76
From the Department of Molecular Biosciences, The Wenner-Gren Institute, Stockholm Sweden Heart Regeneration: Lessons from the Red Spotted Newt Nevin Witman Stockholm 2013

Heart Regeneration - DiVA - Simple search

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Heart Regeneration - DiVA - Simple search

From the Department of Molecular Biosciences, The Wenner-Gren Institute, Stockholm Sweden

Heart Regeneration: Lessons from the Red Spotted Newt

Nevin Witman

Stockholm 2013

Page 2: Heart Regeneration - DiVA - Simple search

All previously published papers herein were reproduced with permission from the publisher. Printed in Sweden by Universitetsservice US-AB Stockholm 2013. Distributor: Stockholm University library © Nevin Witman, 2013 ISBN 978-91-7447-726-9 Front Cover: Shows a photomicrograph of a regenerated newt ventricle immunolabled with MHC (red) and BrdU (green) at 65dpi.

Page 3: Heart Regeneration - DiVA - Simple search

ABSTRACT Unlike mammals, adult salamanders possess an intrinsic ability to regenerate

complex organs and tissue types, making them an exciting and useful model to study

tissue regeneration. The aims of this thesis are two fold, (1) to develop and

characterize a reproducible cardiac regeneration model system in the newt, and (2) to

decipher the cellular and molecular underpinnings involved in regeneration.

In Paper I of this thesis we developed a novel and reproducible heart

regeneration model system in the red-spotted newt and demonstrated for the first time

the newt’s ability to regenerate functional myocardial muscle, following resection

injury, without scarring. The observed findings coincide with an increase in several

developmental cardiac transcription factors, wide-spread cellular proliferation of

cardiomyocytes and non-cardiomyocyte populations in the ventricle and reverse-

remodeling at later time points during regeneration. Of further interest was the

identification of functionally active Islet1+ve and GATA4+ve cardiac precursor cells in

and around regenerating areas. The observation of such cell types further compels the

similarity between mammalian cardiac development and newt cardiac regeneration

and justifies these animals as suitable model organisms for studying heart

regeneration.

In Paper II we wanted to decipher the molecular cues possibly driving

cardiac regeneration in newts. Here we used qualitative and quantitative methods to

delineate the function microRNAs (miRNAs) have in this process. One interesting

candidate, namely miR-128, a known tumor suppressor miRNA and regulator of

myogenesis, was found to have a regulatory role in controlling hyperplasia during

newt cardiac regeneration. We show that treating regenerating newt hearts with

miR128 antagomirs evoked an increase in cellular proliferation within non-

cardiomyocyte populations. Of further interest is the discovery of a novel binding site

of miR-128 in the 3’UTR of Islet1. We speculate that the natural increase in miR-128

expression levels during newt cardiac regeneration functions as a fine-tuning

mechanism to control cellular proliferation of precursor cells.

The question of how newts regenerate has long been postulated. It has been

thought that although gene products found in the newt are highly homologous to those

of mammals, there may be functional differences. It has also been debated whether

newts are able to more efficiently and effectively utilize transcriptional or post-

transcriptional modifiers to diversify their gene pool. In Paper III of my thesis we

Page 4: Heart Regeneration - DiVA - Simple search

sought to examine these questions by exploring if a link exists between RNA editing,

a wide-spread post-transcriptional process and regeneration. We observed that A-to-I

editing enzymes (ADARs) are present and active in regenerating newt tissues. In

regenerating tissues, at times concomitant with increased cellular proliferation, we

discovered a localized nuclear to cytoplasmic shift of ADAR1 expression. This

activity of ADAR1 during regeneration may be partly responsible for driving the

cellular plasticity that is needed during multiple phases of tissue regeneration in the

red-spotted newt.

The major findings of this thesis can be summarized as follows: that the red

spotted newt is able to regenerate cardiac tissue, in the absence of scarring, and

restore function to the heart following resection injury. The response of the newt heart

to injury is governed by an increase in known developmental cardiac transcription

factors, which peak at a time when cellular proliferation is also heightened. We have

also shown that miRNAs are differentially expressed during various phases of heart

regeneration. One novel candidate, miR-128, appears to have some control over

proliferating cell types in the heart. We also show that Islet1 is a novel target of miR-

128, indicating that miR-128 may also contribute to the re-specification of

differentiating cardiac precursors. Finally, we observe that ADAR enzymes are

functionally active in regenerating tissue. This implies a notion that RNA editing may

act as a global regulator involved in several aspects of tissue regeneration. These

discoveries provide additional information as to understanding the molecular

pathways that are activated during regeneration and may one day help to amend the

regeneration deficiencies that are found in humans.

Page 5: Heart Regeneration - DiVA - Simple search

TABLE OF CONTENTS

LIST OF PUBLICATIONS ..................................................................................................... 7 LIST OF ABBREVIATIONS .................................................................................................. 8 PROLOGUE ............................................................................................................................. 9 1. REGENERATION ............................................................................................................. 11

1.1 ORGAN REGENERATION IN A HISTORICAL PERSPECTIVE.................................................. 11 1.2 TYPES OF REGENERATION ............................................................................................... 13

1.2.1 Reparative regeneration ...................................................................................... 14 1.2.1.1 Cellular regeneration ................................................................................................... 15 1.2.1.2 Tissue regeneration ...................................................................................................... 15

1.2.2 Physiological regeneration .................................................................................. 15 1.2.3 Hypertrophic regeneration ................................................................................. 16

2. NEWT AS AN IDEAL REGENERATION MODEL ..................................................... 17 2.1 ECOLOGY AND HABITAT ................................................................................................. 17 2.2 WHY THE NEWT IN RESEARCH ON REGENERATION? ........................................................ 17 2.3 EXAMPLES OF NEWT REGENERATION .............................................................................. 18

2.3.1 Limb regeneration ............................................................................................... 19 2.3.2 Lens regeneration ................................................................................................ 19 2.3.3 Spinal chord and neuronal regeneration ........................................................... 20

3. CARDIOGENESIS ............................................................................................................ 22 3.1 VERTEBRATE CARDIAC DEVELOPMENT .......................................................................... 22

3.1.1 Cardiac progenitors and organogenesis ............................................................. 22 3.1.2 Transcriptional regulation .................................................................................. 24

3.2 CARDIOVASCULAR DISEASE ........................................................................................... 25 3.3 CARDIAC THERAPIES ...................................................................................................... 26

3.3.1 Stem cells .............................................................................................................. 26 3.3.2 Reprogramming ................................................................................................... 28

4. MODEL ORGANISMS OF HEART REGENERATION .............................................. 29 4.1 MAMMALS...................................................................................................................... 29 4.2 FISH ................................................................................................................................ 30 4.3 AMPHIBIANS ................................................................................................................... 31

5. MICRORNAS ..................................................................................................................... 35 5.1 MICRORNA BIOGENESIS ................................................................................................ 35 5.2 MICRORNAS IN REGENERATION .................................................................................... 37 5.3 MICRORNAS AND CARDIAC REGENERATION .................................................................. 38

6. RNA EDITING ................................................................................................................... 42 6.1 A-TO-I RNA EDITING ..................................................................................................... 42

6.1.1 ADAR enzymes .................................................................................................... 42 6.2 EDITING SUBSTRATES ..................................................................................................... 44

6.2.1 Editing in protein-coding regions ....................................................................... 44 6.2.2 Editing of non-coding RNA ................................................................................. 45

6.3 FUNCTIONAL ASPECTS OF RNA EDITING ........................................................................ 46 7. PRESENT INVESTIGATIONS ........................................................................................ 48

7.1 AIMS .............................................................................................................................. 48 7.2 PAPER I ........................................................................................................................... 49

7.2.1 Results and discussion ......................................................................................... 49 7.3 PAPER II ......................................................................................................................... 51

7.3.1 Results and discussion ......................................................................................... 51 7.4 PAPER III ........................................................................................................................ 53

7.4.1 Results and discussion ......................................................................................... 53 7.5 FUTURE PERSPECTIVES ................................................................................................... 55

8. CONCLUSIONS ................................................................................................................ 58

Page 6: Heart Regeneration - DiVA - Simple search

9. ACKNOWLEDGEMENTS ............................................................................................... 59 10. REFERENCES ................................................................................................................. 63

Page 7: Heart Regeneration - DiVA - Simple search

7

LIST OF PUBLICATIONS

I. Nevin Witman, Barri Murtuza, Ben Davis, Anders Arner, and Jamie I.

Morrison (2011). Recapitulation of developmental cardiogenesis governs

the morphological and functional regeneration of newt hearts following

injury. Developmental Biology, 354:1. pp.67-76.

II. Nevin Witman, Jana Heigwer, Barbara Thaler, Weng-Onn Lui, and Jamie I.

Morrison (2013). miR-128 regulates non-myocyte hyperplasia, deposition

of

extracellular matrix and Islet1 expression during newt cardiac

regeneration. Manuscript under consideration by Developmental Biology. Original

submission date: June 7, 2013

III. Nevin M Witman, Mikaela Behm, Marie Öhman, and Jamie I. Morrison

(2013). ADAR-Related activation of adenosine-to-inosine RNA editing

during regeneration. Stem Cells and Development, 22:16. pp.2254-67

Page 8: Heart Regeneration - DiVA - Simple search

8

LIST OF ABBREVIATIONS

ADAR Adenosine Deaminases Acting on RNA

BrdU 5-bromo-2’-deoxyuridine

CVD Cardiovascular Disease

CHF Congestive Heart Failure

CDC Cardiac derived cell

CMC Cardiomyocyte

CPC Cardiac progenitor cell

DNA Deoxyribonucleic acid

dpi days post-injury

dsRBD double-stranded RNA binding domain

dsRNA double-stranded RNA

EPDC Epicardial derived progenitor cells

ESC Embryonic Stem Cell

FGF Fibroblast Growth Factor

FHF First Heart Field

GFP Green Fluorescent Protein

LAD Left anterior descending (artery)

SCI Spinal Chord Injury

SHF Secondary Heart Field

MI Myocardial Infarction

miRNA microRNA

mRNA messenger RNA

ncRNA non-coding RNA

NES nuclear export signal

NLS nuclear localization signal

lncRNA long non-coding RNA

RISC RNA induced silexing complex

RNA Ribonucleic acid

RNAi RNA interference

UTR Untranslated Region

Page 9: Heart Regeneration - DiVA - Simple search

9

PROLOGUE The capacity of the human body to repair injured tissues and organs following

damage, disease or degeneration via the natural ageing process is limited and

decreases with age. The field of regenerative biology aims to repair, replace or in

someway augment damaged tissues and organs. At the time of this thesis, a great

majority of treatments for chronic and life threatening diseases are merely palliative.

That is, they are aimed at alleviating the symptoms associated with disease, not

repairing the injury.

An organ with a very limited ability to self-repair following injury is the heart,

with many cardiovascular diseases (CVD) becoming more prevalent with age. The

onset of vascular disease can be gradual, usually with weak symptoms until the onset

of an acute heart attack, which leaves the heart muscle damaged. Cardiac failure can

also develop as a result of several other conditions, e.g. genetic changes in the

musculature or hypertension. The damaged sustained from heart disease is often

irreversible and patients who are living with CVD may suffer a diminished quality of

life. Novel therapies need to be developed.

Unlike humans, some non-mammalian vertebrates such as fish and

amphibians have emerged as excellent models to study regeneration, as they are

natural regenerators. One such example of these amazing natural regenerators is the

red-spotted newt, Notophthalmus viridescens. In what seems like a boundless list,

newts can regenerate many complex organs and cell types including the limb, lens of

the eye, retina, jaw and neurons of the central nervous system. The regeneration of all

these organs and tissues are done intrinsically, without the delivery of stem cells or

transplantations. Furthermore, these aquatic salamanders do not seem to have a

decline in their regenerative capabilities throughout their adult life. Taken together

these characteristics make newts an ideal model to study the processes underlying

tissue regeneration.

The work carried out during my PhD tenure focused on mechanisms of

regeneration of the newt heart after a resection injury. The heart of the newt is able to

regenerate the myocardium following a resection injury during a 60-day period. The

magnitude of this injury casued by the resection corresponds to the functional loss of

heart cells in patients suffering from severe cardiac injury. Additionally I examined

gene-expression profiles of known cardiac genes during heart regeneration and show

that the newt heart regenerates following a proliferative response. I also examined

regulation of the newt heart response to injury at a molecular level, by studying the

Page 10: Heart Regeneration - DiVA - Simple search

10

involvement microRNAs (miRNAs) had during the regeneration process. Finally I

studied a post-transcriptional regulatory mechanism, RNA editing, which may act to

some degree as a global regulator of tissue regeneration.

In the background of this thesis I will provide a detailed description of

regenerative research, including types of regeneration and mechanisms involved in

regenerating organisms. I will then highlight why the red-spotted newt is an excellent

model organism in which to study regeneration. Next, I will introduce the reader to

cardiac development and disease. The importance of cardiac therapies as well as

benefits from using model organisms to study heart regeneration will be discussed.

Subsequently, I will introduce non-coding RNAs, highlighting miRNAs as a

significant avenue for further exploring molecular mechanisms involved in

regenerative research. Lastly, I will briefly discuss the importance of RNA editing as

a post-transcriptional modifier that may be involved as a global regulator in tissue

regeneration.

Page 11: Heart Regeneration - DiVA - Simple search

11

1. REGENERATION

1.1 Organ regeneration in a historical perspective

The regeneration of body parts has captured the attention of many different

cultures and mythologies for thousands of years. One of the earliest dated tales of

regeneration comes from Greek mythology, when Homer and Hesiod recount stories

of the half human, half god Prometheus. The story goes that after Prometheus stole

fire from Olympus to aide mankind, Zeus punished him by chaining him to a rock

where every night he would have his liver pecked out by an eagle. Doomed by

immortality, the following day his liver would regenerate and so he would endure the

same punishment night after night for thousands of years. Whether the Ancient

Greeks were aware of the livers natural ability to regenerate at that time is uncertain,

but ironically we know today the liver is an organ with a remarkable ability to

regenerate throughout adult life (Pack et al., 1962).

Some of the first documented scientific discoveries of regeneration were

recorded by the well know philosopher and scientist, Aristotle (384-322BC). In his

book “Historia Animalium” Aristotle notes that the tails of lizards and snakes are able

to regenerate (as cited in (Dinsmore, 1996)). Although we now know the regenerated

lizard tail is not a perfect replica, with many anatomical and functional differences to

uninjured tails (Ritzman et al., 2012).

Though there were observational scientific discoveries of regeneration in the

ancient times, experimental documentation of regeneration came much later. In 1712

René-Antoine Ferchault de Réaumur (1683-1757), a French scientist, reported a

descriptive study of limb and claw regeneration in crayfish (as cited in (Dinsmore,

1991)). He concluded that one possible explanation as to why crustaceans can

regenerate appendages and humans cannot regenerate theirs is because crustacean

appendages break easily at the joints. This fascinated other scientists of the 18th

Century such as Abraham Trembley, Charles Bonnet and Lazzaro Spallanzani.

Abraham Trembley (1710-1784) was a Swiss naturalist who would later

become known for his scientific discoveries of regeneration in freshwater polyps. He

discovered when this animal is bisected, the half containing the head will regenerate a

new tail and the half containing the tail a new head. These animals would later be

named Hydra, due to the regenerative characteristics they shared with the Greek

Page 12: Heart Regeneration - DiVA - Simple search

12

mythological creature Hercules battled. Interestingly Hercules was also the hero who

freed Prometheus from his daily hepatectomies.

During the same time regeneration was being discovered in Hydra,

Trembley’s cousin Charles Bonnet (1720-1793) documented the annelid worm’s

capability to regenerate new segments multiple times following resection (as cited in

(Birnbaum and Alvarado, 2008)). He noted the rate in which a part regenerated

correlated to the temperature in which the organism was kept, i.e. the cooler the

temperature the slower the regenerative response. Although Bonnet dabbled in

regeneration biology for a few years, he is perhaps more well known for his discovery

of parthenogenesis – a form of asexual reproduction, which he discovered whilst

studying plant lice (aphids). This concept would later fuel scientists to learn more

about embryo development (Davis, 2012).

In 1769 an Italian scientist named Lazzaro Spallanzani (1729-1799) would be

the first to describe detailed studies on vertebrate regeneration. He showed that

amphibians, such as prematamorphic frogs and toads could regenerate their tails. He

would also be the first to show that salamanders possessed a remarkable ability to

regenerate their limbs, tails and jaws. Spallanzani sketched and documented the

formation of a thin round stump near the injury site that formed shortly after

amputation. Today we know this as the blastema, a small bud-like structure

containing a cell mass which gives rise to the new structure (Tsonis and Fox, 2009).

Regeneration would even attract the attention of the well-known British

scientist, Charles Darwin (1809-1882). On Darwin’s famous voyage aboard the

Beagle he observed the regenerative properties of several species of planarian.

However, even Darwin’s work in this area remained primarily descriptive. In the

years to come questions around regeneration would be depicted in higher resolution

and more detailed experimental studies would surface.

Thomas Hunt Morgan (1866-1945) was an American evolutionary biologist

and is well known for receiving the Nobel Prize in medicine in 1933 for discoveries

linking the chromosome and heredity. However, Morgan also explored regenerative

research and believed that by studying regeneration phenomena one could learn more

about the development of organisms. He hypothesized this because he felt that

regeneration, like development, required differentiation of cell types and thus was part

of a fundamental growth process, (as cited in (Esposito, 2013)). This view was at

arms with other theoretical positions of the time, such as those belonging to August

Weismann (1834-1914), which pinned regeneration as a special adaptation to some

Page 13: Heart Regeneration - DiVA - Simple search

13

organisms, (as cited in (Esposito, 2013)). The latter theory coincides with the notion

that lower organisms adapted regenerative traits particularly in organs that were most

frequently subject to injury. In one investigation Morgan provided evidence against

Weismann’s theory by showing hermit crabs (Eupagurus longicarpus) were able to

regenerate their distal limbs following amputation, even though distal limbs were not

prone to injury (as cited in (Esposito, 2013)). Whether or not regeneration is an

adaptive trait or an evolutionary trait lost in more complex organisms remains widely

debated amongst developmental biologists even today.

So why is the evolution of regeneration important? If regeneration is

evolutionarily ancestral across metazoan species, there should exist conserved traits

making it possible to coax repair mechanisms in mammals to better regenerate.

However, if regeneration has been adapted independently of evolution across different

species, then understanding the species-specific selectivity of regeneration may also

be helpful in identifying new molecular activities that could be adapted in human

molecular pathways.

1.2 Types of regeneration

It was Thomas Hunt Morgan who first began to standardize terminology in

regenerative biology (as cited in (Sanchez Alvarado, 2000)). He classified the

regeneration of Metazoans into two basic groups 1) regeneration that occurs via

rearrangement in the absence of cell proliferation and 2) regeneration that requires

cellular proliferation. The first is referred to as Morphallaxis, a term coined by

Morgan in 1901, which describes the replacement of a missing part via the

reorganization of pre-existing parts without the need for cellular proliferation.

Morphallaxis is most commonly seen in several species of invertebrates such as

Hydra. Although it had previously been shown that Hydra can regenerate parts of the

body when DNA synthesis is inhibited, more recent work supports the notion of

proliferating cells with a blastema-like feature that contribute to the regenerative

process (Chera et al., 2009; Cummings and Bode; 1984, Hicklin and Wolpert, 1973).

Another model organism considered to undergo morphallactic regeneration is

the planaria. Planaria are flatworms with bilateral symmetry and a remarkable ability

to regenerate. Morgan himself demonstrated the ability of these flatworms to

regenerate a whole animal from as little as 1/279th the body (as cited in (Sanchez

Alvarado, 2000)). Recent research has shown a small population of pluripotent stem

cells named neoblasts undergo proliferation and differentiation to replace lost tissue

Page 14: Heart Regeneration - DiVA - Simple search

14

following resection in the planaria (Wagner et al., 2011). Indeed the reorganization of

tissue does occur in regenerating organisms making these two terms not mutually

exclusive. However it would more recently appear that cellular proliferation is a

requirement for most regenerating organisms, making the strict term of morphallaxis

slightly outdated (Agata et al., 2007).

According to Morgan’s specific definition, the second group of regeneration,

epimorphosis, refers to the case of regeneration where the proliferation of material

precedes and interacts with the development of the new part, (as cited in (Reddien and

Alvarado, 2004)). There is a broad range of divergent species capable of epimorphic

regeneration including worms (Wagner et al., 2011), insects (Anderson and French,

1985) and starfish (Thorndyke et al., 2001). Here the activation of proliferating cells

is the underpinning regulatory effect in restoring resected or damaged tissue. The

most prominent examples of epimorphic regeneration are seen in urodele amphibians

(see section on newt limb regeneration).

To date the classification of regeneration has been slightly modified.

Epimorphic regeneration is generally discussed as reparative regeneration that is

mediated through a blastema, utilizing a mechanism involving dedifferentiation,

proliferation and transdifferentiation (Sanchez Alvarado, 2000). Transdifferentiation

describes the process of a committed cell type in one tissue lineage that is able to

convert into a cell of an entirely distinct lineage. These cells lose their tissue-specific

markers along with their tissue specific function of their previous lineage type and

acquire the onset of new markers and functions in the transdifferentiated cell type.

One mammalian example of transdifferentiation occurs during development of the

oesophagus. It has been shown that smooth muscle cells in developing rodents can

switch phenotype favoring that of skeletal muscle immediately after birth

(Patapoutian et al., 1995). In lower vertebrates perhaps the most prominent and well

known example of transdifferentiation occurs during lens regeneration in the newt

(see section on newt lens regeneration and reviewed extensively (Barbosa-Sabanero et

al., 2012)).

1.2.1 Reparative regeneration

The common element in reparative regeneration is generally the exact

functional replacement of a lost or resected part of the body, with no scarring. Aquatic

salamanders, such as the red-spotted newt (Notophthalmus viridescens) possess much

more extensive regenerative capabilities than mammals. Amongst them is cellular

Page 15: Heart Regeneration - DiVA - Simple search

15

dedifferentiation – a process that involves the regression of a differentiated cell to a

simpler, more embryonic-like form (Jopling et al., 2011). Following injury,

dedifferentiation can create a milieu of stem cell-like cells that can proliferate and

undergo re-differentiation to form mature tissue specific cell types. Reparative

regeneration can occur at the single cell level or occur in multicellular body parts.

1.2.1.1 Cellular regeneration

Apart from neurons, few mononuclear cell types are known to regenerate or

repair in mammals. Axonal transections lead to a short period of degeneration,

followed by an extension phase providing new cytoplasmic material and interaction

with its immediate environment (Gillen et al., 1997). Following pathological or

alterations in physiological conditions, oligodendrocyte progenitors have been shown

to proliferate and provide mature oligodendrocytes (Barnabe-Heider et al., 2010).

1.2.1.2 Tissue regeneration

The replacement of damaged or worn out tissue such as bone and skeletal

muscle often rely on cells that are not the mature, dominant cell type in order to

regenerate. For example, satellite cells are adult, tissue-specific stem cells and are the

major contributors to the regeneration of mammalian skeletal muscle (Collins et al.,

2005; Mauro, 1961). Although these cells have been shown to give rise to the

myofiber component of skeletal muscle, muscle regeneration often involves a number

of other cell types, such as fibroblasts, endothelial cells, smooth muscle cells and

Schwann cells (Mauro, 1961).

1.2.2 Physiological regeneration

Physiological regeneration within adult vertebrates generally occurs in organs

or tissue sources that require the reconstitution of differentiated cell types that are

generally short-lived. In mammals, the most prominent examples include the

replacement of blood lineage cell types, skin cells and the gut epithelium, which rely

on regular homeostatic replacement. Other mammalian and non-mammalian examples

include the shedding of exoskeletons in crustaceans and snake skin, the annual cycle

of feather replacement in birds and the regrowth of deer antlers (Stoick-Cooper et al.,

2007). With the large variation of examples of physiological regeneration, what

categorizes the phenomena is reconstitution in order to meet physiological needs to

ultimately maintain the equilibrium of tissues.

Page 16: Heart Regeneration - DiVA - Simple search

16

1.2.3 Hypertrophic regeneration

Most adult organisms have some means of dealing with injury, damage or

disease even if complete tissue regeneration is no alternative. A regulatory process

developed in most mammals (including humans) to deal with such loss or damage

includes the process of hypertrophic responses. For example, humans can survive up

to two-thirds hepatectomy, which results in severe remodeling and an increase in

overall mass in the remaining lobe(s) of the liver in order to compensate for the

increase in functional demand (Michalopoulos, 2007). In mammals, and apart from

the liver, some smooth muscle organs also elicit hypertrophic responses to damage or

disease. Following urinary infection or urinary outlet obstruction in the rat, the

smooth muscle cells that make up muscle bundles of the bladder become larger and

longer. Although no mitoses are generally found, cells with two nuclei are often

present (Gabella and Uvelius, 1990). In mammals, nephrectomy results in a

compensatory renal enlargement of the other kidney (Nagaike et al., 1991). Toxic

damage to the kidney is often counteracted by the growth of the remaining nephrons,

however some evidence supports the presence of kidney stem cells (Reule and Gupta,

2011). Hypertrophic regeneration generally encompasses a restorative response

throughout the entirety of the organ, as opposed to a regenerative response that only

replaces tissue at the precise site of damage.

An adult organism with a remarkable ability to regenerate many complex

tissue types is the red-spotted newt and will be discussed in greater detail in the next

section.

Page 17: Heart Regeneration - DiVA - Simple search

17

2. NEWT AS AN IDEAL REGENERATION MODEL

2.1 Ecology and habitat

Newts are aquatic vertebrates that belong to the Caudata group of urodele

amphibians. Urodele amphibians make up the group of metazoans containing newts

and salamanders, which sets them apart from frogs as they retain their tails as adults.

In particular, the red-spotted newts (Notophthalmus viridescens) (Rafinesque, 1820)

are studied extenisvely in experimental biology due to their natural abilities to

regenerate. These newts are inhabitants of North-Eastern America. They are known

for having a complex life cycle that encompasses three distinct life stages: aquatic

larva or tadpole, red eft or terrestrial juvenile and aquatic adult. Larval newts hatch

with external gills, short front legs but the hind legs are absent. Within a week of

hatching the larvae become active, feeding on small aquatic life and grow quickly

(Hamilton, 1940). During larval metamorphosis the animals reabsorb their gills,

develop lungs and emerge from the water as terrestrial juveniles (Shi and Boucaut,

1995). The juvenile newt usually appear orange in color with bright red spots and this

stage of the life cycle can last anywhere from two to seven years, although two to

three years on land is most common (Hurlburt, 1969). The newt may migrate

relatively far distances until finding a suitable aquatic habitat (Gill, 1978). Once the

newts have discovered a new suitable environment they undergo a second form of

metamorphosis, after which they are sexually mature adults and appear olive-green in

color but keep their juvenile red spots (Figure 1) (Brockes and Kumar, 2005). Newts

will spend the majority of their adult life primarily in water. In the wild newts are

estimated to live up to 15 years, but in captivity newts have been reported to live

longer- an impressive life-span for an animal of its size (Hillman et al,. 2009).

2.2 Why the newt in research on regeneration?

By utilizing organisms that retain the ability to regenerate complex body parts

and tissue-types throughout adult life, scientists can construct model systems in the

hope of understanding natural regeneration and thus understand why it may be

blocked in humans. Contrary to mammalian organisms, which have very little

postnatal regenerative ability, urodele amphibians such as the newt are an ideal model

system for studying regeneration. Urodele amphibians share similar biological

pathways with mammals, thus allowing comparative analysis to be performed with

Page 18: Heart Regeneration - DiVA - Simple search

18

humans. Additionally, by comparing how mammals and salamanders respond to

injuries we can identify pertinent genes or proteins that can either be introduced or

inhibited in the mammalian system in the hope that regeneration efficiency can be

increased. The majority of somatic cell types found in the newt, including those that

make up the heart, express similar markers as those found in humans. Furthermore, as

mentioned previously the newt undergoes several stages of metamorphosis, ultimately

providing us with a comparative adult-vertebrate model system.

Figure 1. Photograph of an adult red spotted newt. An adult newt submerged in an aquatic environment. Note the olive-green color and characteristic red spots. The adult newt grows to approximately 8cm in length with an average weight of approximately 2g. Photograph by Nevin Witman.

2.3 Examples of newt regeneration

Salamanders have been of great interest to experimental biologists, beginning

with the work of previously mentioned Lazzaro Spallanzani. In this section I will

concentrate on discussing some of the most well characterized models of regeneration

in red-spotted newts.

Page 19: Heart Regeneration - DiVA - Simple search

19

2.3.1 Limb regeneration

Newt limbs, whether distally (mid-ulna and radius bone) or proximally

amputated (mid-humerus bone), proceed through three timed stages of regeneration

(Iten and Bryant, 1973). Following amputation, migrating epidermal cells quickly

cover the amputation surface forming the wound epithelium, which protects the

internal structures from the external environment and delivers the signals necessary to

initiate and maintain regeneration. Dedifferentiated cells become concentrated just

below the newly formed wound epidermis that produces a bud-like growth at the tip

of the limb stump called the blastema, which becomes vascular and neurogenic as it

matures (Rageh et al., 2002). The rapid morphologic outgrowth of the blastema leaves

the amputated limb structurally identical to that of the uninjured limb, although

slightly smaller than the original. The regenerate will continue to grow in size until

reaching the same dimensions of an uninjured limb. The exact origin and identity of

the dedifferentiated cells in the blastema remain uncertain, but it was long believed

these cells are multipotent with no lineage restrictions. Interestingly, it was shown

that amputating the forelimbs stimulates a population of satellite cells, which

contribute to new limb tissues in a similar manner to mammalian repair (Morrison et

al., 2006). Moreover through the use of GFP transgenic tracking experiments in

another salamander, the axolotl (Ambystoma mexicanum), a study showed that the

mass of undifferentiated progenitors from each tissue source produces progenitors

that are lineage restricted (Kragl et al., 2009). Further work is necessary to fully

understand the origin of cells needed to regenerate the full complement of cells

present within the newt limbs. Worth mentionining is additional work that has shown

limb regeneration is nerve dependent. A newt-specific secreted nerve factor called

anterior gradient protein (nAG) rescues regeneration in de-nervated newt limbs

(Kumar et al., 2007). This nerve factor is expressed in Schwann cells of the mylinated

nerve sheaths and has a role in improving blastemal cell proliferation in vitro.

2.3.2 Lens regeneration

Adult newts possess the ability to regenerate a new lens following lentectomy

(Tsonis and Del Rio-Tsonis, 2004). Following removal of the lens, pigmented

epithelial cells of the iris transdifferentiate. The epithelial cells of the dorsal iris first

dedifferentiate and subsequently change phenotype by differentiating into lens cells.

However, the same epithelial cells from the ventral iris do not appear to undergo these

events.

Page 20: Heart Regeneration - DiVA - Simple search

20

In an effort to address the question of whether ageing or repetitive injury

stunts tissue regeneration, this same group conducted a 16 year-long experiment on

lens regeneration (Eguchi et al., 2011). No significant changes in growth rates or sizes

of regenerated lenses were reported upon repeated removal of the lens after they had

re-grown. Additionally, no significant differences were observed in morphological

and gene expression characteristics compared between control animals that had never

undergone regeneration and animals undergoing repetitive regeneration. The findings

signify regenerative capabilities are not diminished upon repetition in aged animals.

Fibroblast growth factors (FGFs) have additionally been shown to elicit an effect on

lens regeneration. For example, inhibition of FGF-1 blocks lens regeneration from the

dorsal iris, yet the up-regulation of FGF-1 stimulates excess dedifferentiation of the

iris and retina cells (Del Rio-Tsonis et al., 1998).

2.3.3 Spinal chord and neuronal regeneration

Adult newts possess the ability to recover function following damage to the

spinal cord. Following a complete transection injury, the newt spine and tail

regenerates, with the animals regaining use of their hind limbs in as little as 4 weeks

(Davis et al., 1990). The exact mechanisms of spine regeneration are unclear, but

some elegant work suggests that subsequent to tail amputations a blastema forms, the

spinal cord elongates and axons restore connections as they grow into the newly

developing tissues (Singer et al., 1979). The process appears to be governed in a

manner that reiterates developmental programming and processes. A more recent

model has been developed to follow spinal cord injury (SCI) in newts without the

need to perform tail amputations, with the idea being that this type of SCI mimics

more closely the spinal injuries most commonly seen in humans (Zukor et al., 2011).

Following SCI in the newt, an environment consisting of a loose extracellular matrix

permits axon regeneration across the lesion. However, following SCI in humans, the

environment of the injury site becomes inhibitory to regeneration due to the formation

of dense scarring (Norenberg et al., 2004). Of further interest is the recent discovery

that transplanted cultured neurospheres can contribute to the regeneration of the

damaged spinal chord (McHedlishvili et al., 2011).

To determine tissue regenerative abilities in the newt brain, previous studies

attempted to surgically resect small portions of the frontal lobe. After 8 months it was

reported that most the resected area had been restored, but the origin of new cells was

not thoroughly investigated (Okamoto et al., 2007). Additionally it has been shown

Page 21: Heart Regeneration - DiVA - Simple search

21

that newts can regenerate neurons of the central nervous system following toxin

induced cell death (Parish et al., 2007). The authors here reported a 30-day post-lesion

regeneration time and the presence of BrdU+ve dopaminergic neurons, indicating

regeneration fueled by neurogenesis. Interestingly the same group has recently shown

the importance of dopamine as a senescence activator of stem cells and also a

controller of neuronal homeostasis (Berg et al., 2011).

Over the last 10 years the heart has become another organ of interest in

regenerative biology. I will next discuss why heart regeneration has gained significant

attention in the field of regeneration as well as present current research and

considerable discoveries within the field.

Page 22: Heart Regeneration - DiVA - Simple search

22

3. CARDIOGENESIS

“How can you describe this heart in words, without filling a whole book?” -Leonardo da Vinci, 1513

(Note left by Leonardo da Vinci next to an anatomical sketch of a human heart.)

As this quote illustrates, Leonardo da Vinci was highly fascinated by the heart.

He would become one of the first to recognize the heart as a muscle and begin to

provide detailed medical illustrations of the heart. Apart from discovering anatomical

and physiological characteristics of the heart, da Vinci observed post-mortem

degeneration to the heart and blood vessels in the elderly thus linking the ageing

process to degeneration. A better understanding of developmental cardiac biology

offers the potential to broaden current knowledge of mechanisms involved in both

acquired and congenital heart diseases.

3.1 Vertebrate cardiac development

To discuss vertebrate heart development in great detail across vertebrates

would be too extensive for this thesis. More in-depth information regarding distinct

comparisons between cell types and transcription factor networks across vertebrate

heart development have been made previously (Pandur et al., 2013). However, I will

concentrate on the more general findings, features and mechanisms between

vertebrates, with the main focus relating to mammalian heart development.

3.1.1 Cardiac progenitors and organogenesis

During vertebrate embryonic development the heart is the first organ to form

and function, and is paramount for survival during embryonic life. The precise

development of the vertebrate heart requires highly controlled cellular organization.

The migration and differentiation of cardiac progenitor cells contribute to the

morphology of the heart and it’s distinct compartments. The events leading to heart

formation in vertebrates is homologous. The sequence of events include the formation

and looping of the heart tube, elongation and growth via the extensive addition of

progenitor cells and morphogenesis of the cardiac chambers, cushions and valves

(extensively reviewed (Vincent and Buckingham, 2010)).

In zebrafish, it has been shown that myocardial lineages separate during

midblastula stage and will provide cells needed to create atrial and ventricular

Page 23: Heart Regeneration - DiVA - Simple search

23

myocardium (Stainier et al., 1993). Notable discoveries using the zebrafish have

helped to further describe the cellular and molecular mechanisms activated during

vertebrate heart development (Bakkers, 2011).

In amphibians and mammals two myocardial lineages appear to diverge from

a common progenitor, which contribute to two distinct phases of heart development,

the first heat field (FHF) and the second heart field (SHF) (Figure 2A) (Brade et al.,

2007; Meilhac et al., 2004). The FHF makes up the left ventricle and partial atria

formation, whereas the adjacent SHF makes up the right and left ventricle as well as

the outflow tract (Figure 2) (Buckingham et al., 2005; Kelly et al., 2001). However, it

is speculated that both lineages probably share in some common contributions to

other regions of the heart. Interestingly the FHF is a focused assortment of

differentiating cell types and morphological growth, whereas the SHF actively

contributes proliferating cardiac precursors in the early stages of heart development.

Figure 2. Diagram depicting the major steps during the development of the vertebrate heart. First a cardiac crescent is formed which comprises of two heart fields that contain different populations of cardiac precursors. The first heart field (FHF, red) contributes to the left ventricle (lv) and the secondary heart field (SHF, blue) contributes to the right ventricle (rv) left and right atria (la, ra), and outflow tracts (ot). From (Bruneau, 2008), modified and reprinted with permission.

The cells that make up the early heart originate from mesoderm lineage. These

cells form the primitive streak at the midline of the embryos, creating a contracting

linear heart tube (Figure 2B). The linear tube will proceed to expand and undergo

major twisting and looping events, while septa and valves begin to form between the

chambered compartments of atria and ventricles (Figure 2C). The heart will continue

to undergo morphological changes and maturation events during later stages of fetal

life (Figure 2D).

A B C D

Page 24: Heart Regeneration - DiVA - Simple search

24

T-box transcription factors are expressed in the cells that make up the

primitive streak and first reveal a cardiac niche (Costello et al., 2011). Mesp1 has also

been shown to be a primitive marker of cardiac progenitor cells and may be a master

regulator for early cardiac cell fate (Bondue et al., 2008). Several other early stage

markers of cardiac progenitors include members of the GATA family of zinc finger

transcription factors (i.e. GATA 4/5/6) and the NK-2 class of homeodomain-

containing transcription factors (Nkx2-5). These are amongst the earliest known genes

to be expressed in cells facing cardiac lineage fate and their expression plays

paramount roles in heart formation (Peterkin et al., 2005).

The LIM homeodomain family Islet1 is a critical transcription factor with

essential roles in cardiac development, specifically the SHF. Although recently Islet1

protein has been found in the cardiac crescent, which suggests an earlier role during

the FHF (Prall et al., 2007). Evidence supporting the importance of this transcription

factor was previously reported, showing that murine Islet1 knockouts have severely

deformed hearts (Cai et al., 2003). Despite the function of Islet1 previously being

defined as a progenitor marker recent data suggests Islet1 may regulate multi-potency

of epicardial cells into cell types harboring mesenchymal features (Brønnum et al.,

2013; Bu et al., 2009; Zhou et al., 2008).

It was generally believed that the generation of endothelial, cardiac and

smooth muscle cells arise from distinct embryonic precursors during cardiogenesis.

However, it has recently been demonstrated that Islet1 progenitor cells can generate

all three cardiovascular lineages (Moretti et al., 2006). A subset of Islet1 positive

undifferentiated cells are detected shortly after birth as well as in the postnatal heart in

rodents and humans (Laugwitz et al., 2008; Laugwitz et al., 2005). Islet1 has not been

detected in terminally differentiated cardiomyocytes in the adult mammalian heart.

However, these cells can develop into functional cardiomyocytes in vitro based on

their marker gene expression, contractile properties and intracellular calcium release

(Laugwitz et al., 2005).

3.1.2 Transcriptional regulation

The activation of genes encoding cardiogenic transcription factors

meticulously manages the development of the heart along with its endless function.

Transcription factors are required for regulating cardiogenesis and morphogenesis, as

well as components regulating contractility (Olson, 2006). For example, the

expression of Nkx2-5 is a paramount transcription factor critical for terminal

Page 25: Heart Regeneration - DiVA - Simple search

25

differentiation of myocardial cells (Lyons et al., 1995). It was more recently shown

that Nkx2-5 expression functions to maintain a balance between progenitor

proliferation and differentiation (Prall et al., 2007).

Serum response factor (SRF) and the GATA factors have been shown to

regulate the expression of multiple cardiac genes and initiate cardiac differentiation

(Niu et al., 2008; Zhao et al., 2008). The heart and neural crest derivatives (HAND)

basic helix-loop-helix family of transcription factors control aspects of chamber

differentiation (Srivastava et al., 1997). The HAND proteins are selectively expressed

in adult ventricular CMCs (HAND1 and HAND2) and atrial CMCs (HAND2), which

are severely down-regulated in patients with cardiomyopathies (Natarajan et al.,

2001).

Often cardiac transcription factors work synergistically, by regulating gene

expression to co-regulate heart development (He et al., 2011). Networks of these

interacting transcription factors are slowly being defined, however their full

complement of target genes has yet to be determined. One example is the ability for

Nkx2-5 to interact with the GATA factors, which can regulate cardiac genes such as

Mef-2c and connexin40 (Dodou et al., 2004; Linhares et al., 2004). Gene and protein

expression involved in signal transduction pathways including transcriptional

programs, can be severely affected following biomechanical stress to the heart. As

such, developmental signaling pathways (which involve cardiogenic transcription

factors) and stem cells are emerging as novel promising candidates to study in relation

to improving the treatment of cardiovascular disease.

3.2 Cardiovascular disease

Cardiovascular diseases include a broad range of diseases that affect the

cardiovascular system, which comprises the heart and blood vessels. Cardiovascular

disease continues to account for more deaths then any other disease worldwide.

Ischemic heart disease is the most common debilitating cardiac injury in humans,

which often results in heart failure. Ischemia is usually caused by coronary artery

obstruction and leads to a myocardial infarction (MI). An MI occurs when there is an

interruption of blood supply to the heart, causing heart cells to die due to lack of

oxygen. The subsequent loss of cardiac cell number along with ischemic induction of

necrosis, results in the replacement of damaged heart muscle with fibroblasts, which

help to form a fibrotic scar. Even though the scar tissue elicits some mechanical

support shortly after the infarct, it predisposes for arrhythmias. Additionally, the

Page 26: Heart Regeneration - DiVA - Simple search

26

inability for the newly formed scar formation to contract consequently diminishes the

overall function of the heart and if left untreated can lead to heart failure and death

(Anversa et al., 2006). It is interesting to note that heart attacks were very uncommon before the 20th

Century. It was in 1912 that James Herrick (1861-1954), an American physician

would become one of the first physicians to describe heart disease as a result of

hardening of the arteries (James, 2000). Since then, the global scourge of heart

disease has been described as a pandemic. With this said, rather then relying on heart

transplants it has become essential that we produce novel therapies to combat heart

disease.

3.3 Cardiac therapies

The field of regenerative medicine is rapidly evolving, with much hope

recently being placed into stem cell therapies and molecular interventions. Stem cells

are cell types which exist in many organs of the human body, they are self-renewable

and act to replace worn out or damaged cells. Stem cell transplantations are being

widely used across the globe where injection into the diseased part of a body can

stimulate an endogenous healing response or be used to replace damaged or diseased

cells and tissues. Ultimately, scientists are turning to stem cells to replace damaged

and dead cardiomyocytes (CMCs), the major “powerhouse” cell type of the heart.

3.3.1 Stem cells

Embryonic stem cells (ESCs) are perhaps the most suitable cell type for

cardiovascular regeneration, as they are self-renewing and pluripotent. ESCs are

capable of giving rise to all three embryonic germ layers (the mesoderm, the

endoderm and the ectoderm) and therefore retain the potential to differentiate into any

of the functionally mature cell types of the human body (Thomson et al., 1998).

Along with the legal and ethical hurdles that are currently limiting the use of ESCs,

there are many technical issues to overcome. These challenges include the viability of

the transplanted cells that may lead to aggressive tumor formation, coupling of ESCs

to host cardiomyocytes and overcoming susceptibility to immunological rejection

(Yoon et al., 2005). Although there have been some reports of injected ESCs

improving rodent heart contractility and diminished scar sizes following ischemic

injuries, together the ethical and technical obstacles have hindered the use of ESCs for

in vivo myocardial regeneration in humans (Min et al., 2003; van Laake et al., 2007).

Page 27: Heart Regeneration - DiVA - Simple search

27

The presence of adult cardiac stem cell / progenitor cell (CPCs) populations in

the heart have been reported by several groups. Such cell types have been identified

based on different cell surface markers and in vitro assays (Beltrami et al., 2003; Cai

et al., 2003; Matsuura et al., 2004; Messina et al., 2004; Smart et al., 2011). The use

of such cell types for the potential generation of de novo cardiomyocytes are looking

promising, however the mechanisms by which CPCs facilitate improvement in the

heart may be paracrine affiliated (Gnecchi et al., 2008; Huang et al., 2011). For a

detailed overview of cardiac progenitor cell biology and other adult stem cell

therapies, the reader may refer to the following review (Leri et al., 2011).

Of even greater interest, and avoiding some of the ethical implications that are

embryonic stem cell accountable, is the ability of cardiac derived cells (CDCs) to

form 3-dimensional cardiospheres. Explant cultures of surgical or endomyocardial

biopsies can be cultured in order to isolate cardiospheres (Messina et al., 2004; Smith

et al., 2007). Cardiospheres form spherical clusters of cardiac progenitor cells that are

clonogenic and undifferentiated cells (Figure 3) (Davis et al., 2009). Of critical

importance is the fact that CDCs can be expanded many times over on fibronectin

coated plates, providing suitable cell numbers for cell therapies. Very recently the

cardiosphere approach was used in a phase I clinical trial that yielded positive results

(Makkar et al., 2012). Following delivery of CDCs in a randomized trial, patients with

MI showed improved ejection fractions, reduction in scar formation and increases in

heart mass after 6 months.

Figure 3. Examples of cardiospheres immunostained with different cardiac antigens. Cultured cardiac explants can form 3-dimensional spherical clusters of cells that can either express myogenic lineages such as (A) myosin heavy chain (MHC) and (B) cardiac troponin I (cTnI), or express progenitor cell markers such as Nkx2.5 (C). From (Davis et al., 2009). Reprinted with permission from PLOS.

Page 28: Heart Regeneration - DiVA - Simple search

28

3.3.2 Reprogramming

The technique of reprogramming adult cells into pluripotent cells or into an

alternative cell type could hold great potential for regenerative therapies. The

technology to convert one cell type into another by forced expression of foreign DNA

or transcription factors has been known for sometime (Davis et al., 1987; Takahashi

and Yamanaka, 2006).

For example, it has been shown that by overexpressing fibroblasts with the

skeletal muscle transcription factor MyoD, these cells could be converted to skeletal

muscle cells (Davis et al., 1987). Moreover it has been shown that the forced

expression of myocardin can convert fibroblasts into smooth muscle cells (Wang et

al., 2003). However, to date, no single factor has been shown to force fibroblasts into

a cardiomyocyte phenotype. Previously it was reported that fibroblasts could be

converted into CMC-like cells in vitro by expressing a cocktail of transcription factors

including GATA-4, Mef2c, and Tbx5, albeit only a small percentage of these

fibroblasts appeared to be fully reprogrammed into beating CMC-like cells (Ieda et

al., 2010). Even more recently it was shown that resident cardiac non-myocytes such

as fibroblasts could be reprogrammed with this same cocktail of transcription factors

in vivo (Qian et al., 2012). Intriguingly these in vivo reprogrammed CMC-like cells

showed similar contractile properties to that of normal CMCs and were able to

electrically couple with endogenous CMCs. Despite some functional improvement in

the mice treated with these reprogrammed CMC-like cells following ligation injury,

the efficiency of such cells to be reprogrammed remained very low. In short, the combinations of cell and gene therapies are slowly showing

promise, but many problems are limiting their use in a clinical setting. Over-coming

such shortfalls of the activation of immune responses, improving delivery of

reprogrammed factors and advancing the survival rates of transplanted cells will

accelerate the potential for these therapeutic options.

Page 29: Heart Regeneration - DiVA - Simple search

29

4. MODEL ORGANISMS OF HEART REGENERATION

Due to the limited self-renewal capacity of the mammalian heart following

injury, original research approaches need to be explored in order to find ways to

ameliorate cardiac regeneration in humans. Another such avenue to pursue is to study

vertebrate animal models that can regenerate the heart. Lower vertebrate orders such

as teleost fish, and urodele amphibians contain model organisms that have previously

been shown to possess intrinsic abilities to regenerate a plethora of body parts and

tissues as adult organisms. Here I will discuss some of the vertebrate models of heart

regeneration and key findings in the field.

4.1 Mammals

In spite of the limited regenerative capacity of the four-chambered adult

mammalian heart to regenerate, coronary artery ligation models in rodents have been

employed as a tool to study damage associated with post myocardial infarction. One

of the first reports to describe coronary ligation to induce MI in rodents was produced

over 30 years ago (Zolotareva and Kogan, 1978). However, only more recently have

the morphological and functional effects been documented in more detail (Patten et

al., 1998; Salto-Tellez et al., 2004). The implementation of such a model system has

made it feasible to examine the effects of stem cell transplantations in repairing

damaged myocardium (see section on cardiac therapies). Although initial studies

report encouraging evidence for stem cells to transdifferentiate and functionally repair

ischemic damage, more recent reports have begun questioning these findings and

report new short-comings (Balsam et al., 2004; Murry et al., 2004; Orlic et al., 2001).

Major issues currently being noted in these model systems include the formation of

tumors following stem cell transplantations and the need for immunosuppressive

therapy to prevent host rejection (Swijnenburg et al., 2005). These investigations may

provide an advanced understanding of how to overcome several hurdles employing

stem cell technology in humans.

Given the capacity of neonatal cardiomyocytes to divide, a recent report

investigated whether the heart of the 1-day-old neonatal mouse could overcome

partial resection injury. It was shown that in as little as 3 weeks following resection

injury the morphological architecture of the mouse heart was restored and after 2

months the newly formed apex had normal contractile function (Porrello et al., 2011).

Page 30: Heart Regeneration - DiVA - Simple search

30

Interestingly, this ability to regenerate myocardium is lost after 7 days of age, a time

when cardiomyocytes become binucleate and withdraw from the cell cycle (Li et al.,

1996). Of further interest was the recent release of another report also revealing the

ability of the 1 day old neonatal mouse heart to respond to ischemic LAD – ligation

injury, an ability which is also lost after 7.5 days of age (Haubner et al., 2012). Taken

together these seminal investigations indicate a time when regenerative capacity of

the neonatal mouse heart is lost. By understanding the mechanisms that alter this

regenerative behavior, we may one day be able to restore the regenerative capabilities

that have been lost in adult mammalian hearts.

4.2 Fish

Interestingly, teleosts such as zebrafish (Danio rerio) and the giant danio

(Devario aequipinnatus) also possess a high degree of natural regenerative potential.

For over a decade the descriptive regenerative vigor of the two-chambered zebrafish

heart to overcome approximately 20% apical ventricular resection has been

acknowledged (Poss et al., 2002). The complete regenerative process is largely

unknown, although more molecular and genetic tools are being applied providing

critical clues. Previous evidence supported the notion that the regenerated

myocardium of the zebrafish heart arises from a subclass of epicardial progenitor cells

(Lepilina et al., 2006). It was proposed that the cardiac injury stimulates expansion of

the epicardium, where FGFs crosstalk with migrating epicardial cells to the injury

site, which slowly revascularize the heart. At a similar time myocardial progenitors

originate in the damaged areas where they begin to express pre-cardiac markers and

eventually associate with previously existing muscle.

The notion of resident progenitor cells influencing natural heart regeneration

has more recently been opposed using inducible transgenic zebrafish models to

specify cell types within the heart (Jopling et al., 2010; Kikuchi et al., 2010). These

more recent studies describe a mechanism of cardiomyocyte dedifferentiation and

proliferation to restore lost or damaged cardiac tissue in zebrafish. Interestingly, one

of the studies supported the notion of a heterogeneous population of adult

cardiomyocytes with proliferative potential (Kikuchi et al., 2010). This group found

that a subpopulation of GATA4 inducing cardiomyocytes is a major contributor to the

regeneration of resected zebrafish hearts. Interestingly GATA4 remained expressed

throughout the regeneration process. As GATA4 is a gene required for normal cardiac

development, this may imply that these cells activated a more embryonic program.

Page 31: Heart Regeneration - DiVA - Simple search

31

Cryocauterization injuries have recently been applied to zebrafish and are

believed to be an alternative model to coronary artery ligations based on the damage

sustained by the technique (van den Bos et al., 2005). This injury model in zebrafish

causes massive tissue necrosis, followed by large amounts of collagen deposits, which

are subsequently replaced with new cardiac muscle (Chablais et al., 2011; Gonzalez-

Rosa et al., 2011; Schnabel et al., 2011). The dynamics of regeneration appears to

vary between injury models, as recovery from cryocauterization appears to take

longer and perfect ventricular shape is not restored, as it appears to be following

resection (Gonzalez-Rosa et al., 2011; Poss et al., 2002). Another cauterization

damage model has recently emerged in the giant danio (Lafontant et al., 2011). Some

of the data obtained in this model system reveals the importance of neovascularization

to the damaged areas of the heart. It is proposed that the neovascularization may

mitigate the relationship between diminishing fibrosis and proliferating

cardiomyocytes to regenerate the damaged myocardium. Additional evidence

gathered in this report reveals the presence of proliferating cardiomyocytes, although

the authors do not address whether progenitor cells may also support the regeneration.

In an effort to understand if zebrafish can overcome larger injuries to the

cardiac system that relay signs of acute cardiac failure, a recent study employed an

inducible transgenic ablation system in CMCs (Wang et al., 2011). Following cell

specific depletion of 60% of the ventricular myocardium, activated cellular and

molecular responses rapidly regenerate the myocardium. It is interesting to note that

myocyte death alone is substantial enough to prompt a regenerative response in this

model, as opposed to signals emanating from resected tissue or blood clot formation.

A more recent report employing a similar cardiac specific transgenic ablation model

system showed that zebrafish utilize differentiated atrial cardiomyocytes to

transdifferentiate into ventricular cardiomyocytes ultimately aiding in ventricular

regeneration following ablation damage (Zhang et al., 2013). Although it is unknown

if mammals contain a similar transdifferentiation capacity in atrial myocytes, this

pivotal finding warrants a new possible source for endogenous cellular regenerative

therapy in patients with heart failure.

4.3 Amphibians

Previous studies sought to address whether the three-chambered newt heart in

contrast to mammalian hearts, can efficiently regenerate lost cardiac muscle. Initial

studies established that following apical ventricle resections, the adult newt

Page 32: Heart Regeneration - DiVA - Simple search

32

cardiomyocytes in adjacent areas could re-enter the cell cycle and were capable of

forming new cardiac fibers following injury (Oberpriller and Oberpriller, 1974).

However, this response was minimal with dense connective tissue replacing the

absent cardiac tissue, resulting in scarring. Intriguingly, subsequent work observed

that replacing the area of resected cardiac tissue with minced newt cardiac muscle

gave rise to a continuous ventricular wall and lumina, which was formed in the

absence of scarring (Bader and Oberpriller, 1978). However, this grafted area

appeared to be morphologically and functionally separate from the remaining

ventricle.

Unlike their mammalian counterparts, cultured cardiomyocytes from adult

newt hearts continually proliferate (Bettencourt-Dias et al., 2003; Matz et al., 1998).

Also cultured cardiomyocytes from newt hearts provided evidence for disparate sub

populations of cardiomyocytes (Bettencourt-Dias et al., 2003). The authors reported a

sub-population of cardiomyocytes, which stably arrest following mitosis or

cytokinesis, but uncovered another small percentage of cardiomyocytes that retain

proliferative potential. Understanding why such a block arises in some

cardiomyocytes but not others holds potential for augmenting cardiac regeneration in

mammals.

More current studies implemented a mechanical crush-injury model to the

newt heart, which damages the cardiac tissue without the need for resection. Using

forceps to squeeze newt heart ventricles until hemorrhage, a dramatic decrease of

sarcomeric protein expression levels was observed, coinciding with an increase in

mitotically-active cardiomyocyte-like cells in the damaged areas (Laube et al., 2006).

This group revealed for the first time that the regenerative response of the newt heart

to injury was not limited to a wound-healing and scarring scenario. A current in-depth

investigation lead by the same group analyzed morphological and ultrastructural

changes in crush-injury induced newt hearts. This study has provided insight that

extracellular matrix components may be more important then previously believed, as

this not only provided some mechanical stability, but considerably influenced the

reconstitution of damaged myocardium by providing guidance cues for new CMCs

(Piatkowski et al., 2012). With the previous evidence that newt CMCs have high

plasticity, it was believed that dedifferentiating CMCs provide the regenerate (Laube

et al., 2006). This more recent report supports a notion of immature CMCs as a source

for regeneration, as damaged CMCs immediately appear to undergo necrosis and are

removed by macrophages. This group could not exclude newly emerging CMCs may

Page 33: Heart Regeneration - DiVA - Simple search

33

come from a pool of progenitor cells and as such the source of cells responsible for

regenerating newt hearts remains undetermined.

A previous report revealed that the axolotl, Ambystoma mexicanum, could

positively respond to a partial ventricular resection injury (Cano-Martinez et al.,

2010). The axolotl heart showed recovery of contractile activity in as little as 30dpi,

with full morphological recovery between 30-90dpi. Interestingly, evidence for CMC

and non-CMC proliferation was evident based on proliferative responses seen using

BrdU pulse chase experiments.

Newts can impressively regenerate following substantial ventricular injury,

but how they manage to survive in the first place is just as astonishing. In an effort to

understand newt survival following needle puncture to the ventricle, a recent article

provided evidence that the heart of the Cynops pyrrhogaster, another urodele

amphibian, can redirect blood flow away from the injury site (Miyachi, 2011).

According to this report, the onset of valve hyperplasia immediately following

damage consequently changes the anatomical flow of the ventricular inflow and

outflow tracts, preventing additional hemorrhage, allowing the rapid induction of

repair and regeneration mechanisms to proceed.

In summary, different regeneration models provide different theories, which

endeavor to explain the regeneration phenomena in vertebrate species. Although

growth factors and other major signaling networks are most likely involved, the

bigger question remains – where are the cells coming from that contribute to natural

cardiac regeneration? There are two major speculations; 1) Differentiated cells re-

enter the cell cycle. There has been some evidence for the activation and proliferation

of cardiomyocyte populations contributing to the regenerating zebrafish hearts

following ventricular resection (Jopling et al., 2010; Kikuchi et al., 2010). Additional

information obtained in neo-natal mouse heart resection injuries also propose new

cardiomyocytes coming from pre-existing cardiomyocytes (Porrello et al., 2011). The

mechanism of action would require the mature cells of the heart to undergo

dedifferentiation, where they down-regulate their contractile genes allowing them to

re-enter the cell cycle and proliferate. The cells can then redifferentiate back to adult

cardiomyocytes where they can replenish the missing tissue. 2) Recruitment of

progenitor cells. There is also some support for the recruitment of epicardial

progenitor cells (EPDCs) that become active following cryocauterization induced

damage in zebrafish or induced myocardial infarction in mouse (Gonzalez-Rosa et al.,

2011; Smart et al., 2011). These cells may act to migrate into the myocardium and

Page 34: Heart Regeneration - DiVA - Simple search

34

repopulate dead cells in the damaged heart. Of further interest is the recent discovery

of in vivo reprogramming where atrial cardiomyocytes transdifferentiate and migrate

into ventricular cardiomyocytes providing an alternative method for repopulating

damaged myocardial tissue (Zhang et al., 2013).

Investigating how newts and teleosts remove fibrotic lesions, regulate

molecular pathways and control cell proliferation following severe cardiac damage,

provides an incentive to overcome the limitation of cardiac regeneration in mammals.

There are many aspects driving the control, maintenance and regulation of heart

regeneration in these lower vertebrate species. One such regulator is the involvement

of non-coding RNA’s such as microRNAs and they will be discussed in detail in the

following section.

Page 35: Heart Regeneration - DiVA - Simple search

35

5. MicroRNAs

In humans over 20,000 proteins are produced from just 1.2% of our genome

(Clamp et al., 2007; Mattick, 2011). It is also known that many more genes are

transcribed that give rise to non-protein-coding transcripts. Recently non-coding

RNAs (ncRNAs) have begun to emerge with diverse and significant roles in cellular

pathways that can affect physiological processes, including those found in cardiac

development and regeneration (Hudson and Porrello, 2013).

Two classes of ncRNAs include microRNAs (miRNAs) and long non-coding

RNAs (lncRNAs). The human genome encodes thousands of lncRNAs, many of

which have demonstrated affects on biological processes such as regulating gene

expression (Guttman and Rinn, 2012). Presently miRNAs are more widely

characterized so I will focus on the biogenesis and modulation of miRNAs

particularly in the context of regeneration below.

5.1 MicroRNA biogenesis

Since the initial discovery of miRNAs in the early 1990’s, it has been shown

that these small, evolutionary conserved, non-coding RNAs act as a novel mechanism

of post-transcriptional regulation (Fire et al., 1998; Lee et al., 1993). Recent

speculations have estimated that over one-third of mRNAs in the mammalian genome

are regulated by miRNAs (Lewis et al., 2005).

The processing of miRNAs occurs both in the nucleus and the cytoplasm of

the cell and involves a number of complex regulatory enzymes and binding proteins

(for a detailed review see (Winter et al., 2009)) (Figure 4). miRNAs are first

transcribed in the nucleus by RNA polymerase II and are referred to as primary

miRNAs (pri-miRNAs). These pri-miRNAs are 5’ capped, polyadenylated and range

in size, but on average are approximately 2kb in length (Cai et al., 2004). The pri-

miRNA is further microprocessed in the nucleus into hairpin RNAs by Drosha (an

RNase type-III enzyme) with the help of a double stranded RNA binding protein

DiGeorge syndrome critical region 8 (DGCR8) to form precursor miRNAs (pre-

miRNAs) approximately 70 nucleotide bases in length (Landthaler et al., 2004). The

pre-miRNAs are exported out of the nucleus into the cytoplasm via exportin-5, where

they are further processed by the ribonuclease Dicer, leaving a miRNA duplex. One

strand of the miRNA duplex (approximately 18-25 nucleotides long) guides the RNA-

Page 36: Heart Regeneration - DiVA - Simple search

36

induced silencing complex (RISC) to partial complimentary sequences of target

mRNAs where they then interfere with the regulation of mRNA stability or

translation (Mourelatos et al., 2002; Schwarz et al., 2003). The other strand is

typically but not always degraded (Yang et al., 2011). It is commonly supported that

the 5’ end of the miRNA, specifically from nucleotides 2 to 8 (known as the ‘seed

region’), plays a critical role in the pairing with the target-associated mRNAs in

metazoans (Bartel, 2009). The miRNA interaction with the mRNA transcript is

mediated via Watson-Crick base pairing with most miRNAs binding to the 3’-UTR of

their target mRNAs. Recent evidence suggests miRNAs can also interact with the 5’-

UTR and protein-coding regions of targeted mRNAs (Chi et al., 2009). The current

notion for the primary mechanism of action of miRNA – mRNA interactions are to

induce post-transcriptional gene silencing via decay (Guo et al., 2010) or repression

of translational initiation (Gregory et al., 2005). This is the mechanism of RNA

degradation seen by the RNAi machinery, where the Argonaute protein Ago2 is the

endonuclease responsible for cleaving the RNA target (Fire et al., 1998; Liu et al.,

2004; Meister et al., 2004; Rand et al., 2004; Rivas et al., 2005). Although the classic

dogma is that miRNAs most commonly act to repress their mRNA targets, some

evidence suggests that miRNAs could be involved in stimulating mRNA translation

(Vasudevan et al., 2007).

Page 37: Heart Regeneration - DiVA - Simple search

37

Figure 4. Schematic overview of miRNA biogenesis. The process of miRNA transcription is controlled by RNA-polymerase II, which produces long primary miRNAs (pri-miRNAs) typically with hairpin morphology. These pri-miRNA products are cleaved by a nuclear endonuclease, Drosha, which crops the stem of the pri-miRNA leaving a pre-miRNA. Pre-miRNAs are exported out of the nucleus by Exportin-5 and are further cleaved by Dicer, an RNAse III enzyme leaving a mature double-stranded miRNA. Mature miRNAs interact with Argonaute endonuclease proteins (Ago2) to become incorporated into the RNA-induced silencing complex (RISC), which together are directed towards mRNA targets and bind to their 3’ untranslated region (UTR). Image adapted from (Goettsch & Aikawa, 2013).

5.2 MicroRNAs in regeneration

Regeneration encompasses a plethora of dynamic and rapid alterations in gene

expression. Several modes of regulation must exist in order to meet the demands of

quick-acting and fine-tuned changes in altered gene expression ongoing in

regenerating tissue. Compelling evidence would then nominate miRNAs as one such

candidate for this mediation during complex tissue regeneration.

It has previously been shown through both gain-of-function and loss-of-

function experiments that FGF signalling depeletes miR-133 expression, which is a

critical component for optimal caudal fin regeneration in the zebrafish (Yin et al.,

2008). A more recent study comparing data sets from different species revealed many

Page 38: Heart Regeneration - DiVA - Simple search

38

similarities between axolotl spine regeneration and zebrafish fin regeneration,

including the down-regulation of miR-133 during regeneration (Sehm et al., 2009).

However, this study found miR-196 to be the most up-regulated and strongest

candidate in axolotl tail regeneration. miR-196 down-regulates Pax7 protein levels,

which in turn drove cells in the regeneration zone to proliferate and migrate out of the

blastema. This process helps mediate the development of the new ependymal tube,

illuminating the functional importance of miR-196 in tissue regeneration.

Additionally the inhibition of miR-196 resulted in excessively high levels of Pax7

proteins, which can inversely inhibit proliferation, leaving a defective regenerative

response.

As previously mentioned newt lens regeneration occurs via the

transdifferentiation of the epithelial cells of the dorsal iris, but not the ventral iris. To

try to elucidate the mechanisms underlying such specificity, the function of specific

miRNAs were explored to delineate their function during lens regeneration. Here, it

was shown that the up-regulation of miR-148 and let-7b not only affected cellular

proliferation, but interestingly regulated the levels of other miRNAs, indicating a

possible involvement of a miRNA network (Nakamura et al., 2010).

In order to reveal the importance of miRNAs during axolotl limb regeneration,

a current study employed and paired microarrays to identify key players of regulation.

Reportedly the relationship between miR-21 and its target Jagged1 (member of WNT

signaling pathway), are speculated to crucially regulate the commitment of

proliferating blastemal cells during limb regeneration (Holman et al., 2012). This

group only considered one time point during regeneration (mid-blastemal

development) and as such it would be interesting to identify other important highly

regulated miRNAs at different time points during the regenerative timeline.

5.3 MicroRNAs and cardiac regeneration

The importance of miRNAs in the heart was initially reported via the cardiac-

specific deletion of Dicer, the response of which depleted mature miRNAs

specifically in the heart and resulted in lethality shortly after birth (Chen et al., 2008).

Moreover miRNA profiling studies have provided a range of feedback indicating

miRNAs have regulatory functions in specific cardiac cellular processes such as

cardiac dedifferentiation, cell-cycle control, hypertrophy and remodeling in the adult

heart (Chen et al., 2008; Sayed et al., 2007; Thum et al., 2007). The importance of

several miRNAs during heart regeneration has come to light in mouse and zebrafish

Page 39: Heart Regeneration - DiVA - Simple search

39

models. A recent study identified a novel role for the miR-15 family following

ischemic injuries. Specifically, by over-expressing miR-195, a miR-15 family

member, regenerating neo-natal mouse hearts revealed fewer proliferating

cardiomyocytes (Porrello et al., 2013). The authors speculate these proliferating

cardiomyocytes contribute to the regenerate. Additionally these mice showed severe

fibrosis and had irregular cardiac function, indicating a similar response to adult-like

remodeling in non-regenerating scenarios. Intriguingly, the authors also showed in

adult mouse models undergoing the same procedure, inhibition of miR-15 family

members of miRNAs increased cardiomyocyte proliferation and improved cardiac

function. These results speculate that members of the miR-15 family may be one

potential post-natal regulator responsible for the natural arrest of cardiomyocyte

proliferation shortly after birth.

Rodent hearts that were diminished of the cardiomyocyte-specific miR-133

using antagomirs, resulted in a hypertrophic phenotype (Care et al., 2007). The role of

miR-133 was also recently addressed in zebrafish hearts where it’s expression is

naturally decreased upon injury (Yin et al., 2012). Moreover this group showed that

depleting miR-133 in the damaged heart enhanced the regenerative response, while

hearts that were over-expressed with miR-133 alternatively impeded muscle

regeneration, resulting in increased scar tissue formation. This group concluded miR-

133 acts in part to restrict cardiomyocyte proliferation.

The expression of muscle contractile proteins is tightly regulated and

interesting hallmark discoveries include the finding of a family of miRNAs that

regulate muscle-specific myosin genes. Such microRNAs have been coined

“MyomiRs” and consists of miR-1, miR-133, miR-208a, miR-208b and miR-499. It

has been shown that these miRNAs maintain the balance between βMHC and αMHC

expression, which alter contractility and function of cardiomyocytes in the heart

during injury and repair (Gupta, 2007). Of further interest are miRNAs that have been

shown to target several key cardiac transcription factors. One such example is the

effect miR-1 has on downregulating cardiac growth by targeting HAND2 during

development (Zhao et al., 2005). In addition miR-1 is the most abundant miRNA in

the mammalian heart and although all of the cardiac related functions of this miRNA

are not known, the upregulation of miR-1 has been shown to lead to improved cardiac

growth (Chen et al., 2008; Ivey et al., 2008).

Additionally it is interesting to note the heightened expression of certain

miRNAs in patients with cardiovascular disease. For example, miR-21 has been

Page 40: Heart Regeneration - DiVA - Simple search

40

shown to be the most up-regulated miRNA in cardiac disease, although its role is

largely unknown (Small et al., 2010). Some reports delineate a hypertrophic

phenotype with the induction of miR-21 (Thum et al., 2008). Other studies contradict

these results by finding an antihypertrophic effect of miR-21 on isolated

cardiomyocytes and cardiomyocyte growth (Sayed et al., 2008; Tatsuguchi et al.,

2007). The discrepancies highlight a note of caution when considering using these

small non-coding RNAs as therapeutic tools to overcome cardiovascular illnesses.

Other studies have also indicated the use of miRNAs to reprogram cardiac

fibroblasts into cardiomyocyte like cells. Using a small set of only four miRNAs, one

study showed the efficient reprogramming potential of mouse cardiac fibroblasts into

CMC-like cells (Jayawardena et al., 2012). Another study revealed the importance of

several miRNAs to initiate cardiac gene expression programs via the reprogramming

of human fibroblasts (Nam et al., 2013). Unexpectedly, this report also highlighted

muscle-specific miRNAs that further improved myocardial conversion without the

need for key cardiac transcription factors. Of further importance were novel findings

in patients and experimental MI models in mice, revealing muscle-specific miRNAs

such as miR-1, miR-133, miR-208, and miR-499 elevated in the plasma (Wang et al.,

2010). The exact mechanisms mediating the release of miRNAs into the blood and the

biological benefits or functions of circulating miRNAs has not yet been clearly

defined. A recent study reported findings of a mircrine response, in which miR-499

controls gene expression and is responsible for CSC growth and differentiation by

traversing gap-junctions (Hosoda et al., 2011). The discovery of circulating miRNAs

in the blood not only increases the complexity of miRNA involvement in an inter-

cellular manner, but also opens possibilities for the use of miRNAs as biomarkers

against cardiovascular diseases.

Taken together the reports and findings discussed here cover a multitude of

different species and tissue sources all relying on miRNAs for development,

functional repair and regeneration. The avenue for implementing miRNAs as a novel

therapeutic approach to one-day combat congenital heart disease is looking optimistic.

Meanwhile, using lower vertebrate species such as the newt will allow us to continue

to investigate the functional roles of miRNAs in cardiac biology.

In order to consider mechanisms of global regeneration, we considered how

mRNA transcript expression is controlled and regulated. One of many mechanisms

that regulate mRNA is via the post-transcriptional editing of mRNA. Below I will

Page 41: Heart Regeneration - DiVA - Simple search

41

introduce the enzymes that are responsible for A-to-I RNA editing and explain the

significance of RNA editing in relation to regeneration.

Page 42: Heart Regeneration - DiVA - Simple search

42

6. RNA EDITING

6.1 A-to-I RNA editing

RNA editing is a widespread co- and post-transcriptional molecular

phenomenon where changes in RNA sequences are introduced so that it differs from

the DNA template. This is a strategy for metazoans to increase the protein repertoire

without extending the number of genes. The most prevalent type of RNA editing is A-

to-I RNA editing. Here, adenosine deaminases acting on RNA (ADAR) are the

enzymes that mediate hydrolytic deamination, which are responsible for converting

adenosines to inosines in double stranded RNA substrates (Bass, 2002, Wagner et al.,

1989).

6.1.1 ADAR enzymes

The ADAR family is well conserved in vertebrates and there are three family

members (ADAR1-3) (Chen et al., 2000; Kim et al., 1994; O'Connell et al., 1995).

ADAR1 and ADAR2 have a wide tissue distribution, (Bass, 2002; Valente and

Nishikura, 2007), whereas expression of the ADAR3 enzyme is refined to the brain.

To date no enzymatic activity has been attributed to the ADAR3 enzyme (Chen et al.,

2000).

ADARs appear in great diversity across the metazoan spectrum (Figure 5). A

single ADAR2 homologue (dADAR) exists in Drosophila melanogaster, two ADAR

genes exist in Caenorhabditis elegans, which have shown to be critically important

for normal development and only one ADAR1 homologue expressed in Xenopus

laevis (Hough and Bass, 1994; Palladino et al., 2000; Tonkin et al., 2002). Moreover

ADAR homologues have been found in some invertebrates including squids and sea

urchins but interestingly, ADARs are absent in protozoa, yeast and plants (Jin et al.,

2009; Palavicini et al., 2009).

Members of the ADAR family share a common catalytic deaminase domain in

their C-terminus and two or three double-stranded RNA binding domains (dsRBD) at

the N-terminus (Saunders and Barber, 2003). The dsRBDs are required for the

successful binding to the double stranded RNA targets (Valente and Nishikura, 2007).

Page 43: Heart Regeneration - DiVA - Simple search

43

Figure 5. The ADAR family members compared between mammals and amphibians. Three mammalian ADAR proteins are presented with two isoforms of ADAR1, (p150 and p110), ADAR2 and ADAR3. Xenopus laevis express a common ADAR1 homologue (Xl ADAR1). The red spotted newt has an ADAR1 homologue (Nv ADAR1) as well as an ADAR2 homologue (Nv ADAR2). Common features to all ADAR enzymes are a catalytic deaminase domain (in blue), two or three dsRBDs (in green) and nuclear localization signals (in red). Image adapted from (Wahlstedt and Ohman, 2011). See paper III and section 7.4.

Two protein isoforms of the ADAR1 enzyme have been discovered in

mammals and are referred to as p150 and p110 according to their molecular weight.

The ADAR1(p150) has a promoter that is interferon inducible. The shorter isoform,

ADAR1(p110) is constitutively expressed from two different promoters (George and

Samuel, 1999; Kawakubo and Samuel, 2000). One current axiom is that the longer

isoform is involved in an immune host defense mechanism against RNA viruses, as it

is upregulated upon viral infection (Toth et al., 2009).

All the members of the ADAR enzymes contain a nuclear localization signal

(NLS) and are catalytically active in the nucleus. However, ADAR1(p150) also

contains a nuclear export signal (NES) in the N-terminus and as such has been shown

to shuttle between the nucleus and cytoplasm (Poulsen et al., 2001). Interestingly, it

was recently reported that the short ADAR1(p110) lacking this signal also shuttles

between the nucleus and cytoplasm with the help of exportin (Fritz et al., 2009).

Intriguingly this is the same chaperone protein used to export pre-miRNAs from the

nucleus to the cytoplasm (see section 5.1 microRNA biogenesis).

Page 44: Heart Regeneration - DiVA - Simple search

44

Several splice variants have been reported for both ADAR1 and ADAR2

although it has not yet been fully elucidated if these splice variants are translated into

enzymatically active proteins (Agranat et al., 2010; Lykke-Andersen et al., 2007). The

pre-mRNA of ADAR2 has been shown to be a target of the ADAR2 enzyme and

therefore ADAR2 can regulate its own protein expression by an auto-regulatory

feedback loop mechanism (Rueter et al., 1999). Here, editing creates an alternative

splice site, generating a frame shift, resulting in a non-functional truncated protein.

6.2 Editing substrates

6.2.1 Editing in protein-coding regions

The deamination of adenosines into inosines results in inosines being

recognized as guanosines by the splicing and translational machinery. The

replacement of an adenosine with a guanosine can create codon changes. A-to-I RNA

editing has two different modes of action, site-selective editing and hyper editing,

which occur in different regions of RNA. Site selective editing occurs more

frequently in coding regions, where one or a few adenosines are deaminated into

inosines. Hyper-editing occurs in a more promiscuous manner often in non-coding

sequences where many adenosines are targeted for deamination (Wahlstedt and

Ohman, 2011). The majority of site-selective A-to-I RNA editing events is limited in

number and most have been found in neurotransmitter receptors and ion channels. A

few examples include the glutamate receptor (Higuchi et al., 1993), the alpha subunit

of the GABA receptor (Ohlson et al., 2007), the serotonin receptor (Burns et al.,

1997) and the voltage-gated potassium channel Kv1.1 (Hoopengardner et al., 2003).

The effects of editing on these transcripts change amino acid residues, which results

in altered ion channel permeability as well as having an affect on ligand-binding

affinity (Burns et al., 1997; Higuchi et al., 1993; Seeburg et al., 1998; Wang et al.,

2000).

It is important to mention that mechanisms underlying ADAR substrate

determination are not fully known. Although some reports have indicated certain

neighboring nucleotides are favored over others, there is no strict consensus sequence

in which all ADARs act on, nor are there required co-factors for the deaminase

activity (Lehmann and Bass, 2000). It would also appear ADAR activity is influenced

by certain characteristics of dsRNAs including their size and mismatched base pairs

that create bulges (Enstero et al., 2009; Ohman et al., 2000).

Page 45: Heart Regeneration - DiVA - Simple search

45

6.2.2 Editing of non-coding RNA

As mentioned in the previous section hyper-editing occurs in non-coding

RNAs, often within Alu repeats found in 5’ and 3’ UTRs, but also in other regions of

UTRs and introns. Of further interest is accumulating evidence for A-to-I RNA

editing involved in the RNAi pathway. An interesting notion has recently come to

light that A-to-I RNA editing mechanisms interact with the RNAi pathway via the

competition of dsRNA substrates. As discussed previously, miRNA-mediated gene

silencing can control many different processes in development and regeneration

including differentiation and proliferation. To intensify the complexity of miRNAs

and editing further, pri-miRNAs are generally quite long double stranded RNA

structures with miss-matches and bulges (see section 5.1 miRNA biogenesis and

Figure 4), making them ideal targets of ADARs and editing.

The editing of certain pri-miRNA structures can inhibit or enhance Drosha

activity. For example, the editing of pri-miR-142 effects Drosha processing. This

inability for Drosha to process accordingly interrupts the overall mature micro-

processing of miR-142, which instead becomes rapidly degraded by a specific

ribonuclease (Scadden, 2005). Interestingly, reports have begun to show that pri-

miRNAs that escape editing in the nucleus may meet ADARs in the cytoplasm.

Although ADARs are enzymatically active in the nucleus, they can still bind to pre-

miRNAs in the cytoplasm thus preventing continuation of the miRNA maturation

process. Additionally, editing at the pri-miRNA level may later prevent Dicer

processing, as is the case with pre-miR-151 (Kawahara et al., 2007a). It is

hypothesized that ADAR activity may reduce the production of small RNAs if Dicer

prefers only Watson-Crick base pairs, not I:U wobbles created from editing. The

significance of A-to-I RNA editing in miRNAs is also emphasized by the fact that

editing within the “seed region” may lead to the mature miRNAs silencing different

targets as is the case for miR-376 (Kawahara et al., 2007b). In turn, A-to-I RNA

editing is yet another mechanism the cellular machinery utilizes to control abundance

of miRNA and thus their functions. However, the relationship is reciprocal, as recent

studies have shown mature miRNAs can target ADARs. For example, the expression

of ADAR1 has been shown to be down-regulated by miR-1 (Lim et al., 2005). It is of

further interest to note that miR-1 plays a critical role in the development of the heart,

and that during mammalian development the highest expression of ADAR1 is seen in

Page 46: Heart Regeneration - DiVA - Simple search

46

the heart (Srivastava, 2006; Wang et al., 2004). This implies a symbiotic relationship

between RNA editing, miRNAs and the regulation of heart development.

6.3 Functional aspects of RNA editing

The inactivation of ADAR1 leads to an embryonic lethal phenotype due to

widespread apoptosis and therefore is required for life in mammals (Wang et al.,

2004; Wang et al., 2000). Mice with ADAR2 mutations die shortly after birth and

experience repetitive episodes of epileptic seizures due to under-editing of the

glutamate receptor, a major target of ADAR2 (Higuchi et al., 2000).

Human pathophysiology affiliated with RNA-editing deficiencies is most

closely related to disorders of the central nervous system. Editing deficiency in the

Q/R site of the glutamate receptor can disrupt blood flow to the brain and editing

disorders of the serotonin receptor have been linked to neuropsychiatric disorders

such as depression and schizophrenia (Peng et al., 2006; Schmauss, 2005). In

addition, the effect of reduced RNA editing has been linked to human gliomas and

cancer (Maas et al., 2001; Paz et al., 2007). How this reduced activity of editing

manifests malignant cell growth is not yet known.

Of further interest was a recent report that linked RNA editing to congenital

heart disease (Borik et al., 2011). The gene encoding thyroid hormone related protein

1 is implicated in cardiac hypertrophy, angiogenesis and is targeted by the MyomiR

miR-208. The A-to-I editing of this substrate was significantly higher in children with

cyanotic congenital heart disease. These authors concluded that post-transcriptional

regulations such as RNA editing may play a role in regulating cellular and metabolic

pathways that influence hypoxic conditions in the heart.

The physiological importance of editing transcripts has been documented in

mutant flies with deletions of the dADAR gene, which are known to edit glutamate

channels. Here, brain-related physiological complications arise such as uncoordinated

locomotion and signs of neurodegeneration (Palladino et al., 2000).

It is interesting to note that several seminal studies revealed a critical role for

ADAR1 in maintenance of hematopoietic stem cells. In one such study, the loss of

ADAR1 in mouse hematopoietic cells led to a global upregulation of type-I and type-

II interferon-inducible transcripts resulting in cellular apoptosis (Hartner et al., 2009).

Moreover, it has also been shown that ADAR1 deficient hematopoietic progenitor

cells suffer an inability to form differentiated colonies (XuFeng et al., 2009). The

Page 47: Heart Regeneration - DiVA - Simple search

47

inability to form differentiated colonies in vitro is most likely a result of increased

apoptosis in these cells.

Page 48: Heart Regeneration - DiVA - Simple search

48

7. PRESENT INVESTIGATIONS

7.1 Aims

The aims of this thesis were to elucidate if newt hearts could regenerate

myocardial tissue following a novel resection injury and to decipher the molecular

and cellular involvement in response to tissue/organ injury and regeneration.

The specific aims of the different studies are:

Paper I: Create a standardized cardiac resection injury model. Use this new model

system in order to determine the structural and functional responses the newt heart has

to resection injury. To investigate what cell types and transcriptional regulators may

be involved in regenerating the heart.

Paper II: Delineate a possible role for miRNAs in regenerating adult cardiac tissue by

applying a miRNA micro-array screen. Then create a small miRNA cDNA library to

obtain newt specific sequences. Perform quantitative and qualitative methods to

characterize a role for miRNAs in newt cardiac regeneration.

Paper III: Examine if A-to-I RNA editing is active in a variety of different regeneration

tissues in the red-spotted new. Investigate how this post-transcriptional modification

could be responsible for mediating signaling components necessary to drive the

creation of cellular plasticity and regeneration.

Page 49: Heart Regeneration - DiVA - Simple search

49

7.2 Paper I

7.2.1 Results and discussion

It had initially been reported that the heart of the adult red-spotted newt had a

limited regenerative response to cardiac injury. This was based on apical resections of

the ventricle, which left a partial repair response that included some cardiomyocyte

turnover while the presence of fibrotic lesions and scarring triumphed over tissue

regeneration (Oberpriller and Oberpriller, 1974). More recent reports provided

evidence that newt hearts could restore damage done to the myocardium following

crush injury (Laube et al., 2006). Although this group showed the remarkable

plasticity potential of newt cardiomyocytes to control their identity, the further

investigation of cellular and molecular mechanisms involved in newt cardiac

regeneration could go someway in amending the regeneration deficit found in

mammals. In Paper I, we aimed to produce a standardized and reproducible resection

injury model. As opposed to apical resections that incur variations in damage size,

this new procedure would allow us to obtain precise tissue removal. In addition, by

resecting a lateral portion of the ventricle we minimize the chances of breaching the

ventricular lumen and thus reduce morbidity rates. We could then assess on a more

standardized level, the extent of which the newt heart responds to injury over longer

regeneration times.

First, we employed histological staining methods that were able to show the

newt myocardium undergoes full morphological restoration in as little as 2 months

following injury. The morphological recovery could be characterized as follows; first

around 1 week following resection we observed a thick thrombin clot covering the

amputation plane. Over the course of the next few weeks, fibrin and collagen deposits

receded, as the epicardium appears to restore structure to the myocardium. Infiltrating

cardiomyocytes begin to invaginate the damaged areas and by 45dpi minimal signs of

resection injury are noticed.

In an effort to trace cardiac functional output, we next conducted

echocardiogram studies. Interestingly, the inner cardiac wall measurements provided

fractional shortening values similar to humans, which revealed full functional

recovery in newt hearts in as little as 3 weeks following resection injury. At 60dpi, a

time when we observe morphological recovery, fractional shortening values are

similar to that seen in uninjured hearts.

Page 50: Heart Regeneration - DiVA - Simple search

50

We conducted BrdU pulse chase experiments to assess the cell cycle re-entry

of cell populations in the cardiac milieu. Together with immunohistochemical

approaches, we found that the newt hearts response to injury was a result of increased

and widespread cellular proliferation. We observed a gradual increase in proliferation

in both cardiomyocytes and non-cardiomyocytes throughout the heart. A peak of

cellular proliferation was noted between 21-45dpi, thereafter the numbers of BrdU+ve

cells began to subside.

Next we decided to conduct a molecular signature of several well-known

cardiac transcription factors known to play an important role in cardiogenesis, as

molecular events during regeneration often recapitulate those during development.

These results showed similar mRNA expression patterns for all cardiac transcription

factors studied. In general, immediately following resection, expression levels

decreased but were later heightened at times synonymous with increased cellular

proliferation. Worthy of note was the fact that a significant drop in expression was

also observed at 45dpi, possibly indicating that the cells needed for cardiac

replacement have been reprogrammed and increased cardiac transcriptional activity is

no longer required.

Lastly, we were able to examine separate populations of proliferating

GATA4+ve and Islet1+ve expressing cells localized in and around the regeneration

zones. We identified mature cardiomyocytes that expressed these markers. We also

found Islet1 and GATA4 expression in proliferating non-myocytes. That these cells

were also functionally active provides more evidence that there could be a population

of resident cardiac progenitors becoming activated in response to the resection injury.

Page 51: Heart Regeneration - DiVA - Simple search

51

7.3 Paper II

7.3.1 Results and discussion

As described in section 5 of this thesis, miRNAs have previously been shown

to be involved in many aspects of cardiac biology and regeneration. In Paper II we

wanted to assess what kind of miRNAs may be involved in the regulation of the

different phases of newt cardiac regeneration.

Information regarding miRNAs in the newt was quite limited at the time of

this study. As such we decided to conduct a miRNA microarray screen to gauge

differentially expressed miRNAs involved in cardiac regeneration. As newt miRNA

sequences were not available, detection probes that were complementary to Xenopus,

zebrafish and human miRNAs were employed in order to identify conserved

miRNAs. We looked at miRNA expression levels at two different time points, 7dpi

which corresponds to the early wounding response and 21dpi, which is a time-point

when hyperplasia is at its peak.

The microarray revealed 37 miRNA candidates that were significantly up- or-

down-regulated between 7dpi and 21dpi. We next constructed and sequenced small

RNA libraries in order to secure newt specific sequencing information for

downstream experiments. Sequenced data that was aligned to miRBase revealed 22

sequences corresponding to the previously identified miRNAs from the microarray

screen. From here four highly conserved candidates were selected for a thorough

investigation regarding their expression patterns over the duration of the regeneration

timeline.

One candidate, miR-128, showed the most heightened expression pattern at a

time when cellular proliferation is known to be at its peak. Using in-situ hybridization

techniques we were able to localize the expression pattern of miR-128, which was

primarily restricted to the cardiomyocytes near the epicardial border in uninjured

animals. Of great interest was the specific localization of miR-128 directly in and

adjacent to the regeneration zone several weeks following injury and during

regeneration. Furthermore during regeneration miR-128 expression became localized

in non-myocytes in the regenerating areas.

To examine the effect miR-128 may have in vivo, we ablated miR-128

expression with specific antagomirs and conducted BrdU pulse chase experiments to

observe the effect on proliferating CMCs and non-CMCs during regeneration. Using

immunohistochemistry we were able to observe a significant positive hyperplastic

Page 52: Heart Regeneration - DiVA - Simple search

52

response in non-CMCs in animals that received miR-128 antagomirs. These results

suggest that the natural increase in miR-128 levels during regeneration could serve to

halt over-excessive cellular proliferation.

Interestingly, through the use of several bioinformatics programs we observed

that the 3’UTR of Islet1 contains a novel binding site for miR-128. An earlier study

by our group has shown the existence of Islet1+ve proliferating cells during newt heart

regeneration (PAPER I). To examine if Islet1 expression is altered by miR-128, we

cloned the newt Islet1 3’ UTR into a luciferase vector and conducted 3’UTR miRNA

luciferase assays. We confirmed that miR-128 suppresses Islet1 expression and that

Islet1 transcript and protein levels are decreased at a time when miR-128 is normally

elevated. Furthermore, Islet1 transcript and protein levels were dramatically increased

in animals treated with miR-128 antagomirs. Morphological assessment of the

animals treated with miR-128 antagomirs showed a delay in extracellular matrix

removal and myocardial regeneration following resection injury to the heart.

The role of miR-128 during newt cardiac regeneration could therefore be two-

fold. First miR-128 may act as a global regulator to avoid neoplasia at times of

extensive cellular proliferation in regenerating tissues. Alternatively this may be a

mechanism that allows dedifferentiated cells or proliferating cells to be fine-tuned

into their committed lineages. These results show that differentially expressed

miRNAs may coordinate several factors governing the response of the newt heart to

regenerate.

Page 53: Heart Regeneration - DiVA - Simple search

53

7.4 Paper III

7.4.1 Results and discussion

It has been speculated that the ability for the newt to regenerate many complex

tissue-types may be linked to cellular plasticity in differentiated cells or progenitor

cell types. In Paper III we demonstrated that an important post-transcriptional gene

regulator known as A-to-I RNA editing shows interesting connections with a variety

of regeneration scenarios in the newt.

In order to determine if A-to-I RNA editing is an active post-transcriptional

phenomenon in adult newts, we cloned GABRA3 mRNA, a known target for A-to-I

RNA editing. We found editing of the GABRA3 subunit transcript increased during

neuronal regeneration, as it does during neuronal development in mammals. We next

identified two functionally active RNA editing enzymes in the newt, ADAR1 and

ADAR2. Of further interest was the newt ADAR2, which also edits its own mRNA

transcript as it does in mammals. By targeting its own transcript, ADAR2 regulates its

own level of expression providing a mechanism to regulate editing of ADAR2

substrates, which could alter signaling components necessary to drive cellular

plasticity in regenerating tissue.

Immunohistochemistry analysis revealed an interesting nuclear to cytoplasmic

shift of ADAR1 expression during regeneration in all tissues studied. Of further

interest was the co-localization of ADAR1 within a stem cell population of Pax7+ve

satellite cells in the limb. These results indicate a possible involvement between RNA

editing and the behavior of proliferating cell types needed to regenerate the limb. The

editing efficiency between such satellite cells could be one mechanism used by the

newt to regulate plasticity of proliferating cells or to determine their lineage fate.

To investigate the importance of ADARs in newt tissue regeneration, we

cloned newt ADAR1 and ADAR2 and examined the transcriptional and translational

expression of these enzymes during regeneration. The gene expression profiles of

ADAR1 showed an increase in all tissues investigated, with heightened expression

appearing at times concomitant with increased cellular proliferation in the models

studied. This indicates that RNA editing of particular mRNA transcripts are most

likely targeted at intricate times during regeneration. The gene expression profile of

the functionally active unedited ADAR2 showed variation between all tissues studied.

One explanation for the variation of expression between tissues could be due to the

editing efficiency of tissue-specific substrates of ADAR2. Intriguingly, the protein

Page 54: Heart Regeneration - DiVA - Simple search

54

levels of each enzyme in all tissues studied remained relatively unaltered during

regeneration. This would indicate that protein levels may not have a direct correlation

with the functional quantity of RNA editing, but rather the sub-cellular re-localization

of the enzymes may be a better way to control editing efficiency.

Page 55: Heart Regeneration - DiVA - Simple search

55

7.5 Future Perspectives

The culmination of the work driven during this thesis can be summarized in

two different aspects. First we created a novel heart resection model in which we

could decipher how heart regeneration is regulated in the red spotted newt. Secondly

we sought to explore global regulators of regeneration in these animals. Although we

developed a functional model to study heart regeneration, we have not ascertained

what controls this response or which cell types are needed to harness regeneration.

However, one key finding is the identification of two disparate populations of

GATA4+ve and Islet1+ve proliferating cells in regenerating newt hearts. It would be

interesting to generate transgenic cell lines in order to delineate where the cells are

coming from and how they contribute to the regenerate. To date, there are no

transgenic red-spotted newts. Considering the animals take at least 2 years to reach

sexually maturity (adulthood), the process of creating such animals is not optimal and

other methods need to be adopted. Possible methods could include the induction of a

reporter line with a cardiac enhancer. One could try to transiently or stably integrate

marker genes via viral vectors to label and determine which cells are pertinent for

regeneration. Currently, such techniques have not been optimized for use in the red

spotted newt. Apart from addressing cell types in regeneration newt hearts, we also

found the up-regulation of key cardiac transcription factors during different intervals

along the regeneration time-line. As such, it would be interesting to conduct

functional assessment assays to uncover the importance of these transcripts at

different times during regeneration. Possible methods include silencing the expression

of such factors through the use of small interfering RNAs (siRNA) or morpholinos.

The novel model we constructed removes standardized portions of cardiac

tissue from the lateral portion of newt hearts. This model may one day answer pivotal

questions regarding how new muscle, through the formation of new cardiomyocytes,

can be restored in patients with heart disease. As discussed earlier in this thesis,

cryoinjury models are also becoming more popular in zebrafish heart regeneration

studies. To date there are no reports of cryoinjury models on newt hearts. As different

injury models address different questions, it would be of great interest to see how

newt hearts respond to cryoinjury. These model systems could unlock critical

information in how to reduce scarring, a major concern in post MI patients, as scar

tissue formation diminishes heart function. Together, it is the information gained and

compared across these model systems that would go some way to understanding why

Page 56: Heart Regeneration - DiVA - Simple search

56

mammals have a functional deficit when it comes to regenerating damaged

myocardial tissue.

One major caveat regarding newt regeneration model systems is the lack of

genomic information. The large genome size and the lack of a closely related

reference genome have halted any attempts to assemble the newt genome. Several

groups have now invested energy in de novo sequencing of salamander

transcriptomes, and uncovered novel newt protein families (Abdullayev et al., 2013;

Looso et al., 2013). Recently EST database libraries derived from Notophthalmus

viridescens and Ambystoma mexicanum are beginning to provide large amounts of

sequence information, made publicly available to researchers (Bruckskotten et al.,

2012; Smith et al., 2005). Establishing information such as the newt transcriptome or

proteome may provide molecular insights into whether these animals utilize an

explicit set of genes for tissue regeneration.

In order to gauge a further molecular perspective in heart regeneration we

employed microarray and deep-sequencing strategies, which uncovered a large group

of known and conserved mature miRNAs within the newt heart. Several of these

miRNAs showed different expression patterns along the extended course of the

regeneration time-line. This could highlight the different functional roles miRNAs

play at specific time-points during newt cardiac regeneration. Of further interest was

the discovery of small RNA sequences that did not match those found in miRBase. By

using a reference genome such as the Xenopus, it could be possible to map these small

RNAs to larger pre-mRNA sequences, which could provide an indication of novel

miRNAs in the newt. Furthermore we only selected one conserved miRNA based on

the particular expression at a certain time point of interest. Additionally, the

microarray we conducted represents only two time points during regeneration, one

responding to the early wounding response (7dpi) and the other during the

hyperplastic peak (21dpi). Therefore this array may have only revealed a small subset

of differentially expressed miRNAs during a small window of newt cardiac

regeneration. A further and more detailed analysis of miRNAs at all intricate points

during heart regeneration would undoubtedly uncover novel candidate miRNAs

related to every aspect of heart repair. To delineate functional roles for such

candidates, other methods could include using antagomirs to ablate the expression of

several significantly regulated miRNAs during regeneration and study the impact they

have on functional regeneration. In coordination with these miRNA knockdowns, one

Page 57: Heart Regeneration - DiVA - Simple search

57

could then apply deep-sequencing methods to identify potential and putative targets in

regenerating tissue.

Of greater interest would be to further explore the relationship of Islet1 and

miR-128. Although in vitro we conclusively observe that miR-128 can target the 3’

UTR of Islet1, we can not confirm that the result of the diminished transcriptional

expression of Islet1 during regeneration in vivo is due solely to miR-128 or if there

exists alternative contributing factors. Finally, it would be of interest to examine the

relationship between Islet1 and miR-128 in mammalian organisms. The extensive role

of miR-128 during newt heart regeneration is not completely known. We speculate

that miR-128 may regulate Islet1 expression in order to control proliferation in

potential precursor cells. An interesting notion would be that miR-128 is partly

responsible for down-regulating Islet1 expression in mammalian cardiomyocytes,

limiting the plasticity of CMCs in mammalian hearts. Therefore, it is speculative that

through the use of antagomirs one could stimulate an endogenous proliferative

potential in patients suffering from congestive heart failure by unlocking the

dormancy of Islet1 expression.

To gauge if regenerating tissues are regulated in a globally related manner, we

investigated the involvement of one post-transcriptional mechanism, A-to-I RNA

editing. Although ADAR enzymes are active in a variety of regenerating tissues, the

functional relevance of RNA editing in regeneration remains unclear. It would be of

interest to impair the expression of ADAR enzymes in different regeneration

scenarios and address if this elicits a functional defect in regeneration. One possible

method would include the administration of morpholinos to knock down ADAR

expression during regeneration. This would indicate which tissue types absolutely

require RNA editing activity in order to functionally regenerate. To identify and study

the role of other substrates involved in RNA editing during regeneration in the newt,

one could ablate ADAR enzyme activity as previously mentioned and perform large-

scale mRNA sequencing to detect altered levels of mRNA editing. The expression of

ADAR1 enzymes within stem cell populations in the regenerating limb is an

intriguing find. It would be of further interest to explore which cell types are

harboring ADAR expression in regenerating hearts and brains. Further

immunohistochemistry approaches could be used in order to determine if ADARs

favour cell specific niches or if they are promiscuously expressed in a broad range of

cell types throughout regenerating tissue.

Page 58: Heart Regeneration - DiVA - Simple search

58

8. CONCLUSIONS In this thesis I have shown that the adult red-spotted newt can fully respond to

partial ventricular resection in a process that is fueled by proliferation. A molecular

signature during regeneration revealed developmental cardiogenic transcription

factors most likely play a key role in regulating similar genes involved in cardiac

development. Additionally, I show that different miRNAs are involved in various

phases of newt cardiac regeneration. The inhibition of one miRNA has an affect on

altering the levels of proliferating cells within the heart, ultimately delaying the

regenerative response. Finally, I show that the red spotted newt may harness a global

post-transcriptional process, namely A-to-I RNA editing, which could work to micro-

manage the re-specification and phenotypic plasticity needed for cells to regenerate

injured tissue.

I hope that anyone who has read this far along into my thesis can appreciate

the red spotted newt as an incredible model organism to study heart regeneration.

Currently the biggest drawbacks found in these organisms are 1) the lack of genomic

information and 2) the fact that creating transgenic lines of newts is not optimal due to

its prolonged transition into a sexually mature adult. Due to advances in high-

throughput whole genome sequencing, I can foresee the newt genome being

sequenced and annotated in the near future.

The advantages of using these animals as model organisms for heart

regeneration are many. Apart from being ideal adult vertebrate models, they also

have a very complex cardiopulmonary circuitry, similar to that seen in mammals. In

addition they are also easy to maintain. The morphology and size of the heart is ideal

to study whole mount imaging. Strategies for future work in newt cardiac

regeneration should include efforts addressing which cells are responsible for

repairing the damaged newt heart and from where do they arise. Understanding the

molecular and cellular aspects driving heart repair in lower vertebrate species such as

the red spotted newt may one day help us abridge the regeneration deficiencies found

in humans suffering from cardiac diseases.

Page 59: Heart Regeneration - DiVA - Simple search

59

9. ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to the following individuals who have

been a big part of my life over the last 4.5 years during my PhD.

First, it is difficult to overstate my gratitutde to my supervisor Jamie Morrison.

Thank you for accepting me as your PhD student. You introduced me to the exciting

field of regenerative biology and drove me to become the best scientist that I could

be. Thanks for all the great ideas and helping me to drive projects to completion. I

have thoroughly enjoyed working alongside you over the years and discussing great

science. I am sorry we did not manage to publish in Poultry Science, but you never

know what the future may hold.

To my Co-supervisor Anders Arner, thank you for all the professional and personal

advice over the years. I will never forget some late night evenings in the lab where we

discussed our shared musical interests and maybe did some scientific research too? I

only regret we did not have the opportunity to work on more projects together.

Current and past members of the Morrison group:

Claudia García González, you came at the right time to help look after me in the

final year of my PhD. I’ll never forget to put sliced potatoes on any injury from a

sunburn to an amputated limb. :P It has really been a pleasure working with you.

Christofer Ott, thanks for being a good friend and always being optimistic. I look

forward to more visits in Austria. Jana Heigwer, thanks for keeping me laughing

even during the hard days. Jessika Tibblin, thanks for always being encouraging and

helping me to keep my head up and moving forward. Siavish Javaheri, I am glad we

have stayed in touch and hope to see you in Göteborg. Barbara Thaler thanks for

your help with the microRNA projects.

To present and previous members of the old “Molbio” gang:

In grad school it’s important to have a good support system to help you survive and

stay sane. They may not have kept me sane, but the first few mentions tried, very

hard. Thank you Johan Waldholm, too many good memories to include here. Thanks

for being a great friend, giving me years of laughter and so much helpful advice in

Page 60: Heart Regeneration - DiVA - Simple search

60

and outside of the lab. I don’t know what will happen to Club Kungshamra when we

leave. But I do know although it may change, we certainly won’t! ; ) Ylva Ekdahl,

you’re my half-American sister I’m pretty sure. The offer is still on the table to swap

passports if you’re up for it. I’m pretty sure we’ll go down in history as the best

dynamic teaching duo ever. You’ve become a great friend. Thanks for looking after

me for all of these years. I’m sure we’ll meet up in NYC ; ) Mikaela Behm “doesn’t

care”. I’m not sure if we ever managed even one conversation together? So as normal

here come some familiar noises: “giggity” “bang” (the guns, naturally) and “cheap

sheeps are not sheap cheeps” – I bet you’re trying really hard to say that one right

now. ; ) Fredrik Lackmann, I wish you spent less time in Scotland and more time in

Stockholm. My advice to you if you ever get stuck with one of your projects is to just

ask WWYD “what would yeast do?” ; )

To Anne Beskow, you have no idea how quiet the hallways and labs have become

since you left. But don’t worry, “it’s exactly what it looks like.” Helene Wahlstedt,

thanks for always breaking my phone, stealing my milk and being one of the funniest

people I have ever met. Chammiran Daniel, thank you for volunteering so many

times to cut my hair. It was more a privilege for you then me, but I did appreciate it. :

P Vikoria Hessle, thanks for warmly welcoming me into the department as a

newcomer. I have thoroughly enjoyed and will never forget my relaxing summer stays

in Älvsjö. Widad Dantoft, thanks for the fun tea and coffee fikas på svenska. Om jag

har lärt mig någon svenska, så har jag nog dig att tacka. : )

To Micke Crona thanks for showing me around in the beginning of my PhD and also

teaching me how to climb. David Nord thanks for making me feel welcomed in that

big lab when I first started. To Sara Jupiter, thanks for welcoming me in the

department and always greeting me with a friendly smile.

I also want to thank some other current or previous members from old Molbio for

technical support and advice or just smiling faces that brought me encouragement

every day, including Ylva Engström (thanks for being an extra mentor and always

having time for me), Marie Öhman (sorry if I spent more time in your lab then my

own – but can you blame me? Also sorry for causing all those distractions during your

office meetings), Lars Wieslander (did I ever apologize for flooding “your

department” in my first week here?), Anne von Euler (thanks for introducing me to

Lidingö ice hockey), Petra Björk, Patrick Young, Andrea Eberle, Simei Yu,

Ulrich Theopold, Nadja Neumann, Solveig Hahne, Zhi Wang, Robert Marcus,

Monica Davis, Shiva Esfahani, Badrul Arefin, Gunnel Björklund, Robert

Page 61: Heart Regeneration - DiVA - Simple search

61

Krautz, Thomas Hauling, Josefin Plathan, Margareta Sahlin, Britt-Marie

Sjöberg, Fredrik Tholander, Neus Visa, and Gilad Silberberg.

I wish to thank some others in the department who helped me with all the practical

stuff. Britta Tumlin, ‘ett stort tack’ for making buffers and speaking Swedish with

me. Christina Jansson, thanks for your help with all the paperwork over the years.

Marina Meyer, thank you for helping me with all the paperwork as well as the

additional “American forms” I always threw your way. I appreciated your patience

and help so much.

Thanks to all the others who have passed through Molbio over the years and

influenced the department.

To more distant-colleagues and friends at the Karolinska Institute:

Matt Kirkham, thanks for making collaborations between our groups so easy and for

many good scientific discussions about regeneration. Valentina Paloschi thanks for

all the fun in and out of the labs, as well as the good scientific discussions in the

CMM lunchroom. Your Italian-English has really improved, even if I still don’t

understand you! Lasse Folkersen, thanks for your friendship, fun hat parties and

great discussions about space and science. I will definitely be visiting more in

Copenhagen.

To some new friends in MBW including Mattias Engelbrecht, Olle Dahlberg, Ali

Mirzaei, Alice Sollazo and Daniel Vare, thanks for helping to keep me fit in weekly

innebandy sessions.

To my “Stockholm family” who gave me a great life outside of work:

The other horsemen of the apocalypse: Oliver Adfeldt-Still, I’m so grateful we met

on a course at the KI the first week I was in Stockholm. You have been like a brother

to me, introducing me to Stockholm and many great friends, I am forever grateful.

Gregor McKechnie, thanks for possibly being the only friend I have with a voice of

reason and for taking good care of me. Remember, one can always train “IT”. Jack

Corkette, thanks for being my own personal court jester over the years. Your laugh is

contagious. Other members of my Stockholm family, you know who you are - thank

you for taking care of me, my time in Stockholm would not have been nearly as

memorable or complete without all of you.

Page 62: Heart Regeneration - DiVA - Simple search

62

To my best friends from home, Chris Gant and Andrew Maynard, thank you for

unrestricted support and a friendship that has spanned three decades.

And finally, to my family. To my grandparents, aunts, uncles, and other relatives,

your interest in my life has kept me going.

My heartfelt gratitude goes out to my hard-working parents, my dad Ray Witman

and my mum Kathy Witman, for giving me emotional support and guidance from a

far-distance. They have sacraficed their lives to giving me and my brothers

unconditional love and care. I can not express in words how grateful nor how lucky I

am to be your son, even if I’m just “the middle one”. :P To my two wonderful

brothers, Matt Witman and Colin Witman, you have no idea what I’ve been doing

over the years but thankyou for your understanding and support anyway. You have

always been there for me.

Det har varit en häpnadsväckande möjlighet att utvecklas som forskare i

Sverige. Det har varit ännu mer av ett privilegium att växa som person i

Stockholm. Jag kommer aldrig att glömma min tid här, både i och utanför

labbet. Jag har erhållit professionella allianser, vetenskapliga

erfarenheter och ännu flera goda vänner och minnen under 4 år än de

flesta kan hoppas på under en halv livstid. Dessa år kommer alltid att

saknas men kommer aldrig att glömmas. - Nevin Witman, 2013

Page 63: Heart Regeneration - DiVA - Simple search

63

10. REFERENCES Abdullayev, I., Kirkham, M., Bjorklund, A. K., Simon, A., and Sandberg, R. (2013).

A reference transcriptome and inferred proteome for the salamander Notophthalmus viridescen. Exp Cell Res 319, 1187-97.

Agata, K., Saito, Y., and Nakajima, E. (2007). Unifying principles of regeneration I: Epimorphosis versus morphallaxis. Dev Growth Differ 49, 73-8.

Agranat, L., Sperling, J., and Sperling, R. (2010). A novel tissue-specific alternatively spliced form of the A-to-I RNA editing enzyme ADAR2. RNA Biology 7, 253-262.

Anderson, H., and French, V. (1985). Cell-division during intercalary regeneration in the cockroach leg. Journal of Embryology and Experimental Morphology 90, 57-78.

Anversa, P., Leri, A., and Kajstura, J. (2006). Cardiac regeneration. J Am Coll Cardiol 47, 1769-76.

Bader, D., and Oberpriller, J. O. (1978). Repair and reorganization of minced cardiac muscle in the adult newt (Notophthalmus viridescen). J Morphol 155, 349-57.

Bakkers, J. (2011). Zebrafish as a model to study cardiac development and human cardiac disease. Cardiovascular Research 91, 279-288.

Balsam, L. B., Wagers, A. J., Christensen, J. L., Kofidis, T., Weissman, I. L., and Robbins, R. C. (2004). Haematopoietic stem cells adopt mature haematopoietic fates in ischaemic myocardium. Nature 428, 668-73.

Barbosa-Sabanero, K., Hoffmann, A., Judge, C., Lightcap, N., Tsonis, P. A., and Del Rio-Tsonis, K. (2012). Lens and retina regeneration: new perspectives from model organisms. Biochemical Journal 447, 321-334.

Barnabe-Heider, F., Goritz, C., Sabelstrom, H., Takebayashi, H., Pfrieger, F. W., Meletis, K., and Frisen, J. (2010). Origin of new glial cells in intact and injured adult spinal cord. Cell Stem Cell 7, 470-82.

Bartel, D. P. (2009). MicroRNAs: target recognition and regulatory functions. Cell 136, 215-233.

Bass, B. L. (2002). RNA editing by adenosine deaminases that act on RNA. Annual Review of Biochemistry 71, 817-846.

Beltrami, A. P., Barlucchi, L., Torella, D., Baker, M., Limana, F., Chimenti, S., Kasahara, H., Rota, M., Musso, E., Urbanek, K., Leri, A., Kajstura, J., Nadal-Ginard, B., and Anversa, P. (2003). Adult cardiac stem cells are multipotent and support myocardial regeneration. Cell 114, 763-76.

Berg, D. A., Kirkham, M., Wang, H., Frisen, J., and Simon, A. (2011). Dopamine controls neurogenesis in the adult salamander midbrain in homeostasis and during regeneration of dopamine neurons. Cell Stem Cell 8, 426-33.

Bettencourt-Dias, M., Mittnacht, S., and Brockes, J. P. (2003). Heterogeneous proliferative potential in regenerative adult newt cardiomyocytes. J Cell Sci 116, 4001-9.

Birnbaum, K. D., and Alvarado, A. S. (2008). Slicing across kingdoms: Regeneration in plants and animals. Cell 132, 697-710.

Bondue, A., Lapouge, G., Paulissen, C., Semeraro, C., Lacovino, M., Kyba, M., and Blanpain, C. (2008). Mesp1 acts as a master regulator of multipotent cardiovascular progenitor specification. Cell Stem Cell 3, 69-84.

Borik, S., Simon, A. J., Nevo-Caspi, Y., Mishali, D., Amariglio, N., Rechavi, G., and Paret, G. (2011). Increased RNA editing in children with cyanotic congenital heart disease. Intensive Care Medicine 37, 1664-1671.

Page 64: Heart Regeneration - DiVA - Simple search

64

Brade, T., Gessert, S., Kuhl, M., and Pandur, P. (2007). The amphibian second heart field: Xenopus islet-1 is required for cardiovascular development. Developmental Biology 311, 297-310.

Brockes, J., and Kumar, A. (2005). Newts. Curr Biol 15, R42-4. Brønnum, H., Andersen, D. C., Schneider, M., Nossent, A. Y., Nielsen, S. B., and

Sheikh, S. P. (2013). Islet-1 is a dual regulator of fibrogenic epithelial-to-mesenchymal transition in epicardial mesothelial cells. Experimental Cell Research 319, 424-435.

Bruckskotten, M., Looso, M., Reinhardt, R., Braun, T., and Borchardt, T. (2012). Newt-omics: a comprehensive repository for omics data from the newt Notophthalmus viridescens. Nucleic Acids Research 40, D895-D900.

Bruneau, B.G. (2008). The developmental genetics of congenital heart disease. Nature 451, 943-948.

Bu, L., Jiang, X., Martin-Puig, S., Caron, L., Zhu, S., Shao, Y., Roberts, D. J., Huang, P. L., Domian, I. J., and Chien, K. R. (2009). Human ISL1 heart progenitors generate diverse multipotent cardiovascular cell lineages. Nature 460, 113-7.

Buckingham, M., Meilhac, S., and Zaffran, S. (2005). Building the mammalian heart from two sources of myocardial cells. Nature Reviews Genetics 6, 826-835.

Burns, C. M., Chu, H., Rueter, S. M., Hutchinson, L. K., Canton, H., SandersBush, E., and Emeson, R. B. (1997). Regulation of serotonin-2C receptor G-protein coupling by RNA editing. Nature 387, 303-308.

Cai, C. L., Liang, X., Shi, Y., Chu, P. H., Pfaff, S. L., Chen, J., and Evans, S. (2003). Isl1 identifies a cardiac progenitor population that proliferates prior to differentiation and contributes a majority of cells to the heart. Dev Cell 5, 877-89.

Cai, X. Z., Hagedorn, C. H., and Cullen, B. R. (2004). Human microRNAs are processed from capped, polyadenylated transcripts that can also function as mRNAs. Rna-a Publication of the Rna Society 10, 1957-1966.

Cano-Martinez, A., Vargas-Gonzalez, A., Guarner-Lans, V., Prado-Zayago, E., Leon-Oleda, M., and Nieto-Lima, B. (2010). Functional and structural regeneration in the axolotl heart (Ambystoma mexicanum) after partial ventricular amputation. Arch Cardiol Mex 80, 79-86.

Care, A., Catalucci, D., Felicetti, F., Bonci, D., Addario, A., Gallo, P., Bang, M. L., Segnalini, P., Gu, Y. S., Dalton, N. D., Elia, L., Latronico, M. V. G., Hoydal, M., Autore, C., Russo, M. A., Dorn, G. W., Ellingsen, O., Ruiz-Lozano, P., Peterson, K. L., Croce, C. M., Peschle, C., and Condorelli, G. (2007). MicroRNA-133 controls cardiac hypertrophy. Nature Medicine 13, 613-618.

Chablais, F., Veit, J., Rainer, G., and Jazwinska, A. (2011). The zebrafish heart regenerates after cryoinjury-induced myocardial infarction. BMC Dev Biol 11, 21.

Chen, C. X., Cho, D. S. C., Wang, Q. D., Lai, F., Carter, K. C., and Nishikura, K. (2000). A third member of the RNA-specific adenosine deaminase gene family, ADAR3, contains both single- and double-stranded RNA binding domains. Rna-a Publication of the Rna Society 6, 755-767.

Chen, J. F., Murchison, E. P., Tang, R., Callis, T. E., Tatsuguchi, M., Deng, Z., Rojas, M., Hammond, S. M., Schneider, M. D., Selzman, C. H., Meissner, G., Patterson, C., Hannon, G. J., and Wang, D. Z. (2008). Targeted deletion of Dicer in the heart leads to dilated cardiomyopathy and heart failure. Proceedings of the National Academy of Sciences of the United States of America 105, 2111-2116.

Page 65: Heart Regeneration - DiVA - Simple search

65

Chera, S., Ghila, L., Dobretz, K., Wenger, Y., Bauer, C., Buzgariu, W., Martinou, J. C., and Galliot, B. (2009). Apoptotic cells provide an unexpected source of Wnt3 signaling to drive Hydra head regeneration. Dev Cell 17, 279-89.

Chi, S. W., Zang, J. B., Mele, A., and Darnell, R. B. (2009). Argonaute HITS-CLIP decodes microRNA-mRNA interaction maps. Nature 460, 479-486.

Clamp, M., Fry, B., Kamal, M., Xie, X. H., Cuff, J., Lin, M. F., Kellis, M., Lindblad-Toh, K., and Lander, E. S. (2007). Distinguishing protein-coding and noncoding genes in the human genome. Proceedings of the National Academy of Sciences of the United States of America 104, 19428-19433.

Collins, C. A., Olsen, I., Zammit, P. S., Heslop, L., Petrie, A., Partridge, T. A., and Morgan, J. E. (2005). Stem cell function, self-renewal, and behavioral heterogeneity of cells from the adult muscle satellite cell niche. Cell 122, 289-301.

Costello, I., Pimeisl, I. M., Drager, S., Bikoff, E. K., Robertson, E. J., and Arnold, S. J. (2011). The T-box transcription factor Eomesodermin acts upstream of Mesp1 to specify cardiac mesoderm during mouse gastrulation. Nature Cell Biology 13, 1084-U108.

Cummings, S. G., and Bode, H. R. (1984). Head regeneration and polarity reversal in Hydra-Attenuata can occur in the absence of DNA-synthesis. Wilhelm Rouxs Archives of Developmental Biology 194, 79-86.

Davis, B. M., Ayers, J. L., Koran, L., Carlson, J., Anderson, M. C., and Simpson, S. B., Jr. (1990). Time course of salamander spinal cord regeneration and recovery of swimming: HRP retrograde pathway tracing and kinematic analysis. Exp Neurol 108, 198-213.

Davis, D. R., Zhang, Y. Q., Smith, R. R., Cheng, K., Terrovitis, J., Malliaras, K., Li, T. S., White, A., Makkar, R., and Marban, E. (2009). Validation of the cardiosphere method to culture cardiac progenitor cells from myocardial tissue. Plos One 4.

Davis, G. K. (2012). Cyclical parthenogenesis and viviparity in Aphids as evolutionary novelties. Journal of Experimental Zoology Part B-Molecular and Developmental Evolution 318B, 448-459.

Davis, R. L., Weintraub, H., and Lassar, A. B. (1987). Expression of a single transfected cDNA converts fibroblasts to myoblasts. Cell 51, 987-1000.

Del Rio-Tsonis, K., Trombley, M. T., McMahon, G., and Tsonis, P. A. (1998). Regulation of lens regeneration by fibroblast growth factor receptor 1. Dev Dyn 213, 140-6.

Dinsmore, CE., ed. (1991). “Lazzaro Spallanzani: Concepts of generation and regeneration”. In A history of regeneration research: Milestones in the evolution of a Science, 67-89. Cambridge University Press, New York, pp. 67-89.

Dinsmore, C. E. (1996). Urodele limb and tail regeneration in early biological thought: An essay on scientific controversy and social change. International Journal of Developmental Biology 40, 621-627.

Dodou, E., Verzi, M. P., Anderson, J. P., Xu, S. M., and Black, B. L. (2004). Mef2c is a direct transcriptional target of ISL1 and GATA factors in the anterior heart field during mouse embryonic development. Development 131, 3931-42.

Eguchi, G., Eguchi, Y., Nakamura, K., Yadav, M. C., Millan, J. L., and Tsonis, P. A. (2011). Regenerative capacity in newts is not altered by repeated regeneration and ageing. Nature Communications 2.

Enstero, M., Daniel, C., Wahlstedt, H., Major, F., and Ohman, M. (2009). Recognition and coupling of A-to-I edited sites are determined by the tertiary structure of the RNA. Nucleic Acids Research 37, 6916-6926.

Page 66: Heart Regeneration - DiVA - Simple search

66

Esposito, M. (2013). Weismann versus Morgan revisited: clashing interpretations on animal regeneration. J Hist Biol 46, 511-41.

Fire, A., Xu, S. Q., Montgomery, M. K., Kostas, S. A., Driver, S. E., and Mello, C. C. (1998). Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 391, 806-811.

Fritz, J., Strehblow, A., Taschner, A., Schopoff, S., Pasierbek, P., and Jantsch, M. F. (2009). RNA-Regulated interaction of transportin-1 and exportin-5 with the double-stranded RNA-binding domain regulates nucleocytoplasmic shuttling of ADAR1. Molecular and Cellular Biology 29, 1487-1497.

Gabella, G., and Uvelius, B. (1990). Urinary bladder of rat: fine structure of normal and hypertrophic musculature. Cell Tissue Res 262, 67-79.

George, C. X., and Samuel, C. E. (1999). Characterization of the 5 '-flanking region of the human RNA-specific adenosine deaminase ADAR1 gene and identification of an interferon-inducible ADAR1 promoter. Gene 229, 203-213.

Gill, D.E. (1978). Effective population size and interdemic migration rates in a meta-population of the red-spotted newt, Notophthalmus viridescens (Rafinesque). Evolution 32, 839-849.

Gillen, C., Korfhage, C., and Muller, H. W. (1997). Gene expression in nerve regeneration. Neuroscientist 3, 112-122.

Gnecchi, M., Zhang, Z. P., Ni, A. G., and Dzau, V. J. (2008). Paracrine mechanisms in adult stem sell signaling and therapy. Circulation Research 103, 1204-1219.

Goettsch, C and Aikawa E. (2013). Role of MicroRNAs in cardiovascular calcification, calcific aortic valve disease, Dr. Elena Aikawa (Ed.), ISBN: 978-953-51-1150-4, InTech, DOI: 10.5772/55326. Available from: http://www.intechopen.com/books/calcific-aortic-valve-disease/role-of-micrornas-in-cardiovascular-calcification

Gonzalez-Rosa, J. M., Martin, V., Peralta, M., Torres, M., and Mercader, N. (2011). Extensive scar formation and regression during heart regeneration after cryoinjury in zebrafish. Development 138, 1663-74.

Gregory, R. I., Chendrimada, T. P., Cooch, N., and Shiekhattar, R. (2005). Human RISC couples microRNA biogenesis and posttranscriptional gene silencing. Cell 123, 631-640.

Guo, H. L., Ingolia, N. T., Weissman, J. S., and Bartel, D. P. (2010). Mammalian microRNAs predominantly act to decrease target mRNA levels. Nature 466, 835-U66.

Gupta, M. P. (2007). Factors controlling cardiac myosin-isoform shift during hypertrophy and heart failure. Journal of Molecular and Cellular Cardiology 43, 388-403.

Guttman, M., and Rinn, J. L. (2012). Modular regulatory principles of large non-coding RNAs. Nature 482, 339-346.

Hamilton, W.J. (1940). The feeding habits of larval newts with reference to availability and predilection of food items. Ecology 21, 351-356.

Hartner, J. C., Walkley, C. R., Lu, J., and Orkin, S. H. (2009). ADAR1 is essential for the maintenance of hematopoiesis and suppression of interferon signaling (vol 10, pg 109, 2009). Nature Immunology 10, 551-551.

Haubner, B. J., Adamowicz-Brice, M., Khadayate, S., Tiefenthaler, V., Metzler, B., Aitman, T., and Penninger, J. M. (2012). Complete cardiac regeneration in a mouse model of myocardial infarction. Aging (Albany NY) 4, 966-77.

He, A. B., Kong, S. W., Ma, Q., and Pu, W. T. (2011). Co-occupancy by multiple cardiac transcription factors identifies transcriptional enhancers active in heart.

Page 67: Heart Regeneration - DiVA - Simple search

67

Proceedings of the National Academy of Sciences of the United States of America 108, 5632-5637.

Hicklin, J., and Wolpert, L. (1973). Positional information and pattern regulation in Hydra - effect of gamma-radiation. Journal of Embryology and Experimental Morphology 30, 741-752.

Higuchi, M., Single, F. N., Kohler, M., Sommer, B., Sprengel, R., and Seeburg, P. H. (1993). RNA editing of Ampa receptor subunit Glur-B - a base-paired intron-exon structure determines position and efficiency. Cell 75, 1361-1370.

Higuchi, M., Stefan, M., Single, F. N., Hartner, J., Rozov, A., Burnashev, N., Feldmeyer, D., Sprengel, R., and Seeburg, P. H. (2000). Point mutation in an AMPA receptor gene rescues lethality in mice deficient in the RNA-editing enzyme ADAR2. Nature 406, 78-81.

Hillman, S.S., Withers, P.C., Drewes, C.D., and Hillyard, S.D. (2009). Ecological and environmental physiology of amphibians Vol 1 (Oxford; New York: Oxford University Press, 2009).

Holman, E. C., Campbell, L. J., Hines, J., and Crews, C. M. (2012). Microarray analysis of microRNA expression during axolotl limb regeneration. Plos One 7.

Hoopengardner, B., Bhalla, T., Staber, C., and Reenan, R. (2003). Nervous system targets of RNA editing identified by comparative genomics. Science 301, 832-836.

Hosoda, T., Zheng, H. Q., Cabral-da-Silva, M., Sanada, F., Ide-Iwata, N., Ogorek, B., Ferreira-Martins, J., Arranto, C., D'Amario, D., del Monte, F., Urbanek, K., D'Alessandro, D. A., Michler, R. E., Anversa, P., Rota, M., Kajstura, J., and Leri, A. (2011). Human cardiac stem cell differentiation is regulated by a mircrine mechanism. Circulation 123, 1287-U100.

Hough, R. F., and Bass, B. L. (1994). Purification of the Xenopus-laevis double-stranded-RNA adenosine-deaminase. Journal of Biological Chemistry 269, 9933-9939.

Huang, C. Y., Gu, H. M., Yu, Q., Manukyan, M. C., Poynter, J. A., and Wang, M. J. (2011). Sca-1+ cardiac stem cells mediate acute cardioprotection via paracrine factor SDF-1 following myocardial ischemia/reperfusion. Plos One 6.

Hudson, J. E., and Porrello, E. R. (2013). The Non-coding road towards cardiac regeneration. J Cardiovasc Transl Res.

Hurlbert, S.H. (1969). The breeding migrations and interhabitat wandering of the vermilliion-spotted newt Notophthalmus viridescens (Rafinesque) Ecological Monographs. 39, 465-488.

Ieda, M., Fu, J. D., Delgado-Olguin, P., Vedantham, V., Hayashi, Y., Bruneau, B. G., and Srivastava, D. (2010). Direct reprogramming of fibroblasts into functional cardiomyocytes by defined factors. Cell 142, 375-86.

Iten, L. E., and Bryant, S. V. (1973). Forelimb regeneration from different levels of amputation in newt, Notophthalmus-viridescens - length, rate, and stages. Wilhelm Roux Archiv Fur Entwicklungsmechanik Der Organismen 173, 263-282.

Ivey, K. N., Muth, A., Amold, J., King, F. W., Yeh, R. F., Fish, J. E., Hsiao, E. C., Schwartz, R. J., Conklin, B. R., Bernstein, H. S., and Srivastava, D. (2008). MicroRNA regulation of cell lineages in mouse and human embryonic stem cells. Cell Stem Cell 2, 219-229.

James, T. N. (2000). Homage to James B. Herrick: a contemporary look at myocardial infarction and at sickle-cell heart disease: the 32nd Annual Herrick Lecture of the Council on Clinical Cardiology of the American Heart Association. Circulation 101, 1874-87.

Page 68: Heart Regeneration - DiVA - Simple search

68

Jayawardena, T. M., Egemnazarov, B., Finch, E. A., Zhang, L. N., Payne, J. A., Pandya, K., Zhang, Z. P., Rosenberg, P., Mirotsou, M., and Dzau, V. J. (2012). MicroRNA-mediated in vitro and in vivo direct reprogramming of cardiac fibroblasts to cardiomyocytes. Circulation Research 110, 1465-+.

Jin, Y. F., Zhang, W. J., and Li, Q. (2009). Origins and Evolution of ADAR-mediated RNA Editing. Iubmb Life 61, 572-578.

Jopling, C., Boue, S., and Izpisua Belmonte, J. C. (2011). Dedifferentiation, transdifferentiation and reprogramming: three routes to regeneration. Nat Rev Mol Cell Biol 12, 79-89.

Jopling, C., Sleep, E., Raya, M., Marti, M., Raya, A., and Izpisua Belmonte, J. C. (2010). Zebrafish heart regeneration occurs by cardiomyocyte dedifferentiation and proliferation. Nature 464, 606-9.

Kawahara, Y., Zinshteyn, B., Chendrimada, T. P., Shiekhattar, R., and Nishikura, K. (2007a). RNA editing of the microRNA-151 precursor blocks cleavage by the Dicer-TRBP complex. EMBO Reports 8, 763-769.

Kawahara, Y., Zinshteyn, B., Sethupathy, P., Iizasa, H., Hatzigeorgiou, A. G., and Nishikura, K. (2007b). Redirection of silencing targets by adenosine-to-inosine editing of miRNAs. Science 315, 1137-1140.

Kawakubo, K., and Samuel, C. E. (2000). Human RNA-specific adenosine deaminase (ADAR1) gene specifies transcripts that initiate from a constitutively active alternative promoter. Gene 258, 165-172.

Kelly, R. G., Brown, N. A., and Buckingham, M. E. (2001). The arterial pole of the mouse heart forms from Fgf10-expressing cells in pharyngeal mesoderm. Developmental Cell 1, 435-440.

Kikuchi, K., Holdway, J. E., Werdich, A. A., Anderson, R. M., Fang, Y., Egnaczyk, G. F., Evans, T., Macrae, C. A., Stainier, D. Y., and Poss, K. D. (2010). Primary contribution to zebrafish heart regeneration by gata4(+) cardiomyocytes. Nature 464, 601-5.

Kim, U., Wang, Y., Sanford, T., Zeng, Y., and Nishikura, K. (1994). Molecular-cloning of cDNA for double-stranded-RNA adenosine-deaminase, a candidate enzyme for nuclear-RNA editing. Proceedings of the National Academy of Sciences of the United States of America 91, 11457-11461.

Kragl, M., Knapp, D., Nacu, E., Khattak, S., Maden, M., Epperlein, H. H., and Tanaka, E. M. (2009). Cells keep a memory of their tissue origin during axolotl limb regeneration. Nature 460, 60-5.

Kumar, A., Godwin, J. W., Gates, P. B., Garza-Garcia, A. A., and Brockes, J. P. (2007). Molecular basis for the nerve dependence of limb regeneration in an adult vertebrate. Science 318, 772-7.

Lafontant, P. J., Burns, A. R., Grivas, J. A., Lesch, M. A., Lala, T. D., Reuter, S. P., Field, L. J., and Frounfelter, T. D. (2011). The giant Danio (D. aequipinnatus) as a model of cardiac remodeling and regeneration. Anat Rec (Hoboken) 295, 234-48.

Landthaler, M., Yalcin, A., and Tuschl, T. (2004). The human DiGeorge syndrome critical region gene 8 and its D-melanogaster homolog are required for miRNA biogenesis. Current Biology 14, 2162-2167.

Laube, F., Heister, M., Scholz, C., Borchardt, T., and Braun, T. (2006). Re-programming of newt cardiomyocytes is induced by tissue regeneration. J Cell Sci 119, 4719-29.

Laugwitz, K. L., Moretti, A., Caron, L., Nakano, A., and Chien, K. R. (2008). Islet1 cardiovascular progenitors: a single source for heart lineages? Development 135, 193-205.

Page 69: Heart Regeneration - DiVA - Simple search

69

Laugwitz, K. L., Moretti, A., Lam, J., Gruber, P., Chen, Y., Woodard, S., Lin, L. Z., Cai, C. L., Lu, M. M., Reth, M., Platoshyn, O., Yuan, J. X., Evans, S., and Chien, K. R. (2005). Postnatal isl1+ cardioblasts enter fully differentiated cardiomyocyte lineages. Nature 433, 647-53.

Lee, R. C., Feinbaum, R. L., and Ambros, V. (1993). The C-elegans heterochronic gene Lin-4 encodes small RNAs with antisense complementarity to Lin-14. Cell 75, 843-854.

Lehmann, K. A., and Bass, B. L. (2000). Double-stranded RNA adenosine deaminases ADAR1 and ADAR2 have overlapping specificities. Biochemistry 39, 12875-12884.

Lepilina, A., Coon, A. N., Kikuchi, K., Holdway, J. E., Roberts, R. W., Burns, C. G., and Poss, K. D. (2006). A dynamic epicardial injury response supports progenitor cell activity during zebrafish heart regeneration. Cell 127, 607-19.

Leri, A., Kajstura, J., and Anversa, P. (2005). Cardiac stem cells and mechanisms of myocardial regeneration. Physiological Reviews 85, 1373-1416.

Leri, A., Kajstura, J., and Anversa, P. (2011). Role of cardiac stem cells in cardiac pathophysiology: a paradigm shift in human myocardial biology. Circulation Research 109, 941-961.

Lewis, B. P., Burge, C. B., and Bartel, D. P. (2005). Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 120, 15-20.

Li, F., Wang, X., Capasso, J. M., and Gerdes, A. M. (1996). Rapid transition of cardiac myocytes from hyperplasia to hypertrophy during postnatal development. J Mol Cell Cardiol 28, 1737-46.

Lim, L. P., Lau, N. C., Garrett-Engele, P., Grimson, A., Schelter, J. M., Castle, J., Bartel, D. P., Linsley, P. S., and Johnson, J. M. (2005). Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature 433, 769-773.

Linhares, V. L., Almeida, N. A., Menezes, D. C., Elliott, D. A., Lai, D., Beyer, E. C., Campos de Carvalho, A. C., and Costa, M. W. (2004). Transcriptional regulation of the murine Connexin40 promoter by cardiac factors Nkx2-5, GATA4 and Tbx5. Cardiovasc Res 64, 402-11.

Liu, J. D., Carmell, M. A., Rivas, F. V., Marsden, C. G., Thomson, J. M., Song, J. J., Hammond, S. M., Joshua-Tor, L., and Hannon, G. J. (2004). Argonaute2 is the catalytic engine of mammalian RNAi. Science 305, 1437-1441.

Looso, M., Preussner, J., Sousounis, K., Bruckskotten, M., Michel, C. S., Lignelli, E., Reinhardt, R., Hoeffner, S., Krueger, M., Tsonis, P. A., Borchardt, T., and Braun, T. (2013). A de novo assembly of the newt transcriptome combined with proteomic validation identifies new protein families expressed during tissue regeneration. Genome Biol 14, R16.

Lykke-Andersen, S., Pinol-Roma, S., and Kjems, J. (2007). Alternative splicing of the ADAR1 transcript in a region that functions either as a 5 '-UTR or an ORF. Rna-a Publication of the Rna Society 13, 1732-1744.

Maas, S., Patt, S., Schrey, M., and Rich, A. (2001). Underediting of glutamate receptor GluR-B mRNA in malignant gliomas. Proceedings of the National Academy of Sciences of the United States of America 98, 14687-14692.

Makkar, R. R., Smith, R. R., Cheng, K., Malliaras, K., Thomson, L. E. J., Berman, D., Czer, L. S. C., Marban, L., Mendizabal, A., Johnston, P. V., Russell, S. D., Schuleri, K. H., Lardo, A. C., Gerstenblith, G., and Marban, E. (2012). Intracoronary cardiosphere-derived cells for heart regeneration after myocardial infarction (CADUCEUS): a prospective, randomised phase 1 trial. Lancet 379, 895-904.

Page 70: Heart Regeneration - DiVA - Simple search

70

Matsuura, K., Nagai, T., Nishigaki, N., Oyama, T., Nishi, J., Wada, H., Sano, M., Toko, H., Akazawa, H., Sato, T., Nakaya, H., Kasanuki, H., and Komuro, I. (2004). Adult cardiac Sca-1-positive cells differentiate into beating cardiomyocytes. J Biol Chem 279, 11384-91.

Mattick, J. S. (2011). The central role of RNA in human development and cognition. Febs Letters 585, 1600-1616.

Matz, D. G., Oberpriller, J. O., and Oberpriller, J. C. (1998). Comparison of mitosis in binucleated and mononucleated newt cardiac myocytes. Anatomical Record 251, 245-255.

Mauro, A. (1961). Satellite cell of skeletal muscle fibers. J Biophys Biochem Cytol 9, 493-5.

McHedlishvili, L., Mazurov, V., Grassme, K. S., Goehler, K., Robl, B., Tazaki, A., Roensch, K., Duemmler, A., and Tanaka, E. M. (2011). Reconstitution of the central and peripheral nervous system during salamander tail regeneration. Proc Natl Acad Sci U S A 109, E2258-66.

Meilhac, S. M., Esner, M., Kelly, R. G., Nicolas, J. F., and Buckingham, M. E. (2004). The clonal origin of myocardial cells in different regions of the embryonic mouse heart. Developmental Cell 6, 685-698.

Meister, G., Landthaler, M., Patkaniowska, A., Dorsett, Y., Teng, G., and Tuschl, T. (2004). Human Argonaute2 mediates RNA cleavage targeted by miRNAs and siRNAs. Molecular Cell 15, 185-197.

Messina, E., De Angelis, L., Frati, G., Morrone, S., Chimenti, S., Fiordaliso, F., Salio, M., Battaglia, M., Latronico, M. V., Coletta, M., Vivarelli, E., Frati, L., Cossu, G., and Giacomello, A. (2004). Isolation and expansion of adult cardiac stem cells from human and murine heart. Circ Res 95, 911-21.

Michalopoulos, G. K. (2007). Liver regeneration. J Cell Physiol 213, 286-300. Min, J. Y., Yang, Y., Sullivan, M. F., Ke, Q., Converso, K. L., Chen, Y., Morgan, J.

P., and Xiao, Y. F. (2003). Long-term improvement of cardiac function in rats after infarction by transplantation of embryonic stem cells. J Thorac Cardiovasc Surg 125, 361-9.

Miyachi, Y. (2011). The unusual circulation of the newt heart after ventricular injury and its implications for regeneration. Anat Res Int 2011, 812373.

Moretti, A., Caron, L., Nakano, A., Lam, J. T., Bernshausen, A., Chen, Y., Qyang, Y., Bu, L., Sasaki, M., Martin-Puig, S., Sun, Y., Evans, S. M., Laugwitz, K. L., and Chien, K. R. (2006). Multipotent embryonic isl1+ progenitor cells lead to cardiac, smooth muscle, and endothelial cell diversification. Cell 127, 1151-65.

Morrison, J. I., Loof, S., He, P. P., and Simon, A. (2006). Salamander limb regeneration involves the activation of a multipotent skeletal muscle satellite cell population. Journal of Cell Biology 172, 433-440.

Mourelatos, Z., Dostie, J., Paushkin, S., Sharma, A., Charroux, B., Abel, L., Rappsilber, J., Mann, M., and Dreyfuss, G. (2002). miRNPs: a novel class of ribonucleoproteins containing numerous microRNAs. Genes & Development 16, 720-728.

Murry, C. E., Soonpaa, M. H., Reinecke, H., Nakajima, H., Nakajima, H. O., Rubart, M., Pasumarthi, K. B., Virag, J. I., Bartelmez, S. H., Poppa, V., Bradford, G., Dowell, J. D., Williams, D. A., and Field, L. J. (2004). Haematopoietic stem cells do not transdifferentiate into cardiac myocytes in myocardial infarcts. Nature 428, 664-8.

Nagaike, M., Hirao, S., Tajima, H., Noji, S., Taniguchi, S., Matsumoto, K., and Nakamura, T. (1991). Renotropic functions of hepatocyte growth factor in renal regeneration after unilateral nephrectomy. J Biol Chem 266, 22781-4.

Page 71: Heart Regeneration - DiVA - Simple search

71

Nakamura, K., Maki, N., Trinh, A., Trask, H. W., Gui, J. A., Tomlinson, C. R., and Tsonis, P. A. (2010). miRNAs in newt lens regeneration: Specific control of proliferation and evidence for miRNA networking. Plos One 5.

Nam, Y. J., Song, K. H., Luo, X., Daniel, E., Lambeth, K., West, K., Hill, J. A., DiMaio, J. M., Baker, L. A., Bassel-Duby, R., and Olson, E. N. (2013). Reprogramming of human fibroblasts toward a cardiac fate. Proceedings of the National Academy of Sciences of the United States of America 110, 5588-5593.

Natarajan, A., Yamagishi, H., Ahmad, F., Li, D., Roberts, R., Matsuoka, R., Hill, S., and Srivastava, D. (2001). Human eHAND, but not dHAND, is down-regulated in cardiomyopathies. J Mol Cell Cardiol 33, 1607-14.

Niu, Z., Iyer, D., Conway, S. J., Martin, J. F., Ivey, K., Srivastava, D., Nordheim, A., and Schwartz, R. J. (2008). Serum response factor orchestrates nascent sarcomerogenesis and silences the biomineralization gene program in the heart. Proceedings of the National Academy of Sciences of the United States of America 105, 17824-17829.

Norenberg, M. D., Smith, J., and Marcillo, A. (2004). The pathology of human spinal cord injury: defining the problems. J Neurotrauma 21, 429-40.

O'Connell, M. A., Krause, S., Higuchi, M., Hsuan, J. J., Totty, N. F., Jenny, A., and Keller, W. (1995). Cloning of cDNAs encoding mammalian double-stranded RNA-specific adenosine-deaminase. Molecular and Cellular Biology 15, 1389-1397.

Oberpriller, J. O., and Oberpriller, J. C. (1974). Response of the adult newt ventricle to injury. J Exp Zool 187, 249-53.

Ohlson, J., Pedersen, J. S., Haussler, D., and Ohman, M. (2007). Editing modifies the GABA(A) receptor subunit alpha 3. Rna-a Publication of the Rna Society 13, 698-703.

Ohman, M., Kallman, A. M., and Bass, B. L. (2000). In vitro analysis of the binding of ADAR2 to the pre-mRNA encoding the GluR-B R/G site. Rna-a Publication of the Rna Society 6, 687-697.

Okamoto, M., Ohsawa, H., Hayashi, T., Owaribe, K., and Tsonis, P. A. (2007). Regeneration of retinotectal projections after optic tectum removal in adult newts. Mol Vis 13, 2112-8.

Olson, E. N. (2006). Gene regulatory networks in the evolution and development of the heart. Science 313, 1922-7.

Orlic, D., Kajstura, J., Chimenti, S., Limana, F., Jakoniuk, I., Quaini, F., Nadal-Ginard, B., Bodine, D. M., Leri, A., and Anversa, P. (2001). Mobilized bone marrow cells repair the infarcted heart, improving function and survival. Proc Natl Acad Sci U S A 98, 10344-9.

Pack, G. T., Islami, A. H., Hubbard, J. C., and Brasfield, R. D. (1962). Regeneration of human liver after major hepatectomy. Surgery 52, 617-23.

Palavicini, J. P., O'Connell, M. A., and Rosenthal, J. J. C. (2009). An extra double-stranded RNA binding domain confers high activity to a squid RNA editing enzyme. RNA-a Publication of the RNA Society 15, 1208-1218.

Palladino, M. J., Keegan, L. P., O'Connell, M. A., and Reenan, R. A. (2000). dADAR, a Drosophila double-stranded RNA-specific adenosine deaminase is highly developmentally regulated and is itself a target for RNA editing. RNA-a Publication of the RNA Society 6, 1004-1018.

Pandur, P., Sirbu, I. O., Kuhl, S. J., Philipp, M., and Kuhl, M. (2013). Islet1-expressing cardiac progenitor cells: a comparison across species. Development Genes and Evolution 223, 117-129.

Page 72: Heart Regeneration - DiVA - Simple search

72

Parish, C. L., Beljajeva, A., Arenas, E., and Simon, A. (2007). Midbrain dopaminergic neurogenesis and behavioural recovery in a salamander lesion-induced regeneration model. Development 134, 2881-7.

Patapoutian, A., Wold, B. J., and Wagner, R. A. (1995). Evidence for developmentally programmed transdifferentiation in mouse esophageal muscle. Science 270, 1818-1821.

Patten, R. D., Aronovitz, M. J., Deras-Mejia, L., Pandian, N. G., Hanak, G. G., Smith, J. J., Mendelsohn, M. E., and Konstam, M. A. (1998). Ventricular remodeling in a mouse model of myocardial infarction. Am J Physiol 274, H1812-20.

Paz, N., Levanon, E. Y., Amariglio, N., Heimberger, A. B., Ram, Z., Constantini, S., Barbash, Z. S., Adamsky, K., Safran, M., Hirschberg, A., Krupsky, M., Ben-Dov, I., Cazacu, S., Mikkelsen, T., Brodie, C., Eisenberg, E., and Rechavi, G. (2007). Altered adenosine-to-inosine RNA, editing in human cancer. Genome Research 17, 1586-1595.

Peng, P. L., Zhong, X. F., Tu, W. H., Soundarapandian, M. M., Molner, P., Zhu, D. Y., Lau, L., Liu, S. H., Liu, F., and Lu, Y. M. (2006). ADAR2-dependent RNA editing of AMPA receptor subunit GluR2 determines vulnerability of neurons in forebrain ischemia. Neuron 49, 719-733.

Peterkin, T., Gibson, A., Loose, M., and Patient, R. (2005). The roles of GATA-4, -5 and -6 in vertebrate heart development. Semin Cell Dev Biol 16, 83-94.

Piatkowski, T., Muhlfeld, C., Borchardt, T., and Braun, T. (2012). Reconstitution of the myocardium in regenerating newt hearts is preceded by transient deposition of extracellular matrix components. Stem Cells Dev.

Porrello, E. R., Mahmoud, A. I., Simpson, E., Hill, J. A., Richardson, J. A., Olson, E. N., and Sadek, H. A. (2011). Transient regenerative potential of the neonatal mouse heart. Science 331, 1078-80.

Porrello, E. R., Mahmoud, A. I., Simpson, E., Johnson, B. A., Grinsfelder, D., Canseco, D., Mammen, P. P., Rothermel, B. A., Olson, E. N., and Sadek, H. A. (2013). Regulation of neonatal and adult mammalian heart regeneration by the miR-15 family. Proceedings of the National Academy of Sciences of the United States of America 110, 187-192.

Poss, K. D., Wilson, L. G., and Keating, M. T. (2002). Heart regeneration in zebrafish. Science 298, 2188-90.

Poulsen, H., Nilsson, J., Damgaard, C. K., Egebjerg, J., and Kjems, J. (2001). CRM1 mediates the export of ADAR1 through a nuclear export signal within the Z-DNA binding domain. Molecular and Cellular Biology 21, 7862-7871.

Prall, O. W., Menon, M. K., Solloway, M. J., Watanabe, Y., Zaffran, S., Bajolle, F., Biben, C., McBride, J. J., Robertson, B. R., Chaulet, H., Stennard, F. A., Wise, N., Schaft, D., Wolstein, O., Furtado, M. B., Shiratori, H., Chien, K. R., Hamada, H., Black, B. L., Saga, Y., Robertson, E. J., Buckingham, M. E., and Harvey, R. P. (2007). An Nkx2-5/Bmp2/Smad1 negative feedback loop controls heart progenitor specification and proliferation. Cell 128, 947-59.

Qian, L., Huang, Y., Spencer, C. I., Foley, A., Vedantham, V., Liu, L., Conway, S. J., Fu, J. D., and Srivastava, D. (2012). In vivo reprogramming of murine cardiac fibroblasts into induced cardiomyocytes. Nature 485, 593-8.

Rageh, M. A., Mendenhall, L., Moussad, E. E., Abbey, S. E., Mescher, A. L., and Tassava, R. A. (2002). Vasculature in pre-blastema and nerve-dependent blastema stages of regenerating forelimbs of the adult newt, Notophthalmus viridescens. J Exp Zool 292, 255-66.

Rand, T. A., Ginalski, K., Grishin, N. V., and Wang, X. D. (2004). Biochemical identification of Argonaute 2 as the sole protein required for RNA-induced

Page 73: Heart Regeneration - DiVA - Simple search

73

silencing complex activity. Proceedings of the National Academy of Sciences of the United States of America 101, 14385-14389.

Reddien, P. W., and Alvarado, A. S. (2004). Fundamentals of planarian regeneration. Annual Review of Cell and Developmental Biology 20, 725-757.

Reule, S., and Gupta, S. (2011). Kidney regeneration and resident stem cells. Organogenesis 7, 135-9.

Ritzman, T. B., Stroik, L. K., Julik, E., Hutchins, E. D., Lasku, E., Denardo, D. F., Wilson-Rawls, J., Rawls, J. A., Kusumi, K., and Fisher, R. E. (2012). The Gross Anatomy of the Original and Regenerated Tail in the Green Anole (Anolis carolinensis). Anatomical Record-Advances in Integrative Anatomy and Evolutionary Biology 295, 1596-1608.

Rivas, F. V., Tolia, N. H., Song, J. J., Aragon, J. P., Liu, J. D., Hannon, G. J., and Joshua-Tor, L. (2005). Purified Argonaute2 and an siRNA form recombinant human RISC. Nature Structural & Molecular Biology 12, 340-349.

Rueter, S. M., Dawson, T. R., and Emeson, R. B. (1999). Regulation of alternative splicing by RNA editing. Nature 399, 75-80.

Salto-Tellez, M., Yung Lim, S., El-Oakley, R. M., Tang, T. P., ZA, A. L., and Lim, S. K. (2004). Myocardial infarction in the C57BL/6J mouse: a quantifiable and highly reproducible experimental model. Cardiovasc Pathol 13, 91-7.

Sanchez Alvarado, A. (2000). Regeneration in the metazoans: why does it happen? Bioessays 22, 578-90.

Saunders, L. R., and Barber, G. N. (2003). The dsRNA binding protein family: critical roles, diverse cellular functions. Faseb Journal 17, 961-983.

Sayed, D., Hong, C., Chen, I. Y., Lypowy, J., and Abdellatif, M. (2007). MicroRNAs play an essential role in the development of cardiac hypertrophy. Circulation Research 100, 416-424.

Sayed, D., Rane, S., Lypowy, J., He, M. Z., Chen, I. Y., Vashistha, H., Yan, L., Malhotra, A., Vatner, D., and Abdellatif, M. (2008). MicroRNA-21 targets Sprouty2 and promotes cellular outgrowths. Molecular Biology of the Cell 19, 3272-3282.

Scadden, A. D. J. (2005). The RISC subunit Tudor-SN binds to hyper-edited double-stranded RNA and promotes its cleavage. Nature Structural & Molecular Biology 12, 489-496.

Scadden, A. D. J., and Smith, C. W. J. (2001). RNAi is antagonized by A -> I hyper-editing. EMBO Reports 2, 1107-1111.

Schmauss, C. (2005). Regulation of serotonin 2C receptor pre-mRNA editing by serotonin. International Review of Neurobiology, Vol 63 63, 83-+.

Schnabel, K., Wu, C. C., Kurth, T., and Weidinger, G. (2011). Regeneration of cryoinjury induced necrotic heart lesions in zebrafish is associated with epicardial activation and cardiomyocyte proliferation. PLoS One 6, e18503.

Schwarz, D. S., Hutvagner, G., Du, T., Xu, Z. S., Aronin, N., and Zamore, P. D. (2003). Asymmetry in the assembly of the RNAi enzyme complex. Cell 115, 199-208.

Seeburg, P. H., Higuchi, M., and Sprengel, R. (1998). RNA editing of brain glutamate receptor channels: mechanism and physiology. Brain Research Reviews 26, 217-229.

Sehm, T., Sachse, C., Frenzel, C., and Echeverri, K. (2009). miR-196 is an essential early-stage regulator of tail regeneration, upstream of key spinal cord patterning events. Mechanisms of Development 126, S291-S291.

Shi, D. L., and Boucaut, J. C. (1995). The chronological development of the urodele amphibian Pleurodeles Waltl (Michah) (vol 39, pg 423, 1995). International Journal of Developmental Biology 39, 1060-1060.

Page 74: Heart Regeneration - DiVA - Simple search

74

Singer, M., Nordlander, R. H., and Egar, M. (1979). Axonal guidance during embryogenesis and regeneration in the spinal cord of the newt: the blueprint hypothesis of neuronal pathway patterning. J Comp Neurol 185, 1-21.

Small, E. M., Frost, R. J. A., and Olson, E. N. (2010). MicroRNAs add a new dimension to cardiovascular disease. Circulation 121, 1022-U66.

Smart, N., Bollini, S., Dube, K. N., Vieira, J. M., Zhou, B., Davidson, S., Yellon, D., Riegler, J., Price, A. N., Lythgoe, M. F., Pu, W. T., and Riley, P. R. (2011). De novo cardiomyocytes from within the activated adult heart after injury. Nature 474, 640-4.

Smith, J. J., Kump, D. K., Walker, J. A., Parichy, D. M., and Voss, S. R. (2005). A comprehensive expressed sequence tag linkage map for tiger salamander and Mexican axolotl: enabling gene mapping and comparative genomics in Ambystoma. Genetics 171, 1161-1171.

Smith, R. R., Barile, L., Cho, H. C., Leppo, M. K., Hare, J. M., Messina, E., Giacomello, A., Abraham, M. R., and Marban, E. (2007). Regenerative potential of cardiosphere-derived cells expanded from percutaneous endomyocardial biopsy specimens. Circulation 115, 896-908.

Srivastava, D. (2006). Making or breaking the heart: From lineage determination to morphogenesis. Cell 126, 1037-1048.

Srivastava, D., Thomas, T., Lin, Q., Kirby, M. L., Brown, D., and Olson, E. N. (1997). Regulation of cardiac mesodermal and neural crest development by the bHLH transcription factor, dHAND. Nat Genet 16, 154-60.

Stainier, D. Y. R., Lee, R. K., and Fishman, M. C. (1993). Cardiovascular development in the zebrafish .1. Myocardial fate map and heart tube formation. Development 119, 31-40.

Stoick-Cooper, C. L., Moon, R. T., and Weidinger, G. (2007). Advances in signaling in vertebrate regeneration as a prelude to regenerative medicine. Genes & Development 21, 1292-1315.

Swijnenburg, R. J., Tanaka, M., Vogel, H., Baker, J., Kofidis, T., Gunawan, F., Lebl, D. R., Caffarelli, A. D., de Bruin, J. L., Fedoseyeva, E. V., and Robbins, R. C. (2005). Embryonic stem cell immunogenicity increases upon differentiation after transplantation into ischemic myocardium. Circulation 112, I166-72.

Takahashi, K., and Yamanaka, S. (2006). Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663-76.

Tatsuguchi, M., Seok, H. Y., Callis, T. E., Thomson, J. M., Chen, J. F., Newman, M., Rojas, M., Hammond, S. M., and Wang, D. Z. (2007). Expression of microRNAs is dynamically regulated during cardiomyocyte hypertrophy. Journal of Molecular and Cellular Cardiology 42, 1137-1141.

Thomson, J. A., Itskovitz-Eldor, J., Shapiro, S. S., Waknitz, M. A., Swiergiel, J. J., Marshall, V. S., and Jones, J. M. (1998). Embryonic stem cell lines derived from human blastocysts. Science 282, 1145-7.

Thorndyke, M. C., Patruno, M., Chen, W. C., and Beesley, P. W. (2001). Stem cells and regeneration in invertebrate Deuterostomes. Brain Stem Cells, 107-120.

Thum, T., Galuppo, P., Kneitz, S., Fiedler, J., Wolf, C., Van Laake, L., Mummery, C. L., Engelhardt, S., Ertl, G., and Bauersachs, J. (2007). MicroRNAs in the human heart: a clue to fetal gene reprogramming in heart failure. European Heart Journal 28, 786-786.

Thum, T., Gross, C., Fiedler, J., Fischer, T., Kissler, S., Bussen, M., Galuppo, P., Just, S., Rottbauer, W., Frantz, S., Castoldi, M., Soutschek, J., Koteliansky, V., Rosenwald, A., Basson, M. A., Licht, J. D., Pena, J. T. R., Rouhanifard, S. H., Muckenthaler, M. U., Tuschl, T., Martin, G. R., Bauersachs, J., and

Page 75: Heart Regeneration - DiVA - Simple search

75

Engelhardt, S. (2008). MicroRNA-21 contributes to myocardial disease by stimulating MAP kinase signalling in fibroblasts. Nature 456, 980-U83.

Tonkin, L. A., Saccomanno, L., Morse, D. P., Brodigan, T., Krause, M., and Bass, B. L. (2002). RNA editing by ADARs is important for normal behavior in Caenorhabditis elegans. Embo Journal 21, 6025-6035.

Toth, A. M., Li, Z. Q., Cattaneo, R., and Samuel, C. E. (2009). RNA-specific adenosine deaminase ADAR1 suppresses measles virus-induced apoptosis and activation of protein kinase PKR. Journal of Biological Chemistry 284, 29350-29356.

Tsonis, P. A., and Del Rio-Tsonis, K. (2004). Lens and retina regeneration: transdifferentiation, stem cells and clinical applications. Exp Eye Res 78, 161-72.

Tsonis, P. A., and Fox, T. P. (2009). Regeneration according to Spallanzani. Developmental Dynamics 238, 2357-2363.

Valente, L., and Nishikura, K. (2007). RNA binding-independent dimerization of adenosine deaminases acting on RNA and dominant negative effects of nonfunctional subunits on dimer functions. Journal of Biological Chemistry 282, 16054-16061.

van den Bos, E. J., Mees, B. M., de Waard, M. C., de Crom, R., and Duncker, D. J. (2005). A novel model of cryoinjury-induced myocardial infarction in the mouse: a comparison with coronary artery ligation. Am J Physiol Heart Circ Physiol 289, H1291-300.

van Laake, L. W., Passier, R., Monshouwer-Kloots, J., Verkleij, A. J., Lips, D. J., Freund, C., den Ouden, K., Ward-van Oostwaard, D., Korving, J., Tertoolen, L. G., van Echteld, C. J., Doevendans, P. A., and Mummery, C. L. (2007). Human embryonic stem cell-derived cardiomyocytes survive and mature in the mouse heart and transiently improve function after myocardial infarction. Stem Cell Res 1, 9-24.

Vasudevan, S., Tong, Y. C., and Steitz, J. A. (2007). Switching from repression to activation: MicroRNAs can up-regulate translation. Science 318, 1931-1934.

Vincent, S. D., and Buckingham, M. E. (2010). How to make a heart: The origin and regulation of cardiac progenitor cells. Organogenesis in Development 90, 1-41.

Wagner, D. E., Wang, I. E., and Reddien, P. W. (2011). Clonogenic neoblasts are pluripotent adult stem cells that underlie planarian regeneration. Science 332, 811-6.

Wagner, R. W., Smith, J. E., Cooperman, B. S., and Nishikura, K. (1989). A Double-stranded-RNA unwinding activity introduces structural alterations by means of adenosine to inosine conversions in mammalian-cells and Xenopus eggs. Proceedings of the National Academy of Sciences of the United States of America 86, 2647-2651.

Wahlstedt, H., and Ohman, M. (2011). Site-selective versus promiscuous A-to-I editing. Wiley Interdisciplinary Reviews-Rna 2, 761-771.

Wang, G. K., Zhu, J. Q., Zhang, J. T., Li, Q., Li, Y., He, J., Qin, Y. W., and Jing, Q. (2010). Circulating microRNA: a novel potential biomarker for early diagnosis of acute myocardial infarction in humans. European Heart Journal 31, 659-666.

Wang, J., Panakova, D., Kikuchi, K., Holdway, J. E., Gemberling, M., Burris, J. S., Singh, S. P., Dickson, A. L., Lin, Y. F., Sabeh, M. K., Werdich, A. A., Yelon, D., Macrae, C. A., and Poss, K. D. (2011). The regenerative capacity of zebrafish reverses cardiac failure caused by genetic cardiomyocyte depletion. Development 138, 3421-30.

Page 76: Heart Regeneration - DiVA - Simple search

76

Wang, Q. D., Miyakoda, M., Yang, W. D., Khillan, J., Stachura, D. L., Weiss, M. J., and Nishikura, K. (2004). Stress-induced apoptosis associated with null mutation of ADAR1 RNA editing deaminase gene. Journal of Biological Chemistry 279, 4952-4961.

Wang, Q. D., O'Brien, P. J., Chen, C. X., Cho, D. S. C., Murray, J. M., and Nishikura, K. (2000). Altered G protein-coupling functions of RNA editing isoform and splicing variant serotonin(2C) receptors. Journal of Neurochemistry 74, 1290-1300.

Wang, Z. G., Wang, D. Z., Pipes, G. C. T., and Olson, E. N. (2003). Myocardin is a master regulator of smooth muscle gene expression. Proceedings of the National Academy of Sciences of the United States of America 100, 7129-7134.

Winter, J., Jung, S., Keller, S., Gregory, R. I., and Diederichs, S. (2009). Many roads to maturity: microRNA biogenesis pathways and their regulation. Nature Cell Biology 11, 228-234.

XuFeng, R., Boyer, M. J., Shen, H. M., Li, Y. X., Yu, H., Gao, Y. D., Yang, Q., Wang, Q. D., and Cheng, T. (2009). ADAR1 is required for hematopoietic progenitor cell survival via RNA editing. Proceedings of the National Academy of Sciences of the United States of America 106, 17763-17768.

Yang, J. S., Phillips, M. D., Betel, D., Mu, P., Ventura, A., Siepel, A. C., Chen, K. C., and Lai, E. C. (2011). Widespread regulatory activity of vertebrate microRNA* species. RNA-a Publication of the RNA Society 17, 312-326.

Yin, V. P., Lepilina, A., Smith, A., and Poss, K. D. (2012). Regulation of zebrafish heart regeneration by miR-133. Developmental Biology 365, 319-327.

Yin, V. P., Thomson, J. M., Thummel, R., Hyde, D. R., Hammond, S. M., and Poss, K. D. (2008). Fgf-dependent depletion of microRNA-133 promotes appendage regeneration in zebrafish. Genes & Development 22, 728-733.

Yoon, Y. S., Lee, N., and Scadova, H. (2005). Myocardial regeneration with bone-marrow-derived stem cells. Biol Cell 97, 253-63.

Zhang, R., Han, P., Yang, H., Ouyang, K., Lee, D., Lin, Y. F., Ocorr, K., Kang, G., Chen, J., Stainier, D. Y., Yelon, D., and Chi, N. C. (2013). In vivo cardiac reprogramming contributes to zebrafish heart regeneration. Nature.

Zhao, R., Watt, A. J., Battle, M. A., Li, J. X., Bondow, B. J., and Duncan, S. A. (2008). Loss of both GATA4 and GATA6 blocks cardiac myocyte differentiation and results in acardia in mice. Developmental Biology 317, 614-619.

Zhao, Y., Samal, E., and Srivastava, D. (2005). Serum, response factor regulates a muscle-specific microRNA that targets Hand2 during cardiogenesis. Circulation 112, U107-U107.

Zhou, B., Ma, Q., Rajagopal, S., Wu, S. M., Domian, I., Rivera-Feliciano, J., Jiang, D., von Gise, A., Ikeda, S., Chien, K. R., and Pu, W. T. (2008). Epicardial progenitors contribute to the cardiomyocyte lineage in the developing heart. Nature 454, 109-13.

Zolotareva, A. G., and Kogan, M. E. (1978). Production of experimental occlusive myocardial infarction in mice. Cor Vasa 20, 308-14.

Zukor, K. A., Kent, D. T., and Odelberg, S. J. (2011). Meningeal cells and glia establish a permissive environment for axon regeneration after spinal cord injury in newts. Neural Dev 6, 1.