. Imprint erasure in somatic cells induced by embryonic germ cell fusion.

Embryonic germ (EG) cells are pluripotent in-vitro counterparts of primordial germ

cells (PGCs). They express many pluripotency-associated factors similar to ES cells (Mise

et al., 2008), but interestingly, EG cells exhibit genome-wide DNA hypomethylation that

includes imprinted domains (Labosky et al., 1994; Tada et al., 1998). An additional

difference is that EG cells, unlike ES cells, were reported to induce imprint erasure in

thymocyte DNA upon fusion in reprogrammed hybrids (Tada et al., 1997). With Francesco

Piccolo, I have examined the DNA methylation status of imprinted H19 loci in

reprogrammed hybrids generated between mouse B lymphocytes (bearing Oct4-GFP

transgene) and mouse EG cells. To assess whether imprint resetting would occur and if

so what its kinetics would be, I isolated GFP positive hybrids at 7, 12 and 21 days after

fusion, and conducted bisulfite sequencing analysis on those samples (Figure 4.1). The

hypomethylation observed in 21 day hybrids confirmed the erasure of imprints, induced

by mouse EG cells (1.6%, compared to initial 31.2%). Interestingly, methylation level was

maintained until at least 7 days of fusion (43%), and exhibited a gradual decrease after

this point (17.4% at 12 Days, Figure 4.1).

72

Page 73: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

Figure 4.1. CpG methylation analysis of imprinted H19 locus upon pluripotent reprogramming mediated by mouse EG cells. Genomic bisulfite sequencing of H19 ICR in mEG and mB lymphocytes (1:1 mixture) before fusion (0h) and in GFP sorted hybrids at 7, 12 and 21 Days after fusion. Corresponding methylation levels are depicted as percentage. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

4.3. Using dual reporter (2rB) somatic cells to assess the kinetics of imprint erasure during EG-reprogramming.

To elucidate whether EG cells were able to functionally reset imprinting in somatic

cells, I used an alternative mouse B lymphocyte cell line (2rB) as a fusion partner. The

2rB cells, in addition to Oct4-GFP transgene, possess a maternal LacZ knock-in allele of

imprinted Peg1 gene (Peg1M-β-gal) (Lefebvre et al., 1998). Peg1 is paternally expressed,

while the maternal allele is silent. In this setting β-galactosidase activity would be detected

only if the maternal imprinting is reset and the gene is reactivated. To first assess whether

Peg1 imprint can be reset by EG cell fusion (as in the case of H19), I analysed methylation

status of the Peg1 DMR in hybrids generated between mouse EG and 2rB cells. I isolated

GFP positive hybrids at consecutive time points (7, 12 and 21 days of fusion) and

conducted bisulfite sequencing. Although DNA methylation was maintained until at least

7 day hybrids (30%), it gradually decreased at further time points (13.3% at Day 12 and

7% at Day 21, Figure 4.2). This result demonstrated that, Peg1 imprint can efficiently be

re-set in EG/2rB hybrids.

73

Page 74: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

Figure 4.2. CpG methylation analysis of imprinted Peg1 locus upon pluripotent reprogramming mediated by moue EG cells. Genomic bisulfite sequencing of H19 ICR in mEG and 2rB lymphocytes and in GFP sorted hybrids at 7, 12 and 21 Days after fusion. Corresponding methylation levels are depicted as percentage. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

To assess the functional outcome of EG-mediated imprint erasure at the 2rB

somatic nucleus, Karen Brown and David Landeira investigated Peg1 driven β-

galactosidase (β-gal) activity in hybrids. In GFP positive hybrids at 12 days of fusion, β-

gal activity was not detected (as assessed by X-gal staining, Figure 4.3.A-B), indicating

that Peg1M-β-gal remained silent. This result was expected as Peg1 is only expressed in

differentiated cells (Lefebvre et al., 1998). To this end, EG/2rB hybrids were differentiated

by removing LIF from the culture media. Differentiated 22-day hybrids were positive for β-

gal activity as analysed by X-gal staining (Figure 4.3.A-B). This indicated that Peg1 gene

could be expressed from previously methylated 2rB maternal allele that was re-set in

hybrids by EG-cell fusion, however this was not observed in hybrids formed by ES-cell

fusion (Figure 4.3.C, see Chapter 4.5)

74

Page 75: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

Figure 4.3. Functional resetting of somatic imprints mediated by mouse EG cells. (A) Schematic illustration of the experimental strategy to assess imprint erasure in 2rB nucleus in EG-cell hybrids. In addition to Oct4-GFP transgene, 2rB cells carry a maternal LacZ knock-in allele of Peg1. (B) A mEG/2rB hybrid colony expresses GFP which is driven by an Oct4 promoter, however lacks β-gal activity as assessed by X-gal staining). 10 days after differentiation by LIF removal, EG/2rB hybrids are positive for β-gal activity (blue). (C) mES/2rB hybrids, differentiated by LIF removal for 10 days after 12 days of fusion, negative for β-gal activity as assessed by X-gal staining.

4.4. EG cell capacity to induce demethylation is not restricted to imprinted genes.

To understand whether EG cell-induced erasure of DNA methylation is limited to

imprinted genes or occurs on additional sites in the somatic nucleus, we analysed the

methylation status of long interspersed element (LINE) repeats in hybrids. Mouse EG cells

exhibited less DNA methylation at LINE1 repeats compared to 2rB lymphocytes. In GFP

positive hybrids 21 days of fusion of 2rB and EG cells, LINE1 methylation was significantly

reduced (15.8%) (Figure 4.4).

75

Page 76: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

Figure 4.4. CpG methylation analysis of LINE1 repeats upon pluripotent reprogramming mediated by mouse EG cells. Genomic bisulfite sequencing of LINE1 in mouse EG and 2rB lymphocytes and in GFP sorted hybrids at 21 days after fusion. Corresponding methylation levels are depicted in as percentage. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

4.5. Imprint erasure is not seen in fusions with mouse ES cells or female ES cells that are globally hypomethylated.

In 2001, Tada and colleagues showed that in hybrids generated between

thymocytes and mES cells, although reprogrammed, imprinted domains remained

methylated. By conducting methylation-sensitive restriction enzyme analysis, they showed

that the imprint control region (ICR) of the H19 locus maintained CpG methylation in hybrid

clones (Tada et al., 2001).

With Francesco Piccolo, we confirmed that, in addition to H19, Peg3 and Gtl2/Dlk1

ICRs were also produced in mouse ES x mouse B cell fusions. Bisulfite analysis on H19,

Peg3 and Gtl2/Dlk1 imprinted loci in 1 : 1 mixture of mES and mouse B lymphocytes

bearing an Oct4-GFP transgene (denoted as 0h, before fusion) demonstrated initial level

of methylation (56%, 32% and 35%, respectively, Figure 4.5). It is important to note that

although expected 50% methylation of imprinting is observed in mouse B cells, this level

is around 25-30% in mES cells (Piccolo et al., 2013). Then we conducted bisulfite analysis

at the same three loci in GFP positive reprogrammed hybrids generated between mES

and mB lymphocytes after 21 days of fusion. The survey revealed that ICR methylations

76

Page 77: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

of H19, Peg3 and Gtl2/Dlk1 genes were not altered in pluripotent hybrids. (45%, 27% and

46%, respectively, Figure 4.5).

Figure 4.5. CpG methylation analysis of imprinted H19, Peg3, Gtl2/Dlk1 loci upon pluripotent reprogramming induced by mouse ES cells. Genomic bisulfite sequencing of mES and mB lymphocytes (1:1 mixture) before fusion (0h) and corresponding methylation levels (depicted as percentage) (Top). Genomic bisulfite sequencing of 21 days hybrids of mES fused with mB lymphocytes and corresponding methylation levels (depicted as percentage) (Bottom). Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively.

To understand whether the genome-wide methylation status of the mES cells

would have an effect on imprint erasure, I repeated this analysis in hybrids formed

between mB lymphocytes and a female mouse embryonic stem cell line, Pgk12.1

(Zvetkova et al., 2005). Female mES cell lines were previously shown to exhibit global

DNA hypomethylation, including imprint control regions (Shovlin et al., 2008; Zvetkova et

al., 2005). Bisulfite sequencing revealed that H19 and Peg3 lost CpG methylation and

Gtl2/Dlk1 contained moderate levels of DNA methylation in Pgk12.1 cells (9%, 0% and

36%, respectively, Figure 4.6.A). Then I analysed whether these cells can induce imprint

erasure upon fusion with mouse B lymphocytes. Initial methylation levels of 1 : 1 mixture

of Pgk12.1 and mB cells (before fusion) were 26.1%, 22.1% and 37.1% for H19, Peg3 and

Gtl2/Dlk1, respectively (Figure 4.6.B, black dots). Bisulfite analysis of three different hybrid

77

Page 78: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

clones at 21 days after fusion showed that no significant change in methylation levels

occurred in these three loci (Figure 4.6.B, red dots). Overall, these data demonstrated that

although mES cells (male or female origin) can induce pluripotent conversion in somatic

cells upon fusion, they fail to reset imprinting.

Figure 4.6. CpG methylation analysis of imprinted H19, Peg3, Gtl2/Dlk1 loci upon pluripotent reprogramming induced by Pgk12.1 female mouse ES cells. (A) Genomic bisulfite analysis of H19, Peg3 and Gtl2/Dlk1 in Pgk12.1 female ES cells. Unfilled (white) circles and filled (black) circles represent un-methylated and methylated CpGs respectively. (B) Genomic bisulfite analysis of H19, Peg3 and Gtl2/Dlk1 before (0h, black dots) and after fusion where three individual hybrid clones were examined (21 Days, red dots). Each dot represents the percentage methylation acquired from at least 16 respective bisulfite converted DNA sequences. Percentage methylation values at 21 Days hybrids for H19 are 18.4%, 17.7% and 25%; for Peg3 are 23.1%, 29.3% and 14.7%; for Gtl2/Dlk1 are 17.9%, 31.1% and 29.2%.

4.6. Hydroxymethylation at imprinted loci upon fusion with mouse EG cells.

In mouse EG cell hybrids generated with mouse B lymphocytes, we noticed that

imprint erasure started long after the induction of gene re-expression. The gradual loss of

78

Page 79: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

ICR methylation led us hypothesise that this process might involve active oxidation of 5-

methylcytosine (to 5-hydroxymethylcytosine “5-hmC”) coupled to replication-dependent

passive DNA demethylation (Bagci and Fisher, 2013). To explore this, and to understand

early events that might contribute to imprint erasure, we conducted interspecies fusion

between mouse EG cells and human B lymphocytes. This provided us the opportunity to

specifically analyse modifications that may occur to human DNA. Francesco Piccolo in the

lab conducted restriction enzyme protection assays to quantify 5-hmC levels at several

imprinted loci. These results demonstrated that 5-hmC accumulated at human ICRs of

H19, Peg3 and SNRPN/SNURF imprinted genes, as early as 48 hours after fusion (Figure

4.7), providing a possible route for active conversion of 5-mC to 5-hmC at these ICRs.

Figure 4.7. Acquisition of 5-hmC at human B lymphocyte ICRs upon fusion with mouse EG cells. qPCR analysis of restriction enzyme protection assay (see Materials&Methods) to quantify 5hmC levels at human ICRs before (0h) and at 48 and 72 hours after fusion with mouse EG cells.

4.7. Tet regulated 5-mC oxidation at imprinted loci upon fusion with mouse EG cells.

Oxidation of 5-methylcytosine is catalysed by the mammalian TET family

members, composed of paralogous Tet1, Tet2 and Tet3 proteins that share significant

homology (Tahiliani et al., 2009). Pluripotent cells mainly express Tet1 and Tet2 proteins

and have been shown to be responsible for the genomic abundance of 5-hmC (Ito et al.,

2010), while Tet3 has been reported to play the major role in conversion of 5-mC into 5-

hmC in the pre-implantation embryo (Gu et al., 2011; Iqbal et al., 2011; Wossidlo et al.,

2011). In mouse EG cells, I found that both proteins were as abundant as in mES cells as

detected by western blotting. As expected TET proteins were not detected in differentiated

mB lymphocytes (Figure 4.8.A). To investigate whether Tet1 and Tet2 proteins were

involved in the acquisition of 5-hmC at somatic ICRs in heterokaryons, I used RNA

interference to separately downregulate Tet1 and Tet2 expression in mouse EG cells

79

Page 80: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

(shTet1 and shTet2). This resulted in efficient depletion of Tet1 and Tet2 proteins,

detected both by Western Blotting and qRT-PCR (Figure 4.8.A-B). Upon downregulation

of either of the genes, pluripotency-associated Oct4 and Nanog gene expression were not

significantly altered, implying that these cells retained pluripotency (Figure 4.8.B).

Following this, together with Francesco Piccolo, we fused shRNA transfected mEG cells

with human B lymphocytes, and quantified the change in the levels of 5-hmC at the ICRs

of H19, Peg3 and SNRPN/SNURF imprinted genes (48h and 72h, Figure 4.8.C). The

analysis demonstrated that Tet1 depletion resulted in lack of 5-hmC acquisition at the

corresponding ICRs compared to the control fusion, whereas Tet2 downregulation did not

result in a significant change (Figure 4.8.C).

These data suggested that imprint erasure was dependent, at least to some extent

on TET1 activity. Interestingly, although TET1 was present in mouse ES cells, we did not

observe 5-hmC accumulation at the ICRs when the B cells were fused with mES cells,

rather than EG cells (Piccolo et al., 2013). It is important to note that, Tet1 depletion in

mEG cells did not however interfere with their pluripotent reprogramming activity (Piccolo

et al., 2013).

80

Page 81: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

Figure 4.8. Roles of TET proteins in the acquisition of 5-hmC at human B lymphocyte ICRs upon fusion with mouse EG cells. (A) Western Blot detection of Tet1 (left), Tet2 (right), Oct4, Nanog and Tubulin in whole cell extracts of mES, mB and mEG cells upon transfection with shControl, shTet1 (left), or shTet2 (right) RNA interference plasmids. Tubulin antibody was used as a loading control (See Materials&Methods for details). (B) qRT-PCR analysis of Tet1 (left), Tet2 (right), Oct4 and Nanog in mEG cells, 72 hours after transfection with either empty (shCtrl, grey bars) or shTet1 (green bars, left) or shTet2 (blue bars, right) plasmids. Data were normalised to mouse Ubc and represent mean and SEM of four to five independent experiments. (C) qPCR analysis of restriction enzyme protection assay (see Materials&Methods) to quantify 5hmC levels at human ICRs before (0h) and at 48 and 72 hours after fusion with mEG cells transfected either by shTet1 (green dots, upper) or shTet2 (blue dots, lower) plasmids compared with shControl (grey dots) transfected mEG cells.

4.8. Summary and Discussion

In this Chapter, by using cell fusion system, I showed that although mES and

mouse EG cells can both efficiently reprogram somatic cells towards pluripotency, only

81

Page 82: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

mouse EG cells possess the ability to induce imprint erasure. This contrast is especially

interesting considering that both cell types are similar at the level of their transcriptomes

(Leitch et al., 2013). It is important to note that, mouse EG cells we used in our analyses

were derived late in the PGC development (embryonic day 12.5). We also discovered that

these cells can induce imprint reprogramming only at early passage number (< passage

30) (Piccolo et al., 2013). Although we do not yet understand the reasons behind this loss

of ability, culture conditions have been shown to lead to major transcriptome and

methylome changes in these cells that may alter their intrinsic properties (Leitch et al.,

2013). In addition, other EG cells isolated earlier during PGC development do not appear

to possess this imprint erasing capability although they can reprogram (Piccolo et al.,

2013).

Pluripotent conversion of somatic nucleus upon fusion with mouse EG cells occurs

early at the heterokaryon stage. Using mB lymphocytes that contain Oct4-GFP transgene

as fusion partners, I was able to obtain GFP positive hybrids as early as 3 days of fusion.

Here I showed that the imprinted genes are “tagged” by 5-mC oxidation at the early stages

of mouse EG cell induced reprogramming and that this is mediated by Tet1 protein.

However, I also determined that Tet1 is present at similar levels in mES cells. This shows

that imprint resetting induced by mouse EG cell fusion cannot be attributed solely to the

presence of Tet proteins. Over the last years, several studies reported that Tet family

members can interact with various proteins or be part of protein complexes. Examples

include interactions with SIN3A (Williams et al., 2011), Mbd3/NURD complex (Yildirim et

al., 2011), NANOG (Costa et al., 2013), O-linked N-acetylglucosamine transferase (Chen

et al., 2013b; Vella et al., 2013), and long non-coding RNA TARID together with Gadd45a

(Arab et al., 2014). It would be important to compare Tet interaction partners in mES and

mouse EG cells, also to do profiling by mass spectroscopy of EG cells at early and late

passages, to shed light on the discrepancy of their imprint resetting potentials.

Investigation of imprint erasure kinetics in the hybrids formed between mouse EG

cells and mB lymphocytes revealed an interesting result. It was not until at least 7 days

after fusion that demethylation at the somatic ICRs was detected that gradually continued

until 21 days after fusion. This delay of imprint erasure compared to pluripotent

reprogramming can be considered as a reminiscent of PGC development in the embryo.

PGC precursors can first be detected in the epiblast at E6.25 by the expression of specific

marker Blimp1 (Ohinata et al., 2005). By E7.5, Oct4 expression is restricted to the newly

emerged PGCs in the embryo and PGCs continue to express Oct4 during their migration

to form the genital ridges. (Scholer et al., 1990; Yeom et al., 1996). In the meantime,

82

Page 83: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 4

waves of genome-wide DNA demethylation accompanies major changes in the chromatin

context that is reminiscent of acquisition of pluripotency (Hajkova et al., 2008). It is

important to note that in PGC development DNA demethylation is a temporally ordered

process; first promoters of pluripotency markers and germ cell specific genes undergo

demethylation, which is later followed by loss of methylation at the ICRs (Guibert et al.,

2012; Seisenberger et al., 2012). These events strikingly resemble our observations that

mouse EG cell mediated pluripotent conversion occurs ahead of imprint reprogramming

in somatic cells after fusion. In addition, it has been reported that PGCs express significant

levels of Tet1 and Tet2, which may contribute to the replication-coupled removal of 5-mC

upon conversion into 5-hmC. 5-hmC is no longer be recognised by the UHRF1 (Hackett

et al., 2013; Hashimoto et al., 2012), therefore 5-hmC is replaced by unmodified cytosine

during DNA synthesis. In-vivo analysis Tet1 knock-out is also shown to result in

abnormalities in genomic imprinting, due to the presence of hypermethylated ICRs

(Yamaguchi et al., 2013b).

83

Page 84: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

Chapter 5. Analysis of TET Protein Requirement in

Mouse Embryonic Stem Cell Induced Reprogramming of

Human B Lymphocytes

5.1. Introduction

In this Chapter, I investigate the importance of TET proteins for reprogramming

efficiencies. I have used two strategies to disturb Tet activity. The first approach (Chapter

5.2), is based on RNA interference. Upon knockdown Tet mRNA levels are decreased in

mES cells, and these cells are then fused to hB lymphocytes. The reprogramming

efficiencies of Tet -sh and -control mES cells are compared. The advantage of this system

is that it can circumvent any potential long-term effects of knockdown of Tets on mES cell

identity. For complete withdrawal of TET activity in mES cells, my second approach

(Chapter 5.3) is based on CRISPR/Cas9 genome editing system. Here I report a detailed

work-flow of activity to generate mutant ES cells and analyse the capacity of these knock-

out lines to reprogram hB lymphocytes to pluripotency.

5.2. Tet Knockdown in mouse ES cells and cell fusion

In order to knockdown Tet1 and Tet2 genes separately or together in mES cells, I

took advantage of RNA interference system using shRNA vectors (Tet1 alone, Tet2 alone

and Tet1/Tet2 double knockdown), as previously demonstrated (Chapter 4) (Ito et al.,

2010; Williams et al., 2011). To determine how the Tet knockdown affected the ability of

mES cells to convert hB cells to pluripotency, I compared gene expressions in

heterokaryons.

5.2.1. Tet1 knockdown and fusion

In order to reduce Tet1 expression levels, I electroporated mES cells with

pSUPER.neo.GFP + Tet1 shRNA vector (shTet1), and used FACS to sort successfully

transfected cells after 24 hours based on GFP expression (Figure 5.1.A). I used the same

FACS settings to sort mES cells transfected with empty vector (shCtrl) for comparison. 72

84

Page 85: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

hours after transfection (at the time of fusion), qRT-PCR showed (grey compared to green)

that Tet1 mRNA levels had decreased to 35% of the original value (Figure 5.1.B).

Interestingly, this was accompanied by a slight increase of Tet2 expression (Figure 5.1.B).

Although Oct4 expression was not altered in these cells, Nanog mRNA levels were

augmented (Figure 5.1.B), a result which has been observed in a previous research study

(Williams et al., 2011). Previously Ito et al. demonstrated that upon Tet1 knockdown,

Nanog expression level was substantially decreased (Ito et al., 2010), but Williams et al.

reported that this was due to off-target effects of the shRNA used in their study (Williams

et al., 2011).

ES cells transfected with shCtrl or shTet1 (72 hours) were fused with hB

lymphocytes, and the changes in gene expression in the human B nuclei were analysed

(Figures 5.1.C-D). As expected, silent human pluripotency genes OCT4, NANOG and

CRIPTO were induced upon fusion with shCtrl transfected mES cells (Figure 5.1.C, grey

bars). Likewise, shTet1 transfected mES cells were able to reprogram hB cells upon

fusion, more efficiently (i.e. better than controls, Figure 5.1.C, green bars compared to

grey wild type). In line with expectations, hB specific gene CD19 expression decreased

upon fusion with either shCtrl or shTet1 transfected mES cells (Figure 5.1.D). The

decrease was slightly more prominent with Tet1 knockdown ES cells (also CD45 gene

downregulation) suggesting that Tet1 withdrawal may enhance reprogramming.

Collectively, these results demonstrate that Tet1 downregulation does not worsen

the reprogramming efficiency in heterokaryons, suggesting that Tet1 does not play a

crucial role in the early stages of cell fusion mediated pluripotent conversion. If anything,

reprogramming was slightly enhanced in the absence of TET1, which might be due to

unknown secondary effects of the knockdown, or because Tet2 and Nanog expressions

were elevated in these targeted mES cell lines. This contrasts with what has been shown

in a previous report (Costa et al., 2013).

85

Page 86: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

Figure 5.1. Effect of Tet1 knockdown on reprogramming. (A) FACS profile of pSUPER.neo.GFP + Tet1 (shTet1) transfected mES cells. The gate was chosen to sort medium/high GFP expressing cells (plots in blue). (B) qRT-PCR analysis of Tet1, Tet2, Oct4 and Nanog in mES cells, 72 hours after transfection with either empty (shCtrl, grey bars) or shTet1 (green bars) plasmids. Data were normalised to mouse Ubc and represent mean and SEM of 4 different experiments. (C) qRT-PCR analysis of pluripotency factor gene expression (OCT4, NANOG and CRIPTO) in hB lymphocytes before (0h) and at 48 and 72 hours after fusion with mES cells transfected with either shCtrl (grey bars) or shTet1 plasmids (green bars). Data were normalised to human GAPDH and represent mean and SEM of 2 different experiments. (D) qRT-PCR analysis of lymphocyte specific factor gene expression (CD19 and CD45) in hB lymphocytes before (0h) and at 48 and 72 hours after fusion with mES cells transfected with either shCtrl (grey bars) or shTet1 plasmids (green bars). Data were normalised to human GAPDH and represent mean and SEM of 2 biological replicates.

86

Page 87: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

5.2.2. Tet2 knockdown and fusion

To knockdown the Tet2 gene in mES cells I used a similar approach, where I

transfected the cells with pSUPER.neo.GFP + Tet2 shRNA vector (shTet2) and sorted

them after 24 hours based on GFP expression (Figure 5.2.A). I used the same FACS gate

settings to sort empty vector (shCtrl) transfected mES cells as a control. qRT-PCR

analysis of the transfected cells demonstrated a decrease in Tet2 gene expression level

compared to shCtrl transfection, without a major change in Tet1, Oct4 and Nanog levels

(Figure 5.2.B).

To determine whether the decrease in Tet2 mRNA level compromises the

reprogramming ability of mES cells, I conducted cellular fusions and analysed the gene

expression in resulting heterokaryons. Using shCtrl transfected mES cells as controls,

(Figure 5.2.C, grey bars), shTet2 transfected mES cells induced a similar level of

pluripotency gene expression in hB cells (Figure 5.2.C, blue bars). In accord with this,

silencing of CD19 gene expression was comparable in both samples, and there was no

major difference in the extinction of human CD45 expression. These data suggest that the

reduced Tet2 mRNA does not impair the reprogramming potential of mES cells, but is

implicated in mEG mediated reprogramming (Chapter 4) and may also have roles in iPS

conversion (Doege et al., 2012).

87

Page 88: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

Figure 5.2. Effect of Tet2 knockdown on reprogramming. (A) FACS profile of pSUPER.neo.GFP + Tet2 (shTet2) transfected mES cells. The gate was chosen to sort medium/high GFP expressing cells (plots in blue). (B) qRT-PCR analysis of Tet1, Tet2, Oct4 and Nanog in mES cells, 72 hours after transfection with either empty (shCtrl, grey bars) or shTet2 (blue bars) plasmids. Data were normalised to mouse Ubc and represent mean and SEM of 4 different experiments. (C) qRT-PCR analysis of pluripotency factor gene expression (OCT4, NANOG and CRIPTO) in hB lymphocytes before (0h) and at 48 and 72 hours after fusion with mES cells transfected with either shCtrl (grey bars) or shTet2 plasmids (blue bars). Data were normalised to human GAPDH and represent mean and SEM of 2 different experiments. (D) qRT-PCR analysis of lymphocyte specific factor gene expression (CD19 and CD45) in hB lymphocytes before (0h) and at 48 and 72 hours after fusion with mES cells transfected with either shCtrl (grey bars) or shTet2 plasmids (blue bars). Data were normalised to human GAPDH and represent mean and SEM of 2 biological replicates.

88

Page 89: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

5.2.3. Tet1 and Tet2 double knockdown and fusion

In order to rule out the possible redundant functions of Tet1 and Tet2 in

reprogramming, it was important to simultaneously decrease their expression in mES

cells. Therefore, I modified the Tet1 RNAi plasmid, by replacing the neo.GFP with mCherry

coding sequence and I co-transfected mES cells with pSuper.mCherry + Tet1 (shTet1)

and pSUPER.neo.GFP + Tet2 (shTet2) vectors. After 24 hours of transfection, I sorted the

cells expressing both fluorescent markers. (Figure 5.3.A). I applied the same settings for

mES cells co-transfected with empty vectors expressing GFP and mCherry (shCtrls). After

72 hours of shTet1&shTet2 co-transfection, qRT-PCR analysis revealed that the

expression levels of Tet1 and Tet2 were both decreased (by 75% and 55% respectively),

Oct4 level remained constant, and Nanog level was slightly increased, all compared to

empty vector co-transfected mES cells (Figure 5.3.B, purple bars –shTet1&shTet2-

compared to grey bars –shCtrls-).

To answer whether knocking down both of Tet1 and Tet2 genes had an effect on

reprogramming, I fused these mES cells with hB lymphocytes, and evaluated the

efficiency of pluripotent conversion by qRT-PCR in heterokaryons. Interestingly, shRNA

mediated decrease of Tet1 and Tet2 expression did not have a negative effect on

reprogramming, as judged by OCT4, NANOG and CRIPTO expression (Figure 5.3.C), or

extinction of human B cell genes (CD19 and CD45). Thus, inhibition of Tet1 and Tet2 does

not interfere with the capacity of mES cells to reprogram hB cells upon fusion. These data

raise the question whether 5-methylcytosine oxidation is necessary for experimental

reprogramming by cell fusion. However, it is noteworthy that RNAi-based systems do not

necessarily reduce protein or enzymatic activities of targeted genes rapidly. Although it is

conceivable that even a slight decrease in expression of a crucial factor would result in a

detectable deficiency of pluripotent induction, complete removal of TET proteins may not

be achieved with this approach. For this reason, in the following section I describe how I

generated Tet Knockout mES cell lines using CRISPR/Cas9 mediated genome editing

system, and compare reprogramming potentials of these lines.

89

Page 90: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

Figure 5.3. Effect of Tet1/Tet2 double knockdown on reprogramming. (A) FACS profile of pSUPER.mCherry+ Tet1 (shTet1) and pSUPER.neo.GFP + Tet2 (shTet2) co-transfected mES cells. The gate was chosen to sort medium/high GFP & mCherry expressing cells (plots in blue). (B) qRT-PCR analysis of Tet1, Tet2, Oct4 and Nanog in mES cells, 72 hours after co-transfection with either empty (shCtrls, grey bars) or shTet1&shTet2 (purple bars) plasmids. Data were normalised to mouse Ubc and represent mean and SEM of 3 different experiments. (C) qRT-PCR analysis of pluripotency factor gene expression (OCT4, NANOG and CRIPTO) in hB lymphocytes before (0h) and at 48 and 72 hours after fusion with mES cells co-transfected with either shCtrls (grey bars) or shTet1&shTet2 plasmids (blue bars). Data were normalised to human GAPDH and represent mean and SEM of 2 different experiments. (D) qRT-PCR analysis of lymphocyte specific factor gene expression (CD19 and CD45) in hB lymphocytes before (0h) and at 48 and 72 hours after fusion with mES cells co-transfected with either shCtrls (grey bars) or shTet1&shTet2 plasmids (purple bars). Data were normalised to human GAPDH and represent mean and SEM of 2 different experiments.

90

Page 91: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

5.3. CRISPR/Cas9 mediated Tet gene editing and cellular fusion

First discovered as a constituent of bacterial adaptive immunity, Type II

CRISPR/Cas system has recently been used to modify genomes of higher species (Sorek

et al., 2013; Terns and Terns, 2014). Simplicity and cost-effectiveness are among the

advantages of CRISPR-based technologies, providing the opportunity to target multiple

genes simultaneously both in-vivo and in-vitro. This feature has substantially shortened

the time it would take with conventional methods (Wang et al., 2013a). The main role of

CRISPR/Cas system is to generate double-strand break at a particular location in the

genome. The resulting DNA damage is then recognized by the cell’s own repair

mechanism which in return acts on the break. There are two main repair pathways, and

the system can be repurposed according to the aim. First pathway is the non-homologous

end joining, an error-prone mechanism, leading to nucleotide insertions or deletions at the

break site. This can be particularly useful to rapidly inactivate genes by causing mutations

that lead to frameshifts if coding sequences are targeted, or can be used to disrupt

particular protein binding sites. On the other hand, the second system is a high-fidelity

repair mechanism, mediated by homologous recombination which can be exploited to

insert, remove, alter or replace specific sequences with a provided template.

In this section I describe generation of mES cell lines by CRISPR/Cas9, where I

concomitantly target Tet1 and Tet2 genes for non-homologous end joining and report data

on their ability to reprogram somatic cells in experimental fusion system.

5.3.1. CRISPR/Cas9 system construction against Tet1 and Tet2 genes and delivery into mES cells

The Type II CRISPR system is a ribonucleoprotein complex and has two

complementary parts. The first part is composed of CRISPR-RNA (crRNA), a 20

nucleotide guide sequence complementary to the DNA, and Trans-activating crRNA

(tracrRNA) which is complementary to the crRNA and is involved in RNA processing. The

second part is the Cas protein, which is an RNA-guided endonuclease (Deltcheva et al.,

2011). Recently Jinek et al. reported the merging of the two RNA components into a single

RNA chimera (single guide RNA –sgRNA-), which targets the Cas endonuclease with

similar efficiency to the DNA (Jinek et al., 2012).

91

Page 92: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

The px330 plasmid contains all the components of Type II CRISPR system (Cong

et al., 2013); a human codon-optimized Streeptococcus pyogenes Cas9 under strong

hybrid chicken beta-actin promoter, and chimeric guide RNA under the control of U6

promoter. I separately cloned Tet1 and Tet2 sgRNAs into px330 plasmids (see Materials

& Methods) (Figure 5.4) that have previously been shown to target Tet genes efficiently

(Wang et al., 2013a; Yang et al., 2013).

Figure 5.4. Tet1 and Tet2 targeting by CRISPR/Cas. Schematic representation of mouse Tet1 and Tet2 genes, and the DNA target sites. Nucleotides in green represent the PAM sequence, necessary for the target recognition by Cas9 endonuclease of Streptococcus pyogenes (-NGG). Nucleotides in blue depict sgRNA targeted DNA sites. Red arrows demonstrate the locations (3 base pairs downstream of PAM) of double strand breaks upon Cas9 endonuclease activity, which are positioned within SacI and EcoRV restriction enzyme recognition sites in Tet1 and Tet2 respectively.

I co-transfected mES cells by electroporation with (1) px330+sgTet1 vector, (2)

px330+sgTet2 vector and finally (3) pH2BmCherry-Ires-Puro vector which enables the

selection of successfully transfected cells by either of two different methods; FACS and

Puromycin drug treatment (see strategy outlined in Figure 5.5.A).

92

Page 93: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

Figure 5.5. Workflow for CRISPR/Cas mediated gene targeting in mouse ES cells. (A) mES cells are co-transfected by electroporation. (H2BmCherry-IRES-Puro expression cassette is under the control of highly active chicken beta-actin promoter. Transfected cells are splitted and either sorted by FACS according to mCherry expression or treated with Puromycin for 2 days after 48 hours of transfection. A total of 32 colonies were picked for further analysis. (B) FACS profile for transfected mES cells. The gate has been chosen to sort cells expressing high levels of mCherry.

Two days after Puromycin treatment, I used one plate (low cell density) to pick up

colonies, and the other plate (high cell density) for Surveyor Assay to determine the

mutation efficiency. I used the third plate to apply FACS on the mES cells that express

high level of mCherry (Figure 5.5.B), and sorted one cell per well of a 96-well plate, where

after one week, 20-25% of the wells contained viable clones. I sorted the rest of the cells

and plated back for 4 days and collected them to conduct Surveyor Assay. In total I

93

Page 94: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

obtained 32 clones (24 from Puromycin treated, 8 from mCherry sorted) to be further

analysed for the mutation identities.

5.3.2. Surveyor Assay for analysis of CRISPR/Cas efficiency

Surveyor Assay helps to analyse the presence of mutations at a specific locus on

the DNA and is based on the Surveyor DNA endonuclease (Transgenomic), which is a

member of CEL nuclease family isolated from celery (Qiu et al., 2004). Surveyor nuclease

has the capacity to recognise and cleave the 3’ end side of any mismatch site in a given

DNA duplex. Its sensitivity allows the examination of small mismatches such as single

nuclear polymorphisms (SNPs). As non-homologous end joining upon CRISPR/Cas

endonuclease activity creates random indels at the target site, various mutations are

expected to be encountered in a transfected population of cells. Following the

amplification of the target site by PCR, denaturation and reannealing create a vast amount

of mismatches, which then can be recognised and cleaved by the Surveyor nuclease, as

illustrated in Figure 5.6.

Figure 5.6. Schematic Representation of Surveyor Assay. As a result of CRISPR/Cas endonuclease activity on a specific target, random mutations are created by non-homologous end repair (Left). Upon PCR, denaturation and annealing of the DNA from a population of targeted cells (Centre), DNA duplexes contain mismatches, which will then be digested by CEL-1 (Right). The cleaved and non-cleaved bands can be distinguished by gel electrophoresis, where the upper band will be the WT bands (un-cleaved) and the lower bands will be the mismatch bands (cleaved). To conduct the Surveyor Assay, I PCR amplified Tet1 and Tet2 targeted loci of

from the DNA of ‘Puromycin Treated’ or ‘mCherry Sorted’ samples (Day 6, Figure 5.5). I

denatured the PCR amplicons at high temperature, and let the emerging single stranded

DNA randomly reanneal by gradually decreasing the temperature (See Materials &

Methods). This results in DNA duplexes with mismatches, and to visualise this I migrated

the samples on agarose gel after treating with Surveyor endonuclease. In non-transfected

wild type cells, as Tet1 and Tet2 amplicons do not contain mismatches (unless there are

naturally occurring SNPs on these particular loci), no cleavage took place (Figure 5.7;

94

Page 95: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

WTs for Tet1 and Tet2, 460 bp and 466 bp respectively). On the other hand, in

CRISPR/Cas transfected cell population, efficient cleavage by Surveyor Endonuclease

was observed, unveiling the presence of mismatches that were resulted in non-

homologous end joining at the target sites (Figure 5.7; ~235 bp and ~225 bp for Tet1,

~252 bp and ~214 bp for Tet2). These results indicated successful targeting of Tet1 and

Tet2 genes as previously demonstrated (Wang et al., 2013a).

Figure 5.7. Surveyor Assay on Tet1 and Tet2 in wild type and CRISPR/Cas targeted mES cells, treated with Puroymcin or mCherry sorted upon co-transfection. Tet1 PCR produces 460 bp amplicon, Tet2 PCR produces 466 bp amplicon. PCR primers were designed to centre the sgRNA target sites in both Tet1 and Tet2 (Wang et al., 2013a). In all conditions, the upper bands are the undigested amplifications (either wild type or no mismatch mutant amplicons –see Figure 5.6-). The lower bands demonstrate the cleaved products of the amplicons containing mismatches upon Surveyor Assay.

It is important to note that cleavage efficiencies in ‘Puromycin Treated’ and

‘mCherry Sorted’ populations were very similar for both targets, so clones isolated from

either of the conditions could be used for further analysis.

5.3.3. Restriction Fragment Length Polymorphism screen on CRISPR/Cas9 targeted mES cells for Tet1 and Tet2

In order to analyse whether single clones that I selected from CRISPR/Cas

targeting (a total of 32; 24 from puromycin selection, 8 from mCherry sorting) had acquired

indels, I conducted a Restriction Fragment Length Polymorphism (RFLP) screen. This

allows an investigation to quickly assess mutated alleles, as the targeted Tet1 and Tet2

loci contain restriction enzyme recognition sites that can be used to identify mutations in

the analysis (Figure 5.4). Disruption of these sites by indels demonstrates successful

error-prone non-homologous end joining upon cleavage by Cas9 (Figure 5.8).

95

Page 96: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

Figure 5.8. Schematic Representation of RFLP. A wild type allele is expected to be digested by restriction enzyme. On the contrary, the disruption of the recognition site by indels, prevents the cleavage of the DNA amplicon. Agarose gel electrophoresis reveals the identity of allelic targeting by CRISPR/Cas9.

The screen of the single clones is based on PCR amplification of the targeted Tet1

and Tet2 loci (same primer couples as in the Surveyor Assay, the uncleaved products are

460 bp and 466 bp for Tet1 and Tet2 respectively) and restriction digestion. The Tet1

sgRNA targets SacI recognition site in exon 4, and the Tet2 sgRNA targets EcoRV

recognition site in exon 3. Double strand breaks are expected to take place in these

recognition sites, which are then corrected by non-homologous end joining that results in

indel acquisition. There are three outcomes of the clonal screen by RFLP, (1) total

cleavage of the PCR product proving that both alleles are wild type, (2) partial cleavage

of the PCR product showing that one allele is wild type and one allele is mutated, (3) no

cleavage of the PCR product, demonstrating that the both alleles are likely mutated.

As shown in Figure 5.9, wild type ES cell samples show PCR amplicons for both

Tet1 and Tet2 targeted loci that have been efficiently cleaved by the corresponding

restriction enzymes (Figure 5.9; left WT column). Almost all of the co-targeted clones

contained bi-allelic mutations at the Tet1 and Tet2 target sites as evidenced by the lack

of restriction digestion of the amplicons (Figure 5.9). This observation is in line with the

efficiency determined by the Surveyor Assay at the population level.

96

Page 97: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

Figure 5.9. RFLP Assay on WT and Tet1&Tet2 CRISPR co-targeted 32 clones. Tet1 PCR produces 460 bp amplicon, Tet2 PCR produces 466 bp amplicon. For RFLP on the Tet1 target locus, the PCR amplicons were digested with Sac1 restriction enzyme, which in WT condition results in bands at sizes of ~235 bp and ~225 bp. For RFLP on the Tet2 target locus, the PCR amplicons were digested with EcoRV restriction enzyme, which in WT condition results in bands at sizes of ~252 bp and ~214 bp. PCR amplicon sizes may vary due to the nature of acquired indels.

5.3.4. Sequencing of Tet1&Tet2 CRISPR targeted ES clones

Acquired indels upon error prone non-homologous end-joining may have various

outcomes. As the exons are targeted, base insertion or base deletion may cause a frame-

shift in the coding sequence. Alternatively, these modifications may lead to in-frame

mutations, in which case the protein activity may or may not change, depending on the

length of alteration in the DNA sequence. To characterise the indel identities of the

97

Page 98: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

CRISPR targeting, I sequenced the Tet1 and Tet2 target loci in 6 clones (5A, 5B, 5C, 8A,

8B, 8E). To do so, I cloned the PCR amplicons into pGEM®-T Easy Vector (Promega),

which I transformed into competent bacteria. Per transformation, I used at least 8 bacterial

colonies for sequencing to cover both alleles (Figure 5.10.A-B).

Figure 5.10. DNA sequencing results on Tet1&Tet2 CRISPR co-targeted loci. (A) Table demonstrating the size of indels in each allele in corresponding clones. Purple colour indicates frame-shift acquisition in one allele, green colour indicates frame-shift acquisition in both alleles, and red colour indicates in-frame mutations in both alleles. (B) DNA sequences of the WT (blue) and targeted clones. Green nucleotides in WT indicate the restriction sites SacI and EcoRV in Tet1 and Tet2 respectively. Deletions are indicated as hyphens, insertions are indicated in red, and the indel sizes are indicated on the right for each sequence.

98

Page 99: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

In line with the RFLP screen, sequences of Tet1 and Tet2 CRISPR target loci in all

of the 6 clones underwent error prone repair upon double strand break by Cas9 which

distorted the restriction enzyme recognition sites (Figure 5.10.B, nucleotides in green in

WTs). DNA sequencing revealed various genotypes among 6 clones which may lead to

different phenotypes in the ES cells. For example clone 8B contains indels that result in

bi-allelic frame-shifts both in Tet1 and Tet2 genes, causing early termination of translation

by the stop codons in the new frames. On the other hand, in clones 5A and 5B this is the

case only in one of the alleles of Tet1 and Tet2 genes, while second alleles for these

genes acquired an in-frame mutation.

The majority of the observed indels are deletions, and it is interesting to note that

some mutations reoccur in different clones, which may indicate the presence of micro-

homology directed repair in these loci (McVey and Lee, 2008). Overall, the presence of

mutations in all of the sequenced and screened clones suggests a high efficiency of

CRISPR/Cas co-targeting using the protocol described here.

5.3.5. Reprogramming capacity of CRISPR/Cas9 mediated Tet1 and Tet2 mutant ES cell clones upon cell fusion

As a result of CRISPR/Cas9 mediated gene editing, I obtained ESC clones with

various genotypes for Tet1 and Tet2 as previously described. Clone 5A contains one

frame-shift and one in-frame allele for both Tet1 and Tet2 genes, while in the clone 8B

both alleles for both genes acquired frame-shift mutations which are expected to abrogate

TET1 and TET2 activity. To understand how these changes affect reprogramming

potential of ES cells, I conducted cell fusions in which I fused human B cells with clones

5A or 8B or wild type cells. I analysed human gene expression in heterokaryons formed

after 48 and 72 hours of fusion. As expected, pluripotency genes OCT4, NANOG and

CRIPTO that were not expressed in human B cells at the time of fusion but were induced

in heterokaryons upon fusion with WT mES cells (Figure 5.11.A). Interestingly at 48 hours,

both 5A and 8B clones were able to induce the expression of these factors either as

efficiently as or more efficiently than compared to WT cells. However even though WT

heterokaryons were further reprogrammed at 72 hours of fusion, pluripotency gene

expression levels of clone 5A and clone 8B heterokaryons remained unchanged (Figure

5.11.A). In addition, human B specific CD19 and CD45 genes were downregulated to

99

Page 100: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

comparable levels upon fusion with WT, clone 5A and clone 8B ES cells at 48 and 72

hours (Figure 5.11.B).

Figure 5.11. Effect of Tet1/Tet2 knockout on reprogramming. (A) qRT-PCR analysis of pluripotency factor gene expression (OCT4, NANOG and CRIPTO) in hB lymphocytes before (0h) and at 48 and 72 hours after fusion with WT mES cells, Clone 5A and Clone 8B. Data were normalised to human GAPDH and represent mean and SEM of 2 different experiments. (D) qRT-PCR analysis of lymphocyte specific factor gene expression (CD19 and CD45) in hB lymphocytes before (0h) and at 48 and 72 hours after fusion with WT mES cells, Clone 5A and Clone 8B. Data were normalised to human GAPDH and represent mean and SEM of 2 different experiments.

These results demonstrate that partial or complete ablation of Tet activity do not

hinder ES cells’ capacity of reprogramming B cells upon fusion in heterokaryons.

However, the fact that pluripotent gene expression does not appear to be increasing

between 48 and 72 hours of fusion raises questions. This might be due to defects in

stabilisation of pluripotent induction, or due to the stress the clones have been through in

the course of CRISPR/Cas editing.

5.4. Summary and Discussion

Developing mammalian embryos undergo two waves of genome-wide DNA

demethylation, occurring soon after fertilization and during PGC development. Since the

discovery of TET proteins as catalytic modulators of 5-mC to 5-hmC conversion (Tahiliani

et al., 2009; Kriaucionis and Heintz, 2009), many studies have focused on their roles in

embryonic development, as 5-hmC has been suggested to be an intermediate of

100

Page 101: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

replication-dependent and –independent loss of methylation (Pastor et al., 2013). These

waves of DNA demethylation lead to in-vivo acquisition of pluripotency, and it is crucial to

have a better understanding of how lineage identity is reset.

In this chapter, I have not only shown that TET is dispensable in experimental cell

fusion based reprogramming assay, but also described the CRISPR/Cas system as an

effective gene knock-out tool. Firstly, RNA interference-mediated Tet1 and Tet2

knockdown (separately or together), did not interfere with the capacity of mouse ES cells

to reprogram hB lymphocytes. Since RNA interference does not completely abolish mRNA

level of the target gene, I also created Tet1 and Tet2 knock-out mES cell lines using the

CRISPR/Cas system. Interestingly, these ES cells were as successful as wild type cells

in initiating B cell reprogramming, but the induction of pluripotent gene expression did not

develop fully. Reprogramming was initiated in these cells, apparently even in the presence

of cytosine methylation at OCT4, NANOG and CRIPTO genes. This observation deserves

more attention, and I will return to it later in this document.

Recent in-vitro reprogramming strategies have strived to elucidate the contribution

of Tet proteins to pluripotent conversion. These have however, resulted in different and

often opposing findings. The involvement of Tet activity in iPS cells was first reported by

Doege et al., where they reported an increased global 5-hmC distribution upon factor

induced (Oct4, Sox2, Klf4 and c-Myc –OSKM-) reprogramming of mouse embryonic

fibroblasts (Doege et al., 2012). Further analysis revealed 5-hmC enrichment in

endogenous pluripotency genes (such as Nanog an Esrrb), accompanied by gene

expression. This is in line with the early activation of endogenous Tet2 expression (by day

3 of induction), while interestingly Tet1 remained silent. In addition, Tet2 downregulation

upon shRNA knockdown completely abrogated iPS cell colony formation, suggesting that

TET2 protein is responsible for the global and locus specific DNA hydroxylation, necessary

for successful reprogramming (Doege et al., 2012). In a similar approach, Costa et al.

postulated that Tet1 knockdown inhibited factor-induced iPS cell generation from mouse

embryonic fibroblasts (Costa et al., 2013). Moreover, upon fusion with hB cells, Tet1

Knock-out ES cells were less efficient in their ability to reprogram compared to WT ES

cells. Finally, the authors demonstrated that overexpression of TET1 or TET2 proteins

enhanced iPS cell formation, mainly via downstream effects of physical interaction with

NANOG (Costa et al., 2013). A third study revealed that Tet1 expression was gradually

increased during iPS cell reprogramming (while Tet2 levels remained relatively stable)

and Tet1 downregulation by RNA interference abolished colony formation (Gao et al.,

2013). These investigators showed that in the course of reprogramming, TET1

101

Page 102: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

hydroxylated the methylated Oct4, leading to its demethylation and subsequent activation,

and that TET1 could substitute OCT4, in the OSKM factor cocktail (Gao et al., 2013). A

fourth approach demonstrated that, during human iPS cell reprogramming 5-hmC levels

were significantly increased, mediated mainly by the induced expression of endogenous

TET1 gene; its knockdown resulted in a decreased number of reprogrammed colonies

(Wang et al., 2013b). Another study analysed TET function modulation by Vitamin C

during pluripotent conversion of MEFs, and concluded that Tet1 deficiency resulted in

enhanced reprogramming in the presence of, and reduced reprogramming in the absence

of Vitamin C (Chen et al., 2013a). This was interesting as Vitamin C has previously been

shown to enhance Tet activity (Blaschke et al., 2013; Yin et al., 2013). The authors

concluded that cooperative action of Tet1 and Vitamin C stands as a barrier of

mesenchymal-to-epithelial transition in the course of reprogramming (Chen et al., 2013a).

A recent study has provided data that conflicts with all previous reports. It

demonstrated that Tet1&Tet2 and Tet1&Tet3 double mutant MEFs were successfully

reprogrammed into pluripotent colonies. However, simultaneous depletion of all Tet

constituents completely abolished this conversion, which the authors ascribed to a lack of

hydroxylation of methylated miR-200 family gene promoters, followed by demethylation

(Hu et al., 2014). In wild type cells, these activities are exerted by redundant Tet function

in cooperation with thymine deglycosylation by TDG (Hu et al., 2014). By narrowing down

the Tet requirement in factor-induced reprogramming into a miRNA family, this study has

raised doubt as to the necessity of Tet activity in reactivating pluripotency genes in the

course of reprogramming. In view of this, I have examined whether DNA demethylation of

OCT4 is a prerequisite for pluripotent reprogramming.

DNA demethylation was noted in iPS cell creation (Wernig et al., 2007) and was

also shown to be a perquisite for Oct4 reactivation in nuclear transfer-mediated

reprogramming studies (Simonsson and Gurdon, 2004). It is noteworthy that other

epigenetic markers undergo global changes in pluripotent conversion. For example,

genome-wide remodelling is evident in very early stages of factor induced conversion in

somatic cells, where active histone marks such as H3K4 methylation are rapidly

accumulated in pluripotency and early development genes (Koche et al., 2011). In

addition, our lab has previously shown that upon fusion, various chromatin features

change markedly, such as histone acetylation levels (Piccolo et al., 2011) and

heterochromatin marker HP1α is rapidly redistributed in the somatic hB nucleus (Pereira

et al., 2010). I re-evaluated these findings by using mouse ES cells expressing H2B

tagged with mCherry fused with human fibroblasts expressing HP1α tagged with GFP

102

Page 103: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

(Figure 5.12). Hp1α is a nuclear protein and its distribution in the somatic fibroblast is

punctuate (Figure 5.12, top left), marking the heterochromatic loci, and the H2B

distribution is nuclear in mouse ES cells as expected from a histone protein (middle). In

heterokaryons, the mouse-derived nucleus (here mouse ES) can be distinguished from

human derived nucleus (here human fibroblast) on the basis of DAPI staining profile

(Pereira et al., 2008) (Figure 5.12, bottom right). Imaging reveals that upon fusion with

mES cell, HP1α punctuate profile is lost by fibroblast nuclei (white arrow), but mouse ES

cell nuclei (red arrows) in the same heterokaryon show a punctate H2B distribution (Figure

5.12, bottom left white and red arrows, respectively). Similarly, H2BmCherry originating

from mouse nucleus accumulates in the human fibroblast partner.

Figure 5.12. HP1α redistribution in mESxhF heterokaryons. Top row, human IMR90 fibroblast expressing Hp1αGFP fusion protein. Middle row, mES cells expressing H2BmCherry, Bottom row, a heterokaryon with two mES cell nuclei (demonstrated by red arrows) and one human fibroblast nucleus (demonstrated by white arrow), distinguished by the DAPI staining profile. Right column, merge image of GFP and mCherry as well as DAPI (blue) which stains nuclear DNA and Phalloidin (yellow) which stains filamentous actin. Scale bar = 10µm.

103

Page 104: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

This result recapitulates previous findings that HP1α is re-distributed in somatic nuclei

upon fusion, and demonstrates the global alteration in the epigenetic landscape at the

early stages of reprogramming. Using tagged proteins of this sort allows live cell imaging

of chromatin remodelling events in heterokaryons. Considering the global change in

chromatin signatures, is it possible that pluripotency genes can be re-expressed even in

the presence of DNA methylation? Interestingly a precedent for this has previously been

reported during C/EBPα-induced B cell to macrophage transdifferentiation; where silent

macrophage-specific genes were activated, even though their promoters remained highly

methylated (Rodríguez-Ubreva et al., 2012). In addition, these genes were shown to

acquire H3K4 trimethylation (H3K4me3) as well as H3K9/H3K14 acetylation (features of

active transcription) all of which are active histone marks. With this in mind, the authors

have suggested that repressive effect of DNA methylation may be overcome by additional

epigenetic factors such as active histone modifications (Rodríguez-Ubreva et al., 2012).

I have previously shown that by day 3, heterokaryon populations still show high

levels of DNA methylation at the human OCT4 promoter although the OCT4 gene

expression is detected as early as 24 hours after fusion. To explore whether a locus-

specific histone remodelling is occurring in these cells, I used chromatin precipitation

(ChIP) analysis. This is technically difficult as the number of heterokaryons generated after

fusion is low and can be masked by the presence of unfused cells in these cultures. To

avoid this, I used H2BmCherry expressing mouse ES cells and dye labelled human B cells

(CellTrace Violet, Invitrogen) as fusion partners, and I enriched double-labelled

heterokaryons by FACS. I also adapted a low cell ChIP protocol (see Materials and

Methods) to look for H3K4me3 enrichment at the human OCT4 promoter locus, in both

human B cells and heterokaryons using human specific primers. Preliminary ChIP data

demonstrated that H3K4me3 was not present at the OCT4 promoter in human B cells, in

line with the gene’s repressed status, but was highly enriched at the same region in

heterokaryons (Figure 5.13). Although these are preliminary data, it is important that

acquisition of H3K4me3 is detected at the OCT4 promoter of the somatic nucleus upon

reprogramming. It will be essential to expand this analysis in future to see whether other

active histone marks have also been incorporated, and to assess whether other

pluripotency gene loci are also remodelled early after cell fusion. Eventually we hope to

conduct a genome-wide ChIP sequencing analysis of these changes (for discussion see

Chapter 7).

104

Page 105: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 5

Figure 5.13. H3K4me3 ChIP on human B lymphocytes and Day 3 heterokaryons. H3K4me3 is analysed in human B cells and heterokaryons at β-ACTIN as positive control, at OCT4 and at TSH2B (testis-specific histone 2B) as negative control. H3Kme3 data are normalized to total H3, and as negative control, IgG data are normalized to input.

Overall my results, and examples from the literature, support a rather unexpected

model of gene re-activation upon reprogramming. DNA methylation has been regarded as

a relatively stable modification (due to the presence of strong Carbon-Carbon bond

between the cytosine and the methyl group). In a compact chromatin structure where DNA

methylation and repressive histone marks coexist, histone tails could be viewed as being

more physically and chemically accessible, which may render them prone to modification

by external cues. Accumulation of active marks on histone tails may lead to recruitment of

transcription factors that could kick-start low levels of gene transcription. This may be

followed by DNA demethylation, required to stabilise gene expression. As reprogramming

in heterokaryons is considered fast and efficient, as compared to other systems, it might

be difficult to accurately order the sequence of these events. With developing technology

and sensitive detection systems, single cells have increasingly been used to avoid the

loss of information due to averaging. In the future, such analysis could be done on single

heterokaryons that would provide valuable information on the nature of reprogramming.

105

Page 106: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

Chapter 6. CRISPR/Cas Editing of Jarid2 and Non-

Canonical WNT Pathway Components

6.1. Introduction

Jarid2 (Jumonji, AT rich interactive domain 2) is a component of the Polycomb

Repressor Complex 2 (PRC2) in ES cells and has been implicated in regulating

pluripotency networks in embryonic stem cell differentiation (Landeira and Fisher, 2011).

In addition, recent work by David Landeira and colleagues in our laboratory has

demonstrated that Jarid2 deficient mES cells (Shen et al., 2009) possess elevated levels

of Nanog expression (Landeira et. al, submitted) and are extremely good at

reprogramming through cell fusion (Pereira et al., 2010). Analysis of publically available

mES cell ChIP datasets revealed a lack of JARID2 binding at the Nanog promoter,

suggesting that Nanog was indirectly regulated by JARID2. To identify factors that were

differentially expressed in JARID2 deficient mES cells, gene expression profiling was

performed, and several Wnt signalling components were determined to be significantly

de-regulated. Among these were Prickle1 and Fzd2 (downregulated in the absence of

Jarid2) involved in non-canonical Wnt pathway and Wnt9a, all of which were confirmed to

be directly bound by Jarid2 (Landeira et. al, submitted). To understand whether loss of

these non-canonical pathway components could phenocopy the Jarid2 deficiency in

mouse ES cells, I took advantage of loss-of-function approach by using CRISPR/Cas9

gene editing system. In this Chapter, I describe CRISPR design process for Jarid2,

Prickle1, Fzd2 and Wnt9a, and caution their subsequent modification in mouse ES cells.

6.2. CRISPR/Cas9 editing of Jarid2 and Prickle1/Fzd2/Wnt9a in mouse embryonic stem cells

In Chapter 5, I demonstrated how CRISPR/Cas system can be used to modify Tet1

and Tet2 genes using published sgRNA sequences (Wang et al., 2013a). In this Chapter

I describe de-novo design of targeting sgRNAs to establish knockout mES cell lines for

Jarid2 (single) and for Prickle1, Fzd2 and Wnt9a (triple). The process of design is

especially important in view of the undesired off-target potential of CRISPR/Cas9 system.

106

Page 107: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

6.2.1. Guide RNA design and delivery into mouse ES cells

CRISPR/Cas9 DNA targeting generally leads to acquisition of random mutations

as a result of double strand break followed by error-prone non-homologous end joining

pathway. When an exon of a gene is targeted, acquired mutations can cause frameshifts

in the downstream coding sequence. This results in impaired protein function, an early

stop codon, or degradation. For this reason it is wise to target the start of the coding

sequence, paying attention to the presence of splice variants and alternative start codons

(which can be visualised at the UCSC Genome Browser available at

http://genome.ucsc.edu/).

In order to determine guide RNA sequences for targeting Jarid2, Prickle1, Fzd2

and Wnt9a, I used a bioinformatics tool which is available online at http://crispr.mit.edu/

(Hsu et al., 2013). This tool ranks candidate sgRNAs based on the number of potential

off-targets. Furthermore, I performed analysis of Genome-Wide Tag Scan (Iseli et al.,

2007) to evaluate the identities of the possible off-targets (See Chapter 6.4 for discussion).

I therefore eliminated all the sgRNA sequences with a high number of potential off-targets,

as well as those with off-targets falling into intragenic regions. Using the best candidates,

I constructed the px330 plasmid for the expression of sgRNA along with Cas9

endonuclease (Figure 6.1).

Figure 6.1. Jarid2, Prickle1, Fzd2 and Wnt9a targeting by CRISPR/Cas. Schematic representation of mouse Jarid2, Prickle1, Fzd2 and Wnt9a genes, and the DNA target sites. Nucleotides in red represent the PAM sequence, necessary for the target recognition by Cas9 endonuclease of Streptococcus pyogenes (-NGG). Nucleotides in blue depict sgRNA targeted DNA sites. Red arrows demonstrate the locations (3 base pairs downstream of PAM) of double strand breaks upon Cas9 endonuclease activity. Green arrows represent the transcription start sites. Sequences are not to scale.

107

Page 108: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

Here, the aim is to generate mES cells deficient for JARID2 as a control, and mES

cells simultaneously deficient for PRICKLE1, FZD2 and WNT9a. For this reason, I either

transfected mES cells with px330+sgJarid2 or co-transfected with px330+sgPrickle1,

px330+sgFzd2 and px330+sgWnt9a, along with pH2BmCherry-Ires-Puro vector for

selection. I used the same workflow depicted in Figure 5.5, to obtain samples for Surveyor

Assay and single colonies (32 colonies in each case).

6.2.2. Surveyor Assay for analysis of CRISPR/Cas9 efficiency

To determine whether the designed guide RNAs can efficiently target Cas9 to the

loci of interest and whether this results in random acquisition of mutations, I conducted

Surveyor Assay at population level. Strong appearance of cleavage products of the

targeted Jarid2 locus in mES cells indicated successful indel acquisition (Figure 6.2, left).

Similarly, sgRNA sequences tested for Prickle1, Wnt9a and Fzd2 efficiently targeted

corresponding loci in co-transfected mES cells (Figure 6.2, right).

Figure 6.2. Surveyor Assay on Puromycin treated populations of Jarid2 single and Prickle1, Wnt9a, Fzd2 triple CRISPR/Cas targeted mES cells. Jarid2 PCR produces 503 bp amplicon, Prickle1 PCR produces 499 bp amplicon, Wnt9a PCR produces 500 bp amplicon and Fzd2 PCR produces 488 bp amplicon. Respective cleavage product lengths are: ~203 and ~300, ~224 and ~275, ~302 and ~198, ~284 and ~204. In all conditions, the upper bands are the undigested amplifications (either wild type or no mismatch mutant amplicons).

6.2.3. Clonal screens and sequencing for targeted Jarid2 locus in mouse ES cells

Jarid2 CRISPR/Cas target location does not contain a restriction enzyme cut site.

For this reason, instead of using RFLP, I conducted Surveyor Assay at the clonal level.

108

Page 109: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

To do so, I mixed wild type DNA together with DNA derived from each clone in 1 to 1 ratio.

A major drawback of this screen (contrary to RLFP) is that the presence of cleavage

products does not necessarily demonstrate whether the mutations are mono- or bi-allelic.

The Surveyor Assay showed that the majority of the clones acquired mutations at the

targeted Jarid2 locus, with at least one allele modified (Figure 6.3).

Figure 6.3. Surveyor Assay on CRISPR/Cas targeted single mES cell clones for Jarid2. The assay was conducted on both wild type, and 32 clones selected upon Jarid2 targeting. Jarid2 PCR produces 503 bp amplicon with cleavage products at ~203 and ~300 bp.

To characterise the identities of the random mutations and determine those which

cause frame-shift at the coding sequence of Jarid2, I sequenced the target locus in eight

of the clones (Figure 6.4).

Figure 6.4. DNA sequencing results on Jarid2 CRISPR/Cas targeted locus. DNA sequences of the WT (blue) and targeted clones. Deletions are indicated as hyphens, insertions are indicated in red, and the indel sizes are indicated on the right for each sequence.

Sequencing of the Jarid2 target locus revealed that none of the selected clones

contained a wild type allele. Furthermore most of the acquired mutations caused shift in

109

Page 110: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

the frame of the coding sequence. Interestingly, one of the clones (Clone 9G) contained

three different alleles, which was most probably not due to a karyotypic instability, but due

to the concept of mosaicism. This occurs when the transfected single cell undergoes cell

division before its DNA could be targeted, which results in two genetically different

populations in the same colony.

To demonstrate the effect of frame-shift mutations, together with Amélie Feytout

we analysed the presence of JARID2 protein in three selected clones (9D, 11C and 12C)

by Western Blotting. The lack of full size JARID2 protein in the mutant clones confirmed

efficient CRISPR/Cas targeting in these ES cell lines (Figure 6.5).

Figure 6.5. Western Blot detection of Jarid2 in wild type and CRISPR/Cas targeted clones. Western blotting is conducted on whole-cell extracts of wild type mouse ES cells, and Clones 9D, 11C and 12C, using antibodies against Jarid2, and Lamin B as control.

6.2.4. Clonal screens and sequencing for targeted Prickle1, Fzd2 and Wnt9a loci in mES cells.

I selected 32 colonies from triple targeted mES cells, and screened those clones

for presence of mutations. The Fzd2 target locus contains a Tsp45i restriction enzyme

recognition site, allowing me to conduct RFLP screen on the clones (Figure 6.6).

Figure 6.6. Surveyor Assay on CRISPR/Cas triple targeted single mES cell clones for Fzd2. The assay was conducted on both wild type, and 32 clones selected upon triple targeting. Fzd2 PCR produces 488 bp amplicon with cleavage products at ~284 and ~204 bp, indicated by white arrows (wild type, left).

110

Page 111: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

Even though PCR amplification was not observed in some of the, many others

contained mono- or bi-allelic mutations at the Fzd2 target locus (demonstrated by the lack

of cleavage) (Figure 6.6). To further screen the clones for mutations on Prickle1, and

Wnt9a, I selected 13 clones where Fzd2 gene was modified in both alleles, and I

conducted Surveyor Assay (Figure 6.7.A-B).

Figure 6.7. Surveyor Assay on CRISPR/Cas triple targeted single mES cell clones for Prickle1 and Wnt9a. (A) Surveyor Assay on Prickle1 target locus on wild type (right) and 13 clones. Prickle1 PCR produces 499 bp amplicon with cleavage products at ~224 and ~275 bp. (B) Surveyor Assay on Wnt9a target locus on wild type (right) and 13 clones. Wnt9a PCR produces 500 bp amplicon with cleavage products at ~302 and ~198 bp.

These results revealed that majority of the selected 13 clones contained mutations

at both Prickle1 and Wnt9a target loci (Figure 6.7.A-B). However, it is not known whether

these mutations are mono- or bi-allelic, as this cannot be distinguished by the Surveyor

Assay. To determine the mutations, I went on to sequence co-targeted three loci in five

selected clones (Figure 6.8). Sequencing results revealed that almost all co-targeted sites

acquired indels. Interestingly, 3 of the 5 clones appeared to be mosaic for the targeted

Wnt9a locus (depicted by the presence of more than 2 alleles). Collectively, around two

thirds of the acquired indels resulted in frame-shifts, however it is difficult to predict

whether some of the in-frame mutations would cause a conformational change in the

111

Page 112: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

protein and decrease its functionality (such as 18 base deletions at the Allele 3 of targeted

Wnt9a in Clone 3A; Figure 6.8).

Figure 6.8. DNA sequencing results on Prickle1, Fzd2 and Wnt9a CRISPR/Cas co-targeted loci. DNA sequences of the WTs (blue) and targeted clones. Deletions are indicated as hyphens, insertions are indicated in red, and the indel sizes are indicated on the right for each sequence.

Because antibodies to PRICKLE1, WNT9A and FZD2 are not available, I was not

able to determine the effect of frame-shift mutations at the protein level. Instead, together

with Amélie Feytout we used an alternative approach to visualise gene knock-out. The

shift in the coding frame may lead to appearance of premature stop codons causing early

termination of translation. In turn, this affects the stability of mRNA at the translational

level, resulting in degradation. This concept is known as nonsense mediated decay

(Losson and Lacroute, 1979). We conducted transcript analysis for the targeted genes,

and showed that mutations in Prickle1 cause efficient decrease of mRNA levels. We also

observed this effect in Fzd2 and Wnt9a transcripts, yet not as efficiently as in the case of

Prickle1. This can be explained by the fact that Fzd2 and Wnt9a coding sequence sizes

are much smaller than Prickle1, possibly providing stability and prevention from decay

(Figure 6.9).

112

Page 113: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

Figure 6.9. mRNA levels of CRISPR/Cas targeted Prickle1, Fzd2 and Wnt9a in selected clones and wild type cells. qRT-PCR analysis of Prickle1, Fzd2, and Wnt9a expression in clones 2A, 2D, 3A, 3C, 3H and wild type parental cells. Data were normalised to Hmbs and represent mean and SEM of 3 biological replicates. Asterisks indicate statistical significance (p<0.05; Student’s t-test).

6.3. JARID2 deficiency in mouse ES cells can be phenocopied by Prickle1/Fzd2/Wnt targeting

Recent study by David Landeira (Landeira et. al, submitted) showed that non-

canonical Wnt pathway components were downregulated in previously established Jarid2

deficient mES cells (Shen et al., 2009). To investigate whether I can reproduce these data

in mES cells targeted by CRISPR/Cas9 for Jarid2, together with Amelie Feytout we have

conducted transcript analysis. We analysed the gene expression levels of Prickle1, Fzd2

and Wnt9a in three of the Jarid2 mutant mES cell clones (9D, 11C and 12C).

Figure 6.10. mRNA levels of Prickle1, Fzd2 and Wnt9a in Jarid2 mutant lines and wild type cells. qRT-PCR analysis of Prickle1, Fzd2, and Wnt9a expression in Jarid2 deficient clones 9D, 11C and 12C. Data were normalised to Hmbs and represent mean and SEM of 3 biological replicates. Asterisks indicate statistical significance (p<0.05; Student’s t-test).

113

Page 114: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

Compared to the parental wild type mES cells, Jarid2 mutants showed decreased

expression levels of Prickle1, Fzd and Wnt9a, recapitulating previous analysis (Figure

6.10).

Mouse ES cells that lack Jarid2 contain increased level of Nanog expression

(Landeira et. al, submitted). Based on this observation, together with Jorge-Soza Ried,

we analysed Nanog levels in mES cell clones 12C (Jarid2 mutant) and 2D (non-canonical

Wnt pathway mutant), both generated upon CRISPR/Cas9 editing. Flow cytometry

analysis demonstrated that CRISPR/Cas9 mediated Jarid2 deficiency resulted in

increased number of Nanog-high cells, in agreement with previously observed data

(Figure 6.11, left). Interestingly, knocking-out non-canonical Wnt pathway components

also lead to elevated number of Nanog-high cells in the population (Figure 6.12), as also

observed by immunofluorescence (data not shown). This suggests that Jarid2 might be

controlling steady state level of Nanog expression in wild type mES cells via non-canonical

Wnt pathway.

Figure 6.11. Flow Cytometry analysis of Nanog expression in mES cells Clones 12C and 2D. Flow cytometry analysis demonstrating Nanog expression at the population level in mES cell clones targeted for Jarid2 (12C) or Prickle1/Fzd2/Wnt9a (2D) (green traces) compared to wild-type parental ES cells (filled grey). The black line depicts negative control.

Furthermore, by collaborating with Transgenics Facility and Karen Brown, we

analysed individual contribution of parental wild-type cells and clones 12C and 2D to the

developing embryo. We injected wild-type blastocysts with either parental wild-type,

mutant Clone 12C, or mutant Clone 2D ES cells and cultured for 16 hours before

inspection. Remarkably, we observed that blastocysts injected with 12C or 2D clones

initiated the formation of more than one ICM (in around 40% of the blastocysts for either

of the clones), which was not detected in any of the control blastocysts injected with

parental wild-type cells (Landeira et. al, submitted).

114

Page 115: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

Collectively these data suggest that Jarid2 might be regulating Wnt signalling

components that in turn control steady state levels of Nanog expression and proper

development of the pre-implantation embryo.

6.4. Summary and Discussion

As a tool of genome engineering, CRISPR/Cas9 system has gained extensive

popularity due to its easiness, accuracy and efficiency. These features have enabled

scientists to simultaneously target multiple genes in a very short time, both for in-vivo and

in-vitro studies. In this chapter, I have not only demonstrated CRISPR/Cas9 design

process for DNA targeting, but also shown efficient generation of knock-out mES cell lines

for single (Jarid2) and multiple (Prickle1/Fzd2/Wnt9a) genes.

A major concern on CRISPR/Cas9 system has been the extent of potential off-

target mutagenesis it may cause. Cas endonuclease is brought on to the DNA by

complementation of 20-nucleotide long guide RNA and a study on bacterial genome has

revealed that mismatches outside the 12-base ‘seed region’ can be tolerated for targeting

(Jiang et al., 2013). This rises the chances of encountering off-target mutagenesis, even

though DNA cleavage will occur only if a PAM sequence is situated on the downstream of

the target sequence (-NGG for Cas9 from Streptococcus pyogenes). For this reason it is

important to use bioinformatics tools that help to rank guide RNA sequences according to

the possession of least off-targets, which preferentially fall into inter-genic non-conserved

areas in the genome (Hsu et al., 2013). I designed my targeting sequences according to

this ranking, and conducted a genome-wide homology search by using the ‘seed region’

to identify off-target locations (Iseli et al., 2007) Figure (6.12). This analysis demonstrated

that designed guide RNAs possessed limited number of off-targets, none of which

targeted a coding or conserved region.

As off-target potential is a primary drawback of the CRISPR/Cas9 system, many

studies have strived to minimize this effect. Feng Zhang lab has developed a ‘double

nicking’ system, where Cas9 nickase mutant was used to create single strand break upon

targeting. DNA editing is possible only a second Cas9 nickase mutant is targeted to the

complementary sequence at the same location. With this strategy, single strand off-target

breaks can be repaired with high fidelity by base-excision repair (Ran et al., 2013b). In

addition, Shen and colleagues have demonstrated that double nicking has significantly

reduced the number of double strand off-target cleavage compared to wild type Cas9, as

115

Page 116: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

shown by decreased H2AX recruitment to the cut sites (Shen et al., 2014). Development

of such strategies are crucial, especially for translational usage of CRISPR/Cas9 system.

Figure 6.12. Off-target identification of CRISPR/Cas9 targets. Table showing guide RNA sequences targeting Jarid2, Prickle1, Fzd2 and Wnt9a loci together with in-silico identified off-targets and their corresponding genomic coordinates.

Analysis of published ChIP-seq data for Jarid2 binding in mES cells has revealed

that Jarid2 has a strong preference for binding on Wnt signalling and Wnt-related pathway

components (Pasini et al., 2010). Among these are Prickle1, Fzd2 and Wnt9a, also bound

by Ezh2, which were downregulated in Jarid2 knock-out mES cells (Landeira et. al,

submitted) and upon JARID2 depletion by CRISPR/Cas9. This suggests that Polycomb

group proteins might be positively regulating active transcription of these particular genes,

as part of a recently reported feature that involves PRC2-dependent H3K27 mono/di-

methylation (Ferrari et al., 2014). It is important to note that Nanog gene is not bound by

JARID2, yet Nanog levels were significantly increased in mES cells that lack JARID2. My

results upon Prickle1/Fzd2/Wnt9a depletion suggest that this indirect regulation might

pass through non-canonical Wnt signalling pathway.

How does Wnt signalling control Nanog expression? NANOG is a core element of

pluripotency network and its expression fluctuates in mES cells leading to a bimodal

distribution (Chambers et al., 2007). mES cells expressing constitutively high levels of

Nanog are associated with naïve pluripotency, and this can be achieved by the usage of

Mitogen-activated protein kinase kinase (MEK) and Glycogen Synthase Kinase 3 (Gsk3)

116

Page 117: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 6

inhibitors (known as 2 inhibitors; 2i) (Ying et al., 2008). When not inhibited, Gsk3 leads to

the proteolysis of β-catenin (Aberle et al., 1997), a key component of Wnt signalling that

has been suggested to play major roles in pluripotent self-renewal. Upon Wnt activation,

β-catenin is accumulated in the nucleus where it interacts with Tcf/Lef (T-cell

factor/lymphoid enhancer factor) transcription factors and it has been reported that Gsk3

inhibition diminishes Tcf3 repression exerted on pluripotency network (Wray et al., 2011).

In addition, Tcf3 ablation has been shown to replace the requirement of Gsk3 inhibition in

terms of self-renewal (Yi et al., 2011). In line with these observations, activation of

canonical Wnt signalling reduces the levels of TCF3 which results in increased Nanog

levels and defects in differentiation (Atlasi et al., 2013). This is mainly because TCF3 acts

as a direct repressor of Nanog gene expression, and it is believed that this action provides

steady-state levels of NANOG for the maintenance of differentiation potential of

pluripotent cells (Pereira et al., 2006). It has been suggested that non-canonical Wnt

pathway may be negatively regulating the canonical pathway through the Tcf/Lef

transcription factors (Kühl et al., 2000). Additionally, Osei-Sarfo et al. demonstrated that

retinoic acid mediated mES cell differentiation lead to the activation of non-canonical Wnt

pathway, while canonical pathway is inhibited. This resulted in significant accumulation of

TCF3 on the promoters of pluripotency-associated genes and their subsequent repression

(Osei-Sarfo and Gudas, 2014).

Collectively, my data support a view that JARID2 might be regulating the interplay

between the canonical and non-canonical Wnt pathways that controls the self-renewal

and pluripotency of mES cells. Chromatin immunoprecipitation experiments would unravel

whether Nanog upregulation in Jarid2 deficiency is due to Tcf3 dissociation, regulated by

Prickle1/Fzd2/Wnt9a downregulation. Tcf3 displacement can occur upon β-catenin

mediated phosphorylation by homeodomain interacting protein kinase 2 (HIPK2) (Hikasa

et al., 2010), or by direct interaction with β-catenin (Solberg et al., 2012). Alternatively,

Tcf3 expression might be regulated at the expression level by changes in Wnt Signalling

(Atlasi et al., 2013).

117

Page 118: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

Chapter 7. General Discussion

In mammals, CpG methylation has been associated with gene repression.

Although the mechanisms of DNA methylation (both de-novo and maintenance) are well

characterised, how this is reversed in-vivo remained elusive. Recent studies have started

to unravel how DNA demethylation occurs initially by focusing on biological processes

such as pre-implantation embryo development and germ cell specification. In addition,

investigators have studied DNA demethylation that occurs during in-vitro pluripotent

reprogramming. It is now widely accepted that loss of DNA methylation takes place by two

distinct, but interconnected pathways: (1) passive, DNA replication-dependent

demethylation, and (2) active, DNA replication-independent demethylation. Passive

demethylation is based on a gradual dilution of 5-mC that occurs with successive rounds

of DNA replication and cell division when the methylation maintenance machinery is

disabled. Active demethylation involves enzymatic activity that modifies 5-mC, and

ultimately results its replacement by an unmodified cytosine residue. Here I have focused

on DNA methylation/demethylation dynamics during pluripotent reprogramming, where

DNA methylation has been considered to pose a major roadblock. Cell type specific DNA

methylation needs to be re-set for the acquisition of pluripotency (Pasque et al., 2011). In

this regard, cell fusion-based reprogramming provides a tractable experimental platform

to investigate the first signs of pluripotent conversion (within newly formed heterokaryons)

as the pluripotency-associated transcriptional programme is initiated.

7.1. DNA methylation dynamics in reprogramming

Our laboratory and others have demonstrated that pluripotent stem cells can

dominantly reprogram somatic cells upon fusion [(Soza-Ried and Fisher, 2012) and

recapitulated in this thesis]. These studies have suggested that the direction of conversion

(dominance) can be predicted and probably reflects the action of trans-acting factors that

also maintain the ‘stemness’ of pluripotent cells. Pluripotent cells possess so-called “open

chromatin” structure, which may allow cells to rapidly react to external differentiation

signals (Meshorer and Misteli, 2006). This property is maintained by global chromatin

remodellers that mediate histone tail modifications, nucleosome positioning and

reorganization, and interact pluripotency-associated network components (Gaspar-Maia

118

Page 119: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

et al., 2011). Interestingly, mouse ES cells treated with histone deacetylase (HDAC)

inhibitors were shown to be more efficient in reprogramming upon fusion with somatic cells

(Hezroni et al., 2011). In addition, ES cells that lack PRC activity have reprogramming

defects, consistent with a role for chromatin modifiers in pluripotent conversion (Pereira et

al., 2010).

Following differentiation, lineage restriction normally ensures the stable repression

of genes, particularly those involved in pluripotency. During this process, nucleosome-

depleted regions in mouse ES cells may be subject to nucleosome assembly followed by

acquisition of DNA methylation (You et al., 2011). In Chapter 3 of this study, I confirmed

the hypermethylated status of pluripotency-related OCT4, NANOG and CRIPTO genes in

differentiated human B lymphocytes and in human fibroblasts. Intriguing, although these

genes are rapidly induced upon cell fusion and reprogramming, bisulfite sequencing of the

promoter region of the OCT4 gene in heterokaryons did not reveal any significant loss of

DNA methylation. This could be attributed to the technical limitations (for example, 5-hmC

cannot be differentiated from 5-mC by bisulfite sequencing), or to the presence of

unreprogrammed heterokaryons within the population. Alternatively, loss of DNA

methylation may not be required for the initial activation of OCT4 gene expression. In

mouse, Oct4 gene expression and silencing is tightly controlled during embryogenesis

and silencing occurs via a series of events; repressor binding results in G9A-mediated H3

lysine 9 methylation that recruits HP1 that is eventually followed by de-novo DNA

methylation (Feldman et al., 2006). Gene reactivation may occur through a sequential

reversal of these events; histone remodelling followed by DNA demethylation. Upon

reprogramming by cell fusion, it is possible that the chromatin structure is relaxed and

modifications may help initiate OCT4 gene expression. A rapid global increase in H3K9

acetylation and H4 pan-acetylation have been reported as a distinguishing feature of

somatic cell-reprogramming by cell fusion (Piccolo et al., 2011). In addition, my preliminary

results (Chapter 5) suggest that active H3 lysine 4 tri-methylation is rapidly acquired on

the OCT4 gene promoter of somatic nuclei upon fusion with mouse ES cells. It has also

been shown that similarly, C/EBPα mediated pre-B cell conversion into macrophages

initiates macrophage specific gene expression programme with histone modification while

DNA methylation is maintained (Rodríguez-Ubreva et al., 2011). This could perhaps also

explain why the reprogramming potential of mouse ES cells depleted of TET proteins is

unchanged (at least at early stages), as gene induction may not be dependent on the

removal of DNA demethylation (as demonstrated in Chapter 5). It is noteworthy that a

recent study reported that TET activity was necessary in iPS cells only for the induction of

119

Page 120: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

miR-200 family members (which then facilitates the transition from mesenchymal to

epithelial state), but not for reactivation of pluripotency genes (Hu et al., 2014).

Tet proteins, on the other hand, may be required for mouse EG cell mediated

imprint erasure in mouse B lymphocyte genome upon fusion, (as shown in Chapter 4).

Here I showed that in mEG-hB heterokaryons TET activity resulted in 5-hmC accumulation

at imprinted control region DMRs, yet it is not until at least 7 days of fusion that I detected

DNA demethylation. This suggests an orchestrated action of both active and passive

demethylation processes; where actively hydroxylated 5-mC at DMRs may be passively

diluted with continuous cell replication. This mechanism may also be operating in-vivo

during PGC development, where TET enzymatic action and replication-coupled loss of

methylation has been evoked (Hackett et al., 2013).

7.2. Genome editing and the use of CRISPR/Cas9-based approaches

Genetic engineering is the process of targeted modification of genetic material.

Genome editing is usually achieved by homologous recombination (HR) in which an

externally provided DNA molecule serves as a template. The low efficiency of this event

(Capecchi, 1989) was shown to substantially increase by providing a DNA double-strand

break (DSB) at the target region. This occurs because the cell’s intrinsic DNA repair

mechanism can recognise the externally provided DNA as a template for correction via

homology-directed repair (HDR) pathway (Rouet et al., 1994). In the absence of a repair

template, non-homologous end joining pathway (an error-prone repair mechanism),

ligates cleaved DNA ends by inducing insertion and deletion mutations (indels) at the DSB

site (Moore and Haber, 1996). Both repair pathways have been extensively repurposed

for genome editing, however the major challenge has been to deliberately induce site-

specific DSBs among billions of DNA bases of the eukaryotic genome. For this purpose,

in the last decade four main strategies have been developed which exploit DNA-binding

proteins; meganucleases (Smith et al., 2006b), zinc finger nucleases (Urnov et al., 2005),

transcription activator-like effector nucleases (Miller et al., 2011) and recently, RNA-

guided Cas9 endonuclease, originating from microbial acquired immune system CRISPR

(clustered regularly interspaced short palindromic repeats) (Cong et al., 2013; Mali et al.,

2013). Thanks largely to basic research on bacteria, the CRISPR system has now been

re-purposed for gene editing, and in the last couple of years a plethora of studies have

proven its efficacy, despite initial off-target effects (see below). In chapters 5 and 6 of this

120

Page 121: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

document, I used CRISPR/Cas9 to generate a panel of mouse ES cell lines in which

Jarid2, Tet1/Tet2 and Prickle1/Fzd2/Wnt9a were targeted.

CRISPR originates in studies performed in the 1980s, when Atsuo Nakata and

colleagues encountered mysterious “unusual” repeats in Escherichia Coli while

sequencing the iap gene (Ishino et al., 1987). In the following years these repeats were

discovered in various prokaryotes, however it was not until 2002 that they were united

under the acronym CRISPR (Jansen et al., 2002). Briefly, CRISPR sequences are

composed of direct repeats of 21 to 37 base pairs with a loose dyad symmetry, and are

interspaced by similarly sized variable sequences (spacers) (Jansen et al., 2002). In

addition, adjacent to CRISPR loci, Jansen et. al. identified the presence of CRISPR-

associated (cas) genes which possess nuclease motifs, suggesting a possible functional

relation with the repeats (Jansen et al., 2002). Although well conserved among species,

at that time biological significance of CRISPR and Cas genes had remained elusive. Later,

several in-silico approaches identified spacers as originating from foreign DNA elements,

which led scientists to hypothesise that they may be involved in acquired immunity against

bacteriophages. Finally in 2007, direct evidence came from Horvath laboratory, where the

authors demonstrated that CRISPR is a prokaryotic resistance mechanism against

invading DNA molecules mainly originating from viruses, which involves recognition,

destruction and adaptation (Barrangou et al., 2007; Sorek et al., 2013). Studies on

Streptococcus thermophiles revealed three basic components of Type II CRISPR system

(Type I and Type III have slightly different mechanisms). These are Cas9 (or Cas5)

endonuclease that mediates DNA cleavage (Garneau et al., 2010), CRISPR-RNA (crRNA)

transcribed and processed from CRISPR array and trans-activating crRNA (tracrRNA) that

forms a hybrid with crRNA to guide Cas9 for targeting to the homologous DNA sequence

(Deltcheva et al., 2011). In-vitro studies demonstrated that Type II components can

efficiently be used to target and cleave plasmid DNA and to further simplify the system,

single guide RNA (sgRNA) was generated upon fusion of crRNA with tracrRNA (Jinek et

al., 2012) (Figure 7.1). Soon after, two studies simultaneously reported engineering of

Type II CRISPR system from Streptococcus pyogenes for mammalian genome

manipulation (Cong et al., 2013; Mali et al., 2013).

121

Page 122: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

Figure 7.1. Schematic representation of RNA-guided Cas9 targeting on DNA. The Cas9 endonuclease from Streptococcus Pyogenes (in beige) is recruited to the genomic DNA (in black) by sgRNA which is generated by fusion of 20-nt guiding crRNA (in purple) and scaffolding tracrRNA (in green). The guide RNA forms a heteroduplex with the DNA directly upstream of the 5’-NGG motif (Protospacer Adjacent Motif, PAM), which is a strict requirement for the Cas9 nuclease activity. Red triangles show the DSB locations mediated by Cas9, located at the 3-bp upstream of the PAM sequence.

High-resolution structural investigations showed that RNA binding (crRNA-

tracrRNA complex, or sgRNA) is necessary for structural rearrangement and consequent

activation of Cas9 to exhibit DSB upon targeting (Jinek et al., 2014; Nishimasu et al.,

2014). Single molecule imaging using DNA curtains demonstrated that RNA-guided Cas9

first interacts with Protospacer Adjacent Motif (PAM, Figure 7.1) upon random collisions

along the DNA (Sternberg et al., 2014). PAM recognition is followed by DNA strand

separation and sequential extension of the guide RNA-DNA heteroduplex starting from

the PAM (Sternberg et al., 2014). This mechanism also explains the importance of 8-12

nucleotide-long seed sequences on the downstream of the guide RNA where any

nucleotide mismatch would terminate the heteroduplex formation, while upstream

mismatches can be tolerated (Jiang et al., 2013; Sternberg et al., 2014).

In the last couple of years, numerous laboratories have conducted CRISPR/Cas9

mediated genome manipulation in both cell lines and animal models and the number of

such studies have been expanding at a dazzling pace (Sander and Joung, 2014). One of

the biggest advantage of this technique is the possibility to simultaneously edit multiple

genes -a process that would take a long time using conventional methods. For example

Rudolph Jaenisch and colleagues simultaneously disrupted five genes (Tet1, Tet2, Tet3,

Sry and Uty) in mouse ES cells based on indel acquisition upon NHEJ (Wang et al.,

2013a), and rapidly generated reporter cell lines by HDR (Yang et al., 2013). In addition,

upon zygotic injection of Cas9 mRNA and sgRNAs, the authors have been able to

generate knock-in or knock-out mice without any need for further breeding (Wang et al.,

2013a; Yang et al., 2013). Moreover, it has been possible to correct genetic diseases in

human cells (Schwank et al., 2013) and in mouse zygotes (Long et al., 2014; Wu et al.,

122

Page 123: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

2013b) and new reports demonstrated the proof-of-principle that CRISPR/Cas9 system

can be used in adult mice by hydrodynamic tail injection of the components (Xue et al.,

2014; Yin et al., 2014). Alternatively, catalytically inactive Cas9 can be used to alter

epigenetic signature of a targeted site upon tethering to transcriptional regulators, to

activate (Maeder et al., 2013; Perez-pinera et al., 2013), or to repress (Gilbert et al., 2013)

gene expression. In addition, lentiviral sgRNA libraries enabled genome-wide knock-out

screens (Shalem et al., 2014; Wang et al., 2014).

One major concern in CRISPR based gene editing has been the potential of

unwanted off-target mutagenesis as mismatches in target recognition by guide RNA can

be tolerated to some extent. Several studies have reported that undesired modifications

can occur at different sites in the genome with higher levels than expected (Fu et al., 2013;

Hsu et al., 2013; Pattanayak et al., 2013; Shen et al., 2014). However recent studies

demonstrated that careful design of guide RNA sequence (Cho et al., 2014) can

significantly reduce off-target potential as analysed by whole-genome sequencing in

human pluripotent ES and IPS cells (Smith et al., 2014; Suzuki et al., 2014; Veres et al.,

2014). These studies caution that careful analysis and design are required to minimise the

adverse off-target effects of CRISPR targeting.

Here, I used CRISPR/Cas9 to generate single (Jarid2), double (Tet1/Tet2) and

triple (Prickle1/Fzd2/Wnt9a) knock-out mouse ES cell lines. In-vitro and in-vivo analyses

showed that Prickle1/Fzd2/Wnt9a knock-out mouse ES cells phenocopied the effects of

Jarid2 depletion in terms of altered blastocyst development and aberrant cell sorting.

Using these novel cell reagents, I intend to extend these studies and examine the roles of

non-canonical Wnt pathway components in embryogenesis in future studies.

7.3. Future Studies

To dissect the early molecular events that are required for reprogramming I will

explore two main approaches in the future. The first approach is based on microfluidic

systems and aims to establish a high-throughput platform for heterokaryon generation.

This will be important to allow us to conduct epigenomic (ChIP for histone modifications

and transcription factor binding) studies that require a relatively high number of fused cells.

The second approach is based on non-destructive imaging of single heterokaryons to

elucidate the kinetics and order of events underpinning chromatin reorganisation, and

DNA demethylation during reprogramming.

123

Page 124: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

7.3.1. Droplet-based microfluidics and cell fusion

Although cell fusion-based reprogramming is an efficient process, the major

limitation of conventional fusion protocols has been the low number of 1:1 heterokaryons

obtained. In fusion with PEG, electric pulse or virus-mediated, cell pairing is random, and

the resulting population contain a heterogeneous mixture of both cell types. This

eventually leads to unwanted fusion products as a result of pairing of same kinds of cells.

Fused cells ofthen have to be selected by using chemical selection, or by cell sorting using

flow cytometry. Therefore, the protocol is time consuming and yields only small numbers

of viable heterokaryons of the desired type. To find a solution we decided to look into

microfluidic systems to increase fusion efficiency by improving the cell pairing and

membrane fusion in a high-throughput manner.

Microfluidics is the science of fluids confined to a miniature scale and has been

used to analyse small sample volumes in drug discovery, medical diagnostics, genomics,

molecular biology and high-throughput screening [with the idea of miniaturising a

laboratory into a chip (Whitesides, 2006)]. Microfluidic platforms have recently been

repurposed to precisely control and manipulate cells in microenvironments leading to next-

generation living-cell microarrays (Yarmush and King, 2009). In addition to many

applications, innovative microfluidic designs have been tested in the past for efficient cell

fusion. One study reported a microfluidic electrofusion system, where cell conjugation was

achieved by biotin-streptavidin coating and paired cells were flowed in a specific

microfluidic channel under continuous direct current, which consisted of narrow and wide

sections. The field intensity at the centre of the narrow channel enabled paired cells to

fuse during their passage (Wang and Lu, 2006). Electrofusion has been used in various

microfluidic formats, including micro-electrode arrays (Cao et al., 2008; Hu et al., 2011;

Qu et al., 2011), micro-orifices (Kimura et al., 2011) and in several other platforms (Hu et

al., 2013). An alternative approach has been the fabrication of thousands of microscaled

polydimethyl siloxane traps into a flow-through channel to increase the pairing efficiency.

The traps were designed to capture both cell type that are sequentially loaded into the

chip. Once the cells were paired and immobilized, PEG or electric field were applied for

fusion (Dura et al., 2014; Kemna et al., 2011; Skelley et al., 2009). However as yet high-

throughput reprogramming has not been established. Droplet-based microfluidic platforms

offer the ability to process millions of individual assays in very short times and high

reproducibility (Huebner et al., 2008). These systems are briefly based on combining two

immiscible phases (water and oil) by segmented-flow, where shear force and interfacial

124

Page 125: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

tension spontaneously generate microdroplets at rates of up to several kHz (Huebner et

al., 2008).

Together with Dr. Andrew deMello’s research team we have begun to generate

prototype droplet-based high-throughput cell fusion systems. This involves encapsulation

of either of the fusion partners in droplets (Köster et al., 2008) followed by droplet merging

(Niu et al., 2008), sorting droplets that contain both cell types (Baret et al., 2009) and

passing them through electric field to induce cell fusion. Although cell encapsulation

follows Poisson statistics (Köster et al., 2008), it is possible to defeat stochastic loading

by engineering alternative micro-channel designs to ensure proper cell ordering (Edd et

al., 2008; Hur et al., 2010; Kemna et al., 2012). Together with ultra-high-speed droplet

generation and cell encapsulation, it would be possible to process millions of events only

in a short period of time, significantly increasing the number of cell fusion events. This

could provide us with a bespoke of “pure” heterokaryon population for molecular analyses.

One of these is chromatin immunoprecipitation experiments to determine the

initiation and the kinetics of reprogramming at the chromatin level. Although I have been

able to demonstrate a trend of H3K4me3 acquisition at the human OCT4 promoter in B

cells upon fusion with mouse ES cells, it would be of interest to analyse further histone

modifications in various loci, and eventually to conduct ChIP-sequencing at a global level.

Furthermore, analysis of how mouse specific pluripotency-associated factors originating

from mouse ES cells bind to and act on the human genome of the somatic cell upon fusion

would shed light on how pluripotent conversion is induced. Specifically, our laboratory has

previously shown that mouse Oct4 protein is very rapidly accumulated on the somatic

DNA as early as 6 hours after fusion. ChIP experiments on mouse Oct4 would provide a

valuable spatio-temporal information on the establishment of pluripotency network at the

somatic DNA. It is noteworthy that recent advances in ChIP sensitivity suggest that it may

be possible to conduct similar experiments using relatively smaller number of (fused) cells

(Lara-Astiaso et al., 2014).

A second approach for which high number of heterokaryons would be needed is

the analysis of the chromatin environment of the newly replicated DNA in heterokaryons.

In our laboratory we have shown that cell cycle stages of the fusion partners affect the

efficiency of reprogramming, and require DNA replication by the somatic nucleus

(Tsubouchi et al., 2013). By using 5-ethynyl-2’-deoxyuridine (EdU) labelling of newly

replicated DNA, together with click-chemistry, EdU containing DNA can be pulled-down

(with the surrounding chromatin intact) so that we can characterise the first epigenetic

events in cell-type conversion (Kliszczak et al., 2011).

125

Page 126: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

7.3.2. Single-cell heterokaryon analysis

Although maximizing number of “purified” heterokaryons is a prerequisite for such

techniques, recent developments have enabled valuable information to be gained from

studying single cells. Therefore, during the process of establishing a fully functional

microfluidics based high-throughput cell fusion system, I will also be conducting single cell

transcription and DNA methylation analyses on heterokaryons.

Microfluidics based transcription analysis arrays (Fluidigm) have recently been

used to measure expression levels of tens of genes in hundreds of single cells in parallel.

One example is by Guo et. al. where the authors investigated the differentiation process

in embryogenesis (Guo et al., 2010). They analysed mRNA levels of 48 genes on single

cells starting from 1-cell zygote and have been able to characterize three distinct cell types

in 64-cell blastocysts according to their expression profiles (Guo et al., 2010). Similar

setting was later used in iPS cell reprogramming by Rudolph Jaenisch and colleagues,

where the analysis of 48 genes in single cells helped to identify the order of events during

pluripotent conversion (Buganim et al., 2012). According to that study, reprogramming first

starts as a stochastic event and single cells exhibit great heterogeneity which is then

followed by a hierarchical phase, where activation of key pluripotency factors lead to a

predictable series of events for the acquisition of pluripotency (Buganim et al., 2012).

Single cell mRNA investigation of heterokaryons will provide informative data on the

kinetics of gain of pluripotency-associated and loss of somatic-cell-specific gene

expression profiles. This will help me determine the timing of key events, and identify the

fundamentals of reprogramming. In addition, as the pluripotent conversion is much quicker

compared to iPS cell reprogramming, by using RNAi or CRISPR based approaches we

can easily scan for necessary factors and their consecutive roles upon cell fusion. By

combining single cell microarrays with next-generation sequencing it is now possible to

conduct RNA-sequencing with decreasing complexity and cost (Kalisky et al., 2011;

Shapiro et al., 2013). This will deliver a global view on how single heterokaryons behave

at the transcriptional level, and lead to the identification of the concept of cellular

dominance.

With recent advances in miniaturisation, it is now possible to analyse the status of

DNA methylation at a single cell level. The first example of this technique came from Axel

Schumacher laboratory, where a micro-reaction slide was used to conduct PCR based

methylation-sensitive restriction assay to detect DNA methylation levels at particular loci

on single cells (Kantlehner et al., 2011). A similar approach was used to simultaneously

126

Page 127: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Chapter 7

investigate methylation levels of 6 imprinted genes in each single cell of 8-cell pre-

implantation embryos, in order to determine how the loss of Trim28, involved in DNMT1

recruitment, can result in defects in imprint maintenance (Lorthongpanich et al., 2013).

Based on this idea, I will conduct single-heterokaryon DNA methylation analysis on

important pluripotency loci, to analyse the kinetics of DNA demethylation. Comparison of

single cell gene expression and spatiotemporal methylation levels will give information of

epigenetic dynamics of reprogramming, which is not easy to obtain by averaging the

population. In addition, I will be able to screen for candidate factors thought to be directly

or indirectly involved in DNA demethylation in a robust and controllable system.

Furthermore, in the long term, single cell genome-wide bisulfite sequencing can be

conducted which can presently detect DNA methylation at almost half of the CpG sites

(Smallwood et al., 2014). Finally, imaging of heterokaryons can give clues on the real-time

kinetics of reprogramming. As discussed (in Chapter 5), I have conducted live cell imaging

on heterokaryons generated between mouse ES cells expressing H2BmCherry and

human fibroblasts expressing HP1αGFP and demonstrated fusion of distinct nuclei

(Cantone et al, submitted). I hope that using these cell lines and microfluidic platforms, I

will be able to apply single-cell live imaging to study hundreds of cells in a single chip

(Kellogg et al., 2014). These approaches will aim to identify the initial early events that

dictate the conversion of one cell type into another, and provide a platform that allows

these events to be interrogated and tested.

127

Page 128: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Bibliography

Aberle, H., Bauer, a, Stappert, J., Kispert, a, and Kemler, R. (1997). Beta-Catenin Is a Target for the Ubiquitin-Proteasome Pathway. EMBO J. 16, 3797–3804.

Ahmed, K., Dehghani, H., Rugg-Gunn, P., Fussner, E., Rossant, J., and Bazett-Jones, D.P. (2010). Global chromatin architecture reflects pluripotency and lineage commitment in the early mouse embryo. PLoS One 5, e10531.

Arab, K., Park, Y.J., Lindroth, A.M., Schäfer, A., Oakes, C., Weichenhan, D., Lukanova, A., Lundin, E., Risch, A., Meister, M., Dienemann, H., Dyckhoff, G., Herold-Mende, C., Grummt, I., Niehrs, C., and Plass, C. (2014). Long Noncoding RNA TARID Directs Demethylation and Activation of the Tumor Suppressor TCF21 via GADD45A. Mol. Cell 55, 604=614.

Arita, K., Ariyoshi, M., Tochio, H., Nakamura, Y., and Shirakawa, M. (2008). Recognition of hemi-methylated DNA by the SRA protein UHRF1 by a base-flipping mechanism. Nature 455, 818–821.

Atlasi, Y., Noori, R., Gaspar, C., Franken, P., Sacchetti, A., Rafati, H., Mahmoudi, T., Decraene, C., Calin, G. a, Merrill, B.J., and Fodde, R. (2013). Wnt signaling regulates the lineage differentiation potential of mouse embryonic stem cells through Tcf3 down-regulation. PLoS Genet. 9, e1003424.

Avvakumov, G. V, Walker, J.R., Xue, S., Li, Y., Duan, S., Bronner, C., Arrowsmith, C.H., and Dhe-Paganon, S. (2008). Structural basis for recognition of hemi-methylated DNA by the SRA domain of human UHRF1. Nature 455, 822–825.

Azuara, V., Perry, P., Sauer, S., Spivakov, M., Jørgensen, H.F., John, R.M., Gouti, M., Casanova, M., Warnes, G., Merkenschlager, M., and Fisher, A.G. (2006). Chromatin signatures of pluripotent cell lines. Nat. Cell Biol. 8, 532–538.

Bagci, H., and Fisher, A.G. (2013). DNA demethylation in pluripotency and reprogramming: the role of tet proteins and cell division. Cell Stem Cell 13, 265–269.

Balasubramani, A., and Rao, A. (2013). O-GlcNAcylation and 5-Methylcytosine Oxidation: An Unexpected Association between OGT and TETs. Mol. Cell 49, 618–619.

Bannister, A.J., and Kouzarides, T. (2011). Regulation of chromatin by histone modifications. Cell Res. 21, 381–395.

Baret, J.-C., Miller, O.J., Taly, V., Ryckelynck, M., El-Harrak, A., Frenz, L., Rick, C., Samuels, M.L., Hutchison, J.B., Agresti, J.J., Link, D.R., Weitz, D.A., and Griffiths, A.D. (2009). Fluorescence-activated droplet sorting (FADS): efficient microfluidic cell sorting based on enzymatic activity. Lab Chip 9, 1850–1858.

Barlow, D., Stöger, R., and Herrmann, B. (1991). The mouse insulin-like growth factor type-2 receptor is imprinted and closely linked to the Tme locus. Nature 349, 84-87.

Barrangou, R., Fremaux, C., Deveau, H., Richards, M., Boyaval, P., Moineau, S., Romero, D. a, and Horvath, P. (2007). CRISPR provides acquired resistance against viruses in prokaryotes. Science 315, 1709–1712.

Barreto, G., Schäfer, A., Marhold, J., Stach, D., Swaminathan, S.K., Handa, V., Döderlein, G., Maltry, N., Wu, W., Lyko, F., and Niehrs, C. (2007). Gadd45a promotes epigenetic gene activation by repair-mediated DNA demethylation. Nature 445, 671–675.

Bartolomei, M., Zemel, S., and Tilghman, S. (1991). Parental imprinting of the mouse H19 gene. Nature 351, 153-155

128

Page 129: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Van den Berg, D.L.C., Snoek, T., Mullin, N.P., Yates, A., Bezstarosti, K., Demmers, J., Chambers, I., and Poot, R. a (2010). An Oct4-centered protein interaction network in embryonic stem cells. Cell Stem Cell 6, 369–381.

Bernstein, B.E., Mikkelsen, T.S., Xie, X., Kamal, M., Huebert, D.J., Cuff, J., Fry, B., Meissner, A., Wernig, M., Plath, K., Jaenisch, R., Wagschal, A., Feil, R., Schreiber, S.L., and Lander, E.S. (2006). A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 125, 315–326.

Bestor, T., Laudano, a, Mattaliano, R., and Ingram, V. (1988). Cloning and sequencing of a cDNA encoding DNA methyltransferase of mouse cells. The carboxyl-terminal domain of the mammalian enzymes is related to bacterial restriction methyltransferases. J. Mol. Biol. 203, 971–983.

Bhattacharya, D., Talwar, S., Mazumder, A., and Shivashankar, G.V. (2009). Spatio-temporal plasticity in chromatin organization in mouse cell differentiation and during Drosophila embryogenesis. Biophys. J. 96, 3832–3839.

Bhattacharya, S.K., Ramchandani, S., Cervoni, N., and Szyf, M. (1999). A mammalian protein with specific demethylase activity for mCpG DNA. Nature 397, 579–583.

Bhutani, N., Brady, J.J., Damian, M., Sacco, A., Corbel, S.Y., and Blau, H.M. (2010). Reprogramming towards pluripotency requires AID-dependent DNA demethylation. Nature 463, 1042–1047.

Bhutani, N., Decker, M.N., Brady, J.J., Bussat, R.T., Burns, D.M., Corbel, S.Y., and Blau, H.M. (2013). A critical role for AID in the initiation of reprogramming to induced pluripotent stem cells. FASEB J. 27, 1107–1113.

Bird, A. (1980). DNA methylation and the frequency of CpG in animal DNA. Nucleic Acids Res. 1499–1504.

Bird, A. (2002). DNA methylation patterns and epigenetic memory. Genes Dev. 16, 6–21.

Blackledge, N.P., Zhou, J.C., Tolstorukov, M.Y., Farcas, A.M., Park, P.J., and Klose, R.J. (2010). CpG islands recruit a histone H3 lysine 36 demethylase. Mol. Cell 38, 179–190.

Blackledge, N.P., Farcas, A.M., Kondo, T., King, H.W., McGouran, J.F., Hanssen, L.L.P., Ito, S., Cooper, S., Kondo, K., Koseki, Y., Ishikura, T., Long, H.K., Sheahan, T.W., Brockdorff, N., Kessler, B.B., Koseki, H., and Klose R.J. (2014). Variant PRC1 Complex-Dependent H2A Ubiquitylation Drives PRC2 Recruitment and Polycomb Domain Formation. Cell 157, 1445–1459.

Blaschke, K., Ebata, K.T., Karimi, M.M., Zepeda-Martínez, J. a., Goyal, P., Mahapatra, S., Tam, A., Laird, D.J., Hirst, M., Rao, A., Lorincz, M.C., and Ramalho-Santos, M. (2013). Vitamin C induces Tet-dependent DNA demethylation and a blastocyst-like state in ES cells. Nature 500, 222-226.

Blau, H.M., Chiu, C.P., and Webster, C. (1983). Cytoplasmic activation of human nuclear genes in stable heterocaryons. Cell 32, 1171–1180.

Blau, H.M., Pavlath, G.K., Hardeman, E.C., Chiu, C.P., Silberstein, L., Webster, S.G., Miller, S.C., and Webster, C. (1985). Plasticity of the differentiated state. Science 230, 758–766.

Borgel, J., Guibert, S., Li, Y., Chiba, H., Schübeler, D., Sasaki, H., Forné, T., and Weber, M. (2010). Targets and dynamics of promoter DNA methylation during early mouse development. Nat. Genet. 42, 1093–1100.

Borst, P., and Sabatini, R. (2008). Base J: discovery, biosynthesis, and possible functions. Annu. Rev. Microbiol. 62, 235–251.

Bostick, M., Kim, J.K., Estève, P.-O., Clark, A., Pradhan, S., and Jacobsen, S.E. (2007). UHRF1 plays a role in maintaining DNA methylation in mammalian cells. Science 317, 1760–1764.

129

Page 130: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Boyer, L.A., Lee, T.I., Cole, M.F., Johnstone, S.E., Levine, S.S., Zucker, J.P., Guenther, M.G., Kumar, R.M., Murray, H.L., Jenner, R.G., Gifford, D.K., Melton, D.A., Jaenisch, R., and Young, R.A. (2005). Core transcriptional regulatory circuitry in human embryonic stem cells. Cell 122, 947–956.

Boyer, L.A., Plath, K., Zeitlinger, J., Brambrink, T., Medeiros, L. a, Lee, T.I., Levine, S.S., Wernig, M., Tajonar, A., Ray, M.K., Bell, G.W., Otte, A.P., Vidal, M., Gifford, D.K., Young, R.A., and Jaenisch, R. (2006). Polycomb complexes repress developmental regulators in murine embryonic stem cells. Nature 441, 349–353.

Bransteitter, R., Pham, P., Scharff, M.D., and Goodman, M.F. (2003). Activation-induced cytidine deaminase deaminates deoxycytidine on single-stranded DNA but requires the action of RNase. Proc. Natl. Acad. Sci. U. S. A. 100, 4102–4107.

Briggs, B.R., and King, T.J. (1952). Transplantation of living nuclei from blastula. BLASTULA. PNAS 38, 455–463.

Brookes, E., de Santiago, I., Hebenstreit, D., Morris, K.J., Carroll, T., Xie, S.Q., Stock, J.K., Heidemann, M., Eick, D., Nozaki, N., Kimura, H., Ragoussis, J., Teichmann, S.A., and Pombo, A. (2012). Polycomb associates genome-wide with a specific RNA polymerase II variant, and regulates metabolic genes in ESCs. Cell Stem Cell 10, 157–170.

Buganim, Y., Faddah, D.A., Cheng, A.W., Itskovich, E., Markoulaki, S., Ganz, K., Klemm, S.L., van Oudenaarden, A., and Jaenisch, R. (2012). Single-cell expression analyses during cellular reprogramming reveal an early stochastic and a late hierarchic phase. Cell 150, 1209–1222.

Butler, M.G. (2009). Genomic imprinting disorders in humans: a mini-review. J. Assist. Reprod. Genet. 26, 477–486.

Byrne, J. a, Pedersen, D. a, Clepper, L.L., Nelson, M., Sanger, W.G., Gokhale, S., Wolf, D.P., and Mitalipov, S.M. (2007). Producing primate embryonic stem cells by somatic cell nuclear transfer. Nature 450, 497–502.

Cahan, P., and Daley, G.Q. (2013). Origins and implications of pluripotent stem cell variability and heterogeneity. Nat. Rev. Mol. Cell Biol. 14, 357–368.

Cao, Y., Yang, J., Yin, Z.Q., Luo, H.Y., Yang, M., Hu, N., Yang, J., Huo, D.Q., Hou, C.J., Jiang, Z.Z., Zhang, R.Q., Xu, R., and Zheng X.L. (2008). Study of high-throughput cell electrofusion in a microelectrode-array chip. Microfluid. Nanofluidics 5, 669–675.

Capecchi, M.R. (1989). Altering the genome by homologous recombination. Science 244, 1288–1292.

Cardoso, M.C., and Leonhardt, H. (1999). DNA methyltransferase is actively retained in the cytoplasm during early development. J. Cell Biol. 147, 25–32.

Carninci, P., Sandelin, A., Lenhard, B., Katayama, S., Shimokawa, K., Ponjavic, J., Semple, C. a M., Taylor, M.S., Engström, P.G., Frith, M.C., Forres, A.R.R., Alkema, W.B., Tan, S.L., Plessy, C., Kodzius, R., Ravasi, T., Kasukawa, T., Fukuda, S., Kanamori-Katayama, M., Kitazume, Y., Kawaji, H., Kai, C., Nakamura, M., Konno, H., Nakano, K., Mottagui-Tabar, S., Arner, P., Chesi, A., Gustincich, S., Persichetti, F., Suzuki, H., Grimmond, S.M., Wells, C.A., Orlando, V., Wahlestedt, C., Liu, E.T., Harbers, M., Kawai, J., Bajic, V.B., Hume, D.A., and Hayashizaki, Y.(2006). Genome-wide analysis of mammalian promoter architecture and evolution. Nat. Genet. 38, 626–635.

Cervoni, N., and Szyf, M. (2001). Demethylase activity is directed by histone acetylation. J. Biol. Chem. 276, 40778–40787.

Chambers, I., Silva, J., Colby, D., Nichols, J., Nijmeijer, B., Robertson, M., Vrana, J., Jones, K., Grotewold, L., and Smith, A. (2007). Nanog safeguards pluripotency and mediates germline development. Nature 450, 1230–1234.

130

Page 131: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Chen, J., Guo, L., Zhang, L., Wu, H., Yang, J., Liu, H., Wang, X., Hu, X., Gu, T., Zhou, Z., et al. (2013a). Vitamin C modulates TET1 function during somatic cell reprogramming. Nat. Genet. 45, 1504–1509.

Chen, Q., Chen, Y., Bian, C., Fujiki, R., and Yu, X. (2013b). TET2 promotes histone O-GlcNAcylation during gene transcription. Nature 493, 561–564.

Chen, T., Ueda, Y., Dodge, J., Wang, Z., and Li, E. (2003). Establishment and maintenance of genomic methylation patterns in mouse embryonic stem cells by Dnmt3a and Dnmt3b. Mol. Cell. 23, 5594-5605.

Chen, X., Xu, H., Yuan, P., Fang, F., Huss, M., Vega, V.B., Wong, E., Orlov, Y.L., Zhang, W., Jiang, J., Loh, Y., Yeo, H.C., Yeo, Z.X., Narang, V., Govindarajan, K.R., Leong, B., Shahab, A., Ruan, Y., Bourque, G., Sung, W., Clarke, N.D., Wei, C., and Ng, H. (2008). Integration of external signaling pathways with the core transcriptional network in embryonic stem cells. Cell 133, 1106–1117.

Chiu, C.P., and Blau, H.M. (1985). 5-Azacytidine permits gene activation in a previously noninducible cell type. Cell 40, 417–424.

Cho, S.W., Kim, S., Kim, Y., Kweon, J., Kim, H.S., Bae, S., and Kim, J. (2014). Analysis of off-target effects of CRISPR / Cas-derived RNA-guided endonucleases and nickases. Genome research, 24(1), 132-141.

Chung, T.-L., Brena, R.M., Kolle, G., Grimmond, S.M., Berman, B.P., Laird, P.W., Pera, M.F., and Wolvetang, E.J. (2010). Vitamin C promotes widespread yet specific DNA demethylation of the epigenome in human embryonic stem cells. Stem Cells 28, 1848–1855.

Ciccarone, F., Klinger, F.G., Catizone, A., Calabrese, R., Zampieri, M., Bacalini, M.G., De Felici, M., and Caiafa, P. (2012). Poly(ADP-ribosyl)ation acts in the DNA demethylation of mouse primordial germ cells also with DNA damage-independent roles. PLoS One 7, e46927.

Ciccone, D.N., Su, H., Hevi, S., Gay, F., Lei, H., Bajko, J., Xu, G., Li, E., and Chen, T. (2009). KDM1B is a histone H3K4 demethylase required to establish maternal genomic imprints. Nature 461, 415–418.

Clapier, C.R., and Cairns, B.R. (2009). The biology of chromatin remodeling complexes. Annu. Rev. Biochem. 78, 273–304.

Clouaire, T., Webb, S., Skene, P., Illingworth, R., Kerr, A., Andrews, R., Lee, J.-H., Skalnik, D., and Bird, A. (2012). Cfp1 integrates both CpG content and gene activity for accurate H3K4me3 deposition in embryonic stem cells. Genes Dev. 26, 1714–1728.

Cong, L., Ran, F.A., Cox, D., Lin, S., Barretto, R., Habib, N., Hsu, P.D., Wu, X., Jiang, W., Marraffini, L.A., and Zhang, F. (2013). Multiplex genome engineering using CRISPR/Cas systems. Science 339, 819–823.

Cortázar, D., Kunz, C., Selfridge, J., Lettieri, T., Saito, Y., MacDougall, E., Wirz, A., Schuermann, D., Jacobs, A.L., Siegrist, F., Steinacher, R., Jiricny, J., Bird, A., and Schar, P. (2011). Embryonic lethal phenotype reveals a function of TDG in maintaining epigenetic stability. Nature 470, 419–423.

Cortellino, S., Xu, J., Sannai, M., Moore, R., Caretti, E., Cigliano, A., Le Coz, M., Devarajan, K., Wessels, A., Soprano, D., Abramowitz, L.K., Bartolomei, M.S., Rambow, F., Bassi, M.R., Bruno, T., Fanciulli, M., Renner, C., Klein-Szanto, A.J., Matsumoto, Y., Kobi, D., Davidson, I., Alberti, C., Larue, L., and Bellacosa, A. (2011). Thymine DNA glycosylase is essential for active DNA demethylation by linked deamination-base excision repair. Cell 146, 67–79.

Costa, Y., Ding, J., Theunissen, T.W., Faiola, F., Hore, T. a, Shliaha, P. V, Fidalgo, M., Saunders, A., Lawrence, M., Dietmann, S., Das, S., Levasseur, D.N., Li, Z., Xu, M., Reik, W., Silva, J.C.R., and Wang, J. (2013). NANOG-dependent function of TET1 and TET2 in establishment of pluripotency. Nature 495, 370–374.

Cowan, C. a, Atienza, J., Melton, D. a, and Eggan, K. (2005). Nuclear reprogramming of somatic cells after fusion with human embryonic stem cells. Science 309, 1369–1373.

131

Page 132: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Dalton, S., and Bellacosa, A. (2012). DNA demethylation by TDG. Epigenomics 4, 459–467.

Davidson, R.L. (1972). Regulation of Melanin Synthesis in Mammalian Cells: Effect of Gene Dosage on the Expression of Differentiation. Proc. Natl. Acad. Sci. U. S. A. 69, 951–955.

Davidson, R., Ephrussi, B., and Yamamoto, K. (1966). Regulation of pigment synthesis in mammalian cells, as studied by somatic hybridization. Proc. Natl. Acad. Sci. U. S. A. 56.5 1437–1440.

Davis, R.L., Weintraub, H., and Lassar, A.B. (1987). Expression of a single transfected cDNA converts fibroblasts to myoblasts. Cell 51, 987–1000.

Dawlaty, M.M., Ganz, K., Powell, B.E., Hu, Y.-C., Markoulaki, S., Cheng, A.W., Gao, Q., Kim, J., Choi, S.-W., Page, D.C., and Jaenisch, R. (2011). Tet1 is dispensable for maintaining pluripotency and its loss is compatible with embryonic and postnatal development. Cell Stem Cell 9, 166–175.

Dawlaty, M.M., Breiling, A., Le, T., Raddatz, G., Barrasa, M.I., Cheng, A.W., Gao, Q., Powell, B.E., Li, Z., Xu, M., Faull, K.F., Lyko, F., and Jaenisch, R. (2013). Combined deficiency of Tet1 and Tet2 causes epigenetic abnormalities but is compatible with postnatal development. Dev. Cell 24, 310–323.

Dawlaty, M.M., Breiling, A., Le, T., Barrasa, M.I., Raddatz, G., Gao, Q., Powell, B.E., Cheng, A.W., Faull, K.F., Lyko, F., and Jaenisch, R. (2014). Loss of Tet enzymes compromises proper differentiation of embryonic stem cells. Dev. Cell 29, 102–111.

Deaton, A., and Bird, A. (2011). CpG islands and the regulation of transcription. Genes Dev. 1010–1022.

Deb-Rinker, P., Ly, D., Jezierski, A., Sikorska, M., and Walker, P.R. (2005). Sequential DNA methylation of the Nanog and Oct-4 upstream regions in human NT2 cells during neuronal differentiation. J. Biol. Chem. 280, 6257–6260.

DeChiara, T.M., Robertson, E.J., and Efstratiadis, A. (1991). Parental imprinting of the mouse insulin-like growth factor II gene. Cell 64, 849–859.

Deltcheva, E., Chylinski, K., Sharma, C.M., Gonzales, K., Chao, Y., Pirzada, Z. a, Eckert, M.R., Vogel, J., and Charpentier, E. (2011). CRISPR RNA maturation by trans-encoded small RNA and host factor RNase III. Nature 471, 602–607.

Denholtz, M., Bonora, G., Chronis, C., Splinter, E., de Laat, W., Ernst, J., Pellegrini, M., and Plath, K. (2013). Long-range chromatin contacts in embryonic stem cells reveal a role for pluripotency factors and polycomb proteins in genome organization. Cell Stem Cell 13, 602–616.

Deplus, R., Delatte, B., Schwinn, M.K., Defrance, M., Méndez, J., Murphy, N., Dawson, M. a, Volkmar, M., Putmans, P., Calonne, E., Shih, A.H., Levine, R.L., Bernard, O., Mercher, T., Solary, E., Urh, M., Daniels, D.L., and Fuks, F. (2013). TET2 and TET3 regulate GlcNAcylation and H3K4 methylation through OGT and SET1/COMPASS. EMBO J. 32, 645–655.

Do, J., and Schöler, H. (2004). Nuclei of embryonic stem cells reprogram somatic cells. Stem Cells, 22(6), 941–949.

Doege, C. a, Inoue, K., Yamashita, T., Rhee, D.B., Travis, S., Fujita, R., Guarnieri, P., Bhagat, G., Vanti, W.B., Shih, A., Levine, R.L., Nik, S., Chen, E.I., and Abeliovich, A. (2012). Early-stage epigenetic modification during somatic cell reprogramming by Parp1 and Tet2. Nature 488, 652–655.

Dong, K.B., Maksakova, I. a, Mohn, F., Leung, D., Appanah, R., Lee, S., Yang, H.W., Lam, L.L., Mager, D.L., Schübeler, D., Tachibana, M., Shinkai, Y., and Lorincz, M.C. (2008). DNA methylation in ES cells requires the lysine methyltransferase G9a but not its catalytic activity. EMBO J. 27, 2691–2701.

Dura, B., Liu, Y., and Voldman, J. (2014). Deformability-based microfluidic cell pairing and fusion. Lab Chip 14, 2783–2790.

132

Page 133: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Edd, J.F., Di Carlo, D., Humphry, K.J., Köster, S., Irimia, D., Weitz, D. a, and Toner, M. (2008). Controlled encapsulation of single-cells into monodisperse picolitre drops. Lab Chip 8, 1262–1264.

Eden, E.R., Patel, D.D., Sun, X.-M., Burden, J.J., Themis, M., Edwards, M., Lee, P., Neuwirth, C., Naoumova, R.P., and Soutar, A.K. (2002). Restoration of LDL receptor function in cells from patients with autosomal recessive hypercholesterolemia by retroviral expression of ARH1. J. Clin. Invest. 110, 1695–1702.

Edwards, C. a, and Ferguson-Smith, A.C. (2007). Mechanisms regulating imprinted genes in clusters. Curr. Opin. Cell Biol. 19, 281–289.

Efroni, S., Duttagupta, R., Cheng, J., Dehghani, H., Hoeppner, D.J., Dash, C., Bazett-Jones, D.P., Le Grice, S., McKay, R.D.G., Buetow, K.H., Gingeras, T.R., Misteli, T., and Meshorer, E. (2008). Global transcription in pluripotent embryonic stem cells. Cell Stem Cell 2, 437–447.

Eggan, K., Baldwin, K., Tackett, M., Osborne, J., Gogos, J., Chess, A., Axel, R., and Jaenisch, R. (2004). Mice cloned from olfactory sensory neurons. Nature 428, 44–49.

Egli, D., Chen, A.E., Saphier, G., Ichida, J., Fitzgerald, C., Go, K.J., Acevedo, N., Patel, J., Baetscher, M., Kearns, W.G., Goland, R., Leibel, R.L., Melton, D.A., and Eggan, K. (2011). Reprogramming within hours following nuclear transfer into mouse but not human zygotes. Nat. Commun. 2, 488.

Engel, N., Thorvaldsen, J.L., and Bartolomei, M.S. (2006). CTCF binding sites promote transcription initiation and prevent DNA methylation on the maternal allele at the imprinted H19/Igf2 locus. Hum. Mol. Genet. 15, 2945–2954.

Engel, N., Tront, J.S., Erinle, T., Nguyen, N., Latham, K.E., Sapienza, C., Hoffman, B., and Liebermann, D. a (2009). Conserved DNA methylation in Gadd45a(-/-) mice. Epigenetics 4, 98–99.

Epsztejn-Litman, S., Feldman, N., Abu-Remaileh, M., Shufaro, Y., Gerson, A., Ueda, J., Deplus, R., Fuks, F., Shinkai, Y., Cedar, H., and Bergman, Y. (2008). De novo DNA methylation promoted by G9a prevents reprogramming of embryonically silenced genes. Nat. Struct. Mol. Biol. 15, 1176–1183.

Evans, M.J., and Kaufman, M.H. (1981). Establishment in culture of pluripotential cells from mouse embryos. Nature 292, 154–156.

Farcas, A.M., Blackledge, N.P., Sudbery, I., Long, H.K., McGouran, J.F., Rose, N.R., Lee, S., Sims, D., Cerase, A., Sheahan, T.W., Koseki, H., Brockdorff, N., Ponting, C.P., Kessler, B.M., and Klose, R.J. (2012). KDM2B links the Polycomb Repressive Complex 1 (PRC1) to recognition of CpG islands. Elife 1, e00205.

Farthing, C.R., Ficz, G., Ng, R.K., Chan, C.-F., Andrews, S., Dean, W., Hemberger, M., and Reik, W. (2008). Global mapping of DNA methylation in mouse promoters reveals epigenetic reprogramming of pluripotency genes. PLoS Genet. 4, e1000116.

Feldman, N., Gerson, A., Fang, J., Li, E., Zhang, Y., Shinkai, Y., Cedar, H., and Bergman, Y. (2006). G9a-mediated irreversible epigenetic inactivation of Oct-3/4 during early embryogenesis. Nat. Cell Biol. 8, 188–194.

Ferguson-Smith, A.C. (2011). Genomic imprinting: the emergence of an epigenetic paradigm. Nat. Rev. Genet. 12, 565–575.

Ferguson-Smith, A., and Cattanach, B. (1991). Embryological and molecular investigations of parental imprinting on mouse chromosome 7. Nature 351, 667-670.

Ferrari, K.J., Scelfo, A., Jammula, S., Cuomo, A., Barozzi, I., Stützer, A., Fischle, W., Bonaldi, T., and Pasini, D. (2014). Polycomb-dependent H3K27me1 and H3K27me2 regulate active transcription and enhancer fidelity. Mol. Cell 53, 49–62.

133

Page 134: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Ficz, G., Branco, M.R., Seisenberger, S., Santos, F., Krueger, F., Hore, T. a, Marques, C.J., Andrews, S., and Reik, W. (2011). Dynamic regulation of 5-hydroxymethylcytosine in mouse ES cells and during differentiation. Nature 473, 398–402.

Ficz, G., Hore, T.A., Santos, F., Lee, H.J., Dean, W., Arand, J., Krueger, F., Oxley, D., Paul, Y.-L., Walter, J., Cook, S.J., Andrews, S., Branco, M.R., Reik, W. (2013). FGF Signaling Inhibition in ESCs Drives Rapid Genome-wide Demethylation to the Epigenetic Ground State of Pluripotency. Cell Stem Cell, 13(3), 351–359.

Foshay, K.M., Looney, T.J., Chari, S., Mao, F.F., Lee, J.H., Zhang, L., Fernandes, C.J., Baker, S.W., Clift, K.L., Gaetz, J., Di, C.G., Xiang, A.P., and Lahn, B.T. (2012). Embryonic stem cells induce pluripotency in somatic cell fusion through biphasic reprogramming. Mol. Cell 46, 159–170.

Frauer, C., Hoffmann, T., Bultmann, S., Casa, V., Cardoso, M.C., Antes, I., and Leonhardt, H. (2011). Recognition of 5-hydroxymethylcytosine by the Uhrf1 SRA domain. PLoS One 6, e21306.

Freberg, C., Dahl, J., Timoskanien, S., and Collas, P. (2007). Epigenetic reprogramming of OCT4 and NANOG regulatory regions by embryonal carcinoma cell extract. Biol. Cell 18, 1543–1553.

Freudenberg, J.M., Ghosh, S., Lackford, B.L., Yellaboina, S., Zheng, X., Li, R., Cuddapah, S., Wade, P. a, Hu, G., and Jothi, R. (2012). Acute depletion of Tet1-dependent 5-hydroxymethylcytosine levels impairs LIF/Stat3 signaling and results in loss of embryonic stem cell identity. Nucleic Acids Res. 40, 3364–3377.

Frommer, M., McDonald, L.E., Millar, D.S., Collis, C.M., Watt, F., Grigg, G.W., Molloy, P.L., and Paul, C.L. (1992). A genomic sequencing protocol that yields a positive display of 5-methylcytosine residues in individual DNA strands. Proc. Natl. Acad. Sci. U. S. A. 89, 1827–1831.

Frost, J.M., and Moore, G.E. (2010). The importance of imprinting in the human placenta. PLoS Genet. 6, e1001015.

Fu, Y., Foden, J. a, Khayter, C., Maeder, M.L., Reyon, D., Joung, J.K., and Sander, J.D. (2013). High-frequency off-target mutagenesis induced by CRISPR-Cas nucleases in human cells. Nat. Biotechnol. 31, 822–826.

Gao, Y., Chen, J., Li, K., Wu, T., Huang, B., Liu, W., Kou, X., Zhang, Y., Huang, H., Jiang, Y., Yao, C., Liu, X., Lu, Z., Xu, Z., Kang, L., Chen, J., Wang, H., Cai, T., and Gao, S. (2013). Replacement of Oct4 by Tet1 during iPSC induction reveals an important role of DNA methylation and hydroxymethylation in reprogramming. Cell Stem Cell 12, 453–469.

Garneau, J.E., Dupuis, M.-È., Villion, M., Romero, D. a, Barrangou, R., Boyaval, P., Fremaux, C., Horvath, P., Magadán, A.H., and Moineau, S. (2010). The CRISPR/Cas bacterial immune system cleaves bacteriophage and plasmid DNA. Nature 468, 67–71.

Gaspar-maia, A., Alajem, A., Polesso, F., Sridharan, R., Mason, M.J., Heidersbach, A., Mcmanus, M.T., Plath, K., Meshorer, E., and Ramalho-santos, M. (2009). Chd1 regulates open chromatin and pluripotency of embryonic stem cells. Nature 460, 863-868.

Gaspar-Maia, A., Alajem, A., Meshorer, E., and Ramalho-Santos, M. (2011). Open chromatin in pluripotency and reprogramming. Nat. Rev. Mol. Cell Biol. 12, 36–47.

Gilbert, L. a, Larson, M.H., Morsut, L., Liu, Z., Brar, G. a, Torres, S.E., Stern-Ginossar, N., Brandman, O., Whitehead, E.H., Doudna, J.A., Lim, W.A., Weissman, J.S., and Qi, L.S. (2013). CRISPR-mediated modular RNA-guided regulation of transcription in eukaryotes. Cell 154, 442–451.

Globisch, D., Münzel, M., Müller, M., Michalakis, S., Wagner, M., Koch, S., Brückl, T., Biel, M., and Carell, T. (2010). Tissue distribution of 5-hydroxymethylcytosine and search for active demethylation intermedibates. PLoS One 5, e15367.

134

Page 135: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Goll, M.G., and Bestor, T.H. (2005). Eukaryotic cytosine methyltransferases. Annu. Rev. Biochem. 74, 481–514.

Goll, M.G., Kirpekar, F., Maggert, K. a, Yoder, J. a, Hsieh, C.-L., Zhang, X., Golic, K.G., Jacobsen, S.E., and Bestor, T.H. (2006). Methylation of tRNAAsp by the DNA methyltransferase homolog Dnmt2. Science 311, 395–398.

Graf, T., and Enver, T. (2009). Forcing cells to change lineages. Nature 462, 587–594.

Gu, T.-P., Guo, F., Yang, H., Wu, H.-P., Xu, G.-F., Liu, W., Xie, Z.-G., Shi, L., He, X., Jin, S., Iqbal, K., Shi, Y.G., Deng, X., Szabo, P.E., Pfeifer, G.P., Li, J., and Xu, G. (2011). The role of Tet3 DNA dioxygenase in epigenetic reprogramming by oocytes. Nature 477, 606–610.

Guenther, M.G., Levine, S.S., Boyer, L. a, Jaenisch, R., and Young, R.A. (2007). A chromatin landmark and transcription initiation at most promoters in human cells. Cell 130, 77–88.

Guibert, S., Forné, T., and Weber, M. (2012). Global profiling of DNA methylation erasure in mouse primordial germ cells. Genome Res. 22, 633–641.

Guo, G., Huss, M., Tong, G.Q., Wang, C., Li Sun, L., Clarke, N.D., and Robson, P. (2010). Resolution of cell fate decisions revealed by single-cell gene expression analysis from zygote to blastocyst. Dev. Cell 18, 675–685.

Guo, J.U., Su, Y., Zhong, C., Ming, G.-L., and Song, H. (2011). Hydroxylation of 5-Methylcytosine by TET1 Promotes Active DNA Demethylation in the Adult Brain. Cell 145, 423–434.

Gurdon, J.B. (1962a). The developmental capacity of nuclei taken from intestinal epithelium cells of feeding tadpoles. J. Embryol. Exp. Morphol. 10, 622–640.

Gurdon, J.B. (1962b). Adult frogs derived from the nuclei of single somatic cells. Dev. Biol. 4, 256–273.

Habib, O., Habib, G., Do, J.T., Moon, S.-H., and Chung, H.-M. (2014). Activation-induced deaminase-coupled DNA demethylation is not crucial for the generation of induced pluripotent stem cells. Stem Cells Dev. 23, 209–218.

Habibi, E., Brinkman, A.B., Arand, J., Kroeze, L.I., Kerstens, H.H.D., Matarese, F., Lepikhov, K., Gut, M., Brun-Heath, I., Hubner, N.C., Benedetti, R., Altucci, L., Jansen, J.H., Walter, J., Gut, I.G., Marks, H., and Stunnenberg, H.G. (2013). Whole-Genome Bisulfite Sequencing of Two Distinct Interconvertible DNA Methylomes of Mouse Embryonic Stem Cells. Cell Stem Cell 13, 360-369.

Hackett, J. a, Sengupta, R., Zylicz, J.J., Murakami, K., Lee, C., Down, T. a, and Surani, M.A. (2013). Germline DNA demethylation dynamics and imprint erasure through 5-hydroxymethylcytosine. Science 339, 448–452.

Hajkova, P., Erhardt, S., Lane, N., Haaf, T., El-Maarri, O., Reik, W., Walter, J., and Surani, M.A. (2002). Epigenetic reprogramming in mouse primordial germ cells. Mech. Dev. 117, 15–23.

Hajkova, P., Ancelin, K., Waldmann, T., Lacoste, N., Lange, U.C., Cesari, F., Lee, C., Almouzni, G., Schneider, R., and Surani, M.A. (2008). Chromatin dynamics during epigenetic reprogramming in the mouse germ line. Nature 452, 877–881.

Hajkova, P., Jeffries, S.J., Lee, C., Miller, N., Jackson, S.P., and Surani, M.A. (2010). Genome-wide reprogramming in the mouse germ line entails the base excision repair pathway. Science 329, 78–82.

Han, D.W., Do, J.T., Gentile, L., Stehling, M., Lee, H.T., and Schöler, H.R. (2008). Pluripotential reprogramming of the somatic genome in hybrid cells occurs with the first cell cycle. Stem Cells 26, 445–454.

135

Page 136: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Hanna, J.H., Saha, K., and Jaenisch, R. (2010). Pluripotency and cellular reprogramming: facts, hypotheses, unresolved issues. Cell 143, 508–525.

Harris, H., and Miller, O. (1969). Suppression of malignancy by cell fusion. Nature 223, 363-368.

Harris, H., and Watkins, J. (1965). Hybrid cells derived from mouse and man: artificial heterokaryons of mammalian cells from different species. Nature 205, 640-646.

Hashimoto, H., Horton, J.R., Zhang, X., Bostick, M., Jacobsen, S.E., and Cheng, X. (2008). The SRA domain of UHRF1 flips 5-methylcytosine out of the DNA helix. Nature 455, 826–829.

Hashimoto, H., Liu, Y., Upadhyay, A.K., Chang, Y., Howerton, S.B., Vertino, P.M., Zhang, X., and Cheng, X. (2012). Recognition and potential mechanisms for replication and erasure of cytosine hydroxymethylation. Nucleic Acids Res. 40, 4841–4849.

Hata, K., Okano, M., Lei, H., and Li, E. (2002). Dnmt3L cooperates with the Dnmt3 family of de novo DNA methyltransferases to establish maternal imprints in mice. Development 129, 1983–1993.

Hathaway, N. a, Bell, O., Hodges, C., Miller, E.L., Neel, D.S., and Crabtree, G.R. (2012). Dynamics and memory of heterochromatin in living cells. Cell 149, 1447–1460.

Hawkins, R.D., Hon, G.C., Lee, L., Ngo, Q.-M., Lister, R., Pelizzola, M., Edsall, L., Kuan, S., Luu, Y., Klugman, S., Antosiewicz-Bourquet, J., Ye, Z., Espinoza, C., Agarwahl, S., Shen, L., Ruotti, V., Wang, W., Stewart, R., Thomson, J.A., Ecker, J.R., and Ren, B. (2010). Distinct epigenomic landscapes of pluripotent and lineage-committed human cells. Cell Stem Cell 6, 479–491.

He, Y.-F., Li, B.-Z., Li, Z., Liu, P., Wang, Y., Tang, Q., Ding, J., Jia, Y., Chen, Z., Li, L., Sun, Y., Li, X., Dai, Q., Song, C.X., Zhang, K., He, C., and Xu, G.L. (2011). Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science 333, 1303–1307.

Heitz, E. (1928). Das Heterochromatin der Moose. Jahrb Wiss Botanik. 69, 762–818.

Henckel, A., and Arnaud, P. (2010). Genome-wide identification of new imprinted genes. Brief. Funct. Genomics 9, 304–314.

Henckel, A., Chebli, K., Kota, S.K., Arnaud, P., and Feil, R. (2012). Transcription and histone methylation changes correlate with imprint acquisition in male germ cells. EMBO J. 31, 606–615.

Hendrich, B., Guy, J., Ramsahoye, B., Wilson, V. A., and Bird, A. (2001). Closely related proteins MBD2 and MBD3 play distinctive but interacting roles in mouse development. Genes Dev. 15, 710–723.

Hezroni, H., Tzchori, I., Davidi, A., Mattout, A., Biran, A., Nissim-Rafinia, M., Westphal, H., and Meshorer, E. (2011). H3K9 histone acetylation predicts pluripotency and reprogramming capacity of ES cells. Nucleus 2, 300–309.

Hikasa, H., Ezan, J., Itoh, K., Li, X., Klymkowsky, M.W., and Sokol, S.Y. (2010). Regulation of TCF3 by Wnt-dependent phosphorylation during vertebrate axis specification. Dev. Cell 19, 521–532.

Hirasawa, R., Chiba, H., Kaneda, M., Tajima, S., Li, E., Jaenisch, R., and Sasaki, H. (2008). Maternal and zygotic Dnmt1 are necessary and sufficient for the maintenance of DNA methylation imprints during preimplantation development. Genes Dev. 22, 1607–1616.

Hochedlinger, K., and Jaenisch, R. (2002). Monoclonal mice generated by nuclear transfer from mature B and T donor cells. Nature 415, 1035–1038.

Hochedlinger, K., and Plath, K. (2009). Epigenetic reprogramming and induced pluripotency. Development 136, 509–523.

136

Page 137: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Hooper, M., Hardy, K., Handyside, A., Hunter, S., and Monk, M. (1987). HPRT-deficient (Lesch–Nyhan) mouse embryos derived from germline colonization by cultured cells. Nature 326, 292–295.

Hsu, P.D., Scott, D. a, Weinstein, J. a, Ran, F.A., Konermann, S., Agarwala, V., Li, Y., Fine, E.J., Wu, X., Shalem, O., Cradick, T.J., Marraffini, L.A., Bao, G., and Zhang, F. (2013). DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 31, 827–832.

Hu, N., Yang, J., Yin, Z.-Q., Ai, Y., Qian, S., Svir, I.B., Xia, B., Yan, J.-W., Hou, W.-S., and Zheng, X.-L. (2011). A high-throughput dielectrophoresis-based cell electrofusion microfluidic device. Electrophoresis 32, 2488–2495.

Hu, N., Yang, J., Joo, S.W., Banerjee, A.N., and Qian, S. (2013). Cell electrofusion in microfluidic devices: A review. Sensors Actuators B Chem. 178, 63–85.

Hu, X., Zhang, L., Mao, S.-Q., Li, Z., Chen, J., Zhang, R.-R., Wu, H.-P., Gao, J., Guo, F., Liu, W., Xu, G.F., Dai, H.Q., Shi, Y.G., Li,X., Hu, B., Tang, F., Pei, D., and Xu, G.L. (2014). Tet and TDG Mediate DNA Demethylation Essential for Mesenchymal-to-Epithelial Transition in Somatic Cell Reprogramming. Cell Stem Cell 14, 512–522.

Huang, P., He, Z., Ji, S., Sun, H., Xiang, D., Liu, C., Hu, Y., Wang, X., and Hui, L. (2011). Induction of functional hepatocyte-like cells from mouse fibroblasts by defined factors. Nature 475, 386–389.

Huang, Y., Chavez, L., Chang, X., Wang, X., Pastor, W. a, Kang, J., Zepeda-Martínez, J. a, Pape, U.J., Jacobsen, S.E., Peters, B., and Rao, A. (2014). Distinct roles of the methylcytosine oxidases Tet1 and Tet2 in mouse embryonic stem cells. Proc. Natl. Acad. Sci. U. S. A. 111, 1361–1366.

Huebner, A., Sharma, S., Srisa-Art, M., Hollfelder, F., Edel, J.B., and Demello, A.J. (2008). Microdroplets: a sea of applications? Lab Chip 8, 1244–1254.

Hur, S.C., Tse, H.T.K., and Di Carlo, D. (2010). Sheathless inertial cell ordering for extreme throughput flow cytometry. Lab Chip 10, 274–280.

Ieda, M., Fu, J.-D., Delgado-Olguin, P., Vedantham, V., Hayashi, Y., Bruneau, B.G., and Srivastava, D. (2010). Direct Reprogramming of Fibroblasts into Functional Cardiomyocytes by Defined Factors. Cell 142, 375–386.

Ikeda, Y., and Kinoshita, T. (2009). DNA demethylation: a lesson from the garden. Chromosoma 118, 37–41.

Inoue, A., and Zhang, Y. (2011). Replication-dependent loss of 5-hydroxymethylcytosine in mouse preimplantation embryos. Science 334, 194.

Inoue, A., Shen, L., Dai, Q., He, C., and Zhang, Y. (2011). Generation and replication-dependent dilution of 5fC and 5caC during mouse preimplantation development. Cell Res. 21, 1670–1676.

Inoue, A., Matoba, S., and Zhang, Y. (2012). Transcriptional activation of transposable elements in mouse zygotes is independent of Tet3-mediated 5-methylcytosine oxidation. Cell Res. 22, 1640–1649.

Iqbal, K., Jin, S., Pfeifer, G.P., and Szabó, P.E. (2011). Reprogramming of the paternal genome upon fertilization involves genome-wide oxidation of 5-methylcytosine. Proc. Natl. Acad. Sci. U. S. A. 108, 2–7.

Iseli, C., Ambrosini, G., Bucher, P., and Jongeneel, C.V. (2007). Indexing strategies for rapid searches of short words in genome sequences. PLoS One 2, e579.

Ishino, Y., Shinagawa, H., Makino, K., Amemura, M., and Nakata, A. (1987). Nucleotide sequence of the iap gene, responsible for alkaline phosphatase isozyme conversion in Escherichia coli, and identification of the gene product. J. Bacteriol. 169, 5429–5433.

137

Page 138: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Ito, S., D’Alessio, A.C., Taranova, O. V, Hong, K., Sowers, L.C., and Zhang, Y. (2010). Role of Tet proteins in 5mC to 5hmC conversion, ES-cell self-renewal and inner cell mass specification. Nature 466, 1129–1133.

Ito, S., Shen, L., Dai, Q., Wu, S.C., Collins, L.B., Swenberg, J. a, He, C., and Zhang, Y. (2011). Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science 333, 1300–1303.

Iyengar, S., and Farnham, P.J. (2011). KAP1 protein: an enigmatic master regulator of the genome. J. Biol. Chem. 286, 26267–26276.

Iyer, L., Tahiliani, M., Rao, A., and Aravind, L. (2009). Prediction of novel families of enzymes involved in oxidative and other complex modifications of bases in nucleic acids. Cell Cycle 8, 1698–1710.

Jansen, R., Embden, J.D. a Van, Gaastra, W., and Schouls, L.M. (2002). Identification of genes that are associated with DNA repeats in prokaryotes. Mol. Microbiol. 43, 1565–1575.

Jenuwein, T., and Allis, C.D. (2001). Translating the histone code. Science 293, 1074–1080.

Jiang, W., Bikard, D., Cox, D., Zhang, F., and Marraffini, L. A. (2013). RNA-guided editing of bacterial genomes using CRISPR-Cas systems. Nat. Biotechnol. 31, 233–239.

Jin, S.-G., Guo, C., and Pfeifer, G.P. (2008). GADD45A does not promote DNA demethylation. PLoS Genet. 4, e1000013.

Jin, S.-G., Kadam, S., and Pfeifer, G.P. (2010). Examination of the specificity of DNA methylation profiling techniques towards 5-methylcytosine and 5-hydroxymethylcytosine. Nucleic Acids Res. 38, e125.

Jinek, M., Chylinski, K., Fonfara, I., and Hauer, M. (2012). A programmable dual-RNA–guided DNA endonuclease in adaptive bacterial immunity. Science 337, 816–821.

Jinek, M., Jiang, F., Taylor, D.W., Sternberg, S.H., Kaya, E., Ma, E., Anders, C., Hauer, M., Zhou, K., Lin, S., Kaplan, M., Iavarone, A.T., Carpentier, E., Nogales, E., and Doudna, J.A. (2014). Structures of Cas9 endonucleases reveal RNA-mediated conformational activation. Science 343, 1247997.

John, R.M., and Lefebvre, L. (2011). Developmental regulation of somatic imprints. Differentiation. 81, 270–280.

Jones, P. A. (2012). Functions of DNA methylation: islands, start sites, gene bodies and beyond. Nat. Rev. Genet. 13, 484–492.

Jullien, P.E., Mosquna, A., Ingouff, M., Sakata, T., Ohad, N., and Berger, F. (2008). Retinoblastoma and its binding partner MSI1 control imprinting in Arabidopsis. PLoS Biol. 6, e194.

Kagey, M.H., Newman, J.J., Bilodeau, S., Zhan, Y., Orlando, D. a, van Berkum, N.L., Ebmeier, C.C., Goossens, J., Rahl, P.B., Levine, S.S., Taatjes, D.J., Dekker, J., and Young, R.A. (2010). Mediator and cohesin connect gene expression and chromatin architecture. Nature 467, 430–435.

Kagiwada, S., Kurimoto, K., Hirota, T., Yamaji, M., and Saitou, M. (2013). Replication-coupled passive DNA demethylation for the erasure of genome imprints in mice. EMBO J. 32, 340–353.

Kalb, R., Latwiel, S., Baymaz, H.I., Jansen, P.W.T.C., Müller, C.W., Vermeulen, M., and Müller, J. (2014). Histone H2A monoubiquitination promotes histone H3 methylation in Polycomb repression. Nat. Struct. Mol. Biol. 21, 569–571.

Kalisky, T., Blainey, P., and Quake, S.R. (2011). Genomic analysis at the single-cell level. Annu. Rev. Genet. 45, 431–445.

138

Page 139: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Kaneda, M., Okano, M., Hata, K., Sado, T., Tsujimoto, N., Li, E., and Sasaki, H. (2004). Essential role for de novo DNA methyltransferase Dnmt3a in paternal and maternal imprinting. Nature 429, 900–903.

Kaneko, S., Bonasio, R., Saldaña-Meyer, R., Yoshida, T., Son, J., Nishino, K., Umezawa, A., and Reinberg, D. (2014). Interactions between JARID2 and noncoding RNAs regulate PRC2 recruitment to chromatin. Mol. Cell 53, 290–300.

Kangaspeska, S., Stride, B., Métivier, R., Polycarpou-Schwarz, M., Ibberson, D., Carmouche, R.P., Benes, V., Gannon, F., and Reid, G. (2008). Transient cyclical methylation of promoter DNA. Nature 452, 112–115.

Kantlehner, M., Kirchner, R., Hartmann, P., Ellwart, J.W., Alunni-Fabbroni, M., and Schumacher, A. (2011). A high-throughput DNA methylation analysis of a single cell. Nucleic Acids Res. 39, e44.

Karwacki-Neisius, V., Göke, J., Osorno, R., Halbritter, F., Ng, J.H., Weiße, A.Y., Wong, F.C.K., Gagliardi, A., Mullin, N.P., Festuccia, N., Colby, D., Tomlinson, S.R., Ng, H.H., and Chambers, I. (2013). Reduced Oct4 expression directs a robust pluripotent state with distinct signaling activity and increased enhancer occupancy by Oct4 and Nanog. Cell Stem Cell 12, 531–545.

Kawasaki, Y., Lee, J., Matsuzawa, A., Kohda, T., Kaneko-Ishino, T., and Ishino, F. (2014). Active DNA demethylation is required for complete imprint erasure in primordial germ cells. Sci. Rep. 4, 3658.

Kellogg, R. a, Gómez-Sjöberg, R., Leyrat, A. a, and Tay, S. (2014). High-throughput microfluidic single-cell analysis pipeline for studies of signaling dynamics. Nat. Protoc. 9, 1713–1726.

Kelsey, G., and Feil, R. (2013). New insights into establishment and maintenance of DNA methylation imprints in mammals. R. Soc. B Biol. 368(1609):20110336

Kemna, E.W.M., Wolbers, F., Vermes, I., and van den Berg, A. (2011). On chip electrofusion of single human B cells and mouse myeloma cells for efficient hybridoma generation. Electrophoresis 32, 3138–3146.

Kemna, E.W.M., Schoeman, R.M., Wolbers, F., Vermes, I., Weitz, D. A., and van den Berg, A. (2012). High-yield cell ordering and deterministic cell-in-droplet encapsulation using Dean flow in a curved microchannel. Lab Chip 12, 2881–2887.

Kim, J., Chu, J., Shen, X., Wang, J., and Orkin, S.H. (2008). An extended transcriptional network for pluripotency of embryonic stem cells. Cell 132, 1049–1061.

Kim, K., Doi, a., Wen, B., Ng, K., Zhao, R., Cahan, P., Kim, J., Aryee, M.J., Ji, H., Ehrlich, L.I.R., Yabuuchi, A., Takeuchi, A., Cunniff, K.C., Hongguang, H., McKinney-Freeman, S., Naveiras, O., Yoon, T.J., Irizarry, R.A., Jung, N., Seita, J., Hanna, J., Murakami, P., JAenisch, R., Weissleder, R., Orkin, S.H., Weissman, I.L., Feinberg, A.P., and Daley, G.Q. (2010). Epigenetic memory in induced pluripotent stem cells. Nature 467, 285-290

Kimura, H., Tada, M., Nakatsuji, N., and Tada, T. (2004). Histone Code Modifications on Pluripotential Nuclei of Reprogrammed Somatic Cells. Society 24, 5710–5720.

Kimura, Y., Gel, M., Techaumnat, B., Oana, H., Kotera, H., and Washizu, M. (2011). Dielectrophoresis-assisted massively parallel cell pairing and fusion based on field constriction created by a micro-orifice array sheet. Electrophoresis 32, 2496–2501.

Kishigami, S., Mizutani, E., Ohta, H., Hikichi, T., Thuan, N. Van, Wakayama, S., Bui, H.-T., and Wakayama, T. (2006). Significant improvement of mouse cloning technique by treatment with trichostatin A after somatic nuclear transfer. Biochem. Biophys. Res. Commun. 340, 183–189.

Kishikawa, S., Murata, T., Ugai, H., Yamazaki, T., and Yokoyama, K.K. (2003). Control elements of Dnmt1 gene are regulated in cell-cycle dependent manner. Nucleic Acids Res. Suppl. 307–308.

139

Page 140: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Kliszczak, A.E., Rainey, M.D., Harhen, B., Boisvert, F.M., and Santocanale, C. (2011). DNA mediated chromatin pull-down for the study of chromatin replication. Sci. Rep. 1, 95.

Klose, R.J., Kallin, E.M., & Zhang, Y. (2006). JmjC-domain-containing proteins and histone demethylation. Nature Reviews Genetics, 7(9), 715-727.

Klose, R.J., and Bird, A.P. (2006). Genomic DNA methylation: the mark and its mediators. Trends Biochem. Sci. 31, 89–97.

Ko, M., An, J., Bandukwala, H.S., Chavez, L., Aijö, T., Pastor, W. a, Segal, M.F., Li, H., Koh, K.P., Lähdesmäki, H., Hoga, P.G., Aravind, L., and Rao, A. (2013). Modulation of TET2 expression and 5-methylcytosine oxidation by the CXXC domain protein IDAX. Nature 497, 122–126.

Koche, R.P., Smith, Z.D., Adli, M., Gu, H., Ku, M., Gnirke, A., Bernstein, B.E., and Meissner, A. (2011). Reprogramming factor expression initiates widespread targeted chromatin remodeling. Cell Stem Cell 8, 96–105.

Koh, K.P., Yabuuchi, A., Rao, S., Huang, Y., Cunniff, K., Nardone, J., Laiho, A., Tahiliani, M., Sommer, C. a, Mostoslavsky, G., Lahesmaa, R., Orkin, S.H., Rodig, S.J., Daley, G.Q., Rao, A. (2011). Tet1 and Tet2 regulate 5-hydroxymethylcytosine production and cell lineage specification in mouse embryonic stem cells. Cell Stem Cell 8, 200–213.

Kopp, J.L., Ormsbee, B.D., Desler, M., and Rizzino, A. (2008). Small increases in the level of Sox2 trigger the differentiation of mouse embryonic stem cells. Stem Cells 26, 903–911.

Köster, S., Angilè, F.E., Duan, H., Agresti, J.J., Wintner, A., Schmitz, C., Rowat, A.C., Merten, C. a, Pisignano, D., Griffiths, A.D., and Merten, C.A. (2008). Drop-based microfluidic devices for encapsulation of single cells. Lab Chip 8, 1110–1115.

Krejci, J., Uhlirova, R., Galiova, G., Kozubek, S., Smigova, J., and Bartova, E. (2009). Genome-wide reduction in H3K9 acetylation during human embryonic stem cell. J. Cell. Physiol. 219, 677–687.

Kriaucionis, S., and Heintz, N. (2009). The nuclear DNA base 5-hydroxymethylcytosine is present in Purkinje neurons and the brain. Science 324, 929–930.

Kühl, M., Sheldahl, L., Park, M., Miller, J., and Moon, R. (2000). The Wnt/Ca 2+ pathway: a new vertebrate Wnt signaling pathway takes shape. Trends Genet. 9525, 279–283.

Kulessa, H., Frampton, J., and Graf, T. (1995). GATA-1 reprograms avian myelomonocytic cell lines into eosinophils, thromboblasts, and erythroblasts. Genes Dev. 1250–1262.

Kumar, R., DiMenna, L., Schrode, N., Liu, T.-C., Franck, P., Muñoz-Descalzo, S., Hadjantonakis, A.-K., Zarrin, A. a, Chaudhuri, J., Elemento, O., and Evans, T. (2013). AID stabilizes stem-cell phenotype by removing epigenetic memory of pluripotency genes. Nature 500, 89–92.

Kurimoto, K., Yabuta, Y., Ohinata, Y., Shigeta, M., Yamanaka, K., and Saitou, M. (2008). Complex genome-wide transcription dynamics orchestrated by Blimp1 for the specification of the germ cell lineage in mice. Genes Dev. 22, 1617–1635.

Labosky, P. a, Barlow, D.P., and Hogan, B.L. (1994). Mouse embryonic germ (EG) cell lines: transmission through the germline and differences in the methylation imprint of insulin-like growth factor 2 receptor (Igf2r) gene compared with embryonic stem (ES) cell lines. Development 120, 3197–3204.

Laird, C.D., Pleasant, N.D., Clark, A.D., Sneeden, J.L., Hassan, K.M.A., Manley, N.C., Vary, J.C., Morgan, T., Hansen, R.S., and Stöger, R. (2004). Hairpin-bisulfite PCR: assessing epigenetic methylation patterns on complementary strands of individual DNA molecules. Proc. Natl. Acad. Sci. U. S. A. 101, 204–209.

140

Page 141: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Landeira, D., and Fisher, A.G. (2011). Inactive yet indispensable: the tale of Jarid2. Trends Cell Biol. 21, 74–80.

Landeira, D., Sauer, S., Poot, R., Dvorkina, M., Mazzarella, L., Jørgensen, H.F., Pereira, C.F., Leleu, M., Piccolo, F.M., Spivakov, M., Brookes, E., Pombo, A., Fisher, C., Skarnes, W.C., Snoek, T., Bezstarosti, K., Demmers, J., Klose, R.J., Casanova, M., Tavares, L., Brockdorff, N., Merkenschlager, M., and Fisher, A.G. (2010). Jarid2 is a PRC2 component in embryonic stem cells required for multi-lineage differentiation and recruitment of PRC1 and RNA Polymerase II to developmental regulators. Nat. Cell Biol. 12, 618–624.

Lander, E.S., Linton, L.M., Birren, B., Nusbaum, C., Zody, M.C., Baldwin, J., Devon, K., Dewar, K., Doyle, M., FitzHugh, W., …& International Human Genome Sequencing Consortium (2001). Initial sequencing and analysis of the human genome. Nature 409, 860–921.

Lara-Astiaso, D., Weiner, A., Lorenzo-Vivas, E., Zaretsky, I., Jaitin, D.A., David, E., Keren-Shaul, H., Mildner, A., Winter, D., Jung, S., Friedman, N., and Amit, A. (2014). Chromatin state dynamics during blood formation. Science 345, 943-949.

Lee, J.-S., Smith, E., and Shilatifard, A. (2010). The language of histone crosstalk. Cell 142, 682–685.

Lefebvre, L., Viville, S., Barton, S.C., Ishino, F., Keverne, E.B., and Surani, M. A. (1998). Abnormal maternal behaviour and growth retardation associated with loss of the imprinted gene Mest. Nat. Genet. 20, 163–169.

Lei, H., Oh, S.P., Okano, M., Jüttermann, R., Goss, K. a, Jaenisch, R., and Li, E. (1996). De novo DNA cytosine methyltransferase activities in mouse embryonic stem cells. Development 122, 3195–3205.

Leitch, H.G., McEwen, K.R., Turp, A., Encheva, V., Carroll, T., Grabole, N., Mansfield, W., Nashun, B., Knezovich, J.G., Smith, A., Surani, M.A., and Hajkova, P. (2013). Naive pluripotency is associated with global DNA hypomethylation. Nat. Struct. Mol. Biol. 20, 311–316.

Leonhardt, H., Page, a W., Weier, H.U., and Bestor, T.H. (1992). A targeting sequence directs DNA methyltransferase to sites of DNA replication in mammalian nuclei. Cell 71, 865–873.

Li, E., Bestor, T.H., and Jaenisch, R. (1992). Targeted mutation of the DNA methyltransferase gene results in embryonic lethality. Cell 69, 915–926.

Li, X., Ito, M., Zhou, F., Youngson, N., Zuo, X., Leder, P., and Ferguson-Smith, A.C. (2008). A maternal-zygotic effect gene, Zfp57, maintains both maternal and paternal imprints. Dev. Cell 15, 547–557.

Liu, X., Qinqin G., Pishun L., Qian Z., Jiqin Z., Jiwen L., Haruhiko K., and Jiemin W. (2013) UHRF1 targets DNMT1 for DNA methylation through cooperative binding of hemi-methylated DNA and methylated H3K9. Nature communications 4 (2013): 1563.

Long, C., McAnally, J.R., Shelton, J.M., Mireault, A. a, Bassel-Duby, R., and Olson, E.N. (2014). Prevention of muscular dystrophy in mice by CRISPR/Cas9-mediated editing of germline DNA. Science 345, 1184-1188

Lorthongpanich, C., Cheow, L.F., Balu, S., Quake, S.R., Knowles, B.B., Burkholder, W.F., Solter, D., and Messerschmidt, D.M. (2013). Single-cell DNA-methylation analysis reveals epigenetic chimerism in preimplantation embryos. Science 341, 1110–1112.

Losson, R., and Lacroute, F. (1979). Interference of nonsense mutations with eukaryotic messenger RNA stability. Proc. Natl. Acad. Sci. U. S. A. 76, 5134–5137.

Lucifero, D., Mann, M.R.W., Bartolomei, M.S., and Trasler, J.M. (2004). Gene-specific timing and epigenetic memory in oocyte imprinting. Hum. Mol. Genet. 13, 839–849.

141

Page 142: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Luger, K., Mäder, A., and Richmond, R. (1997). Crystal structure of the nucleosome core particle at 2.8 Å resolution. Nature 7, 251–260.

Ma, D.K., Jang, M.-H., Guo, J.U., Kitabatake, Y., Chang, M.-L., Pow-Anpongkul, N., Flavell, R. a, Lu, B., Ming, G.-L., and Song, H. (2009). Neuronal activity-induced Gadd45b promotes epigenetic DNA demethylation and adult neurogenesis. Science 323, 1074–1077.

Maeder, M.L., Linder, S.J., Cascio, V.M., Fu, Y., Ho, Q.H., and Joung, J.K. (2013). CRISPR RNA-guided activation of endogenous human genes. Nat. Methods 10, 977–979.

Maiti, A., and Drohat, A.C. (2011). Thymine DNA glycosylase can rapidly excise 5-formylcytosine and 5-carboxylcytosine: potential implications for active demethylation of CpG sites. J. Biol. Chem. 286, 35334–35338.

Mali, P., Yang, L., Esvelt, K., and Aach, J. (2013). RNA-guided human genome engineering via Cas9. Science (80), 823–826.

Marks, H., Kalkan, T., Menafra, R., and Denissov, S. (2012). The transcriptional and epigenomic foundations of ground state pluripotency. Cell 149, 590-604

Martin, G. (1981). Isolation of a pluripotent cell line from early mouse embryos cultured in medium conditioned by teratocarcinoma stem cells. Proc. Natl. Acad. Sci. U.S.A. 78, 7634–7638.

Mayer, W., Niveleau, a, Walter, J., Fundele, R., and Haaf, T. (2000). Demethylation of the zygotic paternal genome. Nature 403, 501–502.

McGrath, J., and Solter, D. (1984). Completion of mouse embryogenesis requires both the maternal and paternal genomes. Cell 37, 179–183.

McVey, M., and Lee, S.E. (2008). MMEJ repair of double-strand breaks (director’s cut): deleted sequences and alternative endings. Trends Genet. 24, 529–538.

Meissner, A., Mikkelsen, T.S., Gu, H., Wernig, M., Hanna, J., Sivachenko, A., Zhang, X., Bernstein, B.E., Nusbaum, C., Jaffe, D.B., Gnirke, A., Jaenisch, R., Lander, E.S. (2008). Genome-scale DNA methylation maps of pluripotent and differentiated cells. Nature 454, 766–770.

Mellén, M., Ayata, P., Dewell, S., Kriaucionis, S., and Heintz, N. (2012). MeCP2 binds to 5hmC enriched within active genes and accessible chromatin in the nervous system. Cell 151, 1417–1430.

Mermoud, J.E., Popova, B., Peters, A.H.F.M., Jenuwein, T., and Brockdorff, N. (2002). Histone H3 lysine 9 methylation occurs rapidly at the onset of random X chromosome inactivation. Curr. Biol. 12, 247–251.

Meshorer, E., and Misteli, T. (2006). Chromatin in pluripotent embryonic stem cells and differentiation. Nat. Rev. Mol. Cell Biol. 7, 540–546.

Meshorer, E., Yellajoshula, D., George, E., Scambler, P.J., Brown, D.T., and Misteli, T. (2006). Hyperdynamic plasticity of chromatin proteins in pluripotent embryonic stem cells. Dev. Cell 10, 105–116.

Messerschmidt, D. (2012). Should I stay or should I go: protection and maintenance of DNA methylation at imprinted genes. Epigenetics 969–975.

Messerschmidt, D.M., de Vries, W., Ito, M., Solter, D., Ferguson-Smith, A., and Knowles, B.B. (2012). Trim28 is required for epigenetic stability during mouse oocyte to embryo transition. Science 335, 1499–1502.

142

Page 143: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Métivier, R., Gallais, R., Tiffoche, C., Le Péron, C., Jurkowska, R.Z., Carmouche, R.P., Ibberson, D., Barath, P., Demay, F., Reid, G., Benes, V., Jeltsch, A., Gannon, F., and Salbert, G. (2008). Cyclical DNA methylation of a transcriptionally active promoter. Nature 452, 45–50.

Mikkelsen, T.S., Ku, M., Jaffe, D.B., Issac, B., Lieberman, E., Giannoukos, G., Alvarez, P., Brockman, W., Kim, T.-K., Koche, R.P., Lee, W., Mendenhall, E., O’Donovan, A., Presser, A., Russ, C., Xie, X., Meissner, A., Wernig, M., Jaenisch, R., Nusbaum, C., Lander, E.S., and Bernstein, B.E. (2007). Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 448, 553–560.

Mikkelsen, T.S., Hanna, J., Zhang, X., Ku, M., Wernig, M., Schorderet, P., Bernstein, B.E., Jaenisch, R., Lander, E.S., and Meissner, A. (2008). Dissecting direct reprogramming through integrative genomic analysis. Nature 454, 49–55.

Millar, C.B., Guy, J., Sansom, O.J., Selfridge, J., MacDougall, E., Hendrich, B., Keightley, P.D., Bishop, S.M., Clarke, A.R., and Bird, A. (2002). Enhanced CpG mutability and tumorigenesis in MBD4-deficient mice. Science 297, 403–405.

Miller, R.A., and Ruddle, F.H. (1976). Pluripotent teratocarcinoma-thymus somatic cell hybrids. Cell 9, 45–55.

Miller, J.C., Tan, S., Qiao, G., Barlow, K. a, Wang, J., Xia, D.F., Meng, X., Paschon, D.E., Leung, E., Hinkley, S.J., Dulay, G.P., Hua, K.L., Ankoudinova, I., Cost, G.J., Urnov, D.D., Zhang, H.S., Holmes, M.C., Zhang, Lei, Gregory, P.D., and Rebar, E.J. (2011). A TALE nuclease architecture for efficient genome editing. Nat. Biotechnol. 29, 143–148.

Minor, E. A., Court, B.L., Young, J.I., and Wang, G. (2013). Ascorbate Induces Ten-Eleven Translocation (Tet) Methylcytosine Dioxygenase-mediated Generation of 5-Hydroxymethylcytosine. J. Biol. Chem. 288, 13669–13674.

Mise, N., Fuchikami, T., Sugimoto, M., Kobayakawa, S., Ike, F., Ogawa, T., Tada, T., Kanaya, S., Noce, T., and Abe, K. (2008). Differences and similarities in the developmental status of embryo-derived stem cells and primordial germ cells revealed by global expression profiling. Genes Cells 13, 863–877.

Moore, J.K., and Haber, J.E. (1996). Cell cycle and genetic requirements of two pathways of nonhomologous end-joining repair of double-strand breaks in Saccharomyces cerevisiae. Mol. Cell. Biol. 16, 2164–2173.

Muramatsu, M., Kinoshita, K., Fagarasan, S., Yamada, S., Shinkai, Y., and Honjo, T. (2000). Class switch recombination and hypermutation require activation-induced cytidine deaminase (AID), a potential RNA editing enzyme. Cell 102, 553–563.

Murrell, A., Heeson, S., and Reik, W. (2004). Interaction between differentially methylated regions partitions the imprinted genes Igf2 and H19 into parent-specific chromatin loops. Nat. Genet. 36, 889–893.

Nabel, C.S., Jia, H., Ye, Y., Shen, L., Goldschmidt, H.L., Stivers, J.T., Zhang, Y., and Kohli, R.M. (2012). AID/APOBEC deaminases disfavor modified cytosines implicated in DNA demethylation. Nat. Chem. Biol. 8, 751–758.

Nakamura, T., Arai, Y., Umehara, H., Masuhara, M., Kimura, T., Taniguchi, H., Sekimoto, T., Ikawa, M., Yoneda, Y., Okabe, M., Tanaka, S., Shiota, K., Nakano, T. (2007). PGC7/Stella protects against DNA demethylation in early embryogenesis. Nat. Cell Biol. 9, 64–71.

Nakamura, T., Liu, Y.-J., Nakashima, H., Umehara, H., Inoue, K., Matoba, S., Tachibana, M., Ogura, A., Shinkai, Y., and Nakano, T. (2012). PGC7 binds histone H3K9me2 to protect against conversion of 5mC to 5hmC in early embryos. Nature 486, 415–419.

143

Page 144: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Nativio, R., Wendt, K.S., Ito, Y., Huddleston, J.E., Uribe-Lewis, S., Woodfine, K., Krueger, C., Reik, W., Peters, J.-M., and Murrell, A. (2009). Cohesin is required for higher-order chromatin conformation at the imprinted IGF2-H19 locus. PLoS Genet. 5, e1000739.

Nerlov, C., and Graf, T. (1998). PU.1 induces myeloid lineage commitment in multipotent hematopoietic progenitors. Genes Dev. 12, 2403–2412.

Nettersheim, D., Heukamp, L.C., Fronhoffs, F., Grewe, M.J., Haas, N., Waha, A., Honecker, F., Waha, A., Kristiansen, G., and Schorle, H. (2013). Analysis of TET expression/activity and 5mC oxidation during normal and malignant germ cell development. PLoS One 8, e82881.

Ng, H.-H., and Surani, M.A. (2011). The transcriptional and signalling networks of pluripotency. Nat. Cell Biol. 13, 490–496.

Niehrs, C., and Schäfer, A. (2012). Active DNA demethylation by Gadd45 and DNA repair. Trends Cell Biol. 22, 220–227.

Nishimasu, H., Ran, F.A., Hsu, P.D., Konermann, S., Shehata, S.I., Dohmae, N., Ishitani, R., Zhang, F., and Nureki, O. (2014). Crystal structure of Cas9 in complex with guide RNA and target DNA. Cell 156, 935–949.

Niu, X., Gulati, S., Edel, J.B., and deMello, A.J. (2008). Pillar-induced droplet merging in microfluidic circuits. Lab Chip 8, 1837–1841.

Niwa, H., Miyazaki, J., and Smith, a G. (2000). Quantitative expression of Oct-3/4 defines differentiation, dedifferentiation or self-renewal of ES cells. Nat. Genet. 24, 372–376.

Niwa, H., Masui, S., and Chambers, I. (2002). Phenotypic complementation establishes requirements for specific POU domain and generic transactivation function of Oct-3/4 in embryonic stem cells. Mol. Cell. 22, 1526-1536.

Noggle, S., Fung, H.-L., Gore, A., Martinez, H., Satriani, K.C., Prosser, R., Oum, K., Paull, D., Druckenmiller, S., Freeby, M., Greenberg, E., Zhang, Z., Goland, R., Sauer, M.V., Leibel, R.L., and Egli, D. (2011). Human oocytes reprogram somatic cells to a pluripotent state. Nature 478, 70–75.

Nordhoff, V., Hübner, K., Bauer, A., Orlova, I., Malapetsa, A., and Schöler, H.R. (2001). Comparative analysis of human, bovine, and murine Oct-4 upstream promoter sequences. Mamm. Genome 12, 309–317.

Ohinata, Y., Payer, B., O’Carroll, D., Ancelin, K., Ono, Y., Sano, M., Barton, S.C., Obukhanych, T., Nussenzweig, M., Tarakhovsky, A., Saitou, M., and Surani, M.A. (2005). Blimp1 is a critical determinant of the germ cell lineage in mice. Nature 436, 207–213.

Okano, M., Xie, S., and Li, E. (1998). Cloning and characterization of a family of novel mammalian DNA (cytosine-5) methyltransferases. Nat. Genet. 19, 219–220.

Okano, M., Bell, D.W., Haber, D. A., and Li, E. (1999). DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 99, 247–257.

Okashita, N., Kumaki, Y., Ebi, K., Nishi, M., Okamoto, Y., Nakayama, M., Hashimoto, S., Nakamura, T., Sugasawa, K., Kojima, N., Takada, T., Okano, M., and Seki, Y. (2014). PRDM14 promotes active DNA demethylation through the ten-eleven translocation (TET)-mediated base excision repair pathway in embryonic stem cells. Development 141, 269–280.

Ooi, S.K.T., and Bestor, T.H. (2008). The colorful history of active DNA demethylation. Cell 133, 1145–1148.

144

Page 145: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Ooi, S.K.T., Qiu, C., Bernstein, E., Li, K., Jia, D., Yang, Z., Erdjument-Bromage, H., Tempst, P., Lin, S.-P., Allis, C.D., Cheng, X., and Bestor, T.H. (2007). DNMT3L connects unmethylated lysine 4 of histone H3 to de novo methylation of DNA. Nature 448, 714–717.

Osei-Sarfo, K., and Gudas, L.J. (2014). Retinoic acid suppresses the canonical Wnt signaling pathway in embryonic stem cells and activates the noncanonical Wnt signaling pathway. Stem Cells 32, 2061-2071.

Oswald, J., Engemann, S., Lane, N., Mayer, W., Olek, a, Fundele, R., Dean, W., Reik, W., and Walter, J. (2000). Active demethylation of the paternal genome in the mouse zygote. Curr. Biol. 10, 475–478.

Otani, J., Kimura, H., Sharif, J., Endo, T. a, Mishima, Y., Kawakami, T., Koseki, H., Shirakawa, M., Suetake, I., and Tajima, S. (2013). Cell cycle-dependent turnover of 5-hydroxymethyl cytosine in mouse embryonic stem cells. PLoS One 8, e82961.

Pasini, D., Cloos, P. a C., Walfridsson, J., Olsson, L., Bukowski, J.-P., Johansen, J. V, Bak, M., Tommerup, N., Rappsilber, J., and Helin, K. (2010). JARID2 regulates binding of the Polycomb repressive complex 2 to target genes in ES cells. Nature 464, 306–310.

Pasque, V., Jullien, J., Miyamoto, K., Halley-Stott, R.P., and Gurdon, J.B. (2011). Epigenetic factors influencing resistance to nuclear reprogramming. Trends Genet. 27, 516–525.

Pastor, W. a, Pape, U.J., Huang, Y., Henderson, H.R., Lister, R., Ko, M., McLoughlin, E.M., Brudno, Y., Mahapatra, S., Kapranov, P., Tahiliani, M., Daley, G.Q., Liu, X.S., Ecker, J.R., Milos, P.M., Agarwal, S., and Rao, A. (2011). Genome-wide mapping of 5-hydroxymethylcytosine in embryonic stem cells. Nature 473, 394–397.

Pastor, W.A., Aravind, L., and Rao, A. (2013). TETonic shift: biological roles of TET proteins in DNA demethylation and transcription. Nat. Rev. Mol. Cell Biol. 14, 341–356.

Pattanayak, V., Lin, S., Guilinger, J.P., Ma, E., Doudna, J. a, and Liu, D.R. (2013). High-throughput profiling of off-target DNA cleavage reveals RNA-programmed Cas9 nuclease specificity. Nat. Biotechnol. 31, 839–843.

Patten, M.M., Ross, L., Curley, J.P., Queller, D.C., Bonduriansky, R., and Wolf, J.B. (2014). The evolution of genomic imprinting: theories, predictions and empirical tests. Heredity (Edinb). 113, 119–128.

Penny, G.D., Kay, G.F., Sheardown, S. a, Rastan, S., and Brockdorff, N. (1996). Requirement for Xist in X chromosome inactivation. Nature 379, 131–137.

Pereira, C.F., and Fisher, A.G. (2009). Heterokaryon-based reprogramming for pluripotency. Curr. Protoc. Stem Cell Biol. Chapter 4, Unit 4B.1.

Pereira, C.F., Terranova, R., Ryan, N.K., Santos, J., Morris, K.J., Cui, W., Merkenschlager, M., and Fisher, A.G. (2008). Heterokaryon-based reprogramming of human B lymphocytes for pluripotency requires Oct4 but not Sox2. PLoS Genet. 4, e1000170.

Pereira, C.F., Piccolo, F.M., Tsubouchi, T., Sauer, S., Ryan, N.K., Bruno, L., Landeira, D., Santos, J., Banito, A., and Gil, J. (2010). ESCs Require PRC2 to Direct the Successful Reprogramming of Differentiated Cells toward Pluripotency. Cell Stem Cell 6, 547–556.

Pereira, L., Yi, F., and Merrill, B.J. (2006). Repression of Nanog gene transcription by Tcf3 limits embryonic stem cell self-renewal. Mol. Cell. Biol. 26, 7479–7491.

Perez-pinera, P., Kocak, D.D., Vockley, C.M., Adler, A.F., Kabadi, A.M., Polstein, L.R., Thakore, P.I., Glass, K.A., Ousterout, D.G., Leong, K.W., et al. (2013). RNA-guided gene activation by CRISPR- Cas9–based transcription factors. Nat. Methods 10, 973-976.

145

Page 146: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Peterson, J. a, and Weiss, M.C. (1972). Expression of differentiated functions in hepatoma cell hybrids: induction of mouse albumin production in rat hepatoma-mouse fibroblast hybrids. Proc. Natl. Acad. Sci. U. S. A. 69, 571–575.

Pfaffeneder, T., Hackner, B., Truss, M., Münzel, M., Müller, M., Deiml, C. A., Hagemeier, C., and Carell, T. (2011). The discovery of 5-formylcytosine in embryonic stem cell DNA. Angew. Chem. Int. Ed. Engl. 50, 7008–7012.

Piccolo, F.M., Pereira, C.F., Cantone, I., Brown, K., Tsubouchi, T., Soza-Ried, J., Merkenschlager, M., and Fisher, A.G. (2011). Using heterokaryons to understand pluripotency and reprogramming. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 366, 2260–2265.

Piccolo, F.M., Bagci, H., Brown, K.E., Landeira, D., Soza-Ried, J., Feytout, A., Mooijman, D., Hajkova, P., Leitch, H.G., Tada, T., Kriaucionis, S., Dawlaty, M.M., Jaenisch, R., Merkenschlager, M., and Fisher, A.G. (2013). Different roles for Tet1 and Tet2 proteins in reprogramming-mediated erasure of imprints induced by EGC fusion. Mol. Cell 49, 1023–1033.

Polo, J.M., Anderssen, E., Walsh, R.M., Schwarz, B. a, Nefzger, C.M., Lim, S.M., Borkent, M., Apostolou, E., Alaei, S., Cloutier, J., Bar-Nur, O., Cheloufi, S., Stadtfeld, M., Figueroa, M.E., Robinton, D., Natesan, S., Melnick, A., Zhu, J., Ramaswamy, S., and Hochedlinger, K. (2012). A molecular roadmap of reprogramming somatic cells into iPS cells. Cell 151, 1617–1632.

Popp, C., Dean, W., Feng, S., Cokus, S.J., Andrews, S., Pellegrini, M., Jacobsen, S.E., and Reik, W. (2010). Genome-wide erasure of DNA methylation in mouse primordial germ cells is affected by AID deficiency. Nature 463, 1101–1105.

Qiu, P., Shandilya, H., D’Alessio, J.M., O’Connor, K., Durocher, J., and Gerard, G.F. (2004). Mutation detection using Surveyor nuclease. Biotechniques 36, 702–707.

Qu, Y., Hu, N., Xu, H., Yang, J., Xia, B., Zheng, X., and Yin, Z.Q. (2011). Somatic and stem cell pairing and fusion using a microfluidic array device. Microfluid. Nanofluidics 11, 633–641.

Quenneville, S., Verde, G., Corsinotti, A., Kapopoulou, A., Jakobsson, J., Offner, S., Baglivo, I., Pedone, P. V, Grimaldi, G., Riccio, A., and Trono, D. (2011). ZFP57 / KAP1 Recognize a Methylated Hexanucleotide to Affect Chromatin and DNA Methylation of Imprinting Control Regions. Mol. Cell 44, 361–372.

Radford, E.J., Ferrón, S.R., and Ferguson-Smith, A.C. (2011). Genomic imprinting as an adaptative model of developmental plasticity. FEBS Lett. 585, 2059–2066.

Rai, K., Huggins, I.J., James, S.R., Karpf, A.R., Jones, D. A., and Cairns, B.R. (2008). DNA demethylation in zebrafish involves the coupling of a deaminase, a glycosylase, and gadd45. Cell 135, 1201–1212.

Ramchandani, S., Bhattacharya, S.K., Cervoni, N., and Szyf, M. (1999). DNA methylation is a reversible biological signal. Proc. Natl. Acad. Sci. U. S. A. 96, 6107–6112.

Ran, F.A., Hsu, P.D., Wright, J., Agarwala, V., Scott, D. A., and Zhang, F. (2013a). Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 8, 2281–2308.

Ran, F.A., Hsu, P.D., Lin, C.-Y., Gootenberg, J.S., Konermann, S., Trevino, A.E., Scott, D. a, Inoue, A., Matoba, S., Zhang, Y., Zhang, F. (2013b). Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell 154, 1380–1389.

Rangam, G., Schmitz, K.-M., Cobb, A.J.A., and Petersen-Mahrt, S.K. (2012). AID enzymatic activity is inversely proportional to the size of cytosine C5 orbital cloud. PLoS One 7, e43279.

Rao, P., and Johnson, R. (1970). Mammalian cell fusion: studies on the regulation of DNA synthesis and mitosis. Nature 225, 159-164.

146

Page 147: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Reik, W. (2007). Stability and flexibility of epigenetic gene regulation in mammalian development. Nature 447, 425–432.

Rodríguez-Ubreva, J., Ciudad, L., Gómez-Cabrero, D., Parra, M., Bussmann, L.H., di Tullio, A., Kallin, E.M., Tegnér, J., Graf, T., and Ballestar, E. (2011). Pre-B cell to macrophage transdifferentiation without significant promoter DNA methylation changes. Nucleic Acids Res. 40, 1954-1968.

Rothbart, S.B., Krajewski, K., Nady, N., Tempel, W., Xue, S., Badeaux, A.I., Barsyte-Lovejoy, D., Martinez, J.Y., Bedford, M.T., Fuchs, S.M., Arrowsmith, C.H., and Strahl, B.D. (2012). Association of UHRF1 with methylated H3K9 directs the maintenance of DNA methylation. Nat. Struct. Mol. Biol. 19, 1155–1160.

Rouet, P., Smih, F., and Jasin, M. (1994). Introduction of double-strand breaks into the genome of mouse cells by expression of a rare-cutting endonuclease. Mol. Cell. Biol. 14, 8096-8106.

Sakaue, M., Ohta, H., Kumaki, Y., Oda, M., Sakaide, Y., Matsuoka, C., Yamagiwa, A., Niwa, H., Wakayama, T., and Okano, M. (2010). DNA methylation is dispensable for the growth and survival of the extraembryonic lineages. Curr. Biol. 20, 1452–1457.

Samavarchi-Tehrani, P., Golipour, A., David, L., Sung, H.-K., Beyer, T. a, Datti, A., Woltjen, K., Nagy, A., and Wrana, J.L. (2010). Functional genomics reveals a BMP-driven mesenchymal-to-epithelial transition in the initiation of somatic cell reprogramming. Cell Stem Cell 7, 64–77.

Sander, J.D., and Joung, J.K. (2014). CRISPR-Cas systems for editing, regulating and targeting genomes. Nat. Biotechnol. 32, 347–355.

Santos, F., Hendrich, B., Reik, W., and Dean, W. (2002). Dynamic reprogramming of DNA methylation in the early mouse embryo. Dev. Biol. 241, 172–182.

Saxonov, S., Berg, P., and Brutlag, D.L. (2006). A genome-wide analysis of CpG dinucleotides in the human genome distinguishes two distinct classes of promoters. Proc. Natl. Acad. Sci. U. S. A. 103, 1412–1417.

Schäfer, A., Karaulanov, E., Stapf, U., Döderlein, G., and Niehrs, C. (2013). Ing1 functions in DNA demethylation by directing Gadd45a to H3K4me3. Genes Dev. 27, 261–273.

Schiesser, S., Hackner, B., Pfaffeneder, T., Müller, M., Hagemeier, C., Truss, M., and Carell, T. (2012). Mechanism and stem-cell activity of 5-carboxycytosine decarboxylation determined by isotope tracing. Angew. Chem. Int. Ed. Engl. 51, 6516–6520.

Schnerch, A., Cerdan, C., and Bhatia, M. (2010). Distinguishing between mouse and human pluripotent stem cell regulation: the best laid plans of mice and men. Stem Cells 28, 419–430.

Scholer, H.R., Dressler, G.R., Rohdewohid, H., and Gruss, P. (1990). Oct-t: a germnline-specific transcription factor mapping to the mouse t-complex. Embo J. 9, 2185–2195.

Schwank, G., Koo, B.-K., Sasselli, V., Dekkers, J.F., Heo, I., Demircan, T., Sasaki, N., Boymans, S., Cuppen, E., van der Ent, C.K., Nieuwenhuis, E.E., Beekman, J.M., Clevers, H. (2013). Functional repair of CFTR by CRISPR/Cas9 in intestinal stem cell organoids of cystic fibrosis patients. Cell Stem Cell 13, 653–658.

Seisenberger, S., Andrews, S., Krueger, F., Arand, J., Walter, J., Santos, F., Popp, C., Thienpont, B., Dean, W., and Reik, W. (2012). The dynamics of genome-wide DNA methylation reprogramming in mouse primordial germ cells. Mol. Cell 48, 849–862.

Sekiya, S., and Suzuki, A. (2011). Direct conversion of mouse fibroblasts to hepatocyte-like cells by defined factors. Nature 475, 390–393.

147

Page 148: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Shalem, O., Sanjana, N.E., Hartenian, E., Shi, X., Scott, D. a, Mikkelsen, T.S., Heckl, D., Ebert, B.L., Root, D.E., Doench, J.G., and Zhang F. (2014). Genome-scale CRISPR-Cas9 knockout screening in human cells. Science 343, 84–87.

Shapiro, E., Biezuner, T., and Linnarsson, S. (2013). Single-cell sequencing-based technologies will revolutionize whole-organism science. Nat. Rev. Genet. 14, 618–630.

Sharif, J., Muto, M., Takebayashi, S., Suetake, I., Iwamatsu, A., Endo, T. a, Shinga, J., Mizutani-Koseki, Y., Toyoda, T., Okamura, K., Tajima, S., Mitsuya, K., Okano, M., and Koseki, H. (2007). The SRA protein Np95 mediates epigenetic inheritance by recruiting Dnmt1 to methylated DNA. Nature 450, 908–912.

Shen, B., Zhang, W., Zhang, J., Zhou, J., Wang, J., Chen, L., Wang, L., Hodgkins, A., Iyer, V., Huang, X., and Skarnes W.C. (2014). Efficient genome modification by CRISPR-Cas9 nickase with minimal off-target effects. Nat. Methods. 11(4), 399-402

Shen, L., Wu, H., Diep, D., Yamaguchi, S., D’Alessio, A.C., Fung, H.-L., Zhang, K., and Zhang, Y. (2013). Genome-wide analysis reveals TET- and TDG-dependent 5-methylcytosine oxidation dynamics. Cell 153, 692–706.

Shen, X., Kim, W., Fujiwara, Y., Simon, M.D., Liu, Y., Mysliwiec, M.R., Yuan, G.-C., Lee, Y., and Orkin, S.H. (2009). Jumonji modulates polycomb activity and self-renewal versus differentiation of stem cells. Cell 139, 1303–1314.

Shimamoto, R., Amano, N., Ichisaka, T., Watanabe, A., Yamanaka, S., and Okita, K. (2014). Generation and characterization of induced pluripotent stem cells from Aid-deficient mice. PLoS One 9, e94735.

Shovlin, T.C., Durcova-Hills, G., Surani, A., and McLaren, A. (2008). Heterogeneity in imprinted methylation patterns of pluripotent embryonic germ cells derived from pre-migratory mouse germ cells. Dev. Biol. 313, 674–681.

Silva, J., Chambers, I., Pollard, S., and Smith, A. (2006). Nanog promotes transfer of pluripotency after cell fusion. Nature 441, 997–1001.

Simonsson, S., and Gurdon, J. (2004). DNA demethylation is necessary for the epigenetic reprogramming of somatic cell nuclei. Nat. Cell Biol. 6, 984–990.

Skelley, A.M., Kirak, O., Suh, H., Jaenisch, R., and Voldman, J. (2009). Microfluidic control of cell pairing and fusion. Nat. Methods 6, 147–152.

Slotkin, R.K., Vaughn, M., Borges, F., Tanurdzić, M., Becker, J.D., Feijó, J. a, and Martienssen, R. A. (2009). Epigenetic reprogramming and small RNA silencing of transposable elements in pollen. Cell 136, 461–472.

Smallwood, S. a, Lee, H.J., Angermueller, C., Krueger, F., Saadeh, H., Peat, J., Andrews, S.R., Stegle, O., Reik, W., and Kelsey, G. (2014). Single-cell genome-wide bisulfite sequencing for assessing epigenetic heterogeneity. Nat. Methods 11, 817-820.

Smith, Z.D., and Meissner, A. (2013). DNA methylation: roles in mammalian development. Nat. Rev. Genet. 14, 204–220.

Smith, C., Gore, A., Yan, W., Abalde-Atristain, L., Li, Z., He, C., Wang, Y., Brodsky, R. a, Zhang, K., Cheng, L., and Ye, Z. (2014). Whole-Genome Sequencing Analysis Reveals High Specificity of CRISPR/Cas9 and TALEN-Based Genome Editing in Human iPSCs. Cell Stem Cell 15, 12–13.

Smith, F.M., Garfield, a S., and Ward, A. (2006a). Regulation of growth and metabolism by imprinted genes. Cytogenet. Genome Res. 113, 279–291.

148

Page 149: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Smith, J., Grizot, S., Arnould, S., Duclert, A., Epinat, J.-C., Chames, P., Prieto, J., Redondo, P., Blanco, F.J., Bravo, J., Montoya, G., Paques, F., and Duchateau, P. (2006b). A combinatorial approach to create artificial homing endonucleases cleaving chosen sequences. Nucleic Acids Res. 34, e149.

Smith, S.S., Kaplan, B.E., Sowers, L.C., and Newman, E.M. (1992). Mechanism of human methyl-directed DNA methyltransferase and the fidelity of cytosine methylation. Proc. Natl. Acad. Sci. U. S. A. 89, 4744–4748.

Solberg, N., Machon, O., Machonova, O., and Krauss, S. (2012). Mouse Tcf3 represses canonical Wnt signaling by either competing for β-catenin binding or through occupation of DNA-binding sites. Mol. Cell. Biochem. 365, 53–63.

Son, J., Shen, S.S., Margueron, R., and Reinberg, D. (2013). Nucleosome-binding activities within JARID2 and EZH1 regulate the function of PRC2 on chromatin. Genes Dev. 27, 2663–2677.

Song, C.-X., Szulwach, K.E., Dai, Q., Fu, Y., Mao, S.-Q., Lin, L., Street, C., Li, Y., Poidevin, M., Wu, H., Gao, J., Liu, P., Li, L., Xu, G.L., Jin, P., and He, C. (2013). Genome-wide profiling of 5-formylcytosine reveals its roles in epigenetic priming. Cell 153, 678–691.

Sorek, R., Lawrence, C.M., and Wiedenheft, B. (2013). CRISPR-mediated adaptive immune systems in bacteria and archaea. Annu. Rev. Biochem. 82, 237–266.

Soufi, A., Donahue, G., and Zaret, K.S. (2012). Facilitators and impediments of the pluripotency reprogramming factors’ initial engagement with the genome. Cell 151, 994–1004.

Soza-Ried, J., and Fisher, A.G. (2012). Reprogramming somatic cells towards pluripotency by cellular fusion. Curr. Opin. Genet. Dev. 22, 459–465.

Spruijt, C.G., Gnerlich, F., Smits, A.H., Pfaffeneder, T., Jansen, P.W.T.C., Bauer, C., Münzel, M., Wagner, M., Müller, M., Khan, F., Eber;, H.C., Mensinga, A., Brinkman, A.B., Lephikov, K., Muller, U., Walter, J., Boelens, R., van Ingen, H., Leonhardt, H., Carell, T., and Vermeulen M. (2013). Dynamic readers for 5-(hydroxy)methylcytosine and its oxidized derivatives. Cell 152, 1146–1159.

Sternberg, S.H., Redding, S., Jinek, M., Greene, E.C., and Doudna, J. a (2014). DNA interrogation by the CRISPR RNA-guided endonuclease Cas9. Nature 507, 62–67.

Stock, J.K., Giadrossi, S., Casanova, M., Brookes, E., Vidal, M., Koseki, H., Brockdorff, N., Fisher, A.G., and Pombo, A. (2007). Ring1-mediated ubiquitination of H2A restrains poised RNA polymerase II at bivalent genes in mouse ES cells. Nat. Cell Biol. 9, 1428–1435.

Stresemann, C., and Lyko, F. (2008). Modes of action of the DNA methyltransferase inhibitors azacytidine and decitabine. Int. J. Cancer 123, 8–13.

Suetake, I., Shinozaki, F., Miyagawa, J., Takeshima, H., and Tajima, S. (2004). DNMT3L stimulates the DNA methylation activity of Dnmt3a and Dnmt3b through a direct interaction. J. Biol. Chem. 279, 27816–27823.

Surani, M., Barton, S., and Norris, M. (1984). Development of reconstituted mouse eggs suggests imprinting of the genome during gametogenesis. Nature 308, 548-550.

Suzuki, K., Yu, C., Qu, J., Li, M., Yao, X., Yuan, T., Goebl, A., Tang, S., Ren, R., Aizawa, E., Zhang, F., Xu, X., Soligalla, R.D., Chen, F., Kim, J., Kim, N.Y., Liao, H.K., Benner, C., Esteban, C.R., Jin, Y., Liu, G.H., Li, Y., Izpisua-Belmonte, J.C. (2014). Targeted gene correction minimally impacts whole-genome mutational load in human-disease-specific induced pluripotent stem cell clones. Cell Stem Cell 15, 31–36.

Tachibana, M., Matsumura, Y., Fukuda, M., Kimura, H., and Shinkai, Y. (2008). G9a/GLP complexes independently mediate H3K9 and DNA methylation to silence transcription. EMBO J. 27, 2681–2690.

149

Page 150: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Tachibana, M., Amato, P., Sparman, M., Gutierrez, N.M., Tippner-Hedges, R., Ma, H., Kang, E., Fulati, A., Lee, H.-S., Sritanaudomchai, H., Masterson, K., Larson, J., Eaton, D., Sadler-Fredd, K., Battaglia, D., Lee, D., Wu, D., Jensen, J., Patton, P., Gokhale, S., Stouffer, R.L., Wolf, D., and Mitalipov, S. (2013). Human embryonic stem cells derived by somatic cell nuclear transfer. Cell 153, 1228–1238.

Tada, M., Tada, T., Lefebvre, L., Barton, S.C., and Surani, M. A. (1997). Embryonic germ cells induce epigenetic reprogramming of somatic nucleus in hybrid cells. EMBO J. 16, 6510–6520.

Tada, M., Takahama, Y., Abe, K., Nakatsuji, N., and Tada, T. (2001). Nuclear reprogramming of somatic cells by in vitro hybridization with ES cells. Curr. Biol. 11, 1553–1558.

Tada, T., Tada, M., Hilton, K., Barton, S.C., Sado, T., Takagi, N., and Surani, M. A. (1998). Epigenotype switching of imprintable loci in embryonic germ cells. Dev. Genes Evol. 207, 551–561.

Tahiliani, M., Koh, K.P., Shen, Y., Pastor, W. a, Bandukwala, H., Brudno, Y., Agarwal, S., Iyer, L.M., Liu, D.R., Aravind, L., and Rao, A. (2009). Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 324, 930–935.

Takagi, N. (1993). Variable X chromosome inactivation patterns in near-tetraploid murine EC x somatic cell hybrid cells differentiated in vitro. Genetica 88, 107–117.

Takahashi, K., and Yamanaka, S. (2006). Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676.

Taylor, S.M., and Jones, P.A. (1979). Multiple New Phenotypes Induced in 10t1/2 and 3t3 cells treated with 5-azayctidine. Cell 17, 771–779.

Terns, R.M., and Terns, M.P. (2014). CRISPR-based technologies: prokaryotic defense weapons repurposed. Trends Genet. 30, 111–118.

Theunissen, T.W., and Jaenisch, R. (2014). Molecular Control of Induced Pluripotency. Cell Stem Cell 14, 720–734.

Thomson, J. A. (1998). Embryonic Stem Cell Lines Derived from Human Blastocysts. Science (80-. ). 282, 1145–1147.

Thomson, J.P., Skene, P.J., Selfridge, J., Clouaire, T., Guy, J., Webb, S., Kerr, A.R.W., Deaton, A., Andrews, R., James, K.D., Turner, D.J., Illingworth, R., Bird, A. (2010). CpG islands influence chromatin structure via the CpG-binding protein Cfp1. Nature 464, 1082–1086.

Thuan, N. Van, Kishigami, S., and Wakayama, T. (2010). How to improve the success rate of mouse cloning technology. J. Reprod. Dev. 56, 20–30.

Tsubouchi, T., Soza-Ried, J., Brown, K., Piccolo, F.M., Cantone, I., Landeira, D., Bagci, H., Hochegger, H., Merkenschlager, M., and Fisher, A.G. (2013). DNA synthesis is required for reprogramming mediated by stem cell fusion. Cell 152, 873–883.

Tsumura, A., Hayakawa, T., Kumaki, Y., Takebayashi, S., Sakaue, M., Matsuoka, C., Shimotohno, K., Ishikawa, F., Li, E., Ueda, H.R., Nakayama, J., and Okano, M. (2006). Maintenance of self-renewal ability of mouse embryonic stem cells in the absence of DNA methyltransferases Dnmt1, Dnmt3a and Dnmt3b. Genes Cells 11, 805–814.

Urnov, F.D., Miller, J.C., Lee, Y.-L., Beausejour, C.M., Rock, J.M., Augustus, S., Jamieson, A.C., Porteus, M.H., Gregory, P.D., and Holmes, M.C. (2005). Highly efficient endogenous human gene correction using designed zinc-finger nucleases. Nature 435, 646–651.

Valinluck, V., and Sowers, L.C. (2007). Endogenous cytosine damage products alter the site selectivity of human DNA maintenance methyltransferase DNMT1. Cancer Res. 67, 946–950.

150

Page 151: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Valinluck, V., Tsai, H.-H., Rogstad, D.K., Burdzy, A., Bird, A., and Sowers, L.C. (2004). Oxidative damage to methyl-CpG sequences inhibits the binding of the methyl-CpG binding domain (MBD) of methyl-CpG binding protein 2 (MeCP2). Nucleic Acids Res. 32, 4100–4108.

Vella, P., Scelfo, A., Jammula, S., Chiacchiera, F., Williams, K., Cuomo, A., Roberto, A., Christensen, J., Bonaldi, T., Helin, K., and Pasini, D. (2013). Tet proteins connect the O-linked N-acetylglucosamine transferase Ogt to chromatin in embryonic stem cells. Mol. Cell 49, 645–656.

Veres, A., Gosis, B.S., Ding, Q., Collins, R., Ragavendran, A., Brand, H., Erdin, S., Talkowski, M.E., and Musunuru, K. (2014). Low Incidence of Off-Target Mutations in Individual CRISPR-Cas9 and TALEN Targeted Human Stem Cell Clones Detected by Whole-Genome Sequencing. Cell Stem Cell 15, 27–30.

Vierbuchen, T., Ostermeier, A., Pang, Z.P., Kokubu, Y., Südhof, T.C., and Wernig, M. (2010). Direct conversion of fibroblasts to functional neurons by defined factors. Nature 463, 1035–1041.

Vincent, J.J., Huang, Y., Chen, P.-Y., Feng, S., Calvopiña, J.H., Nee, K., Lee, S. a, Le, T., Yoon, A.J., Faull, K., Fan, G., Rao, A., Jacobsen, S.E., Pellegrini, M., Clark, A.T. (2013). Stage-specific roles for tet1 and tet2 in DNA demethylation in primordial germ cells. Cell Stem Cell 12, 470–478.

Visvader, J., Elefanty, A., Strasser, A., and Adams, J. (1992). GATA-1 but not SCL induces megakaryocytic differentiation in an early myeloid line. EMBO J. 1, 4557–4564.

Wakayama, T., Perry, a C., Zuccotti, M., Johnson, K.R., and Yanagimachi, R. (1998). Full-term development of mice from enucleated oocytes injected with cumulus cell nuclei. Nature 394, 369–374.

Wang, J., and Lu, C. (2006). Microfluidic cell fusion under continuous direct current voltage. Appl. Phys. Lett. 89, 234102.

Wang, H., Yang, H., Shivalila, C.S., Dawlaty, M.M., Cheng, A.W., Zhang, F., and Jaenisch, R. (2013a). One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell 153, 910–918.

Wang, T., Wu, H., Li, Y., Szulwach, K.E., Lin, L., Li, X., Chen, I.-P., Goldlust, I.S., Chamberlain, S.J., Dodd, A., Gong, H., Ananiev, G., Han, J.W., Yoon, Y.S., Rudd, M.K., Yu, M., Song, C.X., He, C., Chang, Q., Warren, S.T., Jin, P. (2013b). Subtelomeric hotspots of aberrant 5-hydroxymethylcytosine-mediated epigenetic modifications during reprogramming to pluripotency. Nat. Cell Biol. 15, 700–711.

Wang, T., Wei, J.J., Sabatini, D.M., and Lander, E.S. (2014). Genetic screens in human cells using the CRISPR-Cas9 system. Science 343, 80–84.

Watanabe, D., Suetake, I., Tada, T., and Tajima, S. (2002). Stage- and cell-specific expression of Dnmt3a and Dnmt3b during embryogenesis. Mech. Dev. 118, 187–190.

Weber, M., Hellmann, I., Stadler, M.B., Ramos, L., Pääbo, S., Rebhan, M., and Schübeler, D. (2007). Distribution, silencing potential and evolutionary impact of promoter DNA methylation in the human genome. Nat. Genet. 39, 457–466.

Weiss, M.C., and Chaplain, M. (1971). Expression of differentiated functions in hepatoma cell hybrids: reappearance of tyrosine aminotransferase inducibility after the loss of chromosomes. Proc. Natl. Acad. Sci. U. S. A. 68, 3026–3030.

Wernig, M., Meissner, A., Foreman, R., Brambrink, T., Ku, M., Hochedlinger, K., Bernstein, B.E., and Jaenisch, R. (2007). In vitro reprogramming of fibroblasts into a pluripotent ES-cell-like state. Nature 448, 318–324.

Whitesides, G.M. (2006). The origins and the future of microfluidics. Nature 442, 368–373.

151

Page 152: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Whyte, W. A., Orlando, D. a, Hnisz, D., Abraham, B.J., Lin, C.Y., Kagey, M.H., Rahl, P.B., Lee, T.I., and Young, R. A. (2013). Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319.

Wilkinson, L.S., Davies, W., and Isles, A.R. (2007). Genomic imprinting effects on brain development and function. Nat. Rev. Neurosci. 8, 832–843.

Williams, K., Christensen, J., Pedersen, M.T., Johansen, J. V, Cloos, P. a C., Rappsilber, J., and Helin, K. (2011). TET1 and hydroxymethylcytosine in transcription and DNA methylation fidelity. Nature 473, 343–348.

Wilmut, I., Schnieke, A., McWhir, J., Kind, a J., and Campbel, K.H. (1997). Viable offspring derived from fetal and adult mammalian cells. Nature 385, 810–813.

De Wit, E., Bouwman, B.A.M., Zhu, Y., Klous, P., Splinter, E., Verstegen, M.J.A.M., Krijger, P.H.L., Festuccia, N., Nora, E.P., Welling, M., Heard, E., Geijsen, N., Poot, R.A., Chambers, I., and de Laat, W. (2013). The pluripotent genome in three dimensions is shaped around pluripotency factors. Nature 501, 227–231.

Wong, C.C., Gaspar-Maia, A., Ramalho-Santos, M., and Reijo Pera, R. A. (2008). High-efficiency stem cell fusion-mediated assay reveals Sall4 as an enhancer of reprogramming. PLoS One 3, e1955.

Wossidlo, M., Arand, J., Sebastiano, V., Lepikhov, K., Boiani, M., Reinhardt, R., Schöler, H., and Walter, J. (2010). Dynamic link of DNA demethylation, DNA strand breaks and repair in mouse zygotes. EMBO J. 29, 1877–1888.

Wossidlo, M., Nakamura, T., Lepikhov, K., Marques, C.J., Zakhartchenko, V., Boiani, M., Arand, J., Nakano, T., Reik, W., and Walter, J. (2011). 5-Hydroxymethylcytosine in the mammalian zygote is linked with epigenetic reprogramming. Nat. Commun. 2, 241.

Wray, J., Kalkan, T., and Smith, A.G. (2010). The ground state of pluripotency. Biochem. Soc. Trans. 38, 1027–1032.

Wray, J., Kalkan, T., Gomez-Lopez, S., Eckardt, D., Cook, A., Kemler, R., and Smith, A. (2011). Inhibition of glycogen synthase kinase-3 alleviates Tcf3 repression of the pluripotency network and increases embryonic stem cell resistance to differentiation. Nat. Cell Biol. 13, 838–845.

Wu, H., and Zhang, Y. (2014). Reversing DNA methylation: mechanisms, genomics, and biological functions. Cell 156, 45–68.

Wu, S.C., and Zhang, Y. (2010). Active DNA demethylation: many roads lead to Rome. Nat. Rev. Mol. Cell Biol. 11, 607–620.

Wu, H., D’Alessio, A.C., Ito, S., Xia, K., Wang, Z., Cui, K., Zhao, K., Eve Sun, Y., and Zhang, Y. (2011). Dual functions of Tet1 in transcriptional regulation in mouse embryonic stem cells. Nature 473, 389-393.

Wu, X., Johansen, J.V., and Helin, K. (2013a). Fbxl10/Kdm2b recruits polycomb repressive complex 1 to CpG islands and regulates H2A ubiquitylation. Mol. Cell 49, 1134–1146.

Wu, Y., Liang, D., Wang, Y., Bai, M., Tang, W., Bao, S., Yan, Z., Li, D., and Li, J. (2013b). Correction of a genetic disease in mouse via use of CRISPR-Cas9. Cell Stem Cell 13, 659–662.

Xie, H., Ye, M., Feng, R., and Graf, T. (2004). Stepwise reprogramming of B cells into macrophages. Cell 117, 663–676.

Xu, C., Inokuma, M.S., Denham, J., Golds, K., Kundu, P., Gold, J.D., and Carpenter, M.K. (2001). Feeder-free growth of undifferentiated human embryonic stem cells. Nat. Biotechnol. 19, 971–974.

152

Page 153: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

Xu, Y., Wu, F., Tan, L., Kong, L., Xiong, L., Deng, J., Barbera, A.J., Zheng, L., Zhang, H., Huang, S., Min, J., Nicholson, T., Chen, T., Xu, G., Shi, Y., Zhang, K., and Shi Y.G. (2011). Genome-wide regulation of 5hmC, 5mC, and gene expression by Tet1 hydroxylase in mouse embryonic stem cells. Mol. Cell 42, 451–464.

Xue, W., Chen, S., Yin, H., Tammela, T., Papagiannakopoulos, T., Joshi, N.S., Cai, W., Yang, G., Bronson, R., Crowley, D.G., Zhang, F., Anderson, D.G., Sharp, P.A., and Jacks, T. (2014). CRISPR-mediated direct mutation of cancer genes in the mouse liver. Nature 514, 380-384.

Yamaguchi, S., Hong, K., Liu, R., Shen, L., Inoue, A., Diep, D., Zhang, K., and Zhang, Y. (2012). Tet1 controls meiosis by regulating meiotic gene expression. Nature 492, 443–447.

Yamaguchi, S., Hong, K., Liu, R., Inoue, A., Shen, L., Zhang, K., and Zhang, Y. (2013a). Dynamics of 5-methylcytosine and 5-hydroxymethylcytosine during germ cell reprogramming. Cell Res. 23, 329-339.

Yamaguchi, S., Shen, L., Liu, Y., Sendler, D., and Zhang, Y. (2013b). Role of Tet1 in erasure of genomic imprinting. Nature 504, 460–464.

Yamanaka, S. (2009). A fresh look at iPS cells. Cell 137, 13–17.

Yamanaka, S., and Blau, H.M. (2010). Nuclear reprogramming to a pluripotent state by three approaches. Nature 465, 704–712.

Yang, H., Wang, H., Shivalila, C.S., Cheng, A.W., Shi, L., and Jaenisch, R. (2013). One-step generation of mice carrying reporter and conditional alleles by CRISPR/Cas-mediated genome engineering. Cell 154, 1370–1379.

Yarmush, M.L., and King, K.R. (2009). Living-cell microarrays. Annu. Rev. Biomed. Eng. 11, 235–257.

Yeom, Y.I., Fuhrmann, G., Ovitt, C.E., Brehm, a, Ohbo, K., Gross, M., Hübner, K., and Schöler, H.R. (1996). Germline regulatory element of Oct-4 specific for the totipotent cycle of embryonal cells. Development 122, 881–894.

Yi, F., Pereira, L., Hoffman, J. a, Shy, B.R., Yuen, C.M., Liu, D.R., and Merrill, B.J. (2011). Opposing effects of Tcf3 and Tcf1 control Wnt stimulation of embryonic stem cell self-renewal. Nat. Cell Biol. 13, 762–770.

Yildirim, O., Li, R., Hung, J.-H., Chen, P.B., Dong, X., Ee, L.-S., Weng, Z., Rando, O.J., and Fazzio, T.G. (2011). Mbd3/NURD complex regulates expression of 5-hydroxymethylcytosine marked genes in embryonic stem cells. Cell 147, 1498–1510.

Yin, H., Xue, W., Chen, S., Bogorad, R.L., Benedetti, E., Grompe, M., Koteliansky, V., Sharp, P. a, Jacks, T., and Anderson, D.G. (2014). Genome editing with Cas9 in adult mice corrects a disease mutation and phenotype. Nat. Biotechnol. 32, 551–553.

Yin, R., Mao, S.-Q., Zhao, B., Chong, Z., Yang, Y., Zhao, C., Zhang, D., Huang, H., Gao, J., Li, Z., Jiao, Y., Li, C., Liu, S., Wu, D., Gu, W., Yang, Y.G., Xu, G.L., and Wang, H. (2013). Ascorbic Acid Enhances Tet-Mediated 5-Methylcytosine Oxidation and Promotes DNA Demethylation in Mammals. J. Am. Chem. Soc. 135, 10396-10403.

Ying, Q.-L., Wray, J., Nichols, J., Batlle-Morera, L., Doble, B., Woodgett, J., Cohen, P., and Smith, A. (2008). The ground state of embryonic stem cell self-renewal. Nature 453, 519–523.

Yoshimizu, T., Sugiyama, N., De Felice, M., Yeom, Y.I., Ohbo, K., Masuko, K., Obinata, M., Abe, K., Schöler, H.R., and Matsui, Y. (1999). Germline-specific expression of the Oct-4/green fluorescent protein (GFP) transgene in mice. Dev. Growth Differ. 41, 675–684.

153

Page 154: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Bibliography

You, J.S., Kelly, T.K., De Carvalho, D.D., Taberlay, P.C., Liang, G., and Jones, P. A. (2011). OCT4 establishes and maintains nucleosome-depleted regions that provide additional layers of epigenetic regulation of its target genes. Proc. Natl. Acad. Sci. U. S. A. 108, 14497–14502.

Yu, M., Hon, G.C., Szulwach, K.E., Song, C.-X., Zhang, L., Kim, A., Li, X., Dai, Q., Shen, Y., Park, B., et al. (2012). Base-resolution analysis of 5-hydroxymethylcytosine in the mammalian genome. Cell 149, 1368–1380.

Zaret, K.S., and Carroll, J.S. (2011). Pioneer transcription factors: establishing competence for gene expression. Genes Dev. 25, 2227–2241.

Zhang, L., Lu, X., Lu, J., Liang, H., Dai, Q., Xu, G.-L., Luo, C., Jiang, H., and He, C. (2012). Thymine DNA glycosylase specifically recognizes 5-carboxylcytosine-modified DNA. Nat. Chem. Biol. 8, 328–330.

Zhou, Q., Brown, J., Kanarek, A., Rajagopal, J., and Melton, D. A. (2008). In vivo reprogramming of adult pancreatic exocrine cells to beta-cells. Nature 455, 627–632.

Zhu, J., Adli, M., Zou, J.Y., Verstappen, G., Coyne, M., Zhang, X., Durham, T., Miri, M., Deshpande, V., De Jager, P.L., Bennett, D.A., Houmard, J.A., Muoio, D.M., Onder, T.T., Camahort, R., Cowan, C.A., Meissner, A., Epstein, C.B., Shoresh, N., Bernstein, B.E. (2013). Genome-wide chromatin state transitions associated with developmental and environmental cues. Cell 152, 642–654.

Zilberman, D., Coleman-Derr, D., Ballinger, T., and Henikoff, S. (2008). Histone H2A.Z and DNA methylation are mutually antagonistic chromatin marks. Nature 456, 125–129.

Zuo, X., Sheng, J., Lau, H.-T., McDonald, C.M., Andrade, M., Cullen, D.E., Bell, F.T., Iacovino, M., Kyba, M., Xu, G., and Li, X. (2012). Zinc finger protein ZFP57 requires its co-factor to recruit DNA methyltransferases and maintains DNA methylation imprint in embryonic stem cells via its transcriptional repression domain. J. Biol. Chem. 287, 2107–2118.

Zvetkova, I., Apedaile, A., Ramsahoye, B., Mermoud, J.E., Crompton, L. a, John, R., Feil, R., and Brockdorff, N. (2005). Global hypomethylation of the genome in XX embryonic stem cells. Nat. Genet. 37, 1274–1279.

154

Page 155: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Appendix

Appendix Oligonucleotides

Table 1. Primers for transcript analysis by quantitative RT-PCR

Human-specific primers Gene Sequence 5’ – 3’

GAPDH Forward TCT GCT CCT CCT GTT CGA CA Reverse AAA AGC AGC CCT GGT GAC C

OCT4 Forward TCG AGA ACC GAG TGA GAG GC Reverse CAC ACT CGG ACC ACA TCC TTC

NANOG Forward CCA ACA TCC TGA ACC TCA GCT AC Reverse GCC TTC TGC GTC ACA CCA TT

CRIPTO Forward AGA AGT GTT CCC TGT GTA AAT GCT G Reverse CAC GAG GTG CTC ATC CAT CA

CD19 Forward GCT CAA GAC GCT GGA AAG TAT TAT T Reverse GAT AAG CCA AAG TCA CAG CTG AGA

CD45 Forward CCC CAT GAA CGT TAC CAT TTG Reverse GAT AGT CTC CAT TGT GAA AAT AGG CC

Mouse-specific primers Gene Sequence 5’ – 3’

Ubc Forward GTC TGC TGT GTG AGG ACT GC Reverse GTC TTG CCT GTC AGG GTC TT

Oct4 Forward CGT GGA GAC TTT GCA GCC TG Reverse GCT TGG CAA ACT GTT CTA GCT CCT

Nanog Forward GAA CTA TTC TTG CTT ACA AGG GTC TGC Reverse GCA TCT TCT GCT TCC TGG CAA

Tet1 Forward GAG CCT GTT CCT CGA TGT GG Reverse CAA ACC CAC CTG AGG CTG TT

Tet2 Forward TGT TGT TGT CAG GGT GAG AAT C Reverse TCT TGC TTC TGG CAA ACT TAC A

Prickle1 Forward ATG GAT TCT TTG GCG TTG TC Reverse TGA CGG TCT TGG CTT GCT

Fzd2 Forward GAC ACC AGG GCT GAA GAG TG Reverse AAG GGC ACT TAG AAA AGT CGA G

155

Page 156: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Appendix

Wnt9a Forward CGA GTG GAC TTC CAC AAC AA Reverse GGC ATT TGC AAG TGG TTT C

Table 2. Primers for bisulfite sequencing analysis

Human-specific primers Locus Sequence 5’ – 3’

OCT4 Forward AAA GTT TTT GTG GGG GAT TTG TAT T Reverse AAA ACC TAA CCC AAC CCC CAA

NANOG Forward TTA ATT TAT TGG GAT TAT AGG GGT G Reverse AAA CCT AAA AAC AAA CCC AAC AAC

CRIPTO Forward GGA GGA TTG AAA TGT TAG GTG AG Reverse AAA TTT ATC TCA ACC TCC CAA CTC

OCT4 (Bhutani et al., 2010) Forward GGA GAG GGG GTT AAG TAT TTG GGT TTT Reverse TCC ACT TTA TTA CCC AAA CTA A

Mouse-specific primers

Locus Sequence 5’ – 3’

Oct4-Gfp Forward GGG GTT AGA GGT TAA GGT TAG AGG Reverse ACC AAA ATA AAC ACC ACC CC

H19 Forward AAG GAG ATT ATG TTT TAT TTT TGG A Reverse AAA AAA ACT CAA TCA ATT ACA ATC C

Peg1 Forward GAT TAG AGA TTT ATA AGG AAA GAG Reverse CAA CAA AAA CAA CAA ACA ACA AC

LINE1 Forward GTT AGA GAA TTT GAT AGT TTT TGG AAT AGG Reverse TCA AAC ACT ATA TTA CTT TAA CAA TTC CCA

Peg3 Forward TTG ATA ATA GTA GTT TGA TTG GTA GGG TGT Reverse ATC TAC AAC CTT ATC AAT TAC CCT TAA AAA

Gtl2/Dlk1 Forward GGA AGG AAA AGA TAA AAT GTA GAA A Reverse CAT AAA TAA ATA AAC CCA TAA TCC C

Table 3. Primers for enzyme protection assay

Human-specific primers Locus Sequence 5’ – 3’

H19 Digestion Forward ACT GAA GCC CTC GGA GTG T

Reverse AGA TCT TCA GGT CGG GCA TT

Normalisation Forward GAT AAT GCC CGA CCT GAA GA Reverse GGG GTC ATC TGG GAA TAG GA

156

Page 157: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Appendix

PEG3 Digestion Forward AAA ACC CCT ACA GGC AGG AC

Reverse GCG AAA ATG CCC CTT CCT

Normalisation Forward GAA AAC CCC TAC AGG CAG GA Reverse TTG TTT GCC GCA GTG GTG

SNRPN/SNURF Digestion Forward ACT GCG GCA AAC AAG CAC

Reverse CTC CTC AGA CAG ATG CGT CA

Normalisation Forward ACT GCG GCA AAC AAG CAC Reverse CAG GCT TCG CAC ACA TCC

Table 4. Primers for ChIP assay

Human-specific primers Locus Sequence 5’ – 3’

β-ACTIN Forward GAT CAG CAA GCA GGA GTA TGA CG Reverse AAG GGT GTA ACG CAA CTA AGT CAT AG

OCT4 Forward TTG CCA GCC ATT ATC ATT CA Reverse TAT AGA GCT GCT GCG GGA TT

TSH2B Forward Diagenode, pp-1041–500 Reverse

Table 5. Primers for genomic DNA amplification for Surveyor and RFLP Assays

Mouse-specific primers Gene Sequence 5’ – 3’

Tet1 Forward TTG TTC TCT CCT CTG ACT GC Reverse TGA TTG ATC AAA TAG GCC TGC

Tet2 Forward CAG ATG CTT AGG CCA ATC AAG Reverse AGA AGC AAC ACA CAT GAA GAT G

Jarid2 Forward GGC ACA GGG TAG AAG GAA AA Reverse ATT CCA GGG GTC CTT GAG TT

Prickle1 Forward TGG CCA TTG GCT TAT TTT TC Reverse AAC ACA ACC CAC AGG AAA GC

Fzd2 Forward ACA TCG CCT ACA ACC AGA CC Reverse GAG ATA GGA CGG CAC CTT GA

Wnt9a Forward GTG CTC TGG CTC CTC TGT TC Reverse TGT GCC CAG TAG AAG GGT TT

157

Page 158: Epigenetic Reprogramming and DNA Demethylation · However reprogramming of imprinting is only induced by EG, but not ES cells, and it requires sequential steps of 5methylcytosine

Publications Part of the work presented here has been published as follows:

Piccolo, F.M., Bagci, H., Brown, K.E., Landeira, D., Soza-Ried, J., Feytout, A., Mooijman, D., Hajkova, P., Leitch, H.G., Tada, T., Kriaucionis, S., Dawlaty, M.M., Jaenisch, R., Merkenschlager, M., and Fisher, A.G. (2013). Different roles for Tet1 and Tet2 proteins in reprogramming-mediated erasure of imprints induced by EGC fusion. Mol. Cell 49, 1023–1033.

Tsubouchi, T., Soza-Ried, J., Brown, K., Piccolo, F.M., Cantone, I., Landeira, D., Bagci, H., Hochegger, H., Merkenschlager, M., and Fisher, A.G. (2013). DNA synthesis is required for reprogramming mediated by stem cell fusion. Cell 152, 873–883.

158