Download Apoptotic and Epigenetic Induction of Embryo Failure Following

Document related concepts

Cell growth wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Cell culture wikipedia , lookup

Mitosis wikipedia , lookup

Apoptosis wikipedia , lookup

Cell cycle wikipedia , lookup

Programmed cell death wikipedia , lookup

Amitosis wikipedia , lookup

List of types of proteins wikipedia , lookup

JADE1 wikipedia , lookup

Cellular differentiation wikipedia , lookup

Somatic cell nuclear transfer wikipedia , lookup

Epigenetics in stem-cell differentiation wikipedia , lookup

Transcript
Utah State University
DigitalCommons@USU
All Graduate Theses and Dissertations
Graduate Studies
2013
Apoptotic and Epigenetic Induction of Embryo
Failure Following Somatic Cell Nuclear Transfer
Aaron Patrick Davis
Utah State University
Follow this and additional works at: http://digitalcommons.usu.edu/etd
Part of the Animal Sciences Commons, and the Biology Commons
Recommended Citation
Davis, Aaron Patrick, "Apoptotic and Epigenetic Induction of Embryo Failure Following Somatic Cell Nuclear Transfer" (2013). All
Graduate Theses and Dissertations. Paper 1941.
This Dissertation is brought to you for free and open access by the
Graduate Studies at DigitalCommons@USU. It has been accepted for
inclusion in All Graduate Theses and Dissertations by an authorized
administrator of DigitalCommons@USU. For more information, please
contact [email protected].
APOPTOTIC AND EPIGENETIC INDUCTION OF EMBRYO FAILURE
FOLLOWING SOMATIC CELL NUCLEAR TRANSFER
by
Aaron Patrick Davis
A dissertation submitted in partial fulfillment
of the requirements for the degree
of
DOCTOR OF PHILOSOPHY
in
Animal, Dairy and Veterinary Science
Approved:
_____________________________
Kenneth L. White, Ph.D.
Major Professor
_____________________________
Gregory J. Podgorski, Ph.D.
Committee Member
_____________________________
Abby D. Benninghoff, Ph.D.
Dissertation Director
_____________________________
S. Clay Isom, Ph.D.
Committee Member
_____________________________
Thomas D. Bunch, Ph.D.
Committee Member
_____________________________
Dirk K. Vanderwall, D.V.M., Ph.D.
ADVS Interim Department Head
_________________________________
Mark R. McLellan, Ph.D.
Vice President for Research and
Dean of the School of Graduate Studies
UTAH STATE UNIVERSITY
Logan, Utah
2013
ii
Copyright © Aaron Patrick Davis 2013
All Rights Reserved
iii
ABSTRACT
Apoptotic and Epigenetic Induction of Embryo Failure
Following Somatic Cell Nuclear Transfer
by
Aaron Patrick Davis, Doctor of Philosophy
Utah State University, 2013
Major Professor: Dr. Kenneth L. White
Department: Animal, Dairy, and Veterinary Sciences
Somatic cell nuclear transfer (SCNT) is a useful tool for selective breeding,
conservation, and production of transgenic animals. Despite the successful cloning of
several species, high rates of embryo failure following SCNT prevent the wide-scale use
of the technique. Embryos produced through cloning have a higher incidence of
developmental arrest, decreased developmental potential, frequent implantation failures,
and increased incidence of abortion. The objective of this dissertation research was to
characterize the factors that lead to SCNT failures by examining epigenetic and apoptotic
pathways that can negatively influence the development of cloned preimplantation
embryos.
Aberrant genome reprogramming is generally considered to be a key factor in the
failure of SCNT embryo development. Therefore, we used bisulfite pyrosequencing
technology to compare DNA methylation patterns of several genes critical for embryonic
iv
development (POU5F1, NANOG, SOX2, and KLF4) in SCNT and in vitro fertilized
(IVF) blastocyst stage embryos. The methylation profiles obtained from these
experiments indicate that methylation patterns of the POU5F1 gene were
undermethylated compared to IVF embryos, suggesting reprogramming did occur, but
that the reduced methylation was inappropriate for the blastocyst stage. Furthermore,
aberrant methylation profiles were detected for SOX2 and NANOG, suggesting that
problems of genome reprogramming following SCNT can be gene-specific or localized.
Because high rates of apoptosis are associated with failure of preimplantation
embryos, we compared the activation of the P53-mediated apoptosis pathway in
individual IVF and SCNT preimplantation embryos at multiple developmental stages.
This pathway is activated in response to cell stress and genomic instability, and in
response to the expression of genes associated with somatic cell reprogramming.
Evidence from gene expression and immunohistochemistry analyses suggests that the
P53 pathway is frequently active in SCNT embryos. Also, we detected expression of
several factors known to induce apoptosis more frequently and at higher levels in SCNT
embryos.
Collectively, the work presented here illuminates some of the molecular
consequences of incomplete or inappropriate genome reprogramming in cloned embryos.
The identification of these factors may lead to interventions that target the apoptosis
pathway during preimplantation development and increase SCNT success rates.
(183 pages)
v
PUBLIC ABSTRACT
Apoptotic and Epigenetic Induction of Embryo Failure
Following Somatic Cell Nuclear Transfer
by
Aaron Patrick Davis, Doctor of Philosophy
Utah State University, 2013
Major Professor: Dr. Kenneth L. White
Department: Animal, Dairy, and Veterinary Sciences
The cloning of domestic species has tremendous potential, as the technology can
be used in selective breeding, conservation, and the production of transgenic animals.
The technique of cloning involves the transplant of DNA from a cell to a recipient
gamete. Following transfer to a surrogate, the cloned embryo may successfully complete
development to a live offspring. Despite intensive research, the success rate of cloning
remains prohibitively low, and the potential benefits of cloning have not yet been
realized. Embryos produced from cloning suffer from high rates of embryo degradation,
implantation failure, and abortion. The goal of this dissertation research project was to
determine the mechanisms that cause the high rates of embryo failure in clones.
In order for embryo development to occur properly following cloning, the donor
cell genome must undergo epigenetic reprogramming, a process that resets the donor
nuclei to an embryonic state. Epigenetic markings alter the expression of genes within a
vi
cell, and are one factor that prevents cells from reverting back to a primordial state. As
part of this research, we measured methylation of DNA, a type of epigenetic mark, to
determine the level of reprogramming that occurs in cloned embryos. Our results
demonstrate that cloned embryos have abnormalities in reprogramming following cloning
for some important genes. This observation provides further support for the notion that
the inefficiency of the cloning process may be a result of the inability of donor cells to
readily reactivate genes that have been epigenetically silenced.
A second area of research into the causes of cloning failure was to examine the
role of programmed cell death within cloned embryos. Every cell contains an innate
defense system that triggers programmed cell death when the cell is under severe stress or
has undergone irreparable DNA damage. We observed increased incidence of
programmed cell death in cloned embryos that contributes to the high rates of
developmental failure. These findings may potentially allow therapeutics targeted at
preventing cell death, and thereby increase the success of cloning.
vii
ACKNOWLEDGMENTS
I give thanks to my advisor, Dr. Ken White, as well as my committee members,
Drs. Abby Benninghoff, Thomas Bunch, Gregory Podgorski, and Clay Isom. I also thank
Drs. Qinggang Meng, Chris Davies, Lee Rickords, Aaron Thomas, and Heloisa
Rutigliano for their frequent assistance. I recognize Dr. Ben Sessions, Dr. Ki Aston, Dr.
Ammon Bayles, Kira Morgado, Bryce Osborne, and Eric Innes for their help throughout
my time at Utah State.
This work was supported by the National Institute of Food and Agriculture, U.S.
Department of Agriculture, under agreement No. 08-34526-19199 and 09-34526-19808.
In addition, I would like to acknowledge the financial support of the Utah Agriculture
Experiment Station.
I would like to give special thanks to my parents, Sam and Robin, for their
continued support and my wife, Jessica, for being there with me all along the way.
Aaron Patrick Davis
viii
CONTENTS
Page
ABSTRACT .................................................................................................................. iii
PUBLIC ABSTRACT ..................................................................................................... v
ACKNOWLEDGMENTS ............................................................................................. vii
LIST OF TABLES........................................................................................................... x
LIST OF FIGURES ........................................................................................................ xi
LIST OF ABBREVIATIONS AND SYMBOLS ........................................................ xiii
CHAPTER
1.
REVIEW OF LITERATURE ................................................................. 1
History of Somatic Cell Nuclear Transfer ......................................... 1
Limitations in Generating Cloned Offspring ..................................... 7
The Epigenetic Landscape ................................................................. 9
DNA Methylation and Somatic Cell Nuclear Transfer ................... 15
Epigenetic Reprogramming ............................................................. 19
The P53-Mediated Apoptosis Pathway ........................................... 24
Summary .......................................................................................... 30
Research Goals and Potential Applications ..................................... 31
References ....................................................................................... 32
2.
DNA METHYLATION REPROGRAMMING OF THE
POU5F1, NANOG, SOX2, AND KLF4 GENES
FOLLOWING SOMATIC CELL NUCLEAR TRANSFER ................ 62
Abstract ............................................................................................
Introduction .....................................................................................
Materials and Methods ....................................................................
Results .............................................................................................
Discussion ........................................................................................
References .......................................................................................
62
62
66
69
74
80
ix
3.
DNA METHYLATION OF THE LIN28 PSEUDOGENE
FAMILY ................................................................................................ 93
Abstract ............................................................................................ 93
Introduction ..................................................................................... 93
Materials and Methods .................................................................... 96
Results ........................................................................................... 100
Discussion ...................................................................................... 105
References ..................................................................................... 108
4.
ACTIVATION OF THE P53 PATHWAY FOLLOWING
BOVINE SOMATIC CELL NUCLEAR TRANSFER ....................... 118
Abstract ..........................................................................................
Introduction ...................................................................................
Materials and Methods ..................................................................
Results ...........................................................................................
Discussion ......................................................................................
References .....................................................................................
5.
118
119
123
128
132
140
CONCLUSION ................................................................................... 161
CURRICULUM VITAE.............................................................................................. 166
x
LIST OF TABLES
Table
Page
1-1
Somatic Cell Nuclear Transfer Success Rates for Live Offspring
in Various Species ............................................................................................... 8
2-1
Primers Used for Bisulfite Sequencing ............................................................. 87
3-1
Primers Used for Bisulfite Sequencing and End-Point RT-PCR .................... 112
4-1
Primers Used for qRT-PCR ............................................................................. 149
4-2
Statistical Analysis of Gene Expression. ......................................................... 150
xi
LIST OF FIGURES
Figure
Page
2-1
Map of CpG islands and regions examined for methylation ............................. 88
2-2
Heat map depicting methylation profile of POU5F1 in donor
cells and blastocyst stage SCNT and IVF embryos ........................................... 89
2-3
Heat map depicting methylation profile of NANOG in donor
cells and blastocyst stage SCNT and IVF embryos ........................................... 90
2-4
Heat map depicting methylation profile of SOX2 in donor
cells and blastocyst stage SCNT and IVF embryos ........................................... 91
2-5
Heat map depicting methylation profile of POU5F1 in donor
cells and blastocyst stage SCNT and IVF embryos .......................................... 92
3-1
The ten processed pseudogenes of LIN28 aligned to the LIN28
processed transcript ......................................................................................... 113
3-2
LIN28 pseudogene locations on each chromosome ......................................... 114
3-3
Heat map of LIN28 and LIN28 pseudogene methylation profile..................... 116
3-4
End-point RT-PCR ......................................................................................... 117
4-1
Heat map of all genes analyzed for expression analysis ................................. 151
4-2
Log2 fold change expression values of BAX, BCL2 and P21WAF1/Cip1 ............. 152
4-3
Average log2 fold expression change values of BAX and BCL2 ..................... 153
4-4
Log2 fold change expression values of IGF2R, CX43, PEG3 and ETS2 ........ 155
4-5
Log2 fold change expression values of C-MYC, LIN28, SOX2 and KLF4 ...... 157
4-6
Representative immunofluorescence images for P53 and phosphorylated
CDK2 in IVF and SCNT embryos .................................................................. 158
4-7
Representative immunofluorescence images for P21 and BAX in IVF
and SCNT embryos ......................................................................................... 159
xii
4-8
Representative images and evaluation of apoptosis in SCNT and IVF
embryos ........................................................................................................... 160
xiii
LIST OF ABBREVIATIONS AND SYMBOLS
Abbreviations
Δ-ΔCt
ΔCt
ACADSB
ACTB
BAX
BCL2
BSA
CDK
CpG
Ct
CX43
C-MYC
DHDDS
DNMT
ETS2
ES Cells
FBS
FSH
GAPDH
ICM
IGF2R
iPCS
IKZF5
IVF
KLF4
LH
LIN28
MAN2A1
NANOG
P21WAF1/Cip1
P53
PBS
PCR
Pc-G/trx-G
PEG3
POU5F1
qRT-PCR
ROS
SCNT
Delta-Delta Ct
Delta Ct
acyl-CoA Dehydrogenase
ß-actin
BCL2 associated X protein
B-cell CLL/lymphoma 2
Bovine serum albumin
Cyclin-dependent kinase
Cytosine-phosphate-guanine
Cycle threshold
Gap junction protein, alpha 1
V-MYC myelocytomatosis viral
Dehydrodolichyl diphosphate synthase
DNA methyltransferase
V-Ets erythroblastosis virus E25
Embryonic stem cells
Fetal bovine serum
Follicle stimulating hormone
Glyceraldehyde-3-phosphate dehydrogenase
Inner cell mass
Insulin-like growth factor 2 receptor
Induced pluripotent stem cell
IKAROS family zinc finger 5 (Pegasus)
In vitro fertilization
Kruppel-like factor 4
Luteinizing hormone
Lin-28 homolog A
Mannosidase, alpha, class 2A, member 1
Nanog homeobox
Cyclin-dependent kinase inhibitor 1A
Tumor protein P53
Phosphate buffered saline
Polymerase chain reaction
Polycomb/trithorax group complex
Paternally expressed 3
POU class 5 homeobox 1
Quantitative reverse transcription polymerase chain reaction
Reactive oxygen species
Somatic cell nuclear transfer
xiv
SOX2
TUNEL
Sex determining region-box 2
Terminal deoxynucleotidyl transferase dUTP nick end labeling
Symbols
C
cm
DC
g
g
hr
mg
min
ml
mm
µsec
mM
sec
U
µg
µl
µM
Celsius
Centimeter
Direct current
Grams
Gravity
Hour
Milligrams
Minutes
Milliliter
Millimeter
Microsecond
Millimolar
Seconds
Units
Micrograms
Microliter
Micromolar
CHAPTER 1
REVIEW OF LITERATURE
History of Somatic Cell Nuclear Transfer
The technique of somatic cell nuclear transfer (SCNT) was originally conceived
in the early twentieth century by developmental biologist Hans Spemann, who sought to
understand the process whereby somatic cells specialize during development. During this
time he proposed an experiment in which a somatic cell would be transferred to an
enucleated oocyte in order to determine the developmental potential of somatic cells.
Although he lacked the appropriate instrumentation to carry out such an experiment,
Spemann is credited with introducing the experimental design now known as SCNT.
Prior to Spemann’s proposed nuclear transfer experiment, Hans Driesch had
demonstrated that each cell of an embryo contained the entire complement of genes
necessary to carry out development. He accomplished this by separating individual
blastomeres from developing sea urchin embryos and observed that each blastomere had
the capacity to develop into a separate embryo (Spemann, 1938). This evidence was
contrary to a competing hypothesis specifying that genes were segregated during
development, and that the unique complement of genes found in an individual cell
resulted in the unique phenotype of the somatic population. Using frog embryos,
Spemann carried out a similar set of experiments in which early-stage embryos were
artificially split resulting in two identical embryos, each capable of carrying out
development (Spemann, 1938). Collectively, the work of Dreisch and Spemann affirmed
the hypothesis of complete genome replication, although their experiments were
2
restricted to early-stage blastomeres. The hypothesis that somatic cells lost genes during
development could not be directly tested for several years until the instrumentation
required for the nuclear transfer experiment originally devised by Spemann was
developed.
The first experiment to use nuclear transfer as a technique was completed in 1952.
The North American leopard frog (Rana pipens) was utilized for nuclear transfer using
late blastula cells as a donor cell source (Briggs and King, 1952). The successful cloning
of a frog affirmed the concept of nuclear equivalence, showing that genetic material was
not lost during development and demonstrating that a differentiated nucleus could return
to an undifferentiated state. Briggs and King eventually expanded their work to include
different donor cell types to determine whether or not all cells have the same potential for
dedifferentiation (King and Briggs, 1955). Further experiments showed that cloning
efficiency declined in proportion to the degree of differentiation. Although the success
rate was low, late gastrula cells could be used as a donor source to successfully generate
tadpoles (Briggs and King, 1960). Building on these studies, Gurdon demonstrated that
epithelial cells derived from early Xenopus tadpoles could be used as a source of donor
cells to generate clones (Gurdon, 1962). In a later study, Gurdon successfully used
keratinocytes from adult Xenopus as donor cells (Gurdon et al., 1975). While these
experiments produced tadpoles, neither group was successful in generating adult frogs
from cloning experiments.
In 1975, Bromhall was the first to apply the cloning procedure to mammals.
Using rabbits, he observed that cell divisions could occur following nuclear transfer,
3
although development never continued beyond early cleavage stages (Bromhall, 1975).
Using enucleated mouse zygotes as recipients for zygotic nuclei, McGrath and Solter
generated mouse offspring that developed completely to term (McGrath and Solter,
1983). However, subsequent cloning attempts that used mouse blastomeres as a donor
source failed to generate embryos that could initiate development (McGrath and Solter,
1984). Eventually, the cloning of domestic species using blastomeres as donor cells
resulted in the birth of clones (Willadsen, 1986; Prather et al., 1987; Prather et al., 1989).
Each successful cloning attempt used enucleated oocytes as donor cell recipients, rather
than zygotes which had failed to produce clones in mice (McGrath and Solter, 1984).
Willadsen successfully generated cloned cattle using blastomeres as donor cells
(Willadsen et al., 1991), and Sims and First successfully cloned four calves from inner
cell mass (ICM) cells that were briefly cultured in vitro prior to use as a donor cell (Sims
and First, 1994).
The work of Sims and First established the use of cultured cells in SCNT
procedures to generate mammalian offspring, but the benchmark of generating a cloned
offspring derived from a somatic cell had yet to be achieved. Following several failed
attempts, it was unknown if somatic nuclei were capable of being reprogrammed to an
undifferentiated state.
In 1996, a research team produced a sheep cloned from the mammary epithelial
cell of an adult sheep (Wilmut et al., 1997). Dolly, the offspring of these experiments,
represented a major scientific breakthrough, as the cloned ewe was the first mammalian
offspring produced from a somatic cell. Her birth indicated that somatic cells were, in
4
fact, capable of dedifferentiation to a totipotent state and could sustain embryo
development following transfer to an oocyte. Since the birth of Dolly, several somatic
cell lines have been used for SCNT in mice, including terminally differentiated cells
(Wakayama et al., 1998; Hochedlinger and Jaenisch, 2002).
To date, many species have been cloned from embryonic and/or somatic cell
lines, including: sheep (Wilmut et al., 1997), cattle (Cibelli et al., 1998a; Kato et al.,
1998; Wells et al., 1999; Kato et al., 2000), goat (Baguisi et al., 1999; Keefer et al.,
2001), pig (Betthauser et al., 2000; Onishi et al., 2000; Polejaeva et al., 2000), mouse
(Wakayama et al., 1998; Wakayama and Yanagimachi, 1999; Hosaka et al., 2000),
monkey (Mitalipov et al., 2002), gaur (Lanza et al., 2000b), mouflon (Loi et al., 2001),
cat (Shin et al., 2002; Gomez et al., 2004), rabbit (Chesne et al., 2002; Challah-Jacques et
al., 2003), zebrafish (Lee et al., 2002), rat (Zhou et al., 2003), mule (Woods et al., 2003),
horse (Galli et al., 2003; Hinrichs et al., 2006), water buffalo (Suteevun et al., 2006), dog
(Lee et al., 2005), banteng (Sansinena et al., 2005), ferret (Li et al., 2006), deer (Berg et
al., 2007), pyrenean ibex (Folch et al., 2009), and camel (Wani et al., 2010).
Additionally, human embryos have been produced by SCNT, but have not been allowed
to develop past the blastocyst stage (Hwang et al., 2004; Tachibana et al., 2013).
All SCNT experiments have used variations of the first procedure established by
Briggs and King. Oocyte enucleation is achieved through the removal of the pronucleus
along with the metaphase spindle. Briggs and King performed transfer of the donor
nucleus by aspirating the blastomere into a narrow micropipette that ensured the cell
membrane was sheared, while leaving the nucleus intact, and then microinjecting the
5
nucleus into the oocyte (Briggs and King, 1952). Amphibian eggs do not require any
form of oocyte activation beyond microinjection.
The SCNT technique is similar for mammals and amphibians, although cell fusion
via electrofusion is preferred to microinjection for introduction of the donor cell to the
mammalian recipient oocyte. However, microinjection of the donor cell nucleus remains
the standard practice in the mouse, as murine oocytes tend to self-activate upon
stimulation with an electric pulse (Wakayama et al., 1998). For the electrofusion
procedure, the zona pellucida and oocyte membrane are pierced, and the first polar body,
pronuclei and metaphase plate are removed. The donor cell, which is similar in size to
the polar body, is then inserted between the oocyte membrane and zona pellucida.
Electrofusion is used to fuse the two membranes. The cloned embryo is briefly cultured
in vitro, then transferred to a surrogate for fetal development. In procedures using bovine
oocytes, activation is initiated using the calcium ionophore compound ionomycin along
with cycloheximide to briefly suspend protein synthesis.
Cloning has been a valuable tool for advancing science in several areas of biology
including nuclear differentiation, nuclear reprogramming, genomic imprinting, and
cellular aging. Many researchers have taken advantage of genetic engineering techniques
to manipulate the donor cell genome and subsequently perform SCNT to produce
transgenic offspring. One such example was the generation of a transgenic sheep that
produced human clotting factor IX in its milk as a pharmaceutical for treatment of
hemophilia (Schnieke et al., 1997). Cloning has been used to generate transgenic animals
by random gene addition (Schnieke et al., 1997), gene knockout (McCreath et al., 2000),
6
and gene mutation (Rogers et al., 2008). Aside from the production of human proteins in
transgenic cloned animals, SCNT also provides researchers the opportunity to modify an
animal’s genome to reproduce desired genetic traits, such as increased casein content in
the milk of transgenic cattle (Brophy et al., 2003) or the removal of antigens for potential
xenotransplantation (Lai et al., 2002; Phelps et al., 2003). Transgenic clones can also be
used for research into human disease (Denning et al., 2001).
Cloning can also be applied for animal reproduction. SCNT can be used to
reproduce genetically superior animals as well as for the preservation of endangered
species. Some endangered species have been cloned for this purpose, such as gaur
(Lanza et al., 2000b), mouflon (Loi et al., 2001), banteng (Sansinena et al., 2005), and
water buffalo (Suteevun et al., 2006). This procedure has also been used to clone an
extinct species. The pyrenean ibex, a species that went extinct in 2000, was cloned in
2009 using cells frozen from one of the last remaining survivors (Folch et al., 2009).
Even though the cloned animal survived throughout gestation, it died shortly after birth.
Cloning has also been applied in therapeutic applications including the generation
of embryonic stem (ES) cells (Rideout et al., 2002; Barberi et al., 2003). ES cells have
been derived from clones in mouse (Munsie et al., 2000; Wakayama, 2003) and human
(Hwang et al., 2004; Tachibana et al., 2013), and ES-like cells have been generated in
cattle (Cibelli et al., 1998b) and rabbit (Fang et al., 2006). Additionally, cloning could be
used for pharmacology purposes, such as generating identical individuals for drug testing,
although this potential has yet to be realized.
7
SCNT has also contributed to advances in basic biology. Some examples include
cloning for the study of allelic regulation (Gerdes and Wabl, 2004), the study of
secondary rearrangements of the Immunoglobulin genes (Koralov et al., 2005), and the
determination that olfactory receptor neurons do not generate variation through gene
rearrangements (Eggan et al., 2004; Li et al., 2004).
Limitations in Generating Cloned Offspring
The largest barrier to cloning is the high loss of cloned embryos during gestation.
Most pregnancy failures occur early after implantation during the period critical to
placental establishment (Heyman et al., 2002). Across most species, the success rate of
reconstructed embryos that result in the birth of healthy offspring is less than 3%
(Thibault, 2003). Reported success rates of various donor cells reported in Table 1-1.
Cloned embryos that successfully implant regularly develop placental abnormalities
(Yang et al., 2007). In mice, several problems associated with placental formation have
been observed in SCNT pups (Wakayama and Yanagimachi, 1999; Tanaka et al., 2001;
Jouneau et al., 2006; Wakisaka-Saito et al., 2006; Wakisaka et al., 2008) and similar
observations have been made in cattle (Chavatte-Palmer et al., 2012). An oversized and
dysfunctional placenta is frequently observed in both aborted clones as well as those that
have survived to term. This abnormal placenta likely contributes to the high rates of fetal
loss (Wakayama et al., 1998; Wakayama and Yanagimachi, 1999; De Sousa et al., 2001).
SCNT animals often die shortly after birth, with the most common cause of
neonatal death thought to be respiratory distress and problems associated with circulation
(Wilmut et al., 1997; Young et al., 1998; Hill et al., 1999). Placental deficiencies,
Adult Mammary
Fetal Fibroblasts
Fibroblasts
Cumulus Cells
Sertoli Cells
Natural Killer T-Cells
B and T Cells
Neurons
Fibroblasts
Fetal Fibroblasts
Granulosa Cells
Fetal Fibroblasts
Fibroblasts
Cumulus Cells
Fetal Fibroblasts
Fibroblasts
Fibroblasts
Cumulus Cells
Sheep
Cat
Rabbit
Rat
Horse
Dog
Ferret
Goat
Pigs
Cattle
Mouse
Cell Type
Species
3.4
5 to 20
9 to 20
1 to 3
6
1 to 2
<1
<1
1
3 to 10
1.2
4
1.1
1.6
1.7
5.9
0.2
1.8
Success Rate (%)
(Wilmut et al., 1997)
(Schnieke et al., 1997)
(Wells et al., 1999; Kato et al., 2000; Brophy et al., 2003)
(Wakayama et al., 2000; Wakayama et al., 2005)
(Ogura et al., 2000; Wakayama et al., 2005)
(Inoue et al., 2005),
(Hochedlinger and Jaenisch, 2002; Eggan et al., 2004; Li et al., 2004)
(Eggan et al., 2004; Li et al., 2004)
(Wakayama and Yanagimachi, 1999; Wakayama et al., 2005)
(Baguisi et al., 1999; Keefer et al., 2001)
(Polejaeva et al., 2000)
(Lai et al., 2002)
(Shin et al., 2002)
(Chesne et al., 2002)
(Zhou et al., 2003)
(Galli et al., 2003)
(Lee et al., 2005)
(Li et al., 2006)
Reference
Table 1-1. Somatic Cell Nuclear Transfer Success Rates for Live Offspring in Various Species
8
9
neonatal edema, urogenital tract defects, and problems with the respiratory and
cardiovascular systems are typical abnormalities in cloned offspring that survive to term
(Hill et al., 1999; Renard et al., 1999; Hill et al., 2000a; De Sousa et al., 2001; ChavattePalmer et al., 2002). Reduced immunity, kidney malformation (Lanza et al., 2000a;
McCreath et al., 2000), predisposition to infection, immature lungs, and general weakness
(Zakhartchenko et al., 2001) are also commonly observed in cloned neonates. In
addition, clones often suffer from a collection of growth malformations known as large
offspring syndrome, which is characterized by large size and a general failure to thrive
(Young et al., 1998).
SCNT offspring have a much higher incidence of abnormal phenotypes than
offspring produced in vivo or by in vitro fertilization (IVF) (Han et al., 2003; Thibault,
2003). The abnormal phenotypes are generally held to be the result of incomplete
epigenetic reprogramming of the somatic cell genome after introduction into the oocyte
(Wilmut et al., 1998; Colman, 1999; Kikyo and Wolffe, 2000; Solter, 2000; Dean et al.,
2001; Humpherys et al., 2001; Reik et al., 2001; Niemann et al., 2008).
The Epigenetic Landscape
Epigenetics is defined as heritable changes in gene expression that are not due to
changes to the DNA sequence. Epigenetic factors include DNA methylation, histone
modifications, and polycomb-trithorax complex, all of which modify the physical
characteristics of chromatin, and lock transcriptional regions in either an active or silent
expression state. Epigenetic signals are stably inherited after cell division and enable
daughter cells to recapitulate the transcriptional program of the parent cell. The
10
development of multicellular organisms requires a specific transcriptional program of
gene expression for each individual cell type. In animals, over 50% of genes are inactive
in any given cell (Bergman and Cedar, 2013) The unique epigenetic program of each cell
maintains both the long-term silencing of inappropriate genes and the active transcription
of genes that are necessary to carry out the functions of the cell. Epigenetic patterns are
established during embryogenesis and early development and are maintained within adult
somatic cells.
Epigenetic memory is maintained through methylation of DNA. Once established,
the pattern of methylation can be copied in each cell generation to maintain a long-term
epigenetic signal. Vertebrates have the highest levels of DNA methylation in the animal
kingdom and contain methylated regions dispersed throughout most of their genomes
(Bird, 2002). The widespread distribution of DNA methylation patterns suggests
methylation is involved in multiple aspects of epigenetic control (Tweedie et al., 1999).
Researchers have postulated that DNA methylation originated as a genomic defense
system that allowed a cell to inactivate viral and transposable elements inserted into DNA
(Liu et al., 1994; Woodcock et al., 1997; Yoder et al., 1997; Walsh et al., 1998). DNA
methylation has evolved to play major roles in X-chromosome inactivation (Mohandas et
al., 1981; Csankovszki et al., 2001), genomic imprinting (Bartolomei et al., 1993; Li et
al., 1993; Stoger et al., 1993), the maintenance of chromosome stability (Chen et al.,
1998; Moarefi and Chedin, 2011), and possibly in RNA splicing variation (Schwartz et
al., 2009; Laurent et al., 2010; Shukla et al., 2011). Transcription of genes on the
inactive X-chromosome is induced within cells treated with agents that cause global
11
demethylation (Mohandas et al., 1981; Graves, 1982; Venolia et al., 1982). Induction of
demethylation also causes the aberrant transcription of retroviral elements normally
dormant within the genome (Stewart et al., 1982; Jaenisch et al., 1985).
DNA methylation is also involved in the long-term silencing of tissue-specific
genes, as methylation of the transcriptional start site is associated with transcriptional
inactivation of genes (De Smet et al., 1996; De Smet et al., 1999; Jaenisch and Bird,
2003; Shiota, 2004; Han et al., 2011; You et al., 2011). Embryos and cultured cells with
reduced methylation aberrantly express a number of genes that are normally repressed
(Stancheva and Meehan, 2000; Jackson-Grusby et al., 2001). Some tissue-specific genes
maintain methylation throughout development and are correspondingly transcriptionally
inactive until a demethylation event relieves inhibition to allow gene expression (Han et
al., 2011). Although DNA methylation is an important mechanism for long-term
silencing of tissue-specific genes, less than 10% of genes are regulated by DNA
methylation (Illingworth and Bird, 2009). Multiple tissue-specific genes undergo longterm transcriptional inactivation without changes to DNA methylation patterns (McKeon
et al., 1982; Bird, 1987). However, some genes require the presence of methylation for
long-term gene inactivation (Venolia and Gartler, 1983; Kass et al., 1997; Hashimshony
et al., 2003).
In mammals, DNA methylation occurs almost exclusively on cytosine nucleotides
in the context of CpG dinucleotides. Approximately 75% of CpG dinucleotides are
methylated (Ehrlich et al., 1982), most of which are concentrated in well-defined regions.
However, any CpG dinucleotide can be methylated at any time under special
12
circumstances or in abnormal cell types, such as in cancer cells (Bird, 2002).
Approximately 98% of CpG dinucleotides are dispersed throughout the genome at low
density and are highly methylated. The remaining 2% of CpG dinucleotides are in
densely packed regions consisting of approximately one CpG dinucleotide per ten base
pairs. These dense clusters of CpG dinucleotides are referred to as CpG islands, and are
generally void of methylation (Brandeis et al., 1994; Straussman et al., 2009).
CpG islands are a unique feature of vertebrates and are defined as regions of DNA
with a high density of CpG dinucleotides (Illingworth and Bird, 2009). These islands are
associated with approximately 60% of genes and are commonly found at the 5’-end of
genes (Antequera and Bird, 1993). CpG islands are often associated with transcription
factor binding regions, and methylation of these regions is linked to the transcriptional
regulation of the associated gene (Bird, 1986; Bird, 1987; Antequera and Bird, 1993;
Cuadrado et al., 2001; Bird, 2002; Illingworth et al., 2008).
Repression by DNA methylation may be weak or strong depending on the
methylation density (Boyes and Bird, 1992; Hsieh, 1994). Methylation of DNA leads to
inhibition of gene expression via one of two mechanisms. The first involves the direct
blocking of transcription factor binding to DNA by the methyl groups attached to CpG
dinucleotides (Prendergast and Ziff, 1991; Bell and Felsenfeld, 2000; Hark et al., 2000;
Szabo et al., 2000; Holmgren et al., 2001). The second, and likely more prevalent mode
of gene silencing, involves proteins that bind selectively to methylated DNA and either
block transcription factor access to DNA, or recruit chromatin condensation complexes
that silence the chromatin region (Meehan et al., 1989; Jones et al., 1998; Nan et al.,
13
1998; Ng et al., 1999; Wade et al., 1999; Feng and Zhang, 2001). Several proteins have
been discovered that selectively bind methylated DNA (Nan et al., 1993; Cross et al.,
1997; Nan et al., 1997; Hendrich and Bird, 1998). One such protein, KAISO, has been
found to initiate methylation-dependent repression upon binding to methylated DNA
(Prokhortchouk et al., 2001). Furthermore, histone deacetylases can be recruited to
methylated DNA in order to further reinforce repression (Wade and Wolffe, 2001; Lin et
al., 2007).
The family of DNA methyltransferases (DNMT) (Bestor, 1992; Miniou et al., 1994;
Okano et al., 1998a,b; Pradhan et al., 1999; Hansen et al., 2000; Kondo et al., 2000; Rhee
et al., 2000) is responsible for establishing and maintaining patterns of DNA methylation.
DNMTs are essential for normal development (Li et al., 1992; Okano et al., 1999). Both
mice and humans with deficiencies in DNMT lack methylation on the inactive Xchromosome and pericentromeric repetitive DNA sequences and express genes that are
normally inactive (Bestor, 2000; Sado et al., 2000; Stancheva and Meehan, 2000).
During DNA replication, DNMTs associate with the replication machinery to duplicate
methylation patterns in newly synthesized DNA (Leonhardt et al., 1992). In addition to
DNMTs, methylation requires a number of accessory proteins that likely give DNMTs
access to heterochromatic regions (Gibbons et al., 2000; Dennis et al., 2001). The de
novo establishment of DNA methylation can be targeted to CpG islands by a RNAdirected mechanism (Wassenegger et al., 1994; Bender, 2001; Matzke et al., 2001).
DNMTs are also attracted to specific genomic regions (Turker, 1999) by modified
histone H3 proteins in chromatin (Tamaru and Selker, 2001), and by histone modifying
14
complexes that recruit DNMTs to condensed regions (Feldman et al., 2006; EpsztejnLitman et al., 2008).
DNA methylation does not directly stop gene transcription in actively expressed
genes, but rather acts as a secondary mechanism that locks in a transcriptionally silent
state after the region has been rendered transcriptionally inactive by some other
mechanism (Gautsch and Wilson, 1983; Niwa et al., 1983; Lock et al., 1987; Keohane et
al., 1996; Pannell et al., 2000; Wutz and Jaenisch, 2000). Transgenes that contain a CpG
island can maintain active transcription following genome insertion. However, in the
event that promoter function is impaired, DNA methylation will accumulate upon the
transgene and prevent future expression (Macleod et al., 1994). Aberrant de novo
methylation, such that occurs in cancer, primarily accumulates on regions after they have
become transcriptionally inactive. This reinforces the view of DNA methylation as a
‘lock’ rather than a trigger (Ohm et al., 2007; Widschwendter et al., 2007; Gal-Yam et
al., 2008). However, the role played by DNA methylation in maintaining silencing is
essential for eliminating plasticity and preventing a cell from reverting back to a less
differentiated state (Wareham et al., 1987; Epsztejn-Litman et al., 2008).
The distribution and extent of DNA methylation varies among species.
Caenorhabditis elegans lacks detectible cytosine methylation or DNMT expression (Bird,
2002). Drosophila melengaster contains a DNA methyltransferase (Hung et al., 1999;
Tweedie et al., 1999) and maintains low levels of DNA methylation, primarily in CpT
dinucleotides (Gowher et al., 2000; Lyko et al., 2000). Both of these species maintain
epigenetic memory predominantly through the Polycomb/Trithorax Group Complex (Pc-
15
G/trx-G) (Paro et al., 1998; Pirrotta, 1999; Francis and Kingston, 2001). Pc-G/trx-G is a
multiprotein complex that establishes a transcriptional state at targeted genomic regions
and stably maintains either an active or repressive state of chromatin throughout
development. The Pc-G/trx-G achieves a similar outcome as DNA methylation in
mammals (Yoder et al., 1997; Birchler et al., 2000). Although mammals also utilize the
Pc-G/trx-G system, they appear to have evolved a more complex use of DNA
methylation that is utilized in conjunction with the Pc-G/trx-G system (Sado et al., 2000;
Wang et al., 2001).
DNA Methylation and Somatic Cell Nuclear Transfer
Modest improvements to cloning efficiencies have been achieved by manipulating
factors such as SCNT activation procedure, donor cell passage number, culture
conditions, and cell cycle stage. Changes in these factors have been shown to alter gene
expression in SCNT embryos (Wrenzycki et al., 2001). However, developmental success
still remains diminished compared to in vivo and IVF-produced embryos, likely because
the problems associated with clones are epigenetic in nature.
DNA methylation establishes chromatin structure required for successful
development (Hashimshony et al., 2003). Failure to maintain a suitable methylation
program leads to multiple developmental abnormalities, as embryos deficient in DNA
methylation can either fail to develop or abort during early gestation (Bestor, 2000; Bird,
2002). The establishment of DNA methylation patterns is a complex, coordinated
process that occurs progressively throughout development. Following fertilization, the
paternal pronucleus undergoes active demethylation that occurs by enzymatic removal of
16
methyl groups from CpG sites. This process occurs exclusively on the paternal genome
and is completed prior to the DNA replication at the first cleavage division (Mayer et al.,
2000; Oswald et al., 2000). The maternal pronucleus does not undergo active
demethylation. Instead, DNA methylation levels are gradually diluted through cleavage
divisions, due to inhibition of methylation on the nascent DNA strand (Reik et al., 2001;
Santos et al., 2002). The result is global demethylation during preimplantation
development during which methylation levels drop approximately 30% following
fertilization (Monk et al., 1987; Kafri et al., 1992). Demethylation as part of embryonic
development likely occurs to remove any methylation marks imposed during
gametogenesis (Smith et al., 2012) as well as to remove methylation patterns from
developmentally significant genes that may be methylated in the sperm (Farthing et al.,
2008). By the time the embryo develops to the morula stage, methylation is primarily
found in imprinted genes and repetitive elements (Sanford et al., 1987; Walsh et al.,
1998; Reik and Walter, 2001). However, upon cellular differentiation at the blastocyst
stage, new methylation patterns begin to be established, and clear differences in
methylation signature can be seen between the trophectoderm and inner cell mass (ICM)
(Turker, 1999; Santos et al., 2002).
A major challenge in reprogramming a somatic cell during SCNT is that
reprogramming of donor cell methylation patterns must recapitulate the removal and
reestablishment of methylation patterns that occur in normal embryos. Clones must
epigenetically reset genes required for embryo development prior to the maternal-toembryo transition, after which developmental progress is dependent on the transcriptional
17
program of the donor cell genome. Several studies have observed abnormal levels of
DNA methylation in cloned embryos (Dean et al., 1998; Dean et al., 2001; Kang et al.,
2001; Ohgane et al., 2001; Kang et al., 2002). The mechanism responsible for the
abnormalities in methylation patterns following SCNT is not clear, but it is likely that the
donor cell genome responds differently to the oocyte cytoplasm than the pronucleus.
Examples of methylation defects in clones include errors in X-chromosome
inactivation (Keohane et al., 1996; Eggan et al., 2000; Xue et al., 2002; Senda et al.,
2004; Nolen et al., 2005), irregularities in the methylation patterns of imprinted genes and
their expression patterns (Humpherys et al., 2001; Inoue et al., 2002; Mann et al., 2003),
dysregulation of DNMTs (Chung et al., 2003), and overall abnormalities in patterns of
DNA methylation (Bourc'his et al., 2001; Dean et al., 2001; Humpherys et al., 2001;
Kang et al., 2001).
Analysis of the organs of cloned neonate mice reveals that a large number of
genes are aberrantly methylated (Ohgane et al., 2001; Humpherys et al., 2002). In
addition to alterations in DNA methylation, clones exhibit unusual patterns of histone
acetylation and methylation (Santos et al., 2003), alterations in gene expression
(Humpherys et al., 2002), and failure to transcribe key developmental genes (Boiani et
al., 2002; Boiani et al., 2003). These abnormalities reduce the developmental potential of
clones and likely terminate the development of many cloned embryos prior to
implantation.
Imprinted genes are established by methylation marks that are placed on genes
during late stages of gametogenesis that are specific to the parent of origin (Constancia et
18
al., 1998; Tilghman, 1999). Once disrupted, imprinting patterns cannot be reestablished
unless cells pass through germ line (Tucker et al., 1996) as imprinted patterns are
removed in the germ cell line and then reset prior to gametogenesis (Tada et al., 1998).
Clones display abnormalities in most imprinted genes, and the disruption of imprints is
linked to the overgrowth commonly seen in clones (Humpherys et al., 2001). The large
offspring syndrome seen in clones could be a result of abnormal patterns of imprinting
(Constancia et al., 1998; Tilghman, 1999). Similar overgrowth abnormalities have been
observed in humans that harbor defects in imprinted genes, as well as in mice in which
targeted mutagenesis has altered the expression of imprinted genes (Jaenisch, 1997).
Aberrations in DNA methylation also contribute to placental abnormalities
commonly observed in clones. Deficient placental formation has been characterized in
SCNT sheep (De Sousa et al., 2001), cattle (Hill et al., 2000a; Chavatte-Palmer et al.,
2002), and mice (Wakayama and Yanagimachi, 1999; Tanaka et al., 2001). Perinatal
problems have also been observed in clones (Hill et al., 1999; Chavatte-Palmer et al.,
2002; Pace et al., 2002), and these may be the result of poor placental development. In
normal embryos, the trophectoderm is globally undermethylated (Monk et al., 1987;
Santos et al., 2002). However, in clones this pattern is reversed, with the trophectoderm
being highly methylated globally (Bourc'his et al., 2001; Dean et al., 2001; Kang et al.,
2001; Kang et al., 2002). The abnormal methylation and gene expression patterns in the
trophectoderm of clones likely contribute to abnormal placental development.
Interestingly, the epigenetic signature of cloned embryos deviates not only from
IVF and in vivo control embryos, but also differ significantly among cloned embryos
19
(Dean et al., 2003; Jouneau and Renard, 2003; Kang et al., 2003; Santos et al., 2003;
Hochedlinger et al., 2004). This observation suggests that methylation patterns following
SCNT are established stochastically. The random patterns of epigenetic marks most
likely contribute to the aberrant patterns of gene expression (Ohgane et al., 2001;
Humpherys et al., 2002).
Further evidence that cloning failure is largely epigenetic in nature is that the types
of epigenetic abnormalities in cloned species, at least among farm animals, are not seen
in the offspring of clones (Tamashiro et al., 2002; Jaenisch, 2004). Offspring born to
clones are normal and do not exhibit any of the developmental abnormalities or high
abortion rates associated with clones, indicating that one generation is sufficient to
reestablish normal epigenetic architecture (Zhang et al., 2004).
Epigenetic Reprogramming
Developmental failure of SCNT embryos is thought to be the result of incomplete
epigenetic reprogramming following transfer of the somatic cell (Wilmut et al., 1998;
Colman, 1999; Kikyo and Wolffe, 2000; Solter, 2000; Dean et al., 2001; Humpherys et
al., 2001; Reik et al., 2001; Niemann et al., 2008). Nuclear reprogramming refers to
transcriptional and epigenetic modifications required to achieve totipotency. The desired
outcome of nuclear transfer is for the somatic nucleus to achieve the same transcriptional
pattern as a zygotic nucleus required to successfully direct embryogenesis. This process
requires silencing of genes expressed in somatic cells, and expression of zygotic-specific
genes.
20
As part of normal mammalian development, several major reprogramming events
alter the genomic state and clear the genome of epigenetic markings as a prelude to
development. The first reprogramming event occurs during primordial germ cell
formation in which imprinted genes have their methylation signatures removed, and
differentiation is reestablished to a totipotent state as part of gametogenesis (Tada et al.,
1997; Reik and Walter, 2001). A second reprogramming event occurs upon fertilization
when gamete-specific marks are erased. In addition to reprogramming associated with
development, a partial reprogramming event is associated with cancer development, in
which cells undergo partial dedifferentiation and the epigenetic signature is altered in
ways that accommodate tumor development and growth. Reprogramming can also be
artificially induced via ectopic expression of key reprogramming factors in order to
generate induced pluripotent stem cells (iPSC) (Takahashi et al., 2007; Yu et al., 2007).
Reprogramming during preimplantation development is likely a dynamic process.
Rather than setting the donor cell nuclei to a ‘zero’ state, reprogramming appears to
continue throughout embryo development and potentially beyond. The demethylation
event common to both the paternal and maternal genome does not occur in a donor cell
nucleus to the same degree (Bourc'his et al., 2001; Dean et al., 2001). Potentially this
difference is because gametes are in some way poised for reprogramming, whereas the
donor nuclei are not. Data suggest that the greater the amount of differentiation, the less
successful is the SCNT process. For example, in amphibians approximately 30% of
nuclear transfer performed with blastula cells are successful, compared to only about 2%
for differentiated cells (Gurdon and Wilmut, 2011). A similar pattern has been observed
21
in mammals. Less differentiated cells, such as blastomeres or fetal fibroblasts (Rideout et
al., 2000), have greater cloning success than adult fibroblasts or terminally differentiated
cells (Hochedlinger and Jaenisch, 2002; Eggan et al., 2004; Li et al., 2004; Inoue et al.,
2005). Reprogramming of the oocyte genome is directed by as yet unknown factors in
the cytoplasm. The reprogramming capacity of oocytes diminishes shortly after cleavage
begins. Beyond the two to four-cell stage, the cytoplasm of blastomeres does not appear
capable of reprogramming donor cell nuclei (Wakayama et al., 2000).
Microarray analysis of gene expression immediately following nuclear transfer
indicates that an overwhelming majority (95%) of genes are expressed normally in cloned
embryos (Humpherys et al., 2002; Smith et al., 2005; Beyhan et al., 2007; Vassena et al.,
2007). Based on this evidence, one can reason that reprogramming is largely properly
achieved. However, the lack of expression of a few key genes within the minority of
non-programmed genes may be sufficient to induce abnormal development (Boiani et al.,
2002; Bortvin et al., 2003; Aston et al., 2010).
Cloned mouse embryos derived from cumulus cells can successfully silence
cumulus-specific genes, demonstrating the capacity for reprogramming the gene
expression patterns of donor cell-specific genes (Tong et al., 2007). This study only
examined four genes, and other studies have indicated donor-specific genes maintain
expression (Gao et al., 2003; Tong et al., 2006). Taken together, these findings suggest
that reprogramming occurs, although reprogramming of specific genes can be
incomplete.
22
In amphibians it has been shown that transferred donor cells retain epigenetic
memory. Donor cells derived from the mesoderm, ectoderm or endoderm maintain
expression of lineage-specific genes upon reaching the blastocyst stage in over half of
cloned blastocysts (Ng and Gurdon, 2008). Such studies highlight that lineage-specific
identity of donor nuclei is retained throughout preimplantation development. A
comparable experiment has yet to be carried out in mammals, although a study from the
mouse suggests that the same may hold true of mammalian cloning. The use of myoblast
nuclei as a donor source, followed by embryo culture in myoblast culture medium,
triggers the expression of muscle-specific cell markers in cloned embryos (Gao et al.,
2003).
The successful development of cloned embryos may hinge on the proper expression
of pluripotency factors such as POU5F1 (Boiani et al., 2002; Bortvin et al., 2003; Byrne
et al., 2003; Armstrong et al., 2006; Lee et al., 2006) as preimplantation embryos require
POU5F1 in order to develop (Nichols et al., 1998; Pantazis and Bollenbach, 2012).
POU5F1 is a transcription factor that directs the expression of multiple developmental
genes and is essential for early development (Niwa et al., 2000). SCNT embryos fail to
upregulate embryo-specific genes including POU5F1 (Boiani et al., 2002; Bortvin et al.,
2003; Kishigami et al., 2006). It has been reported in some mouse studies that the use of
ES cells as a donor source have higher cloning efficiency over somatic cells (Wakayama
et al., 1999; Rideout et al., 2000; Eggan et al., 2001), which may be due, in part, to
expression of POU5F1 in ES cells (Nichols et al., 1998; Pantazis and Bollenbach, 2012).
Exposure of donor cells to cell extracts designed to induce reprogramming prior to
23
transfer has had limited success (Landsverk et al., 2002; Hansis et al., 2004), likely
because these agents do not discriminate against terminally silenced regions and are not
specific to the few genomic regions that require epigenetic activation. Recent work has
demonstrated that injecting mouse donor cells into Xenopus eggs leads to demethylation
of the mouse POU5F1 gene (Simonsson and Gurdon, 2004). Such an outcome is
unexpected, given that the Xenopus zygote does not undergo active demethylation
(Stancheva et al., 2002). The results suggest that a conserved, yet to be identified
mechanism exists that is capable of genome reprogramming.
Genetic mutations accumulate during aging and during prolonged periods of in
vitro culture (DePinho, 2000). This opens the possibility that cloning failure is in part
caused by genetic abnormalities. However, the fact that cloning abnormalities are shared
among all species, as well the observation that cloned animals are able to give birth to
offspring without any mutant phenotype strongly suggests that the deficiencies in cloned
organisms are epigenetic in nature. It is unclear however, how much of reprogramming
process is mediated by the oocyte environment versus the epigenetic state of the
individual donor cell at the time of transfer (Roemer et al., 1997).
In cattle, no difference has been seen between cloning success using donor cells
collected from young versus old individuals or low versus high passage fibroblasts (Hill
et al., 2000b; Kubota et al., 2000). However a conflicting report from sheep suggests that
rates of cloning failure are higher with high passage number fibroblasts (Schnieke et al.,
1997; Wilmut et al., 1997; McCreath et al., 2000). However, it is important to note that
in these experiments, the overall success rate of any sample was extremely low, making it
24
difficult to draw statistically significant conclusions.
Although several examples demonstrate that cloned embryos are characterized by
reprogramming failures a limited number of cloned offspring appear normal in all
regards, suggesting that appropriate reprogramming can occur (Chavatte-Palmer et al.,
2002; Cibelli et al., 2002). Additionally, some aspects of embryo development appear to
be normal in clones. X-chromosome inactivation occurs following the formation of the
zygote. X chromosomes are randomly silenced within ICM cells, but in trophoblast cells,
an unknown imprinting mechanism causes preferential inactivation of the paternal Xchromosome. Cloned embryos follow a similar pattern of paternal X-chromosome
inactivation in trophoblasts (Eggan et al., 2000), indicating that clones retain this
hierarchical silencing process.
The P53-Mediated Apoptosis Pathway
The transcription factor P53 plays a central role in cellular health and ensures
genomic integrity by initiating programmed cell death within defective and aberrant cells.
During P53-mediated apoptosis, the cell shrinks, DNA is fragmented and membrane
blebbing occurs. These features allow the degraded cell contents to be phagocytosed,
thus preventing inflammation in surrounding tissue that would occur through cell death
without apoptosis (Kerr et al., 1972). P53 acts as a sensor to eliminate unneeded cells
and prevent overgrowth, as well as eliminating cells under stress, with DNA damage and
that grow uncontrollably. Due to its role in enforcing genomic stability, P53 is
commonly referred to as the guardian of the genome. By preventing mitotic division of
cells with DNA damage or oncogene expression, P53 prevents tumorigenic growth. The
25
critical role of P53 in preventing growth of aberrant cells is clearly shown by the high
rate and early development of tumors in P53 knockout mice (Donehower et al., 1992).
Under normal conditions, cells maintain a constant but low level of P53 within the
cytoplasm. Because of its potential to terminate progression through the cell cycle,
mechanisms that maintain P53 in an inactive form are strictly regulated. The regulator
MDM2 forms a negative feedback loop to prevent P53 activation by consistently
targeting P53 for degradation, thereby preventing accumulation of P53 within the cell
(Moll and Petrenko, 2003; Kruse and Gu, 2009).
Activation of P53 can be induced by a variety of triggers including DNA damage,
hypoxia, nucleotide depletion, reactive oxidative species, and oncogene activation, with
the activation method determining the action carried out by P53. Activation of the P53
pathway may occur directly via various input signals, or activation can arise indirectly
through the inhibition of MDM2 that leads to the accumulation and subsequent activation
of P53. Upon activation, P53 homodimerizes and translocates to the nucleus where it
upregulates a number of genes to carry out cell cycle arrest, senescence or apoptosis
(Vousden and Lu, 2002).
A number of genes upregulated by P53 have been identified (Riley et al., 2008).
Upon activation, P53 can either induce cell cycle arrest or initiate programmed cell death.
Under normal conditions, cell cycle progression is induced by cyclin-dependent kinases
(CDK), which are upregulated in preparation for progression through different stages of
the cell cycle. CDKs form a complex with cyclin which acts as an inducer for the cell
cycle. P21WAF1/Cip1 (CDKN1A) is upregulated by P53 (Brugarolas et al., 1999) and acts
26
by inhibiting cyclin/CDK to prevent cell cycle progression (Niculescu et al., 1998).
Thus, P53 upregulation of P21WAF1/Cip1 induces cell cycle arrest and prevents cell
division.
Aside from cell cycle arrest, activation of P53-dependent gene expression leads to
apoptosis. Alternatively, P53 can directly activate apoptosis in a transcriptionindependent manner. Either pathway ultimately leads to the release of cytochrome c
from the mitochondrial membrane, which acts as a mediator of programmed cell death
(Wei et al., 2001). In both transcription-dependent and transcription-independent
pathways, the downstream effects of P53 at the mitochondria involve members of the
BCL2 family of proteins (Cory and Adams, 2002) that can either activate or inhibit
apoptosis by controlling release of cytochrome c from the mitochondria.
Pro-apoptotic members of the BCL2 protein family include BAX, BAK, BID
NOXA, and PUMA. These pro-apoptotic proteins can be upregulated by P53 (Oltvai et
al., 1993; Miyashita and Reed, 1995) and oligomerize to form a channel on the
membrane of mitochondria through which cytochrome c is released (Miyashita and Reed,
1995; Chipuk et al., 2004). Anti-apoptotic members of the BCL2 protein family include
BCL-2, BCL-xl, and MCL-1. These proteins antagonize pro-apoptosis inducing proteins.
BAX activity is inhibited by BCL2, which heterodimerizes with BAX and inhibits BAXmediated induction of apoptosis (Yin et al., 1994). The BCL2:BAX ratio controls
whether apoptosis is induced or inhibited (Basu and Haldar, 1998). BCL2 is expressed
continually and maintains a constant inhibition of BAX. However, upon upregulation of
27
BAX by P53, the BCL2:BAX ratio is tipped in favor of BAX, allowing BAX to
oligomerize and initiate apoptosis.
The P53 pathway may also act in a transcription-independent manner to induce
apoptosis (Caelles et al., 1994; Wagner et al., 1994; Yan et al., 1997; Gao and Tsuchida,
1999). Deletion of the DNA-binding domain of P53 maintains its apoptosis-inducing
ability under certain conditions (Haupt et al., 1995; Chen et al., 1996; Haupt et al., 1997).
When acting as a direct activator of apoptosis, P53 translocates to the mitochondria
(Marchenko et al., 2000; Sansome et al., 2001; Mihara et al., 2003) where it initiates
apoptosis by either the activation of pro-apoptosis proteins, or the inhibition of antiapoptosis proteins (Mihara et al., 2003; Tomita et al., 2006). P53 is able to both directly
activate BAX and to inhibit BCL2 (Wang et al., 1993; Chiou et al., 1994; Marin et al.,
1994; Froesch et al., 1999).
The transcription-independent pathway of P53-directed apoptosis has been linked
to apoptosis induction by reactive oxygen species (Han et al., 2008). However, DNA
damage response and tumor suppression requires P53 to act in a transcription-dependent
manner (Chao et al., 2000; Jimenez et al., 2000; Johnson et al., 2005; Gaidarenko and Xu,
2009). In cells that undergo cell cycle arrest, P53 does not localize to the mitochondria.
In this case, cell cycle arrest is a transcription-dependent process (Marchenko et al., 2000;
Erster et al., 2004) and involves the upregulation of P21WAF1/Cip1.
Via its role in regulating apoptosis, P53 is a vital component of normal
development and maintenance of tissue and organs. For example, P53-null mice have a
28
high incidence of birth defects, such as polydactylism, problems in bone development,
and abnormalities in neural tube formation (Armstrong et al., 1995).
P53-mediated apoptosis also plays a role during preimplantation development
(Hardy et al., 2001; Gjorret et al., 2003). P53 prevents the integration of blastomeres
with genetic abnormalities (Fabian et al., 2005). Conversely, excessive accumulation of
P53 within the developing embryo is a direct impediment to embryo viability and
implantation (Ganeshan et al., 2010). In mouse preimplantation embryos, decreased
apoptosis levels result in significantly higher rates of fetal development following
embryo transfer (Li et al., 2007).
Recently, P53 was revealed to play a major role in the efficiency of
reprogramming iPSCs, as the elimination of P53 dramatically improved reprogramming
efficiency (Zhao et al., 2008). P53-null cell lines have dramatically higher success rates
of reprogramming (Hong et al., 2009; Kawamura et al., 2009). Additionally, cell lines
that have previously failed to reprogram have been able to successfully do so following
p53 inhibition (Utikal et al., 2009).
The reprogramming pathway common in iPSC formation induces apoptosis,
senescence, and cell cycle arrest. Reprogramming causes DNA damage, which can
activate the P53 apoptosis pathway (Marion et al., 2009). Additionally, the
reprogramming factors used to induce somatic cell potency also have oncogenic effects
that likely induce apoptosis. The six factors regularly used in reprogramming
experiments, C-MYC, LIN28, KLF4, POU5F1, NANOG, and SOX2 (Takahashi et al.,
2007; Yu et al., 2007; Silva et al., 2009), are also involved in the induction of apoptosis.
29
C-MYC is an established oncogene that induces apoptosis when overexpressed
(Pelengaris et al., 2002; Rowland and Peeper, 2006). P53-null cell lines can be induced
to become stem cells by the addition of the reprogramming factors POU5F1 and SOX2
alone, suggesting that one function of C-MYC in reprogramming may be to select for
cells with reduced P53 functionality (Eischen et al., 1999; Kawamura et al., 2009).
LIN28 is a translational enhancer of various oncogenes, including C-MYC (West et al.,
2009), and its presence likely enhances C-MYC expression and increases apoptosis.
KLF4 suppresses P53 (Rowland and Peeper, 2006), while simultaneously inducing the
expression of P21WAF1/Cip1 to inhibit cell cycle progression (Zhang et al., 2000; Chen et
al., 2003; Rowland et al., 2005; Rowland and Peeper, 2006). NANOG, POU5F1, and
SOX2 are required for potency and are overexpressed in some types of cancers (Almstrup
et al., 2004; Ezeh et al., 2005; Hart et al., 2005; Hochedlinger et al., 2005). P53 is
activated by these factors either collectively (Hong et al., 2009; Kawamura et al., 2009;
Marion et al., 2009), or individually (Kawamura et al., 2009). Additionally, SOX2 and
POU5F1 act to trigger cell cycle arrest by upregulating P21WAF1/Cip1 via P53 induction
(Kawamura et al., 2009).
As SCNT requires reprogramming of the somatic donor cell, the activation of
P53-mediated apoptosis may be responsible for increased cell death following somatic
cell nuclear transfer. However, no studies have established the expression of
reprogramming associated genes or the upregulation of P53 targets following SCNT.
Apoptosis has been reported to occur with higher frequency in embryos produced by
SCNT than embryos produced by IVF (Fahrudin et al., 2002; Hao et al., 2003; Jang et al.,
30
2004; Gjorret et al., 2005; Im et al., 2006; Kumar et al., 2007), and elevated apoptosis
continues throughout development (Hao et al., 2003). However, it is unclear if increased
apoptosis is the result of sub-optimal culture conditions, the stress associated with the
cloning procedure, or possibly the byproduct of donor cell reprogramming. The cloning
procedure may induce higher levels of apoptosis. For example, porcine clones activated
by electroporation alone have reduced apoptosis relative to clones activated by
electroporation with chemical activation (Im et al., 2006).
In addition to increased levels of apoptosis, cloned embryos have a reduced
number of blastocyst cells that likely is due to increased apoptosis (Koo et al., 2000;
Fahrudin et al., 2002; Gjorret et al., 2005; Terashita et al., 2011). A reduced blastomere
number likely results in reduced embryo viability and implantation potential (van Soom
et al., 1997; Wuu et al., 1999). Both apoptosis and total cell count reflect important
developmental parameters that influence the developmental potential of preimplantation
embryos (Brison and Schultz, 1997; Brison and Schultz, 1998; Koo et al., 2000; Knijn et
al., 2003). Should the reprogramming process in SCNT embryos activate P53-mediated
apoptosis, this would suggest that the high rates of embryo failure might, in part, be the
result of donor cell reprogramming.
Summary
A great deal of effort by many research groups has been applied to improving the
efficiency of SCNT. A large body of evidence supports the hypothesis that clones suffer
from a failure to reset the epigenetic program of the somatic donor cell. There is also
evidence that preimplantation SCNT embryos experience elevated levels of apoptosis,
31
which likely contributes to failed embryo development and implantation. A separate line
of inquiry has demonstrated that somatic cell reprogramming activates the P53 apoptosis
pathway and induces programmed cell death in cells that are aberrantly reprogrammed
within the somatic cell genome. Should the same pattern of reprogramming-induced
apoptosis occur in cloned embryos, then SCNT embryo failure may be the result of
successful reprogramming, rather than a reprogramming failure.
Research Goals and Potential Applications
The first objective of my dissertation was to characterize the patterns of DNA
methylation in SCNT embryos, in order to identify the epigenetic reprogramming status
of several important developmental genes following SCNT, and to understanding the
regulation of DNA methylation. The second objective was to characterize p53-mediated
apoptosis following SCNT.
For the first objective, I was able to demonstrate that reprogramming is
incomplete following SCNT, and that methylation patterns at some CpG sites in key
genes of SCNT embryos are more similar those found in the donor cell than those seen in
IVF embryos. In a separate study related to DNA methylation, I characterized the
methylation status of a family of LIN28 pseudogenes. These studies hold the potential to
better understand how the addition and maintenance of methylation patterns are
regulated.
For the second objective, I confirmed the activation by P53-mediated apoptosis
within SCNT preimplantation embryos, and identified several P53-activated genes that
are upregulated in clones. Importantly, my work demonstrated that the reprogramming
32
factors C-MYC, SOX2, KLF4, and LIN28 are upregulated in clones. My findings suggest
that SCNT embryos have elevated expression of reprogramming factors. The expression
of these factors likely activates P53, leading to increased apoptosis and cell cycle arrest in
SCNT embryos. These findings identify several genes that contribute to blastomere
depletion and failed development.
SCNT holds immense potential in the production of transgenic animals, the
generation of stem cells, and the production of species with superior traits. However, all
potential benefits are constrained by the low efficiency of the technique. My research
strengthens our understanding of the nature of embryo failure, and reveals potential
interventions to inhibit apoptosis to allow SCNT with higher developmental rates.
Additionally, my findings in the area of DNA methylation contribute to greater
understanding of the reprogramming process and have application in several areas of
biology, including stem cell biology, reprogramming, epigenetics, and cancer.
References
Almstrup K, Hoei-Hansen CE, Wirkner U, Blake J, Schwager C, Ansorge W, Nielsen JE,
Skakkebaek NE, Rajpert-De Meyts E, Leffers H. 2004. Embryonic stem cell-like
features of testicular carcinoma in situ revealed by genome-wide gene expression
profiling. Cancer Res 64(14):4736-4743.
Antequera F, Bird A. 1993. Number of CpG islands and genes in human and mouse. Proc
Natl Acad Sci USA 90(24):11995-11999.
Armstrong JF, Kaufman MH, Harrison DJ, Clarke AR. 1995. High-frequency
developmental abnormalities in p53-deficient mice. Curr Biol 5(8):931-936.
Armstrong L, Lako M, Dean W, Stojkovic M. 2006. Epigenetic modification is central to
genome reprogramming in somatic cell nuclear transfer. Stem Cells 24(4):805814.
33
Aston KI, Li GP, Hicks BA, Sessions BR, Davis AP, Rickords LF, Stevens JR, White
KL. 2010. Abnormal levels of transcript abundance of developmentally important
genes in various stages of preimplantation bovine somatic cell nuclear transfer
embryos. Cell Reprogram 12(1):23-32.
Baguisi A, Behboodi E, Melican DT, Pollock JS, Destrempes MM, Cammuso C,
Williams JL, Nims SD, Porter CA, Midura P, Palacios MJ, Ayres SL, Denniston
RS, Hayes ML, Ziomek CA, Meade HM, Godke RA, Gavin WG, Overstrom EW,
Echelard Y. 1999. Production of goats by somatic cell nuclear transfer. Nat
Biotechnol 17(5):456-461.
Barberi T, Klivenyi P, Calingasan NY, Lee H, Kawamata H, Loonam K, Perrier AL,
Bruses J, Rubio ME, Topf N, Tabar V, Harrison NL, Beal MF, Moore MA,
Studer L. 2003. Neural subtype specification of fertilization and nuclear transfer
embryonic stem cells and application in parkinsonian mice. Nat Biotechnol
21(10):1200-1207.
Bartolomei MS, Webber AL, Brunkow ME, Tilghman SM. 1993. Epigenetic mechanisms
underlying the imprinting of the mouse H19 gene. Genes Dev 7(9):1663-1673.
Basu A, Haldar S. 1998. The relationship between BcI2, Bax and p53: consequences for
cell cycle progression and cell death. Mol Hum Reprod 4(12):1099-1109.
Bell AC, Felsenfeld G. 2000. Methylation of a CTCF-dependent boundary controls
imprinted expression of the Igf2 gene. Nature 405(6785):482-485.
Bender J. 2001. A vicious cycle: RNA silencing and DNA methylation in plants. Cell
106(2):129-132.
Berg DK, Li C, Asher G, Wells DN, Oback B. 2007. Red deer cloned from antler stem
cells and their differentiated progeny. Biol Reprod 77(3):384-394.
Bergman Y, Cedar H. 2013. DNA methylation dynamics in health and disease. Nat Struct
Mol Biol 20(3):274-281.
Bestor TH. 1992. Activation of mammalian DNA methyltransferase by cleavage of a Zn
binding regulatory domain. EMBO J 11(7):2611-2617.
Bestor TH. 2000. The DNA methyltransferases of mammals. Hum Mol Genet
9(16):2395-2402.
Betthauser J, Forsberg E, Augenstein M, Childs L, Eilertsen K, Enos J, Forsythe T,
Golueke P, Jurgella G, Koppang R, Lesmeister T, Mallon K, Mell G, Misica P,
Pace M, Pfister-Genskow M, Strelchenko N, Voelker G, Watt S, Thompson S,
Bishop M. 2000. Production of cloned pigs from in vitro systems. Nat Biotechnol
18(10):1055-1059.
34
Beyhan Z, Ross PJ, Iager AE, Kocabas AM, Cunniff K, Rosa GJ, Cibelli JB. 2007.
Transcriptional reprogramming of somatic cell nuclei during preimplantation
development of cloned bovine embryos. Dev Biol 305(2):637-649.
Birchler JA, Bhadra MP, Bhadra U. 2000. Making noise about silence: repression of
repeated genes in animals. Curr Opin Genet Dev 10(2):211-216.
Bird A. 1987. CpG islands as gene markers in the vertebrate nucleus. Trends Genet
3:342-347.
Bird A. 2002. DNA methylation patterns and epigenetic memory. Genes Dev 16(1):6-21.
Bird AP. 1986. CpG-rich islands and the function of DNA methylation. Nature
321(6067):209-213.
Boiani M, Eckardt S, Leu NA, Scholer HR, McLaughlin KJ. 2003. Pluripotency deficit in
clones overcome by clone-clone aggregation: epigenetic complementation?
EMBO J 22(19):5304-5312.
Boiani M, Eckardt S, Scholer HR, McLaughlin KJ. 2002. Oct4 distribution and level in
mouse clones: consequences for pluripotency. Genes Dev 16(10):1209-1219.
Bortvin A, Eggan K, Skaletsky H, Akutsu H, Berry DL, Yanagimachi R, Page DC,
Jaenisch R. 2003. Incomplete reactivation of Oct4-related genes in mouse
embryos cloned from somatic nuclei. Development 130(8):1673-1680.
Bourc'his D, Le Bourhis D, Patin D, Niveleau A, Comizzoli P, Renard JP, ViegasPequignot E. 2001. Delayed and incomplete reprogramming of chromosome
methylation patterns in bovine cloned embryos. Curr Biol 11(19):1542-1546.
Boyes J, Bird A. 1992. Repression of genes by DNA methylation depends on CpG
density and promoter strength: evidence for involvement of a methyl-CpG
binding protein. EMBO J 11(1):327-333.
Brandeis M, Frank D, Keshet I, Siegfried Z, Mendelsohn M, Nemes A, Temper V, Razin
A, Cedar H. 1994. Sp1 elements protect a CpG island from de novo methylation.
Nature 371(6496):435-438.
Briggs R, King TJ. 1952. Transplantation of living nuclei from blastula cells into
enucleated frogs' eggs. Proc Natl Acad Sci USA 38(5):455-463.
Briggs R, King TJ. 1960. Nuclear transplantation studies on the early gastrula (Rana
pipiens). I. Nuclei of presumptive endoderm. Dev Biol 2:252-270.
35
Brison DR, Schultz RM. 1997. Apoptosis during mouse blastocyst formation: evidence
for a role for survival factors including transforming growth factor alpha. Biol
Reprod 56(5):1088-1096.
Brison DR, Schultz RM. 1998. Increased incidence of apoptosis in transforming growth
factor alpha-deficient mouse blastocysts. Biol Reprod 59(1):136-144.
Bromhall JD. 1975. Nuclear transplantation in the rabbit egg. Nature 258(5537):719-722.
Brophy B, Smolenski G, Wheeler T, Wells D, L'Huillier P, Laible G. 2003. Cloned
transgenic cattle produce milk with higher levels of beta-casein and kappa-casein.
Nat Biotechnol 21(2):157-162.
Brugarolas J, Moberg K, Boyd SD, Taya Y, Jacks T, Lees JA. 1999. Inhibition of cyclindependent kinase 2 by p21 is necessary for retinoblastoma protein-mediated G1
arrest after gamma-irradiation. Proc Natl Acad Sci USA 96(3):1002-1007.
Byrne JA, Simonsson S, Western PS, Gurdon JB. 2003. Nuclei of adult mammalian
somatic cells are directly reprogrammed to oct-4 stem cell gene expression by
amphibian oocytes. Curr Biol 13(14):1206-1213.
Caelles C, Helmberg A, Karin M. 1994. p53-dependent apoptosis in the absence of
transcriptional activation of p53-target genes. Nature 370(6486):220-223.
Challah-Jacques M, Chesne P, Renard JP. 2003. Production of cloned rabbits by somatic
nuclear transfer. Cloning Stem Cells 5(4):295-299.
Chao C, Saito S, Kang J, Anderson CW, Appella E, Xu Y. 2000. p53 transcriptional
activity is essential for p53-dependent apoptosis following DNA damage. EMBO
J 19(18):4967-4975.
Chavatte-Palmer P, Camous S, Jammes H, Le Cleac'h N, Guillomot M, Lee RS. 2012.
Review: Placental perturbations induce the developmental abnormalities often
observed in bovine somatic cell nuclear transfer. Placenta 33 Suppl:S99-S104.
Chavatte-Palmer P, Heyman Y, Richard C, Monget P, LeBourhis D, Kann G, Chilliard Y,
Vignon X, Renard JP. 2002. Clinical, hormonal, and hematologic characteristics
of bovine calves derived from nuclei from somatic cells. Biol Reprod 66(6):15961603.
Chen RZ, Pettersson U, Beard C, Jackson-Grusby L, Jaenisch R. 1998. DNA
hypomethylation leads to elevated mutation rates. Nature 395(6697):89-93.
Chen X, Ko LJ, Jayaraman L, Prives C. 1996. p53 levels, functional domains, and DNA
damage determine the extent of the apoptotic response of tumor cells. Genes Dev
10(19):2438-2451.
36
Chen X, Whitney EM, Gao SY, Yang VW. 2003. Transcriptional profiling of Kruppellike factor 4 reveals a function in cell cycle regulation and epithelial
differentiation. J Mol Biol 326(3):665-677.
Chesne P, Adenot PG, Viglietta C, Baratte M, Boulanger L, Renard JP. 2002. Cloned
rabbits produced by nuclear transfer from adult somatic cells. Nat Biotechnol
20(4):366-369.
Chiou SK, Rao L, White E. 1994. Bcl-2 blocks p53-dependent apoptosis. Mol Cell Biol
14(4):2556-2563.
Chipuk JE, Kuwana T, Bouchier-Hayes L, Droin NM, Newmeyer DD, Schuler M, Green
DR. 2004. Direct activation of Bax by p53 mediates mitochondrial membrane
permeabilization and apoptosis. Science 303(5660):1010-1014.
Chung YG, Ratnam S, Chaillet JR, Latham KE. 2003. Abnormal regulation of DNA
methyltransferase expression in cloned mouse embryos. Biol Reprod 69(1):146153.
Cibelli JB, Campbell KH, Seidel GE, West MD, Lanza RP. 2002. The health profile of
cloned animals. Nat Biotechnol 20(1):13-14.
Cibelli JB, Stice SL, Golueke PJ, Kane JJ, Jerry J, Blackwell C, Ponce de Leon FA, Robl
JM. 1998a. Cloned transgenic calves produced from nonquiescent fetal
fibroblasts. Science 280(5367):1256-1258.
Cibelli JB, Stice SL, Golueke PJ, Kane JJ, Jerry J, Blackwell C, Ponce de Leon FA, Robl
JM. 1998b. Transgenic bovine chimeric offspring produced from somatic cellderived stem-like cells. Nat Biotechnol 16(7):642-646.
Colman A. 1999. Somatic cell nuclear transfer in mammals: progress and applications.
Cloning 1(4):185-200.
Constancia M, Pickard B, Kelsey G, Reik W. 1998. Imprinting mechanisms. Genome Res
8(9):881-900.
Cory S, Adams JM. 2002. The Bcl2 family: regulators of the cellular life-or-death switch.
Nat Rev Cancer 2(9):647-656.
Cross SH, Meehan RR, Nan X, Bird A. 1997. A component of the transcriptional
repressor MeCP1 shares a motif with DNA methyltransferase and HRX proteins.
Nat Genet 16(3):256-259.
Csankovszki G, Nagy A, Jaenisch R. 2001. Synergism of Xist RNA, DNA methylation,
and histone hypoacetylation in maintaining X chromosome inactivation. J Cell
Biol 153(4):773-784.
37
Cuadrado M, Sacristan M, Antequera F. 2001. Species-specific organization of CpG
island promoters at mammalian homologous genes. EMBO Rep 2(7):586-592.
De Smet C, De Backer O, Faraoni I, Lurquin C, Brasseur F, Boon T. 1996. The activation
of human gene MAGE-1 in tumor cells is correlated with genome-wide
demethylation. Proc Natl Acad Sci USA 93(14):7149-7153.
De Smet C, Lurquin C, Lethe B, Martelange V, Boon T. 1999. DNA methylation is the
primary silencing mechanism for a set of germ line- and tumor-specific genes
with a CpG-rich promoter. Mol Cell Biol 19(11):7327-7335.
De Sousa PA, King T, Harkness L, Young LE, Walker SK, Wilmut I. 2001. Evaluation of
gestational deficiencies in cloned sheep fetuses and placentae. Biol Reprod
65(1):23-30.
Dean W, Bowden L, Aitchison A, Klose J, Moore T, Meneses JJ, Reik W, Feil R. 1998.
Altered imprinted gene methylation and expression in completely ES cell-derived
mouse fetuses: association with aberrant phenotypes. Development 125(12):22732282.
Dean W, Santos F, Reik W. 2003. Epigenetic reprogramming in early mammalian
development and following somatic nuclear transfer. Semin Cell Dev Biol
14(1):93-100.
Dean W, Santos F, Stojkovic M, Zakhartchenko V, Walter J, Wolf E, Reik W. 2001.
Conservation of methylation reprogramming in mammalian development:
aberrant reprogramming in cloned embryos. Proc Natl Acad Sci USA
98(24):13734-13738.
Denning C, Burl S, Ainslie A, Bracken J, Dinnyes A, Fletcher J, King T, Ritchie M,
Ritchie WA, Rollo M, de Sousa P, Travers A, Wilmut I, Clark AJ. 2001. Deletion
of the alpha(1,3)galactosyl transferase (GGTA1) gene and the prion protein (PrP)
gene in sheep. Nat Biotechnol 19(6):559-562.
Dennis K, Fan T, Geiman T, Yan Q, Muegge K. 2001. Lsh, a member of the SNF2
family, is required for genome-wide methylation. Genes Dev 15(22):2940-2944.
DePinho RA. 2000. The age of cancer. Nature 408(6809):248-254.
Donehower LA, Harvey M, Slagle BL, McArthur MJ, Montgomery CA, Jr., Butel JS,
Bradley A. 1992. Mice deficient for p53 are developmentally normal but
susceptible to spontaneous tumours. Nature 356(6366):215-221.
Eggan K, Akutsu H, Hochedlinger K, Rideout W, 3rd, Yanagimachi R, Jaenisch R. 2000.
X-Chromosome inactivation in cloned mouse embryos. Science 290(5496):15781581.
38
Eggan K, Akutsu H, Loring J, Jackson-Grusby L, Klemm M, Rideout WM, 3rd,
Yanagimachi R, Jaenisch R. 2001. Hybrid vigor, fetal overgrowth, and viability of
mice derived by nuclear cloning and tetraploid embryo complementation. Proc
Natl Acad Sci USA 98(11):6209-6214.
Eggan K, Baldwin K, Tackett M, Osborne J, Gogos J, Chess A, Axel R, Jaenisch R.
2004. Mice cloned from olfactory sensory neurons. Nature 428(6978):44-49.
Ehrlich M, Gama-Sosa MA, Huang LH, Midgett RM, Kuo KC, McCune RA, Gehrke C.
1982. Amount and distribution of 5-methylcytosine in human DNA from different
types of tissues of cells. Nucleic Acids Res 10(8):2709-2721.
Eischen CM, Weber JD, Roussel MF, Sherr CJ, Cleveland JL. 1999. Disruption of the
ARF-Mdm2-p53 tumor suppressor pathway in Myc-induced lymphomagenesis.
Genes Dev 13(20):2658-2669.
Epsztejn-Litman S, Feldman N, Abu-Remaileh M, Shufaro Y, Gerson A, Ueda J, Deplus
R, Fuks F, Shinkai Y, Cedar H, Bergman Y. 2008. De novo DNA methylation
promoted by G9a prevents reprogramming of embryonically silenced genes. Nat
Struct Mol Biol 15(11):1176-1183.
Erster S, Mihara M, Kim RH, Petrenko O, Moll UM. 2004. In vivo mitochondrial p53
translocation triggers a rapid first wave of cell death in response to DNA damage
that can precede p53 target gene activation. Mol Cell Biol 24(15):6728-6741.
Ezeh UI, Turek PJ, Reijo RA, Clark AT. 2005. Human embryonic stem cell genes OCT4,
NANOG, STELLAR, and GDF3 are expressed in both seminoma and breast
carcinoma. Cancer 104(10):2255-2265.
Fabian D, Koppel J, Maddox-Hyttel P. 2005. Apoptotic processes during mammalian
preimplantation development. Theriogenology 64(2):221-231.
Fahrudin M, Otoi T, Karja NW, Mori M, Murakami M, Suzuki T. 2002. Analysis of
DNA fragmentation in bovine somatic nuclear transfer embryos using TUNEL.
Reproduction 124(6):813-819.
Fang ZF, Gai H, Huang YZ, Li SG, Chen XJ, Shi JJ, Wu L, Liu A, Xu P, Sheng HZ.
2006. Rabbit embryonic stem cell lines derived from fertilized, parthenogenetic or
somatic cell nuclear transfer embryos. Exp Cell Res 312(18):3669-3682.
Farthing CR, Ficz G, Ng RK, Chan CF, Andrews S, Dean W, Hemberger M, Reik W.
2008. Global mapping of DNA methylation in mouse promoters reveals
epigenetic reprogramming of pluripotency genes. PLoS Genet 4(6):e1000116.
39
Feldman N, Gerson A, Fang J, Li E, Zhang Y, Shinkai Y, Cedar H, Bergman Y. 2006.
G9a-mediated irreversible epigenetic inactivation of Oct-3/4 during early
embryogenesis. Nat Cell Biol 8(2):188-194.
Feng Q, Zhang Y. 2001. The MeCP1 complex represses transcription through preferential
binding, remodeling, and deacetylating methylated nucleosomes. Genes Dev
15(7):827-832.
Folch J, Cocero MJ, Chesne P, Alabart JL, Dominguez V, Cognie Y, Roche A,
Fernandez-Arias A, Marti JI, Sanchez P, Echegoyen E, Beckers JF, Bonastre AS,
Vignon X. 2009. First birth of an animal from an extinct subspecies (Capra
pyrenaica pyrenaica) by cloning. Theriogenology 71(6):1026-1034.
Francis NJ, Kingston RE. 2001. Mechanisms of transcriptional memory. Nat Rev Mol
Cell Biol 2(6):409-421.
Froesch BA, Aime-Sempe C, Leber B, Andrews D, Reed JC. 1999. Inhibition of p53
transcriptional activity by Bcl-2 requires its membrane-anchoring domain. J Biol
Chem 274(10):6469-6475.
Gaidarenko O, Xu Y. 2009. Transcription activity is required for p53-dependent tumor
suppression. Oncogene 28(49):4397-4401.
Gal-Yam EN, Egger G, Iniguez L, Holster H, Einarsson S, Zhang X, Lin JC, Liang G,
Jones PA, Tanay A. 2008. Frequent switching of Polycomb repressive marks and
DNA hypermethylation in the PC3 prostate cancer cell line. Proc Natl Acad Sci
USA 105(35):12979-12984.
Galli C, Lagutina I, Crotti G, Colleoni S, Turini P, Ponderato N, Duchi R, Lazzari G.
2003. Pregnancy: a cloned horse born to its dam twin. Nature 424(6949):635.
Ganeshan L, Li A, O'Neill C. 2010. Transformation-related protein 53 expression in the
early mouse embryo compromises preimplantation embryonic development by
preventing the formation of a proliferating inner cell mass. Biol Reprod
83(6):958-964.
Gao C, Tsuchida N. 1999. Activation of caspases in p53-induced transactivationindependent apoptosis. Jpn J Cancer Res 90(2):180-187.
Gao S, Chung YG, Williams JW, Riley J, Moley K, Latham KE. 2003. Somatic cell-like
features of cloned mouse embryos prepared with cultured myoblast nuclei. Biol
Reprod 69(1):48-56.
Gautsch JW, Wilson MC. 1983. Delayed de novo methylation in teratocarcinoma
suggests additional tissue-specific mechanisms for controlling gene expression.
Nature 301(5895):32-37.
40
Gerdes T, Wabl M. 2004. Autoreactivity and allelic inclusion in a B cell nuclear transfer
mouse. Nat Immunol 5(12):1282-1287.
Gibbons RJ, McDowell TL, Raman S, O'Rourke DM, Garrick D, Ayyub H, Higgs DR.
2000. Mutations in ATRX, encoding a SWI/SNF-like protein, cause diverse
changes in the pattern of DNA methylation. Nat Genet 24(4):368-371.
Gjorret JO, Knijn HM, Dieleman SJ, Avery B, Larsson LI, Maddox-Hyttel P. 2003.
Chronology of apoptosis in bovine embryos produced in vivo and in vitro. Biol
Reprod 69(4):1193-1200.
Gjorret JO, Wengle J, Maddox-Hyttel P, King WA. 2005. Chronological appearance of
apoptosis in bovine embryos reconstructed by somatic cell nuclear transfer from
quiescent granulosa cells. Reprod Domest Anim 40(3):210-216.
Gomez MC, Pope CE, Giraldo A, Lyons LA, Harris RF, King AL, Cole A, Godke RA,
Dresser BL. 2004. Birth of African Wildcat cloned kittens born from domestic
cats. Cloning Stem Cells 6(3):247-258.
Gowher H, Leismann O, Jeltsch A. 2000. DNA of Drosophila melanogaster contains 5methylcytosine. EMBO J 19(24):6918-6923.
Graves JA. 1982. 5-azacytidine-induced re-expression of alleles on the inactive X
chromosome in a hybrid mouse cell line. Exp Cell Res 141(1):99-105.
Gurdon JB. 1962. The developmental capacity of nuclei taken from intestinal epithelium
cells of feeding tadpoles. J Embryol Exp Morphol 10:622-640.
Gurdon JB, Laskey RA, Reeves OR. 1975. The developmental capacity of nuclei
transplanted from keratinized skin cells of adult frogs. J Embryol Exp Morphol
34(1):93-112.
Gurdon JB, Wilmut I. 2011. Nuclear transfer to eggs and oocytes. Cold Spring Harb
Perspect Biol 3(6).
Han H, Cortez CC, Yang X, Nichols PW, Jones PA, Liang G. 2011. DNA methylation
directly silences genes with non-CpG island promoters and establishes a
nucleosome occupied promoter. Hum Mol Genet 20(22):4299-4310.
Han MK, Song EK, Guo Y, Ou X, Mantel C, Broxmeyer HE. 2008. SIRT1 regulates
apoptosis and Nanog expression in mouse embryonic stem cells by controlling
p53 subcellular localization. Cell Stem Cell 2(3):241-251.
Han YM, Kang YK, Koo DB, Lee KK. 2003. Nuclear reprogramming of cloned embryos
produced in vitro. Theriogenology 59(1):33-44.
41
Hansen RS, Stoger R, Wijmenga C, Stanek AM, Canfield TK, Luo P, Matarazzo MR,
D'Esposito M, Feil R, Gimelli G, Weemaes CM, Laird CD, Gartler SM. 2000.
Escape from gene silencing in ICF syndrome: evidence for advanced replication
time as a major determinant. Hum Mol Genet 9(18):2575-2587.
Hansis C, Barreto G, Maltry N, Niehrs C. 2004. Nuclear reprogramming of human
somatic cells by xenopus egg extract requires BRG1. Curr Biol 14(16):14751480.
Hao Y, Lai L, Mao J, Im GS, Bonk A, Prather RS. 2003. Apoptosis and in vitro
development of preimplantation porcine embryos derived in vitro or by nuclear
transfer. Biol Reprod 69(2):501-507.
Hardy K, Spanos S, Becker D, Iannelli P, Winston RM, Stark J. 2001. From cell death to
embryo arrest: mathematical models of human preimplantation embryo
development. Proc Natl Acad Sci USA 98(4):1655-1660.
Hark AT, Schoenherr CJ, Katz DJ, Ingram RS, Levorse JM, Tilghman SM. 2000. CTCF
mediates methylation-sensitive enhancer-blocking activity at the H19/Igf2 locus.
Nature 405(6785):486-489.
Hart AH, Hartley L, Parker K, Ibrahim M, Looijenga LH, Pauchnik M, Chow CW, Robb
L. 2005. The pluripotency homeobox gene NANOG is expressed in human germ
cell tumors. Cancer 104(10):2092-2098.
Hashimshony T, Zhang J, Keshet I, Bustin M, Cedar H. 2003. The role of DNA
methylation in setting up chromatin structure during development. Nat Genet
34(2):187-192.
Haupt Y, Rowan S, Shaulian E, Kazaz A, Vousden K, Oren M. 1997. p53 mediated
apoptosis in HeLa cells: transcription dependent and independent mechanisms.
Leukemia 11 Suppl 3:337-339.
Haupt Y, Rowan S, Shaulian E, Vousden KH, Oren M. 1995. Induction of apoptosis in
HeLa cells by trans-activation-deficient p53. Genes Dev 9(17):2170-2183.
Hendrich B, Bird A. 1998. Identification and characterization of a family of mammalian
methyl-CpG binding proteins. Mol Cell Biol 18(11):6538-6547.
Heyman Y, Chavatte-Palmer P, LeBourhis D, Camous S, Vignon X, Renard JP. 2002.
Frequency and occurrence of late-gestation losses from cattle cloned embryos.
Biol Reprod 66(1):6-13.
Hill JR, Burghardt RC, Jones K, Long CR, Looney CR, Shin T, Spencer TE, Thompson
JA, Winger QA, Westhusin ME. 2000a. Evidence for placental abnormality as the
42
major cause of mortality in first-trimester somatic cell cloned bovine fetuses. Biol
Reprod 63(6):1787-1794.
Hill JR, Roussel AJ, Cibelli JB, Edwards JF, Hooper NL, Miller MW, Thompson JA,
Looney CR, Westhusin ME, Robl JM, Stice SL. 1999. Clinical and pathologic
features of cloned transgenic calves and fetuses (13 case studies). Theriogenology
51(8):1451-1465.
Hill JR, Winger QA, Long CR, Looney CR, Thompson JA, Westhusin ME. 2000b.
Development rates of male bovine nuclear transfer embryos derived from adult
and fetal cells. Biol Reprod 62(5):1135-1140.
Hinrichs K, Choi YH, Love CC, Chung YG, Varner DD. 2006. Production of horse foals
via direct injection of roscovitine-treated donor cells and activation by injection of
sperm extract. Reproduction 131(6):1063-1072.
Hochedlinger K, Jaenisch R. 2002. Monoclonal mice generated by nuclear transfer from
mature B and T donor cells. Nature 415(6875):1035-1038.
Hochedlinger K, Rideout WM, Kyba M, Daley GQ, Blelloch R, Jaenisch R. 2004.
Nuclear transplantation, embryonic stem cells and the potential for cell therapy.
Hematol J 5 Suppl 3:S114-117.
Hochedlinger K, Yamada Y, Beard C, Jaenisch R. 2005. Ectopic expression of Oct-4
blocks progenitor-cell differentiation and causes dysplasia in epithelial tissues.
Cell 121(3):465-477.
Holmgren C, Kanduri C, Dell G, Ward A, Mukhopadhya R, Kanduri M, Lobanenkov V,
Ohlsson R. 2001. CpG methylation regulates the Igf2/H19 insulator. Curr Biol
11(14):1128-1130.
Hong H, Takahashi K, Ichisaka T, Aoi T, Kanagawa O, Nakagawa M, Okita K,
Yamanaka S. 2009. Suppression of induced pluripotent stem cell generation by
the p53-p21 pathway. Nature 460(7259):1132-1135.
Hosaka K, Ohi S, Ando A, Kobayashi M, Sato K. 2000. Cloned mice derived from
somatic cell nuclei. Hum Cell 13(4):237-242.
Hsieh CL. 1994. Dependence of transcriptional repression on CpG methylation density.
Mol Cell Biol 14(8):5487-5494.
Humpherys D, Eggan K, Akutsu H, Friedman A, Hochedlinger K, Yanagimachi R,
Lander ES, Golub TR, Jaenisch R. 2002. Abnormal gene expression in cloned
mice derived from embryonic stem cell and cumulus cell nuclei. Proc Natl Acad
Sci USA 99(20):12889-12894.
43
Humpherys D, Eggan K, Akutsu H, Hochedlinger K, Rideout WM, 3rd, Biniszkiewicz D,
Yanagimachi R, Jaenisch R. 2001. Epigenetic instability in ES cells and cloned
mice. Science 293(5527):95-97.
Hung MS, Karthikeyan N, Huang B, Koo HC, Kiger J, Shen CJ. 1999. Drosophila
proteins related to vertebrate DNA (5-cytosine) methyltransferases. Proc Natl
Acad Sci USA 96(21):11940-11945.
Hwang WS, Ryu YJ, Park JH, Park ES, Lee EG, Koo JM, Jeon HY, Lee BC, Kang SK,
Kim SJ, Ahn C, Hwang JH, Park KY, Cibelli JB, Moon SY. 2004. Evidence of a
pluripotent human embryonic stem cell line derived from a cloned blastocyst.
Science 303(5664):1669-1674.
Illingworth R, Kerr A, Desousa D, Jorgensen H, Ellis P, Stalker J, Jackson D, Clee C,
Plumb R, Rogers J, Humphray S, Cox T, Langford C, Bird A. 2008. A novel CpG
island set identifies tissue-specific methylation at developmental gene loci. PLoS
Biol 6(1):e22.
Illingworth RS, Bird AP. 2009. CpG islands--'a rough guide'. FEBS Lett 583(11):17131720.
Im GS, Seo JS, Hwang IS, Kim DH, Kim SW, Yang BC, Yang BS, Lai L, Prather RS.
2006. Development and apoptosis of pre-implantation porcine nuclear transfer
embryos activated with different combination of chemicals. Mol Reprod Dev
73(9):1094-1101.
Inoue K, Kohda T, Lee J, Ogonuki N, Mochida K, Noguchi Y, Tanemura K, KanekoIshino T, Ishino F, Ogura A. 2002. Faithful expression of imprinted genes in
cloned mice. Science 295(5553):297.
Inoue K, Wakao H, Ogonuki N, Miki H, Seino K, Nambu-Wakao R, Noda S, Miyoshi H,
Koseki H, Taniguchi M, Ogura A. 2005. Generation of cloned mice by direct
nuclear transfer from natural killer T cells. Curr Biol 15(12):1114-1118.
Jackson-Grusby L, Beard C, Possemato R, Tudor M, Fambrough D, Csankovszki G,
Dausman J, Lee P, Wilson C, Lander E, Jaenisch R. 2001. Loss of genomic
methylation causes p53-dependent apoptosis and epigenetic deregulation. Nat
Genet 27(1):31-39.
Jaenisch R. 1997. DNA methylation and imprinting: why bother? Trends Genet
13(8):323-329.
Jaenisch R. 2004. Human cloning - the science and ethics of nuclear transplantation. N
Engl J Med 351(27):2787-2791.
44
Jaenisch R, Bird A. 2003. Epigenetic regulation of gene expression: how the genome
integrates intrinsic and environmental signals. Nat Genet 33 Suppl:245-254.
Jaenisch R, Schnieke A, Harbers K. 1985. Treatment of mice with 5-azacytidine
efficiently activates silent retroviral genomes in different tissues. Proc Natl Acad
Sci USA 82(5):1451-1455.
Jang G, Park ES, Cho JK, Bhuiyan MM, Lee BC, Kang SK, Hwang WS. 2004.
Preimplantational embryo development and incidence of blastomere apoptosis in
bovine somatic cell nuclear transfer embryos reconstructed with long-term
cultured donor cells. Theriogenology 62(3-4):512-521.
Jimenez GS, Nister M, Stommel JM, Beeche M, Barcarse EA, Zhang XQ, O'Gorman S,
Wahl GM. 2000. A transactivation-deficient mouse model provides insights into
Trp53 regulation and function. Nat Genet 26(1):37-43.
Johnson TM, Hammond EM, Giaccia A, Attardi LD. 2005. The p53QS transactivationdeficient mutant shows stress-specific apoptotic activity and induces embryonic
lethality. Nat Genet 37(2):145-152.
Jones PL, Veenstra GJ, Wade PA, Vermaak D, Kass SU, Landsberger N, Strouboulis J,
Wolffe AP. 1998. Methylated DNA and MeCP2 recruit histone deacetylase to
repress transcription. Nat Genet 19(2):187-191.
Jouneau A, Renard JP. 2003. Reprogramming in nuclear transfer. Curr Opin Genet Dev
13(5):486-491.
Jouneau A, Zhou Q, Camus A, Brochard V, Maulny L, Collignon J, Renard JP. 2006.
Developmental abnormalities of NT mouse embryos appear early after
implantation. Development 133(8):1597-1607.
Kafri T, Ariel M, Brandeis M, Shemer R, Urven L, McCarrey J, Cedar H, Razin A. 1992.
Developmental pattern of gene-specific DNA methylation in the mouse embryo
and germ line. Genes Dev 6(5):705-714.
Kang YK, Koo DB, Park JS, Choi YH, Chung AS, Lee KK, Han YM. 2001. Aberrant
methylation of donor genome in cloned bovine embryos. Nat Genet 28(2):173177.
Kang YK, Lee KK, Han YM. 2003. Reprogramming DNA methylation in the
preimplantation stage: peeping with Dolly's eyes. Curr Opin Cell Biol 15(3):290295.
Kang YK, Park JS, Koo DB, Choi YH, Kim SU, Lee KK, Han YM. 2002. Limited
demethylation leaves mosaic-type methylation states in cloned bovine preimplantation embryos. EMBO J 21(5):1092-1100.
45
Kass SU, Landsberger N, Wolffe AP. 1997. DNA methylation directs a time-dependent
repression of transcription initiation. Curr Biol 7(3):157-165.
Kato Y, Tani T, Sotomaru Y, Kurokawa K, Kato J, Doguchi H, Yasue H, Tsunoda Y.
1998. Eight calves cloned from somatic cells of a single adult. Science
282(5396):2095-2098.
Kato Y, Tani T, Tsunoda Y. 2000. Cloning of calves from various somatic cell types of
male and female adult, newborn and fetal cows. J Reprod Fertil 120(2):231-237.
Kawamura T, Suzuki J, Wang YV, Menendez S, Morera LB, Raya A, Wahl GM, Izpisua
Belmonte JC. 2009. Linking the p53 tumour suppressor pathway to somatic cell
reprogramming. Nature 460(7259):1140-1144.
Keefer CL, Baldassarre H, Keyston R, Wang B, Bhatia B, Bilodeau AS, Zhou JF, Leduc
M, Downey BR, Lazaris A, Karatzas CN. 2001. Generation of dwarf goat (Capra
hircus) clones following nuclear transfer with transfected and nontransfected fetal
fibroblasts and in vitro-matured oocytes. Biol Reprod 64(3):849-856.
Keohane AM, O'Neill L P, Belyaev ND, Lavender JS, Turner BM. 1996. X-Inactivation
and histone H4 acetylation in embryonic stem cells. Dev Biol 180(2):618-630.
Kerr JF, Wyllie AH, Currie AR. 1972. Apoptosis: a basic biological phenomenon with
wide-ranging implications in tissue kinetics. Br J Cancer 26(4):239-257.
Kikyo N, Wolffe AP. 2000. Reprogramming nuclei: insights from cloning, nuclear
transfer and heterokaryons. J Cell Sci 113 ( Pt 1):11-20.
King TJ, Briggs R. 1955. Changes in the nuclei of differentiating gastrula cells, as
demonstrated by nuclear transplantation. Proc Natl Acad Sci USA 41(5):321-325.
Kishigami S, Hikichi T, Van Thuan N, Ohta H, Wakayama S, Bui HT, Mizutani E,
Wakayama T. 2006. Normal specification of the extraembryonic lineage after
somatic nuclear transfer. FEBS Lett 580(7):1801-1806.
Knijn HM, Gjorret JO, Vos PL, Hendriksen PJ, van der Weijden BC, Maddox-Hyttel P,
Dieleman SJ. 2003. Consequences of in vivo development and subsequent culture
on apoptosis, cell number, and blastocyst formation in bovine embryos. Biol
Reprod 69(4):1371-1378.
Kondo T, Bobek MP, Kuick R, Lamb B, Zhu X, Narayan A, Bourc'his D, ViegasPequignot E, Ehrlich M, Hanash SM. 2000. Whole-genome methylation scan in
ICF syndrome: hypomethylation of non-satellite DNA repeats D4Z4 and NBL2.
Hum Mol Genet 9(4):597-604.
46
Koo DB, Kang YK, Choi YH, Park JS, Han SK, Park IY, Kim SU, Lee KK, Son DS,
Chang WK, Han YM. 2000. In vitro development of reconstructed porcine
oocytes after somatic cell nuclear transfer. Biol Reprod 63(4):986-992.
Koralov SB, Novobrantseva TI, Hochedlinger K, Jaenisch R, Rajewsky K. 2005. Direct
in vivo VH to JH rearrangement violating the 12/23 rule. J Exp Med 201(3):341348.
Kruse JP, Gu W. 2009. Modes of p53 regulation. Cell 137(4):609-622.
Kubota C, Yamakuchi H, Todoroki J, Mizoshita K, Tabara N, Barber M, Yang X. 2000.
Six cloned calves produced from adult fibroblast cells after long-term culture.
Proc Natl Acad Sci USA 97(3):990-995.
Kumar BM, Jin HF, Kim JG, Ock SA, Hong Y, Balasubramanian S, Choe SY, Rho GJ.
2007. Differential gene expression patterns in porcine nuclear transfer embryos
reconstructed with fetal fibroblasts and mesenchymal stem cells. Dev Dyn
236(2):435-446.
Lai L, Kolber-Simonds D, Park KW, Cheong HT, Greenstein JL, Im GS, Samuel M,
Bonk A, Rieke A, Day BN, Murphy CN, Carter DB, Hawley RJ, Prather RS.
2002. Production of alpha-1,3-galactosyltransferase knockout pigs by nuclear
transfer cloning. Science 295(5557):1089-1092.
Landsverk HB, Hakelien AM, Kuntziger T, Robl JM, Skalhegg BS, Collas P. 2002.
Reprogrammed gene expression in a somatic cell-free extract. EMBO Rep
3(4):384-389.
Lanza RP, Cibelli JB, Blackwell C, Cristofalo VJ, Francis MK, Baerlocher GM, Mak J,
Schertzer M, Chavez EA, Sawyer N, Lansdorp PM, West MD. 2000a. Extension
of cell life-span and telomere length in animals cloned from senescent somatic
cells. Science 288(5466):665-669.
Lanza RP, Cibelli JB, Diaz F, Moraes CT, Farin PW, Farin CE, Hammer CJ, West MD,
Damiani P. 2000b. Cloning of an endangered species (Bos gaurus) using
interspecies nuclear transfer. Cloning 2(2):79-90.
Laurent L, Wong E, Li G, Huynh T, Tsirigos A, Ong CT, Low HM, Kin Sung KW,
Rigoutsos I, Loring J, Wei CL. 2010. Dynamic changes in the human methylome
during differentiation. Genome Res 20(3):320-331.
Lee BC, Kim MK, Jang G, Oh HJ, Yuda F, Kim HJ, Hossein MS, Kim JJ, Kang SK,
Schatten G, Hwang WS. 2005. Dogs cloned from adult somatic cells. Nature
436(7051):641.
47
Lee E, Lee SH, Kim S, Jeong YW, Kim JH, Koo OJ, Park SM, Hashem MA, Hossein
MS, Son HY, Lee CK, Hwang WS, Kang SK, Lee BC. 2006. Analysis of nuclear
reprogramming in cloned miniature pig embryos by expression of Oct-4 and Oct4 related genes. Biochem Biophys Res Commun 348(4):1419-1428.
Lee KY, Huang H, Ju B, Yang Z, Lin S. 2002. Cloned zebrafish by nuclear transfer from
long-term-cultured cells. Nat Biotechnol 20(8):795-799.
Leonhardt H, Page AW, Weier HU, Bestor TH. 1992. A targeting sequence directs DNA
methyltransferase to sites of DNA replication in mammalian nuclei. Cell
71(5):865-873.
Li A, Chandrakanthan V, Chami O, O'Neill C. 2007. Culture of zygotes increases TRP53
[corrected] expression in B6 mouse embryos, which reduces embryo viability.
Biol Reprod 76(3):362-367.
Li E, Beard C, Jaenisch R. 1993. Role for DNA methylation in genomic imprinting.
Nature 366(6453):362-365.
Li E, Bestor TH, Jaenisch R. 1992. Targeted mutation of the DNA methyltransferase
gene results in embryonic lethality. Cell 69(6):915-926.
Li J, Ishii T, Feinstein P, Mombaerts P. 2004. Odorant receptor gene choice is reset by
nuclear transfer from mouse olfactory sensory neurons. Nature 428(6981):393399.
Li Z, Sun X, Chen J, Liu X, Wisely SM, Zhou Q, Renard JP, Leno GH, Engelhardt JF.
2006. Cloned ferrets produced by somatic cell nuclear transfer. Dev Biol
293(2):439-448.
Lin JC, Jeong S, Liang G, Takai D, Fatemi M, Tsai YC, Egger G, Gal-Yam EN, Jones
PA. 2007. Role of nucleosomal occupancy in the epigenetic silencing of the
MLH1 CpG island. Cancer Cell 12(5):432-444.
Liu WM, Maraia RJ, Rubin CM, Schmid CW. 1994. Alu transcripts: cytoplasmic
localisation and regulation by DNA methylation. Nucleic Acids Res 22(6):10871095.
Lock LF, Takagi N, Martin GR. 1987. Methylation of the Hprt gene on the inactive X
occurs after chromosome inactivation. Cell 48(1):39-46.
Loi P, Ptak G, Barboni B, Fulka J, Jr., Cappai P, Clinton M. 2001. Genetic rescue of an
endangered mammal by cross-species nuclear transfer using post-mortem somatic
cells. Nat Biotechnol 19(10):962-964.
48
Lyko F, Ramsahoye BH, Jaenisch R. 2000. DNA methylation in Drosophila
melanogaster. Nature 408(6812):538-540.
Macleod D, Charlton J, Mullins J, Bird AP. 1994. Sp1 sites in the mouse aprt gene
promoter are required to prevent methylation of the CpG island. Genes Dev
8(19):2282-2292.
Mann MR, Chung YG, Nolen LD, Verona RI, Latham KE, Bartolomei MS. 2003.
Disruption of imprinted gene methylation and expression in cloned
preimplantation stage mouse embryos. Biol Reprod 69(3):902-914.
Marchenko ND, Zaika A, Moll UM. 2000. Death signal-induced localization of p53
protein to mitochondria. A potential role in apoptotic signaling. J Biol Chem
275(21):16202-16212.
Marin MC, Hsu B, Meyn RE, Donehower LA, el-Naggar AK, McDonnell TJ. 1994.
Evidence that p53 and bcl-2 are regulators of a common cell death pathway
important for in vivo lymphomagenesis. Oncogene 9(11):3107-3112.
Marion RM, Strati K, Li H, Murga M, Blanco R, Ortega S, Fernandez-Capetillo O,
Serrano M, Blasco MA. 2009. A p53-mediated DNA damage response limits
reprogramming to ensure iPS cell genomic integrity. Nature 460(7259):11491153.
Matzke M, Matzke AJ, Kooter JM. 2001. RNA: guiding gene silencing. Science
293(5532):1080-1083.
Mayer W, Niveleau A, Walter J, Fundele R, Haaf T. 2000. Demethylation of the zygotic
paternal genome. Nature 403(6769):501-502.
McCreath KJ, Howcroft J, Campbell KH, Colman A, Schnieke AE, Kind AJ. 2000.
Production of gene-targeted sheep by nuclear transfer from cultured somatic cells.
Nature 405(6790):1066-1069.
McGrath J, Solter D. 1983. Nuclear transplantation in the mouse embryo by microsurgery
and cell fusion. Science 220(4603):1300-1302.
McGrath J, Solter D. 1984. Inability of mouse blastomere nuclei transferred to enucleated
zygotes to support development in vitro. Science 226(4680):1317-1319.
McKeon C, Ohkubo H, Pastan I, de Crombrugghe B. 1982. Unusual methylation pattern
of the alpha 2 (l) collagen gene. Cell 29(1):203-210.
Meehan RR, Lewis JD, McKay S, Kleiner EL, Bird AP. 1989. Identification of a
mammalian protein that binds specifically to DNA containing methylated CpGs.
Cell 58(3):499-507.
49
Mihara M, Erster S, Zaika A, Petrenko O, Chittenden T, Pancoska P, Moll UM. 2003.
p53 has a direct apoptogenic role at the mitochondria. Mol Cell 11(3):577-590.
Miniou P, Jeanpierre M, Blanquet V, Sibella V, Bonneau D, Herbelin C, Fischer A,
Niveleau A, Viegas-Pequignot E. 1994. Abnormal methylation pattern in
constitutive and facultative (X inactive chromosome) heterochromatin of ICF
patients. Hum Mol Genet 3(12):2093-2102.
Mitalipov SM, Yeoman RR, Nusser KD, Wolf DP. 2002. Rhesus monkey embryos
produced by nuclear transfer from embryonic blastomeres or somatic cells. Biol
Reprod 66(5):1367-1373.
Miyashita T, Reed JC. 1995. Tumor suppressor p53 is a direct transcriptional activator of
the human bax gene. Cell 80(2):293-299.
Moarefi AH, Chedin F. 2011. ICF syndrome mutations cause a broad spectrum of
biochemical defects in DNMT3B-mediated de novo DNA methylation. J Mol
Biol 409(5):758-772.
Mohandas T, Sparkes RS, Shapiro LJ. 1981. Reactivation of an inactive human X
chromosome: evidence for X inactivation by DNA methylation. Science
211(4480):393-396.
Moll UM, Petrenko O. 2003. The MDM2-p53 interaction. Mol Cancer Res 1(14):10011008.
Monk M, Boubelik M, Lehnert S. 1987. Temporal and regional changes in DNA
methylation in the embryonic, extraembryonic and germ cell lineages during
mouse embryo development. Development 99(3):371-382.
Munsie MJ, Michalska AE, O'Brien CM, Trounson AO, Pera MF, Mountford PS. 2000.
Isolation of pluripotent embryonic stem cells from reprogrammed adult mouse
somatic cell nuclei. Curr Biol 10(16):989-992.
Nan X, Campoy FJ, Bird A. 1997. MeCP2 is a transcriptional repressor with abundant
binding sites in genomic chromatin. Cell 88(4):471-481.
Nan X, Meehan RR, Bird A. 1993. Dissection of the methyl-CpG binding domain from
the chromosomal protein MeCP2. Nucleic Acids Res 21(21):4886-4892.
Nan X, Ng HH, Johnson CA, Laherty CD, Turner BM, Eisenman RN, Bird A. 1998.
Transcriptional repression by the methyl-CpG-binding protein MeCP2 involves a
histone deacetylase complex. Nature 393(6683):386-389.
50
Ng HH, Zhang Y, Hendrich B, Johnson CA, Turner BM, Erdjument-Bromage H, Tempst
P, Reinberg D, Bird A. 1999. MBD2 is a transcriptional repressor belonging to
the MeCP1 histone deacetylase complex. Nat Genet 23(1):58-61.
Ng RK, Gurdon JB. 2008. Epigenetic memory of an active gene state depends on histone
H3.3 incorporation into chromatin in the absence of transcription. Nat Cell Biol
10(1):102-109.
Nichols J, Zevnik B, Anastassiadis K, Niwa H, Klewe-Nebenius D, Chambers I, Scholer
H, Smith A. 1998. Formation of pluripotent stem cells in the mammalian embryo
depends on the POU transcription factor Oct4. Cell 95(3):379-391.
Niculescu AB, 3rd, Chen X, Smeets M, Hengst L, Prives C, Reed SI. 1998. Effects of
p21(Cip1/Waf1) at both the G1/S and the G2/M cell cycle transitions: pRb is a
critical determinant in blocking DNA replication and in preventing
endoreduplication. Mol Cell Biol 18(1):629-643.
Niemann H, Tian XC, King WA, Lee RS. 2008. Epigenetic reprogramming in embryonic
and foetal development upon somatic cell nuclear transfer cloning. Reproduction
135(2):151-163.
Niwa H, Miyazaki J, Smith AG. 2000. Quantitative expression of Oct-3/4 defines
differentiation, dedifferentiation or self-renewal of ES cells. Nat Genet 24(4):372376.
Niwa O, Yokota Y, Ishida H, Sugahara T. 1983. Independent mechanisms involved in
suppression of the Moloney leukemia virus genome during differentiation of
murine teratocarcinoma cells. Cell 32(4):1105-1113.
Nolen LD, Gao S, Han Z, Mann MR, Gie Chung Y, Otte AP, Bartolomei MS, Latham
KE. 2005. X chromosome reactivation and regulation in cloned embryos. Dev
Biol 279(2):525-540.
Ogura A, Inoue K, Ogonuki N, Noguchi A, Takano K, Nagano R, Suzuki O, Lee J, Ishino
F, Matsuda J. 2000. Production of male cloned mice from fresh, cultured, and
cryopreserved immature Sertoli cells. Biol Reprod 62(6):1579-1584.
Ohgane J, Wakayama T, Kogo Y, Senda S, Hattori N, Tanaka S, Yanagimachi R, Shiota
K. 2001. DNA methylation variation in cloned mice. Genesis 30(2):45-50.
Ohm JE, McGarvey KM, Yu X, Cheng L, Schuebel KE, Cope L, Mohammad HP, Chen
W, Daniel VC, Yu W, Berman DM, Jenuwein T, Pruitt K, Sharkis SJ, Watkins
DN, Herman JG, Baylin SB. 2007. A stem cell-like chromatin pattern may
predispose tumor suppressor genes to DNA hypermethylation and heritable
silencing. Nat Genet 39(2):237-242.
51
Okano M, Bell DW, Haber DA, Li E. 1999. DNA methyltransferases Dnmt3a and
Dnmt3b are essential for de novo methylation and mammalian development. Cell
99(3):247-257.
Okano M, Xie S, Li E. 1998a. Cloning and characterization of a family of novel
mammalian DNA (cytosine-5) methyltransferases. Nat Genet 19(3):219-220.
Okano M, Xie S, Li E. 1998b. Dnmt2 is not required for de novo and maintenance
methylation of viral DNA in embryonic stem cells. Nucleic Acids Res
26(11):2536-2540.
Oltvai ZN, Milliman CL, Korsmeyer SJ. 1993. Bcl-2 heterodimerizes in vivo with a
conserved homolog, Bax, that accelerates programmed cell death. Cell 74(4):609619.
Onishi A, Iwamoto M, Akita T, Mikawa S, Takeda K, Awata T, Hanada H, Perry AC.
2000. Pig cloning by microinjection of fetal fibroblast nuclei. Science
289(5482):1188-1190.
Oswald J, Engemann S, Lane N, Mayer W, Olek A, Fundele R, Dean W, Reik W, Walter
J. 2000. Active demethylation of the paternal genome in the mouse zygote. Curr
Biol 10(8):475-478.
Pace MM, Augenstein ML, Betthauser JM, Childs LA, Eilertsen KJ, Enos JM, Forsberg
EJ, Golueke PJ, Graber DF, Kemper JC, Koppang RW, Lange G, Lesmeister TL,
Mallon KS, Mell GD, Misica PM, Pfister-Genskow M, Strelchenko NS, Voelker
GR, Watt SR, Bishop MD. 2002. Ontogeny of cloned cattle to lactation. Biol
Reprod 67(1):334-339.
Pannell D, Osborne CS, Yao S, Sukonnik T, Pasceri P, Karaiskakis A, Okano M, Li E,
Lipshitz HD, Ellis J. 2000. Retrovirus vector silencing is de novo methylase
independent and marked by a repressive histone code. EMBO J 19(21):58845894.
Pantazis P, Bollenbach T. 2012. Transcription factor kinetics and the emerging
asymmetry in the early mammalian embryo. Cell Cycle 11(11):2055-2058.
Paro R, Strutt H, Cavalli G. 1998. Heritable chromatin states induced by the Polycomb
and trithorax group genes. Novartis Found Symp 214:51-61; discussion 61-56,
104-113.
Pelengaris S, Khan M, Evan G. 2002. c-MYC: more than just a matter of life and death.
Nat Rev Cancer 2(10):764-776.
Phelps CJ, Koike C, Vaught TD, Boone J, Wells KD, Chen SH, Ball S, Specht SM,
Polejaeva IA, Monahan JA, Jobst PM, Sharma SB, Lamborn AE, Garst AS,
52
Moore M, Demetris AJ, Rudert WA, Bottino R, Bertera S, Trucco M, Starzl TE,
Dai Y, Ayares DL. 2003. Production of alpha 1,3-galactosyltransferase-deficient
pigs. Science 299(5605):411-414.
Pirrotta V. 1999. Polycomb silencing and the maintenance of stable chromatin states.
Results Probl Cell Differ 25:205-228.
Polejaeva IA, Chen SH, Vaught TD, Page RL, Mullins J, Ball S, Dai Y, Boone J, Walker
S, Ayares DL, Colman A, Campbell KH. 2000. Cloned pigs produced by nuclear
transfer from adult somatic cells. Nature 407(6800):86-90.
Pradhan S, Bacolla A, Wells RD, Roberts RJ. 1999. Recombinant human DNA (cytosine5) methyltransferase. I. Expression, purification, and comparison of de novo and
maintenance methylation. J Biol Chem 274(46):33002-33010.
Prather RS, Barnes FL, Sims MM, Robl JM, Eyestone WH, First NL. 1987. Nuclear
transplantation in the bovine embryo: assessment of donor nuclei and recipient
oocyte. Biol Reprod 37(4):859-866.
Prather RS, Sims MM, First NL. 1989. Nuclear transplantation in early pig embryos. Biol
Reprod 41(3):414-418.
Prendergast GC, Ziff EB. 1991. Methylation-sensitive sequence-specific DNA binding by
the c-Myc basic region. Science 251(4990):186-189.
Prokhortchouk A, Hendrich B, Jorgensen H, Ruzov A, Wilm M, Georgiev G, Bird A,
Prokhortchouk E. 2001. The p120 catenin partner Kaiso is a DNA methylationdependent transcriptional repressor. Genes Dev 15(13):1613-1618.
Reik W, Dean W, Walter J. 2001. Epigenetic reprogramming in mammalian
development. Science 293(5532):1089-1093.
Reik W, Walter J. 2001. Evolution of imprinting mechanisms: the battle of the sexes
begins in the zygote. Nat Genet 27(3):255-256.
Renard JP, Chastant S, Chesne P, Richard C, Marchal J, Cordonnier N, Chavatte P,
Vignon X. 1999. Lymphoid hypoplasia and somatic cloning. Lancet
353(9163):1489-1491.
Rhee I, Jair KW, Yen RW, Lengauer C, Herman JG, Kinzler KW, Vogelstein B, Baylin
SB, Schuebel KE. 2000. CpG methylation is maintained in human cancer cells
lacking DNMT1. Nature 404(6781):1003-1007.
Rideout WM, 3rd, Hochedlinger K, Kyba M, Daley GQ, Jaenisch R. 2002. Correction of
a genetic defect by nuclear transplantation and combined cell and gene therapy.
Cell 109(1):17-27.
53
Rideout WM, 3rd, Wakayama T, Wutz A, Eggan K, Jackson-Grusby L, Dausman J,
Yanagimachi R, Jaenisch R. 2000. Generation of mice from wild-type and
targeted ES cells by nuclear cloning. Nat Genet 24(2):109-110.
Riley T, Sontag E, Chen P, Levine A. 2008. Transcriptional control of human p53regulated genes. Nat Rev Mol Cell Biol 9(5):402-412.
Roemer I, Reik W, Dean W, Klose J. 1997. Epigenetic inheritance in the mouse. Curr
Biol 7(4):277-280.
Rogers CS, Hao Y, Rokhlina T, Samuel M, Stoltz DA, Li Y, Petroff E, Vermeer DW,
Kabel AC, Yan Z, Spate L, Wax D, Murphy CN, Rieke A, Whitworth K, Linville
ML, Korte SW, Engelhardt JF, Welsh MJ, Prather RS. 2008. Production of
CFTR-null and CFTR-DeltaF508 heterozygous pigs by adeno-associated virusmediated gene targeting and somatic cell nuclear transfer. J Clin Invest
118(4):1571-1577.
Rowland BD, Bernards R, Peeper DS. 2005. The KLF4 tumour suppressor is a
transcriptional repressor of p53 that acts as a context-dependent oncogene. Nat
Cell Biol 7(11):1074-1082.
Rowland BD, Peeper DS. 2006. KLF4, p21 and context-dependent opposing forces in
cancer. Nat Rev Cancer 6(1):11-23.
Sado T, Fenner MH, Tan SS, Tam P, Shioda T, Li E. 2000. X inactivation in the mouse
embryo deficient for Dnmt1: distinct effect of hypomethylation on imprinted and
random X inactivation. Dev Biol 225(2):294-303.
Sanford JP, Clark HJ, Chapman VM, Rossant J. 1987. Differences in DNA methylation
during oogenesis and spermatogenesis and their persistence during early
embryogenesis in the mouse. Genes Dev 1(10):1039-1046.
Sansinena MJ, Hylan D, Hebert K, Denniston RS, Godke RA. 2005. Banteng (Bos
javanicus) embryos and pregnancies produced by interspecies nuclear transfer.
Theriogenology 63(4):1081-1091.
Sansome C, Zaika A, Marchenko ND, Moll UM. 2001. Hypoxia death stimulus induces
translocation of p53 protein to mitochondria. Detection by immunofluorescence
on whole cells. FEBS Lett 488(3):110-115.
Santos F, Hendrich B, Reik W, Dean W. 2002. Dynamic reprogramming of DNA
methylation in the early mouse embryo. Dev Biol 241(1):172-182.
Santos F, Zakhartchenko V, Stojkovic M, Peters A, Jenuwein T, Wolf E, Reik W, Dean
W. 2003. Epigenetic marking correlates with developmental potential in cloned
bovine preimplantation embryos. Curr Biol 13(13):1116-1121.
54
Schnieke AE, Kind AJ, Ritchie WA, Mycock K, Scott AR, Ritchie M, Wilmut I, Colman
A, Campbell KH. 1997. Human factor IX transgenic sheep produced by transfer
of nuclei from transfected fetal fibroblasts. Science 278(5346):2130-2133.
Schwartz S, Meshorer E, Ast G. 2009. Chromatin organization marks exon-intron
structure. Nat Struct Mol Biol 16(9):990-995.
Senda S, Wakayama T, Yamazaki Y, Ohgane J, Hattori N, Tanaka S, Yanagimachi R,
Shiota K. 2004. Skewed X-inactivation in cloned mice. Biochem Biophys Res
Commun 321(1):38-44.
Shin T, Kraemer D, Pryor J, Liu L, Rugila J, Howe L, Buck S, Murphy K, Lyons L,
Westhusin M. 2002. A cat cloned by nuclear transplantation. Nature
415(6874):859.
Shiota K. 2004. DNA methylation profiles of CpG islands for cellular differentiation and
development in mammals. Cytogenet Genome Res 105(2-4):325-334.
Shukla S, Kavak E, Gregory M, Imashimizu M, Shutinoski B, Kashlev M, Oberdoerffer
P, Sandberg R, Oberdoerffer S. 2011. CTCF-promoted RNA polymerase II
pausing links DNA methylation to splicing. Nature 479(7371):74-79.
Silva J, Nichols J, Theunissen TW, Guo G, van Oosten AL, Barrandon O, Wray J,
Yamanaka S, Chambers I, Smith A. 2009. Nanog is the gateway to the pluripotent
ground state. Cell 138(4):722-737.
Simonsson S, Gurdon J. 2004. DNA demethylation is necessary for the epigenetic
reprogramming of somatic cell nuclei. Nat Cell Biol 6(10):984-990.
Sims M, First NL. 1994. Production of calves by transfer of nuclei from cultured inner
cell mass cells. Proc Natl Acad Sci USA 91(13):6143-6147.
Smith SL, Everts RE, Tian XC, Du F, Sung LY, Rodriguez-Zas SL, Jeong BS, Renard
JP, Lewin HA, Yang X. 2005. Global gene expression profiles reveal significant
nuclear reprogramming by the blastocyst stage after cloning. Proc Natl Acad Sci
USA 102(49):17582-17587.
Smith ZD, Chan MM, Mikkelsen TS, Gu H, Gnirke A, Regev A, Meissner A. 2012. A
unique regulatory phase of DNA methylation in the early mammalian embryo.
Nature 484(7394):339-344.
Solter D. 2000. Mammalian cloning: advances and limitations. Nat Rev Genet 1(3):199207.
Spemann H. 1938. Embryonic Development and Induction. New Haven: Yale University
Press.
55
Stancheva I, El-Maarri O, Walter J, Niveleau A, Meehan RR. 2002. DNA methylation at
promoter regions regulates the timing of gene activation in Xenopus laevis
embryos. Dev Biol 243(1):155-165.
Stancheva I, Meehan RR. 2000. Transient depletion of xDnmt1 leads to premature gene
activation in Xenopus embryos. Genes Dev 14(3):313-327.
Stewart CL, Stuhlmann H, Jahner D, Jaenisch R. 1982. De novo methylation, expression,
and infectivity of retroviral genomes introduced into embryonal carcinoma cells.
Proc Natl Acad Sci USA 79(13):4098-4102.
Stoger R, Kubicka P, Liu CG, Kafri T, Razin A, Cedar H, Barlow DP. 1993. Maternalspecific methylation of the imprinted mouse Igf2r locus identifies the expressed
locus as carrying the imprinting signal. Cell 73(1):61-71.
Straussman R, Nejman D, Roberts D, Steinfeld I, Blum B, Benvenisty N, Simon I,
Yakhini Z, Cedar H. 2009. Developmental programming of CpG island
methylation profiles in the human genome. Nat Struct Mol Biol 16(5):564-571.
Suteevun T, Parnpai R, Smith SL, Chang CC, Muenthaisong S, Tian XC. 2006.
Epigenetic characteristics of cloned and in vitro-fertilized swamp buffalo
(Bubalus bubalis) embryos. J Anim Sci 84(8):2065-2071.
Szabo P, Tang SH, Rentsendorj A, Pfeifer GP, Mann JR. 2000. Maternal-specific
footprints at putative CTCF sites in the H19 imprinting control region give
evidence for insulator function. Curr Biol 10(10):607-610.
Tachibana M, Amato P, Sparman M, Gutierrez NM, Tippner-Hedges R, Ma H, Kang E,
Fulati A, Lee HS, Sritanaudomchai H, Masterson K, Larson J, Eaton D, SadlerFredd K, Battaglia D, Lee D, Wu D, Jensen J, Patton P, Gokhale S, Stouffer RL,
Wolf D, Mitalipov S. 2013. Human embryonic stem cells derived by somatic cell
nuclear transfer. Cell 153(6):1228-1238.
Tada M, Tada T, Lefebvre L, Barton SC, Surani MA. 1997. Embryonic germ cells induce
epigenetic reprogramming of somatic nucleus in hybrid cells. EMBO J
16(21):6510-6520.
Tada T, Tada M, Hilton K, Barton SC, Sado T, Takagi N, Surani MA. 1998. Epigenotype
switching of imprintable loci in embryonic germ cells. Dev Genes Evol
207(8):551-561.
Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, Yamanaka S.
2007. Induction of pluripotent stem cells from adult human fibroblasts by defined
factors. Cell 131(5):861-872.
56
Tamaru H, Selker EU. 2001. A histone H3 methyltransferase controls DNA methylation
in Neurospora crassa. Nature 414(6861):277-283.
Tamashiro KL, Wakayama T, Akutsu H, Yamazaki Y, Lachey JL, Wortman MD, Seeley
RJ, D'Alessio DA, Woods SC, Yanagimachi R, Sakai RR. 2002. Cloned mice
have an obese phenotype not transmitted to their offspring. Nat Med 8(3):262267.
Tanaka S, Oda M, Toyoshima Y, Wakayama T, Tanaka M, Yoshida N, Hattori N,
Ohgane J, Yanagimachi R, Shiota K. 2001. Placentomegaly in cloned mouse
concepti caused by expansion of the spongiotrophoblast layer. Biol Reprod
65(6):1813-1821.
Terashita Y, Sugimura S, Kudo Y, Amano R, Hiradate Y, Sato E. 2011. Improving the
quality of miniature pig somatic cell nuclear transfer blastocysts: aggregation of
SCNT embryos at the four-cell stage. Reprod Domest Anim 46(2):189-196.
Thibault C. 2003. Recent data on the development of cloned embryos derived from
reconstructed eggs with adult cells. Reprod Nutr Dev 43(4):303-324.
Tilghman SM. 1999. The sins of the fathers and mothers: genomic imprinting in
mammalian development. Cell 96(2):185-193.
Tomita Y, Marchenko N, Erster S, Nemajerova A, Dehner A, Klein C, Pan H, Kessler H,
Pancoska P, Moll UM. 2006. WT p53, but not tumor-derived mutants, bind to
Bcl2 via the DNA binding domain and induce mitochondrial permeabilization. J
Biol Chem 281(13):8600-8606.
Tong GQ, Heng BC, Ng SC. 2007. Cumulus-specific genes are transcriptionally silent
following somatic cell nuclear transfer in a mouse model. J Zhejiang Univ Sci B
8(8):533-539.
Tong GQ, Heng BC, Tan LG, Ng SC. 2006. Aberrant profile of gene expression in
cloned mouse embryos derived from donor cumulus nuclei. Cell Tissue Res
325(2):231-243.
Tucker KL, Beard C, Dausmann J, Jackson-Grusby L, Laird PW, Lei H, Li E, Jaenisch R.
1996. Germ-line passage is required for establishment of methylation and
expression patterns of imprinted but not of nonimprinted genes. Genes Dev
10(8):1008-1020.
Turker MS. 1999. The establishment and maintenance of DNA methylation patterns in
mouse somatic cells. Semin Cancer Biol 9(5):329-337.
Tweedie S, Ng HH, Barlow AL, Turner BM, Hendrich B, Bird A. 1999. Vestiges of a
DNA methylation system in Drosophila melanogaster? Nat Genet 23(4):389-390.
57
Utikal J, Polo JM, Stadtfeld M, Maherali N, Kulalert W, Walsh RM, Khalil A, Rheinwald
JG, Hochedlinger K. 2009. Immortalization eliminates a roadblock during cellular
reprogramming into iPS cells. Nature 460(7259):1145-1148.
van Soom A, Ysebaert MT, de Kruif A. 1997. Relationship between timing of
development, morula morphology, and cell allocation to inner cell mass and
trophectoderm in in vitro-produced bovine embryos. Mol Reprod Dev 47(1):4756.
Vassena R, Han Z, Gao S, Baldwin DA, Schultz RM, Latham KE. 2007. Tough
beginnings: alterations in the transcriptome of cloned embryos during the first two
cell cycles. Dev Biol 304(1):75-89.
Venolia L, Gartler SM. 1983. Comparison of transformation efficiency of human active
and inactive X-chromosomal DNA. Nature 302(5903):82-83.
Venolia L, Gartler SM, Wassman ER, Yen P, Mohandas T, Shapiro LJ. 1982.
Transformation with DNA from 5-azacytidine-reactivated X chromosomes. Proc
Natl Acad Sci USA 79(7):2352-2354.
Vousden KH, Lu X. 2002. Live or let die: the cell's response to p53. Nat Rev Cancer
2(8):594-604.
Wade PA, Gegonne A, Jones PL, Ballestar E, Aubry F, Wolffe AP. 1999. Mi-2 complex
couples DNA methylation to chromatin remodelling and histone deacetylation.
Nat Genet 23(1):62-66.
Wade PA, Wolffe AP. 2001. ReCoGnizing methylated DNA. Nat Struct Biol 8(7):575577.
Wagner AJ, Kokontis JM, Hay N. 1994. Myc-mediated apoptosis requires wild-type p53
in a manner independent of cell cycle arrest and the ability of p53 to induce
p21waf1/cip1. Genes Dev 8(23):2817-2830.
Wakayama S, Ohta H, Kishigami S, Thuan NV, Hikichi T, Mizutani E, Miyake M,
Wakayama T. 2005. Establishment of male and female nuclear transfer embryonic
stem cell lines from different mouse strains and tissues. Biol Reprod 72(4):932936.
Wakayama T. 2003. Cloned mice and embryonic stem cell lines generated from adult
somatic cells by nuclear transfer. Oncol Res 13(6-10):309-314.
Wakayama T, Perry AC, Zuccotti M, Johnson KR, Yanagimachi R. 1998. Full-term
development of mice from enucleated oocytes injected with cumulus cell nuclei.
Nature 394(6691):369-374.
58
Wakayama T, Rodriguez I, Perry AC, Yanagimachi R, Mombaerts P. 1999. Mice cloned
from embryonic stem cells. Proc Natl Acad Sci USA 96(26):14984-14989.
Wakayama T, Tateno H, Mombaerts P, Yanagimachi R. 2000. Nuclear transfer into
mouse zygotes. Nat Genet 24(2):108-109.
Wakayama T, Yanagimachi R. 1999. Cloning of male mice from adult tail-tip cells. Nat
Genet 22(2):127-128.
Wakisaka N, Inoue K, Ogonuki N, Miki H, Sekita Y, Hanaki K, Akatsuka A, KanekoIshino T, Ishino F, Ogura A. 2008. Ultrastructure of placental hyperplasia in mice:
comparison of placental phenotypes with three different etiologies. Placenta
29(8):753-759.
Wakisaka-Saito N, Kohda T, Inoue K, Ogonuki N, Miki H, Hikichi T, Mizutani E,
Wakayama T, Kaneko-Ishino T, Ogura A, Ishino F. 2006. Chorioallantoic
placenta defects in cloned mice. Biochem Biophys Res Commun 349(1):106-114.
Walsh CP, Chaillet JR, Bestor TH. 1998. Transcription of IAP endogenous retroviruses is
constrained by cytosine methylation. Nat Genet 20(2):116-117.
Wang J, Mager J, Chen Y, Schneider E, Cross JC, Nagy A, Magnuson T. 2001. Imprinted
X inactivation maintained by a mouse Polycomb group gene. Nat Genet
28(4):371-375.
Wang Y, Szekely L, Okan I, Klein G, Wiman KG. 1993. Wild-type p53-triggered
apoptosis is inhibited by bcl-2 in a v-myc-induced T-cell lymphoma line.
Oncogene 8(12):3427-3431.
Wani NA, Wernery U, Hassan FA, Wernery R, Skidmore JA. 2010. Production of the
first cloned camel by somatic cell nuclear transfer. Biol Reprod 82(2):373-379.
Wareham KA, Lyon MF, Glenister PH, Williams ED. 1987. Age related reactivation of
an X-linked gene. Nature 327(6124):725-727.
Wassenegger M, Heimes S, Riedel L, Sanger HL. 1994. RNA-directed de novo
methylation of genomic sequences in plants. Cell 76(3):567-576.
Wei MC, Zong WX, Cheng EH, Lindsten T, Panoutsakopoulou V, Ross AJ, Roth KA,
MacGregor GR, Thompson CB, Korsmeyer SJ. 2001. Proapoptotic BAX and
BAK: a requisite gateway to mitochondrial dysfunction and death. Science
292(5517):727-730.
Wells DN, Misica PM, Tervit HR. 1999. Production of cloned calves following nuclear
transfer with cultured adult mural granulosa cells. Biol Reprod 60(4):996-1005.
59
West JA, Viswanathan SR, Yabuuchi A, Cunniff K, Takeuchi A, Park IH, Sero JE, Zhu
H, Perez-Atayde A, Frazier AL, Surani MA, Daley GQ. 2009. A role for Lin28 in
primordial germ-cell development and germ-cell malignancy. Nature
460(7257):909-913.
Widschwendter M, Fiegl H, Egle D, Mueller-Holzner E, Spizzo G, Marth C,
Weisenberger DJ, Campan M, Young J, Jacobs I, Laird PW. 2007. Epigenetic
stem cell signature in cancer. Nat Genet 39(2):157-158.
Willadsen SM. 1986. Nuclear transplantation in sheep embryos. Nature 320(6057):63-65.
Willadsen SM, Janzen RE, McAlister RJ, Shea BF, Hamilton G, McDermand D. 1991.
The viability of late morulae and blastocysts produced by nuclear transplantation
in cattle. Theriogenology 35(1):161–170.
Wilmut I, Schnieke AE, McWhir J, Kind AJ, Campbell KH. 1997. Viable offspring
derived from fetal and adult mammalian cells. Nature 385(6619):810-813.
Wilmut I, Young L, Campbell KH. 1998. Embryonic and somatic cell cloning. Reprod
Fertil Dev 10(7-8):639-643.
Woodcock DM, Lawler CB, Linsenmeyer ME, Doherty JP, Warren WD. 1997.
Asymmetric methylation in the hypermethylated CpG promoter region of the
human L1 retrotransposon. J Biol Chem 272(12):7810-7816.
Woods GL, White KL, Vanderwall DK, Li GP, Aston KI, Bunch TD, Meerdo LN, Pate
BJ. 2003. A mule cloned from fetal cells by nuclear transfer. Science
301(5636):1063.
Wrenzycki C, Wells D, Herrmann D, Miller A, Oliver J, Tervit R, Niemann H. 2001.
Nuclear transfer protocol affects messenger RNA expression patterns in cloned
bovine blastocysts. Biol Reprod 65(1):309-317.
Wutz A, Jaenisch R. 2000. A shift from reversible to irreversible X inactivation is
triggered during ES cell differentiation. Mol Cell 5(4):695-705.
Wuu YD, Pampfer S, Becquet P, Vanderheyden I, Lee KH, De Hertogh R. 1999. Tumor
necrosis factor alpha decreases the viability of mouse blastocysts in vitro and in
vivo. Biol Reprod 60(2):479-483.
Xue F, Tian XC, Du F, Kubota C, Taneja M, Dinnyes A, Dai Y, Levine H, Pereira LV,
Yang X. 2002. Aberrant patterns of X chromosome inactivation in bovine clones.
Nat Genet 31(2):216-220.
60
Yan Y, Shay JW, Wright WE, Mumby MC. 1997. Inhibition of protein phosphatase
activity induces p53-dependent apoptosis in the absence of p53 transactivation. J
Biol Chem 272(24):15220-15226.
Yang X, Smith SL, Tian XC, Lewin HA, Renard JP, Wakayama T. 2007. Nuclear
reprogramming of cloned embryos and its implications for therapeutic cloning.
Nat Genet 39(3):295-302.
Yin XM, Oltvai ZN, Korsmeyer SJ. 1994. BH1 and BH2 domains of Bcl-2 are required
for inhibition of apoptosis and heterodimerization with Bax. Nature
369(6478):321-323.
Yoder JA, Walsh CP, Bestor TH. 1997. Cytosine methylation and the ecology of
intragenomic parasites. Trends Genet 13(8):335-340.
You JS, Kelly TK, De Carvalho DD, Taberlay PC, Liang G, Jones PA. 2011. OCT4
establishes and maintains nucleosome-depleted regions that provide additional
layers of epigenetic regulation of its target genes. Proc Natl Acad Sci USA
108(35):14497-14502.
Young LE, Sinclair KD, Wilmut I. 1998. Large offspring syndrome in cattle and sheep.
Rev Reprod 3(3):155-163.
Yu J, Vodyanik MA, Smuga-Otto K, Antosiewicz-Bourget J, Frane JL, Tian S, Nie J,
Jonsdottir GA, Ruotti V, Stewart R, Slukvin, II, Thomson JA. 2007. Induced
pluripotent stem cell lines derived from human somatic cells. Science
318(5858):1917-1920.
Zakhartchenko V, Mueller S, Alberio R, Schernthaner W, Stojkovic M, Wenigerkind H,
Wanke R, Lassnig C, Mueller M, Wolf E, Brem G. 2001. Nuclear transfer in
cattle with non-transfected and transfected fetal or cloned transgenic fetal and
postnatal fibroblasts. Mol Reprod Dev 60(3):362-369.
Zhang S, Kubota C, Yang L, Zhang Y, Page R, O'Neill M, Yang X, Tian XC. 2004.
Genomic imprinting of H19 in naturally reproduced and cloned cattle. Biol
Reprod 71(5):1540-1544.
Zhang W, Geiman DE, Shields JM, Dang DT, Mahatan CS, Kaestner KH, Biggs JR,
Kraft AS, Yang VW. 2000. The gut-enriched Kruppel-like factor (Kruppel-like
factor 4) mediates the transactivating effect of p53 on the p21WAF1/Cip1
promoter. J Biol Chem 275(24):18391-18398.
Zhao Y, Yin X, Qin H, Zhu F, Liu H, Yang W, Zhang Q, Xiang C, Hou P, Song Z, Liu
Y, Yong J, Zhang P, Cai J, Liu M, Li H, Li Y, Qu X, Cui K, Zhang W, Xiang T,
Wu Y, Liu C, Yu C, Yuan K, Lou J, Ding M, Deng H. 2008. Two supporting
61
factors greatly improve the efficiency of human iPSC generation. Cell Stem Cell
3(5):475-479.
Zhou Q, Renard JP, Le Friec G, Brochard V, Beaujean N, Cherifi Y, Fraichard A, Cozzi
J. 2003. Generation of fertile cloned rats by regulating oocyte activation. Science
302(5648):1179.
62
CHAPTER 2
DNA METHYLATION REPROGRAMMING OF THE POU5F1, NANOG, SOX2 AND
KLF4 GENES FOLLOWING SOMATIC CELL NUCLEAR TRANSFER
Abstract
Aberrant methylation patterns have been implicated as a contributing factor in the
poor development of embryos following somatic cell nuclear transfer (SCNT). Embryos
produced via SCNT that are unable to epigenetically reprogram somatic cell-specific
methylation patterns are at a developmental disadvantage, as the donor cell methylation
signature retained within cloned embryos reinforces inappropriate gene silencing. Using
bisulfite sequencing, we compared multiple methylation sites in the POU5F1 (OCT4),
NANOG, SOX2, KLF4, C-MYC, and LIN28 genes within bovine IVF and SCNT
blastocyst embryos and donor cells. Prior to implantation, the methylation patterns of
NANOG and SOX2 in SCNT embryos are more similar to patterns observed in somatic
cells than patterns seen in IVF blastocysts. We also observed that methylation of
POU5F1 is lower in SCNT blastocysts, suggesting that reestablishment of methylation of
this gene prior to implantation of SCNT embryos is delayed. Our results demonstrate
that methylation patterns vary among developmental genes in SCNT embryos, with
reprogramming at some CpG sites and incomplete reprogramming at others.
Introduction
Since the cloning of the first animal from somatic cells, the technique of somatic
cell nuclear transfer (SCNT) has had limited success in embryo development and
63
production of viable offspring (Wilmut et al., 1997). Despite intense research, the
benefits of cloned offspring remain limited due to the failed development of embryos and
fetuses generated via SCNT (Wells, 2005). Following SCNT, many cloned neonates die
shortly after birth and often exhibit phenotypic abnormalities (Hill et al., 2000a; Farin et
al., 2004; Li et al., 2005). SCNT-derived organisms face challenges at each stage of
development. The low success rate of the SCNT procedure appears to be due in large
part to incomplete embryonic reprogramming of the somatic genome (Wilmut et al.,
1998; Colman, 1999; Kikyo and Wolffe, 2000; Solter, 2000; Dean et al., 2001;
Humpherys et al., 2001; Reik and Walter, 2001; Niemann et al., 2008).
Several phenotypic abnormalities have been observed in clones, including placental
deficiency, fetal loss, respiratory problems, and cardiovascular failure (Hill et al., 1999;
Hill et al., 2000b; De Sousa et al., 2001; Chavatte-Palmer et al., 2002; Yang et al., 2007).
Neonates are predisposed to infection and weakness (Zakhartchenko et al., 2001), as well
as a condition referred to as large offspring syndrome, which is characterized by
abnormally large body and organ size (Young et al., 1998; Farin et al., 2004). Large
offspring syndrome shares many similarities with human epigenetic disorders, such as
Beckwith-Wiedman syndrome (DeBaun et al., 2003; Everts et al., 2008). In addition,
altered levels of gene expression have been observed in SCNT embryos (Sebastiano et
al., 2005; Aston et al., 2010). These abnormalities are likely due to the failure of the
SCNT embryo to appropriately reprogram the epigenome of the donor cell nucleus to an
embryonic state.
64
DNA methylation in mammals involves the covalent addition of a methyl group
to the cytosine within cytosine-guanine (CpG) dinucleotides. DNA methylation is
responsible for the inactivation of transposons within the genome (Yoder et al., 1997), Xchromosome inactivation (Mohandas et al., 1981; Csankovszki et al., 2001),
establishment and silencing of imprinted genes (Bartolomei et al., 1993; Li et al., 1993;
Stoger et al., 1993), and maintenance of chromosome stability (Chen et al., 1998).
Additionally, DNA methylation plays a role in the silencing of gene transcription
(Jaenisch and Bird, 2003; Shiota, 2004) through one of two mechanisms: 1) the presence
of methyl groups in DNA directly blocks the binding of transcription factors (Watt and
Molloy, 1988; Hendrich and Bird, 1998; Prokhortchouk et al., 2001) and 2) methylation
of DNA targets the region for chromatin silencing that prevents gene transcription (Jones
et al., 1998; Wade et al., 1999; Zhang et al., 1999; Sarraf and Stancheva, 2004).
Although any CpG site within the mammalian genome can incur methylation (Bird,
2002), CpG dinucleotides often cluster within CpG dense regions known as CpG islands
(Illingworth and Bird, 2009). These islands are commonly found at the 5’-end of genes,
have a high density of CpG sites (about one site per 10 bp), and are associated with
transcriptional start sites (Antequera and Bird, 1993).
Abnormal levels of DNA methylation present a significant impediment to animal
cloning. Cloned bovine embryos have higher levels of global methylation than in vivo or
in vitro fertilized (IVF) embryos (Kikyo and Wolffe, 2000; Kang et al., 2001).
Additionally, specific genes have been shown to undergo gradual loss of methylation in
cloned embryos throughout embryo development (Yamazaki et al., 2006), indicating that
65
methylation patterns of somatic cells are not correctly reprogrammed via the same
mechanism undergone by in vivo and IVF embryos.
The discovery of transcription factors capable of inducing pluripotency in
differentiated cells may provide insights into the challenges of SCNT development
(Takahashi et al., 2007; Yu et al., 2007). Pluripotency-inducing transcription factors
include POU5F1 (OCT-4) (Pesce and Scholer, 2001), NANOG (Chambers et al., 2003;
Mitsui et al., 2003), SOX2 (Avilion et al., 2003), KLF4 (Evans and Liu, 2008), C-MYC
(Adhikary and Eilers, 2005), and LIN28 (Darr and Benvenisty, 2009). When expressed
ectopically in various combinations, these pluripotency genes are able to fully reprogram
somatic cells (Takahashi et al., 2007; Yu et al., 2007). Additionally, POU5F1, SOX2,
KLF4, C-MYC, and NANOG expression is essential for the creation of induced
pluripotent stem cells in cattle (Sumer et al., 2011). Either the absence or irregular
expression of these factors likely impedes successful development following SCNT.
To determine whether or not DNA methylation patterns in pre-implantation stage
SCNT embryos are appropriate for embryonic development, we characterized DNA
methylation patterns of POU5F1, NANOG, SOX2, KLF4, C-MYC, and LIN28 genes in
bovine SCNT and IVF blastocysts and compared them to patterns observed in the donor
cell line. We determined that the methylation patterns of SCNT embryos are more
similar to those of the donor cell for the NANOG and SOX2 genes. Additionally we
determined that the methylation patterns of CpG sites in POU5F1 were abnormally
regulated in SCNT embryos. The aberrant epigenetic patterns may lead to irregularities
in gene expression patterns and restrict reprogramming in SCNT embryos.
66
Materials and Methods
Donor Cell Culture
Bovine fibroblasts were collected from skin samples and cultured in DMEM F12
(Thermo Scientific HyClone Laboratories, Logan, UT) supplemented with 15% fetal
bovine serum (FBS) (Thermo Scientific HyClone Laboratories), 100 U/ml penicillin, and
100 mg/ml streptomycin. Cells were seeded in T25 culture flasks and cultured at 37°C
with 5% CO2. Prior to SCNT cultured cells were treated with 0.25% trypsin and
resuspended in phosphate buffered saline (PBS).
Oocyte Maturation
Cumulus oocyte complexes were aspirated from 3 to 8 mm follicles from ovaries
collected at a local abattoir (E.A. Miller, Hyrum, UT). Cumulus oocyte complexes were
cultured 18 to 22 hr in TCM 199 maturation medium (Thermo Scientific HyClone
Laboratories) containing 10% FBS, 0.05 mg/ml follicle stimulating hormone (FSH), 5
mg/ml luteinizing hormone (LH), 100 U/ml penicillin, and 100 mg/ml streptomycin.
In Vitro Fertilization
Following 18 to 22 hr of oocyte maturation, cryopreserved bovine semen
(Hoffman AI, Logan, UT) was thawed and live sperm were collected via centrifugation
with a 45%/90% percoll gradient. Live sperm were suspended in Tyrode’s albumin
lactate pyruvate (TALP) to be used for oocyte fertilization. Twenty-four hours following
fertilization, cumulus oocyte complexes were vortexed in PBS with 0.32 mM sodium
pyruvate, 5.55 mM glucose, 3 mg/ml BSA, and 10 mg/ml hyaluronidase to remove
67
cumulus cells from the oocyte. Oocytes were subsequently washed with PBS and
cultured in CR2 medium atop a layer of cumulus cells.
SCNT Production
Cumulus cells were removed following maturation as described above.
Metaphase II oocytes with an extruded first polar body were selected as recipient
cytoplasts. Following enucleation, removal of the metaphase plate and polar body, a
single donor cell was placed in the perivitelline space. Oocyte and donor cell fusion was
carried out with two electric DC pulses of 2.2 kV/cm for 25 µsec applied to embryos
placed in a fusion chamber suspended in mannitol fusion medium (Wells et al., 1999).
Following fusion, embryos were placed in CR2 medium for 1 to 3 hr prior to activation.
Embryo activation was performed 23 to 25 hr following maturation of oocytes by
treatment with 5 mM ionomycin for 5 min and 10 mg/ml cyclohexamide for 5 hr.
Activated embryos were then placed in co-culture dishes with CR2 medium atop a layer
of cumulus cells.
Bisulfite Conversion
A pool of 25 SCNT or IVF embryos was collected over the course of 3 IVF or
cloning sessions. Embryos were collected on day eight of development, and embryos of
similar size and quality were selected. Following collection embryos were snap frozen
with liquid nitrogen and stored at -80°C until direct bisulfite conversion. Approximately
1 million donor cells were collected and DNA was isolated using the DNeasy Blood and
Tissue Kit (Qiagen, Germantown, MD) and 1 µg of DNA was used for bisulfite
68
conversion. Embryos were used directly for bisulfite conversion. Bisulfite conversion
was carried out using the EZ DNA Methylation Kit (Zymo Research, Irvine, CA)
according to the manufacturer’s protocol. Converted DNA was immediately used in PCR
reactions.
PCR and 454 Sequencing
Following bisulfite conversion, converted embryo pools or donor cell DNA were
used as template in 25 µl PCR reactions using 1 to 3 µl bisulfite converted genomic DNA
with 0.6 µM primers in GoTaq Master Mix (Promega, Madison, WI) reactions. The
same cycling parameters were used for all primer pairs: 94°C for 1 min, annealing
temperature ranging from 54 to 59°C for 20 sec, and 72°C for 30 sec for 30 cycles. A
second PCR was performed using identical primers including the 454 adapter and key
sequence (adapter A CCATCTCATCCCTGCGTGTCTCCGACTCAG on the forward
primer and adapter B CCTATCCCCTGTGTGCCTTGGCAGTCTCAG on the reverse
primer). This second round of PCR used 1 to 3 µl of the PCR product from the first
round and 0.3 µM primer in GoTaq Master Mix (Promega) and the following cycling
parameters: 94°C for 30 sec, annealing temperature ranging from 55 to 60°C for 20 sec,
and 72°C for 30 sec for 15 cycles. After purification with AMPure beads (Beckman
Coulter, Brea, CA) and quantification with PicoGreen (Life Technologies, Carlsbad,
CA), PCR products were pooled and finally sequenced on the 454 GS FLX Titanium
DNA sequencer (Roche, Indianapolis, IN) according to the manufacturer’s protocols.
Briefly, amplicon libraries were subjected to emulsion PCR to produce DNA-coated
69
beads, loaded onto a 70X75 PicoTiterPlate, and sequenced using the FLX Titanium DNA
sequencing Kit.
Sequence Analysis
Sequencing data was analyzed for frequency of methylation using BISMA
analysis software (Rohde et al., 2010). Data are expressed as the percentage of
methylated CpG sites by normalizing methylation frequency data to the number of
sequence reads obtained.
Statistical Analysis
To determine whether methylation levels were significantly different among
donor cells, IVF, and SCNT embryos, we performed Fisher’s exact tests on a per CpG
site basis using the number of methylated transcripts and the number of unmethylated
transcripts in a categorical analyses; thus, these tests are influenced by the total number
of transcripts obtained for each amplicon. The resulting raw P values were then
Bonferroni-adjusted using proc multtest (SAS 9.3, Cary, NC) to account for multiple
testing.
Results
DNA methylation patterns were examined using bisulfite sequencing of 336 CpG
sites at single base pair resolution across six genes. DNA methylation within the donor
cell line was observed in the POU5F1, NANOG, and SOX2 genes, whereas limited
methylation was observed within the gene body of KLF4. The LIN28 and C-MYC genes
had no methylation within any of the regions examined and were excluded from further
70
analysis. Once the donor cell methylation patterns were examined, patterns of POU5F1,
NANOG, SOX2, and KLF4 within IVF and SCNT blastocyst stage embryos were
examined. This analysis revealed aberrant methylation of DNA within SCNT embryos.
Methylation of POU5F1
Analysis of the POU5F1 CpG island revealed an island beginning immediately
upstream from the translation start site that spans the entirety of the first exon (Fig. 2-1).
The methylation status of the POU5F1 CpG island is summarized in Figure 2-2. Overall,
54% of the CpG sites were methylated within this region. Interestingly the methylation
levels were not uniform across the length of the CpG island, as relatively low methylation
was detected at the 5’ end and comparatively high methylation at the 3’ end of the
examined sequence. Overall methylation in IVF embryos was low, with 35% of all CpG
sites methylated. Methylation in bovine SCNT embryos was substantially lower than
both donor cells and IVF blastocysts, with 18% of all CpG sites methylated (Fig. 2-2).
A statistical comparison of these methylation profiles revealed specific CpG sites
that are differentially methylated. CpG site 4 is among the most frequently methylated
sites in the donor cell line (Fig. 2-2). Seventy-one percent of donor cell sequence reads
are methylated, compared to 49% in IVF embryos. By contrast, none of the SCNT
sequence reads were methylated at this site. Similar observations were observed at sites
26 and 40, all of which were differentially methylated in SCNT embryos and donor cells
(Fig. 2-2). These CpG sites appear to have undergone successful reprogramming within
the SCNT genome, although based on methylation patterns observed within IVF samples,
the reduction of methylation levels appears to be inappropriate for the blastocyst stage.
71
In SCNT embryos, methylation at the 3’ end of the CpG island is more prevalent
compared to the 5’ end, although methylation of this 3’ region remains lower than levels
detected in IVF embryos. Of note, methylation at CpG sites 140, 164, and 263 is
significantly different lower in SCNT embryos compared to IVF embryos or to donor
cells (Fig. 2-2). At a number of other CpG sites in this 3’ region, methylation levels are
significantly different among all sample types (e.g., sites 232, 250, 284, 320, 355).
When considering the entire CpG island, average methylation levels are highest in
donor cells (54%), lower in IVF embryos (35%), and lowest in SCNT embryos (18%).
However, three closely aligned CpG sites break from this apparent pattern. Specifically
CpG dinucleotides at sites 56, 64, and 79 are more frequently methylated in SCNT
embryos than IVF, yet methylation at these sites is lower in both embryo types compared
to donor cells (Fig. 2-2).
Methylation of NANOG
NANOG contains a small CpG island within the first intron that begins 270 bp
downstream from the transcription start site (Fig. 2-1). The methylation status for this
region is summarized in Figure 2-3. Within this region, some SCNT CpG sites more
closely resemble the methylation pattern of the donor cell, while others are more similar
to IVF. The overall average DNA methylation for this region is higher in the SCNT
samples than donor cell as well as IVF. Within the donor cell samples, 25% of CpG sites
are methylated; in IVF blastocysts, 21% of CpG dinucleotides are methylated; and within
SCNT blastocysts, 31% of CpG sites are methylated (Fig. 2-3). Results of the statistical
analyses of methylation frequency were not significant for any particular CpG site, likely
72
a consequence of the low number of sequence reads obtained. Even so, it is interesting to
note that for seven CpG loci, methylation is completely absent in IVF embryos, while
several of these sites (e.g., 303, 308, 487) are moderately methylated in SCNT embryos
and donor cells. Only one site was not methylated in SCNT embryos, while every CpG
locus examined was methylated to some extent in donor cells.
Another interesting feature of the examined region in NANOG was an apparent
spike in the percentage of methylated cytosines at CpG site 523. A methylation spike is
defined here as a sharp elevation in the methylation of a CpG dinucleotide that is closely
associated with other CpG loci that retain low levels of methylation. Within the NANOG
CpG island site 523 was methylated at a percentage of 83% in SCNT embryos, 70% in
donor cells, and 55% in IVF embryos (Fig. 2-3).
Methylation of SOX2
SOX2 is a single exonic gene and contains a 3000 bp CpG dense region that
begins 1500 bp upstream from the transcriptional start site and continues throughout the
gene. Three CpG islands were identified within this region, and several CpG sites were
examined for methylation (Fig. 2-1). Our analysis identified two regions that were
positive for methylation within donor cells, and both regions were subsequently
examined for methylation patterns within IVF and SCNT embryos. Results are
summarized in Figure 2-4.
The first region is upstream from the translational start site and ranges from -743
bp to -480 bp. This region is hypomethylated at all CpG sites with the exception of a
cluster of four CpG dinucleotides from -598 bp to -585 bp. For all three sample types, a
73
similar pattern of methylation was observed in the sequence of these four CpG sites
contained within a 14 bp segment of DNA, where sites -598, -589, and -585 are
methylated with moderate to high frequency (38 to 100%), while site -593 is not
methylated for any sequence read examined in any of the tissue types (Fig. 2-4).
Interestingly, methylation of these sites seems to be maintained in a specific pattern
across tissue types, where sites -598, -589, and -585 are methylated with 100% frequency
in SCNT embryos, 63 to 70% frequency in donor cells, and only 38 to 50% frequency in
IVF embryos. Also within this region, CpG site -701 is appears more frequently
methylated in SCNT embryos and donor cells compared to no methylation in IVF
embryos.
A second region within SOX2 ranging from 600 bp to 1400 bp downstream from
the transcriptional start site was methylated to varying degrees in donor cells, IVF
embryos and SCNT embryos. A second cluster of highly methylated sites exists within
this region, and methylation for this sequence of CpG sites (562-669) is consistent for
IVF and SCNT embryos and higher than that of donor cells for sites 652 and 669. Few
statistical differences were detected for CpG sites in the SOX2 gene due to the relatively
low number of reads obtained.
Methylation of KLF4
Two large CpG islands are associated with the KLF4 gene ranging from 1800 bp
upstream to 2200 bp downstream from the transcriptional start site in the third intron
(Fig. 2-1). In all sample types, methylation of the regions from -1218 to -361 and +1250
to +1900 was quite low, averaging from 1 to 9% (Fig. 2-5). The region upstream from
74
the transcriptional start site was not methylated at any CpG site in any of the tissue types
examined (data not shown). The region downstream of the transcriptional start site, from
1553 to 1884, was methylated to a limited extent in donor cells, but only minimally
methylated at a few loci in IVF or SCNT embryos. The donor cell likewise has minimal
methylation throughout the region (Fig. 2-5). Interestingly, significant differences in
frequency of methylation were observed for many CpG sites among the tissue types, even
though the differences were not large in magnitude. Most notable were differences in
donor cells and IVF embryos from sites 1731 to 1806, as well as differences between
SCNT embryos or donor cells compared to IVF embryos in the region from 1281 to 1495
(Fig. 2-5).
Discussion
The aim of this project was to determine whether the DNA methylation patterns
of POU5F1, NANOG, SOX2, KLF4, LIN28, and C-MYC were appropriately
reprogrammed in SCNT embryos prior to implantation. We accomplished this objective
by measuring DNA methylation of the CpG islands associated with each gene in SCNT
and IVF blastocysts. We also measured the methylation status of the donor cell line to
determine whether methylation patterns of SCNT embryos more closely resembled those
of the donor cell line or IVF embryos. The results of this study suggest that DNA
methylation patterns of POU5F1, NANOG, and SOX2 in SCNT embryos are abnormally
established at several CpG sites in pre-implantation SCNT embryos. Across the broad
CpG islands examined, levels of methylation of the NANOG and SOX2 genes were
elevated and more closely matched the methylation patterns observed in donor cells. On
75
the other hand, POU5F1 is relatively hypomethylated compared to donor cells and IVF
embryos; this observation suggests that while reprogramming of POU5F1 occurs in
SCNT embryos, the timing may be delayed compared to IVF embryos.
Several studies have examined the methylation status of CpG sites proximal to the
transcriptional start sites for POU5F1, NANOG, and SOX2 genes and observed abnormal
methylation patterns in SCNT embryos (Yamazaki et al., 2006). Aberrant epigenetic
patterns likely contribute to cloning failure during preimplantation development and early
gestation (Bourc'his et al., 2001; Dean et al., 2001; Kang et al., 2001; Ohgane et al.,
2001; Kang et al., 2002). In this study, we examined previously uncharacterized CpG
islands associated with the POU5F1, NANOG, and SOX2 genes, as well as three other
genes not previously studied in bovine embryos, KLF4, LIN28, and C-MYC. These later
three genes were selected because they are likely to be involved in donor cell
reprogramming (Takahashi et al., 2007; Yu et al., 2007). CpG islands may or may not
directly regulate expression, but can be methylated and are useful in determining the
reprogramming status of an associated gene.
DNA methylation is tightly regulated during embryonic development in order to
enable proper gene expression and to reestablish appropriate epigenetic patterns during
tissue differentiation. Following fertilization, the maternal genome is passively
demethylated throughout early embryonic stages (Monk et al., 1987; Howlett and Reik,
1991; Rougier et al., 1998), while the paternal genome is actively demethylated
immediately following fertilization (Mayer et al., 2000; Oswald et al., 2000). SCNT
embryos are susceptible to disruption of methylation patterns (Kikyo and Wolffe, 2000;
76
Kang et al., 2001; Bonk et al., 2008). The demethylation that occurs in normal
developing embryos is incomplete and abnormally timed in clones (Armstrong et al.,
2006).
As expected, CpG sites associated with the POU5F1 gene are more frequently
methylated in donor cells than in IVF embryos. This difference suggests that
reprogramming of the epigenome for POU5F1 is underway in IVF embryos, likely a
consequence of ongoing cellular differentiation, yet the process is not quite complete.
Alternatively, the overall lower level of methylation for CpG sites associated with
POU5F1 in SCNT embryos suggests that the timing of reprogramming in clones may be
delayed compared to IVF blastocysts. Moreover, the disrupted methylation pattern may
contribute to the abnormal expression of POU5F1 observed in clones (Boiani et al.,
2002; Bortvin et al., 2003; Aston et al., 2010). Because POU5F1 is critical for early
development (Nichols et al., 1998; Niwa et al., 2000), abnormal expression of this gene
imposes an impediment to normal development. Yamazaki et al. (2006) showed that
reprogramming of POU5F1 occurs gradually throughout development, and our results
confirm that reprogramming of methylation patterns does indeed occur. The relatively
low level of overall methylation for SCNT embryos may be the result of gradual
demethylation of the POU5F1 gene during early embryo development. Alternatively, the
low abundance of methylated CpG sites in SCNT embryos may be a consequence of
fewer trophectoderm cells present in cloned embryos (Koo et al., 2002). Due to ongoing
active cellular differentiation, one would expect higher levels of methylation for
pluripotency factors in trophectoderm cells (Hattori et al., 2004). A loss of
77
trophectoderm cells in cloned embryos, which are often smaller in size compared to IVF
embryos, could conceivably influence the proportion of methylated CpG sites acquired
from whole embryo samples. Interestingly, one cluster of CpG sites within POU5F1 was
identified for which methylation frequency was substantially higher in SCNT embryos
compared to IVF. However, the role of these specific CpG sites in regulating gene
expression is, as yet, unclear nor is it known whether these loci are resistant to
reprogramming or sensitive to the reestablishment of methylation within the blastocyst
stage. Overall, the methylation patterns observed for POU5F1 are atypical of those
observed in IVF and highlight the aberrant nature of the SCNT reprogramming among
specific genes.
NANOG is involved in the maintenance of pluripotency during embryo
development, prevents differentiation (Chambers et al., 2003), and is essential to early
development. Aberrant NANOG methylation may hinder expression in the early embryo
and perturb SCNT embryo development. However, in this study, the overall patterns of
methylation across donor cells, IVF and SCNT embryos were broadly similar and
relatively moderately methylated (with a few exceptions for IVF embryos where no
methylation was detected). Interestingly, NANOG contains a CpG site at location 523
characterized by a sudden increase in methylation, whereas neighboring sites were only
moderately methylated. That higher levels of methylation were consistently maintained
for this site in all sample types suggest that CpG site 523 may be functionally important,
yet additional experiments would be necessary to establish a specific role for this
particular locus in regulating NANOG expression.
78
SOX2 maintains potency and directs cell lineage fate in combination with
POU5F1 (Avilion et al., 2003). SOX2 contains a CpG island upstream from the
translational start site that is largely hypomethylated in donor cells, IVF and SCNT
embryos. An exception to this regional hypomethylation is a 14 bp sequence
approximately 600 bp upstream from the translational start site that contains four CpG
sites. Three of these loci are moderately methylated in donor cells and IVF embryos and
highly methylated in SCNT embryos, whereas the fourth site is void of methylation in all
tissue types. From this data set, it is not possible to know whether or not the elevated
methylation at the three CpG sites in this region is a consequence of incomplete
reprogramming or premature establishment of hypermethylation for this pluripotency
gene in cloned embryos. The apparent tight regulation of methylation within this small
region of DNA, as well as the absence of methylation on surrounding CpG dinucleotides
suggests that this sequence of CpG sites may play a key regulatory role for SOX2
expression. An analysis of this region of the SOX2 gene promoter using PROMO
(Messeguer et al., 2002) revealed that eight transcription factors have affinity for this
sequence: SP1, Mad, fα-fε, BTEB3, ADR1, GAGA Factor, P53, and Pax-6.
Interestingly, Pax-6 is known to regulate SOX2 expression (Wen et al., 2008). Future
studies should focus on the impact of high methylation at these three CpG sites in the
promoter for SOX2 on recruitment of critical transcription factors, such as Pax-6, and
subsequent expression of the SOX2 gene.
Within the donor cell line, we observed a high degree of variation of methylation
among individual cells. There was an average of 125 reads among donor samples, which
79
generated a large enough data set to see the methylation pattern of several hundred cells
gathered from a pool of potential donor cells. This allowed us to look at the possible
effect of random donor cell selection. What we observe is that among the genes
examined, a small percentage of donor cells lack methylation. In the case of the
upstream region of SOX2, 26% of donor cells contain no DNA methylation at any site.
As methylation at these locations may carry over from the donor cells within SOX2, the
lack on methylation on in 26% of donor cells suggests that some donor cells transferred
during SCNT are unmethylated at these CpG sites and the clone would inherit an
unmethylated pattern. This may account for why some clones are better prepared for
early development.
Overall, we examined 336 CpG sites across six genes involved in reprogramming
and early development SCNT and IVF embryos. We found that SCNT embryos harbor
some aberrations in patterns of DNA methylation, although these abnormalities were
often localized to small regions or specific CpG sites. The most profound evidence for
inappropriate genome reprogramming was obtained for the POU5F1 gene, which was
overall hypomethylated compared to IVF embryos. Future work should aim to determine
the mechanism(s) responsible for appropriate reprogramming of NANOG and SOX2, and
to determine whether or not this mechanism is defective or inefficient for reprogramming
the POU5F1 gene. Furthermore, future work should establish the consequences of
aberrant methylation following implantation and determine the regulatory nature of the
three CpG sites identified within the promoter region of SOX2.
80
References
Adhikary S, Eilers M. 2005. Transcriptional regulation and transformation by Myc
proteins. Nat Rev Mol Cell Biol 6(8):635-645.
Antequera F, Bird A. 1993. Number of CpG islands and genes in human and mouse. Proc
Natl Acad Sci USA 90(24):11995-11999.
Armstrong L, Lako M, Dean W, Stojkovic M. 2006. Epigenetic modification is central to
genome reprogramming in somatic cell nuclear transfer. Stem Cells 24(4):805814.
Aston KI, Li GP, Hicks BA, Sessions BR, Davis AP, Rickords LF, Stevens JR, White
KL. 2010. Abnormal levels of transcript abundance of developmentally important
genes in various stages of preimplantation bovine somatic cell nuclear transfer
embryos. Cell Reprogram 12(1):23-32.
Avilion AA, Nicolis SK, Pevny LH, Perez L, Vivian N, Lovell-Badge R. 2003.
Multipotent cell lineages in early mouse development depend on SOX2 function.
Genes Dev 17(1):126-140.
Bartolomei MS, Webber AL, Brunkow ME, Tilghman SM. 1993. Epigenetic mechanisms
underlying the imprinting of the mouse H19 gene. Genes Dev 7(9):1663-1673.
Bird A. 2002. DNA methylation patterns and epigenetic memory. Genes Dev 16(1):6-21.
Boiani M, Eckardt S, Scholer HR, McLaughlin KJ. 2002. Oct4 distribution and level in
mouse clones: consequences for pluripotency. Genes Dev 16(10):1209-1219.
Bonk AJ, Li R, Lai L, Hao Y, Liu Z, Samuel M, Fergason EA, Whitworth KM, Murphy
CN, Antoniou E, Prather RS. 2008. Aberrant DNA methylation in porcine in
vitro-, parthenogenetic-, and somatic cell nuclear transfer-produced blastocysts.
Mol Reprod Dev 75(2):250-264.
Bortvin A, Eggan K, Skaletsky H, Akutsu H, Berry DL, Yanagimachi R, Page DC,
Jaenisch R. 2003. Incomplete reactivation of Oct4-related genes in mouse
embryos cloned from somatic nuclei. Development 130(8):1673-1680.
Bourc'his D, Le Bourhis D, Patin D, Niveleau A, Comizzoli P, Renard JP, ViegasPequignot E. 2001. Delayed and incomplete reprogramming of chromosome
methylation patterns in bovine cloned embryos. Curr Biol 11(19):1542-1546.
Chambers I, Colby D, Robertson M, Nichols J, Lee S, Tweedie S, Smith A. 2003.
Functional expression cloning of Nanog, a pluripotency sustaining factor in
embryonic stem cells. Cell 113(5):643-655.
81
Chavatte-Palmer P, Heyman Y, Richard C, Monget P, LeBourhis D, Kann G, Chilliard Y,
Vignon X, Renard JP. 2002. Clinical, hormonal, and hematologic characteristics
of bovine calves derived from nuclei from somatic cells. Biol Reprod 66(6):15961603.
Chen RZ, Pettersson U, Beard C, Jackson-Grusby L, Jaenisch R. 1998. DNA
hypomethylation leads to elevated mutation rates. Nature 395(6697):89-93.
Colman A. 1999. Somatic cell nuclear transfer in mammals: progress and applications.
Cloning 1(4):185-200.
Csankovszki G, Nagy A, Jaenisch R. 2001. Synergism of Xist RNA, DNA methylation,
and histone hypoacetylation in maintaining X chromosome inactivation. J Cell
Biol 153(4):773-784.
Darr H, Benvenisty N. 2009. Genetic analysis of the role of the reprogramming gene
LIN-28 in human embryonic stem cells. Stem Cells 27(2):352-362.
De Sousa PA, King T, Harkness L, Young LE, Walker SK, Wilmut I. 2001. Evaluation of
gestational deficiencies in cloned sheep fetuses and placentae. Biol Reprod
65(1):23-30.
Dean W, Santos F, Stojkovic M, Zakhartchenko V, Walter J, Wolf E, Reik W. 2001.
Conservation of methylation reprogramming in mammalian development:
aberrant reprogramming in cloned embryos. Proc Natl Acad Sci USA
98(24):13734-13738.
DeBaun MR, Niemitz EL, Feinberg AP. 2003. Association of in vitro fertilization with
Beckwith-Wiedemann syndrome and epigenetic alterations of LIT1 and H19. Am
J Hum Genet 72(1):156-160.
Evans PM, Liu C. 2008. Roles of Krupel-like factor 4 in normal homeostasis, cancer and
stem cells. Acta Biochim Biophys Sin (Shanghai) 40(7):554-564.
Everts RE, Chavatte-Palmer P, Razzak A, Hue I, Green CA, Oliveira R, Vignon X,
Rodriguez-Zas SL, Tian XC, Yang X, Renard JP, Lewin HA. 2008. Aberrant gene
expression patterns in placentomes are associated with phenotypically normal and
abnormal cattle cloned by somatic cell nuclear transfer. Physiol Genomics
33(1):65-77.
Farin CE, Farin PW, Piedrahita JA. 2004. Development of fetuses from in vitro-produced
and cloned bovine embryos. J Anim Sci 82 E-Suppl:E53-62.
Hattori N, Nishino K, Ko YG, Ohgane J, Tanaka S, Shiota K. 2004. Epigenetic control of
mouse Oct-4 gene expression in embryonic stem cells and trophoblast stem cells.
J Biol Chem 279(17):17063-17069.
82
Hendrich B, Bird A. 1998. Identification and characterization of a family of mammalian
methyl-CpG binding proteins. Mol Cell Biol 18(11):6538-6547.
Hill JR, Burghardt RC, Jones K, Long CR, Looney CR, Shin T, Spencer TE, Thompson
JA, Winger QA, Westhusin ME. 2000a. Evidence for placental abnormality as the
major cause of mortality in first-trimester somatic cell cloned bovine fetuses. Biol
Reprod 63(6):1787-1794.
Hill JR, Roussel AJ, Cibelli JB, Edwards JF, Hooper NL, Miller MW, Thompson JA,
Looney CR, Westhusin ME, Robl JM, Stice SL. 1999. Clinical and pathologic
features of cloned transgenic calves and fetuses (13 case studies). Theriogenology
51(8):1451-1465.
Hill JR, Winger QA, Long CR, Looney CR, Thompson JA, Westhusin ME. 2000b.
Development rates of male bovine nuclear transfer embryos derived from adult
and fetal cells. Biol Reprod 62(5):1135-1140.
Howlett SK, Reik W. 1991. Methylation levels of maternal and paternal genomes during
preimplantation development. Development 113(1):119-127.
Humpherys D, Eggan K, Akutsu H, Hochedlinger K, Rideout WM, 3rd, Biniszkiewicz D,
Yanagimachi R, Jaenisch R. 2001. Epigenetic instability in ES cells and cloned
mice. Science 293(5527):95-97.
Illingworth RS, Bird AP. 2009. CpG islands--'a rough guide'. FEBS Lett 583(11):17131720.
Jaenisch R, Bird A. 2003. Epigenetic regulation of gene expression: how the genome
integrates intrinsic and environmental signals. Nat Genet 33 Suppl:245-254.
Jones PL, Veenstra GJ, Wade PA, Vermaak D, Kass SU, Landsberger N, Strouboulis J,
Wolffe AP. 1998. Methylated DNA and MeCP2 recruit histone deacetylase to
repress transcription. Nat Genet 19(2):187-191.
Kang YK, Koo DB, Park JS, Choi YH, Chung AS, Lee KK, Han YM. 2001. Aberrant
methylation of donor genome in cloned bovine embryos. Nat Genet 28(2):173177.
Kang YK, Park JS, Koo DB, Choi YH, Kim SU, Lee KK, Han YM. 2002. Limited
demethylation leaves mosaic-type methylation states in cloned bovine preimplantation embryos. EMBO J 21(5):1092-1100.
Kikyo N, Wolffe AP. 2000. Reprogramming nuclei: insights from cloning, nuclear
transfer and heterokaryons. J Cell Sci 113 ( Pt 1):11-20.
83
Koo DB, Kang YK, Choi YH, Park JS, Kim HN, Oh KB, Son DS, Park H, Lee KK, Han
YM. 2002. Aberrant allocations of inner cell mass and trophectoderm cells in
bovine nuclear transfer blastocysts. Biol Reprod 67(2):487-492.
Li E, Beard C, Jaenisch R. 1993. Role for DNA methylation in genomic imprinting.
Nature 366(6453):362-365.
Li S, Li Y, Du W, Zhang L, Yu S, Dai Y, Zhao C, Li N. 2005. Aberrant gene expression
in organs of bovine clones that die within two days after birth. Biol Reprod
72(2):258-265.
Mayer W, Niveleau A, Walter J, Fundele R, Haaf T. 2000. Demethylation of the zygotic
paternal genome. Nature 403(6769):501-502.
Messeguer X, Escudero R, Farre D, Nunez O, Martinez J, Alba MM. 2002. PROMO:
detection of known transcription regulatory elements using species-tailored
searches. Bioinformatics 18(2):333-334.
Mitsui K, Tokuzawa Y, Itoh H, Segawa K, Murakami M, Takahashi K, Maruyama M,
Maeda M, Yamanaka S. 2003. The homeoprotein Nanog is required for
maintenance of pluripotency in mouse epiblast and ES cells. Cell 113(5):631-642.
Mohandas T, Sparkes RS, Shapiro LJ. 1981. Reactivation of an inactive human X
chromosome: evidence for X inactivation by DNA methylation. Science
211(4480):393-396.
Monk M, Boubelik M, Lehnert S. 1987. Temporal and regional changes in DNA
methylation in the embryonic, extraembryonic and germ cell lineages during
mouse embryo development. Development 99(3):371-382.
Nichols J, Zevnik B, Anastassiadis K, Niwa H, Klewe-Nebenius D, Chambers I, Scholer
H, Smith A. 1998. Formation of pluripotent stem cells in the mammalian embryo
depends on the POU transcription factor Oct4. Cell 95(3):379-391.
Niemann H, Tian XC, King WA, Lee RS. 2008. Epigenetic reprogramming in embryonic
and foetal development upon somatic cell nuclear transfer cloning. Reproduction
135(2):151-163.
Niwa H, Miyazaki J, Smith AG. 2000. Quantitative expression of Oct-3/4 defines
differentiation, dedifferentiation or self-renewal of ES cells. Nat Genet 24(4):372376.
Ohgane J, Wakayama T, Kogo Y, Senda S, Hattori N, Tanaka S, Yanagimachi R, Shiota
K. 2001. DNA methylation variation in cloned mice. Genesis 30(2):45-50.
84
Oswald J, Engemann S, Lane N, Mayer W, Olek A, Fundele R, Dean W, Reik W, Walter
J. 2000. Active demethylation of the paternal genome in the mouse zygote. Curr
Biol 10(8):475-478.
Pesce M, Scholer HR. 2001. Oct-4: gatekeeper in the beginnings of mammalian
development. Stem Cells 19(4):271-278.
Prokhortchouk A, Hendrich B, Jorgensen H, Ruzov A, Wilm M, Georgiev G, Bird A,
Prokhortchouk E. 2001. The p120 catenin partner Kaiso is a DNA methylationdependent transcriptional repressor. Genes Dev 15(13):1613-1618.
Reik W, Walter J. 2001. Genomic imprinting: parental influence on the genome. Nat Rev
Genet 2(1):21-32.
Rohde C, Zhang Y, Reinhardt R, Jeltsch A. 2010. BISMA--fast and accurate bisulfite
sequencing data analysis of individual clones from unique and repetitive
sequences. BMC Bioinformatics 11:230.
Rougier N, Bourc'his D, Gomes DM, Niveleau A, Plachot M, Paldi A, Viegas-Pequignot
E. 1998. Chromosome methylation patterns during mammalian preimplantation
development. Genes Dev 12(14):2108-2113.
Sarraf SA, Stancheva I. 2004. Methyl-CpG binding protein MBD1 couples histone H3
methylation at lysine 9 by SETDB1 to DNA replication and chromatin assembly.
Mol Cell 15(4):595-605.
Sebastiano V, Gentile L, Garagna S, Redi CA, Zuccotti M. 2005. Cloned preimplantation mouse embryos show correct timing but altered levels of gene
expression. Mol Reprod Dev 70(2):146-154.
Shiota K. 2004. DNA methylation profiles of CpG islands for cellular differentiation and
development in mammals. Cytogenet Genome Res 105(2-4):325-334.
Solter D. 2000. Mammalian cloning: advances and limitations. Nat Rev Genet 1(3):199207.
Stoger R, Kubicka P, Liu CG, Kafri T, Razin A, Cedar H, Barlow DP. 1993. Maternalspecific methylation of the imprinted mouse Igf2r locus identifies the expressed
locus as carrying the imprinting signal. Cell 73(1):61-71.
Sumer H, Liu J, Malaver-Ortega LF, Lim ML, Khodadadi K, Verma PJ. 2011. NANOG
is a key factor for induction of pluripotency in bovine adult fibroblasts. J Anim
Sci 89(9):2708-2716.
85
Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, Yamanaka S.
2007. Induction of pluripotent stem cells from adult human fibroblasts by defined
factors. Cell 131(5):861-872.
Wade PA, Gegonne A, Jones PL, Ballestar E, Aubry F, Wolffe AP. 1999. Mi-2 complex
couples DNA methylation to chromatin remodelling and histone deacetylation.
Nat Genet 23(1):62-66.
Watt F, Molloy PL. 1988. Cytosine methylation prevents binding to DNA of a HeLa cell
transcription factor required for optimal expression of the adenovirus major late
promoter. Genes Dev 2(9):1136-1143.
Wells DN. 2005. Animal cloning: problems and prospects. Rev Sci Tech 24(1):251-264.
Wells DN, Misica PM, Tervit HR. 1999. Production of cloned calves following nuclear
transfer with cultured adult mural granulosa cells. Biol Reprod 60(4):996-1005.
Wen J, Hu Q, Li M, Wang S, Zhang L, Chen Y, Li L. 2008. Pax6 directly modulate Sox2
expression in the neural progenitor cells. Neuroreport 19(4):413-417.
Wilmut I, Schnieke AE, McWhir J, Kind AJ, Campbell KH. 1997. Viable offspring
derived from fetal and adult mammalian cells. Nature 385(6619):810-813.
Wilmut I, Young L, Campbell KH. 1998. Embryonic and somatic cell cloning. Reprod
Fertil Dev 10(7-8):639-643.
Yamazaki Y, Fujita TC, Low EW, Alarcon VB, Yanagimachi R, Marikawa Y. 2006.
Gradual DNA demethylation of the Oct4 promoter in cloned mouse embryos. Mol
Reprod Dev 73(2):180-188.
Yang X, Smith SL, Tian XC, Lewin HA, Renard JP, Wakayama T. 2007. Nuclear
reprogramming of cloned embryos and its implications for therapeutic cloning.
Nat Genet 39(3):295-302.
Yoder JA, Walsh CP, Bestor TH. 1997. Cytosine methylation and the ecology of
intragenomic parasites. Trends Genet 13(8):335-340.
Young LE, Sinclair KD, Wilmut I. 1998. Large offspring syndrome in cattle and sheep.
Rev Reprod 3(3):155-163.
Yu J, Vodyanik MA, Smuga-Otto K, Antosiewicz-Bourget J, Frane JL, Tian S, Nie J,
Jonsdottir GA, Ruotti V, Stewart R, Slukvin, II, Thomson JA. 2007. Induced
pluripotent stem cell lines derived from human somatic cells. Science
318(5858):1917-1920.
86
Zakhartchenko V, Mueller S, Alberio R, Schernthaner W, Stojkovic M, Wenigerkind H,
Wanke R, Lassnig C, Mueller M, Wolf E, Brem G. 2001. Nuclear transfer in
cattle with non-transfected and transfected fetal or cloned transgenic fetal and
postnatal fibroblasts. Mol Reprod Dev 60(3):362-369.
Zhang Y, Ng HH, Erdjument-Bromage H, Tempst P, Bird A, Reinberg D. 1999. Analysis
of the NuRD subunits reveals a histone deacetylase core complex and a
connection with DNA methylation. Genes Dev 13(15):1924-1935.
87
Table 2-1. Primers Used for Bisulfite Sequencing
Gene
POU5F1
Forward
Reverse
Forward
Reverse
Forward
Reverse
Forward
Reverse
NANOG
Forward
Reverse
SOX2
Forward
Reverse
Forward
Reverse
Forward
Reverse
Forward
Reverse
Forward
Reverse
KLF4
Forward
Reverse
Forward
Reverse
Forward
Reverse
Forward
Reverse
Forward
Reverse
LIN28
Forward
Reverse
C-MYC
Forward
Reverse
Primer
GTTAGAGGTTAAGGTTAGTGGG
GTTTTATGATTTGTGTGGAGGGATG
TTTGGATGAGTTTTTAAGGGTTTT
TTGGATAAGGAGAAGTTGGAG
TGGAGTYGAATTTTGAGGAGG
TTTTAGGGTTAAGTGTGGTTT
TGGTGAAAGTTTTAGTGTAGTGTA
GTATTTAAATTGAGGTGTTTGT
ATGTTTGAAGAAAGTTAYGTGTT
AAGGTGTGTTAGAGATGGTAT
Location
-110
218
99
389
384
827
3961
4507
253
746
TAAGAGAGTGGAAGGAAATTTAG
GTTTTGTTTGGTTTGTATTTGTG
TGGTTTGATTTTAGTTTTGGAGG
AGTGGGATAGTTTTATTGAGTTAT
AGTGGGATAGTTTTATTGAGTTAT
TTGTTAAGGTAGAGAAGAGAGT
GATGGTTTAAGAGAATTTTAAGATG
TAGTATGATGTAGGATTAGTTGGG
TAGTATGATGTAGGATTAGTTGGG
TTAAAGAAAAAGAGGGAAATGGAG
-1587
-1277
-770
-477
-471
20
580
943
919
1424
GTGTTTTGTAAAGTGTTTTGAGAT
GGGATTAGAGTAGGATTTTTGG
GAGTTTAAGAAGGATTATGYGAG
GGTAGTTTTTYGGTTTGAGAGAG
GGTAGTTTTTYGGTTTGAGAGAG
GTTTTGGGGTTYGGGTATT
GGTTTGAATTGGATTTAGTGTATA
TAGAGGTTTATTTGGGTATTGGAT
GTGTTTTAAGATTAAGTAGGAGGT
GAGTTTAAATTAAAGAGGGGAAGA
-1614
-1274
-1347
-725
-748
-178
1250
1532
1455
1912
TGGAGTTAGGAATTTTGGTGTTTTA
AGGGGAGATTTAGGGTTTTTATTGT
196
796
TTTTGTTTTAGATGATTTAAGAATATAGGA
TTATTATTGTTTGGAAGGGTAGGGT
-506
-85
88
POU5F1!
NANOG!
SOX2!
KLF4!
LIN28!
c-MYC!
-2000!
-1500!
-1000!
-500!
0!
500!
1000!
1500!
2000!
Figure 2-1. Map of CpG islands and regions examined for methylation. Horizontal
line shows the first 2000 bp of the gene as well as 2000 bp upstream from the
translational start site. White boxes show exons. Vertical lines represent CpG sites and
grey boxes represent CpG islands. Arrows indicate primers of the regions examined for
methylation analysis.
89
Figure 2-2. Heat map depicting methylation profile of POU5F1 in donor cells and
blastocyst stage SCNT and IVF embryos. The reads column indicates the total number
of reads associated with each amplicon. The average percent methylation across all CpG
sites examined is also shown. Position number specifies the bp location of the CpG site
in relation to the translational start site. Each heat map cell indicates the percentage of
CpG sites positive for methylation. Red corresponds to high methylation and blue to low
methylation. The significant difference table below each heat map panel provides the
results of Fisher’s exact tests for each CpG site for the indicated pair-wise comparison
with Bonferroni correction for multiple testing (SAS). A stared box indicates a
statistically significance difference in methylation (P <0.05).
90
Figure 2-3. Heat map depicting methylation profile of NANOG in donor cells and
blastocyst stage SCNT and IVF embryos. The reads column indicates the total number
of reads associated with each amplicon. The average percent methylation across all CpG
sites examined is also shown. Position number specifies the bp location of the CpG site
in relation to the translational start site. Each heat map cell indicates the percentage of
CpG sites positive for methylation. Red corresponds to high methylation and blue to low
methylation. The significant difference table below each heat map panel provides the
results of Fisher’s exact tests for each CpG site for the indicated pair-wise comparison
with Bonferroni correction for multiple testing (SAS). A stared box indicates a
statistically significance difference in methylation (P <0.05).
91
Figure 2-4. Heat map depicting methylation profile of SOX2 in donor cells and
blastocyst stage SCNT and IVF embryos. The reads column indicates the total number
of reads associated with each amplicon. The average percent methylation across all CpG
sites examined is also shown. Position number specifies the bp location of the CpG site
in relation to the translational start site. Each heat map cell indicates the percentage of
CpG sites positive for methylation. Red corresponds to high methylation and blue to low
methylation. The significant difference table below each heat map panel provides the
results of Fisher’s exact tests for each CpG site for the indicated pair-wise comparison
with Bonferroni correction for multiple testing (SAS). A stared box indicates a
statistically significance difference in methylation (P <0.05).
92
Figure 2-5. Heat map depicting methylation profile of KLF4 in donor cells and
blastocyst stage SCNT and IVF embryos. The reads column indicates the total number
of reads associated with each amplicon. The average percent methylation across all CpG
sites examined is also shown. Position number specifies the bp location of the CpG site
in relation to the translational start site. Each heat map cell indicates the percentage of
CpG sites positive for methylation. Red corresponds to high methylation and blue to low
methylation. The significant difference table below each heat map panel provides the
results of Fisher’s exact tests for each CpG site for the indicated pair-wise comparison
with Bonferroni correction for multiple testing (SAS). A stared box indicates a
statistically significance difference in methylation (P <0.05).
93
CHAPTER 3
DNA METHYLATION OF THE LIN28 PSEUDOGENE FAMILY
Abstract
DNA methylation is a covalent attachment on DNA that directs epigenetic
silencing of selected regions of DNA. Methylation is widespread throughout the genome
and contributes to various forms of genomic regulation. Pseudogenes are decayed copies
of duplicated genes that spread throughout the genome by retrotransposition.
Pseudogenes are transcriptionally silenced by DNA methylation, but little is know about
the mode of epigenetic regulation of pseudogenes. Using bisulfite next generation
sequencing we examined the methylation status of four pseudogenes of the LIN28 family
along with the parent gene to determine the methylation status of dispersed pseudogenes
sharing a common origin. Our aim was to determine whether pseudogenes derived from
LIN28 maintain the same pattern of methylation as the LIN28 parent gene or acquire a
methylation pattern that reflects the region of insertion. Results of this study indicate that
pseudogenes are methylated in patterns specific to the local genomic environment and do
not recapitulate the methylation signature of the parent gene.
Introduction
DNA methylation is utilized to control diverse aspects of genome regulation and
transcriptional activity. Methylation of mammalian DNA involves the addition of a
methyl group to the 5’-carbon the cytosine in cytosine-guanine dinucleotides (termed
CpG sites). This system of methylation likely evolved from a genomic defense system
94
responsible for preventing the spread of parasitic genetic elements. DNA methylation
has since evolved to play an active role in maintaining genetic structure and genome
regulation (Chen et al., 1998). Methylation is involved in X-chromosome inactivation
(Mohandas et al., 1981; Csankovszki et al., 2001), silencing of transposable elements
(Liu et al., 1994; Woodcock et al., 1997; Yoder et al., 1997; Walsh et al., 1998), tissuespecific gene expression (De Smet et al., 1996; De Smet et al., 1999; Jaenisch and Bird,
2003; Shiota, 2004), and gene imprinting (Bartolomei et al., 1993; Li et al., 1993; Stoger
et al., 1993; Bird, 2002). DNA methylation is widespread throughout the genome, and
the maintenance of methylation patterns is highly regulated and tissue-specific (Eckhardt
et al., 2006; Illingworth et al., 2008). Patterns of methylation are established and
maintained by DNA methyltransferase (DNMT) enzymes (Holliday and Pugh, 1975;
Riggs, 2002; Jones and Liang, 2009). The absence of DNA methylation results in
embryonic lethality, which highlights the crucial role of DNA methylation in
development (Li et al., 1992; Okano et al., 1999; Goll and Bestor, 2005).
Methylation regulates pseudogenes within the genome (Grunau et al., 2000;
Eckhardt et al., 2006). Pseudogenes are decayed copies of active genes that have arisen
from either a duplication event in which the entire gene, or portion of a gene, is
duplicated (non-processed pseudogenes) or from the retrotransposition of a gene
transcript back into the genome (processed pseudogenes). An analysis of the human
genome estimates that as many as 19,000 pseudogenes are evenly distributed throughout
the genome, and approximately 70% of these are processed pseudogenes (Riggs, 2002;
Rauch et al., 2009). Pseudogenes primarily arise from parent genes that are
95
transcriptionally active within the germ line, and 10% of genes within the human genome
have at least one corresponding pseudogene (Riggs, 2002; Laurent et al., 2010).
The same regulatory network that inhibits transposable element movement likely
induces DNA methylation on pseudogenes. Methylation of pseudogenes is elevated in
embryos, likely as a mechanism for preventing the spread of retrotransposible elements
during embryogenesis (Eckhardt et al., 2006; Zhang et al., 2006). In plants, the
inactivation of methyltransferases resulted in the widespread activation of transposable
elements and pseudogenes (Zhang et al., 2006), demonstrating that DNA methylation is
sufficient to prevent the activation of pseudogenes. In humans, pseudogenes are highly
methylated, presumably to prevent transposition (Eckhardt et al., 2006).
In order to better understand how methylation patterns are established and
maintained on pseudogenes, we examined the methylation status of four pseudogenes
derived from the translational enhancer LIN28. This gene is involved in early
development and acts as a switch for developmental timing (Moss et al., 1997). LIN28
can also act as a reprogramming factor in the production of induced pluripotent stem cells
(Yu et al., 2007; Darr and Benvenisty, 2009).
The protein-coding region of LIN28 contains a high concentration of CpG sites,
making the gene a potential target for DNA methylation once inserted elsewhere in the
genome as a pseudogene. LIN28 has given rise to 10 processed pseudogenes within the
bovine genome that vary in length between 100 to 4000 bp. By measuring the
methylation levels of selected pseudogenes and the gene of origin, we sought to
determine whether the same regulatory mechanism that maintains methylation of the
96
LIN28 parent gene also controls the methylation status of LIN28 pseudogenes.
Additionally we examined the expression of genes near the insertion site to determine
whether pseudogene methylation is involved in transcriptional control of adjacent genes.
Materials and Methods
Fibroblast Cell Culture
Bovine fibroblasts isolated from skin were cultured in DMEM F12 (Thermo
Scientific HyClone Laboratories, Logan, UT) supplemented with 15% fetal bovine serum
(FBS) (Thermo Scientific HyClone Laboratories), 100 U/ml penicillin, and 100 mg/ml
streptomycin. Cells were cultured at 37°C with 5% CO2. For cell collection fibroblasts
were treated with 0.25% trypsin prior to collection for RNA and DNA isolation.
Oocyte Maturation
Bovine ovaries were collected at a local abattoir (E.A. Miller, Hyrum, UT) and
used for collecting oocytes. Oocytes with 3 to 8 mm follicles were aspirated along with
cumulus complexes. Following aspiration, cumulus oocyte complexes were cultured at
37°C with 5% CO2 for 18 to 22 hr in TCM 199 maturation medium containing 10% FBS
(Thermo Scientific HyClone Laboratories), 0.05 mg/ml follicle stimulating hormone, 5
mg/ml luteinizing hormone, 100 U/ml penicillin, and 100 mg/ml streptomycin.
In Vitro Fertilization
Following 18 to 22 hr of oocyte maturation, cryopreserved bovine semen
(Hoffman AI, Logan, UT) was thawed in a 37°C water bath. Live sperm were isolated by
centrifugation through a 45%/90% percoll gradient. Sperm were suspended in Tyrode’s
97
albumin lactate pyruvate (TALP) and used for oocyte fertilization. Twenty-four hr post
fertilization, cumulus cells were removed by vortexing the cumulus oocyte complex in
phosphate buffered saline (PBS) containing 0.32 mM sodium pyruvate, 5.55 mM
glucose, 3 mg/ml BSA, and 10 mg/ml hyaluronidase. Oocytes were washed through six
drops of PBS and placed in co-culture dishes plated with cultured cumulus cells and
cultured in CR2 medium.
Tissue Collection and RNA Isolation
Twenty-five pooled blastocyst embryos were collected after 8 days of culture in
CR2 medium, snap frozen with liquid nitrogen, and stored at -80°C. Twenty-five pooled
oocytes were collected after 22 hr of maturation, vortexed for 5 min in PBS to remove
cumulus cells, snap frozen with liquid nitrogen, and stored at -80°C. RNA was isolated
from oocyte and blastocyst embryo samples using the RNeasy Mini Kit (Qiagen,
Germantown, MD) following manufacturer’s instructions. Bovine brain, liver, testes
tissue samples were collected immediately after slaughter and suspended in RNALater
(Ambion, Austin, TX). Samples were stored overnight at 4°C. Fibroblasts were
collected as described above. RNA was isolated from brain, liver, testes, and fibroblasts
using TRIzol reagent (Life Technologies, Carlsbad, CA). Tissue samples were
homogenized in 3 ml TRIzol with a tissue homogenizer. Cells were incubated for 5 min
at room temperature, combined with 0.6 ml chloroform, and mixed by inversion.
Samples were centrifuged at 12,000 × g for 15 min at 4°C, the upper aqueous phase was
removed and combined with 1.5 ml isopropyl alcohol, then centrifuged at 12,000 × g for
10 min at 4°C. Supernatant was removed and the RNA was washed with 75% ethanol
98
and centrifuged again, then dried and resuspended in H2O. Isolated RNA was
immediately converted to cDNA using the Superscript III Reverse Transcriptase Kit (Life
Technologies) following manufacturer’s protocol. Samples were stored at -20°C until
use.
DNA Isolation and Bisulfite Conversion
Twenty-five pooled oocytes and blastocyst embryos were snap frozen with liquid
nitrogen and stored at -80°C until direct bisulfite conversion. DNA from brain, liver,
testes, and fibroblast samples were collected from the interphase of the TRIzol treatment
used for RNA collection. The collected interphase was combined with 0.9 ml ethanol,
mixed by inversion, and centrifuged at 2000 × g for 5 min at 4°C. The pellet was
suspended with 75% ethanol and washed by centrifugation three times. The DNA pellet
was resuspended in 0.1 µM sodium citrate. The isolated DNA underwent bisulfite
conversion using the EZ DNA Methylation Kit (Zymo Research, Irvine, CA) according
to manufacturer’s recommendation. DNA was stored at -20°C until use.
Bisulfite PCR and 454 Sequencing
Primers for bisulfite-converted DNA were designed for each pseudogene and the
LIN28 parent gene. Primers covered CpG sites within and immediately surrounding the
protein-coding sequence of LIN28. As the four examined pseudogenes maintain high
sequence identity to the LIN28 parent gene and one another, primers were designed
specific to each pseudogene to ensure only amplification of the target pseudogene.
Bisulfite-converted DNA was used as a template for 25 µl PCR reactions using 1 to 3 µl
99
DNA. Reactions were carried out using the Mastercycler thermal cycler (Eppendorf,
New York, NY). Primer concentrations were 0.6 µM for all reactions. All reactions
were carried out using the following cycling parameters: 30 cycles of 94°C for 1 min,
annealing temperature ranging from 54 to 59°C for 1 min, and 72°C for 1 min for 30
cycles. A second PCR was preformed using primers including the original primer
sequence with the addition of the 454 adapter sequence, key sequence, and molecular
identification tags (designated as N) to differentiate individual tissue samples (adapter A
CGTATCGCCTCCCTCGCGCCATCAGNNNNNNNNNN attached to the forward
primer and adapter B CTATGCGCCTTGCCAGCCCGCTCAGNNNNNNNNNN
attached to the reverse primer). The following cycling parameters were used: 94°C for
30 sec, annealing temperature ranging from 55 to 60°C for 30 sec, and 72°C for 30 sec
for 15 cycles. The second reaction used 1 µl of PCR product from the first PCR and 0.3
µM primer. All PCR samples used GoTaq Green Master Mix (Promega, Madison, WI).
All PCR reactions were carried out on a Mastercycler thermal cycler (Eppendorf). PCR
reactions were purified with AMPure beads (Beckman Coulture, Brea, CA) and
quantified using PicoGreen (Life Technologies). PCR product was sequenced using the
454 GD FLX Titanium DNA sequencer (Roche, Indianapolis, IN). Amplicon libraries
were subjected to emulsion PCR to generate DNA-coated beads, loaded onto a
PicoTiterPlate, and sequenced with a FLX Titanium DNA sequencing Kit according to
the manufacturer’s protocol.
100
End-Point RT-PCR
End-point RT-PCR reactions targeting bovine cDNA were carried out using the
Mastercycler thermal cycler (Eppendorf). GoTaq Green Master Mix (Promega) was used
on all reactions with primer concentrations of 0.6 µM using the following cycling
parameters: 40 cycles of 94°C 15 sec, annealing temperature ranging from 58 to 60°C for
15 sec, and 72°C for 30 sec for 35 cycles. Following cycling, samples were
electrophoresed on a 1% agarose gel and imaged.
Sequence Analysis
Sequencing data was analyzed to generate the frequency of methylation using
BISMA analysis software (Rohde et al., 2010). Data are expressed as the percentage of
methylated CpG sites by normalizing methylation frequency data to the number of
sequence reads obtained.
Results
We assessed the methylation status of four LIN28 processed pseudogenes in six
bovine tissue samples, including brain, liver, testes, fibroblast cells, IVF blastocyst stage
embryo, and oocyte. Three of the LIN28 pseudogenes contain the entire protein-coding
sequence as well as transcribed regions up and downstream from the protein-coding
region. These three pseudogenes are likely the products of retrotransposition of the
LIN28 transcript following intron excision (Fig. 3-1). A fourth pseudogene contains only
the terminal portion of the protein-coding region and downstream transcript (Fig. 3-1).
These four pseudogenes were selected for examination based on their retention of the
101
protein-coding region, as well as insertion location and number of CpG sites. The
pseudogenes examined were as follows: LIN28P-Ch:3 (LOC784466) inserted into
chromosome 3 approximately 20Kbp upstream of the UBQLN4 gene (84% identity with
LIN28), LIN28P-Ch:7 (LOC781442) on chromosome 7 inserted into the fourth intron of
the MAN2A1 gene (97% identity with LIN28) and LIN28P-Ch:26 (LOC539705) on
chromosome 26 inserted in the first intron of ACADSB (97% identity with LIN28) (Fig.
3-2). In addition, an identical pair of processed pseudogenes LIN28P-Ch28 (LOC785096
and LOC785075) on chromosome 28 was included; these pseudogenes have no proximity
to any gene (90% identity with the LIN28 gene) (Fig. 3-2). The LIN28P-Ch:28
pseudogenes likely arose from the insertion of a single pseudogene into chromosome 28,
followed by a duplication event that generated a second copy of the initial pseudogene.
The two pseudogenes are nearly 100% identical to one another, and differ by only three
nucleotides and cannot be distinguished from one another. Primer amplification was
indiscriminate of either pseudogene and analysis was carried out as though it were a
single targeted site.
Primer sets were designed to profile each individual pseudogene for a sequence
within the protein-coding region. Due to the high sequence identity between the
pseudogenes and parent gene, the overlap of examined CpG sites was limited by our
ability to amplify each individual pseudogene. Pseudogenes LIN28P-Ch:3, LIN28P-Ch:7
and LIN28P-Ch:26 all contain considerable overlap between identical CpG sites.
LIN28P-Ch:26 contains overlap of only two CpG sites with the LIN28 parent gene. The
LIN28 parent gene also shares only three overlapping CpG sites with LIN28P-Ch:28 (Fig.
102
3-3). Because of the restricted overlap of CpG specific sites, our assessment regarding
differences between individual CpG dinucleotides was limited. Rather we focused our
analysis to consider patterns of methylation among the amplified regions as a whole.
Methylation Status of the LIN28 Family
Our results indicate that LIN28 and all examined pseudogenes were methylated in
all bovine cell and tissue types examined. Seven CpG dinucleotides within the LIN28
parent gene were analyzed. These dinucleotides are located in the third exon of LIN28.
All of the CpG sites were methylated in each tissue to varying degrees (Fig. 3-3). In
addition, all pseudogenes examined were methylated in all tissue samples. However,
each pseudogene demonstrated a distinct methylation pattern that differs sharply from the
LIN28 gene, and that correlates to the pseudogene insertion location.
The pseudogenes LIN28P-Ch:3 and LIN28P-Ch:28 were inserted into a location
not associated with any gene, and both of these pseudogenes share a similar overall
pattern of methylation (Fig. 3-3). We analyzed methylation at nine CpG dinucleotides
within LIN28P-Ch:3 and eight CpG dinucleotides in LIN28P-Ch:28. In both
pseudogenes, oocytes had the lowest levels of overall methylation in these pseudogenes
with 52% average methylation across all CpG sites for LIN28P-Ch:3 and 41% for
LIN28P-Ch:28. In contrast, average methylation was highest in blastocysts, at 81% for
LIN28P-Ch:3 and 92% for LIN28P-Ch:28. The methylation patterns of both
pseudogenes apparently deviated from patterns observed in the LIN28 parent gene.
Within the LIN28 parent gene, oocytes had the highest frequency of methylation with an
average of 93% across all CpG sites examined. Alternatively, the methylation of the
103
parent gene was lowest in blastocyst embryos, with only 56% of CpG sites carrying the
methyl mark (Fig. 3-3). In LIN28P-Ch:28 we observed three overlapping CpG
dinucleotides with the LIN28 parent gene. Although limited, within these three
overlapping sites, the inverse pattern of methylation between the LIN28 parent gene and
pseudogenes LIN28P-Ch:3 and LIN28P-Ch:28 is maintained.
In contrast to the tissue-specific methylation patterns, pseudogenes LIN28P-Ch:7
and LIN28P-Ch:26 both have a hypermethylated pattern. Both LIN28P-Ch:7 and
LIN28P-Ch:26 were inserted into gene bodies. LIN28P-Ch:7 is inserted into the fourth
intron of the gene MAN2A1. Over 90% of the 28 CpG sites inspected within this
pseudogene were methylated in all tissue samples. The LIN28P-Ch:26 pseudogene lies
within the first intron of the gene ACADSB and is 20 kbp from the gene IKZF5. As for
LIN28P-Ch:7, all 28 CpG sites examined for the LIN28P-Ch:26 pseudogene were highly
methylated at greater than 90% frequency (Fig. 3-3). The hypermethylation observed in
both of these pseudogenes is in sharp contrast to the moderate levels of methylation of
the LIN28 parent gene and pseudogenes not inserted into gene bodies, none of which
were as consistently hypermethylated for all tissue samples.
These apparent differences in methylation correlate to the location of insertion of
the pseudogene. LIN28 pseudogenes inserted into gene bodies were hypermethylated.
Alternatively, LIN28 pseudogenes with an insertion location distant from a gene varied in
the pattern of methylation in a tissue-specific manner, although these tissue-specific
patterns were inverse with respect to the LIN28 parent gene (Fig. 3-3).
104
Single CpG Demethylation
Both LIN28P-Ch:7 and LIN28P-Ch:26 contain a single CpG dinucleotide that is
characterized by a low frequency of methylation in a tissue-specific manner. The
LIN28P-Ch:7 CpG dinucleotide at site 23 (Fig. 3-3) is under-methylated in three tissues.
Methylation of this site is only 45% in the testes and blastocyst, whereas this site is
devoid of any methylation in fibroblast cells. The same CpG dinucleotide 23 was
additionally measured within the LIN28P-Ch:3 sequence, yet this site retained a high
level of methylation for all tissues examined (Fig. 3-3).
LIN28P-Ch26 contains a similar CpG dinucleotide that is undermethylated in a
tissue-specific manner at CpG site 16 in IVF blastocyst and oocyte samples. Both
samples are hypomethylated with a 13% methylation frequency in IVF blastocyst and 5%
frequency in oocytes, yet highly methylated for other tissue types (Fig. 3-3).
Expression of Genes Adjacent to LIN28 Pseudogenes
End-point RT-PCR was performed to examine expression of LIN28 and genes
closely adjacent to the four LIN28 pseudogenes examined to determine whether
methylation of pseudogenes correlated with differences in gene expression. Expression
of the genes DHDDS, MAN2A1, ACADSB, and IKZF5 were measured in testes, liver,
brain, fibroblast, IVF blastocyst, and oocyte samples (Fig. 3-4). DHDDS, MAN2A1, and
ACADSB expression did not correspond to methylation patterns of adjacent pseudogenes
(Figs. 3-3 and 3-4). IKZF5 is proximal to LIN28P-Ch:26, which does contain a tissuespecific methylation pattern of a single CpG site. Interestingly, the expression pattern for
IKZF5 mirrors the pattern of methylation in that this gene is not expressed in oocytes or
105
blastocyst embryos (corresponding to samples with low methylation at site 16), but
highly expressed in other tissue types (corresponding to samples with high methylation at
site 16) (Figs. 3-3 and 3-4).
Discussion
The objective of our study was to determine whether the methylation of
pseudogenes followed the same patterns as the functional parent gene by examining
LIN28 as a case study. An observation that the LIN28 pseudogenes maintained an
identical or highly similar methylation pattern as the LIN28 parent gene would suggest
that regulation of LIN28 methylation was specific to the sequence of the gene and that
methylation of pseudogenes with high sequence identity was maintained by the same
mechanism(s) that maintained methylation of the LIN28 parent gene. Alternatively, an
observation that pseudogene methylation patterns deviated from the parent gene would
indicate that pseudogenes are subject to local regulation of methylation patterns. In this
study, we observed that LIN28 pseudogenes do not recapitulate the same methylation
status as LIN28, but rather appear to acquire methylation patterns independent of the
parent gene. Furthermore, we observed that methylation levels of the examined
pseudogenes correlate to the location of insertion. LIN28 pseudogenes inserted into gene
bodies were hypermethylated in all tissues examined. In contrast, pseudogenes inserted
into genomic regions that are not proximal to genes had reduced overall methylation and
the methylation pattern was dependent upon the tissue type. Additionally, pseudogenes
not associated with genes had less methylation in tissue samples that were highly
methylated in the parent gene.
106
We measured methylation of seven CpG dinucleotides located within the third
exon of the LIN28 gene. Methylation of CpG dinucleotides within the gene body is
generally associated with transcriptionally active genes (Rauch et al., 2009; Laurent et al.,
2010). This pattern is in contrast to methylation of the 5’ upstream and promoter regions
of genes, which are typically associated with transcriptional silencing. Methylation of
these seven CpG dinucleotides was dependent upon tissue type. Notably, oocytes had the
highest levels of methylation within the LIN28 parent gene. Oocytes generally maintain
low levels of global methylation relative to somatic cells (Razin et al., 1984; Monk et al.,
1987). Following fertilization, global methylation levels decline further and are then
reestablished during embryonic and somatic cell development (Mayer et al., 2000;
Oswald et al., 2000; Hajkova et al., 2002; Morgan et al., 2005). The methylation level
observed in the LIN28 parent gene is counter to this pattern. However, methylation levels
measured in the LIN28P-Ch:3 and LIN28P-Ch:28 pseudogenes are consistent with
changes in methylation levels on a global level. Within both these pseudogenes, oocytes
maintained low levels of methylation, and somatic cells maintained high levels of
methylation. It is possible that the methylation patterns observed for LIN28P-Ch:3 and
LIN28P-Ch:28 are maintained by the same mechanism that maintains global levels of
methylation, while methylation of the LIN28 parent gene is maintained by a separate
mechanism.
We also observed that methylation of pseudogenes depends on the genomic
context into which the pseudogene was inserted. Both pseudogenes inserted into a gene,
LIN28P-Ch:7 and LIN28P-Ch:26, are hypermethylated in all tissues. On the other hand,
107
both pseudogenes not associated with a gene, LIN28P-Ch:3 and LIN28P-Ch:28, vary in
the pattern of methylation within different tissue samples, but tissue methylation levels
were similar between the two pseudogenes. The similarity of methylation levels among
all tissue types between similar pseudogenes suggests that pseudogenes share a common
regulatory mechanism that establishes and maintains the methylation signature.
Within both pseudogenes inserted into gene bodies, we observed single CpG
dinucleotides that exhibit demethylation in a tissue specific manner. Pseudogene
LIN28P-Ch:7 was inserted into the fourth intron of MAN2A1. Within the investigated
CpG dinucleotides, a single CpG site at position 23 demonstrated a sharp reduction in
methylation in the testes and blastocysts (both of which have 45% methylation), as well
as fibroblasts (0% methylation). The same CpG site measured in LIN28P-Ch:3 had no
such reduction in methylation observed in any tissue studied, so it is likely that the
apparent decline in methylation was specific to this pseudogene. This hypomethylated
site exists within the context of three closely aligned CpG dinucleotides, both of which
are highly methylated in all tissues. This site- and tissue-specific decrease in methylation
points to a possible deliberate regulation of methylation at this locus that is unexplained
in our study.
The LIN28P-Ch:26 pseudogene is located within the first intron of ACADSB and
20 Kbp upstream from IKZF5. Within the LIN28P-Ch:26 pseudogene, a single CpG
dinucleotide at site 16 is hypomethylated in the oocyte and blastocyst samples, a sharp
contrast to the hypermethylated status of all surrounding CpG dinucleotides as well as the
same CpG site in the four remaining tissues. IKZF5 was not expressed in either oocytes
108
or blastocyst embryos. Although the expression pattern of IKZF5 was correlated with the
tissue-specific methylation patterns observed for the pseudogene, it is doubtful that this
single CpG site is involved in the transcriptional regulation of IKZF5. Although
demethylation has been associated with transcriptional silencing (Hark et al., 2000; Lim
and van Oudenaarden, 2007), the involvement of this specific CpG site in regulation of
IKZF5 is unlikely, as the site is removed from the transcriptional start site by 20kbp.
In conclusion, the four LIN28 pseudogenes examined in this study maintained
methylation patterns that differed from the parent gene and from one another according to
their location of insertion. New knowledge on the regulation of pseudogenes via DNA
methylation could contribute to greater understanding of the maintenance of global
and/or regional patterns of methylation. Future work on this topic should focus on
defining methylation patterns for other pseudogene families to determine whether all
pseudogenes are maintained in a similar manner or whether sequence specific patterns
can be identified through analysis of pseudogenes. Additionally, future studies should
investigate expression of pseudogenes to determine how patterns of methylation correlate
with transcript levels.
References
Bartolomei MS, Webber AL, Brunkow ME, Tilghman SM. 1993. Epigenetic mechanisms
underlying the imprinting of the mouse H19 gene. Genes Dev 7(9):1663-1673.
Bird A. 2002. DNA methylation patterns and epigenetic memory. Genes Dev 16(1):6-21.
Chen RZ, Pettersson U, Beard C, Jackson-Grusby L, Jaenisch R. 1998. DNA
hypomethylation leads to elevated mutation rates. Nature 395(6697):89-93.
109
Csankovszki G, Nagy A, Jaenisch R. 2001. Synergism of Xist RNA, DNA methylation,
and histone hypoacetylation in maintaining X chromosome inactivation. J Cell
Biol 153(4):773-784.
Darr H, Benvenisty N. 2009. Genetic analysis of the role of the reprogramming gene
LIN-28 in human embryonic stem cells. Stem Cells 27(2):352-362.
De Smet C, De Backer O, Faraoni I, Lurquin C, Brasseur F, Boon T. 1996. The activation
of human gene MAGE-1 in tumor cells is correlated with genome-wide
demethylation. Proc Natl Acad Sci USA 93(14):7149-7153.
De Smet C, Lurquin C, Lethe B, Martelange V, Boon T. 1999. DNA methylation is the
primary silencing mechanism for a set of germ line- and tumor-specific genes
with a CpG-rich promoter. Mol Cell Biol 19(11):7327-7335.
Eckhardt F, Lewin J, Cortese R, Rakyan VK, Attwood J, Burger M, Burton J, Cox TV,
Davies R, Down TA, Haefliger C, Horton R, Howe K, Jackson DK, Kunde J,
Koenig C, Liddle J, Niblett D, Otto T, Pettett R, Seemann S, Thompson C, West
T, Rogers J, Olek A, Berlin K, Beck S. 2006. DNA methylation profiling of
human chromosomes 6, 20 and 22. Nat Genet 38(12):1378-1385.
Goll MG, Bestor TH. 2005. Eukaryotic cytosine methyltransferases. Annu Rev Biochem
74:481-514.
Grunau C, Hindermann W, Rosenthal A. 2000. Large-scale methylation analysis of
human genomic DNA reveals tissue-specific differences between the methylation
profiles of genes and pseudogenes. Hum Mol Genet 9(18):2651-2663.
Hajkova P, Erhardt S, Lane N, Haaf T, El-Maarri O, Reik W, Walter J, Surani MA. 2002.
Epigenetic reprogramming in mouse primordial germ cells. Mech Dev 117(12):15-23.
Hark AT, Schoenherr CJ, Katz DJ, Ingram RS, Levorse JM, Tilghman SM. 2000. CTCF
mediates methylation-sensitive enhancer-blocking activity at the H19/Igf2 locus.
Nature 405(6785):486-489.
Holliday R, Pugh JE. 1975. DNA modification mechanisms and gene activity during
development. Science 187(4173):226-232.
Illingworth R, Kerr A, Desousa D, Jorgensen H, Ellis P, Stalker J, Jackson D, Clee C,
Plumb R, Rogers J, Humphray S, Cox T, Langford C, Bird A. 2008. A novel CpG
island set identifies tissue-specific methylation at developmental gene loci. PLoS
Biol 6(1):e22.
Jaenisch R, Bird A. 2003. Epigenetic regulation of gene expression: how the genome
integrates intrinsic and environmental signals. Nat Genet 33 Suppl:245-254.
110
Jones PA, Liang G. 2009. Rethinking how DNA methylation patterns are maintained. Nat
Rev Genet 10(11):805-811.
Laurent L, Wong E, Li G, Huynh T, Tsirigos A, Ong CT, Low HM, Kin Sung KW,
Rigoutsos I, Loring J, Wei CL. 2010. Dynamic changes in the human methylome
during differentiation. Genome Res 20(3):320-331.
Li E, Beard C, Jaenisch R. 1993. Role for DNA methylation in genomic imprinting.
Nature 366(6453):362-365.
Li E, Bestor TH, Jaenisch R. 1992. Targeted mutation of the DNA methyltransferase
gene results in embryonic lethality. Cell 69(6):915-926.
Lim HN, van Oudenaarden A. 2007. A multistep epigenetic switch enables the stable
inheritance of DNA methylation states. Nat Genet 39(2):269-275.
Liu WM, Maraia RJ, Rubin CM, Schmid CW. 1994. Alu transcripts: cytoplasmic
localisation and regulation by DNA methylation. Nucleic Acids Res 22(6):10871095.
Mayer W, Niveleau A, Walter J, Fundele R, Haaf T. 2000. Demethylation of the zygotic
paternal genome. Nature 403(6769):501-502.
Mohandas T, Sparkes RS, Shapiro LJ. 1981. Reactivation of an inactive human X
chromosome: evidence for X inactivation by DNA methylation. Science
211(4480):393-396.
Monk M, Boubelik M, Lehnert S. 1987. Temporal and regional changes in DNA
methylation in the embryonic, extraembryonic and germ cell lineages during
mouse embryo development. Development 99(3):371-382.
Morgan HD, Santos F, Green K, Dean W, Reik W. 2005. Epigenetic reprogramming in
mammals. Hum Mol Genet 14 Spec No 1:R47-58.
Moss EG, Lee RC, Ambros V. 1997. The cold shock domain protein LIN-28 controls
developmental timing in C. elegans and is regulated by the lin-4 RNA. Cell
88(5):637-646.
Okano M, Bell DW, Haber DA, Li E. 1999. DNA methyltransferases Dnmt3a and
Dnmt3b are essential for de novo methylation and mammalian development. Cell
99(3):247-257.
Oswald J, Engemann S, Lane N, Mayer W, Olek A, Fundele R, Dean W, Reik W, Walter
J. 2000. Active demethylation of the paternal genome in the mouse zygote. Curr
Biol 10(8):475-478.
111
Rauch TA, Wu X, Zhong X, Riggs AD, Pfeifer GP. 2009. A human B cell methylome at
100-base pair resolution. Proc Natl Acad Sci USA 106(3):671-678.
Razin A, Webb C, Szyf M, Yisraeli J, Rosenthal A, Naveh-Many T, Sciaky-Gallili N,
Cedar H. 1984. Variations in DNA methylation during mouse cell differentiation
in vivo and in vitro. Proc Natl Acad Sci USA 81(8):2275-2279.
Riggs AD. 2002. X chromosome inactivation, differentiation, and DNA methylation
revisited, with a tribute to Susumu Ohno. Cytogenet Genome Res 99(1-4):17-24.
Rohde C, Zhang Y, Reinhardt R, Jeltsch A. 2010. BISMA--fast and accurate bisulfite
sequencing data analysis of individual clones from unique and repetitive
sequences. BMC Bioinformatics 11:230.
Shiota K. 2004. DNA methylation profiles of CpG islands for cellular differentiation and
development in mammals. Cytogenet Genome Res 105(2-4):325-334.
Stoger R, Kubicka P, Liu CG, Kafri T, Razin A, Cedar H, Barlow DP. 1993. Maternalspecific methylation of the imprinted mouse Igf2r locus identifies the expressed
locus as carrying the imprinting signal. Cell 73(1):61-71.
Walsh CP, Chaillet JR, Bestor TH. 1998. Transcription of IAP endogenous retroviruses is
constrained by cytosine methylation. Nat Genet 20(2):116-117.
Woodcock DM, Lawler CB, Linsenmeyer ME, Doherty JP, Warren WD. 1997.
Asymmetric methylation in the hypermethylated CpG promoter region of the
human L1 retrotransposon. J Biol Chem 272(12):7810-7816.
Yoder JA, Walsh CP, Bestor TH. 1997. Cytosine methylation and the ecology of
intragenomic parasites. Trends Genet 13(8):335-340.
Yu J, Vodyanik MA, Smuga-Otto K, Antosiewicz-Bourget J, Frane JL, Tian S, Nie J,
Jonsdottir GA, Ruotti V, Stewart R, Slukvin, II, Thomson JA. 2007. Induced
pluripotent stem cell lines derived from human somatic cells. Science
318(5858):1917-1920.
Zhang X, Yazaki J, Sundaresan A, Cokus S, Chan SW, Chen H, Henderson IR, Shinn P,
Pellegrini M, Jacobsen SE, Ecker JR. 2006. Genome-wide high-resolution
mapping and functional analysis of DNA methylation in arabidopsis. Cell
126(6):1189-1201.
112
Table 3-1. Primers Used for Bisulfite Sequencing and End-Point RT-PCR
Bisulfite Primers
LIN28P-Ch:7
LIN28P-Ch:28
LIN28P-Ch:3
LIN28P-Ch:26
LIN28
Forward
Reverse
Forward
Reverse
Forward
Reverse
Forward
Reverse
Forward
Reverse
End-Point RT-PCR Primers
MAN2A1
Forward
Reverse
ACADSB
Forward
Reverse
IKZF5
Forward
Reverse
DHDDS
Forward
Reverse
LIN28
Forward
Reverse
ACTB
Forward
Reverse
GAAAAGTTTATGAAAGAGGTTAGGGTT
AAAACCCTCCATATACAACTTACTC
TTTTAAGAAGTTTATTAAGGGTTTGGAA
TAACCCCCACCCACTATAACTTAAT
TGTATATAATGGGGAGTAGGGGTT
CAAAAACCCTCCAAATACAACTTAC
AAAGATAGTGGTGATATTGTGT
CCCAAAAACCAATAACTCTAA
TTTTTTTAGAGTAAGTTGTATATGGAGGG
TTACCAAAAACCACAAACTTCACTT
GACCCATTTGGACATTCACC
TTTAGGATCAGGCCCACAAG
TGGAAAACTCCTCCTCATGC
CAACTTGTTCCTGGGCAAAT
GAGAAGAAACCGGAGCCTTT
AGGTCCTTTCAAACCCGTCT
GTCATTTTGGGAGCGGTTCT
GCCAGTTTGTTGAAGCCTTG
TGCAGAAACGCAGATCAAAG
TTCTTCCTCCTCCCGAAAGT
ATGGGCCAGAAGGACTCGTA
CTTCTCCATGTCGTCCCAGT
113
Chromosome%
Start%%
Codon%
Stop%
Codon%
Iden?ty%
2%
LIN28&
26%
LIN28P(Ch:26&
97%
7%
LIN28P(Ch:7&
97%
28%
LIN28P(Ch:28&
90%
28%
LIN28P(Ch:28&
90%
3%
84%
2%
LIN28P(Ch:3&
LOC100295351%
7%
N/A%
99%
3%
N/A%
96%
29%
N/A%
86%
3%
N/A%
84%
100%
85%
!150%
350%
850%
1350%
1850%
2350%
2850%
3350%
3850%
Figure 3-1. The ten processed pseudogenes of LIN28 aligned to the LIN28 processed
transcript. The horizontal filled in lines represent each pseudogene with the location
relative to the start of the LIN28 protein-coding region indicated at the bottom and
measured in base pairs. Vertical lines represents start and stop codons. Sequence
identity of each pseudogene to the LIN28 parent transcript is indicated on the right
column and the chromosome of insertion is indicated on the left. The dashed box shows
the pseudogenes examined for methylation analysis
114
Figure 3-2. LIN28 pseudogene locations on each chromosome. Genes are represented
by boxes with horizontal lines indicating approximate exon start locations. Pseudogene
location is indicated by the darkened box. Arrows indicate gene transcription direction.
115
Figure 3-3. Heat map of LIN28 and LIN28 pseudogene methylation profile.
(following page) LIN28 and pseudogenes include the examined tissues: testes, liver,
brain, fibroblast, IVF blastocyst embryos, and oocytes. The total number of sequencing
reads is indicated next to each sample. The average percentage methylation of CpG sites
within the entire amplicon is shown for each pseudogene. Each heat map cell indicates
the percentage of CpG sites positive for methylation. Red indicates high methylation and
blue low methylation. Each CpG site corresponds to a number indicated above each
sample that indicates the placement of each CpG site in relation to the protein-coding
region. CpG sites of pseudogenes that do not correspond to a CpG location within the
LIN28 parent gene are represented by ‘n’.
116
117
B-Actin
IKZF5
DHDDS
MAN2A1
ACADSB
Lin28
Testes
Liver
Brain
Fibroblast
IVF Blastocyst
Oocyte
Figure 3-4. End-point RT-PCR. Expression of genes LIN28, ACADSB, MAN2A1,
DHDDS, IKZF5 and ACTB in the testes, liver, brain, fibroblast, IVF blastocyst embryos
and oocytes.
118
CHAPTER 4
ACTIVATION OF THE P53 PATHWAY FOLLOWING BOVINE
SOMATIC CELL NUCLEAR TRANSFER
Abstract
Embryos produced via somatic cell nuclear transfer have impaired development,
reduced implantation rates, and have reduced viability as neonates. Failure of cloned
organisms to thrive is largely believed to be the product of failed somatic cell
reprogramming. Evidence points to a role for the P53 pathway in modulating
reprogramming of the genome to a pluripotent state. The tumor suppressor P53
maintains genome integrity by activating apoptosis and inducing cell cycle arrest in
response to cell stress or oncogene expression. Activation of the P53 pathway is contrary
to the needs of a pre-implantation embryo, which should express numerous genes
responsible for cell growth and proliferation. The aim of this study was to determine the
extent of apoptosis in cloned bovine embryos throughout early development. Our results
indicate that the P53 pathway is activated in cloned embryos. Additionally we observed
that the cell cycle arrest activator P21WAF1/Cip1 (CDKN1A) was commonly expressed in
cloned embryos, suggesting that P53 also induces cell cycle arrest in SCNT embryos.
Activation of the P53 pathway may be the result of upregulation of the reprogramming
factors C-MYC, SOX2, LIN28, and KLF4, as all of these genes are upregulated in cloned
embryos. These observations suggest that SCNT embryos are susceptible to blastomere
depletion via apoptosis and cell cycle arrest in response to expression of reprogramming
factors.
Introduction
119
Somatic cell nuclear transfer (SCNT) is an established method for the production
of transgenic animals and reproduction of domestic species with phenotypically valuable
traits. The procedure utilizes an enucleated oocyte as a recipient for a donor nucleus.
Cloning has been successfully performed in a number of species. Nevertheless, the
benefits of cloning are hindered due to the limited success SCNT development (Wells et
al., 1999). Reconstructed SCNT embryos suffer from high rates of embryo failure and
increased rates of fetal and neonate death. The success rate of reconstructed oocytes
resulting in live births is less than 3% (Thibault, 2003), and the low success limits the
widespread use of the technique. The low success rate is thought to be primarily the
result of incomplete reprogramming of the donor cell genome (Wilmut et al., 1998;
Colman, 1999; Kikyo and Wolffe, 2000; Solter, 2000; Dean et al., 2001; Humpherys et
al., 2001; Reik et al., 2001; Niemann et al., 2008). However, studies also point to high
numbers of apoptotic cells in embryos as another factor that contributes to
preimplantation and early gestational failure (Fahrudin et al., 2002; Hao et al., 2003; Jang
et al., 2004; Gjorret et al., 2005; Im et al., 2006; Kumar et al., 2007).
Programmed cell death by apoptosis is characterized by the morphological
features of membrane blebbing, cell shrinkage, nuclear condensation, and DNA
fragmentation (Kerr et al., 1972). Apoptosis plays an active role during early embryonic
development (Jurisicova et al., 1998; Hardy et al., 2001; Liu et al., 2002; Neuber et al.,
2002; Gjorret et al., 2003; Jurisicova et al., 2003). Within preimplantation embryos,
apoptosis is responsible for the removal of blastomeres damaged by exogenous stress,
120
carrying genetic abnormalities or that have reduced developmental potential (Fabian et
al., 2005).
Apoptosis in preimplantation embryos has been detected in multiple species and
has been observed at higher frequency in embryos produced by SCNT than those
produced by in vitro fertilization (IVF) (Fahrudin et al., 2002; Hao et al., 2003; Jang et
al., 2004; Gjorret et al., 2005; Im et al., 2006; Kumar et al., 2007). Cloned blastocysts
have lower overall cell numbers than IVF embryos, an observation consistent with
elevated levels of apoptosis in SCNT embryos (Koo et al., 2000; Fahrudin et al., 2002;
Gjorret et al., 2005; Terashita et al., 2011). The frequency of apoptotic cells and the cell
count within preimplantation embryos are important indicators of developmental
potential (Brison and Schultz, 1997; Brison and Schultz, 1998; Koo et al., 2000; Knijn et
al., 2003). Embryos exposed to factors that increase apoptosis have reduced implantation
rates and decreased fetal birth weight (Wuu et al., 1999). These observations suggest that
increased apoptosis in early developing embryos can negatively affect post-implantation
development.
Apoptosis is regulated by P53, which is activated by cell stress, DNA damage,
and oncogene expression. Upon activation, P53 can induce pathways leading to
senescence, cell cycle arrest, or apoptosis (Vousden and Lu, 2002). Upon activation, P53
translocates to the nucleus where it then induces expression of numerous genes that can
either inhibit cell cycle progression or initiate programmed cell death. In the event of cell
cycle arrest, P21WAF1/Cip1 (CDKN1A) is transcriptionally upregulated by P53 (Brugarolas
et al., 1999). P21 is an inhibitor of cyclin-dependent kinases (CDKs) that are required for
cell cycle progression. Alternatively, P53 can initiate programmed cell death by
121
upregulating the expression of genes that initiate apoptosis. BAX is one such proapoptotic gene (Oltvai et al., 1993; Miyashita and Reed, 1995). BAX oligomerizes to
form a channel on the mitochondrial membrane that induces the release of cytochrome c
(Wei et al., 2001). In healthy cells, BAX is inhibited by BCL2, which forms a
heterodimer with BAX (Yin et al., 1994). Under normal conditions, levels of BCL2
exceed BAX levels. This favors BCL2:BAX heterodimer formation and prevents
activation of apoptosis. However upon upregulation of BAX by P53, BAX oligomers
become common, allowing for activation of the intrinsic apoptotic pathway via release of
cytochrome c from the mitochondria (Basu and Haldar, 1998).
Recently, the P53 pathway has been shown to modulate somatic cell
reprogramming. Reprogramming is known to activate P53-mediated apoptosis and cell
cycle arrest (Qin et al., 2007; Mali et al., 2008; Zhao et al., 2008). Knockout of P53
dramatically improves the efficiency of generating induced pluripotent stem cells (iPSCs)
(Zhao et al., 2008). P53 null cell lines manifest dramatically higher reprogramming
success rates (Hong et al., 2009; Kawamura et al., 2009). Several genes upregulated by
P53 prevent survival of reprogrammed cells (Hong et al., 2009). Among these is
P21WAF1/Cip1, an observation that suggests cell cycle arrest is an additional consequence of
somatic cell reprogramming. Cell lines that are recalcitrant to reprogramming can be
successfully reprogrammed following P53 inhibition, providing further evidence for a
role of P53 in somatic genome reprogramming (Utikal et al., 2009).
122
The observations that P53 plays a role in modulating reprogramming somatic
DNA to a pluripotent state for the generation of iPSCs has relevance for SCNT and the
high rate of failure for SCNT embryos. The donor cell genome must undergo somatic
cell reprogramming as a part of embryonic development. However, reprogramming the
somatic genome may activate P53 and trigger either cell cycle arrest or apoptosis. P53
accumulation within IVF embryos has been observed to act as a direct impediment to
embryo viability and implantation (Ganeshan et al., 2010). Decreased P53 expression led
to significantly higher rates of fetal development following transfer of IVF embryos (Li
et al., 2007). Furthermore, the number of apoptotic cells within an embryo was
negatively correlated with the quality and overall viability of the embryo (Hao et al.,
2004). Apoptosis has been observed to be higher in cloned embryos as measured by
DNA degradation and morphological changes associated with apoptotic cells (Fahrudin et
al., 2002; Hao et al., 2003; Jang et al., 2004; Gjorret et al., 2005; Im et al., 2006; Kumar
et al., 2007). However, previous work associating SCNT with apoptosis have not
measured the transcriptional activity of genes associated with the P53 pathway nor
determined the cause of increased apoptosis following SCNT.
The aims of this study were to determine the extent of P53-mediated apoptosis
within SCNT embryos and to ascertain the factors that activate apoptosis following
SCNT. P53 expression can alter the developmental trajectory of SCNT preimplantation
embryos and may function to remove inappropriately programmed or otherwise abnormal
blastomeres from contributing to SCNT embryo development. Our findings indicate that
123
P53-mediated apoptosis is activated within SCNT embryos and that activation is likely in
response to aberrant expression of reprogramming factors.
Materials and Methods
Oocyte Maturation and In Vitro Fertilization
Bovine ovaries were collected from a local abattoir (E.A. Miller, Hyrum, UT).
Cumulus-oocyte complexes were isolated from 3 to 8 mm follicles by aspiration using an
18-gauge needle. TCM 199 maturation medium (Thermo Scientific HyClone
Laboratories, Logan, UT) containing 10% FBS, 0.05 mg/ml follicle stimulating hormone,
5 mg/ml leutinizing hormone, 100 U/ml penicillin, and 100 mg/ml streptomycin was used
to culture aspirated cumulus-oocyte complexes for 18 to 22 hr. Following oocyte
maturation, cryopreserved bovine semen (Hoffman AI, Logan, UT) was thawed in a
water bath, and live sperm were segregated through centrifugation in a 45%/90% percoll
gradient. Collected live sperm were resuspended in Tyrode’s albumin lactate pyruvate
(TALP) and used to fertilize matured oocytes. Twenty-four hr post-fertilization cumulus
cells were removed from oocytes by vortexing in phosphate buffered saline (PBS) with
0.32 mM sodium pyruvate, 5.55 mM glucose, 3 mg/ml BSA, and 10 mg/ml
hyaluronidase. Following vortexing, oocytes were washed in PBS and cultured in CR2
medium atop a layer of cumulus cells.
Donor Cell Culture and SCNT Production
Bovine fibroblasts collected from skin samples were cultured in DMEM F12
(Thermo Scientific HyClone Laboratories). Culture medium was supplemented with
124
15% fetal bovine serum (Thermo Scientific HyClone Laboratories), 100 U/ml penicillin,
and 100 mg/ml streptomycin. Cells were seeded in a T25 culture flask and cultured at
37ºC with 5% CO2. Donor cells were collected prior to SCNT by treatment with 0.25%
trypsin and resuspended in HEPES-buffered synthetic oviduct fluid. Oocytes had
cumulus cells removed following maturation as described above. Oocytes in metaphase
II were chosen for SCNT. Oocytes were enucleated and a single donor cell was
transferred to the perivitelline space. Oocyte-donor cell fusion occurred in mannitol
fusion medium (Wells et al., 1999) with two electric DC pulses of 2.2 kV/cm for 25 µsec.
Following fusion, embryos were cultured in CR2 medium for 1 to 3 hr prior to activation,
which was performed 23 to 25 hr post-maturation. SCNT embryos were activated with 5
min incubation in 5 mM ionomycin, followed by 4 hr incubation in 10 mg/ml
cyclohexamide. SCNT embryos were placed in CR2 medium atop a layer of cumulus
cells.
Fluorescence Immunohistochemistry
IVF and SCNT embryos were collected at the 8-cell and expanded blastocyst
stages. Embryos were collected on three separate IVF/SCNT days. Embryos were
incubated overnight at 4ºC in fixative consisting of PBS with 4% formaldehyde and fixed
overnight. Embryos were then washed at room temperature in PBS with 0.05% Tween
20 (Sigma-Aldrich, St. Louis, MO) (PBST) for 30 min, permeabilized at room
temperature in PBS with 1% Triton X-100 for 30 min, and incubated overnight at 4ºC in
blocking buffer consisting of PBST with 1% BSA. Primary antibodies used were antiCDK2 (ab79379), anti-P53 (ab80645) (Abcam, San Francisco, CA), anti- P21WAF1/Cip1
(LS-C88148), and anti-BAX (LS-C63137) (Lifespan Biosciences, Seattle, WA).
125
Embryos were incubated with the primary antibody at 37ºC for 3 hr, and then washed in
PBST at room temperature for 30 min. Secondary antibodies used were: Alexa Fluor 594
IgG donkey anti-rabbit and Alexa Fluor 488 IgG donkey anti mouse (Life Technologies,
Carlsbad, CA). Embryos were incubated with the appropriate secondary antibody at
37ºC for 2 hr, and then washed in PBST at room temperature for 30 min. During each
immunofluorescence analysis, two stage-appropriate IVF embryos were used as a
negative control and incubated in the absence of the primary antibody, and underwent
secondary antibody incubation. Following primary and secondary antibody incubation,
embryos were incubated with Hoechst 33342 at room temperature for 20 min and then
washed in PBST at room temperature for 10 min. Embryos were mounted onto slides in
PBS with 70% glycerol and covered with a cover slip. Slides were imaged with a Zeiss
Axioobserver equipped with an Axiocam MRm camera
TUNEL Assay
Terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) assay
was carried out using the Click-iT TUNEL Alexa Fluor Imaging Assay (Life
Technologies) according to the manufacturer’s protocol. In brief, IVF and SCNT
embryos were collected from a minimum of three separate IVF and SCNT production
days and collected at the 8-cell and expanded blastocyst stages. Embryos were fixed in
PBS with 4% formaldehyde at room temperature for 30 min and then permeabilized in
PBS with 1% Triton X-100 at room temperature for 30 min. Embryos were then washed
with PBST. Embryos were incubated in the TdT reaction cocktail to incorporate EdUTP
into dsDNA breaks. This was followed by incubation in Click-iT reaction cocktail to
126
induce fluorescence of EdUTP. For each TUNEL assay two stage-appropriate embryos
were used as a positive control by incubation in DNase I prior to performing the TUNEL
assay. Additionally, two stage-appropriate IVF embryos were used as a negative control
by incubation in the TdT reaction cocktail, but were withheld from Click-iT reaction
cocktail. Following cocktail incubation, embryos were incubated with Hoechst 33342
and mounted to slides and imaged as described above. All blastomere nuclei positive for
DNA degradation were counted in each embryo and the total number of apoptotic nuclei
were divided by the total number of imaged embryos at the 8-cell (IVF, N = 23; SCNT, N
= 14) and blastocyst (IVF, N = 48; SCNT, N = 39) stages to determine the average
number of apoptotic blastomeres. A two-way ANOVA with Bonferroni post-hoc test for
multiple comparisons was used to measure significance between stage-matched IVF and
SCNT embryos.
ROS Detection
ROS detection was carried out using the Image-iT LIVE Green Reactive Oxygen
Species Detection Kit (Life Technologies) following the manufacturer’s instructions.
Briefly, embryos were washed in PB1+ (phosphate-buffered saline with Ca2+ and Mg2+
with 5.55 mM glucose, 0.32 mM sodium pyruvate, and 3 mg/ml BSA). Cells were
labeled following manufacturer’s protocol. Embryos were stained with Hoechst and
mounted to slides and imaged as described above.
Quantitative Reverse Transcriptase PCR (qRT-PCR)
127
Individual embryos were suspended in 3 µl H2O, snap frozen in liquid nitrogen,
and frozen at -80ºC until reverse transcription. A total of 12 individual embryos were
analyzed at the 2-cell, 8-cell, morula and blastocyst stages for both IVF and SCNT
embryos. Reverse transcription was carried out using the Cellsdirect One-Step qRT-PCR
Kit (Life Technologies) according to manufacturer’s protocol. Complementary DNA
samples were immediately processed using the Biomark 96.96 Dynamic Array (Fluidigm,
San Francisco, CA) using SYBR Green.
The delta-delta Ct method (Δ-ΔCt) was used for analysis of gene expression. The
average Ct value for the housekeeping genes ACTB and GAPDH was subtracted from the
Ct value of the gene of interest to yield a ΔCt value. Then, the average ΔCt value for
each gene of interest for all embryos at each IVF stage was subtracted from the ΔCt of an
individual embryo for each gene of interest to yield the Δ-ΔCt value. These Δ-ΔCt
values were then used to calculate the expression level for SCNT embryos relative to IVF
embryos at each developmental stage. A number of samples lacked a positive Ct value
for certain genes. In order to calculate fold change any sample lacking a Ct value was
assigned a Ct value of 40, the number of maximum cycles run. These assigned Ct values
were used to generate fold change using the ∆-∆Ct method. Following fold change, all
samples with failed reads were removed for expression data and statistical analysis. Any
embryo that failed for the majority of PCR assays (i.e., no transcript detected for most
genes in the array) or any PCR assay that failed in the majority of embryo samples (i.e.,
no transcript for a PCR assay for most of the embryo samples, IVF or SCNT) was
128
excluded from further analysis.
Statistical Analyses
We employed a two-way ANOVA with Bonferroni post-hoc test for multiple
comparisons (GraphPad Prism v 5.0, La Jolla, CA) to determine whether expression of a
given gene was significantly different in SCNT embryos compared to their stage-matched
IVF counterparts and to determine whether or not apparent differences in relative
expression of a given gene in SCNT and IVF embryos were consistent across all embryo
stages. A probability of P < 0.05 was considered significant.
Results
Gene Expression in Cloned Embryos
Gene transcription levels were measured for several apoptosis-specific genes
using the BioMark Dynamic Array (Fluidigm). This method allowed the examination of
relative gene expression levels for individual embryos. Figure 4-1 summarizes the
expression data for all genes included in the Fluidigm array. An examination of
expression patterns across individual embryos revealed high variability for nearly all
genes assessed for both IVF and SCNT embryos. The gene set for this array included
transcripts involved in early embryo development, pluripotency factors, genome
reprogramming, apoptosis, cell cycle arrest, and several imprinted genes. For the
purposes of this project, the analysis was focused on a subset of transcripts relevant to
apoptosis and cell cycle arrest, as well as several pluripotency factors. Normalized gene
129
expression data and results of statistical analyses for the entire gene set are available in
Figures 4-1 and Table 4-1, respectively.
Several genes relevant to the apoptosis pathway were frequently detected in
SCNT embryos, but not expressed at measurable levels in IVF embryos. BAX, BCL2,
and P21WAF1/Cip1, all of which play a role in regulating apoptosis or cell cycle arrest, are
elevated in SCNT embryos compared to their IVF counterparts, although at specific
developmental stages (Fig. 4-2). For example, BAX was expressed in 42%, 36%, or 58%
of SCNT embryos at the 2-cell, 8-cell, and blastocyst stages, but not in morula SCNT
embryos. Alternatively, fewer IVF embryos were positive for BAX expression at these
developmental stages. Expression of BAX was comparatively higher in SCNT embryos
than IVF at the 8-cell and blastocyst stages, although these differences were not
statistically significant (Fig. 4-3, Table 2). BCL2 is an anti-apoptosis protein that inhibits
BAX, and the BCL2:BAX ratio is an important regulator of apoptosis. The BCL2
transcript was detected in most embryos examined (Fig. 4-2), and levels of BCL2
expression were significantly elevated in cloned blastocysts compared to IVF
counterparts (P < 0.0001 by two-way ANOVA) (Table 4-2 and Fig. 4-3).
The P21WAF1/Cip1 transcript was commonly detected in SCNT embryos at the 2cell, 8-cell and blastocyst developmental stages, whereas this gene was infrequently
detected in IVF embryos (Fig. 4-2). The most marked difference in relative expression
was observed at the 8-cell development stage, where all SCNT embryos expressed
comparatively higher levels of P21WAF1/Cip1 than their IVF counter parts (P < 0.01, twoway ANOVA, Table 4-2).
130
Additional genes involved in apoptosis that are more commonly expressed in
SCNT embryos included IGF2R, CX43, and PEG3 (Fig. 4-4). PEG3 and IGF2R were
expressed at significantly higher levels in SCNT clones at the 2-cell or 8-cell stages
respectively (P <0.01 and < 0.05, respectively, two-way ANOVA, Table 4-2). CX43
transcript was detected in cloned embryos more frequently than IVF embryos at early
developmental stages; an overall significant difference between IVF and cloned embryos
was evident (P < 0.05), although the apparent levels of gene expression were not
statistically different for any particular stage of development (Table 4-2).
Several genes that are involved in P53 activation or somatic cell genome
reprogramming were also examined. Although not detected in any 2-cell or morula stage
embryos, C-MYC was expressed in 73% of SCNT 8-cell embryos compared to only 9%
for their IVF counterparts, and this expression was significantly higher (P <0.05, twoway ANOVA, Fig. 4-5, Table 4-2). Similarly, LIN28 was more frequently expressed in
cloned embryos compared to IVF (82% versus 9% for 8-cell; 50% versus 17% for
blastocyst stage) (Fig. 4-5), expression levels were not significantly different between
SCNT and IVF embryos at any of the developmental stages. SOX2 transcript was
detected in most embryos without any apparent differences between embryos types (Fig.
4-5). However, levels of SOX2 expression were significantly higher in morula stage
clones compared to IVF embryos (P < 0.01, two-way ANOVA, Table 4-2).
Finally, expression of two genes that may function to alleviate apoptotic pressure
in SCNT embryos was examined. ETS2 can block activation of P53, and more SCNT
embryos expressed ETS2 transcripts than did IVF embryos at the 8-cell and blastocyst
131
stages (Fig. 4-4); expression of ETS2 was significantly greater in blastocyst stage clones
compared to IVF embryos (P < 0.05, two-way ANOVA, Table 4-2). Additionally,
although KLF4 was not detected in any IVF embryos, about half of the SCNT embryos at
the 8-cell stage expressed this gene at detectable levels (P <0.0001, two-way ANOVA,
Fig. 4-5, Table 4-2).
Fluorescence Immunohistochemistry
Fluorescence immunohistochemistry was employed to examine protein
expression for active P53, phosphorylated CDK2, P21, and BAX proteins in 8-cell and
blastocyst stage IVF and SCNT embryos. The 8-cell stage was selected based upon the
large number of genes that appeared differentially regulated at this critical developmental
checkpoint; blastocyst stage embryos were assessed as this is the developmental stage at
which embryos are evaluated prior to transfer for subsequent in vivo experiments.
Activated P53 accumulated within SCNT embryos to a greater extent than in IVF
embryos at both the 8-cell and blastocyst stages (Fig. 4-6). Phosphorylated cyclindependent kinase 2 (phospho-CDK2) is a marker of cell cycle arrest. Immunostaining of
phospho-CDK2 revealed higher levels in SCNT embryos at both the 8-cell and blastocyst
stages (Fig. 4-6). Likewise, expression of P21 and BAX proteins were elevated in SCNT
embryos at 8-cell and blastocyst stages (Fig. 4-7).
TUNEL Assay
The TUNEL assay was used to identify apoptotic cells in 8-cell and blastocyst
stage SCNT and IVF embryos. Results of the TUNEL assay indicate that there is no
132
apparent difference in DNA degradation between IVF and SCNT embryos at the 8-cell
stage (Fig. 4-8). However, at the blastocyst stage, SCNT embryos have significantly
more apoptotic nuclei than do IVF embryos (Fig. 4-8). The mean number of apoptotic
nuclei in IVF embryos was 3.9 in IVF compared to 6.8 in SCNT blastocysts. TUNEL
measures DNA degradation in late-stage apoptotic cells, and the elevated incidence of
DNA degradation in SCNT blastocysts confirms our expression and protein analysis of
apoptotic factors.
ROS Detection
Reactive oxidative species (ROS) is an activator of apoptosis and may possibly
induce apoptosis during preimplantation development. In order to determine if ROS was
elevated in SCNT embryos and potentially responsible for activating apoptosis, we
compared levels of ROS in SCNT and IVF embryos. ROS levels were highly variable
among embryos, with no apparent patterns distinctive for SCNT or IVF embryos (data
not shown).
Discussion
To determine whether the P53 pathway is activated in bovine cloned preimplantation embryos, we compared gene and protein expression of key regulators of
apoptosis, cell cycle progression and genome reprogramming in SCNT and IVF embryos.
Herein, we report that a number of genes and proteins involved directly or indirectly in
activation of P53 or subsequent cell cycle and apoptosis pathways are aberrantly
expressed in SCNT embryos. We also report that apoptosis is elevated in blastocyst stage
133
bovine clones compared to their IVF counterparts. To our knowledge, this study is the
first to employ the Fluidigm platform, which incorporates nanofluidics for nano-scale
quantitative RT-PCR, to investigate gene expression in single bovine embryos. Results
of these PCR analyses reveal a high level of inter-embryo variability in gene expression,
an important observation when considering that success of SCNT is dependent on a
single, high quality embryo. Collectively, the observations of this study point to the
involvement of the P53 pathway in mediating apoptosis in SCNT embryos.
Apoptosis plays a key role in preimplantation development and is likely necessary
to ensure the overall quality of blastomeres by eliminating defective cells from the
embryo (Brison and Schultz, 1997; Hardy, 1997). Several studies have shown that
apoptosis in SCNT embryos is elevated compared to embryos generated by IVF
(Fahrudin et al., 2002; Hao et al., 2003; Jang et al., 2004; Gjorret et al., 2005; Im et al.,
2006; Kumar et al., 2007). Additionally, SCNT embryos have a lower cell count
compared to IVF embryos (Koo et al., 2000; Fahrudin et al., 2002; Gjorret et al., 2005;
Terashita et al., 2011), which is likely the result of higher levels apoptosis in clones.
Although apoptosis occurs in IVF embryos, it occurs at minimal levels, whereas
apoptosis in cloned embryos increases throughout pre-implantation development (Hao et
al., 2003).
P53 expression was elevated in SCNT embryos compared to their IVF
counterparts, and this elevated expression may account for the lower blastomere count in
clones observed in other studies (Devreker and Hardy, 1997; Byrne et al., 1999; Hardy et
al., 2001) as well as the increased DNA degradation we observed in SCNT blastocysts in
134
this study. Although embryos with reduced cell counts are developmentally competent
(Machaty et al., 1998), embryos with greater cell numbers are better prepared for
implantation and developmental success (van Soom et al., 1997). Loss of SCNT embryos
after transfer to surrogates and low pregnancy rates may be a consequence of high rates
of apoptosis and reduced blastomere number during the preimplantation stage (Fahrudin
et al., 2002).
P21WAF1/Cip1 is transcriptionally regulated by P53 and is responsible for cell cycle
arrest (Brugarolas et al., 1999). Upon transcriptional upregulation, P21 acts as an
inhibitor of cyclin/CDK2 to block the progression of the cell cycle (Niculescu et al.,
1998). Results of this study establish that P21WAF1/Cip1 is expressed in a majority of
SCNT embryos at the 2-cell, 8-cell, and blastocyst stage as measured by gene expression
and protein immunofluorescence. As an additional measure of cell cycle arrest, we
performed immunostaining of the phosphorylated form of CDK2, which is a regulator of
cell cycle progression and is inhibited by phosphorylation (Morgan, 1995). CDKs are a
target of inhibition by P21 (Poon et al., 1996), and phospho-CDK2 is higher in apoptotic
cells. Our results indicate that phospho-CDK2 is prevalent in SCNT embryos, yet
present at low levels in embryos produced via IVF. These results further support the role
of cell cycle arrest as a factor that likely influences developmental competence of SCNTderived embryos. While several research groups have examined the apoptotic status of
SCNT embryos using the TUNEL assay and blastomere morphological changes (Hao et
al., 2003; Jang et al., 2004; Gjorret et al., 2005; Im et al., 2006), this study was the first to
investigate P53 activation as the trigger for apoptosis in SCNT embryos. Additionally,
135
cell cycle arrest cannot be detected using assays such as TUNEL. Our findings suggest
that cell cycle arrest should be included as a standard in assessing SCNT embryo
viability, as cessation of cell replication likely has grave consequences for development
of SCNT preimplantation embryos. The arrested development of a blastomere at the 8cell stage may reduce the total number of cells capable of contributing to embryo
development. Indeed, the prior observations that SCNT embryo cell counts are
comparatively low (Koo et al., 2000; Fahrudin et al., 2002; Gjorret et al., 2005; Terashita
et al., 2011) may be the result of blastomere arrest at earlier embryo stages.
In addition to analyzing genes related to apoptosis, we also investigated
reprogramming factors that may activate P53-mediated apoptosis. Overexpression of CMYC has been shown to lead to the activation of apoptosis (Eischen et al., 1999;
Pelengaris et al., 2002; Zhao et al., 2008; Kawamura et al., 2009). In this study, C-MYC
was detected in a majority of 8-cell SCNT embryos, but in only a single IVF embryo. CMYC may act as a trigger in activating apoptosis following SCNT. LIN28 was originally
characterized as a heterchronic gene in C. elegans that acts as a switch that controls the
timing and developmental fate of cells (Moss et al., 1997). LIN28 is a translational
enhancer of C-MYC, and elevated levels of this gene, as observed in 8-cell clones, could
contribute to the activation of P53-mediated apoptosis (West et al., 2009).
Expression of the pluripotency factor SOX2 was elevated in 2-cell and morula
stage SCNT embryos compared to their IVF counterparts. This gene is critical during
early embryo development and in somatic cell genome reprogramming (Takahashi et al.,
2007; Yu et al., 2007; Silva et al., 2009). When expressed ectopically, SOX2 can induce
136
expression of p21WAF1/Cip1 via the P53 pathway (Kawamura et al., 2009). Interestingly, in
this study, the elevated expression of SOX2 appears to precede P21WAF1/Cip1 expression in
latter developmental stages. Although largely absent in most embryos examined, KLF4
expression was detected in more than half of the 8-cell cloned embryos examined.
Because this protein has been shown to induce P21WAF1/Cip1 in other studies (Zhang et al.,
2000; Chen et al., 2003; Rowland et al., 2005), it stands to reason that KLF4 may have
contributed to the higher levels of p21WAF1/Cip1 expression also detected in 8-cell SCNT
embryos in this study. Interestingly, while KLF4 can upregulate P21WAF1/Cip1, it also
functions as a repressor of P53 (Li et al., 2001; Rowland et al., 2005). Expression of CMYC and KLF4 does not appear to be required for preimplantation development, as
neither gene was detected with regular frequency in embryos produced via IVF. Both
genes are expressed primarily at the 8-cell stage, and expression varied widely among
SCNT embryos. It is possible that the activation of genes such as C-MYC, SOX2, KLF4,
and LIN28 is a necessary step in donor cell reprogramming. However, the upregulation
of these genes in SCNT embryos may contribute to activation of P53-mediated apoptosis
and cell cycle arrest.
Our study also demonstrates that BAX protein expression is elevated in SCNT
embryos at the 8-cell and blastocyst stages. Melka et al. (2010) observed higher
expression of BAX in low quality embryos and suggested that BAX could serve as a
marker for embryo quality in bovine embryos. BAX can be transcriptionally upregulated
by P53 (Oltvai et al., 1993; Miyashita and Reed, 1995), or P53 can directly activate BAX
in a transcription-independent manner (Schuler et al., 2000; Chipuk et al., 2005).
Whichever the mechanism, increased BAX expression is important for understanding
137
SCNT embryo failure. It is however, essential to interpret BAX expression in context of
BCL2. In bovine IVF development, embryos with high cytoplasmic fragmentation had
elevated levels of BAX, whereas embryos with limited cytoplasmic fragmentation had
low BAX and elevated BLC2 (McCurrach et al., 1997; Yang and Rajamahendran, 2002).
BCL2 confers an anti-apoptotic response (Levine, 1997; Yang et al., 1997), but activated
P53 can directly inhibit BCL2 to facilitate apoptosis (Wang et al., 1993; Chiou et al.,
1994; Marin et al., 1994; Froesch et al., 1999). Analysis of BCL2 transcription indicates
that the gene is expressed within all embryos at all stages, but a statistically significant
upregulation occurs within SCNT embryos at the blastocyst stage. BAX expression was
observed in a subset of embryos and was more commonly expressed in SCNT embryos
that IVF. Aberrant expression of genes involved in apoptosis was predominant in 8-cell
SCNT embryos, which could induce a compensatory response to this apoptotic pressure
leading to elevated expression of BCL2 in blastocyst embryos as a mechanism to
conserve blastomeres. This response may account for the development of SCNT
embryos to the blastocyst stage despite the presence of apoptosis inducing factors such as
BAX. Additionally, the ETS2 gene is capable of suppressing P53. ETS2 acts as a
proliferation factor that blocks apoptosis (Venanzoni et al., 1996; Sevilla et al., 1999).
Indeed, the expression of ETS2, KLF4, and BCL2 in clones may explain why some SCNT
embryos are able to proceed through development and successfully develop despite
activation of the P53 pathway.
138
In this study, we also detected elevated expression of IGF2R, CX43, and PEG3 in
a subset of SCNT embryos. IGF2R knockdown mice are typified by overgrowth during
fetal development and perinatal death (Lau et al., 1994; Wang et al., 1994). Additionally,
IGF2R has functions as a tumor suppressor, as loss of function mutations to this gene
lead to tumor growth (De Souza et al., 1995; Hankins et al., 1996; Chappell et al., 1997;
Ouyang et al., 1997; Piao et al., 1997; Wang et al., 1997; Xu et al., 1997; Yamada et al.,
1997). During apoptosis, IGF2R acts as a death receptor through activation of
mitochondrial-dependent apoptosis (Motyka et al., 2000). IGF2R overexpression
increases cell susceptibility to apoptosis (Kang et al., 1999). Likewise CX43 contributes
to apoptosis and is elevated in SCNT embryos. CX43 is a component of gap junctions
and has been implicated in apoptosis (Huang et al., 2001; Kameritsch et al., 2013) and
cell cycle arrest (Zhang et al., 2003). PEG3 facilitates the translocation of BAX to the
mitochondria and participates in P53-mediated apoptosis (Deng and Wu, 2000). PEG3 is
expressed in all stages of embryo development, but is elevated in SCNT embryos at the
2-cell and morula stages. The elevated expression of IGF2R, CX43, and PEG3 further
highlights apoptotic environment under which SCNT embryos develop. Collectively, our
observations point to a scenario in which SCNT embryos experience apoptotic pressure
that may contribute to either activation of the programmed cell death pathway or cell
cycle arrest. A consequence of either cell response would be the under-representation of
the embryo cell population, placing that embryo at a selective disadvantage for robust
placentation and fetal maintenance.
139
The TUNEL assay detects 3’-OH DNA strand breaks that occur during late-stage
apoptosis (Gavrieli et al., 1992). TUNEL assay has been previously used to assess
embryo viability during preimplantation development (Brison and Schultz, 1997; Byrne
et al., 1999; Matwee et al., 2000; Gjorret et al., 2003; Gjorret et al., 2005). Earlier reports
have shown that DNA fragmentation as measured by TUNEL is higher in SCNT embryos
as compared to IVF (Fahrudin et al., 2002; Hao et al., 2003; Jang et al., 2004; Gjorret et
al., 2005; Im et al., 2006), and our data is consistent with these findings. We observed no
difference between IVF and SCNT at the 8-cell stage, which is expected as apoptosis
related genes are minimally present at the 2-cell stage and onset of activation of the P53
pathway appears to occur at the 8-cell stage. Because TUNEL measures late-stage
apoptosis, onset of apoptosis would not be expected until later in development, as we
observed in blastocyst SCNT embryos.
The potential benefits of SCNT are currently hindered by the low success rate of
the technique. A clearer understanding of the pathways implicated in the poor
development of SCNT embryos may yield important knowledge that can be leveraged to
develop interventions to overcome embryonic failure and improve developmental rates of
cloned animals. Future studies should examine other downstream genes in the P53
pathway in order to gain additional insight regarding the mechanisms responsible for
triggering P53-mediated apoptosis. We did not explore chromosomal aberrations within
cloned embryos, which have been associated with reprogramming (Marion et al., 2009),
and cannot rule out that increase apoptosis is the result of chromosomal anomalies.
Inhibition of P53 during preimplantation development may give a developmental window
140
of opportunity in which blastomeres can proliferate and development of the embryo can
occur prior to implantation.
References
Basu A, Haldar S. 1998. The relationship between BcI2, Bax and p53: consequences for
cell cycle progression and cell death. Mol Hum Reprod 4(12):1099-1109.
Brison DR, Schultz RM. 1997. Apoptosis during mouse blastocyst formation: evidence
for a role for survival factors including transforming growth factor alpha. Biol
Reprod 56(5):1088-1096.
Brison DR, Schultz RM. 1998. Increased incidence of apoptosis in transforming growth
factor alpha-deficient mouse blastocysts. Biol Reprod 59(1):136-144.
Brugarolas J, Moberg K, Boyd SD, Taya Y, Jacks T, Lees JA. 1999. Inhibition of cyclindependent kinase 2 by p21 is necessary for retinoblastoma protein-mediated G1
arrest after gamma-irradiation. Proc Natl Acad Sci USA 96(3):1002-1007.
Byrne AT, Southgate J, Brison DR, Leese HJ. 1999. Analysis of apoptosis in the
preimplantation bovine embryo using TUNEL. J Reprod Fertil 117(1):97-105.
Chappell SA, Walsh T, Walker RA, Shaw JA. 1997. Loss of heterozygosity at the
mannose 6-phosphate insulin-like growth factor 2 receptor gene correlates with
poor differentiation in early breast carcinomas. Br J Cancer 76(12):1558-1561.
Chen X, Whitney EM, Gao SY, Yang VW. 2003. Transcriptional profiling of Kruppellike factor 4 reveals a function in cell cycle regulation and epithelial
differentiation. J Mol Biol 326(3):665-677.
Chiou SK, Rao L, White E. 1994. Bcl-2 blocks p53-dependent apoptosis. Mol Cell Biol
14(4):2556-2563.
Chipuk JE, Bouchier-Hayes L, Kuwana T, Newmeyer DD, Green DR. 2005. PUMA
couples the nuclear and cytoplasmic proapoptotic function of p53. Science
309(5741):1732-1735.
Colman A. 1999. Somatic cell nuclear transfer in mammals: progress and applications.
Cloning 1(4):185-200.
De Souza AT, Hankins GR, Washington MK, Orton TC, Jirtle RL. 1995. M6P/IGF2R
gene is mutated in human hepatocellular carcinomas with loss of heterozygosity.
Nat Genet 11(4):447-449.
Dean W, Santos F, Stojkovic M, Zakhartchenko V, Walter J, Wolf E, Reik W. 2001.
Conservation of methylation reprogramming in mammalian development:
aberrant reprogramming in cloned embryos. Proc Natl Acad Sci USA
98(24):13734-13738.
141
Deng Y, Wu X. 2000. Peg3/Pw1 promotes p53-mediated apoptosis by inducing Bax
translocation from cytosol to mitochondria. Proc Natl Acad Sci USA
97(22):12050-12055.
Devreker F, Hardy K. 1997. Effects of glutamine and taurine on preimplantation
development and cleavage of mouse embryos in vitro. Biol Reprod 57(4):921928.
Eischen CM, Weber JD, Roussel MF, Sherr CJ, Cleveland JL. 1999. Disruption of the
ARF-Mdm2-p53 tumor suppressor pathway in Myc-induced lymphomagenesis.
Genes Dev 13(20):2658-2669.
Fabian D, Koppel J, Maddox-Hyttel P. 2005. Apoptotic processes during mammalian
preimplantation development. Theriogenology 64(2):221-231.
Fahrudin M, Otoi T, Karja NW, Mori M, Murakami M, Suzuki T. 2002. Analysis of
DNA fragmentation in bovine somatic nuclear transfer embryos using TUNEL.
Reproduction 124(6):813-819.
Froesch BA, Aime-Sempe C, Leber B, Andrews D, Reed JC. 1999. Inhibition of p53
transcriptional activity by Bcl-2 requires its membrane-anchoring domain. J Biol
Chem 274(10):6469-6475.
Ganeshan L, Li A, O'Neill C. 2010. Transformation-related protein 53 expression in the
early mouse embryo compromises preimplantation embryonic development by
preventing the formation of a proliferating inner cell mass. Biol Reprod
83(6):958-964.
Gavrieli Y, Sherman Y, Ben-Sasson SA. 1992. Identification of programmed cell death in
situ via specific labeling of nuclear DNA fragmentation. J Cell Biol 119(3):493501.
Gjorret JO, Knijn HM, Dieleman SJ, Avery B, Larsson LI, Maddox-Hyttel P. 2003.
Chronology of apoptosis in bovine embryos produced in vivo and in vitro. Biol
Reprod 69(4):1193-1200.
Gjorret JO, Wengle J, Maddox-Hyttel P, King WA. 2005. Chronological appearance of
apoptosis in bovine embryos reconstructed by somatic cell nuclear transfer from
quiescent granulosa cells. Reprod Domest Anim 40(3):210-216.
142
Hankins GR, De Souza AT, Bentley RC, Patel MR, Marks JR, Iglehart JD, Jirtle RL.
1996. M6P/IGF2 receptor: a candidate breast tumor suppressor gene. Oncogene
12(9):2003-2009.
Hao Y, Lai L, Mao J, Im GS, Bonk A, Prather RS. 2003. Apoptosis and in vitro
development of preimplantation porcine embryos derived in vitro or by nuclear
transfer. Biol Reprod 69(2):501-507.
Hao Y, Lai L, Mao J, Im GS, Bonk A, Prather RS. 2004. Apoptosis in parthenogenetic
preimplantation porcine embryos. Biol Reprod 70(6):1644-1649.
Hardy K. 1997. Cell death in the mammalian blastocyst. Mol Hum Reprod 3(10):919925.
Hardy K, Spanos S, Becker D, Iannelli P, Winston RM, Stark J. 2001. From cell death to
embryo arrest: mathematical models of human preimplantation embryo
development. Proc Natl Acad Sci USA 98(4):1655-1660.
Hong H, Takahashi K, Ichisaka T, Aoi T, Kanagawa O, Nakagawa M, Okita K,
Yamanaka S. 2009. Suppression of induced pluripotent stem cell generation by
the p53-p21 pathway. Nature 460(7259):1132-1135.
Huang R, Liu YG, Lin Y, Fan Y, Boynton A, Yang D, Huang RP. 2001. Enhanced
apoptosis under low serum conditions in human glioblastoma cells by connexin
43 (Cx43). Mol Carcinog 32(3):128-138.
Humpherys D, Eggan K, Akutsu H, Hochedlinger K, Rideout WM, 3rd, Biniszkiewicz D,
Yanagimachi R, Jaenisch R. 2001. Epigenetic instability in ES cells and cloned
mice. Science 293(5527):95-97.
Im GS, Seo JS, Hwang IS, Kim DH, Kim SW, Yang BC, Yang BS, Lai L, Prather RS.
2006. Development and apoptosis of pre-implantation porcine nuclear transfer
embryos activated with different combination of chemicals. Mol Reprod Dev
73(9):1094-1101.
Jang G, Park ES, Cho JK, Bhuiyan MM, Lee BC, Kang SK, Hwang WS. 2004.
Preimplantational embryo development and incidence of blastomere apoptosis in
bovine somatic cell nuclear transfer embryos reconstructed with long-term
cultured donor cells. Theriogenology 62(3-4):512-521.
Jurisicova A, Antenos M, Varmuza S, Tilly JL, Casper RF. 2003. Expression of
apoptosis-related genes during human preimplantation embryo development:
potential roles for the Harakiri gene product and Caspase-3 in blastomere
fragmentation. Mol Hum Reprod 9(3):133-141.
143
Jurisicova A, Latham KE, Casper RF, Varmuza SL. 1998. Expression and regulation of
genes associated with cell death during murine preimplantation embryo
development. Mol Reprod Dev 51(3):243-253.
Kameritsch P, Khandoga N, Pohl U, Pogoda K. 2013. Gap junctional communication
promotes apoptosis in a connexin-type-dependent manner. Cell Death Dis 4:e584.
Kang JX, Bell J, Beard RL, Chandraratna RA. 1999. Mannose 6-phosphate/insulin-like
growth factor II receptor mediates the growth-inhibitory effects of retinoids. Cell
Growth Differ 10(8):591-600.
Kawamura T, Suzuki J, Wang YV, Menendez S, Morera LB, Raya A, Wahl GM, Izpisua
Belmonte JC. 2009. Linking the p53 tumour suppressor pathway to somatic cell
reprogramming. Nature 460(7259):1140-1144.
Kerr JF, Wyllie AH, Currie AR. 1972. Apoptosis: a basic biological phenomenon with
wide-ranging implications in tissue kinetics. Br J Cancer 26(4):239-257.
Kikyo N, Wolffe AP. 2000. Reprogramming nuclei: insights from cloning, nuclear
transfer and heterokaryons. J Cell Sci 113 ( Pt 1):11-20.
Knijn HM, Gjorret JO, Vos PL, Hendriksen PJ, van der Weijden BC, Maddox-Hyttel P,
Dieleman SJ. 2003. Consequences of in vivo development and subsequent culture
on apoptosis, cell number, and blastocyst formation in bovine embryos. Biol
Reprod 69(4):1371-1378.
Koo DB, Kang YK, Choi YH, Park JS, Han SK, Park IY, Kim SU, Lee KK, Son DS,
Chang WK, Han YM. 2000. In vitro development of reconstructed porcine
oocytes after somatic cell nuclear transfer. Biol Reprod 63(4):986-992.
Kumar BM, Jin HF, Kim JG, Ock SA, Hong Y, Balasubramanian S, Choe SY, Rho GJ.
2007. Differential gene expression patterns in porcine nuclear transfer embryos
reconstructed with fetal fibroblasts and mesenchymal stem cells. Dev Dyn
236(2):435-446.
Lau MM, Stewart CE, Liu Z, Bhatt H, Rotwein P, Stewart CL. 1994. Loss of the
imprinted IGF2/cation-independent mannose 6-phosphate receptor results in fetal
overgrowth and perinatal lethality. Genes Dev 8(24):2953-2963.
Levine AJ. 1997. p53, the cellular gatekeeper for growth and division. Cell 88(3):323331.
Li A, Chandrakanthan V, Chami O, O'Neill C. 2007. Culture of zygotes increases TRP53
[corrected] expression in B6 mouse embryos, which reduces embryo viability.
Biol Reprod 76(3):362-367.
Li C, Ahlborn TE, Tokita K, Boxer LM, Noda A, Liu J. 2001. The critical role of the
PE21 element in oncostatin M-mediated transcriptional repression of the p53
tumor suppressor gene in breast cancer cells. Oncogene 20(57):8193-8202.
144
Liu L, Trimarchi JR, Keefe DL. 2002. Haploidy but not parthenogenetic activation leads
to increased incidence of apoptosis in mouse embryos. Biol Reprod 66(1):204210.
Machaty Z, Day BN, Prather RS. 1998. Development of early porcine embryos in vitro
and in vivo. Biol Reprod 59(2):451-455.
Mali P, Ye Z, Hommond HH, Yu X, Lin J, Chen G, Zou J, Cheng L. 2008. Improved
efficiency and pace of generating induced pluripotent stem cells from human
adult and fetal fibroblasts. Stem Cells 26(8):1998-2005.
Marin MC, Hsu B, Meyn RE, Donehower LA, el-Naggar AK, McDonnell TJ. 1994.
Evidence that p53 and bcl-2 are regulators of a common cell death pathway
important for in vivo lymphomagenesis. Oncogene 9(11):3107-3112.
Marion RM, Strati K, Li H, Murga M, Blanco R, Ortega S, Fernandez-Capetillo O,
Serrano M, Blasco MA. 2009. A p53-mediated DNA damage response limits
reprogramming to ensure iPS cell genomic integrity. Nature 460(7259):11491153.
Matwee C, Betts DH, King WA. 2000. Apoptosis in the early bovine embryo. Zygote
8(1):57-68.
McCurrach ME, Connor TM, Knudson CM, Korsmeyer SJ, Lowe SW. 1997. baxdeficiency promotes drug resistance and oncogenic transformation by attenuating
p53-dependent apoptosis. Proc Natl Acad Sci USA 94(6):2345-2349.
Melka MG, Rings F, Holker M, Tholen E, Havlicek V, Besenfelder U, Schellander K,
Tesfaye D. 2010. Expression of apoptosis regulatory genes and incidence of
apoptosis in different morphological quality groups of in vitro-produced bovine
pre-implantation embryos. Reprod Domest Anim 45(5):915-921.
Miyashita T, Reed JC. 1995. Tumor suppressor p53 is a direct transcriptional activator of
the human bax gene. Cell 80(2):293-299.
Morgan DO. 1995. Principles of CDK regulation. Nature 374(6518):131-134.
Moss EG, Lee RC, Ambros V. 1997. The cold shock domain protein LIN-28 controls
developmental timing in C. elegans and is regulated by the lin-4 RNA. Cell
88(5):637-646.
145
Motyka B, Korbutt G, Pinkoski MJ, Heibein JA, Caputo A, Hobman M, Barry M,
Shostak I, Sawchuk T, Holmes CF, Gauldie J, Bleackley RC. 2000. Mannose 6phosphate/insulin-like growth factor II receptor is a death receptor for granzyme
B during cytotoxic T cell-induced apoptosis. Cell 103(3):491-500.
Neuber E, Luetjens CM, Chan AW, Schatten GP. 2002. Analysis of DNA fragmentation
of in vitro cultured bovine blastocysts using TUNEL. Theriogenology 57(9):21932202.
Niculescu AB, 3rd, Chen X, Smeets M, Hengst L, Prives C, Reed SI. 1998. Effects of
p21(Cip1/Waf1) at both the G1/S and the G2/M cell cycle transitions: pRb is a
critical determinant in blocking DNA replication and in preventing
endoreduplication. Mol Cell Biol 18(1):629-643.
Niemann H, Tian XC, King WA, Lee RS. 2008. Epigenetic reprogramming in embryonic
and foetal development upon somatic cell nuclear transfer cloning. Reproduction
135(2):151-163.
Oltvai ZN, Milliman CL, Korsmeyer SJ. 1993. Bcl-2 heterodimerizes in vivo with a
conserved homolog, Bax, that accelerates programmed cell death. Cell 74(4):609619.
Ouyang H, Shiwaku HO, Hagiwara H, Miura K, Abe T, Kato Y, Ohtani H, Shiiba K,
Souza RF, Meltzer SJ, Horii A. 1997. The insulin-like growth factor II receptor
gene is mutated in genetically unstable cancers of the endometrium, stomach, and
colorectum. Cancer Res 57(10):1851-1854.
Pelengaris S, Khan M, Evan G. 2002. c-MYC: more than just a matter of life and death.
Nat Rev Cancer 2(10):764-776.
Piao Z, Choi Y, Park C, Lee WJ, Park JH, Kim H. 1997. Deletion of the M6P/IGF2r gene
in primary hepatocellular carcinoma. Cancer Lett 120(1):39-43.
Poon RY, Jiang W, Toyoshima H, Hunter T. 1996. Cyclin-dependent kinases are
inactivated by a combination of p21 and Thr-14/Tyr-15 phosphorylation after
UV-induced DNA damage. J Biol Chem 271(22):13283-13291.
Qin H, Yu T, Qing T, Liu Y, Zhao Y, Cai J, Li J, Song Z, Qu X, Zhou P, Wu J, Ding M,
Deng H. 2007. Regulation of apoptosis and differentiation by p53 in human
embryonic stem cells. J Biol Chem 282(8):5842-5852.
Reik W, Dean W, Walter J. 2001. Epigenetic reprogramming in mammalian
development. Science 293(5532):1089-1093.
146
Rowland BD, Bernards R, Peeper DS. 2005. The KLF4 tumour suppressor is a
transcriptional repressor of p53 that acts as a context-dependent oncogene. Nat
Cell Biol 7(11):1074-1082.
Schuler M, Bossy-Wetzel E, Goldstein JC, Fitzgerald P, Green DR. 2000. p53 induces
apoptosis by caspase activation through mitochondrial cytochrome c release. J
Biol Chem 275(10):7337-7342.
Sevilla L, Aperlo C, Dulic V, Chambard JC, Boutonnet C, Pasquier O, Pognonec P,
Boulukos KE. 1999. The Ets2 transcription factor inhibits apoptosis induced by
colony-stimulating factor 1 deprivation of macrophages through a Bcl-xLdependent mechanism. Mol Cell Biol 19(4):2624-2634.
Silva J, Nichols J, Theunissen TW, Guo G, van Oosten AL, Barrandon O, Wray J,
Yamanaka S, Chambers I, Smith A. 2009. Nanog is the gateway to the pluripotent
ground state. Cell 138(4):722-737.
Solter D. 2000. Mammalian cloning: advances and limitations. Nat Rev Genet 1(3):199207.
Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, Yamanaka S.
2007. Induction of pluripotent stem cells from adult human fibroblasts by defined
factors. Cell 131(5):861-872.
Terashita Y, Sugimura S, Kudo Y, Amano R, Hiradate Y, Sato E. 2011. Improving the
quality of miniature pig somatic cell nuclear transfer blastocysts: aggregation of
SCNT embryos at the four-cell stage. Reprod Domest Anim 46(2):189-196.
Thibault C. 2003. Recent data on the development of cloned embryos derived from
reconstructed eggs with adult cells. Reprod Nutr Dev 43(4):303-324.
Utikal J, Polo JM, Stadtfeld M, Maherali N, Kulalert W, Walsh RM, Khalil A, Rheinwald
JG, Hochedlinger K. 2009. Immortalization eliminates a roadblock during cellular
reprogramming into iPS cells. Nature 460(7259):1145-1148.
van Soom A, Ysebaert MT, de Kruif A. 1997. Relationship between timing of
development, morula morphology, and cell allocation to inner cell mass and
trophectoderm in in vitro-produced bovine embryos. Mol Reprod Dev 47(1):4756.
Venanzoni MC, Robinson LR, Hodge DR, Kola I, Seth A. 1996. ETS1 and ETS2 in p53
regulation: spatial separation of ETS binding sites (EBS) modulate protein: DNA
interaction. Oncogene 12(6):1199-1204.
Vousden KH, Lu X. 2002. Live or let die: the cell's response to p53. Nat Rev Cancer
2(8):594-604.
147
Wang S, Souza RF, Kong D, Yin J, Smolinski KN, Zou TT, Frank T, Young J, Flanders
KC, Sugimura H, Abraham JM, Meltzer SJ. 1997. Deficient transforming growth
factor-beta1 activation and excessive insulin-like growth factor II (IGFII)
expression in IGFII receptor-mutant tumors. Cancer Res 57(13):2543-2546.
Wang Y, Szekely L, Okan I, Klein G, Wiman KG. 1993. Wild-type p53-triggered
apoptosis is inhibited by bcl-2 in a v-myc-induced T-cell lymphoma line.
Oncogene 8(12):3427-3431.
Wang ZQ, Fung MR, Barlow DP, Wagner EF. 1994. Regulation of embryonic growth
and lysosomal targeting by the imprinted Igf2/Mpr gene. Nature 372(6505):464467.
Wei MC, Zong WX, Cheng EH, Lindsten T, Panoutsakopoulou V, Ross AJ, Roth KA,
MacGregor GR, Thompson CB, Korsmeyer SJ. 2001. Proapoptotic BAX and
BAK: a requisite gateway to mitochondrial dysfunction and death. Science
292(5517):727-730.
Wells DN, Misica PM, Tervit HR. 1999. Production of cloned calves following nuclear
transfer with cultured adult mural granulosa cells. Biol Reprod 60(4):996-1005.
West JA, Viswanathan SR, Yabuuchi A, Cunniff K, Takeuchi A, Park IH, Sero JE, Zhu
H, Perez-Atayde A, Frazier AL, Surani MA, Daley GQ. 2009. A role for Lin28 in
primordial germ-cell development and germ-cell malignancy. Nature
460(7257):909-913.
Wilmut I, Young L, Campbell KH. 1998. Embryonic and somatic cell cloning. Reprod
Fertil Dev 10(7-8):639-643.
Wuu YD, Pampfer S, Becquet P, Vanderheyden I, Lee KH, De Hertogh R. 1999. Tumor
necrosis factor alpha decreases the viability of mouse blastocysts in vitro and in
vivo. Biol Reprod 60(2):479-483.
Xu YQ, Grundy P, Polychronakos C. 1997. Aberrant imprinting of the insulin-like
growth factor II receptor gene in Wilms' tumor. Oncogene 14(9):1041-1046.
Yamada T, De Souza AT, Finkelstein S, Jirtle RL. 1997. Loss of the gene encoding
mannose 6-phosphate/insulin-like growth factor II receptor is an early event in
liver carcinogenesis. Proc Natl Acad Sci USA 94(19):10351-10355.
Yang J, Liu X, Bhalla K, Kim CN, Ibrado AM, Cai J, Peng TI, Jones DP, Wang X. 1997.
Prevention of apoptosis by Bcl-2: release of cytochrome c from mitochondria
blocked. Science 275(5303):1129-1132.
148
Yang MY, Rajamahendran R. 2002. Expression of Bcl-2 and Bax proteins in relation to
quality of bovine oocytes and embryos produced in vitro. Anim Reprod Sci 70(34):159-169.
Yin XM, Oltvai ZN, Korsmeyer SJ. 1994. BH1 and BH2 domains of Bcl-2 are required
for inhibition of apoptosis and heterodimerization with Bax. Nature
369(6478):321-323.
Yu J, Vodyanik MA, Smuga-Otto K, Antosiewicz-Bourget J, Frane JL, Tian S, Nie J,
Jonsdottir GA, Ruotti V, Stewart R, Slukvin, II, Thomson JA. 2007. Induced
pluripotent stem cell lines derived from human somatic cells. Science
318(5858):1917-1920.
Zhang W, Geiman DE, Shields JM, Dang DT, Mahatan CS, Kaestner KH, Biggs JR,
Kraft AS, Yang VW. 2000. The gut-enriched Kruppel-like factor (Kruppel-like
factor 4) mediates the transactivating effect of p53 on the p21WAF1/Cip1
promoter. J Biol Chem 275(24):18391-18398.
Zhang YW, Nakayama K, Morita I. 2003. A novel route for connexin 43 to inhibit cell
proliferation: negative regulation of S-phase kinase-associated protein (Skp 2).
Cancer Res 63(7):1623-1630.
Zhao Y, Yin X, Qin H, Zhu F, Liu H, Yang W, Zhang Q, Xiang C, Hou P, Song Z, Liu
Y, Yong J, Zhang P, Cai J, Liu M, Li H, Li Y, Qu X, Cui K, Zhang W, Xiang T,
Wu Y, Liu C, Yu C, Yuan K, Lou J, Ding M, Deng H. 2008. Two supporting
factors greatly improve the efficiency of human iPSC generation. Cell Stem Cell
3(5):475-479.
149
Table 4-1. Primers Used for qRT-PCR
Gene
Forward Primer
Reverse Primer
ACTB
BAX
BCL2
CDX2
C-MYC
CX43
DNMT3A
ELF5
EOMES
ESRRB
ETS2
G9A
GAPDH
GATA6
GNAS
HDAC1
HDAC2
IGF2R
JARID2
KAT2B
KDM3A
KDM5A
KLF4
LIN28
MASH2
MDM2
P21WAF1/Cip1
PEG3
RASGRF1
SOX2
TEAD4
UBE3A
ZIM2
TTCCGCTGCCCTGAGGCTCT
ACCAAGAAGCTGAGCGAGTGTC
TTCGCCGAGATGTCCAGTCAGC
ACTACGGCGGATACCATGTGGC
TAGTGGAAATACGGGCTTGC
AACCGCTCCTCTCTCGCCCA
CTGAGAAGAAAGCCAAGGTGATCGC
ACAGGAGGAGTTCATGGAGGCG
CTCAAATTCCACCGCCACCAAAC
TGCTCAACGCCATCCCCAAG
TCCCTGCTTGCGGTTTTCCC
CAGCAGCAGGGCCTGAACCC
CACCTTCTGCCGATGCCCCC
ACCGTGGATGGCCTTGACTGAC
TGCTGCTGGGTGCTGGAGAATC
AAATTCTTACGCTCCATCCGCCC
AGGGATATTGGTGCAGGAAAAGGC
GCCAAGTCACAGACCCCGCC
TCCGTGGTCAGAAGAAAGGGTGG
TTGAACAGCAGAACGGAGGGAGC
GCAAGGACAAGGAGCAGCAGAAAG
CTGAATGAGCTTGAGGCGATGACC
GGCCGCCTGCTCACGACTTT
AAACGCAGATCAAAGGGAGA
CCTCGTCGTCCCCTGGTCGT
GATGAAGATGATGAGGTGTATC
GCCACAGGTGCCATGTCTG
CGCCAAAGTCAGGGAGAG
TTCCTGAGCATCGACTTCCT
GTGAACCAGCGCATGGACAG
CCGCAGCCTTCCACAGTAAAATG
ACTGAGGGCTGTGGAAATGAGGC
AGCCAGAGATCCACACCAAG
TGTGGACAGCGAGGCCAGGA
CGAAGGAAGTCCAATGTCCAGCC
TTGACGCTCTCCACACACATGACC
TTCCGCATCCACTCGCACAG
TTCTTCCAGATATCCTCGCTG
CTGGCTCTGCTGGAAGGCCG
ATGACCCAACGAGTGCCTTCCG
GTCAACCCGTTCCAAAATTCCG
TCATCTGGGTGTTGTTGTTGTTTGC
GGTGGAGGCATGGCATACAGTTTG
CCCTGTCATGCCAAACCTCTGC
CCCACTCCTCCAGGGACCCG
CACGTTGGGAGTGGGGACGC
CGAAGGGATGCGAGGCGTAG
ACGCACCCCTTCATCCTCCC
AATATCCCGTAGGTCCCCAGTTCC
GGGAATCTCACAATCAAGGGCAAC
GTCGGGGAGCCCGTTGTGTG
CATTGTTGGTGGCTGCTTTGACG
CCTACCAGCCACATGAGCACCTTC
GTGCAGGTTGAAGATGGTGGTGTC
TGGCGAAGTTTGGATGTCAGTGC
GTGGCCGTCCTTTTCCGGGG
TGCAGAAACGCAGATCAAAG
CTGGCTCGGTGGTGGCTTGG
CCCAACATCTGTTGCAGTGA
TCTCGGTGACAAAGTCGAAG
CTTAACTGCCAGGACACC
AGCTTGTTGTTCCCACT
GTGCCCTGCTGAGAATAGGAC
ATGTCCACGGCTTCGAGGTAGG
TCTTGCTCCTTGCCCGTCCTTC
CGGAAGCAGCACTTCCTATG
150
Table 4-2. Statistical Analysis of Gene Expression.
Two-way ANOVA
Gene
Embryo type
Stage
Interaction
Bonferroni post hoc tests
comparing IVF vs. SCNT
2-cell
8-cell
Morula
Blast.
BAX
0.2828
0.1660
0.9309
ns
ns
ns
ns
BCL2
0.1664
<0.0001
<0.0001
ns
ns
ns
****
CDX2
0.0008
0.9596
0.7664
ns
ns
ns
ns
C-MYC
0.6040
0.1570
0.0222
ns
*
ns
ns
CX43
0.0297
0.7419
0.8253
ns
ns
ns
ns
DNMT3A
0.0311
0.0131
0.0119
ns
**
ns
ns
ELF5
<0.0001
0.0020
0.0019
****
ns
**
ns
EOMES
0.0233
<0.0001
0.0022
ns
****
ns
ns
ESRRB
0.0844
0.6047
0.2943
ns
ns
ns
ns
ETS2
0.0323
0.0060
0.2611
ns
ns
ns
*
G9A
0.0393
0.0492
0.6920
ns
ns
ns
ns
GATA6
0.0044
0.0336
0.0006
ns
**
***
ns
GNAS
0.3710
0.0001
0.1090
ns
ns
ns
ns
HDAC1
0.8995
0.0338
0.6881
ns
ns
ns
ns
HDAC2
0.0016
<0.0001
0.0245
ns
ns
**
ns
IGF2R
0.0873
0.0040
0.1380
ns
*
ns
ns
JARID2
0.8297
<0.0001
0.1254
ns
ns
ns
ns
KAT2B
0.9355
0.0001
0.8978
ns
ns
ns
ns
KDM3A
0.0122
0.0175
0.2170
ns
ns
ns
*
KDM5A
0.1368
0.0066
0.3187
ns
ns
ns
ns
KLF4
0.0216
0.0019
0.0019
ns
****
ns
ns
LIN28
0.0498
0.9456
0.8495
ns
ns
ns
ns
MASH2
0.0172
0.4921
0.4740
ns
ns
ns
ns
MDM2
0.3576
0.0864
0.1530
ns
ns
ns
ns
P21WAF1/Cip1
<0.0001
0.1330
0.0820
ns
**
ns
ns
PEG3
0.0092
<0.0001
0.0282
**
ns
ns
ns
RASGFR
0.0352
0.1162
0.0006
ns
****
ns
ns
SOX2
0.0021
0.0487
0.0494
ns
ns
**
ns
TEAD4
0.0939
<0.0001
0.0430
ns
**
ns
ns
UBE3A
0.3657
0.0109
0.3088
ns
ns
ns
ns
ZIM2
0.5831
0.0032
0.0292
ns
*
ns
ns
Note: Fold changes were calculated using the Δ-ΔCt method using the averaged values of
GAPDH and ACTB as an internal control. Individual fold changes for each IVF and SCNT
embryos were measured against the mean IVF ΔCt at the corresponding embryo stage. The
maximum Ct value of 40 was assigned to all embryos negative for gene expression in order to
calculate fold change. Statistical analysis employed a two-way ANOVA with Bonferroni posthoc test for multiple comparisons to measure significance between stage-matched IVF and
SCNT embryos (Bonferroni post-hoc tests comparing IVF vs. SCNT column) as well as across
embryo stages for both IVF and SCNT samples (Two-way ANOVA column). *, P < 0.05; **,
P < 0.01; ***, P < 0.001; ****, P < 0.0001. Additional analyses were performed using the
Student’s t-test with Welch’s correction compared IVF to SCNT embryos at each
developmental stage (not shown); results of those analyses confirmed those obtained by the
ANOVA shown here.
151
Figure 4-1. Heat map of all genes analyzed for expression analysis. Each cell of the
heat map represents the relative expression of individual SCNT embryos and IVF
embryos normalized to the mean expression for IVF at each developmental stage. The
number of embryos included in the analyses is shown in parentheses. Colors represent
the log2 fold change for embryos positive for expression for each gene, with red
indicating elevated expression and blue indicating low expression. Cells colored gray
represent embryos for which the corresponding transcript was not detected.
152
BAX
2-Cell
15
8-Cell
Morula
Blastocyst
Log2 Ratio
10
5
0
-5
-10
-15
IVF
27%
SCNT
42%
IVF
0%
SCNT
36%
IVF
0%
SCNT
0%
IVF
25%
SCNT
58%
BCL2
15
2-Cell
8-Cell
Morula
Blastocyst
Log2 Ratio
10
5
0
-5
-10
-15
IVF
100%
SCNT
83%
IVF
73%
SCNT
91%
IVF
100%
SCNT
80%
IVF
67%
SCNT
100%
P21WAF1/Cip1
15
2-Cell
8-Cell
Morula
Blastocyst
Log2 Ratio
10
5
0
-5
-10
-15
IVF
27%
SCNT
50%
IVF
9%
SCNT
100%
IVF
0%
SCNT
0%
IVF
0%
SCNT
42%
Figure 4-2. Log2 fold change expression values of BAX, BCL2 and P21WAF1/Cip1.
Open bars represent individual IVF embryos and solid bars represent individual SCNT
embryos at the 2-cell, 8-cell, morula, and blastocyst stages. Fold change was calculated
using the Δ-ΔCt method using the averaged values of GAPDH and ACTB as an internal
control as described in Materials and Methods. Fold changes for individual IVF and
SCNT embryos were calculated relative to the mean IVF ΔCt at the corresponding
embryo stage. For those embryos with no transcript detection, a maximum Ct value of 40
was assigned in order to calculate the fold change. The percentage of embryos for which
the indicated transcript was detected is shown below each Figure (N = 10 to 12 embryos).
153
BAX
BCL2
Relative expression
(Log2 fold change)
10
10
5
5
0
0
-5
2-Cell
8-Cell
Morula
Blast.
Embryo development stage
-5
IVF
SCNT
2-Cell
8-Cell
Morula
Blast.
Embryo development stage
Figure 4-3. Average log2 fold change expression values of BAX and BCL2. Open
bars represent individual IVF embryos and solid bars represent individual SCNT embryos
at the 2-cell, 8-cell, morula, and blastocyst stages. Fold change was calculated using the
Δ-ΔCt method using the averaged values of GAPDH and ACTB as an internal control as
described in Materials and Methods. Fold changes for individual IVF and SCNT
embryos were calculated relative to the mean IVF ΔCt at the corresponding embryo
stage. For those embryos with no transcript detection, a maximum Ct value of 40 was
assigned in order to calculate the fold change (N = 10 to 12 embryos).
154
Figure 4-4. Log2 fold change expression values of IGF2R, CX43, PEG3 and ETS2.
(following page) Open bars represent individual IVF embryos and solid bars represent
individual SCNT embryos at the 2-cell, 8-cell, morula, and blastocyst stages. Fold
change was calculated using the Δ-ΔCt method using the averaged values of GAPDH and
ACTB as an internal control as described in Materials and Methods. Fold changes for
individual IVF and SCNT embryos were calculated relative to the mean IVF ΔCt at the
corresponding embryo stage. For those embryos with no transcript detection, a maximum
Ct value of 40 was assigned in order to calculate the fold change. The percentage of
embryos for which the indicated transcript was detected is shown below each Figure (N =
10 to 12 embryos).
155
IGF2R
2-Cell
Log2 Ratio
20
8-Cell
Morula
Blastocyst
10
0
-10
IVF
0%
SCNT
0%
IVF
0%
SCNT
45%
IVF
0%
SCNT
0%
IVF
25%
SCNT
33%
CX43
2-Cell
Log2 Ratio
20
8-Cell
Morula
Blastocyst
10
0
-10
IVF
27%
SCNT
50%
IVF
18%
SCNT
45%
IVF
27%
SCNT
50%
IVF
58%
SCNT
58%
PEG3
2-Cell
Log2 Ratio
20
8-Cell
Morula
Blastocyst
10
0
-10
IVF
82%
SCNT
100%
IVF
100%
SCNT
100%
IVF
100%
SCNT
100%
IVF
100%
SCNT
92%
ETS2
Log2 Ratio
20
8-Cell
Morula
Blastocyst
10
0
-10
2-Cell
IVF
46%
SCNT
58%
IVF
9%
SCNT
36%
IVF
0%
SCNT
0%
IVF
17%
SCNT
58%
156
Figure 4-5. Log2 fold change expression values of C-MYC, LIN28, SOX2 and KLF4.
(following page) Open bars represent individual IVF embryos and solid bars represent
individual SCNT embryos at the 2-cell, 8-cell, morula, and blastocyst stages. Fold
change was calculated using the Δ-ΔCt method using the averaged values of GAPDH and
ACTB as an internal control as described in Materials and Methods. Fold changes for
individual IVF and SCNT embryos were calculated relative to the mean IVF ΔCt at the
corresponding embryo stage. For those embryos with no transcript detection, a maximum
Ct value of 40 was assigned in order to calculate the fold change. The percentage of
embryos for which the indicated transcript was detected is shown below each Figure (N =
10 to 12 embryos).
157
C-MYC
2-Cell
20
8-Cell
Morula
Blastocyst
Log2 Ratio
15
10
5
0
-5
IVF
0%
SCNT
0%
IVF
9%
SCNT
73%
IVF
0%
SCNT
0%
IVF
25%
SCNT
33%
LIN28
2-Cell
20
8-Cell
Morula
Blastocyst
Log2 Ratio
15
10
5
0
-5
IVF
73%
SCNT
50%
IVF
9%
SCNT
82%
IVF
64%
SCNT
40%
IVF
17%
SCNT
50%
SOX2
2-Cell
20
8-Cell
Morula
Blastocyst
Log2 Ratio
15
10
5
0
-5
IVF
91%
SCNT
92%
IVF
100%
SCNT
91%
IVF
100%
SCNT
100%
IVF
100%
SCNT
92%
KLF4
2-Cell
20
8-Cell
Morula
Blastocyst
Log2 Ratio
15
10
5
0
-5
IVF
0%
SCNT
0%
IVF
0%
SCNT
55%
IVF
0%
SCNT
0%
IVF
0%
SCNT
0%
158
Figure 4-6. Representative immunofluorescence images for P53 and phosphorylated
CDK2 in IVF and SCNT embryos. Alexa Fluor 488 detection of P53 (green) and
Alexa Fluor 594 detection of phosphorylated CDK2 (pCDK2, red) are shown in 8-cell
and blastocyst stage embryos. Nuclei were stained using Hoeschst (blue). A merged
image for all three fluorescence detection channels is shown. Images are representative
of 12 to 18 embryos.
159
Figure 4-7. Representative immunofluorescence images for P21 and BAX in IVF
and SCNT embryos. Alexa Fluor 488 detection of P21 (green) and Alexa Fluor 594
detection of BAX (red) are shown in 8-cell and blastocyst stage embryos. Nuclei were
stained using Hoeschst (blue). A merged image for all three fluorescence detection
channels is shown. Images are representative of 12 to 18 embryos.
160
B
Apoptotic cells/embryo
10
8
IVF
SCNT
*
6
4
2
0
8 cell
Blastocyst
Figure 4-8. Representative images and evaluation of apoptosis in SCNT and IVF
embryos. (A) Cells undergoing apoptosis in 8-cell and blastocyst stage embryos were
detected by TUNEL assay with Alexa Fluor 594 (red). Nuclei were stained using
Hoeschst (blue). A merged image for both fluorescence detection channels is shown.
Images are representative of embryos at the 8-cell (IVF, N = 23; SCNT, N = 14) and
blastocyst (IVF, N = 48; SCNT, N = 39) embryos. (B) A bar graph indicates the average
number of DNA fragmented nuclei per embryo at the 8-cell and blastocyst stages. Open
bars represent IVF embryos and solid bars represent SCNT embryos. Data is expressed
as mean and ± SEM. *, P < 0.05 indicates statistically significant difference between
IVF and SCNT groups according to two-way ANOVA with Bonferroni post-hoc test for
multiple comparisons.
161
CHAPTER 5
CONCLUSION
Over the past two decades, researchers have focused intensely on the problems
associated with SCNT in an effort to improve the success rate of the technique. Several
advances in the cloning procedure have been made, but have only led to modest
improvements in success rates. Despite these collective efforts, SCNT remains
inefficient, and the potential benefits from generating cloned offspring remain unrealized.
My dissertation research was concerned with determining the factors that affect
SCNT embryo development. One line of inquiry of my work characterized the degree of
apoptosis in SCNT embryos to determine whether programmed cell death is responsible
for reduced embryo cell number and clone failure. We compared gene and protein
expression of several genes involved in apoptosis in cloned embryos, as well as measured
several factors that may play a role in activating apoptosis. Additionally, we assessed
DNA degradation and confirmed that SCNT embryos have higher rates of apoptosis
during preimplantation development. Results of these studies indicate that SCNT
embryos have higher rates of apoptosis at the late developmental stage. These
observations were confirmed by gene and protein expression of several apoptosis factors,
including the apoptosis regulator P53, which was expressed at higher levels in SCNT
embryos compared to embryos produced via IVF. Our results also showed upregulation
of several genes that likely trigger apoptosis within SCNT embryos, including C-MYC,
LIN28, SOX2, and KLF4. These findings help illuminate the cause of embryo failure and
162
may lead to interventions that increase the success rate for development of SCNT
embryos by targeting the apoptosis pathway.
A second line of inquiry of my research was to determine whether the methylation
signatures of the developmental genes POU5F1, NANOG, and SOX2 were appropriately
reprogrammed in SCNT embryos prior to implantation. These genes are essential for
early development, and the aberrant reprogramming of methylation patterns would likely
lead to embryo degradation. Using bisulfite sequencing, we measured the methylation
status of CpG islands associated with each gene and compared SCNT embryos to IVF
and donor cell samples. We determined that specific CpG sites of all three genes were
abnormally established prior to implantation. The genes NANOG and SOX2 contained
CpG sites that were more similar to those of the donor cell than patterns observed in IVF
embryos. POU5F1 also contained aberrant levels of methylation, but POU5F1 was
hypomethylated relative to IVF embryos. Our results suggest that genes essential for
development are aberrantly methylated, which may affect embryo developmental
success.
One constraint of methylation studies is the limited data that can be obtained from
experiments. The technique of choice for methylation analysis uses sodium bisulfite, a
chemical that alters the sequence of DNA. Although the technique allows for single base
pair analysis of a sequence of interest, sodium bisulfite degrades DNA and renders it
unstable. Because of the problem of degraded DNA, amplified regions must be short,
thus limiting the amount of data that can be generated during sequencing. These
163
problems are compounded when using preimplantation embryos as a source of DNA, as
the low amount of starting material markedly limits the scope of experiments.
As few as five years ago, analysis of methylation required expansion of DNA
fragments in bacterial colonies, followed by pyrosequencing; at most, researchers would
gather 80 to 100 sequences for analysis. This limitation has recently been overcome with
the advent of high-throughput sequencing technology. Next generation sequencing is
contributing to more robust experiments and generating data sufficient to draw
conclusions regarding the targeting and maintenance of methylation patterns. Such
advances allow for more expansive data collection that can be applied to research
questions involving cloned embryos. Most research in the area of DNA methylation has
focused on methylation associated with CpG islands, but advances in the scope of data
collection allow for the study of methylation in non-CpG island regions, including
intragenic DNA. Knowledge of the methylation status of all CpG sites for the genome of
a cloned embryo would facilitate a greater understanding of epigenetic reprogramming
following SCNT.
A second challenge in methylation studies is the limited understanding of the
biological consequences of DNA methylation. It is well understood that methylation of
regions near transcriptional start sites is associated with transcriptional silencing, but the
mechanism(s) directing the targeted methylation of specific genes or loci within a gene
promoter region remain poorly understood. While experiments may reveal the pattern of
methylation levels associated with genes, the elucidation of apparent differences in
methylation status remains difficult due to a lack of knowledge about the interpretation of
164
this methylation code by transcription regulators. For example, do specific methylation
sites direct activation of transcription or is regional methylation critical for a specific
gene? Is there a necessary threshold of CpG methylation frequency within a regulatory
region? Without a better understanding of the mode of silencing following DNA
methylation, interpretation of results is limited.
In order to address some of these problems and to expand our knowledge
regarding how methylation targets and maintains a patterns within cells, we characterized
the methylation pattern of a pseudogene family in order to assess how the genome treats
identical sequences of DNA. We examined four pseudogenes derived form the LIN28
gene, as well as methylation of the parent gene itself. Results of these analyses showed
that pseudogenes are not methylated with the same pattern as the parent gene, but rather
maintain methylation levels specific to the region into which the pseudogene was
inserted. These findings indicate that methylation maintenance enzymes are targeted to
different genomic regions within the cell, and may ultimately lead to greater insights
regarding the methylation machinery.
The research reported in this dissertation has added to the body of knowledge of
the mechanisms of methylation regulation and helps to answer the question of why SCNT
embryos fail. Understanding the role of apoptosis in SCNT development may lead to
better procedures, culture conditions or treatments that will better enable embryos to
develop following SCNT. Targeted disruption of the apoptosis pathway may yield
improvements in the cloning procedure. If apoptosis could be briefly abated during
preimplantation development, SCNT implantation and birth rates may increase.
165
However, it is possible that apoptosis functions to selectively remove dysfunctional or
damaged blastomeres from the developing embryo and that inhibition of this process
could allow for the inappropriate expansion of aberrant cells. Additionally, by
characterizing the reprogramming inefficiencies of SCNT embryos, this work contributes
to the body of knowledge of epigenetic regulation. Identifying epigenetic factors that
may be responsible for reduced embryo viability and introducing information that
suggests how methylation occurs differently in different regions of DNA may reveal
better procedures to partially reprogram donor cells prior to nuclear transfer. Continued
research based on these conclusions may lead to more effective procedures for SCNT and
greater success of the technique.
166
CURRICULUM VITAE
Aaron Patrick Davis
Utah State University
College of Agriculture and Applied Sciences
Department of Animal, Dairy and Veterinary Sciences
4815 Old Main Hill
Logan, UT 84321-4815
[email protected]
(435) 730-1117
Education
2006 – Present
2002 – 2006
Doctor of Philosophy: Molecular Biology
Utah State University
Bachelor of Arts
Utah State University
Professional Experience
2007 – 2012
Teaching Assistant
Utah State University
Department of Animal, Dairy and Veterinary Science
2008
Visiting Researcher
College of Veterinary Medicine, Texas A&M University
Learned techniques relating to RNA Interference
2006 – 2007
Research Assistant
Utah State University
Department of Animal, Dairy and Veterinary Science
2004 – 2006
Research Technician
Laboratory of Kenneth White, Utah State University
Teaching Experience
2011 – 2012
Lab Instructor
Animal Molecular Biology
2007 – 2010
Lab Instructor
Animal Anatomy and Physiology
167
2010
Summer Academy Mentor
Mentored Visiting Student in Independent Research Project
2008 – 2011
Teaching Assistant
Introduction to Animal Science
2007 – 2009
Lab Instructor
Reproductive Physiology
Awards
2010
Distinguished Service Award
Presented by Associated Students of Utah State University
2007 – 2012
Departmental Assistantship
Animal Dairy and Veterinary Sciences, Utah State
2006 – 2010
Doctoral Student Tuition Award
Utah State University
2005
Undergraduate Researcher of the Year
College of Agriculture, Utah State University
Professional Service
2010
Vice President of Research
Graduate Student Senate
Associated Students of Utah State University
2010
Symposium Chair
2010 Intermountain Graduate Research Symposium
2009
Vice President of Research
Graduate Student Senate
Associated Students of Utah State University
2009
Symposium Chair
2009 Intermountain Graduate Research Symposium
2007 – 2008
College of Agriculture Senator
Graduate Student Senate
Associated Students of Utah State University
168
Professional Societies
American Society for Cell Biology
Publications
Aston KI, LI GP, Hicks BA, Sessions BR, Davis AP, Rickords LF, Stevens JR,
White KL. 2010. Abnormal levels of transcript abundance of developmentally
important genes in various stages of preimplantation bovine somatic cell nuclear
transfer embryos. Cell Reprogram. 12(1):23-32.
Aston KI, LI GP, Hicks BA, Sessions BR, Davis AP, Winger QA, Rickords LF,
Stevens JR, White KL. 2009. Global gene expression analysis of bovine somatic
cell nuclear transfer blastocysts and cotyledons. Mol Repro Dev. 76(5):471-82.
Sessions BR, Aston KI, Davis AP, Pate BJ, White KL. 2006. Effects of amino
acid substitutions in and around the arginine-glycine-aspartic acid (RGD)
sequence on fertilization and parthenogenetic development in mature bovine
oocytes. Mol Reprod Dev. 73(5):651-7.
Publications in Preparation
Davis AP, Benninghoff AD, Sessions BR, Meng Q, Thomas AJ, White KL. (In
preparation) DNA methylation reprogramming of the POU5F1, NANOG, SOX2
and KLF4 genes following somatic cell nuclear transfer. Mol Repro Dev
Davis AP, Benninghoff AD, Sessions BR, Thomas AJ, White KL. (In
preparation) DNA methylation of the Lin28 pseudogene family. BMC Genomics
Davis AP, Meng Q, Sessions BR, White KL, Benninghoff AD. (In preparation)
Activation of the P53 pathway following somatic cell nuclear transfer.
Reproduction
Abstracts
Davis AP, Benninghoff AD, Meng Q, Sessions BR, Thomas AJ, White KL.
Activation of the P53 apoptosis Pathway following somatic cell nuclear transfer
Presented at the ADVS Graduate Student Symposium. August 6, 2012. Logan,
UT.
Davis AP, White KL. Gene expression of P53 and MDM2 following somatic cell
nuclear transfer. Presented at the Intermountain Graduate Research Symposium.
March 31, 2010. Logan, UT.
169
Davis AP, Sessions BR, Meng Q, Innes E, Osborne B, White KL. Identification
of methyl-specific DNA binding proteins of the bovine Nanog and POU5F1
promoter regions using electrophoretic mobility shift assay. Presented at the
American Society of Cell Biology 49th Annual Meeting. December 5-9 2009, San
Diego, CA.
Davis AP, Bayles AH, Davies C, Innes EW, Meng Q, Osborne B, Sessions BR,
White KL. Bisulfite sequencing analysis of POU5F1, C-MYC, LIN28 and
NANOG following bovine somatic cell nuclear transfer. Presented at the
American Society of Cell Biology 48th Annual Meeting. December 13-17 2008,
San Francisco, CA.
Invited Presentations
Davis AP, White KL. Sodium bisulfite sequencing shows aberrant epigenetic
reprogramming following somatic cell nuclear transfer. Presented at the Center
for Integrated BioSystems Research Seminar. November 2008 Logan, UT.
Davis AP, White KL. DNA methylation patterns of POU5F1 and NANOG in
single in vitro fertilized embryos. Presented at the Center for Integrated
BioSystems Research Seminar. October 2007 Logan, UT.