07.01.2014 Views

LIFE01200604005 Shri Somnath Ghosh - Homi Bhabha National ...

LIFE01200604005 Shri Somnath Ghosh - Homi Bhabha National ...

LIFE01200604005 Shri Somnath Ghosh - Homi Bhabha National ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

RADIATION INDUCED SIGNALING IN MAMMALIAN CELLS<br />

By<br />

SOMNATH GHOSH<br />

<strong>Bhabha</strong> Atomic Research Centre, Mumbai<br />

A thesis submitted to the Board of Studies in Life Sciences<br />

In partial fulfillment of requirements<br />

For the Degree of<br />

DOCTOR OF PHILOSOPHY<br />

of<br />

HOMI BHABHA NATIONAL INSTITUTE<br />

April, 2012


STATEMENT BY AUTHOR<br />

This dissertation has been submitted in partial fulfillment of requirements for an advanced<br />

degree at <strong>Homi</strong> <strong>Bhabha</strong> <strong>National</strong> Institute (HBNI) and is deposited in the Library to be made<br />

available to borrowers under rules of the HBNI.<br />

Brief quotations from this dissertation are allowable without special permission, provided that<br />

accurate acknowledgement of source is made. Requests for permission for extended quotation<br />

from or reproduction of this manuscript in whole or in part may be granted by the Competent<br />

Authority of HBNI when in his or her judgment the proposed use of the material is in the<br />

interests of scholarship. In all other instances, however, permission must be obtained from the<br />

author.<br />

<strong>Somnath</strong> <strong>Ghosh</strong>


DECLARATION<br />

I, hereby declare that the investigation presented in the thesis has been carried out by me. The<br />

work is original and has not been submitted earlier as a whole or in part for a degree / diploma at<br />

this or any other Institution / University.<br />

<strong>Somnath</strong> <strong>Ghosh</strong>


ACKNOWLEDGEMENTS<br />

I wish to express my heartfelt gratitude to my mentor and guide Dr (Mrs.) Malini Krishna for introducing<br />

me into the wonderful world of signal transduction. From my first steps in the vast maize of pathways to<br />

the more assured strides, she was always there, showing the way with her able guidance, keen<br />

supervision, constant encouragement and support. Her parental affection, infectious optimism, bright<br />

spirits, never say die attitude and belief in me were invaluable at every stumble or wrong turn. She<br />

provided a pleasant and optimum working environment in the lab, where a lack of resources was<br />

something unimaginable due to her managerial skills and foresight. Words can never ever be enough to<br />

thank her and I will remain indebted forever.<br />

I wish to express my deep gratitude to Dr. K.B. Sainis, Director, Biomedical Group, and<br />

Chairman of my doctoral committee for supporting and mentoring me from the day I joined training<br />

school and for always being approachable. I would like to take this opportunity to thank Dr. S.K. Apte,<br />

Associate Director (B), Biomedical Group and member of my doctoral committee for his kind support,<br />

encouragement, critical analysis of the results and valuable suggestion during my presentation. Words of<br />

thanks are due to Dr. M. Seshadri, Head, RB&HSD and member of my doctoral committee for his kind<br />

support, encouragement and for providing the facilities and pleasant working environment in the division.<br />

Many thanks are due to Dr. (Smt.) S. Zingde, Dy. Director, ACTREC, Kharghar and member of my<br />

doctoral committee for her valuable suggestion during my presentation.<br />

I wish to acknowledge all my seniors and colleagues in my division and group for all the help,<br />

support and encouragement that they have provided. I take this opportunity to thank Dr Rakesh Kumar<br />

Singh and Dr. Anirban Kumar Mitra.<br />

Special thanks are due to my lab mates and friends, Dr. (Smt.) Himanshi N Mishra and Fatema<br />

for their cooperation, help and pleasure of their company.<br />

I would like to acknowledge and thank Mr Paresh for his excellent technical assistance and<br />

Manjoor Ali for confocal microscopy. I will be failing in my duty if I do not mention the administrative<br />

staff of this department for their timely help.<br />

On the personal front, words are not enough to thank Baba and Maa for making me what I am.<br />

Special thanks are due to Raju Da, Ravi Da, Botan and Buni for their cooperation<br />

Above all I would reserve my heartiest thanks for my better half ‘Ants’, for being my pillar of<br />

strength through thick and thin and loving me for who I am. I am at a loss of words to thank Siddhant, my<br />

bundle of joy who has brought so much sunshine in our lives. My thesis could never have seen the light of<br />

the day without them.<br />

<strong>Somnath</strong> <strong>Ghosh</strong>


CONTENTS<br />

Page No.<br />

SYNOPSIS…………………………………………………………………………. 1-28<br />

LIST OF FIGURES<br />

Fig. 1.1 Pie-chart showing the percentage of radiation exposure…….. 31<br />

Fig. 1.3A Regulation of Cell Cycle by p53 and p21…………………. 50<br />

Fig. 1.3B Pathways of DSB repair……………………………………. 53<br />

Fig. 1.4 Effects of ionizing radiation on cellular response systems….. 62<br />

Fig.1.5. Biological consequences of clustered DNA damage………... 71<br />

Fig. 1.6 Radiation induced bystander effect…………………………. 76<br />

Fig. 3.1.1 Relative radioresistance of three cell line…………………. 96<br />

Fig. 3.1.2 Relative expression of genes in A549 cells……………….. 98<br />

Fig. 3.1.3 Validation of microarray data by RT-PCR………………… 126<br />

Fig.3.1.4 Gene expression of ATM, DNA-PK, Rad52 and MLH1…… 129<br />

Fig. 3.1.5 Repair kinetics as determined by γ-H2AX foci……………… 130<br />

Fig. 3.1.6A Radiation induced foci of DNA damage response proteins... 131<br />

Fig. 3.1.7 Translocation of phospho-p53……………………………….. 133<br />

Fig. 3.1.8 Response of A549 and MCF-7………………………………. 134<br />

Fig.3.1.9 Stabilization of DNA-PK mRNA in A549…………………… 136<br />

Fig.3.1.10 Effect of Rad52 and MLH1…………………………………. 137<br />

Fig.3.2.1 Clonogenic Cell Survival of A549……………………………. 140<br />

Fig.3.2.2 Gene expression of ATM, DNA-PK, ATR, Chk1 and Chk2…. 141<br />

Fig.3.2.3 Immunofluorescence staining and Image analysis…………….. 142<br />

Fig.3.2.4 Gene expression of Bax and Bcl-2…………………………….. 143<br />

Fig.3.3.1.1 Clonogenic Cell Survival of A549 cells……………………... 145<br />

Fig.3.3.1.2 Formation of γ-H2AX foci…………………………………... 147<br />

Fig.3.3.1.3 Phosphorylation of ATM……………………………………. 149<br />

Fig.3.3.1.4 Phosphorylation of ATR…………………………………….. 150<br />

Fig.3.3.1.5 Phosphorylation of BRCA1…………………………………. 151<br />

Fig.3.3.1.6 Phosphorylation of Chk1 and Chk2…………………………. 153<br />

Fig.3.3.1.7a Phosphorylation of ERK…………………………………… 154<br />

Fig.3.3.1.7b Phosphorylation of JNK…………………………………… 155<br />

Fig.3.3.1.8 Apoptosis in carbon irradiated A549 cells…………………... 156<br />

Fig. 3.3.2.1 Clonogenic Cell Survival of A549 cells……………………. 158<br />

Fig. 3.3.2.2 Phosphorylation of H2AX…………………………………. 160<br />

Fig.3.3.2.3 Phosphorylation of ATM…………………………………… 163<br />

Fig.3.3.2.4 Phosphorylation of ATR……………………………………. 164<br />

Fig.3.3.2.5 Phosphorylation of TP53……………………………………. 165<br />

Fig.3.3.2.6 Phosphorylation of Chk2……………………………………. 166<br />

Fig.3.3.2.7 Apoptosis in oxygen irradiated A549 cells………………….. 167<br />

Fig.3.4.1Bystander effect in similar cells (A549 Vs A549)……………... 170<br />

Fig.3.4.2 Bystander effect in dissimilar cells (A549 Vs U937)…………. 171


Fig.3.4.3.1 Gene expression of iNOS, p21, p53 and NF-kB (p65)……… 174<br />

Fig.3.4.3.2 NO production from EL-4 cells……………………………... 175<br />

Fig.3.4.3.3 DNA damage in EL-4 cells………………………………….. 176<br />

Fig. 3.4.3.4a DNA fragmentation assay of EL-4 cells…………………... 177<br />

Fig. 3.4.3.4b Apoptosis assay of EL-4 cells……………………………... 178<br />

Fig.3.4.4a Gene expression of NF-kB (p65) and iNOS gene…………… 180<br />

Fig.3.4.4b DNA damage in EL-4 cells…………………………………... 181<br />

Fig.3.4.5a Gene expression of NF-kB (p65) and iNOS gene…………… 183<br />

Fig.3.4.5b DNA damage in EL-4 cells…………………………………... 184<br />

LIST OF TABLES<br />

Table 2.6 The sequences of gene specific primers………………………. 87<br />

Table 3.1.2A List of Up-regulated Genes: Control Vs 5X2Gy…………. 99<br />

Table 3.1.2B List of Down-regulated Genes: Control Vs 5X2Gy……..... 107<br />

Table 3.1.2C List of Up-regulated genes: 2Gy Vs 5X2Gy……………… 111<br />

Table 3.1.2D List of Down-regulated genes: 2Gy Vs 5X2Gy…………... 116<br />

Table 3.1.2E List of Up-regulated genes: 10Gy Vs 5X2Gy…………….. 117<br />

Table 3.1.2F List of Down-regulated genes of 10Gy Vs 5X2Gy……….. 123<br />

CHAPTER 1- INTRODUCTION<br />

1.1. Ionizing Radiation………………………………………………….. 32<br />

1.2. Radiation Therapy…………………………………………………. 34<br />

1.3. DNA damage signaling following irradiation…………………….. 37<br />

1.3.1 Alarm Sensors…………………………………………… 37<br />

1.3.2 Cell Cycle Regulation…………………………………… 49<br />

1.3.3 DNA Repair……………………………………………… 52<br />

1.4. Overview of Radiation induced cell signaling…………………….. 61<br />

1.5. Charged particle irradiation induced Signaling………………….. 70<br />

1.6. Radiation induced Bystander Signaling…………………………… 75<br />

1.7. Aims and Objectives………………………………………………… 78<br />

CHAPTER 2 - MATERIALS & METHODS<br />

2.1 Chemicals............................................................................................. 80<br />

2.2 Cell Lines…………………………………………………………….. 81<br />

2.3 Irradiation............................................................................................ 81<br />

2.4 Clonogenic Cell Survival Assay……………………………………. 84<br />

2.5 Microarray…………………………………………………………... 84<br />

2.6 Semiquantitative RT-PCR…………………………………………. 86<br />

2.7 Immunofluorescence staining………………………………………. 86<br />

2.8 Image analysis using ImageJ software…………………………….. 88<br />

2.9 Western Blotting…………………………………………………….. 89<br />

2.10 Measurement of DNA damage……………………………………. 90


2.11 Measurement of NO production………………………………….. 91<br />

2.12 Measurement of Apoptosis by DNA fragmentation……………... 91<br />

2.13 Transfection with ShRNA………………………………………… 92<br />

2.14 Statistical Analysis…………………………………………………. 92<br />

CHAPTER 3 - RESULTS<br />

3.1 Fractionated irradiation induced signaling..................................... 94<br />

3.2 Proton beam induced signaling…………………………………….. 138<br />

3.3 High LET irradiation induced signaling…………………………… 144<br />

3.2.1 Carbon beam induced signaling………………………… 144<br />

3.2.2 Oxygen beam induced signaling………………………… 157<br />

3.4 Radiation induced bystander signalling............................................. 168<br />

CHAPTER 4 - DISCUSSION<br />

4.1 Fractionated irradiation induced signaling....................................... 187<br />

4.2 Proton beam induced signaling…………………………………….. 193<br />

4.3 High LET irradiation induced signaling…………………………. 196<br />

4.3.1 Carbon beam induced signaling…………………………. 196<br />

4.3.2 Oxygen beam induced signaling…………………………. 200<br />

4.4 Radiation induced bystander signaling............................................. 203<br />

4.5 Summary and Conclusion.................................................................. 208<br />

CHAPTER 5 – REFERENCES……………………………………………….. 212<br />

Publications and Presentations………………………………………………… 253


SYNOPSIS<br />

SYNOPSIS


PREAMBLE<br />

SYNOPSIS<br />

Ionising radiation leads to a cascade of signaling events which include activation<br />

of alarm signals, cytotoxic and cytoprotective signaling pathways, nitric oxide production<br />

etc. These events culminate in the death or survival of the irradiated cell. The end result<br />

seems to depend upon the pathway predominantly activated. Minute observation of these<br />

complex signaling networks reveals that this is not a random activation of all the<br />

pathways but a very specific, organized and regulated process that is controlled at<br />

multiple steps.<br />

The cellular responses to various forms of radiation are manifested as irreversible<br />

and reversible structural and functional changes in cells and cell organelles. A widely<br />

held view is that the initial reaction of a cell to DNA damage is to repair the damage.<br />

However, with increasing levels of DNA damage the cell switches to cell cycle arrest or<br />

to apoptosis. Cell cycle arrest is sometimes permanent, but ordinarily reversible allowing<br />

time for further DNA repair.<br />

The fact that radiation can lead to cell death has long been exploited effectively in<br />

the therapy of cancer. However, the survival of the radioresistant cell has been the major<br />

drawback of radiotherapy which is delivered in fractionated doses and the cells surviving<br />

the initial dose tend to develop radioresistance making them refractory to subsequent<br />

doses of radiation. Fractionated irradiation induced signaling has not been studied in<br />

detail.<br />

The work embodied in this thesis has been divided into five chapters.<br />

1. Introduction<br />

2. Materials and Methods<br />

3. Results<br />

4. Discussion<br />

5. References<br />

CHAPTER 1: INTRODUCTION<br />

The first chapter is a general introduction to radiation induced signaling with a historical<br />

perspective on radiation biology. It scans the development of radiation biology and also<br />

reviews the current literature relevant to the work embodied in this thesis. It also lists the<br />

aims and objectives of the present work. A brief overview of the chapter is outlined<br />

below.<br />

Viewed historically, all radiation science is recent, proving that science does<br />

indeed make giant strides in a brief period. Roentgen in the year 1895 initiated the<br />

process by announcing the discovery of a “new kind of penetrating ray” or x-rays [1].<br />

This was followed soon after by Becquerel’s discovery of natural radioactivity in 1896<br />

[2]. Unfortunately the biological effects of ionising radiation and radioactivity lay<br />

unexplored for quite some time after Roentgen’s initial report [3].<br />

When a radiation beam passes through a biological material or other absorbing<br />

media, some energy is imparted to the medium while some of it leaves the volume of<br />

interaction. The absorption of energy from radiation may lead to either excitation or<br />

ionisation. Radiations which lead to the latter effect are called ionising radiation and can<br />

be further classified into electromagnetic and particulate. Ionising radiation can directly<br />

interact with and damage the target molecule. This is known as the direct effect. If it<br />

interacts with the medium to produce secondary particles which, in turn, damages the<br />

target molecule, then it is known as indirect effect.<br />

1


SYNOPSIS<br />

The average energy locally imparted to the medium per unit length of the path<br />

traversed is the Linear Energy Transfer (LET) of the ionising radiation. On this basis<br />

ionising radiation can be classified into high and low LET radiation. High LET radiations<br />

are predominantly causes direct effect whereas low LET radiations are predominantly<br />

causes indirect effect [3, 4]. Direct effects occur when an ionizing particle interacts with a<br />

macromolecule in a cell (DNA, RNA, protein etc.) where as indirect effects involve the<br />

interaction of ionizing radiation with the medium in which the molecules are suspended,<br />

especially water, to produce free radicals and reactive oxygen species (ROS), in presence<br />

of oxygen, that damage critical targets.<br />

Ionising radiation exposure causes a spectrum of lesions within the cell. These<br />

include, at the DNA level, single strand breaks (ssb), double strand breaks (dsb), base<br />

damage, DNA-protein crosslinks etc. Apart from DNA damage, other lesions in cellular<br />

macromolecules (lipids, proteins etc) are also generated. DNA as a target assumes added<br />

importance due to the very limited redundancy of information in the molecule. Any<br />

irreversible damage may lead to loss of genetic information vital for cellular function and<br />

survival. However, the non DNA lesions, probably not as life threatening to the cell, can<br />

stimulate various signaling pathways which determine the final fate of the cell [5].To<br />

complicate matters a number of non-targeted and delayed effects associated with<br />

radiation exposure, referred to as “bystander effect”, may also contribute to mutagenic<br />

events and genome instability in adjacent unexposed cells [6].<br />

The very obvious application of this was the use of ionising radiation as a potent<br />

tool in cancer therapy. This took advantage of inherent radiosensitivity of tumors over<br />

surrounding normal tissue. Moreover, introduction of fractionated radiotherapy by<br />

Coutard was a major step in enhancing the efficacy of radiotherapy [7]. Radiation therapy<br />

has enjoyed tremendous success in the field of cancer, so much so that it is used in the<br />

treatment of two thirds of all cancer patients in India today. However, one of the major<br />

hurdles in the success rate is radioresistance of tumors which may be innate or acquired.<br />

Therefore, study of fractionated irradiation induced signaling in mammalian cells will be<br />

of great importance.<br />

Exposure of cells to ionising radiation leads to the activation of existing cellular<br />

response pathways [8]. These pathways are predominantly either cytoprotective or<br />

cytotoxic. The activation of the former leads to repair, survival and proliferation whereas,<br />

the activation of the latter leads to cell death. Radiation effects are mediated through the<br />

activation of damage sensors which transduce the signal through the signaling cascades.<br />

These in turn direct the response elucidated, depending upon the signaling pathway that is<br />

predominantly activated. Various studies have shown the activation of cytoprotective<br />

factors contribute to radioresistance in the tumors.<br />

Another, albeit less studied, field is high LET radiation induced signal<br />

transduction. This assumes greater importance because of the increased application of<br />

charged particle beam in radiotherapy and the emergence of the phenomenon of radiation<br />

induced bystander effect. The latter is especially relevant in case of low dose exposure to<br />

high LET radiation (e.g. From Radon) where the damage seen in much more than would<br />

be expected by extrapolating from higher doses. It would be of interest to study the effect<br />

of high LET radiation on the signaling pathways since the damage produced by high LET<br />

radiation (clustered damage) is very different from that by low LET (scattered damage)<br />

[9].<br />

2


SYNOPSIS<br />

Aims and Objectives: To Study<br />

1. Fractionated irradiation induced signaling events. The contribution of<br />

these events to the increased cell survival.<br />

2. High LET induced signaling and variance from low LET gamma<br />

radiation.<br />

3. Contribution of radiation induced bystander effect to cell survival and its<br />

mechanism.<br />

CHAPTER 2: MATERIALS AND METHODS<br />

The second chapter lists the reagents and describes in detail the methods<br />

employed in this study. A brief summary of the same is given below.<br />

2.1 Chemicals and Reagents<br />

All fine chemicals used were from Sigma, USA. Primary antibodies were from<br />

Cell Signaling Technology, USA; Calbiochem, USA; Chemicon, USA; Santa Cruz<br />

Biotechnology, USA; Sigma, USA and Transduction Laboratories, USA.<br />

2.2 Cell Lines<br />

All the cell lines used in this study were cultured in 25 cm 2 / 75 cm 2 plastic flasks<br />

in the appropriate medium supplemented with 10% fetal bovine serum (Sigma, USA.).<br />

Cells were kept at 37 0 C in humidified atmosphere with 5% CO 2 .<br />

Following are the list/table of the cell lines used in the experiments<br />

SN Cell line Description p53 status<br />

1 A549 Human Lung Adenocarcinoma Wild type<br />

2 MCF-7 Human breast adenocarcinoma Wild type<br />

3 INT 407 human intestinal epithelial cell line Wild type<br />

4 U937 Human histiocytic lymphoma (Monocyte) Negative<br />

5 EL-4 Mouse lymphoma Wild type<br />

6 RAW 264.7 Mouse macrophages Wild type<br />

3


SYNOPSIS<br />

2.3 Irradiation<br />

2.3.1] Low LET: For gamma irradiation, cells were exposed to required dose of<br />

60 Co γ radiation in in-house facility Gamma Cell 220 at dose rate of 3.5 Gy/min.<br />

Proton beam irradiation was carried out using the Radiation Biology beam line of<br />

in-house 4 MeV Folded Tandem Ion Accelerator (FOTIA) facility.<br />

2.3.2] High LET: Heavy ion irradiation was carried out using the Radiation<br />

Biology beam line of 16 MV 15UD Pelletron at Inter University Accelerator Centre, New<br />

Delhi.<br />

2.4 Clonogenic Cell Survival Assay<br />

After treatment, cells were seeded out in appropriate dilutions (500 cells/60 mm<br />

petriplate) to form colonies in 14 days. Colonies were fixed with glutaraldehyde (6.0%<br />

v/v), stained with crystal violet (0.5% w/v) and counted using a stereomicroscope.<br />

Colonies containing more than 50 cells were scored as survivors.<br />

2.5 Semiquantitative Reverse transcriptional polymerase chain reaction (RT-PCR)<br />

Total RNA from 1 x 10 6 cells was extracted after required time periods of irradiation and<br />

RT-PCR was carried out. Briefly, total RNA from A549 cells was extracted after required<br />

time periods of 1, 2, 3 and 4 hours of irradiation and eluted in 30 µl RNAase free water<br />

using RNA tissue kit (Roche). Equal amount of RNA in each group was reverse<br />

transcribed using cMaster RT kit (Eppendorf). Equal amount (2µl) of cDNA in each<br />

group was used for specific amplification of required genes using gene specific primers.<br />

The PCR condition were 94 o C, 5 min initial denaturation followed by 30 cycles of 94 o C<br />

45 s, 55 – 64 o C 45 s (depending on the annealing temperature of specific primers) 72 o C<br />

45 s and final extension at 72 o C for 10 min. Equal amount of each PCR product (10 µl)<br />

was run on 1.5 % agarose gel containing ethidium bromide in tris borate EDTA buffer at<br />

60 V. The bands in the gel were visualized under UV lamp and relative intensities were<br />

quantified using Gel doc software (Syngene).<br />

4


SYNOPSIS<br />

2.6 Immunofluorescence staining<br />

Cells were grown in cover slips which were kept in 35mm petri dishes and irradiated as<br />

mentioned above. Cell layer were washed 4 h after irradiation in PBS and fixed for 20<br />

min in 4% paraformaldehyde at room temperature. Afterwards, the cells were washed<br />

twice in PBS. For immunofluorescence staining, cells were permeabilized for 3 min in<br />

0.25% Triton X-100 in PBS, washed two times in PBS and blocked for 1 h with 5% BSA<br />

in PBS. Antibodies were diluted (1:200) in 1% BSA in PBS. Cells were incubated with<br />

primary antibodies for 1 h 30 min at room temperature, washed three times in PBS and<br />

incubated with secondary antibodies for 1 h at room temperature. Finally, cells were<br />

rinsed and mounted with ProLong Gold antifede with DAPI mounting media (Molecular<br />

Probe, USA). Images were captured using Carl Zeiss confocal microscope. Acquisition<br />

settings were optimized to obtain maximal signal in immunostained cells with minimal<br />

background.<br />

2.7 Image analysis using ImageJ software<br />

The captured images were analyzed for relative quantification of phosphorylation using<br />

ImageJ software [10]. A 25 x 25 pixel box was positioned over the fluorescent image of<br />

each cell, and the average intensity within the selection (the sum of the intensities of all<br />

the pixels in the selection divided by the number of pixels) was measured. At least 100<br />

cells per experiment were analyzed from three independent experiments.<br />

2.8 Western Blotting<br />

For the samples that had to be probed with specific antibodies, the proteins were<br />

separated by SDS-PAGE (10%) followed by transfer to nitrocellulose membrane<br />

(Hybond ECL, Amersham Pharmacia Biotech), and probed with specific antibodies. The<br />

membranes were then probed with horseradish peroxidase conjugated secondary antibody<br />

against mouse/rabbit (Roche Molecular Biochemicals, Germany) and developed using<br />

BM Chemiluminiscence Western Blotting Kit Mouse/Rabbit (Roche Molecular<br />

Biochemicals, Germany). Densitometry was done using Shimadzu CS 9000 Dual<br />

wavelength flying spot scanner. Statistical analysis was done using ANOVA.<br />

2.9 Measurement of DNA damage<br />

DNA damage in all the groups of EL-4 cells 2 h after exposure to different conditioned<br />

medium, was measured by alkaline single cell gel electrophoresis (comet assay) [11]. In<br />

brief, frosted microscope slides (Gold Coin, Mumbai, India) were covered with 200 µl of<br />

1% normal melting agarose (NMA) in phosphate buffered saline (PBS) at 45 o C, covered<br />

with a cover slip and kept at 4 o C for 10 min. After solidification a second layer of 200 µl<br />

of 0.5% low melting agarose (LMA) containing approximately 105 cells at 37 o C was<br />

layered over it. The cover slips were replaced immediately and the slides were kept at 4<br />

o C. After solidification of the LMA, the cover-slips were removed and slides were placed<br />

in the chilled lysing solution (2.5M NaCl, 100mM Na 2 -EDTA, 10mM Tris–HCl, pH 10,<br />

and 1% DMSO, 1% Triton X100 and 1% sodium sarcosinate) for 1 h at 4 o C. The slides<br />

were removed from the lysing solution and placed on a horizontal electrophoresis tank<br />

filled with freshly prepared alkaline buffer (300mM NaOH, 1mM Na 2 -EDTA and 0.2%<br />

DMSO, pH≥13.0) and were equilibrated in the above buffer for 20 min and<br />

electrophoresis was carried out at 25V for 20 min. After electrophoresis the slides were<br />

washed gently with 0.4M Tris–HCl buffer, pH 7.4, to remove the alkali, stained with 50<br />

µl of propidium iodide (PI, 20 µg/ml) and visualized using a Carl Zeiss Fluorescent<br />

5


SYNOPSIS<br />

microscope (Axioskop). The images (40–50 cells/slide) were captured with highperformance<br />

GANZ (model: ZCY11PH 4) color video camera. The integral frame<br />

grabber used in this system (Cvfbo1p) is a PC based card and it accepts color composite<br />

video output of the camera. The quantification of the DNA strand breaks of the stored<br />

images was done using the CASP software by which %DNA in tail, tail length, tail<br />

moment and Olive tail moment could be obtained directly [12].<br />

2.10 Measurement of NO production in EL-4 cells<br />

Control unirradiated cells, cells exposed to radiation and the cells exposed to different<br />

conditioned media as mentioned above were cultured for 24 h at 37 0 C in 5%CO 2 / 95%<br />

air. The production of NO in the culture supernatants was measured by assaying NO 2<br />

-<br />

using colorimetric Griess reaction [13].<br />

In brief, 100 µl of culture supernatants was incubated with an equal volume of Griess<br />

reagent [1% sulfanilamide and 0.1% N-(1-naphthyl-ethylenediamine) in 2.5% phosphoric<br />

acid; Sigma, USA] at room temperature for 10 min in a 96 well plate. The absorbance at<br />

550 nm was measured using a microplate reader (Fluostar Optima, Biotron Healthcare).<br />

Absorbance measurements were converted to µ moles of NO 2<br />

-<br />

using a standard curve of<br />

NaNO 2.<br />

2.11 Measurement of Apoptosis by DNA fragmentation<br />

Apoptosis in control unirradiated cells, cells exposed to radiation and the cells exposed to<br />

different conditioned media as mentioned above was measured by DNA fragmentation<br />

assay [14]. After transfer of medium from irradiated cells, all the above mentioned<br />

groups of EL-4 cells were cultured for 24 h at 37 0 C in 5%CO 2 , 95% air. The cells were<br />

harvested and lysed in lysis buffer (10 mM EDTA, 50 mM Tris–HCl, pH 8.0, 0.5%<br />

sodium lauryl sarcosine) containing the 100 mg/ml proteinase K at 55 o C for 2 h. The<br />

DNA was extracted with phenol/chloroform and precipitated with ethanol. The RNA was<br />

removed by RNAse A treatment and DNA was electrophorased on agarose gel to<br />

visualize the fragmented DNA.<br />

CHAPTER 3: RESULTS<br />

In this study the relative radioresistance of various cell lines (MCF-7, INT 407<br />

and A 549) was first established. Then the differences in the signaling pattern were<br />

established. The variance in signaling with the change in LET was investigated. Finally,<br />

the contribution of the bystanding cell, which had not been exposed to irradiation, to the<br />

cell survival was also looked at. The results obtained in the study are divided into the<br />

following subsections.<br />

3.1 Fractionated irradiation induced signaling and repair in mammalian cells.<br />

Three human cell lines viz. MCF-7, INT 407 and A 549 were exposed to different<br />

doses of γ irradiation; the first set was irradiated with acute dose of 2 Gy or 10 Gy. The<br />

second set was exposed to 5 fraction of 2 Gy each. Among the three cell line the lung<br />

carcinoma cell line A 549 was found to be relatively more radioresistant if the 10 Gy<br />

dose was delivered as a fractionated regimen. The MCF-7 cell line was found to be<br />

relatively more radiosensitive (Fig. 3.1A). This was further confirmed by analysis of<br />

6


SYNOPSIS<br />

ATM gene expression, which was found to significantly upregulated in A 549 cell line as<br />

compared to the MCF-7 (Fig. 3.1B). DNA-PK gene was also significantly upregulated in<br />

A 549 cell line which had been exposed to 5 fraction of 2 Gy γ-radiation (Fig. 3.1C). As a<br />

consequence of this the Phospho-p53 was found to be translocated to the nucleus of A<br />

549 cell line which had been exposed to fractionated regimen. The expression of<br />

Phospho- p53 was confirmed by western blotting (Fig. 3.1D).<br />

The fractionated dose regimen led to a significant upregulation of ATM and DNA-PK in<br />

A549 cells line. The differences in the two cell lines (A549 and MCF-7) could be<br />

summarized as significant differences in the repair capability of the two. Significant<br />

activation of BRCA1, phospho ATM and Rad52 seems to be causative factors in A549<br />

cells. None of this activation was apparent in MCF-7 (Fig. 3.1E).<br />

The activation of DNA-PK was accompanied by the stabilization of its mRNA and was<br />

regulated by the p38 MAP Kinase pathway but not by ERK or JNK MAP Kinase<br />

pathway (Fig. 3.1F). The ATM and Rad 52 mRNA were not stabilized indicating a<br />

specific and selective stabilization of DNA-PK mRNA. Finally the inhibition of Rad52<br />

with its ShRNA reversed the radioresistance of the A549 cells and made it sensitive to<br />

fractionated irradiation (Fig. 3.1G).<br />

3.2 Proton beam induced signaling in mammalian cells and its comparison with<br />

Gamma irradiation.<br />

Having established that A549 cells are relatively more radioresistant. The<br />

variance in signaling pattern and effectiveness of cell killing with the change in LET was<br />

looked at. A549 Lung Adenocarcinoma cells were irradiated with 2 Gy proton beam or γ-<br />

radiation. Proton beam was found to be more cytotoxic than γ-radiation (Fig. 3.2A).<br />

Proton beam irradiated cells showed phosphorylation of H2AX, ATM, Chk2 and p53<br />

(Fig. 3.2B). The mechanism of excessive cell killing in proton beam irradiated cells was<br />

found to be upregulation of Bax and down-regulation of Bcl-2 (Fig. 3.2C). The<br />

noteworthy finding of this study is the biphasic activation of the sensor proteins, ATM<br />

and DNA-PK and no activation of ATR by proton irradiation (Fig. 3.2D).<br />

3.3 High LET irradiation induced signaling in mammalian cells and its comparison<br />

with Gamma irradiation.<br />

After establishing the signaling differences between proton beam and gamma ray in lung<br />

carcinoma cells. The variance in signaling pattern with higher LET (carbon and oxygen)<br />

was studied.<br />

Lung carcinoma cell line A549 were irradiated with 1 or 2 Gy doses of carbon ions,<br />

oxygen ions or gamma rays and ensuing activation of few important components<br />

involved in DNA repair, cell cycle arrest, apoptosis and survival pathways were<br />

followed.<br />

Expectedly the carbon beam was found to be three times more cytotoxic than γ-radiation<br />

despite the fact that the number of H2AX foci was the same with both the radiations. The<br />

repair, although activated in carbon beam-irradiated cells, was not effective in reviving<br />

the cells. The carbon beam irradiated cells showed increased phosphorylation of primary<br />

regulators of Homologous Recombination Repair (HRR) pathway i.e. H2AX, ATM,<br />

BRCA1 and Chk2 as compared to gamma irradiation (Fig. 3.3). The noteworthy finding<br />

of this study was the biphasic activation of the pH2AX and the unusually large foci of<br />

BRCA1. Apoptosis was found to be p53 independent.<br />

7


SYNOPSIS<br />

Comparison of carbon and oxygen beam irradiation induced signaling<br />

The differences in activation of signaling components with respect to LET and fluence of<br />

radiation can be very helpful in elucidating repair mechanisms in cells. Foci formation of<br />

pH2AX, pATM and pATR was also compared between carbon and oxygen at 1 Gy,<br />

which differ on the basis of LET by a factor of 2. The dose was kept constant since LET<br />

of oxygen is double of carbon, the fluence of oxygen should be half that of carbon ions.<br />

The probability is that with carbon irradiation one cell may be hit by 2.8 carbon ions and<br />

with oxygen irradiation one cell may be hit by 1.3 oxygen ions. Thus the probability that<br />

a cell is not being hit by oxygen increases the bystander responses with oxygen ion<br />

irradiation. pH2AX foci were observed in 60% of the cells at 15 minutes indicating that<br />

only 60% of the cells might have been hit by oxygen (Fig. 3.3I). However, presence of<br />

significant number of pATM foci in 96% of the cells indicates activation of ATM in<br />

bystander cells. pATR foci were also significantly high in oxygen irradiated cells and was<br />

found in 36% of the cells unlike in carbon irradiated cells where only10% cells showed<br />

the foci. This could indicate that complex damage may lead to increased activation of<br />

ATR (Fig. 3.3I).<br />

3.4 Radiation induced bystander signaling in mammalian cells.<br />

To establish the contribution of the bystander effect in the survival of the<br />

neighbouring cells, the lung carcinoma A549 cells were exposed to 2 Gy γ-irradiation.<br />

The medium from irradiated cells was transferred to unirradiated cells. Irradiated A549<br />

cells as well as those cultured in the presence of medium from γ-irradiated A 549 cells<br />

showed decrease in survival, increase in γ-H2AX and pATM foci (Fig. 3.4A). Bystander<br />

signaling was also found to exist between U937 cells (human monocyte) and Lung<br />

carcinoma A549 cells. PMA stimulated and irradiated U937 cells induced bystander<br />

response in unirradiated A549 cells. The latter showed a decrease in survival, increase in<br />

γ-H2AX and pATM foci (Fig. 3.4B). The unstimulated and/or irradiated U937 cells did<br />

not induce bystander effect in unirradiated A549 cells. The mechanism of such a cross<br />

bystander effect was investigated in mouse cell line.<br />

EL-4 and RAW 264.7 cells were exposed to 5 Gy γ-irradiation. The medium from<br />

irradiated cells was transferred to unirradiated cells. Irradiated EL-4 cells as well as those<br />

cultured in the presence of medium from γ-irradiated EL-4 cells showed an upregulation<br />

of NF-κB, iNOS, p53 and p21/waf1 genes (Fig. 3.4C). The directly irradiated and the<br />

bystander EL-4 cells showed an increase in NO production (Fig. 3.4D), DNA damage<br />

(Fig. 3.4E) and apoptosis. Bystander signaling was also found to exist between RAW<br />

264.7 (macrophage) and EL-4 (lymphoma) cells. Unstimulated and/or irradiated RAW<br />

264.7 cells did not induce bystander effect in unirradiated EL-4 cells but LPS stimulated<br />

and irradiated RAW 264.7 cells induced an upregulation of NF-κB and iNOS genes (Fig.<br />

3.4F) and increased the DNA damage in bystander EL-4 cells. Treatment of EL-4 or<br />

RAW 264.7 cells with L-NAME significantly reduced the induction of gene expression<br />

and DNA damage in the bystander EL-4 cells while treatment with cPTIO only partially<br />

reduced the induction of gene expression and DNA damage in the bystander EL-4 cells<br />

(Fig. 3.4G). It was concluded that active iNOS in the irradiated cells was essential for<br />

bystander response.<br />

8


SYNOPSIS<br />

Percentage Survival<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

MCF7<br />

INT407<br />

A549<br />

0Gy 2Gy 5X2Gy 10Gy<br />

Radiation Dose<br />

Fig. 3.1A. Relative radioresistance of three cell line viz. MCF-7, INT 407 and A 549<br />

Data represents means ± SE of three independent experiments. Lane 1, control<br />

unirradiated A549 cell line. Key; Lane 2, 2 Gy acute dose ( single dose that was<br />

delivered in each fraction ) of 60 Co γ-irradiation Lane 3 2 Gy of 60 Co γ-irradiation daily<br />

over a period of 5 days, Lane 4, 10 Gy acute (which is the sum total of all the fraction )<br />

60 Co γ-irradiation.<br />

Ratio of ATM to Actin<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

Con 2 Gy 5X2 Gy 10 Gy<br />

Treatment<br />

MCF<br />

INT407<br />

A549<br />

Fig. 3.1B. Gene expression of ATM in MCF-7, INT 407 and A 549 cells. Key;<br />

Lane 1, control unirradiated A549 cell line. Lane 2, 2 Gy acute dose ( single dose that<br />

was delivered in each fraction ) of 60 Co γ-irradiation Lane 3 2 Gy of 60 Co γ-irradiation<br />

daily over a period of 5 days, Lane 4, 10 Gy acute (which is the sum total of all the<br />

fraction ) 60 Co γ-irradiation.<br />

9


SYNOPSIS<br />

Ratio of DNA-PK to Actin<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

1 2 3 4<br />

Lane<br />

Fig. 3.1C. Gene expression of DNA-PK in A 549 cells. Key; Lane 1, control unirradiated<br />

A549 cell line. Lane 2, 2 Gy acute dose ( single dose that was delivered in each fraction )<br />

of 60 Co γ-irradiation Lane 3 2 Gy of 60 Co γ-irradiation daily over a period of 5 days,<br />

Lane 4, 10 Gy acute (which is the sum total of all the fraction ) 60 Co γ-irradiation.<br />

Relative phosphorylation of p53<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

p-TP53<br />

Con 2Gy 5X2Gy 10Gy<br />

Loading<br />

control<br />

(ponceau)<br />

Con 2Gy 5X2Gy 10Gy<br />

Fig. 3.1D Expression of phospho-p53 in A 549 cells. Key; Lane 1, control unirradiated<br />

A549 cell line. Lane 2, 2 Gy acute dose ( single dose that was delivered in each fraction )<br />

of 60 Co γ-irradiation Lane 3 2 Gy of 60 Co γ-irradiation daily over a period of 5 days,<br />

Lane 4, 10 Gy acute (which is the sum total of all the fraction ) 60 Co γ-irradiation.<br />

10


SYNOPSIS<br />

Ratio of Rad52 to Actin<br />

Ratio of MLH1 to Actin<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

Role of Rad 52 and MLH1 in relative radioresistance<br />

Con 2Gy 5X2Gy 10Gy<br />

Con 2Gy 5X2Gy 10Gy<br />

Rad52<br />

Actin<br />

MLH1<br />

Ratio of Rad52 to Actin<br />

Ratio of MLH1 to Actin<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

Con 2Gy 5X2Gy 10Gy<br />

Treatment<br />

Con 2Gy 5X2Gy 10Gy<br />

Treatment<br />

Rad52<br />

MLH1<br />

Rad52<br />

Actin<br />

MLH1<br />

Actin<br />

A549 cells<br />

Actin<br />

MCF-7 cells<br />

Fig. 3.1E. Gene expression of Rad52 and MLH1 in MCF-7 and A 549 cells. Key;<br />

Lane 1, control unirradiated A549 cell line. Lane 2, 2 Gy acute dose ( single dose that<br />

was delivered in each fraction ) of 60 Co γ-irradiation Lane 3 2 Gy of 60 Co γ-irradiation<br />

daily over a period of 5 days, Lane 4, 10 Gy acute (which is the sum total of all the<br />

fraction ) 60 Co γ-irradiation.<br />

mRNA stability of DNA-PK<br />

Relative expression of DNA-PK<br />

30000<br />

25000<br />

20000<br />

15000<br />

10000<br />

5000<br />

0<br />

2h+Act 2h+Act+In 4h+Act 4h+Act+In 6h+Act 6h+Act+In 8h+Act 8h+Act+In<br />

Time (h) 2 2 4 4 6 6 8 8<br />

Actinomycin (5mg/ml) + + + + + + + +<br />

p38 MAPK inhibitor (1µM) - + - + - + - +<br />

DNA-PK<br />

Fig. 3.1F Stabilization of DNA-PK mRNA in A549, which had been exposed to 2 Gy of<br />

60 Co γ-irradiation daily over a period of 5 days and its stabilization is control by p38<br />

MAP Kinase.<br />

11


Role of Rad52 and MLH1 In relative radioresistance of A549 cells<br />

SYNOPSIS<br />

Control 5X2Gy 5X2Gy + MLH1 5X2Gy + Rad52<br />

Fig. 3.1G Inhibition of Rad52 with its SiRNA reversed the radioresistance of the A549<br />

cells and made it sensitive to fractionated irradiation. Key; Lane 1, control unirradiated<br />

A549 cell line. Lane 2, 2 Gy of 60 Co γ-irradiation daily over a period of 5 days Lane 3, A<br />

549 cells transfected with MLH1 SiRNA and exposed to 2 Gy of 60 Co γ-irradiation daily<br />

over a period of 5 days, Lane 4, A 549 cells transfected with Rad52 SiRNA and exposed<br />

to 2 Gy of 60 Co γ-irradiation daily over a period of 5 days.<br />

Fig. 3.2A Clonogenic Cell Survival of A 549 cells irradiated with 2Gy Proton Beam or<br />

2Gy γ-radiation. Data shown from representative experiment of three independent<br />

experiments.<br />

12


Relative Phosphorylation<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Control<br />

Gamma<br />

Proton<br />

y-H2AX p-ATM p-Chk2 p-TP53<br />

Signaling Molecules<br />

SYNOPSIS<br />

Fig.3.2B Immunofluorescence staining and Image analysis of γ-H2AX, phospho-ATM,<br />

phospho-Chk2 and phospho-p53 4h after irradiation in A 549 cells irradiated with 2Gy<br />

Proton Beam or 2Gy γ-radiation. Graph represents relative phosphorylation of H2AX,<br />

ATM, Chk2 and p53 as determined by ImageJ software. At least 100 cells per experiment<br />

were analyzed from three independent experiments. Data represents means ± SE of three<br />

independent experiments; significantly different from unirradiated controls: *P < 0.05,<br />

**P < 0.01.<br />

Fig 3.2C Gene expression of Bax and Bcl-2 12h after irradiation in A 549 cells irradiated<br />

with 2Gy Proton Beam or 2Gy γ-radiation. Con means unirradiated control, H means<br />

proton. Data represents means ± SE of three independent experiments; significantly<br />

different from unirradiated controls: *P < 0.05, **P < 0.01.<br />

13


SYNOPSIS<br />

Fig. 3.2D Gene expression of ATM, DNA-PK, ATR, Chk1 and Chk2 in A 549 cells<br />

irradiated with 2Gy Proton Beam or 2Gy γ-radiation at 2, 3 or 4h after irradiation. Total<br />

RNA from A 549 cells was isolated and reverse transcribed. RT-PCR analysis of ATM,<br />

DNA-PK, ATR, Chk1 and Chk2 genes was carried out as described in materials and<br />

methods. PCR products were resolved on 1.5% agarose gels containing ethidium<br />

bromide. β-Actin gene expression in each group was used as an internal control. Ratio of<br />

intensities of (A & B) ATM, DNA-PK or ATR, (C & D) Chk1 and Chk2 band to that of<br />

respective β-actin band as quantified from gel pictures are shown above each gel picture.<br />

Con means unirradiated control. Data represents means ± SE of three independent<br />

experiments; significantly different from unirradiated controls: *P < 0.05, **P < 0.01.<br />

14


SYNOPSIS<br />

Fig. 3.3A Clonogenic Cell Survival of A 549 cells irradiated with 1, 2 or 3Gy carbon<br />

Beam or γ-radiation. Data represents means ± SD of three independent experiments.<br />

(A)<br />

(B)<br />

Average y-H2AX foci per cell<br />

Relative γ−H2AX intensity<br />

24 Con<br />

22<br />

20<br />

18<br />

15 m<br />

4 hrs<br />

100%<br />

16<br />

100%<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

57%<br />

2<br />

10%<br />

10%<br />

22%<br />

0<br />

Gamma<br />

Carbon<br />

1 Gy Dose<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

Con<br />

15m<br />

4h<br />

1<br />

0<br />

Gamma<br />

Carbon<br />

1Gy Dose<br />

Relative Intensity pH2AX<br />

(C)<br />

Lane<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

pH2AX Carbon<br />

pH2AX Gamma<br />

Con 0.2 0.5 1 2 3 4 6<br />

Time (hours)<br />

Gamma<br />

Carbon<br />

C 10 30 1 2 3 4 6<br />

Fig. 3.3B Phosphorylation of H2AX (γ-H2AX) at different time points after exposure to 1<br />

Gy gamma or carbon irradiation in A 549 cells. (A) Graph represents average numbers of<br />

foci per cell, percentage of cells showing the foci is marked above the bars. (B) Graph<br />

represents relative phosphorylation of H2AX as determined by ImageJ software. At least<br />

100 cells per experiment were analyzed from three independent experiments. (C)<br />

Represenative image of immunoblot and histogram showing phosphorylation of H2AX at<br />

various time periods (0.2, 0.5, 1, 2, 3, 4 and 6 hours). Relative pH2Ax band intensity<br />

shown in histogram was divided by ponceau as loading control. Data represents means ±<br />

SD of three independent experiments.<br />

15


SYNOPSIS<br />

(A)<br />

Average pATM foci per cell<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Con<br />

1Gy<br />

52%<br />

88%<br />

25%<br />

22%<br />

Gamma<br />

Carbon<br />

(B)<br />

Relative intensity of phospho ATM<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Con<br />

1Gy<br />

Gamma<br />

Carbon<br />

Fig.3.3C Phosphorylation of ATM at 4h after exposure to 1 Gy gamma or carbon<br />

irradiation in A 549 cells. (A) Graph represents average numbers of foci per cell,<br />

percentage of cells showing the foci is marked above the bars. (B) Graph represents<br />

relative phosphorylation of ATM as determined by ImageJ software. At least 100 cells<br />

per experiment were analyzed from three independent experiments. Data represents<br />

means ± SD of three independent experiments.<br />

(A)<br />

Average ATR foci per cell<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

53%<br />

Gamma<br />

Con<br />

1Gy<br />

10%<br />

Carbon<br />

(B)<br />

Relative intensity of pATR<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Gamma<br />

Con<br />

1Gy<br />

Carbon<br />

Fig.3.3D Phosphorylation of ATR at 4h after exposure to 1 Gy gamma or carbon<br />

irradiation in A 549 cells. (A) Graph represents average numbers of foci per cell,<br />

percentage of cells showing the foci is marked above the bars. (B) Graph represents<br />

relative phosphorylation of ATR as determined by ImageJ software. At least 100 cells per<br />

experiment were analyzed from three independent experiments. Data represents means ±<br />

SD of three independent experiments.<br />

16


SYNOPSIS<br />

(A)<br />

Av No. of phospho BRCA1 Foci per cell<br />

22<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

55%<br />

Gamma<br />

Con<br />

1 Gy<br />

100%<br />

Carbon<br />

(B)<br />

Relative phosphorylation of BRCA<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

Gamma<br />

Con<br />

1Gy<br />

Carbon<br />

Fig. 3.3E Phosphorylation of BRCA1 at 4h after exposure to 1 Gy gamma or carbon<br />

irradiation in A 549 cells. (A) Graph represents average numbers of foci per cell,<br />

percentage of cells showing the foci is marked above the bars. (B) Graph represents<br />

relative phosphorylation of BRCA1 as determined by ImageJ software. At least 100 cells<br />

per experiment were analyzed from three independent experiments. Data represents<br />

means ± SD of three independent experiments.<br />

(A)<br />

Relative intensity of pChk1<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Con<br />

1Gy<br />

(C)<br />

Relative intensity of Chk2<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Con<br />

1Gy<br />

0.0<br />

Gamma<br />

Carbon<br />

0.0<br />

Gamma<br />

Carbon<br />

Control<br />

1Gy<br />

(B)<br />

p-Chk2 (Gamma) p-Chk2 (Carbon)<br />

Fig.3.3F Phosphorylation of Chk1 and Chk2 at 4h after exposure to 1 Gy gamma or carbon<br />

irradiation in A 549 cells. (A) Graph represents relative phosphorylation of Chk1 as determined<br />

by ImageJ software (B) Representative image showing phosphorylation and foci formation of<br />

Chk2 shown by “arrow” between 1Gy carbon beam and 1Gy γ-radiation. Each phospho-Chk2<br />

antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green). All<br />

images were captured using Carl Zeiss confocal microscope with the same exposure time (C)<br />

Graph represents relative phosphorylation of Chk2 as determined by ImageJ software. At least<br />

100 cells per experiment were analyzed from three independent experiments. Data represents<br />

means ± SD of three independent experiments.<br />

17


SYNOPSIS<br />

Relative phosphorylation<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

(A)<br />

pERK(gamma)<br />

pERK(carbon)<br />

Con 0.1 0.5 1 2 3 4 6<br />

Time after irradiation (hrs)<br />

Relative phosphorylation<br />

(B)<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

pJNK(gamma)<br />

pJNK(carbon)<br />

Con 0.1 0.5 1 2 3 4 6<br />

Time after irradiation (Hrs)<br />

Carbon<br />

Gamma<br />

Con 10 30 1 2 3 4 6 Lane Con 10 30 1 2 3 4 6<br />

pERK<br />

pJNK<br />

Fig.3.3G Phosphorylation of ERK (pERK) and JNK (pJNK) in A549 cells cells at<br />

different time periods (hours) after exposure to 1 Gy dose of gamma or carbon ion<br />

irradiation. Treated and untreated cells were immunoblotted and detected using antiphospho<br />

ERK (panel A) or anti phospho JNK (panel B). One representative image of<br />

three independent experiments is shown here.<br />

Bar graph represents the relative pERK (panel A) and pJNK (panel B) band intensities<br />

plotted against time. The intensities of pERK/pJNK were obtained after dividing them<br />

with the respective loading control intensities and have been normalized for comparison.<br />

(A)<br />

30<br />

Control 1Gy Gamma 1Gy Carbon<br />

(B)<br />

% Apoptotic Nuclei at 4h<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Con<br />

Gamma<br />

Carbon<br />

1 Gy<br />

Con 2h 3h 4h 6h<br />

(C)<br />

Bax<br />

Lane<br />

Fig.3.3H Apoptosis in carbon irradiated A549 cells. Panel A, Apoptotic nuclei of A549 cells were<br />

visualized and quantified by using DAPI staining, 4 hours after carbon or gamma irradiation<br />

(1Gy). Panel B, The percentage of cells with apoptotic nuclei are represented as graph. At least<br />

100 cells per experiment were analyzed from three independent experiments. Panel C,<br />

Representative immunoblot image depicting upregulation of Bax, observed at various time<br />

periods (2, 3, 4 and 6 hours) after carbon irradiation (1 Gy).<br />

18


SYNOPSIS<br />

Percentage Survival<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

(A)<br />

Carbon beam<br />

Oxygen beam<br />

Con 1 Gy 2 Gy<br />

Irradiation Dose<br />

Average H2Ax foci per cell<br />

26<br />

24<br />

22<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

(B)<br />

100%<br />

22%<br />

10%<br />

Carbon<br />

Con<br />

1Gy 15m<br />

1Gy 4h<br />

61%<br />

23%<br />

10%<br />

Oxygen<br />

Average pATM foci per cell<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

(C)<br />

22%<br />

Carbon<br />

52%<br />

Con<br />

1Gy<br />

96%<br />

19%<br />

Oxygen<br />

Average ATR foci per cell<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

(D)<br />

10%<br />

Carbon<br />

Con<br />

1Gy<br />

36%<br />

Oxygen<br />

Fig.3.3I Comparison of carbon and oxygen beam irradiation induced signaling. (A)<br />

Graph represents Clonogenic Cell Survival of A 549 cells irradiated with 1 or 2Gy carbon<br />

or oxygen beam. Data represents means ± SD of three independent experiments. (B-D)<br />

Graphs represent average numbers of foci per cell of γ-H2AX, p-ATM and p-ATR<br />

respectively, percentage of cells showing the foci is marked above the bars. At least 100<br />

cells per experiment were analyzed from three independent experiments. Data represents<br />

means ± SD of three independent experiments.<br />

19


SYNOPSIS<br />

Fig.3.4A. Irradiated A549 cells as well as those cultured in the presence of medium from<br />

γ-irradiated A 549 cells (Bystander) showed decrease in survival, increase in γ-H2AX<br />

and pATM foci<br />

20


SYNOPSIS<br />

Fig.3.4B. Existance of cross bystander effect between Human Lung carcinoma A549<br />

cells and Human Monocyte U937 cells. Key: Lane 1, control unirradiated A549 cells;<br />

Lane 2, unirradiated A549 cells receiving medium from PMA-stimulated (32nM) and γ<br />

irradiated U937 cells; Lane 3, unirradiated A549 cells receiving medium from γ irradiated<br />

U937 cells without PMA-stimulation; Lane 4, unirradiated A549 cells receiving medium<br />

from PMA-stimulated U937 cells; Lane 5, unirradiated A549 cells receiving medium<br />

from unirradiated U937 cells without PMA-stimulation.<br />

21


SYNOPSIS<br />

Fig. 3.4C Gene expression of iNOS, p21, p53 and NF-kB (p65) in EL-4 cells cultured for<br />

2 h in presence of different conditioned media. Key: Lane 1, control unirradiated EL-4<br />

cells; Lane 2, γ irradiated EL-4 cells; Lane 3, unirradiated EL-4 cells receiving medium<br />

from γ irradiated EL-4 cells; Lane 4, unirradiated EL-4 cells receiving medium irradiated<br />

with γ rays in the absence of cells; Lane 5, unirradiated EL-4 cells receiving medium<br />

from unirradiated EL-4 cells. Data represents means ± SE of three independent<br />

experiments; significantly different from unirradiated controls: *P < 0.05, **P < 0.01.<br />

Fig. 3.4D NO production from EL-4 cells cultured for 24 h in the presence of different<br />

conditioned media. The culture supernatant from each group of EL-4 cells was used for<br />

determination of NO 2<br />

-<br />

with Griess reagent. Key: Lane 1, control unirradiated EL-4 cells; Lane 2,<br />

γ irradiated EL-4 cells; Lane 3, unirradiated EL-4 cells receiving medium from γ irradiated EL-4<br />

cells; Lane 4, unirradiated EL-4 cells receiving medium irradiated with γ rays in the absence of<br />

cells; Lane 5, unirradiated EL-4 cells receiving medium from unirradiated EL-4 cells. Data<br />

represents means ± SE of three independent experiments; significantly different from unirradiated<br />

controls: *P < 0.05, **P < 0.01.<br />

22


SYNOPSIS<br />

Fig. 3.4EDNA damage in EL-4 cells cultured for 2 h in presence of different conditioned media<br />

was measured using single cell gel electrophoresis (‘comet assay’), (A) Tail length and (B) %<br />

DNA in tail. Key: Lane 1, control unirradiated EL-4 cells; Lane 2, γ irradiated EL-4 cells; Lane<br />

3, unirradiated EL-4 cells receiving medium from γ irradiated EL-4 cells; Lane 4, unirradiated<br />

EL-4 cells receiving medium irradiated with γ rays in the absence of cells; Lane 5, unirradiated<br />

EL-4 cells receiving medium from unirradiated EL-4 cells. Data represents means ± SE of three<br />

independent experiments; significantly different from unirradiated controls: *P < 0.05, **P <<br />

0.01.<br />

Fig. 3.4F Gene expression of NF-kB (p65) and iNOS gene in EL-4 cells which had received<br />

medium from irradiated RAW 264.7 cells. EL-4 cells cultured for 2 h in presence of different<br />

conditioned media. Key: Lane 1, control EL-4 cells; Lane 2, EL-4 cells receiving medium from<br />

LPS-stimulated (500ng/ml) and γ irradiated RAW 264.7 cells; Lane 3, EL-4 cells receiving<br />

medium from γ irradiated RAW 264.7 cells without LPS-stimulation; Lane 4, EL-4 cells<br />

receiving medium from LPS-stimulated RAW 264.7 cells; Lane 5, EL-4 cells receiving medium<br />

from unirradiated RAW 264.7 cells. Data represents means ± SE of three independent<br />

experiments; significantly different from unirradiated controls: *P < 0.05, **P < 0.01.<br />

23


SYNOPSIS<br />

Fig. 3.4G Gene expression of NF-kB (p65) and iNOS gene in EL-4 cells which had<br />

received medium from irradiated cells. EL-4 cells cultured for 2 h in presence of different<br />

conditioned media. Key: Lane 1, control EL-4 cells; Lane 2, γ irradiated EL-4 cells;<br />

Lane 3, EL-4 cells receiving medium from γ irradiated EL-4 cells; Lane 4, EL-4 cells<br />

receiving medium from L-NAME treated and γ irradiated EL-4 cells; Lane 5, EL-4 cells<br />

receiving medium from cPTIO treated and γ irradiated EL-4 cells; Lane 6, EL-4 cells<br />

receiving medium from LPS-stimulated (500ng/ml), L-NAME treated and γ irradiated<br />

RAW 264.7 cells; Lane 7, EL-4 cells receiving medium from LPS-stimulated<br />

(500ng/ml), cPTIO treated and γ irradiated RAW 264.7 cells. Data represents means ± SE<br />

of three independent experiments; significantly different from unirradiated controls: *P <<br />

0.05, **P < 0.01.<br />

CHAPTER 4: DISCUSSION<br />

The work embodied in this thesis clearly shows that the fractionated doses led to a<br />

signaling response that was different from a single dose of irradiation. The work also<br />

stresses on the signaling differences between Low and High LET radiation and the fine<br />

tuning of bystander signaling. The observed results have been extensively discussed in<br />

this chapter and their significance and impact has been discussed in context with the<br />

current literature. The conclusions drawn from this study has also been included in this<br />

chapter.<br />

4.1 Fractionated irradiation induced signaling and repair in mammalian cells.<br />

In tumor radiotherapy, radiation is given as fractionated doses [4]. This leads to<br />

development of radioresistance in the surviving tumor cells. Chronic exposure of cells to<br />

ionizing radiation induces an adaptive response that results in the increased tolerance to<br />

24


SYNOPSIS<br />

subsequent cytotoxicity caused by the same [15-17]. However the mechanism and the<br />

factors that contribute to radioresistance of an irradiated cell are yet not known. It was of<br />

interest to study the signaling factors that contribute to this phenomenon. A549 cells were<br />

found to be relatively more radioresistant if the 10Gy dose was delivered as a fractionated<br />

regimen. Microarray analysis showed upregulation of DNA repair and cell cycle arrest<br />

genes in fractionated regimen. We have also observed intense activation of DNA repair<br />

pathway (DNA-PK, ATM, Rad52, MLH1 and BRCA1), efficient DNA repair and<br />

phospho-p53 was found to be translocated to the nucleus of A549 cells (5X2 Gy<br />

treatment group). When we silenced the Rad52 gene in A549 cells, the sensitivity of the<br />

cells (5X2 Gy treatment group) increased by more than 50%. The above results indicate<br />

that fractionated dose led to increased radioresistance in A549 cells and Rad52 gene is<br />

one of the factors responsible for increased radioresistance in A549 cells.<br />

4.2 Proton beam induced signaling in mammalian cells and its comparison with<br />

Gamma irradiation.<br />

Exposure of mammalian cells to ionising radiation leads to the activation of several<br />

signaling pathways that result in apoptosis. Although, the end point, apoptosis is<br />

effectively activated by conventional X-ray treatment, the development of radioresistance<br />

following therapy remains a major cause for concern. Clinicians are now favoring proton<br />

beam therapy to avoid the issue [18-23]. However, the mechanism of cell killing by<br />

proton beam is not yet well delineated.<br />

In this study, A549 cells were irradiated with either 2Gy proton beam or 2Gy γ-radiation<br />

and ensuing activation of few important components involved in DNA repair, cell cycle<br />

arrest, and apoptosis pathways were followed to elucidate the mechanisms behind<br />

observed cellular response. Increased phosphorylation of p53, upregulation of Bax and<br />

down-regulation of Bcl-2 in proton beam irradiated A549 cells as compared to γ-<br />

irradiated cells indicate the activation of apoptotic pathways. The noteworthy finding of<br />

this study is the biphasic activation of the sensor proteins, ATM and DNA-PK and no<br />

activation of ATR by proton irradiation and the significant activation of Chk2 even at the<br />

gene level only in the proton beam irradiated cells. Unlike gamma irradiation, the<br />

biphasic induction of ATM and DNA-PK following proton beam irradiation could be<br />

responsible for the activation of apoptotic machinery, however, further studies in this<br />

direction will reveal the importance of such biphasic response.<br />

4.3 High LET irradiation induced signaling in mammalian cells and its comparison<br />

with Gamma irradiation.<br />

Because of the effective cell killing clinicians now favor charged particle beam therapy<br />

over conventional photons for tumors that are radioresistant and when the survival of<br />

nearby cells cannot be compromised [24-26]. Although charged particle beams like<br />

carbon etc are currently being used for the treatment of many cancers [25-27], the<br />

signaling pathways activated and their variance from γ -irradiation is not well understood.<br />

These results suggest that while ATM has been shown to be the critical regulator of low<br />

LET radiation induced responses, BRCA1 seems to dominate the high LET induced<br />

response. The subtle differences leading to different pathways of initiation of downstream<br />

responses with respect to radiation quality seem to lie in pattern of the phosphorylationdephosphorylation<br />

of early response proteins such as H2AX rather than their immediate<br />

phosphorylation per se.<br />

25


SYNOPSIS<br />

4.4 Radiation induced bystander signaling in mammalian cells.<br />

It is now apparent that the target for biological effects of ionizing radiation is not solely<br />

the irradiated cell but also the surrounding cells and tissues. Preliminary evidence<br />

indicates that similar signal transduction pathways are activated in directly irradiated<br />

cells and the bystander cells and contribute to the induction of bystander DNA damage<br />

[28]. Many genes have been shown to be activated in bystander cells [29]but evidence for<br />

genes that are involved in the signaling pathways is lacking.<br />

The present study was carried out to investigate what genes are activated in the<br />

bystanding cell and whether the state of activation of the irradiated cell alters the<br />

bystander response. Bystander signaling was found to exist between U937 cells (human<br />

monocyte) and Lung carcinoma A549 cells. The mechanism of such a cross bystander<br />

signaling was investigated in mouse cell line. Bystander signaling was also found to exist<br />

between RAW 264.7 (macrophage) and EL-4 (lymphoma) cells. Unstimulated and/or<br />

irradiated RAW 264.7 cells did not induce bystander effect in unirradiated EL-4 cells but<br />

LPS stimulated and irradiated RAW 264.7 cells induced an upregulation of NF-κB and<br />

iNOS genes and increased the DNA damage in bystander EL-4 cells. Treatment of EL-4<br />

or RAW 264.7 cells with L-NAME (NOS inhibitor) significantly reduced the induction<br />

of gene expression and DNA damage in the bystander EL-4 cells while treatment with<br />

cPTIO (NO scavenger) only partially reduced the induction of gene expression and DNA<br />

damage in the bystander EL-4 cells. These results showed that active iNOS in the<br />

irradiated cells was essential for bystander response and some long-lived bioactive<br />

factors downstream of NO are most likely involved in this bystander response.<br />

Summary:<br />

The irradiation of cells with fractionated doses led to a signaling response that was<br />

different from a single dose of irradiation. The reasons for the radioresistance of the<br />

A549 cells lay in the repair pathway, the Rad52, the inhibition of which could revert the<br />

cells to radiosensitivity leading to a decrease in survival.<br />

The contribution of the bystander effect to the decrease in survival of A549 cells<br />

cannot be ruled out. There was a significant bystander response by unirradiated A549<br />

cells. The mechanism of which may be via NO generation.<br />

The effectiveness of heavy ions (carbon and oxygen) in decreasing the survival of<br />

A549 cells could be due to the lack of efficient repair despite the activation of the repair<br />

pathways (HRR). The pathway to apoptosis seems to be p53 independent.<br />

Conclusion:<br />

The lung carcinoma cell line A549, a highly radioresistant cell line could be<br />

effectively made radiosensitive by inhibiting Rad52, one of the components of its repair<br />

pathway. The survival of the cells decreased significantly after doing so.<br />

The survival of the A549 cell line could also be significantly decreased by heavy<br />

ion irradiation. The mechanism of which seems to be a lack of repair, despite the<br />

activation of some of the component of the repair pathway.<br />

CHAPTER 5: REFERENCES<br />

[1] Rontgen, W. C. On a New Kind of Rays. Science 3:227-231; 1896.<br />

[2] Becquerel, H. Emission of New Radiations by Metallic Uranium. Compt. Ren.<br />

122:1086-1088; 1896.<br />

[3] Alpen, E. L. Radiation Biophysics. Prentice Hall, New Jersey; 1990.<br />

26


SYNOPSIS<br />

[4] Hall, E. J. Radiobiology for the radiobiologist. J.B. Lippincott Company,<br />

Philadelphia.; 1994.<br />

[5] Criswell, T.; Leskov, K.; Miyamoto, S.; Luo, G.; Boothman, D. A. Transcription<br />

factors activated in mammalian cells after clinically relevant doses of ionizing radiation.<br />

Oncogene 22:5813-5827; 2003.<br />

[6] Morgan, W. F. Non-targeted and delayed effects of exposure to ionizing radiation:<br />

I. Radiation-induced genomic instability and bystander effects in vitro. Radiat Res<br />

159:567-580; 2003.<br />

[7] Coutard, H. The Results and Methods of Treatment of Cancer by Radiation. Ann<br />

Surg 106:584-598; 1937.<br />

[8] Schmidt-Ullrich, R. K.; Dent, P.; Grant, S.; Mikkelsen, R. B.; Valerie, K. Signal<br />

transduction and cellular radiation responses. Radiat Res 153:245-257; 2000.<br />

[9] Prise, K. M.; Ahnstrom, G.; Belli, M.; Carlsson, J.; Frankenberg, D.; Kiefer, J.;<br />

Lobrich, M.; Michael, B. D.; Nygren, J.; Simone, G.; Stenerlow, B. A review of dsb<br />

induction data for varying quality radiations. Int J Radiat Biol 74:173-184; 1998.<br />

[10] Rasband, W. S.; Image, J. U. S. <strong>National</strong> Institutes of Health, Bethesda, USA.<br />

http://rsb.info.nih.gov/ij/. 1997-2006.<br />

[11] Maurya, D. K.; Salvi, V. P.; Nair, C. K. Radiation protection of DNA by ferulic<br />

acid under in vitro and in vivo conditions. Mol Cell Biochem 280:209-217; 2005.<br />

[12] Konca, K.; Lankoff, A.; Banasik, A.; Lisowska, H.; Kuszewski, T.; Gozdz, S.;<br />

Koza, Z.; Wojcik, A. A cross-platform public domain PC image-analysis program for the<br />

comet assay. Mutat Res 534:15-20; 2003.<br />

[13] Green, L. C.; Wagner, D. A.; Glogowski, J.; Skipper, P. L.; Wishnok, J. S.;<br />

Tannenbaum, S. R. Analysis of nitrate, nitrite, and [15N]nitrate in biological fluids. Anal<br />

Biochem 126:131-138; 1982.<br />

[14] Jeon, S. H.; Kang, M. G.; Kim, Y. H.; Jin, Y. H.; Lee, C.; Chung, H. Y.; Kwon,<br />

H.; Park, S. D.; Seong, R. H. A new mouse gene, SRG3, related to the SWI3 of<br />

Saccharomyces cerevisiae, is required for apoptosis induced by glucocorticoids in a<br />

thymoma cell line. J Exp Med 185:1827-1836; 1997.<br />

[15] Russell, J.; Wheldon, T. E.; Stanton, P. A radioresistant variant derived from a<br />

human neuroblastoma cell line is less prone to radiation-induced apoptosis. Cancer Res<br />

55:4915-4921; 1995.<br />

[16] Dahlberg, W. K.; Azzam, E. I.; Yu, Y.; Little, J. B. Response of human tumor<br />

cells of varying radiosensitivity and radiocurability to fractionated irradiation. Cancer<br />

Res 59:5365-5369; 1999.<br />

[17] Pearce, A. G.; Segura, T. M.; Rintala, A. C.; Rintala-Maki, N. D.; Lee, H. The<br />

generation and characterization of a radiation-resistant model system to study<br />

radioresistance in human breast cancer cells. Radiat Res 156:739-750; 2001.<br />

[18] Krengli, M.; Hug, E. B.; Adams, J. A.; Smith, A. R.; Tarbell, N. J.; Munzenrider,<br />

J. E. Proton radiation therapy for retinoblastoma: comparison of various intraocular<br />

tumor locations and beam arrangements. Int J Radiat Oncol Biol Phys 61:583-593; 2005.<br />

[19] St Clair, W. H.; Adams, J. A.; Bues, M.; Fullerton, B. C.; La Shell, S.; Kooy, H.<br />

M.; Loeffler, J. S.; Tarbell, N. J. Advantage of protons compared to conventional X-ray<br />

or IMRT in the treatment of a pediatric patient with medulloblastoma. Int J Radiat Oncol<br />

Biol Phys 58:727-734; 2004.<br />

[20] Chang, J. Y.; Zhang, X.; Wang, X.; Kang, Y.; Riley, B.; Bilton, S.; Mohan, R.;<br />

Komaki, R.; Cox, J. D. Significant reduction of normal tissue dose by proton<br />

radiotherapy compared with three-dimensional conformal or intensity-modulated<br />

27


SYNOPSIS<br />

radiation therapy in Stage I or Stage III non-small-cell lung cancer. Int J Radiat Oncol<br />

Biol Phys 65:1087-1096; 2006.<br />

[21] Fukumitsu, N.; Sugahara, S.; Nakayama, H.; Fukuda, K.; Mizumoto, M.; Abei,<br />

M.; Shoda, J.; Thono, E.; Tsuboi, K.; Tokuuye, K. A prospective study of<br />

hypofractionated proton beam therapy for patients with hepatocellular carcinoma. Int J<br />

Radiat Oncol Biol Phys 74:831-836; 2009.<br />

[22] Jones, B.; Dale, R. G.; Carabe-Fernandez, A. Charged particle therapy for cancer:<br />

the inheritance of the Cavendish scientists? Appl Radiat Isot 67:371-377; 2009.<br />

[23] Miralbell, R.; Crowell, C.; Suit, H. D. Potential improvement of three dimension<br />

treatment planning and proton therapy in the outcome of maxillary sinus cancer. Int J<br />

Radiat Oncol Biol Phys 22:305-310; 1992.<br />

[24] Ogata, T.; Teshima, T.; Kagawa, K.; Hishikawa, Y.; Takahashi, Y.; Kawaguchi,<br />

A.; Suzumoto, Y.; Nojima, K.; Furusawa, Y.; Matsuura, N. Particle irradiation suppresses<br />

metastatic potential of cancer cells. Cancer Res 65:113-120; 2005.<br />

[25] Schulz-Ertner, D.; Nikoghosyan, A.; Thilmann, C.; Haberer, T.; Jakel, O.; Karger,<br />

C.; Kraft, G.; Wannenmacher, M.; Debus, J. Results of carbon ion radiotherapy in 152<br />

patients. Int J Radiat Oncol Biol Phys 58:631-640; 2004.<br />

[26] Tsuji, H.; Ishikawa, H.; Yanagi, T.; Hirasawa, N.; Kamada, T.; Mizoe, J. E.;<br />

Kanai, T.; Tsujii, H.; Ohnishi, Y. Carbon-ion radiotherapy for locally advanced or<br />

unfavorably located choroidal melanoma: a Phase I/II dose-escalation study. Int J Radiat<br />

Oncol Biol Phys 67:857-862; 2007.<br />

[27] Miyamoto, T.; Yamamoto, N.; Nishimura, H.; Koto, M.; Tsujii, H.; Mizoe, J. E.;<br />

Kamada, T.; Kato, H.; Yamada, S.; Morita, S.; Yoshikawa, K.; Kandatsu, S.; Fujisawa, T.<br />

Carbon ion radiotherapy for stage I non-small cell lung cancer. Radiother Oncol 66:127-<br />

140; 2003.<br />

[28] Azzam, E. I.; de Toledo, S. M.; Gooding, T.; Little, J. B. Intercellular<br />

communication is involved in the bystander regulation of gene expression in human cells<br />

exposed to very low fluences of alpha particles. Radiat Res 150:497-504; 1998.<br />

[29] Azzam, E. I.; De Toledo, S. M.; Spitz, D. R.; Little, J. B. Oxidative metabolism<br />

modulates signal transduction and micronucleus formation in bystander cells from alphaparticle-irradiated<br />

normal human fibroblast cultures. Cancer Res 62:5436-5442; 2002.<br />

List of Publications:<br />

1. <strong>Ghosh</strong>, S.; Narang, H.; Sarma, A.; Krishna, M. DNA damage response signaling<br />

in lung adenocarcinoma A549 cells following gamma and carbon beam<br />

irradiation. Mutation Research/Fundamental and Molecular Mechanisms of<br />

Mutagenesis; 2011 (In Press doi:10.1016/j.mrfmmm.2011.07.015).<br />

2. <strong>Ghosh</strong>, S.; Narang, H.; Sarma, A.; Kaur, H.; Krishna, M. Activation of DNA<br />

damage response signaling in lung adenocarcinoma A549 cells following oxygen<br />

beam irradiation. Mutat Res 723:190-198; 2011.<br />

3. <strong>Ghosh</strong>, S.; Bhat, N. N.; Santra, S.; Thomas, R. G.; Gupta, S. K.; Choudhury, R.<br />

K.; Krishna, M. Low energy proton beam induces efficient cell killing in A549<br />

lung adenocarcinoma cells. Cancer Invest 28:615-622; 2010.<br />

4. <strong>Ghosh</strong>, S.; Maurya, D. K.; Krishna, M. Role of iNOS in bystander signaling<br />

between macrophages and lymphoma cells. Int J Radiat Oncol Biol Phys<br />

72:1567-1574; 2008.<br />

28


CHAPTER 1<br />

INTRODUCTION<br />

INTRODUCTION<br />

29


CHAPTER 1<br />

INTRODUCTION<br />

In physics, radiation describes a process in which energetic particles or waves travel<br />

through a medium or space. Radiation can be classified according to the effects it<br />

produces on matter, into ionizing and non-ionizing radiation. Ionizing radiation includes<br />

cosmic rays, X-rays and the radiation from radioactive materials. Non-ionizing radiation<br />

includes ultraviolet light, radiant heat and microwaves. The word radiation is commonly<br />

used in reference to ionizing radiation only (i.e., having sufficient energy to ionize an<br />

atom), but it may also refer to non-ionizing radiation (e.g., radio waves or visible light).<br />

The energy radiates (i.e., travels outward in straight lines in all directions) from its<br />

source. This geometry naturally leads to a system of measurements and physical units<br />

that are equally applicable to all types of radiation. Both ionizing and non-ionizing<br />

radiation can be harmful to organisms and the natural environment. All life on Earth has<br />

evolved in presence of this radiation. Figure 1.1 shows the percentage contribution of<br />

various sources of ionizing radiation to which human beings are exposed. Over 85<br />

percent of total exposure is from natural resources with about half coming from radon<br />

decay products in the home. Medical exposure of patients accounts for 14 percent of the<br />

total, whereas all other artificial sources – fallout, consumer products, occupational<br />

exposure, and discharges from nuclear industries account for less than 1 percent of the<br />

total value. Of all the exposures, medical exposure is of the main interest to scientists all<br />

over the world because of the ability of radiation to treat dreaded disease like cancer, for<br />

which, ironically it is a cause as well.<br />

30


CHAPTER 1<br />

INTRODUCTION<br />

15.8%<br />

9.56%<br />

Natural Radon<br />

Natural Cosmic<br />

Natural External<br />

Natural Internal<br />

Medical<br />

Nuclear<br />

12.9%<br />

19.8%<br />

0.461%<br />

41.5%<br />

Fig.1.1 Pie-chart showing the percentage of radiation exposure from various sources.<br />

[Data taken from UNSCEAR (United Nations Scientific community on the effects of<br />

Atomic Radiation report 2008].<br />

31


CHAPTER 1<br />

INTRODUCTION<br />

This thesis deals with radiation induced signaling in mammalian cells. This introductory<br />

chapter therefore provides an outline of our current knowledge in the same area.<br />

1.1 Ionizing Radiation:<br />

Whenever the energy of a particle or photon exceeds the ionizing potential of a molecule,<br />

a collision with the molecule might lead to its ionization. Therefore, ionizing radiation,<br />

upon interaction with matter, has an ability to remove electrons from atoms resulting in<br />

the formation of ions. The result of this interaction is the production of negatively<br />

charged free electrons and positively charged ions and hence the term “ionizing<br />

radiation”. In other words it can be described as radiation is considered to be ionizing, if<br />

it has a wavelength


CHAPTER 1<br />

INTRODUCTION<br />

far apart and therefore the resulting damage produced are scattered. Contrarily, the<br />

lesions produced by high LET radiations are clustered because the energy transfer events<br />

are very close together. Thus high LET radiation is more damaging because of the very<br />

high density of ionizing events and the subsequent high likelihood that the ionization<br />

event will fall in the volume of an important biomolecule [2].<br />

The ionizing event involves the ejection of an orbital electron from a molecule, producing<br />

a free electron and a positively charged ion. These ions are highly unstable and rapidly<br />

dissipate their energy and undergo chemical change. This results in the production of free<br />

radicals containing unpaired electrons. Free radicals are an extremely reactive species<br />

having a very short half life. They react very rapidly with other surrounding molecules<br />

and may lead to permanent damage of the affected molecule, or the energy may be<br />

transferred to another molecule. Injury due to radiation is caused mainly by ionization<br />

within the tissues of the body. When radiation interacts with a cell, ionizations and<br />

excitations are produced in either biological macromolecules or in the medium, in which<br />

the cellular organelles are suspended (predominantly water). Based on the site of<br />

interaction, the radiation-cellular interactions may be termed as direct or indirect.<br />

Direct effects occur when an ionizing particle interacts with a macromolecule in a cell<br />

(DNA, RNA, protein etc.). The resulting damage to the macromolecules leads to<br />

abnormalities, which initiate the events that lead to biological changes. Direct effects are<br />

predominant with high linear energy transfer (LET) radiations, such as neutrons, protons,<br />

α-particles and heavy nuclei.<br />

Indirect effects involve the interaction of ionizing radiation with the medium in which the<br />

molecules are suspended, especially water, to produce free radicals and reactive oxygen<br />

33


CHAPTER 1<br />

INTRODUCTION<br />

species (ROS), in presence of oxygen, that damage critical targets. This is known as<br />

indirect action of radiation. With low LET radiations such as X-rays and γ-rays, most of<br />

the damage induced in the biological systems is indirect and mediated by reactive oxygen<br />

species (ROS) generated by radiolytic products of water. These include oxygen radical<br />

(O •- 2 ), hydrogen radical (H • ), hydroxyl radical (OH • ), hydrogen peroxide (H 2 O 2 ),<br />

aqueous electron (e - aq) etc [3]. These ROS are known to cause damage to important<br />

macromolecules in cellular systems leading to a loss of function.<br />

The very obvious application of this was the use of ionizing radiation as a potent tool in<br />

cancer therapy.<br />

1.2 Radiation Therapy:<br />

Ionizing radiation has been an important part of cancer treatment for almost a century,<br />

being used in external-beam radiotherapy, brachytherapy, and targeted radionuclide<br />

therapy. Radiotherapy is based on the idea that exposure to a sufficient quantity (dose) of<br />

ionizing radiation kills. The goal is to deliver a measured dose of radiation to a defined<br />

volume with minimal damage to surrounding normal tissue, resulting in eradication of the<br />

tumor. The first clinical use of radiation for the treatment of tumors was recorded in<br />

1897, and during the past 50 years radiation biology has led to the development of ideas<br />

that form a mechanistic framework of the predicted pathways between energy deposition<br />

in a tumor cell and probability of cell survival [4]. Ionizing` radiation deposits energy<br />

that injures or destroys cells in the area being treated (the "target tissue") by damaging<br />

their genetic material, making it impossible for these cells to continue to grow. Although<br />

radiation damages both cancer cells and normal cells, proliferating cells are more<br />

radiosensitive and have a greater cell loss/turnover rate. Tumors are rapidly proliferating<br />

34


CHAPTER 1<br />

INTRODUCTION<br />

and are more likely to be irradiated at the radiosensitive phase of the cell cycle. Cells are<br />

most sensitive to ionizing radiation during M (mitosis) and G2 phases of the cell cycle<br />

and most resistant in late S phase (DNA synthesis). Radiotherapy may be used to treat<br />

localized solid tumors, such as cancers of the skin, tongue, larynx, brain, breast, or<br />

uterine cervix [5]. It can also be used to treat leukemia and lymphoma (cancers of the<br />

blood-forming cells and lymphatic system, respectively) [6, 7].<br />

Though use of radiation therapy is one of the major and widely used modality in<br />

treatment of cancer, there are several limiting factors that hamper the success of<br />

radiotherapy. These limiting factors are as follows:<br />

1. Limitation to use high radiation dose to maximize tumor regression.<br />

2. Detrimental effects of radiation doses on the normal tissues/organ falling within<br />

radiation field and surrounding the tumor region.<br />

3. Radioresistance of hypoxic cells in tumor tissues.<br />

4. Radioresistance of cell population in S-phase in tumor tissue<br />

5. Induced radioresistance in the tumor cells.<br />

To overcome some of the above, radiotherapy is given as fractionated doses ranging from<br />

2-4 Gy, per fraction. This is done primarily to enhance tumor cell killing over normal<br />

tissues. This is achieved because of hypoxic cell reoxygenation which renders them<br />

sensitive to radiation. The redistribution or reassortment of cells in the cell cycle is<br />

another reason behind fractionation of the radiotherapy dose. Dividing the radiation dose<br />

into multiple fractions allows cells to reassort to more sensitive phases of the cell cycle<br />

before the next treatment. The total dose and hence the period of treatment may vary<br />

depending on the type of tumor and radiation used [8]. Further in order to optimize the<br />

35


CHAPTER 1<br />

INTRODUCTION<br />

tumor cell killing, clinicians use hyper fractionation and hypofractionation regimens,<br />

depending on the type and radioresistance of the tumors and the nature of the surrounding<br />

normal tissues. However, as a result of the repeated doses of radiation delivered to the<br />

tumor cells, it results in an adaptive response in them, leading to the development of<br />

induced radioresistance, rendering the tumor refractory to subsequent doses of radiation.<br />

Several new approaches to radiation therapy are being evaluated to determine their<br />

effectiveness in treating cancer. One such investigational approach is particle beam<br />

radiation therapy. This type of therapy differs from photon radiotherapy in that it involves<br />

the use of fast-moving charged particles to treat localized cancers. This mode has two<br />

major advantages over conventional radiotherapy. First, because of the high LET, the<br />

damage produced is much more and hence the tumor cell eradication is greatly enhanced.<br />

The second advantage is that the energy deposition by the charged particles is highly<br />

localized in a particular depth in the tissue. This is called the Bragg Peak, which confers<br />

that advantage of selective, localized irradiation of the tumor tissue and the surrounding<br />

normal cells are spared.<br />

Though radiotherapy is a prime modality of cancer treatment, a major obstacle in its<br />

successful execution is the development of radioresistance in tumor cells following<br />

treatment. One of the factors contributing to the development of radioresistance is the<br />

activation of the signaling molecules following irradiation. Modulation of these activated<br />

signaling molecules with a view to overcome radioresistance could perhaps be an<br />

effective approach.<br />

36


CHAPTER 1<br />

INTRODUCTION<br />

1.3 DNA damage signaling following irradiation<br />

The pathways that signal DNA damage can be envisioned as transduction cascades in<br />

which the DNA lesion acts as the signal. Much of our understanding of these pathways is<br />

based on genetic studies of the yeasts Saccharomyces cerevisiae and<br />

Schizosaccharomyces pombe; these studies have identified a series of genes crucial for<br />

signaling. There are mammalian homologues of virtually all of these genes, indicating<br />

that DNA damage signaling is highly conserved.<br />

In all organisms examined, signaling involves proteins of the phosphatidylinositol 3-<br />

kinase-like (PIKK) family — ATR and ATM in humans, Mec1p and Tel1p in S.<br />

cerevisiae, and Rad3 and Tel1 in S. pombe. The prevailing view is that DNA damage<br />

somehow activates these kinases, which then amplify and channel the signal by activating<br />

downstream kinases (hChk1 and hChk2 in humans). These, in turn, phosphorylate target<br />

proteins such as p53, Cdc25A and Cdc25C [9-11] to delay the cell cycle, a process called<br />

the DNA damage checkpoint. In addition to checkpoint control, the downstream kinases<br />

also modulate DNA repair and trigger apoptosis [12, 13].<br />

1.3.1 Alarm Sensors<br />

The first response to radiation induced DNA damage is the activation of the alarm<br />

sensors. As the name indicates, these are proteins which detect the damage and set off the<br />

alarm signals, thereafter, the cell readies itself for subsequent action. Interestingly, it is<br />

the proteins involved in DNA repair, like (DNA-PK, ATR, ATM, BRCA-1, PARP etc.),<br />

which scan the genome, detect the damage and act as alarm sensors.<br />

37


CHAPTER 1<br />

INTRODUCTION<br />

1.3.1.1 DNA-PK: a DNA damage sensor for DNA non-homologous end-joining<br />

DNA-PK is a DSB repair enzyme composed of a catalytic subunit, DNA-PKcs that is<br />

part of the PIKK family [14, 15] and a targeting subunit, the Ku70–Ku80 heterodimer<br />

[16, 17]. DNA-PK was discovered more than a decade ago and has since been studied<br />

extensively at a biochemical level. DNA-PK is characterized by its DNA activated<br />

serine/threonine-directed protein kinase activity and can be considered as a paradigm for<br />

the biochemical functions of PIKK family kinases [15]. DNA-PKcs possesses an<br />

intrinsic, albeit weak, DNA-end binding activity that is greatly stimulated and stabilized<br />

by the Ku70–Ku80 heterodimer [17, 18]. Whether DNA-PK signals DSBs to the<br />

checkpoint machinery [19, 20] remains controversial, but the emerging consensus is that<br />

it does not [21-23]. DNA-PK may, however, be involved in signaling DNA damage to<br />

the apoptosis machinery [24]. Genetically, the main role of DNA-PK seems to be in<br />

promoting the rejoining of DSBs by NHEJ. Thus, cells and animals deficient in DNA-PK<br />

are defective in DSB repair and in V(D)J recombination, a process by which antibody<br />

and T-cell receptor diversity is generated, and are consequently radiosensitive and<br />

immunodeficient [15]. The kinase activity of DNA-PK is required for these processes,<br />

indicating that reversible protein phosphorylation is important and that DNA-PK plays a<br />

catalytic role [25]. However, the proteins that have to be phosphorylated by DNA-PK for<br />

it to exert its effects are unknown.<br />

1.3.1.2 ATR is the main sensor of stalled replication forks<br />

The ATR kinase appears to be the master activator of the replication stress response [26,<br />

27]. ATR is a member of a superfamily of serine-threonine kinases, which include ATM,<br />

DNA-PK, mTOR, TRRAP and SMG-1 [28, 29]. ATR is an essential kinase [30, 31] that<br />

38


CHAPTER 1<br />

INTRODUCTION<br />

associates with chromatin specifically during S-phase and it is thought that it plays a<br />

functional role during normal replication [32, 33]. Transient inhibition of ATR results in<br />

(i) inappropriate firing of replication origins [34]; (ii) fork stalling and (iii) replication<br />

induced chromosome breakage [35, 36]. Recently, some individuals with the Seckel<br />

syndrome were found to have very low levels of ATR kinase in their cells due to a splice<br />

mutation in the ATR gene [37]. These individuals share phenotypes with DNA damage<br />

response syndromes such as the Nijmegen breakage syndrome<br />

Since the intrinsic kinase activity of ATR does not change appreciably following DNA<br />

damage, it is thought that ATR activity is primarily regulated through its cellular<br />

localization. ATR is known to form distinct nuclear foci following blockage of<br />

replication [38] and it has been shown that the single strand-binding protein RPA and the<br />

ATR-interacting protein ATRIP are important for the recruitment of ATR to sites of<br />

blocked replication forks [26, 27]. It is thought that stretches of single-stranded DNA are<br />

formed in and around the stalled replication fork and that these sequences become coated<br />

with RPA proteins. ATR/ATRIP is then recruited to the site of RPA-coated DNA where<br />

ATR phosphorylates its substrates. Some studies have shown that ATR is capable of<br />

phosphorylating its targets even in cells depleted of RPA suggesting that in some<br />

situation RPA may not be required for ATR activity [39, 40].<br />

The Nijmegen breakage syndrome (NBS) has overlapping phenotypes with both ataxia<br />

telangiectasia (AT) and the ATR-Seckel syndrome [37, 41]. Following treatment with the<br />

replication blocking agent hydroxyurea (HU), Chk1 kinase and the tumor suppressor p53<br />

are phosphorylated in an ATR-dependent manner and cells that make it through S-phase<br />

will arrest in the G2 phase of the cell cycle [41]. In NBS1-defective cells, however, the<br />

39


CHAPTER 1<br />

INTRODUCTION<br />

activity of ATR signaling is diminished following replication blockage leading to blunted<br />

phosphorylation of Chk1 and p53 and attenuated cell cycle arrests. Thus, NBS1 may play<br />

some role regulating ATR activity following replication blockage [41]. One potential role<br />

of NBS1 in ATR signaling is to retain the ATR kinase at the sites of replication blockage<br />

as to obtain a stronger induction of phosphorylation of its substrates [41].<br />

The ATR kinase orchestrates DNA repair, apoptosis, S-phase and G2 arrest via the Chk1,<br />

p53 and the Fanconi pathways<br />

The outcomes of the activation of the ATR signaling pathway by replication stress is<br />

many fold . First, ATR induces an intra-S checkpoint via phosphorylation of the Chk1<br />

kinase, which targets the Cdc25A phosphatase for degradation [42, 43]. Loss of Cdc25A<br />

results in the inhibition of replication firing [44, 45]. Furthermore, the FANCD2 protein,<br />

which if defective causes Fanconi anemia (FA), becomes monoubiquitylated following<br />

replication stalling in an ATR-dependent manner by a complex made up by at least six<br />

different FA proteins [46-48]. This monoubuitylation of FANCD2 is critical for its<br />

interaction with the tumor suppressor BRCA1 protein, the relocalization of these proteins<br />

to sites of replication blockage and for the induction of S-phase arrest [46]. Second, the<br />

activation of the Chk1 pathway by ATR following replication stress inhibits cells from<br />

prematurely progressing into mitosis with unreplicated DNA [30, 35]. Third, ATR can<br />

stimulate DNA repair through the activation of the FA pathway [49] by activation of p53<br />

[50-53] or by Chk1-mediated activation of Rad51 [54]. The activation of p53 by ATR is<br />

both direct, by phosphorylation of the ser15 site of p53, and indirect, by an inhibitory<br />

phosphorylation of the MDM2 protein that normally negatively regulates p53 [55].<br />

Fourth, activation of p53 may lead to the induction of apoptosis that eliminates cells that<br />

40


CHAPTER 1<br />

INTRODUCTION<br />

are unable to resolve the blockage of replication. Taken together, the replication<br />

machinery could be considered as an effective scanning apparatus that when blocked by<br />

DNA damage will recruit ATR kinase to sites of RPA-coated single-stranded DNA. This<br />

recruitment will then induce phosphorylation of numerous ATR substrates leading to the<br />

activation of S-phase and G2 cell cycle checkpoints and induction of DNA repair and/or<br />

apoptosis.<br />

1.3.1.3 ATM: a DNA damage sensor for DNA homologous recombination repair and<br />

cell cycle arrest<br />

Ataxia-telangiectasia (A-T) is a rare human disease characterized by cerebellar<br />

degeneration, immune system defects and cancer predisposition [29, 56, 57]. The disease<br />

has been the subject of intense scientific scrutiny, particularly since the identification in<br />

1995 of the gene mutated in A-T, ATM [58, 59]. The ataxia-telangiectasia mutated<br />

protein (ATM) has emerged as a central player in the cellular response to ionizing<br />

radiation (IR), in which it plays a critical role in the activation of cell cycle checkpoints<br />

that lead to DNA damage-induced arrest at G1/S, S, and G2/M. ATM is a member of the<br />

phosphatidylinositol 3-kinaselike family of serine/threonine protein kinases (PIKKs).<br />

Other members of this protein family include ATM- and Rad3-related (ATR), DNAdependent<br />

protein kinase catalytic subunit (DNA-PKcs), mammalian target of rapamycin<br />

(mTOR), and ATX/hSMG-1 [29, 60]. The PIKKs represent a subclass of “atypical”<br />

protein kinases within the overall eukaryotic protein kinase family [61]. ATM, like other<br />

members of this protein kinase family, phosphorylates its substrates on serine or<br />

threonine that is followed by glutamine, i.e. an SQ or TQ motif [62, 63]. ATM shares<br />

several features with other members of the PIKK family, including a FAT domain<br />

41


CHAPTER 1<br />

INTRODUCTION<br />

(named due to amino acid conservation between the PIKK family members FRAP, ATR,<br />

and TRRAP), a phosphoinositide 3,4-kinase (PI3K) domain and a FAT carboxy-terminal<br />

(FAT-C) domain [64]. One function of the FAT domain may be to interact with the<br />

kinase domain to stabilize the carboxy-terminal region of the protein [65]. Bioinformatics<br />

analysis indicates that the amino-terminal portions of ATM, ATR, mTOR and, likely,<br />

DNA-PKcs are composed of multiple huntingtin, elongation factor 3, A subunit of protein<br />

phosphatase 2A and TOR1 (HEAT) repeats [66]. Each HEAT repeat is composed of two<br />

interacting, anti-parallel alpha helices linked by a flexible loop. The amino-terminal<br />

regions of the PIKK proteins, therefore, are predicted to consist of multiple, interacting<br />

HEAT repeats that form a massive, superhelical scaffolding matrix [66]. While the<br />

function of the HEAT repeat domain remains unknown, it is speculated that it could<br />

provide a surface for interactions with other proteins. The first indications of the overall<br />

shape and dimensions of the ATM protein have recently emerged. Single particle electron<br />

microscopic analysis reveals that ATM has a large “head” domain of approximately<br />

115Å by 75–140 Å, as well as a long “arm” structure that protrudes from the head region<br />

[67]. The overall shape of ATM is, thus, similar to that of the related protein, DNA-PKcs,<br />

at similar resolution [68]. Both proteins have the ability to interact with ends of doublestranded<br />

(ds) DNA[69-72], and this interaction may be facilitated by a conformational<br />

change that causes the arm region to wrap around the DNA [67]. From the low-resolution<br />

(30 Å) structure of ATM, it is speculated that the HEAT domain corresponds to the<br />

head structure and the carboxy-terminal kinase domain to the arm. However, elucidation<br />

of the precise molecular architecture of ATM will require a higher resolution structure.<br />

42


CHAPTER 1<br />

INTRODUCTION<br />

Perhaps one of the most important recent developments in the field has been<br />

insight into the mechanism of ATM activation in response to DNA damage. ATM is a<br />

predominantly nuclear protein, the levels of which do not change when cells are exposed<br />

to IR [73-76]. However, the protein kinase activity of ATM, as determined using<br />

immunoprecipitation (IP) kinase assays, increases two- to three-fold following cellular<br />

exposure to IR [75] or the radiomimetic neocarzinostatin (NCS) [73]. IR also induces<br />

increased incorporation of phosphate into ATM [65, 77]. DNA damage-induced<br />

phosphate incorporation was not observed in cells expressing kinase-dead ATM,<br />

suggesting that IR induces autophosphorylation of ATM. Indeed, serine 1981, an SQ site<br />

located in the amino-terminal region of the FAT domain, is rapidly phosphorylated in<br />

cells that have been exposed to even very low doses of IR [65]. ATM containing a serine<br />

to alanine mutation at amino acid 1981 failed to support IR-induced phosphorylation of<br />

p53 or cell cycle arrest, suggesting that serine 1981 phosphorylation is critical for the<br />

function of ATM [65]. trans-Phosphorylation and dissociation of ATM from an inactive<br />

dimeric (or higher order) form to an active monomeric form has, thus, emerged as one of<br />

the earliest events in the activation of ATM [65]. While phosphorylation of serine 1981 is<br />

certainly critical to the activation of ATM, it is also possible that ATM undergoes<br />

autophosphorylation at other sites important for the DNA damage response [77].<br />

Although these elegant experiments place ATM autophosphorylation as an<br />

important upstream event in the activation pathway, they do not address whether ATM is<br />

the primary sensor of DNA damage. If it were a direct sensor of DNA damage, ATM<br />

would be expected to interact directly with the damaged DNA. Addition of dsDNA does<br />

not stimulate ATM protein kinase activity in IP kinase assays [73], although addition of<br />

43


CHAPTER 1<br />

INTRODUCTION<br />

DNA does stimulate the activity of the purified ATM protein under some conditions in<br />

vitro [70, 78]. Purified ATM binds ends of dsDNA in vitro [70], but the affinity of this<br />

interaction is not known. In crude cell extracts, ATM binds preferentially to irradiated<br />

DNA cellulose resin compared to undamaged DNA cellulose suggesting that ATM<br />

interacts with, or is recruited to, damaged DNA [79]. DNA damage also causes a portion<br />

of the total ATM to associate with a chromatin fraction that is resistant to detergent<br />

extraction [80], and serine 1981-phosphorylated ATM localizes to nuclear foci in<br />

response to DNA damage [65]. Together, these studies suggest that ATM may interact<br />

with, or be recruited to, DNA-damaged chromatin. However, these studies, carried out in<br />

intact cells or crude extracts, do not address whether ATM interacts with damaged DNA<br />

directly or indirectly. The MRN complex (composed of Mre11, Rad50, and Nbs1 in<br />

humans, or xrs2 in yeast) fulfills many of the criteria for a DNA damage sensor [81, 82].<br />

Cells that are defective in Mre11 or Nbs1 have many features in common with A-T cells,<br />

including radiosensitivity and cell cycle checkpoint defects. Indeed, mutations in the<br />

human Mre11 gene are responsible for an A-T like disorder (ATLD), which clinically<br />

resembles A-T [83], suggesting that the MRN complex and ATM function in similar<br />

pathways in vivo. Indeed, several recent reports have placed the MRN complex as an<br />

upstream activator of ATM [84, 85]. Autophosphorylation of ATM on serine 1981 is<br />

defective in cells that are compromised for Nbs1 or Mre11 [84, 85], and phosphorylation<br />

of some downstream targets of ATM is partially dependent on a functional MRN<br />

complex [84]. Intriguingly, the nuclease activity of Mre11 is required for activation of<br />

ATM, suggesting that DNA double-strand breaks (DSBs) may require processing prior to<br />

activation of ATM [84]. Also, ATM-mediated phosphorylation of Nbs1 on serine 343<br />

44


CHAPTER 1<br />

INTRODUCTION<br />

occurs at DSBs, suggesting that the proteins co-localize at sites of DNA damage [86].<br />

Together, these studies suggest an attractive model in which (i) the MRN complex binds<br />

directly to damaged DNA, (ii) Mre11 processes the DNA ends, (iii) ATM is recruited and<br />

activated through autophosphorylation and monomerization, and (iv) kinase-active,<br />

monomeric ATM phosphorylates substrates present at, or recruited to, the DSB. Once<br />

activated, ATM may be released from the site of the DSB, enabling phosphorylation of<br />

more distal substrates, such as chromatin-bound histone H2AX. This might account for<br />

the rapid DNA damage-induced phosphorylation of histone H2AX that occurs over<br />

several megabases surrounding the site of a DSB [87]. Indeed, the downstream<br />

checkpoint protein kinase Chk2 is recruited only transiently to sites of DNA damage [86].<br />

It remains to be seen whether this is the only mechanism for activation of ATM by all<br />

DNA-damaging agents, or whether ATM can, under some circumstances, be activated by<br />

direct DNA binding, or by interaction with other DNA-binding partners. A very recent<br />

report suggests that the mediator of DNA damage checkpoint protein, MDC1 (also called<br />

NFBD1), which is known to interact with the MRN complex [88], is required for MRNdependent<br />

activation of ATM at high doses of IR, whereas the p53-binding protein<br />

53BP1 is required for activation of ATM at low doses of IR [89], suggesting that<br />

additional proteins may indeed regulate the activation of ATM. Another fascinating clue<br />

to emerge in recent years is that the BRCA1 protein is required for IR-induced<br />

phosphorylation of some ATM substrates, including p53, c-jun, Nbs1, and Chk2 [90].<br />

BRCA1 is also required for phosphorylation of p53 on serine 15 at later times after DNA<br />

damage in A-T cells, consistent with a requirement of BRCA1 for some ATR-mediated<br />

events [90]. In addition, BRCA1 is required for IR-induced phosphorylation of the<br />

45


CHAPTER 1<br />

INTRODUCTION<br />

structural maintenance of chromosomes protein, SMC1 [91], Chk1 [92], and, is itself a<br />

substrate of ATM[93]. However, BRCA1 is not required for the activation of ATM (as<br />

judged by ATM activity in IP kinase assays) [90]. DNA damage-induced phosphorylation<br />

of histone H2AX and human Rad17 (which recruits the Hus1–Rad1–Rad9 complex) also<br />

do not require the presence of BRCA1 [90]. Together, these studies suggest that BRCA1<br />

acts as a scaffolding protein that makes some, but not all, non-chromatin-associated<br />

substrates accessible to the activated ATM protein [90]. BRCA1 has also been identified<br />

as part of a larger protein complex, BRCA1-associated genome surveillance complex<br />

(BASC), which contains ATM, the mismatch repair proteins MSH2, MSH6, MLH1, the<br />

Bloom syndrome helicase (BLM), the MRN complex, and replication factor C (RFC)<br />

[94], but how BASC functions in the DNA damage response remains to be determined.<br />

BRCA1 contains two carboxy-terminal BRCT repeats, motifs that have been shown<br />

recently to bind phosphoproteins [95-97]. It is, therefore, possible that phosphorylation of<br />

BRCA1 could regulate its association with partner proteins, and thus, further regulate<br />

phosphorylation of BRCA1-associated proteins by ATM. Interestingly, BRCA1 has also<br />

been shown to interact with protein phosphatase 1α [98], providing another possible<br />

mechanism for regulation. Activation of ATM results in the phosphorylation of a plethora<br />

of downstream targets. Some of these proteins are direct substrates of ATM, for example,<br />

ATM likely directly phosphorylates serine 15 of p53 and serine 139 of histone H2AX in<br />

response to DNA damage. Others may be phosphorylated indirectly, through ATMmediated<br />

regulation of other protein kinases, such as Chk1, Chk2 [29, 60], IКB kinase<br />

(IKK) [99], LKB1 [100], c-Abl [101], and the cyclin-dependent protein kinases (cdk1<br />

and cdk2). It was previously thought that ATM signaled directly only to Chk2, while the<br />

46


CHAPTER 1<br />

INTRODUCTION<br />

related protein kinase, ATR, signaled to Chk1. Recent studies, however, have<br />

demonstrated that IR induces ATM-dependent phosphorylation of Chk1 on serine 317<br />

[102] and serine 345 [100], suggesting that ATM and ATR-dependent signaling pathways<br />

could overlap rather than exist in parallel as originally proposed. The most well<br />

characterized cellular response to activation of ATM is activation of G1/S, intra-S, and<br />

G2/M cell cycle checkpoints. However, ATM also plays important roles in other aspects<br />

of the DNA damage response. One of the primary functions of cell cycle arrest may be to<br />

allow the cell more time to repair DNA damage. Thus, activation of ATM-dependent cell<br />

cycle checkpoints would be expected to engage mechanisms for the repair of IR-induced<br />

DNA damage. A very recent study [103], suggests that one way in which this occurs is<br />

via ATM-dependent, Chk2-mediated phosphorylation of BRCA1 on serine 988, and<br />

upregulation of DSB repair via homologous recombination [103, 104]. It is also possible<br />

that ATM-dependent activation of c-Abl [101] and c-Abl-induced phosphorylation of<br />

Rad51 on tyrosine 315 [105], play a role in regulation of DSB repair. c-Abl also<br />

phosphorylates BRCA1 in an ATM-dependent manner [106], while BRCA1 interacts<br />

with BRCA2 which, in turn, interacts with Rad51 [107]. In addition, ATM also<br />

influences DNA damage-induced changes in gene expression through its effects on the<br />

IKK/NF-КB pathway [99] and possibly c-jun [90]. How ATM regulates DNA damageinduced<br />

apoptosis is still uncertain. One possibility is via p53-mediated upregulation of<br />

pro-apoptotic genes, such as Noxa and Puma [108]. Given the importance of ATM in the<br />

cellular response to DSBs, it is perhaps surprising that ATM is not an essential gene in<br />

mammalian cells. This likely reflects the fact that other members of the PIKK family,<br />

such as ATR, DNA-PKcs, or ATX, have similar substrate specificities and may be able to<br />

47


CHAPTER 1<br />

INTRODUCTION<br />

compensate for loss or dysfunction of ATM. However, the inter-relationship between the<br />

various PIKK family members remains to be clarified.<br />

Despite the abundance of physiological substrates of ATM, one of the most<br />

frequently used “read-outs” of ATM activity is the phosphorylation of p53 on serine 15.<br />

IR-induced phosphorylation of p53 is both attenuated and delayed in A-T cells [109].<br />

However, serine 15 phosphorylation of p53 is clearly evident in ATM-deficient cell lines<br />

at later times after exposure to IR, indicating that ATM is required for serine 15<br />

phosphorylation predominantly during the initial phase of the DNA damage response,<br />

and that other serine 15 specific protein kinases, such as ATR, can probably compensate<br />

for the absence of ATM at later times [109]. IR also induces phosphorylation of p53 at<br />

serines 6, 9, 20, 33, 46, 315, and 392, and ATM is required for efficient phosphorylation<br />

on serines 9, 20, and 46, as well as 15 [110]. Indeed, phosphorylation of p53 on serines<br />

20 and 46 is far more dependent on the presence of ATM than is phosphorylation on<br />

serine 15 [110]. Therefore, in studying the requirement for ATM in the cellular response<br />

to a given stressor, the analysis of other ATM-dependent p53 phosphorylation sites, in<br />

particular serines 20 and 46, may be informative. Although p53 is an important target of<br />

ATM, activation of ATM results in the phosphorylation of many other substrates and<br />

regulation of multiple cellular processes. Therefore, analysis of the ATM-dependent<br />

phosphorylation of one substrate cannot provide an accurate picture of the complexity of<br />

a given cellular response. A thorough understanding of the role of ATM will require a<br />

multifactorial approach in which the expression, activity, and phosphorylation of multiple<br />

substrates are analyzed both temporally and spatially. As discussed above, ATM is<br />

predominantly required for the initial response to DNA damage; therefore, time after<br />

48


CHAPTER 1<br />

INTRODUCTION<br />

damage must be an important experimental consideration when examining the potential<br />

role of ATM in a given response. Another important experimental consideration is the<br />

dose of a given DNA-damaging agent used, as alternative or redundant non-ATMdependent<br />

pathways may be engaged at high levels of damage or cellular stress [84, 89].<br />

1.3.2 Cell Cycle Regulation<br />

Once the alarm sensors have set off the alarm, the cell reacts to the damage by<br />

stopping cell cycle progression. This allows the cell to assess the damage and then<br />

accordingly either repair it or undergo apoptosis. The cell cycle is predominantly blocked<br />

at either G 1 -S transition (G 1 -S block) or at the G 2 -M transition (G 2 -M block). Under<br />

normal conditions such transitions are orchestrated by specific cyclins and cyclin<br />

dependent kinases. However, following damage to the DNA, the blocking of the cell<br />

cycle is achieved primarily by p53 and p21 proteins (Fig. 1.3A).<br />

p53<br />

Multicellular organisms elicit a complex response to IR, which is hard to separate from<br />

the fact that they have evolved to have p53 in checkpoint pathways. The checkpoint thus<br />

is not equated with ‘arrest for repair’ that was derived from the extensive studies in yeast.<br />

Instead, an additional cell fate of apoptosis is included in the possible outcomes of cell<br />

fate in higher eukaryotic systems in response to a variety of stress conditions. p53 plays<br />

an important role in the response of IR. It is stabilized in vitro and in vivo after IR. The<br />

level of p53 protein accumulation in response to IR primarily results from the intensity of<br />

DNA damage. Posttranslational modifications of p53 and its interacting proteins<br />

determine its stability and transactivation activity. This stabilization and activation is<br />

important for its roles in inducing growth arrest when reversible, which is coupled with<br />

49


CHAPTER 1<br />

INTRODUCTION<br />

DNA damage repair to protect cells from the IR damage. Apoptosis occurs in response to<br />

a high dose of IR when the damage is irreparable, in order to eliminate the damaged cells.<br />

The mechanism by which p53 mediates cell cycle arrest and apoptosis is largely mediated<br />

through its target genes, although some transcriptional-independent phenomena have<br />

been reported in other systems [111]. The dosage of IR predominantly contributes to the<br />

outcomes of cell fate, either survival or death. The role of p53 in DSB repair may involve<br />

p53 directly. The role(s) of p53 in the S phase arrest/delay are elusive.<br />

Fig. 1.3A: Regulation of Cell Cycle by p53 and p21. Following DNA damage, blocking<br />

of the cell cycle is orchestrated by p53 along with p21.<br />

50


CHAPTER 1<br />

INTRODUCTION<br />

At the present, much insight is being gained regarding the roles of p53 in DSB repair.<br />

Further studies on p53 functioning as the switch from reversible arrest to triggering<br />

apoptosis may offer a greater understanding of radioresistance and radiosensitivity [111].<br />

p21<br />

p21 Waf1/Cip1/Sdi1 is the first identified inhibitor of cyclin/cyclin-dependent kinase (CDK)<br />

complexes, which regulate transitions between different phases of the cell cycle. p21 was<br />

independently isolated as a CDK-binding protein and as a growth-inhibitory gene which<br />

is upregulated by wild-type p53 [112] or overexpressed in senescent fibroblasts [113].<br />

Although many other CDK inhibitors have since been discovered [114], p21 appears to<br />

be the only inhibitor capable of interacting with essentially all of the CDK complexes.<br />

The effects of p21 on CDKs, however, are not limited to simple binding and inhibition.<br />

For example, p21 binding, depending on its stoichiometry, may not only inhibit but also<br />

stimulate CDK4/6 complexes [115]. On the other hand, p21 affinity towards<br />

CDC2/CDK1 (the primary regulator of cellular entry into mitosis) is low, suggesting that<br />

p21 may not bind CDC2 under physiological conditions [116]. Nevertheless, p21<br />

expression efficiently inhibits CDC2 in vivo; this effect appears to be mediated at least in<br />

part through the interference with the activating phosphorylation of CDC2 at Thr 161 [117,<br />

118].<br />

The first transcriptional target of p53 identified was the Cdk inhibitor p21. That p53<br />

transactivation of p21 does not necessarily yield increased p21 protein levels is another<br />

important aspect of the study described by Jascur et al. [119]. DNA damage by ionising<br />

radiation results in stabilization of wild-type p53 and subsequent transcriptional<br />

51


CHAPTER 1<br />

INTRODUCTION<br />

activation of p21. However, p21 protein levels are not increased when WISp39 is<br />

downmodulated with siRNA or Hsp90 is inhibited with the drug geldanamycin following<br />

DNA damage. This exciting observation suggests that p21 levels are increased<br />

transcriptionally by p53 and at the level of protein stability by WISp39 and Hsp90<br />

suggesting that two events are required to stabilize p21. Counterintuitive to the role of<br />

p21 in cell cycle arrest, p21 has also been implicated as a positive regulator of cell<br />

survival. For example, loss of p21 sensitizes cells to undergo uncoordinated DNA<br />

replication and death induced by anticancer drugs [120].<br />

1.3.3 DNA Repair<br />

Following cell cycle arrest and triggering of the alarm sensors, many DNA repair<br />

pathways are activated to repair the damaged DNA. The five major DNA repair pathways<br />

are homologous recombination repair (HRR); non-homologous end joining (NHEJ);<br />

nucleotide excision repair (NER); base excision repair (BER); and mismatch repair<br />

(MMR)[121].<br />

1.3.3.1 DNA double strand break repair<br />

DNA double-strand breaks (DSBs) are a common form of DNA damage and DSB<br />

rejoining is a fundamental mechanism of genome protection. Breaks arise through direct<br />

action of ionizing radiation or some chemicals, and indirectly as a product of blocked<br />

replication forks. The ability to repair DSBs and to ensure that repair is performed with<br />

appropriate fidelity is a fundamental part of genome protection. The three known DSB<br />

repair pathways are outlined in Fig. 1.3B. The pathways are conserved between<br />

Saccharomyces cerevisiae and mammalian cells although the relative importance in each<br />

differs considerably. The prevailing view has been that homologous recombination is the<br />

52


CHAPTER 1<br />

INTRODUCTION<br />

predominant pathway of DSB repair in yeast whereas mammalian cells presented with<br />

similar substrates use non-homologous or illegitimate pathways. This still seems to be the<br />

consensus although there is mounting evidence that DSB repair via homologous<br />

recombination in mammalian cells is more significant than was previously thought [122,<br />

123]. In addition, there are indications that some proteins may participate in more than<br />

one of the three repair pathways.<br />

Fig. 1.3B Pathways of DSB repair. The termini of a DNA DSB introduced by ionising<br />

radiation or other means are bound either by the Ku heterodimer/DNA-PKcs complex or<br />

by hRad52. In the NHEJ rejoining pathway, repair is completed by DNA ligase IV and<br />

XRCC4. DNA strand invasion of the intact sister chromatid, facilitated by hRad51,<br />

initiates repair by homologous recombination. Resection and annealing of short regions<br />

of complementary sequence initiates repair by the SSA pathway in which ligation is<br />

preceded by the trimming of non-complementary single-stranded DNA tails.<br />

53


CHAPTER 1<br />

INTRODUCTION<br />

Non-homologous end rejoining<br />

Non-homologous end rejoining (NHEJ) has been considered the major pathway of DSB<br />

repair in mammalian cells (Fig. 1.3B). Repair is achieved without the need for extensive<br />

homology between the DNA ends to be joined. NHEJ processes the site-specific DSBs<br />

introduced during V(D)J (variable [division] joining) recombination and its importance is<br />

emphasised by the severe combined immune deficiency (SCID) and ionizing radiation<br />

sensitivity of mice with NHEJ defects. NHEJ involves a DNA end-binding heterodimer<br />

of the Ku70 and Ku80 proteins which activates the catalytic subunit (DNA-PKcs) of<br />

DNA-dependent protein kinase (DNA-PK) by stabilizing its interaction with DNA ends.<br />

This facilitates rejoining by a DNA ligase IV/Xrcc4 (X-ray cross-complementing 4)<br />

heterodimer [15]. The presence of constitutively high steady-state levels of DNA-PK is<br />

consistent with a role as a primary DNA damage recognition factor. The family of<br />

kinases to which DNA-PKcs belongs contains the important ATM (ataxia telangiectasia<br />

mutated) and ATR (ataxia telangiectasia Rad3-related) signalling proteins. All three<br />

proteins are serine/ threonine protein kinases which have some sequence similarity to<br />

phosphatidylinositol kinases. DNA-PK can phosphorylate p53 but the p53 response is<br />

apparently intact in DNA-PK-defective mouse cells which express wild-type p53 [124].<br />

These rules out DNA-PK as an essential requirement for activation of the p53 DNA<br />

damage response. Other plausible targets for DNA-PK include the single stranded<br />

binding protein RPA (replication protein A), the DNA ligase IV cofactor Xrcc4, Ku, and<br />

DNA-PKcs itself. It is noteworthy that, although the stimulation of DNA ligase IV<br />

activity by phosphorylated and unphosphorylated forms of Xrcc4 is similar,<br />

phosphorylation of Xrcc4 may prevent its direct association with DNA [125]. This could<br />

54


CHAPTER 1<br />

INTRODUCTION<br />

provide a means to regulate rejoining by modulating the interactions among Xrcc4, DNA<br />

and DNA ligase IV. Both Ku subunits and DNA-PKcs are essential for repair by NHEJ<br />

although defects in DNA-PKcs generally confer a milder phenotype than defects in Ku.<br />

DNA-PK activity is undetectable in Ku-defective cell lines, indicating that DNA binding<br />

by the Ku70:80 heterodimer is essential for its activation. The carboxy-terminal portion<br />

of Ku80 is also important for DNA-PK function and distinct regions are responsible for<br />

Ku70 interaction and DNA-PKcs activation. Deletion of the carboxy-terminal region of<br />

Chinese hamster Ku80 imparts a phenotype similar to DNA-PK deficiency even when<br />

high levels of DNA-PKcs are present[126]. The requirement for this carboxy-terminal<br />

region in kinase activation is consistent with the absence of an analogous Ku80 carboxyterminal<br />

tail in S. cerevisiae which also lack a homologous DNA-PKcs protein. The<br />

molecular architecture of DNA-PKcs suggests a structural role for the protein in the<br />

rejoining process. It has a potential DNA-binding groove and an enclosed cavity with<br />

three apertures through which single-stranded DNA could pass [68]. These findings are<br />

consistent with DNA-PKcs mediating the alignment of short stretches of single stranded<br />

DNA prior to ligation. This would be in addition to any role of DNA-PK in signaling.<br />

Repair by NHEJ is completed by the DNA ligase IV/Xrcc4 complex. Mice with targeted<br />

inactivation of either component exhibit embryonic lethality [127-129]. Consistent with<br />

their joint participation in the NHEJ pathway, the sensitivity of DNA ligase IV or Xrcc4<br />

knockout mouse cells to ionizing radiation, and their defects in V(D)J recombination<br />

mimic those of Ku null mice. As Ku and DNA-PKcs deficient mice are viable, however,<br />

there appears to be an additional requirement for DNA ligase IV and Xrcc4 during<br />

embryonic development. Reduced NHEJ activity may confer radiosensitivity in the<br />

55


CHAPTER 1<br />

INTRODUCTION<br />

absence of overt defects in immune function. Radiation sensitivity and a modest defect in<br />

DSB repair, but no significant impairment in V(D)J recombination, are associated with a<br />

mutated DNA ligase IV in cells from a radiation-sensitive leukaemia patient [130]. The<br />

absence of a detectable effect on V(D)J recombination indicates that an alternative DNArejoining<br />

activity might partially substitute for defective Xrcc4/DNA ligase IV in these<br />

cells. The occurrence of a DNA ligase IV defect in a leukaemia patient suggests that<br />

reduced NHEJ might be a factor in the incidence of this disease.<br />

Homologous recombination<br />

S. cerevisiae has provided a useful model for homology directed recombination (HR)<br />

processes. HR is performed by the RAD52 epistasis group of proteins which includes the<br />

products of RAD50–55, RAD57, and RAD59, MRE11 and XRS2 [131]. Although<br />

mammalian cells are considered to rely less on HR, they do perform mitotic<br />

recombination and preferentially repair DSBs by HR in late S and G2 phases of the cell<br />

cycle when an undamaged sister chromatid is available. Indeed, the DSB repair defect of<br />

murine scid (severe combined immunodeficiency) cells is only apparent in G1/early S<br />

phases [132]. The mammalian Rad51 protein is a homolog of the Escherichia coli RecA<br />

protein and is involved in both meiotic and mitotic recombination. The S. cerevisiae and<br />

human Rad51 proteins catalyse strand exchange in a reaction which is stimulated by<br />

Rad52 and RPA [133, 134]. Human Rad52 has a DNA double-strand end binding activity<br />

like Ku [135] and it probably co-operates with hRad51 in DSB repair but this is unlikely<br />

to be the only important function of Rad51. Targeted inactivation of mRad51 results in<br />

early embryonic lethality [136] whereas Rad52–/– mice are viable and fertile [137].<br />

Rad52–/– murine stem cells, although deficient in recombination, are not detectably<br />

56


CHAPTER 1<br />

INTRODUCTION<br />

hypersensitive to the presence of DSBs. Animals which are null for a third member of the<br />

RAD52 family, mRad54, are also viable and fertile but, in this case, their cells are<br />

hypersensitive to DSB-inducing agents [138]. Unraveling the biochemical functions of<br />

these mammalian RAD52 family members — and defining the essential functions of the<br />

Rad51 and Rad54 proteins — is clearly a high priority. At present, it is clear that they are<br />

not functionally identical to their yeast counterparts. Human Rad51 forms discrete foci in<br />

the nuclei of cells exposed to ionising radiation (but not UV) or chemicals. In rodents,<br />

DNA damage-induced mRad51 focus formation requires mRad54 [139]. As both<br />

mRad51 and mRad54 null cells are sensitive to ionising radiation, the direct participation<br />

of mRad54 and mRad51 in DSB repair is implied. Human cells are known to express two<br />

other Rad51-related proteins — Xrcc2 and Xrcc3 — both of which interact with Rad51<br />

and influence DSB repair by HR [140, 141]. Confusingly, although mRad54-null and<br />

Xrcc3-deficient cells are both sensitive to ionising radiation, only the latter are crosssensitive<br />

to UV light. An important recent development has been the realization that the<br />

products of the human breast cancer susceptibility genes BRCA1 and BRCA2 are involved<br />

in DNA repair. In fact, the carboxy-terminal region (BRCT domain) of these proteins<br />

defines a common motif among DNA-repair proteins [142]. Loss of functional Brca1<br />

results in sensitivity (albeit slight) to radiation and DNA-damaging chemicals [143, 144].<br />

A fraction of hRad51 colocalises with Brca1 and Brca2 in mitotic cells. Following DNA<br />

damage, the three proteins are relocated to structures which also contain PCNA<br />

(proliferating cell nuclear antigen) and may represent sites of active repair [145]. Both<br />

Brca2–/– and Brca1–/– mouse fibroblasts develop spontaneous chromosome aberrations<br />

consistent with the participation of these proteins in the repair of spontaneous DSBs<br />

57


CHAPTER 1<br />

INTRODUCTION<br />

[146]. Brca1 is phosphorylated by the ATM protein [144]. ATM, which is mutated in the<br />

radiationsensitive disorder ataxia-telangiectasia, is a candidate primary sensor of certain<br />

types of DNA damage, particularly DSBs induced by ionizing radiation. Brca1 in<br />

therefore a target for a key protein which signals the response to ionizing radiation.<br />

Brca2–/– cells are also sensitive to DNA-damaging agents, although, curiously, display a<br />

much more pronounced sensitivity to UV than to g-radiation. In addition, the massive<br />

sensitization to mitomycin C which accompanies loss of Xrcc2 or Xrcc3 is not seen in the<br />

Brca2–/– cells [146]. The likely involvement of Brca1 in HR, which is suggested by its<br />

association with mRad51, is supported by the observation that there is relatively more<br />

DSB repair by non-homologous pathways in Brca1–/– cells [123]. This is unlikely to be<br />

the full story, however, because Brca1 also displays a DNA-damage-dependent<br />

association with the hRad50/Mre11/Nbs1 complex [147]. This important tumor<br />

suppressor therefore appears to participate in the two pathways of DSB repair.<br />

Brca1 is also implicated in the removal of oxidized DNA bases from DNA. Damage is<br />

selectively removed from the transcribed DNA strand by a process known as<br />

transcription-coupled repair (TCR). Brca1-null cells are deficient in the TCR of DNA<br />

thymine glycol produced by ionizing radiation or hydrogen peroxide but perform TCR of<br />

UV induced DNA damage normally [143, 148]. It is unclear whether Brca1 participates<br />

directly in TCR or indirectly via a signaling role. The involvement of Brca1 suggests that<br />

the TCR of oxidized bases may represent a form of recombinational repair.<br />

Cellular signaling functions and DSB repair<br />

Many of the participants in DSB repair — including hMRE11, hRad50, DNA-PK, Ku<br />

and Xrcc4 — are phosphoproteins. Brca1 and Brca2 are both phosphorylated in a cell-<br />

58


CHAPTER 1<br />

INTRODUCTION<br />

cycle-specific fashion and the former is a target for the ATM protein kinase [144]. ATM<br />

and the tyrosine kinase c- Abl protein appear to be central to the control of DSB repair.<br />

The c-Abl protein is itself activated by phosphorylation. This appears to be achieved by<br />

its interaction with the ATM protein [149]. Human Rad51 is phosphorylated by c-Abl,<br />

possibly via the formation of a tripartite complex with the ATM protein [105]]. ATM/c-<br />

Abl-dependent phosphorylation promotes the interaction between hRad51 and hRad52,<br />

suggesting that the HR recombination repair pathway is subject, at least in part, to fine<br />

control by these interactions. c-Abl is also implicated in activation of c-Jun aminoterminal<br />

kinase. Significantly, this activation is impaired in Nbs1 (and hMre11) as well as<br />

ATM-defective human cells. The ATM protein is implicated in the p53-related DNA<br />

damage response. p53 plays a crucial role in determining the fate of cells in which DSBs<br />

persist. In particular, the embryonic lethality of both the Rad51 null [136] and Brca1 null<br />

[148] animals is attenuated in a p53–/– background and some double knockout animals<br />

survive to full term. This suggests that the massive failure of cellular proliferation, which<br />

is a feature of Rad51–/– and Brca1–/– embryos, is to some degree a direct consequence<br />

of an active p53. Put another way, cells which fail to repair DSBs by the Rad51- and<br />

Brca1-dependent HR pathways die in a p53-dependent fashion. Promoting the deletion of<br />

cells which might otherwise perform mutation-prone DSB repair may be a facet of p53’s<br />

role as ‘guardian of the genome’.<br />

1.3.3.2Single strand DNA repair<br />

Nucleotide excision repair (NER) repairs DNA with helix-distorting damages, including<br />

the damages of cyclobutane pyrimidine dimers and 6–4 photoproducts produced by UV<br />

light, and adducts produced by the chemotherapeutic agents cisplatin and 4-<br />

59


CHAPTER 1<br />

INTRODUCTION<br />

nitroquinoline oxide [150, 151]. About 30 polypeptides are involved in NER, and the<br />

NER process has been reconstituted with purified components [150].<br />

Key steps of NER include: (i) recognition of a DNA defect; (ii) recruitment of a repair<br />

complex; (iii) preparation of the DNA for repair through action of helicases; (iv) incision<br />

of the damaged strand on each side of the damage, with release of the damage in a singlestrand<br />

fragment about 24–32 nucleotides long; (v) filling in of the gap by repair<br />

synthesis; (vi) ligation to form the final phosphodiester bond [151, 152].<br />

Base excision repair (BER) is a major DNA repair pathway protecting mammalian cells<br />

against single-base DNA damage caused by methylating and oxidizing agents, other<br />

genotoxicants, and a large number (about 10,000 per cell per day) of spontaneous<br />

depurinations [153]. BER is mediated through at least two subpathways, one involving<br />

single nucleotide BER and the other involving longer patch BER of 2–15 nucleotides.<br />

A highly conserved set of MMR proteins in humans is primarily responsible for the postreplication<br />

correction of nucleotide mispairs and extra-helical loops. Mutational defects<br />

in MMR genes in humans give rise to a mutator phenotype, microsatellite instability, and<br />

a predisposition to cancer. Mouse embryonic fibroblast and human epithelial cell lines<br />

lacking the MMR protein, MLH1, are more resistant than wild-type cells to two inducers<br />

of oxidative stress, hydrogen peroxide and tert-butyl hydroperoxide [154]. Analysis of<br />

this resistance indicates that it results from a defect in apoptosis, as the consequence of a<br />

requirement for wild-type MLH1 in the transduction of apoptotic signals by a<br />

mitochondrial pathway, although the details of this pathway are currently unknown.<br />

60


CHAPTER 1<br />

INTRODUCTION<br />

1.4 Overview of Radiation induced cell signaling<br />

Exposure of cells to ionizing radiation results in complex cellular responses resulting in<br />

cell death and altered proliferation states. The underlying cytotoxic, cytoprotective and<br />

cellular stress responses to radiation are mediated by existing signaling pathways,<br />

activation of which may be amplified by intrinsic cellular radical production systems.<br />

These signaling responses include the activation of plasma membrane receptors, the<br />

stimulation of cytoplasmic protein kinases, transcriptional activation, and altered cell<br />

cycle regulation. There is increasing evidence for the functional links between cellular<br />

signal transduction responses and DNA damage recognition and repair, cell survival, or<br />

cell death through apoptosis or reproductive mechanisms.<br />

Recent radiobiological studies have demonstrated that the exposure of<br />

mammalian cells to ionizing radiation over a wide dose range results in activation of<br />

existing cellular response pathways. These pathways, dominantly involving protein<br />

kinases, mediate the cytoprotective and cytotoxic responses of cell survival and cell<br />

death, respectively. Cytoprotective responses involve pathways of the mitogen activated<br />

protein kinase (MAPK) and phosphatidyl inositol- 3-phosphate kinase (PI3 kinase)<br />

cascades which activate the machinery of biosynthesis and may stimulate cell<br />

proliferation if radiation-induced damage is successfully repaired. The most direct<br />

consequence of cytotoxic or stress responses is thought to involve Jun N-terminal kinase<br />

(JNK, now known as MAPK8) and results in apoptosis and/or other forms of cell death.<br />

Although general statements may be premature, current data suggest that cells of the<br />

hematopoietic lineage and fibroblasts or other normal cells are substantially more prone<br />

to undergo radiation-induced apoptosis than many carcinoma cells.<br />

61


CHAPTER 1<br />

INTRODUCTION<br />

The goal of this chapter is to describe the complexity of the responses of cells to<br />

exposure to ionizing radiation. Links between the mechanisms of various sensors of<br />

radiation effects and the activation of major cellular response pathways will be<br />

emphasized where possible (Fig. 1.4). The response networks include cytokines and<br />

plasma membrane receptors, effector protein kinases, and phosphatases in the cytoplasm<br />

or at the interfaces of plasma membrane and nucleus. The extent of radiation-induced<br />

changes in proteins, lipids and nucleic acids, the latter recognized by defined proteins as<br />

DNA damage, is likely to determine the relative balance between plasma membrane<br />

events, mitochondrial reactions, and nuclear responses<br />

EFFECT OF IONIZING RADIATION<br />

ON CELLULAR RESPONSE SYSTEM<br />

RADIATION SENSOR/ROS, RNS AMPLIFIER<br />

SIGNAL TRANSDUCTION<br />

CELL CYCLE<br />

CONTROL<br />

CYTOPROTECTIVE<br />

SURVIVAL<br />

CYTOTOXICITY<br />

DEATH: APOPTOSIS/<br />

REPRODUCTIVE DEATH<br />

FIG. 1.4 Effects of ionizing radiation on cellular response systems. Radiation effects are<br />

mediated through the interaction of radicals and reactive oxygen and nitrogen species<br />

(ROS/RNS), with proteins, lipids and nucleic acids; these radicals may be generated from primary<br />

ionization events or through secondary amplification systems. Biological molecules that are<br />

modified by radicals and generate or transmit intracellular signals are components of existing<br />

cellular signal transduction pathways. Effectors of these systems are linked to cell cycle<br />

regulation and DNA repair, which determine the ultimate fate of cells exposed to radiation.<br />

62


CHAPTER 1<br />

INTRODUCTION<br />

1.4.1 Radiation induced radicals:<br />

In studies of the responses of cells to radiation, it is unclear how a few ionization events,<br />

2000 per gray per cell, can induce the rapid and robust activation of diverse signal<br />

transduction pathways [155]. Recent evidence suggests that cells are endowed with<br />

cytoplasmic amplification mechanisms involving reactive oxygen (ROS) and nitrogen<br />

(RNS) species and that these systems are responsive to relatively low radiation doses.<br />

The primary radical generated as a consequence of initial ionization events is the<br />

. OH, which is short-lived and only diffuses about 4 nm before reacting [156]. Of the<br />

secondary ROS, O - 2 and H 2 O 2 , the latter can react via Fenton chemistry with cellular<br />

metal ions to produce additional . OH. Since the amounts of secondary ROS generated<br />

after irradiation are considerably lower than those produced by normal cell metabolism, it<br />

is unlikely that they contribute significantly to a potential amplification process [155].<br />

Less information is available about RNS that may also be produced after irradiation and<br />

may be a consequence of radiation-induced stimulation of nitric oxide synthase activity in<br />

cells enriched in this enzyme [157]. Reaction of nitric oxide with O 2<br />

-<br />

results in the<br />

formation of peroxynitrite, a membrane-permeant and relatively stable RNS. When<br />

protonated, peroxynitrite isomerizes to trans-peroxynitrous acid, which can cause protein<br />

and DNA damage similar to . OH [158].<br />

Cellular Amplification of Radiation-Induced Generation of ROS/RNS<br />

Indirect evidence for an extranuclear radical amplification mechanism has come from<br />

studies with high-LET particles. For example, bone marrow progenitor cells exposed to<br />

neutrons show both an enhanced ability to oxidize a fluorescent probe for ROS/RNS and<br />

increased 8-hydroxy- 2-deoxyguanosine levels indicative of oxidative DNA base damage<br />

63


CHAPTER 1<br />

INTRODUCTION<br />

[159]. Direct evidence for cytoplasmic ionization events affecting nuclear processes has<br />

come from studies in which individual cells were irradiated with a microbeam of particles<br />

that permitted selective irradiation of the cytoplasm or nucleus. Irradiation of the<br />

cytoplasm was shown to be mutagenic but not cytotoxic [160]. The major class of<br />

mutations was similar to that generated spontaneously and thought to arise from DNA<br />

damage due to endogenous production of ROS/RNS [161] and differed from that induced<br />

after irradiation of the nucleus.<br />

Studies with fluorescent dyes demonstrated generation of ROS/RNS in cells<br />

within 15 min after irradiation with less than one a particle per cell [162]. The<br />

intracellular production of ROS/RNS was 50-fold greater than the extracellular<br />

production. The intracellular source was inhibited by diphenyliodonium, an inhibitor of<br />

flavoproteins, suggesting the involvement of a plasma membrane NADPH oxidase.<br />

However, flavoproteins are also localized to the mitochondria and endoplasmic<br />

reticulum, possible alternative cellular organelles that generate ROS/RNS. Enhanced<br />

cellular production of ROS/RNS has also been observed after exposure to low-LET<br />

radiation [163]. Cobalt- 60 γ-irradiation, at 1–4 Gy, stimulated production of ROS/RNS<br />

in HepG2 cells and depletion of GSH, which radiosensitized cells due to enhanced radical<br />

generation. These studies do not discriminate between possible signal-amplifying<br />

mechanisms or an irreversible mitochondrial permeability transition as a late<br />

consequence of radiation exposure.<br />

Consequences of Radiation-Induced Generation of ROS/RNS<br />

Some ROS/RNS, e.g. H 2 O 2 , nitric oxide and peroxynitrite, are membrane-permeant and<br />

are sufficiently stable to diffuse significant distances within cells, facilitating their<br />

64


CHAPTER 1<br />

INTRODUCTION<br />

interaction with biological molecules, including proteins, lipids and DNA. Of particular<br />

interest for regulation of cytoplasmic signal transduction pathways are proteins SH<br />

residues. The interconversion of oxidized and reduced Cys represents a reversible<br />

controlling element in the regulation of protein functions [164]. For example, H 2 O 2 or<br />

more reactive ROS have been shown to reversibly oxidize a Cys to a cysteine sulfenic<br />

acid in the catalytic site of Tyr phosphatase 1B, resulting in transient inhibition of<br />

phosphatase activity. A possible consequence of this is the enhanced phosphorylation and<br />

activation of target proteins, such as EGFR (also known as ErbB1) [165, 166]. Similarly,<br />

nitric oxide stimulates guanine nucleotide exchange on RAS by S-nitrosylation of a<br />

critical Cys [167].<br />

Irradiation of cells induces lipid peroxidation and changes in membrane structure<br />

in intact cells [168]. How these structural changes modulate membrane function, e.g.<br />

activation of EGFR, and the cellular response to radiation currently are not known.<br />

However, they provide potential mechanisms, through free radical propagation or<br />

changes in membrane fluidity, for the amplification of localized perturbations of the<br />

membrane structure to signals throughout the cell.<br />

1.4.2 Cytokines and Plasma Membrane Receptors<br />

Ionizing radiation activates cytokine receptors, such as EGFR [169] and tumor necrosis<br />

factor receptor (TNFRSF1A, formerly known as TNFR)[170], with consequences<br />

currently indistinguishable from the effects of the physiological cytokine/growth factor<br />

ligands. The downstream effects of receptor activation are the initiation and amplification<br />

of signals along growth and stress pathways, involving MAPK, PI3 kinase and MAPK8,<br />

resulting in modulation of the proliferation and survival states of cells [170, 171].<br />

65


CHAPTER 1<br />

INTRODUCTION<br />

Cytoprotective Response Cascade<br />

The term cytoprotective describes the summation of responses which favor cell survival<br />

and may span from enhanced repair functions to stimulation of proliferation. These<br />

responses may be initiated at the level of the plasma membrane through activation of<br />

ERBB receptor tyrosine kinases and other related molecules [172]. Several lines of<br />

investigation support that the loss of EGFR/ERBB2 function radiosensitizes tumor cells<br />

[172-174].<br />

The radiation-induced activation of EGFR has been linked mechanistically to<br />

enhanced signaling through critical components of the MAPK and MAPK8 pathways<br />

[172, 175]. These links are based on inhibition of the function of EGFR by over<br />

expression of dominant-negative mutants [173, 176] or the use of specific<br />

pharmacological inhibitors, such as the tyrphostin AG1478 [172]. The disruption of<br />

function of ERBB2 may exert similar effects. The radiation-induced activation of EGFR<br />

at doses between 1 and 5 Gy [169] results in the activation of several immediate<br />

downstream effectors. Tyr phosphorylation of EGFR at positions Tyr1068, Tyr1148 and<br />

Tyr1173 (30) permits interaction with adapter proteins GRB2 and SHC, which in turn act<br />

through factors to stimulate GTP for GDP exchange in RAS. Radiation-induced<br />

activation of EGFR also results in the stimulation of PLCγ with production of<br />

diacylglyceride and inositol trisphosphate (IP3)[172, 177]. Diacylglyceride activates<br />

protein kinase C (PRKC) and IP3 induces release of Ca 2+ from intracellular stores [172,<br />

178, 179]. RAS-GTP recruits RAF1 to the plasma membrane; this activation is dependent<br />

on the release of Ca 2+ from intracellular stores [177, 178, 180]. RAF1 primarily signals<br />

along the MAPK cascade involving MAP2K1/2 and p90S6 kinase [180]. MAPK and<br />

66


CHAPTER 1<br />

INTRODUCTION<br />

p90S6 kinase regulate diverse targets in cells including transcription factors [181, 182].<br />

RAF1 has also been shown to phosphorylate the antiapoptosis protein BCL2, a<br />

modification that may counteract the anti-apoptosis function of BCL2 [183]. The<br />

importance of RAS lies in its potential cross-communication into the MAPK8 pathway<br />

[173], and the critical role of MAPK as downstream effector of EGFR is demonstrated by<br />

the finding that inhibition of MAPK by PD98059 [173, 184, 185] disrupts the<br />

cytoprotective response against radiation. Also in line with these conclusions are the<br />

findings that both the inhibition of RAS function through inhibition of prenylation [186]<br />

and the down-regulation of RAF1 with antisense-RAF oligonucleotides [187] result in<br />

radiosensitization of human tumor cells.<br />

Increased signaling through the MAPK pathway is cytoprotective after exposure<br />

to radiation, although the precise mechanism(s) by which this occurs is unclear [172, 175,<br />

184, 185]. There is new evidence that the regulatory functions of MAPK on cell<br />

proliferation compared to differentiation vary with cell type and depend on the magnitude<br />

and duration of MAPK activation [188-191]. A short activation of MAPK correlated with<br />

increased proliferation, potentially through coordinated expression induction of cyclin D1<br />

and the cyclin-dependent kinase (CDK) inhibitor CDKN1A (formerly known as p21)<br />

[188, 190, 191]. In contrast, prolonged stimulation of MAPK activity was linked to<br />

decreased DNA synthesis, potentially through super-induction of CDKN1A [188, 189].<br />

This establishes MAPK as an important target for both radiation- and growth factorinduced<br />

activation of EGFR and transient increases in CDKN1A protein levels [184,<br />

192]. High levels of CDKN1A expression would potentially lead to growth arrest at the<br />

G1/S- and G2/ M-phase boundaries, as has been reported for cells exposed to radiation<br />

67


CHAPTER 1<br />

INTRODUCTION<br />

[193]. Since the activation of EGFR and MAPK occurs in cells expressing wild-type or<br />

mutant TP53 (formerly known as p53), these responses are not necessarily independent<br />

of TP53. Collectively, these studies provide strong evidence that radiation-induced<br />

signaling of MAPK through CDKN1A modulates cell cycle progression and cell<br />

radiosensitivity.<br />

Irradiated cells, in which the activation of MAPK is inhibited, remain arrested in<br />

the G2/M phase of the cell cycle [185]. This arrest is associated with increased Tyr15<br />

phosphorylation of CDK1 and its reduced activity [194, 195]. The prolonged arrest<br />

correlates closely with the previously observed MAPK-dependent enhancement of<br />

radiation-induced apoptosis. Removal of MAP2K1/2 inhibitors or caffeine treatment of<br />

cells 6 h after irradiation abrogates the effects of MAPK inhibition on radiation-induced<br />

G2/Mphase arrest and potentiates apoptosis [185].<br />

Activation of ERBB receptor tyrosine kinases can also enhance PI3 kinase<br />

activity. This kinase plays a critical role in protecting cells from apoptosis during growth<br />

factor deprivation [196], and may fulfill an important cytoprotective function after<br />

irradiation. Downstream of PI3 kinase, an IP3-activated protein kinase termed 3-<br />

phosphoinositide- dependent protein kinase (PDK1/2) has been identified [197]. PDK1/2<br />

phosphorylates and activates the protein kinase AKT. Targets of AKT include glycogen<br />

synthase kinase 3 (GSK3A) and p70S6 kinase [196] regulating transcription factor<br />

function and protein synthesis, respectively.<br />

Cytotoxic Response Pathway<br />

The cytotoxic pathway summarizes responses of cells to a variety of stresses, including<br />

radiation. Two forms of the tumor necrosis factor-a receptor have been characterized as<br />

68


CHAPTER 1<br />

INTRODUCTION<br />

CD4 (formerly known as p55) and TNFRSF1B (formerly known as p75); only CD4<br />

contains a domain which can directly signal the caspase cascade that leads to apoptosis<br />

[198]. Both receptors can activate the lipid-cleaving enzyme acid sphingomyelinase<br />

(ASMase) with the production of ceramide. Ceramide, through the family of GTPbinding<br />

molecules RHO/RAC and RAS, leads to the activation of downstream signaling<br />

molecules MEKK1/2, which are analogous to RAF1 in the MAPK pathway. Active<br />

MEKK1 in turn can phosphorylate and activate MAP2K4/7 and MAP2K3/6. These<br />

molecules share considerable sequence similarity to MAP2K in the MAPK pathway.<br />

MAP2K4/7 [199] signals to MAPK8, which phosphorylates the transcription factor JUN,<br />

and has been shown to facilitate apoptosis in certain cell types. For example, in several<br />

studies using cells of hematopoietic lineage, enhanced MAPK8 signaling has been<br />

closely associated with apoptosis in response to cytotoxic stimuli [200-202]. In other<br />

nonhematopoietic cells, such as carcinoma cells, increased MAPK8 signaling can result<br />

in increased proliferation and protection from radiation and other cytotoxic stresses [203,<br />

204]. Thus the precise role of MAPK8 activation in response to exposure of cells to<br />

radiation may vary substantially with the cell type analyzed. Alternatively, MAP2K3/6<br />

may signal along a rescue pathway involving MAPK1 (formerly known as p38)<br />

reactivating kinase, which facilitates phosphorylation of heat-shock proteins and cell<br />

survival.<br />

Although much research has been done on low LET induced signaling, there have been<br />

few sporadic and diverse studies with high LET radiation.<br />

69


CHAPTER 1<br />

INTRODUCTION<br />

1.5 Charged particle irradiation induced Signaling<br />

Apart from X-rays, biological effects from particulate radiation are of immense interest<br />

due to their use in treatment of certain solid tumors and also due to their abundance in<br />

space. High LET radiations, such as heavy ions or neutrons, have an increased biological<br />

effectiveness compared to X-rays for gene mutation, genomic instability and<br />

carcinogenesis. The amount of induction of DSBs is weakly dependent on LET of the<br />

radiation [205]. However, the degree of lesion complexity increases with increasing LET<br />

[206]. There has been a consensus about the clustering of damage in case of high LET<br />

radiation along individual radiation tracks crossing the DNA<br />

The extensive studies using synthetic clustered DNA lesions have largely contributed to<br />

our understanding of the biological consequences of clustered lesions in cells or tissues.<br />

Clustered damage sites are of large diversity, they are processed differently depending on<br />

the nature of the base modification, the inter lesion spacing, the presence or not of strand<br />

breaks. The expected biological consequences range from point mutations and loss of<br />

genetic material to cellular lethality due to repair impairment and lesion or repairintermediate<br />

persistency (Fig. 1.5) In fact, the deleterious effect of repair intermediates<br />

may be amplified at replication and cause cell killing. Tandem lesions have been isolated<br />

and have been shown to be poorly repaired or by-passed by DNA polymerases, and may<br />

be lethal lesions. Bistranded AP sites are more likely to result in DSBs, the outcome of<br />

which will depend on the presence of damage bases around the DSB, and subsequent<br />

deletions have been observed in human cells. The repair of bistranded oxidized bases is<br />

mostly impaired, causing a point mutation at the unrepaired damaged base. DSBs are<br />

unlikely to be formed by BER at these clusters, but may occur at replication forks if the<br />

70


CHAPTER 1<br />

INTRODUCTION<br />

unrepaired damage is a block to replication as is the case for thymidine glycol or an AP<br />

site. Homologous recombination or single strand annealing will overcome the collapse of<br />

the replication fork and may produce deletions. Importantly, the cell will survive with<br />

mutations or loss of genetic material. The more complex the non-DSB cluster (i.e. several<br />

oxidized base damages or SSB/gap), the more complex the subsequent mutations. The<br />

presence of two or more base substitutions or ±1 insertions/deletions can be considered<br />

the signature of complex non-DSB clusters. The repair of complex DSBs, either formed<br />

by the radiation track or originating from “repair” of non-DSB clusters, is likely<br />

compromised and very slow, potentially forming large deletions. Lethal events may be<br />

expected from such lesions, even though this has not yet been demonstrated.<br />

Fig.1.5. Biological consequences of clustered DNA damage in eukaryotes. Clustered damage in<br />

the form of non-DSB bistranded lesions, tandem lesions or complex DSBs can consist of AP sites<br />

(A), SSBs or base damage (O is 8-oxoG, fU is 5-formyluracil, hU is 5-hydroxyuracil, and FA is<br />

formylamine). Two opposing AP sites can be converted to a DSB and hence complex DSBs can<br />

be formed either directly by the DNA damaging agent or by abortive BER. Complex DSBs can<br />

decrease the efficiency of NHEJ and be inaccurately repaired. Lesions of high complexity can<br />

result in mutagenesis: DSBs are not formed due to inhibition of repair enzymes by near-by<br />

damage. Tandem lesions can block DNA replication, but can also be mutagenic as found for<br />

bistranded lesions consisting of base damage. Partial processing of these latter two types of<br />

lesions could also result in persistent SSBs. Clustered lesions can therefore be mutagenic, result<br />

in DSBs and inaccurate repair, or block replication, and could be cytotoxic to the eukaryote cell.<br />

71


CHAPTER 1<br />

INTRODUCTION<br />

Early observations showed that densely ionizing radiation like high LET radiation was<br />

more cytotoxic than sparsely ionizing radiation and it was anticipated that a greater<br />

proportion of non-repairable strand breaks originating from clustered lesions was<br />

responsible for the increased biological effectiveness of densely ionizing radiation.<br />

Larger proportions of DSBs remain unrejoined after exposure to high LET radiation than<br />

after exposure to sparsely ionizing radiation, and chromosomal damage is more severe<br />

and complex. The frequency of chromosome breaks and of complex rearrangements<br />

increases up to a LET of 100–150 keV/µm and seems to plateau at higher LETs. In most<br />

cellular systems examined but not all, high LET radiation is observed to generate larger<br />

deletions (over Mbp size). A high frequency of complex deletion events and complex<br />

rearrangements at deletion junctions have been observed with high LET radiation and<br />

have not been reported for low LET radiation [207-209]. Notably, an unusually high<br />

proportion of radon-induced mutants exhibited two or more base substitutions and<br />

insertions/deletions within 3–14 bp at the HPRT locus in T lymphocytes and this has<br />

been suggested as a signature of exposure to densely ionizing radiation [210]. The<br />

carcinogenic potential of high LET radiation has also been proven. In addition, exposures<br />

to densely ionizing radiation have been shown to lead to a persistent, transmissible<br />

genomic instability in a variety of biological systems. With regard to DNA repair, it has<br />

recently been shown that DNA-PKcs participates in the repair of some frank DSBs and<br />

some non-DSB clustered damages that are converted into DSB by replication in tumor<br />

cells exposed to high LET radiation [211]. In addition, intact homologous recombination<br />

is also required to ensure DNA repair and cell survival after exposure to high-energy iron<br />

ions [212]. It appears that the biological features specific to high LET radiation do relate<br />

72


CHAPTER 1<br />

INTRODUCTION<br />

well to the known processing of the clustered lesions examined. We still need to<br />

reconstitute the path leading to complex large deletions.<br />

Radiation exposure has been associated with risk of diseases and in particular cancer. On<br />

the other hand, radiotherapy for cancer treatment is used advantageously for about 70%<br />

of cancer patients. The clinical applications of high LET radiation have a long history.<br />

High LET radiotherapy is increasing worldwide, with respect to proton therapy and<br />

hadron therapy, even though a complete understanding of the mechanisms underlying the<br />

biological action has not yet been determined. An important feature of high LET<br />

radiation which can be anticipated from studies on clustered DNA lesions is the increase<br />

of point mutations and clusters of mutations at non-DSB clusters. If a tumor suppressor<br />

gene is thus inactivated in normal tissue located near the irradiated tumor and hit by the<br />

beam, it may be a step towards the development of a secondary, radio-induced cancer.<br />

This situation would particularly apply to proliferating tissues. Even though the energy<br />

deposition of high LET radiation in normal tissue situated near the tumor is not elevated,<br />

the effect is not yet known. In tumor cells, such mutations may contribute to increase<br />

genomic instability and eventually lead to cell death. Low LET radiation also induces<br />

point mutations from dispersed base damages, but the probability of formation and the<br />

complexity of clustered lesions are lower. The frequency of mutation in normal tissue is<br />

expected to be higher than with high LET radiation. Delayed repair of clustered DNA<br />

damage possibly induced by high LET radiation could cause mutations but may also<br />

generate DSBs from stalled replication forks [213], as in rapidly replicating tumor cells.<br />

Such complex DSBs may undergo mutagenic repair via homologous recombination and<br />

may reflect the large (over Mbp) and complex deletions observed in culture cells exposed<br />

73


CHAPTER 1<br />

INTRODUCTION<br />

to high LET radiation. Such genome loss will contribute to genome instability of the<br />

tumor cells, and a probable final outcome is cell death. High LET-induced non-DSB<br />

clusters containing opposing AP sites would be particularly toxic via the formation of<br />

DSBs which will be poorly repaired, or due to the presence of unrepaired AP sites. A<br />

dead cell is a good cell not only for tumor eradication, but also for normal tissue since it<br />

prevents replication of damaged cells with radiation-induced mutations. With respect to<br />

the biological effectiveness of clustered DNA damage, high LET radiotherapy seems to<br />

present a relatively better benefit over risk to the patient than low LET radiotherapy.<br />

Indeed, in combination with a low dose deposited in the entrance channel, fewer as well<br />

as more easily repairable damages are produced in normal tissue, whereas a large dose is<br />

deposited in the tumor, accompanied by complex and poorly repairable lesion production.<br />

In addition, our understanding of repair processes at clustered lesions lets us predict that<br />

inhibiting the late steps of BER should increase toxic DSBs and repair intermediates and<br />

consequently would be beneficial for tumor cell killing and a combination of inhibitors<br />

for the late steps of BER, HR and/or NHEJ would lead to additional cell killing.<br />

Many studies on signaling pathways after low and high LET radiations have found the<br />

activation to be similar except in the intensity [214] but since the end result is different,<br />

there must be a divergence of pathways at some stage.<br />

Since the production of ROS is a major feature of irradiation where some of these could<br />

be permeable, it was of interest to look at the effect of irradiation on cells that had not<br />

exposed to radiation, i.e. the Bystander effect.<br />

74


CHAPTER 1<br />

INTRODUCTION<br />

1.6 Radiation induced Bystander Signaling<br />

The response of a cell or tissue to ionizing radiation has been shown to be mediated by<br />

direct damage to cellular components like DNA, lipids, proteins and small molecules as<br />

well as indirect damage mediated by radiolysis of water ([215]. It is now apparent that the<br />

target for the biological effects of ionizing radiation is not solely the irradiated cells but<br />

also includes the surrounding cells and tissues. Radiation-induced bystander effect is<br />

defined by the observance of biological effects of radiation in cells that were not<br />

themselves in the field of irradiation.<br />

Estimates of the various damage types at sub-cellular, cellular and supra-cellular<br />

level induced by low doses, where no experimental data are available, are generally based<br />

on linear back-extrapolations from data at higher doses. However, a large amount of data<br />

produced during the last decade seem to indicate higher damage yields than expected,<br />

possibly due to the so-called “bystander effect”. One of the most intriguing aspects of<br />

these experiments lies in that bystander damage is observed even at very low doses, down<br />

to the mGy level, and it does not significantly increase with dose. No clear-cut<br />

conclusions can be drawn on the mechanisms underlying bystander effect observations.<br />

Different pathways may be involved, depending on the specific conditions. In any case<br />

the various forms of cellular communication seem to play a key role, since damage in<br />

non-directly hit cells may represent a response to signals coming from cells that were<br />

directly hit by radiation (Fig. 1.6). If these findings are confirmed, radiobiology<br />

experiments may need to be re-interpreted and the models of low-dose radiation action<br />

may need to be revised. Furthermore, a significant role of bystander effects in vivo would<br />

imply the re-evaluation of models of damage at tissue and organ level. A typical example<br />

75


CHAPTER 1<br />

INTRODUCTION<br />

is provided by low-dose cancer risk, which up to now has generally been estimated by<br />

assuming a linear, no-threshold dose-response.<br />

Fig. 1.6: Radiation induced bystander effect. Irradiation of target cells leads to<br />

elicitation of response from unirradiated bystander cells. This takes place through the<br />

transduction of the signal from the irradiated cell to the bystander cell via gap junctions<br />

or through the release of the signaling molecules into the surrounding medium.<br />

A large amount of data on bystander effects has been obtained with the so-called ICM<br />

(Irradiated conditioned medium) treatment [216-219]. This technique consists in<br />

replacing the culture medium of non-irradiated cells with medium taken from cell<br />

cultures previously exposed to radiation. The main finding of these studies, lie in the fact<br />

that such treatment can reduce survival of unexposed cells. This suggests that, as a result<br />

of a radiation insult, certain cell types can release factors in the medium, which therefore<br />

76


CHAPTER 1<br />

INTRODUCTION<br />

becomes potentially cytotoxic. Therefore the onus lies on non gap junction mediated cell<br />

communication.<br />

Evidence for some forms of bystander effect has been suggested also by conventional<br />

irradiation with low doses. One of the first studies in this field has showed a significant,<br />

unexpected increase in the frequency of sister chromatid exchanges (SCEs) after<br />

exposure of Chinese hamster ovary (CHO) cells to very low doses of alpha particles,<br />

from 0.03 to 0.25 cGy [220]. While about 1% or less of the cells were actually traversed<br />

by a particle track, 30–45% of the individual cells showed increased levels of SCE. This<br />

suggests that genetic damage following exposure to very low doses of light ions may<br />

occur also in non directly-irradiated bystander cells.<br />

The studies discussed above strongly suggest a role of extranuclear- and possibly<br />

extracellular-targets in the propagation of certain damage types after irradiation with very<br />

low doses of light ions or treatment with ICM. Experiments allowing the identification of<br />

specific targets for such phenomena can be of great help to better understand the<br />

underlying mechanisms and to quantify the relative contributions of different targets.<br />

This kind of studies is now possible due to the increasing number of microbeam facilities.<br />

Such facilities in which cells are individually irradiated by a predefined exact number of<br />

particles allow the effects of individual particle traversals to be assessed, and these<br />

methods have become a useful tool in the study of bystander responses [221, 222]. Such<br />

studies suggest a major role of cell-to-cell communication via gap-junctions [223].<br />

77


CHAPTER 1<br />

INTRODUCTION<br />

1.7 Aims and Objectives: To Study<br />

1. Fractionated irradiation induced signaling events. The contribution of<br />

these events to the increased cell survival.<br />

2. High LET induced signaling and variance from low LET gamma<br />

radiation.<br />

3. Contribution of radiation induced bystander effect to cell survival and its<br />

mechanism.<br />

78


CHAPTER 2<br />

MATERIALS AND METHODS<br />

MATERIALS AND METHODS<br />

79


CHAPTER 2<br />

MATERIALS AND METHODS<br />

2.1 Chemicals<br />

Ammonium persulphate (APS), sucrose, 4-(2-hydroxyethyl)-1-<br />

piperazineethanesulfonicacid (HEPES), ethylene (oxyethylenenitrilo) tetraacetic acid (EGTA),<br />

phenylmethylsulphonyl fluoride (PMSF), aprotinin, leupeptin, pepstatin,<br />

ethylenediaminetetraaceticacid (EDTA), β-mercaptoethanol, Triton X-100, dithiothreitol (DTT),<br />

dimethy sulphoxide (DMSO), Acrylamide, N, N’-methylenebisacrylamide, Sodium dodecyl<br />

sulfate (SDS), Trizma base, N, N,N’N’-tetramethylethylenediamine (TEMED), Glycine,<br />

ponceau S and coomassie brilliant blue, bromophenol blue, paraformaldehyde,<br />

Polyoxyethylenesorbitan monolaurate (Tween 20), sulfanilamide, N-(1-naphthylethylenediamine),<br />

phosphoric acid, proteinase K, RPMI 1640 medium, Dulbecco’s Modified<br />

Eagle’s Medium (DMEM), Fetal Bovine Serum (FBS), Streptomycin and Penicillin, Trypan<br />

Blue, Sodium bicarbonate, agarose, bovine serum albumin (BSA), Sarcosyl and ethidium were<br />

procured from Sigma Chemicals (USA). PVDF (poly vinylidene difluoride) membrane and Nitro<br />

cellulose membrane were from Amersham Pharmacia Biotech (U.K). Chemiluminescence<br />

western blotting kits (Mouse/rabbit), RNA tissue kit were obtained from Roche Molecular<br />

Biochemicals (Germany). cMaster RT kit was procured from Eppendorf, Germany. Konica AX<br />

medical X-Ray films and Kodak developer and fixer and other chemicals were procured locally.<br />

The primary antibodies used were mouse anti-pATM (S1981), rabbit anti-γ-H2AX<br />

(S139), rabbit anti-pChk1 (Ser296), rabbit anti-pChk2 (Thr68), rabbit anti-pBRCA1 (Ser1524),<br />

rabbit anti-pATR (Ser428) and mouse anti-phospho-p53 (S15) were procured from Cell<br />

Signaling, USA. Secondary antibodies used were Alexa Fluor 488 goat anti-rabbit IgG and<br />

Alexa Fluor 488 goat anti-mouse IgG were obtained from Invitrogen, Molecular Probe, USA.<br />

80


CHAPTER 2<br />

MATERIALS AND METHODS<br />

2.2 Cell Lines<br />

Human lung adenocarcinoma A549 cells, human breast carcinoma MCF-7 cells, human<br />

monocyte U937 cells and human intestinal cell line INT 407 cells were cultured in 75 cm 2 tissue<br />

culture flasks (Falcon, USA) in in Dulbecco’s modified eagles medium (Sigma) supplemented<br />

with 10% fetal calf serum (Sigma). Mouse lymphoma cell line, EL-4 and mouse macrophage<br />

RAW 264.7 cells were grown in RPMI 1640 (Sigma, USA), supplemented with 10% fetal calf<br />

serum (Sigma). Cells were kept at 37 0 C in humidified atmosphere with 5% CO 2 in Nu-Air water<br />

jacketed CO 2 Incubator (USA).<br />

2.3 Irradiation<br />

2.3.1 Low LET ( 60 Co γ rays): A549, INT 407 and MCF-7 cells were exposed to different doses<br />

of γ irradiation using Gamma Cell 220 irradiator (Atomic Energy of Canada Ltd), at a dose rate<br />

of 3.5 Gy/min; the first set was irradiated with acute dose of 2 Gy or 10 Gy. The second set was<br />

exposed to 5 fraction of 2 Gy each over a period of five days.<br />

A549 cells and PMA stimulated U937 cells were exposed to 2 Gy γ-irradiation. The<br />

medium from irradiated cells was transferred to unirradiated cells. EL-4 and RAW 264.7 cells<br />

were resuspended in RPMI medium at a density of 1x10 6 cells/ml and exposed to 5 Gy γ<br />

irradiation using Gamma Cell 220 irradiator (Atomic Energy of Canada Ltd), at a dose rate of<br />

4.74 Gy/min. Medium from irradiated cells was transferred to unirradiated cells (1 h post<br />

irradiation). Some group of RAW 264.7 cells were stimulated with 500 ng/ml of bacterial<br />

Lipopolysaccharide (LPS, Sigma) for 3 h and then the medium was replaced before irradiation.<br />

2.3.2 Proton beam irradiation: Proton beam irradiation was carried out using the Radiation<br />

Biology beam line of in-house 4 MeV Folded Tandem Ion Accelerator (FOTIA) facility. The<br />

81


CHAPTER 2<br />

MATERIALS AND METHODS<br />

primary proton beam from the FOTIA was collimated using an adjustable slit to reduce the<br />

fluence and then diffused using a gold foil. The diffused beam was channeled to the exit window<br />

made of 20 micrometer titanium foil of 3 cm diameter to get uniformly distributed irradiation<br />

area. The samples were thus irradiated under normal atmospheric pressure at 24 o C. The target to<br />

be irradiated was positioned at a distance of 11 mm from the exit window, that being the closest<br />

possible place to mount the target. Before the irradiation, a silicon surface barrier (SSB) detector<br />

was positioned at the same position where samples were irradiated and the beam energy as well<br />

as the uniformity was measured. Another SSB detector placed inside the scattering chamber at a<br />

forward angle of 40 o to the primary beam, after the gold foil, served as a monitor detector. The<br />

ratio of the monitor detector counts to that of the flux measured using SSB detector at the sample<br />

position was measured by multiple trials and the calibration factor was obtained. Monitor<br />

detector counts and the measured ratio were used for delivering the required fluence to the<br />

samples. Flux was calculated to be 10 8 and the fluence was kept such that each cell is hit by at<br />

least 100 proton particles ensuring that bystander effect is kept to minimum. Fluence was then<br />

used to calculate dose with the help of specific ionization curve constructed for the given energy<br />

distribution and composition of the cellline. The average energy of the proton beam used in this<br />

study was measured to be 3.2 MeV and corresponding estimated LET is 12.5 keV/µm within the<br />

cell layer. Initial portion of Bragg’s curve was used for dose estimation since the cell layer used<br />

for irradiation was in monolayer geometry with thickness less than 10µm, which helped to avoid<br />

steep increase in LET within the sample.<br />

The cells were grown in monolayer geometry on 35mm petri dishes. Media was removed and<br />

petridishes were mounted on the exit window vertically for the irradiation, after which they were<br />

82


CHAPTER 2<br />

MATERIALS AND METHODS<br />

resupplemented with the medium. On an average samples were irradiated for 1 minute to get the<br />

required fluence. The cells were exposed to 2 Gy dose with the help of monitor detector counts.<br />

2.3.3 High LET (heavy ion): For the experiments plateau phase A549 cells were irradiated with<br />

12 C and 16 O ions from 15 UD Pelletron accelerator at the Inter University Accelerator Centre,<br />

New Delhi. During carbon and oxygen irradiation, 35 mm petri plates [NUNC] with cellular<br />

monolayers were kept perpendicular to the particle beam coming out of the exit window. The<br />

petri dishes were kept in the serum free medium in a plexiglass magazine during the irradiation.<br />

The computer controlled irradiation system is such that during the actual irradiation the dish was<br />

removed from medium.<br />

At the cell surface the energy of C ions was 62 MeV (5.16 MeV/u , LET=290 keV/ µm] and the<br />

energy of O ions was 55.8 MeV [LET=614 keV/ µm]. Both the energy and LET were calculated<br />

using the Monte Carlo Code SRIM code [224, 225]. The corresponding fluence for the dose of 1<br />

Gy of carbon was 2.2x10 6 particles cm -2 , thus each cell nuclei can be expected to be physically<br />

traversed at least once by carbon ion. The corresponding fluence for the dose of 1Gy of oxygen<br />

was 1.01x10 6 particles cm -2 , thus each cell nuclei can be expected to be physically traversed at<br />

least once by oxygen ion.<br />

Fluence was calculated using the relation given below [226], approximating the density ρ of<br />

cellular matrix as 1.0:<br />

Dose [Gy] = 1.6 x 10 -9 x LET (keV/ µm) x particle fluence (cm -2 ) x 1/ρ (cm 3 /g)<br />

83


CHAPTER 2<br />

MATERIALS AND METHODS<br />

2.4 Clonogenic Cell Survival Assay<br />

After treatment, cells were seeded out in appropriate dilutions (500 cells/60 mm petriplate, 1000<br />

cells/60 mm petriplate or 2000 cells/60 mm petriplate) to form colonies in 14 days. Colonies<br />

were fixed with glutaraldehyde (6.0% v/v), stained with crystal violet (0.5% w/v) and counted<br />

using a stereomicroscope. Colonies containing more than 50 cells were scored as survivors.<br />

Absolute plating efficiency of A549 cells at 0 Gy (sham irradiated control) was 32.3±2.4 %.<br />

Absolute plating efficiency of MCF-7 cells at 0 Gy (sham irradiated control) was 45.2±2.4 %.<br />

2.5 Microarray<br />

Two independent series of experiments were performed with Human Genome U-133 Plus 2.0<br />

microarrays containing 54675 probe sets (Affymetrix, Santa Clara, CA). Microarray analysis was<br />

carried out at Oscimum Biosolution, Hyderabad on paid basis. Total RNA was isolated from<br />

~3 × 10 6 A549 cells with RNeasy Mini Kits (Qiagen) including digestion with RNase-free<br />

DNase I, and its quantity and integrity were checked spectrophotometrically and by<br />

electrophoresis in 1% agarose gels. Materials and methods for microarrays were from<br />

Affymetrix (Santa Clara); double-stranded cDNA was prepared with the GeneChip Expression<br />

3’-Amplification One-Cycle cDNA Synthesis Kit, cleaned using the Sample Cleanup Module,<br />

and biotinylated cRNA was synthesized with GeneChip Expression 3’-Amplification Reagents<br />

for IVT Labeling, cleaned on RNA Sample Cleanup columns, and fragmented at 94 °C for<br />

35 min in Fragmentation Buffer. The biotinylated cRNA was hybridized first to a control Test3<br />

microarray to evaluate its quality and then to a Human Genome U-133 Plus 2.0 microarrays<br />

containing 54675 probe sets, using GeneChip Expression 3’-Amplification Reagents.<br />

Microarrays were stained with streptavidin–phycoerythrin conjugate and scanned in a Gene<br />

84


CHAPTER 2<br />

MATERIALS AND METHODS<br />

Array G2500A scanner (Agilent) and the expression data on human Affymetrix chip was<br />

generated. Since the numbers of samples in each treatment type were sparse, we used a bivariate<br />

simulation approach to identify differentially expressed (DE) genes for the specified threshold<br />

fold change (FC). The DE candidates across each comparison were further evaluated for their<br />

biological relevance through Gene Ontology and Pathways studies. The statistical analysis was<br />

carried out using R package and the biological analysis was carried out using Genowiz TM<br />

software.<br />

For comparative evaluation, an un-irradiated control cell was used. The expression data on<br />

Affymetrix Human 133AB chip was obtained for each sample. The primary objective of the<br />

study was to obtain differentially expressed (DE) probe sets (genes) across various comparisons.<br />

Also, the interest was to study the clustering of genes and samples and the functional relevance<br />

of the differentially expressed genes.<br />

Upon identifying the DE genes, the next interest was to study the biological relevance of selected<br />

marker genes through GO and Pathway analysis.<br />

The data analysis process involves Robust Multichip Average (RMA) normalization, quality<br />

check analysis, differential expression analysis and the gene enrichment analysis.<br />

The correlation between the expression values for samples was studied through Pearson's<br />

coefficient. A pairwise correlation analysis was performed to generate a correlation matrix<br />

indicating how the samples are related with each other. The coefficient tells about the similarity<br />

of expression trend between the two arrays. A coefficient value close to 1.0 indicates the linear<br />

85


CHAPTER 2<br />

MATERIALS AND METHODS<br />

expression between the arrays. The minimum correlation between the samples was 0.987, which<br />

was above the acceptable criterion of 0.8.<br />

2.6 Semiquantitative Reverse transcriptional polymerase chain reaction (RT-PCR)<br />

Total RNA from all the groups of cells (1 x 10 6 cells) was extracted after required time periods<br />

of irradiation and eluted in 30 µl RNAase free water using RNA tissue kit (Roche). Equal<br />

amount of RNA in each group was reverse transcribed using cMaster RT kit (Eppendorf). Equal<br />

amount (2µl) of cDNA in each group was used for specific amplification of β-actin, NF-κB<br />

(p65), iNOS, p21, p53, ATM, DNA-PK, ATR, MLH1, Rad52, GADD45α, Toll-like receptor 3<br />

(TLR3), Ferredoxin reductase (FDXR), Chk1, Chk2, Bcl-2 or Bax gene using gene specific<br />

primers (Table 2.6). The PCR condition were 94 o C, 5 min initial denaturation followed by 30<br />

cycles of 94 o C 45 s, 55 – 62 o C 45 s (depending on the annealing temperature of specific primers<br />

given in Table 2.6), 72 o C 45 s and final extension at 72 o C for 10 min. Equal amount of each<br />

PCR product (10 µl) was run on 1.5 % agarose gel containing ethidium bromide in tris borate<br />

EDTA buffer at 60 V. The bands in the gel were visualized under UV lamp and relative<br />

intensities were quantified using Gel doc software (Syngene).<br />

2.7 Immunofluorescence staining<br />

Cells were grown in cover slips which were kept in 35mm petri dishes and irradiated as<br />

mentioned above. Cell layer were washed at different time period after irradiation in PBS and<br />

fixed for 20 min in 4% paraformaldehyde at room temperature. Afterwards, the cells were<br />

washed twice in PBS. For immunofluorescence staining, cells were permeabilized for 3 min in<br />

0.25% Triton X-100 in PBS, washed two times in PBS and blocked for 1 h with 5% BSA in<br />

86


CHAPTER 2<br />

MATERIALS AND METHODS<br />

PBS. Antibodies were diluted (1:200) in 1% BSA in PBS. Cells were incubated with primary<br />

antibodies for 1 h 30 min at room temperature, washed three times in PBS and incubated with<br />

secondary antibodies for 1 h at room temperature. Finally, cells were rinsed and mounted with<br />

ProLong Gold antifede with DAPI mounting media (Molecular Probe, USA). Images were<br />

captured using Carl Zeiss confocal microscope. Acquisition settings were optimized to obtain<br />

maximal signal in immunostained cells with minimal background.<br />

Table 2.6: The sequences of gene specific primers and their annealing temperature (Tm) used in<br />

RT-PCR experiments<br />

Gene Primer Sequence Tm ( o C)<br />

β-Actin<br />

F 5’-TGGAATCCTGTGGCATCCATGAAAC-3’<br />

R 5’- TAAAACGCAGCTCAGTAACAGTCCG-3’<br />

55<br />

iNOS F 5’- CTCTGACAGCCCAGAGTTCC - 3’<br />

62<br />

R 5’- AGGCAAAGGAGGAGAAGGAG- 3’<br />

p21 F 5’- GCCTTAGCCCTCACTCTGTG- 3’<br />

58<br />

R 5’-GGTTGGGAGGGGCTTAAATA- 3’<br />

p53 F 5’- CGGGTGGAAGGAAATTTGTA- 3’<br />

58<br />

R 5’- CTTCTGTACGGCGGTCTCTC - 3’<br />

p65 F 5’- ACAACTGAGCCCATGCTGAT- 3’<br />

50<br />

ATM<br />

ATR<br />

R 5’- GAGAGGTCCATGTCCGCAAT- 3’<br />

F 5’-GGACAGTGGAGGCACAAAAT-3’<br />

R 5’-GTGTCGAAGACAGCTGGTGA-3’<br />

F 5’-CTCGCTGAACTGTACGTGGA-3’<br />

R 5’-GCATAGCTCGACCATGGATT-3’<br />

57<br />

57<br />

DNA-PK F 5’-TGCCAATCCAGCAGTCATTA-3’ 64<br />

87


CHAPTER 2<br />

MATERIALS AND METHODS<br />

Chk2<br />

Chk1<br />

Bax<br />

Bcl-2<br />

Rad52<br />

MLH1<br />

CDKN1A<br />

(p21)<br />

GADD45α<br />

TLR3<br />

FDXR<br />

R 5’-GGTCCATCAGGCACTTCACT-3’<br />

F 5’-GGCTTCAGGATGAAGACATGA-3’<br />

R 5’-CACAACACAGCAGCACACAC-3’<br />

F 5’-TGTCAGAGTCTCCCAGTGGA-3’<br />

R 5’-AGGGGCTGGTATCCCATAAG-3’<br />

F 5’-TCCCCCCGAGAGGTCTTT-3’<br />

R 5’-CGGCCCCAGTTGAAGTTG-3’<br />

F 5’-GAGGATTGTGGCCTTCTTTG-3’<br />

R 5’-ACAGTTCCACAAAGGCATCC-3’<br />

F 5’-AGTTTTGGGAATGCACTTGG-3’<br />

R 5’-TCGGCAGCTGTTGTATCTTG-3’<br />

F 5’-GAGGTGAATTGGGACGAAGA-3’<br />

R 5’-TCCAGGAGTTTGGAATGGAG-3’<br />

F 5’-CAGCAGAGGAAGACCATGTG-3’<br />

R 5’-GGCGTTTGGAGTGGTAGAAA-3’<br />

F 5’-TGCGAGAACGACATCAACAT-3’<br />

R 5’-TCCCGGCAAAAACAAATAAG-3’<br />

F 5’-GCCTCTTCGTAACTTGACCA-3’<br />

R 5’-AAGGATGTGGAGGTGAGACA-3’<br />

F 5’-AGAGAACGGACATCACGAAG-3’<br />

R 5’-GTCCTGGAGACCCAAGAAAT-3’<br />

64<br />

59<br />

56<br />

59<br />

50<br />

52<br />

59<br />

58<br />

57<br />

57<br />

F- Forward and R- Reverse<br />

2.8 Image analysis using ImageJ software<br />

The captured images were analyzed for relative quantification of phosphorylation using ImageJ<br />

software [227]. A 25 x 25 pixel box was positioned over the fluorescent image of each cell, and<br />

88


CHAPTER 2<br />

MATERIALS AND METHODS<br />

the average intensity within the selection (the sum of the intensities of all the pixels in the<br />

selection divided by the number of pixels) was measured. All the foci were counted manually; at<br />

least 100 cells per experiment were analyzed from three independent experiments.<br />

2.9 Western Blotting<br />

Proteins were separated by 8 % SDS-PAGE using minigel apparatus (Technosource, India) run<br />

at a constant voltage of 150 volts for 3 hr. The resolved proteins were then transferred to<br />

nitrocellulose membrane using an immunoblot apparatus (Technosource, India) at a constant<br />

voltage of 70 volts for 90 min. Following transfer, membranes were stained using Ponceau S<br />

solution to confirm complete transfer of proteins from the gel. Thereafter, the gel was also<br />

stained with 0.2 % coomassie blue to reconfirm the same. The membranes were blocked<br />

overnight with wash buffer (0.05 M Tris, 0.15 M NaCl and 0.1 % Tween 20 pH8.0) containing 5<br />

% BSA with gentle rocking at 4 0 C. The blocked membranes were probed with specific primary<br />

antibodies (dilutions used are mentioned in the respective figures) followed by washing with<br />

wash buffer four times for 5 min each. Thereafter the membranes were reprobed with horse<br />

radish peroxidase conjugated secondary antibody (Roche Molecular, Biochemicals. Germany) at<br />

a dilution of 1:2000 in wash buffer containing 1 % BSA, washed 4x5 min with wash buffer and<br />

developed with Mouse/Rabbit Western Blotting Kit (Roche Molecular, Biochemicals. Germany).<br />

Chemiluminescence was recorded on Konica AX medical X-Ray films. Images were digitized<br />

using HP Scan Jet ADF Scanner and the figures were assembled using Gelquant Version 2.7<br />

Software.<br />

89


CHAPTER 2<br />

MATERIALS AND METHODS<br />

2.10 Measurement of DNA damage<br />

DNA damage in all the groups of EL-4 cells 2 h after exposure to different conditioned medium,<br />

was measured by alkaline single cell gel electrophoresis (comet assay) [228]. In brief, frosted<br />

microscope slides (Gold Coin, Mumbai, India) were covered with 200 µl of 1% normal melting<br />

agarose (NMA) in phosphate buffered saline (PBS) at 45 o C, covered with a cover slip and kept<br />

at 4 o C for 10 min. After solidification a second layer of 200 µl of 0.5% low melting agarose<br />

(LMA) containing approximately 105 cells at 37 o C was layered over it. The cover slips were<br />

replaced immediately and the slides were kept at 4 o C. After solidification of the LMA, the<br />

cover-slips were removed and slides were placed in the chilled lysing solution (2.5M NaCl,<br />

100mM Na 2 -EDTA, 10mM Tris–HCl, pH 10, and 1% DMSO, 1% Triton X100 and 1% sodium<br />

sarcosinate) for 1 h at 4 o C. The slides were removed from the lysing solution and placed on a<br />

horizontal electrophoresis tank filled with freshly prepared alkaline buffer (300mM NaOH, 1mM<br />

Na 2 -EDTA and 0.2% DMSO, pH≥13.0) and were equilibrated in the above buffer for 20 min and<br />

electrophoresis was carried out at 25V for 20 min. After electrophoresis the slides were washed<br />

gently with 0.4M Tris–HCl buffer, pH 7.4, to remove the alkali, stained with 50 µl of propidium<br />

iodide (PI, 20 µg/ml) and visualized using a Carl Zeiss Fluorescent microscope (Axioskop). The<br />

images (40–50 cells/slide) were captured with high-performance GANZ (model: ZCY11PH 4)<br />

color video camera. The integral frame grabber used in this system (Cvfbo1p) is a PC based card<br />

and it accepts color composite video output of the camera. The quantification of the DNA strand<br />

breaks of the stored images was done using the CASP software by which %DNA in tail, tail<br />

length, tail moment and Olive tail moment could be obtained directly [229].<br />

90


CHAPTER 2<br />

MATERIALS AND METHODS<br />

2.11 Measurement of NO production in EL-4 cells<br />

Control unirradiated cells, cells exposed to radiation and the cells exposed to different<br />

conditioned media as mentioned above were cultured for 24 h at 37 0 C in 5%CO 2 / 95% air. The<br />

production of NO in the culture supernatants was measured by assaying NO - 2 using colorimetric<br />

Griess reaction [230].<br />

In brief, 100 µl of culture supernatants was incubated with an equal volume of Griess reagent<br />

[1% sulfanilamide and 0.1% N-(1-naphthyl-ethylenediamine) in 2.5% phosphoric acid; Sigma,<br />

USA] at room temperature for 10 min in a 96 well plate. The absorbance at 550 nm was<br />

measured using a microplate reader (Fluostar Optima, Biotron Healthcare). Absorbance<br />

measurements were converted to µ moles of NO 2<br />

-<br />

using a standard curve of NaNO 2.<br />

2.12 Measurement of Apoptosis by DNA fragmentation<br />

Apoptosis in control unirradiated cells, cells exposed to radiation and the cells exposed to<br />

different conditioned media as mentioned above was measured by DNA fragmentation assay<br />

[231]. After transfer of medium from irradiated cells, all the above mentioned groups of EL-4<br />

cells were cultured for 24 h at 37 0 C in 5%CO 2 , 95% air. The cells were harvested and lysed in<br />

lysis buffer (10 mM EDTA, 50 mM Tris–HCl, pH 8.0, 0.5% sodium lauryl sarcosine) containing<br />

the 100 mg/ml proteinase K at 55 o C for 2 h. The DNA was extracted with phenol/chloroform and<br />

precipitated with ethanol. The RNA was removed by RNAse A treatment and DNA was<br />

electrophorased on agarose gel to visualize the fragmented DNA.<br />

91


CHAPTER 2<br />

MATERIALS AND METHODS<br />

2.13 Transfection with ShRNA<br />

MLH1, Rad52 and Control shRNA plasmids were procured from Santa cruz biotechnology (Cat<br />

No. sc-35943-SH, sc-37399-SH and sc-108060 respectively). In a six well tissue culture plate,<br />

A549 cells were grown to 50-70% confluency in antibiotic-free normal growth medium<br />

supplemented with FBS and transfection was carried out as per supplier’s instruction. 48 hours<br />

post-transfection, medium was aspirated and replaced with fresh medium containing puromycin<br />

(2µg/ml) for selection of stably transfected cells. Semi-quantitative RT-PCR was performed to<br />

monitor MLH1 and Rad52 gene expression knockdown using gene specific primers (Table 1).<br />

2.14 Statistical Analysis<br />

The data were imported to excel work sheets, and graphs were made using Origin version 5.0.<br />

One-way ANOVA with Tukey–Kramer Multiple Comparisons as post-test for P < 0.05 used to<br />

study the significant level. Data were insignificant at P > 0.05. Each point represented as<br />

mean±S.E. (standard error of the mean).<br />

92


CHAPTER 3<br />

RESULTS<br />

RESULTS<br />

93


CHAPTER 3<br />

RESULTS<br />

Exposure of cells to ionizing radiation leads to a cascade of signaling events in cells. Until<br />

recently, the molecular mechanisms by which cells recognize ionizing radiation damage and<br />

respond to it were unknown, but now they are steadily emerging. Cells use complex protein<br />

signaling systems, which recognize radiation induced oxidative damage to DNA and plasma<br />

membrane lipids, and stimulate intracellular signaling pathways. The balance between various<br />

pathways and the ever-changing cell’s microenvironment together determines the ultimate fate of<br />

the cell which may be death or survival. On surveying the literature on radiation induced signal<br />

transduction, most studies seemed to have been done at doses of choice, very often with<br />

supralethal doses of radiation with the assumption that the same effects can be attributed to lower<br />

or therapeutic doses of radiation. Moreover, the time of observation chosen by the investigators<br />

also varied according to their own past experiences. Most, if not all the studies examined<br />

signaling after a single dose of irradiation. The work embodied in this thesis; therefore, attempts<br />

to bridge these gaps in our understanding of radiation induced signaling.<br />

3.1 Fractionated irradiation induced signaling and repair in mammalian cells.<br />

Radiation therapy is delivered clinically in the form of fractionated doses; therefore the effect of<br />

fractionated doses of γ-irradiation (2Gy per fraction over 5 days), as delivered in cancer<br />

radiotherapy was compared with acute doses of 10 and 2Gy, in three human cell lines viz. MCF-<br />

7, INT 407 and A549. Cell lines were divided into 4 Groups. While a) controls were sham<br />

irradiated, the rest were subjected to b) 2 Gy of 60 Co γ irradiation daily over a period of 5 days c)<br />

10 Gy acute dose (which is the sum total of all the fractions i.e. 2 Gy x 5) or d) 2 Gy acute dose<br />

(single dose that was delivered in each fraction) of 60 Co γ irradiation. The 2 Gy dose was<br />

included to compare the effect of a single fraction that was delivered with the effect of the total<br />

94


CHAPTER 3<br />

RESULTS<br />

fractionated regimen and the 10 Gy acute dose, which is the sum total of all the fractions<br />

delivered. It was of interest to compare both these doses with the fractionated doses.<br />

3.1.1 Clonogenic Cell Survival<br />

To determine the sensitivity of cells to different doses of radiation, their ability to form colonies<br />

after irradiation was observed. There was 52.4±2.5% survival of MCF-7 cells that had been<br />

exposed to 2 Gy acute dose where as there was 25±1.5% of cell survival in those that had been<br />

exposed to 5 fractions of 2 Gy and 0.5±0.05% of cells survived after exposure to 10 Gy acute<br />

dose (Fig.3.1.1).<br />

54±2.1% of INT 407 cells survived after exposure to 2 Gy where as only 30±1.8% survived after<br />

exposure to 5 fractions of 2 Gy and only 1.5±0.08% of cells survived after exposure to 10 Gy<br />

(Fig.3.1.1).<br />

The A549 cells responded differently to radiation. While 60±2.5% of A549 cells survived after<br />

2 Gy. 48±2.8% of cells survived after exposure to 5 fractions of 2 Gy and only 4±0.5% of cells<br />

survived after exposure 10 Gy (Fig.3.1.1).<br />

95


CHAPTER 3<br />

RESULTS<br />

Percentage Survival<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

MCF7<br />

INT407<br />

A549<br />

0Gy 2Gy 5X2Gy 10Gy<br />

Radiation Dose<br />

Fig. 3.1.1 Relative radioresistance of three cell line viz. MCF-7, INT 407 and A549. Data represents<br />

means ± SE of three independent experiments. Key; Lane 1, control unirradiated cells. Lane 2, 2 Gy<br />

acute dose ( single dose that was delivered in each fraction ) of 60 Co γ-irradiation Lane 3 2 Gy of<br />

60 Co γ-irradiation daily over a period of 5 days, Lane 4, 10 Gy acute (which is the sum total of all the<br />

fraction ) 60 Co γ-irradiation; 2000 cells/60 mm petriplate plated for this treatment group.<br />

3.1.2 Human Genome U-133 Plus 2.0 microarray data analysis.<br />

Having established that A549 was relatively more radioresistant, global expression analysis of<br />

54675 probe set using Human Genome U-133 Plus 2.0 microarray was performed in A549 cells.<br />

For microarray analysis, 4 h after irradiation was selected as the time point to monitor the early<br />

response of cells to irradiation and to identify differentially expressed early genes that mediate<br />

cellular events such as DNA repair and apoptosis.<br />

96


CHAPTER 3<br />

RESULTS<br />

On comparison of control Vs 5X2 Gy (henceforth called fractionated dose), 461 genes were<br />

identified as differentially-expressed; 312 genes up-regulated (defined as a 2.0-fold or greater<br />

increase, Table 3.1.2A) and 149 down-regulated (defined as a 2.0-fold or greater decrease, Table<br />

3.1.2B). Up-regulated genes were associated with p53 signaling pathway, cytokine-cytokine<br />

receptor interaction, hematopoietic cell lineage, Toll-like receptor signaling pathway, Jak-STAT<br />

signaling pathway, MAPK signaling pathway, cell-cycle check point, B cell receptor signaling<br />

pathway. Down-regulated genes were associated with TGF-beta signaling pathway and tyrosine<br />

metabolism.<br />

In the second group (2 Gy Vs 5X2 Gy), 234 genes were identified as differentially-expressed;<br />

172 genes as up-regulated (Table 3.1.2C) and 62 as down-regulated (Table 3.1.2D). Genes<br />

involved in cytokine-cytokine receptor interaction, toll-like receptor signaling pathway,<br />

hematopoietic cell lineage, Jak-STAT signaling pathway, MAPK signaling pathway were upregulated.<br />

Gene involved in folate biosynthesis, glycine, serine and threonine metabolisms were<br />

down-regulated.<br />

In the 10 Gy Vs fractionated group, 289 genes were identified as differentially-expressed; 211<br />

genes as up-regulated (Table 3.1.2E) and 78 as down-regulated (Table 3.1.2F). Genes involved<br />

in cytokine-cytokine receptor interaction, toll-like receptor signaling pathway, cell cycle, DNA<br />

replication, mismatch repair, homologous recombination were up-regulated. Genes involved in<br />

regulation of autophagy, base excision repair, folate biosynthesis were down-regulated.<br />

Cluster analysis of the expression of these genes showed that the fractionated group differed<br />

most in gene expression compared to the other three groups (control, 2Gy or 10Gy) (Fig. 3.1.2<br />

A, B & C).<br />

97


CHAPTER 3<br />

RESULTS<br />

Fig. 3.1.2 Relative expression of genes in A549 cells exposed to 0Gy, 2Gy, 5X2Gy or 10Gy (A-<br />

C) Hierarchical clustering of all experimental groups based on the expression of all genes. (A)<br />

Heatmap showing similarity of genes profile for the comparison Control Vs 5X2Gy. (B)<br />

Heatmap showing similarity of genes profile for the comparison 2Gy Vs 5X2Gy. (C) Heatmap<br />

showing similarity of genes profile for the comparison 10Gy Vs 5X2Gy.<br />

98


CHAPTER 3<br />

RESULTS<br />

Table 3.1.2A List of Up-regulated Genes: Control Vs 5X2Gy<br />

S.No Public ID Gene Symbol Gene Title<br />

1 M21121 CCL5 chemokine (C-C motif) ligand 5<br />

2 NM_153838 GPR115 G protein-coupled receptor 115<br />

3 NM_134268 CYGB cytoglobin<br />

4 NM_005764 PDZK1IP1 PDZK1 interacting protein 1<br />

5 BC021286 C1orf187 chromosome 1 open reading frame 187<br />

6 AF355465 ZMAT3 zinc finger, matrin type 3<br />

7 AB014766 --- ---<br />

8 AF043341 CCL5 chemokine (C-C motif) ligand 5<br />

9 AK094613 RPS24 Ribosomal protein S24<br />

10 AI332397 EIF4A2 eukaryotic translation initiation factor 4A, isoform 2<br />

11 AA580691 RBM25 RNA binding motif protein 25<br />

12 BC039509 LOC643401 hypothetical protein LOC643401<br />

13 BC007947 PHLDB3 pleckstrin homology-like domain, family B, member 3<br />

14 BG389789 --- ---<br />

15<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

BE708432 MALAT1<br />

16 AK090412 LOC375010 hypothetical LOC375010<br />

17 BC040303 LOC727916 hypothetical protein LOC727916<br />

18 BC043411 --- Homo sapiens, clone IMAGE:6155889, mRNA<br />

19 AL832227 BCL8 B-cell CLL/lymphoma 8<br />

20 AL832227 BCL8 B-cell CLL/lymphoma 8<br />

21 AF086134 --- Full length insert cDNA clone ZA88B06<br />

22 AF147425 SF3B14 Splicing factor 3B, 14 kDa subunit<br />

23 AK098337 LOC375010 hypothetical LOC375010<br />

24<br />

serpin peptidase inhibitor, clade E (nexin, plasminogen<br />

activator inhibitor type 1), member 1<br />

BC020765 SERPINE1<br />

25 BC017771 CCDC90B coiled-coil domain containing 90B<br />

26 BC018787 MGC59937 Similar to RIKEN cDNA 2310002J15 gene<br />

27 BC023629 PDLIM7 PDZ and LIM domain 7 (enigma)<br />

28 BC023629 PDLIM7 PDZ and LIM domain 7 (enigma)<br />

29 BC041011 EME2 essential meiotic endonuclease 1 homolog 2 (S. pombe)<br />

30 NM_001613 ACTA2 actin, alpha 2, smooth muscle, aorta<br />

31 BG339064 BTG2 BTG family, member 2<br />

32 NM_006763 BTG2 BTG family, member 2<br />

33 BF111821 WSB1 WD repeat and SOCS box-containing 1<br />

34 BG491844 JUN jun oncogene<br />

35 NM_002228 JUN jun oncogene<br />

36 NM_002276 KRT19 keratin 19<br />

37 NM_006762 LAPTM5 lysosomal associated multispanning membrane protein 5<br />

38<br />

NM_002462 MX1<br />

myxovirus (influenza virus) resistance 1, interferoninducible<br />

protein p78 (mouse)<br />

99


CHAPTER 3<br />

RESULTS<br />

39 NM_014734 KIAA0247 KIAA0247<br />

40 NM_005562 LAMC2 laminin, gamma 2<br />

41<br />

NM_002615 SERPINF1<br />

serpin peptidase inhibitor, clade F (alpha-2 antiplasmin,<br />

pigment epithelium derived factor), member 1<br />

42 NM_000389 CDKN1A cyclin-dependent kinase inhibitor 1A (p21, Cip1)<br />

43 NM_001710 CFB complement factor B<br />

44 NM_001674 ATF3 activating transcription factor 3<br />

45<br />

angiotensinogen (serpin peptidase inhibitor, clade A,<br />

member 8)<br />

NM_000029 AGT<br />

46 NM_014905 GLS glutaminase<br />

47 NM_005541 INPP5D inositol polyphosphate-5-phosphatase, 145kDa<br />

48 NM_001345 DGKA diacylglycerol kinase, alpha 80kDa<br />

49 NM_005502 ABCA1 ATP-binding cassette, sub-family A (ABC1), member 1<br />

50 NM_005103 FEZ1 fasciculation and elongation protein zeta 1 (zygin I)<br />

51 NM_001924 GADD45A growth arrest and DNA-damage-inducible, alpha<br />

52<br />

carbohydrate (N-acetylglucosamine-6-O) sulfotransferase<br />

2<br />

NM_004267 CHST2<br />

53 NM_004585 RARRES3 retinoic acid receptor responder (tazarotene induced) 3<br />

54 NM_006851 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

55 NM_021127 PMAIP1 phorbol-12-myristate-13-acetate-induced protein 1<br />

56 NM_000229 LCAT lecithin-cholesterol acyltransferase<br />

57 NM_000698 ALOX5 arachidonate 5-lipoxygenase<br />

58 NM_004961 GABRE gamma-aminobutyric acid (GABA) A receptor, epsilon<br />

59 NM_002985 CCL5 chemokine (C-C motif) ligand 5<br />

60 NM_001549 IFIT3 interferon-induced protein with tetratricopeptide repeats 3<br />

61 AA164751 FAS Fas (TNF receptor superfamily, member 6)<br />

62 NM_000043 FAS Fas (TNF receptor superfamily, member 6)<br />

63 NM_004165 RRAD Ras-related associated with diabetes<br />

64 NM_004165 RRAD Ras-related associated with diabetes<br />

65 NM_006307 SRPX sushi-repeat-containing protein, X-linked<br />

66 NM_000203 IDUA iduronidase, alpha-L-<br />

67 NM_000576 IL1B interleukin 1, beta<br />

68 AB046692 AOX1 aldehyde oxidase 1<br />

69 NM_001159 AOX1 aldehyde oxidase 1<br />

70 NM_000600 IL6 interleukin 6 (interferon, beta 2)<br />

71 AF026303 SULT1C2 sulfotransferase family, cytosolic, 1C, member 2<br />

72 NM_002392 MDM2 Mdm2 p53 binding protein homolog (mouse)<br />

73 NM_005101 ISG15 ISG15 ubiquitin-like modifier<br />

74 NM_005658 TRAF1 TNF receptor-associated factor 1<br />

75 NM_001785 CDA cytidine deaminase<br />

76 NM_000756 CRH corticotropin releasing hormone<br />

77 NM_003733 OASL 2'-5'-oligoadenylate synthetase-like<br />

78 NM_000777 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

79 NM_001197 BIK BCL2-interacting killer (apoptosis-inducing)<br />

100


CHAPTER 3<br />

RESULTS<br />

80 NM_002185 IL7R interleukin 7 receptor<br />

81 NM_016352 CPA4 carboxypeptidase A4<br />

82 NM_014668 GREB1 GREB1 protein<br />

83<br />

GABBR1 ///<br />

NM_006398 UBD<br />

84 NM_005393 PLXNB3 plexin B3<br />

85 NM_001086 AADAC arylacetamide deacetylase (esterase)<br />

86 NM_001941 DSC3 desmocollin 3<br />

87 NM_003265 TLR3 toll-like receptor 3<br />

88 NM_005531 IFI16 interferon, gamma-inducible protein 16<br />

gamma-aminobutyric acid (GABA) B receptor, 1 ///<br />

ubiquitin D<br />

89<br />

carcinoembryonic antigen-related cell adhesion molecule<br />

NM_001712 CEACAM1 1 (biliary glycoprotein)<br />

90 NM_015364 LY96 lymphocyte antigen 96<br />

91 NM_003811 TNFSF9 tumor necrosis factor (ligand) superfamily, member 9<br />

92<br />

NM_001974 EMR1<br />

egf-like module containing, mucin-like, hormone<br />

receptor-like 1<br />

93 NM_001561 TNFRSF9 tumor necrosis factor receptor superfamily, member 9<br />

94 NM_001531 MR1 major histocompatibility complex, class I-related<br />

95 NM_001531 MR1 major histocompatibility complex, class I-related<br />

96 NM_013314 BLNK B-cell linker<br />

97 NM_004110 FDXR ferredoxin reductase<br />

98<br />

SAA1 ///<br />

NM_030754 SAA2 serum amyloid A1 /// serum amyloid A2<br />

99 AF204231 GOLGA8A golgi autoantigen, golgin subfamily a, 8A<br />

100 AF208043 IFI16 interferon, gamma-inducible protein 16<br />

101 AF247168 C1orf63 chromosome 1 open reading frame 63<br />

102 AF007162 CRYAB crystallin, alpha B<br />

103 BC001743 C7orf44 chromosome 7 open reading frame 44<br />

104 AF237813 ABAT 4-aminobutyrate aminotransferase<br />

105 AF237813 ABAT 4-aminobutyrate aminotransferase<br />

106<br />

X16354 CEACAM1<br />

carcinoembryonic antigen-related cell adhesion molecule<br />

1 (biliary glycoprotein)<br />

107<br />

DNA segment on chromosome 4 (unique) 234 expressed<br />

NM_014392 D4S234E sequence<br />

108<br />

DNA segment on chromosome 4 (unique) 234 expressed<br />

BC001745 D4S234E sequence<br />

109 BC001422 PGF placental growth factor<br />

110<br />

M87507 CASP1<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

111 M15329 IL1A interleukin 1, alpha<br />

112 AF010446 MR1 major histocompatibility complex, class I-related<br />

113 AF031469 MR1 major histocompatibility complex, class I-related<br />

114 M11734 CSF2 colony stimulating factor 2 (granulocyte-macrophage)<br />

115 AB007458 TP53AP1 TP53 activated protein 1<br />

101


CHAPTER 3<br />

RESULTS<br />

116 L76224 GRIN2C glutamate receptor, ionotropic, N-methyl D-aspartate 2C<br />

117<br />

AF164622<br />

GOLGA8A<br />

///<br />

GOLGA8B<br />

golgi autoantigen, golgin subfamily a, 8A /// golgi<br />

autoantigen, golgin subfamily a, 8B<br />

118 M13436 INHBA inhibin, beta A<br />

119 AF119863 MEG3 maternally expressed 3<br />

120 AF063612 OASL 2'-5'-oligoadenylate synthetase-like<br />

121 AF064771 DGKA diacylglycerol kinase, alpha 80kDa<br />

122<br />

U13698 CASP1<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

123<br />

caspase 1, apoptosis-related cysteine peptidase<br />

124<br />

U13699<br />

CASP1<br />

(interleukin 1, beta, convertase)<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

U13700 CASP1<br />

125 AF186255 SULT1C2 sulfotransferase family, cytosolic, 1C, member 2<br />

126 U20489 PTPRO protein tyrosine phosphatase, receptor type, O<br />

127 AF201370 MDM2 Mdm2 p53 binding protein homolog (mouse)<br />

128<br />

129<br />

130<br />

M76742<br />

D12502<br />

CEACAM1<br />

CEACAM1<br />

carcinoembryonic antigen-related cell adhesion molecule<br />

1 (biliary glycoprotein)<br />

carcinoembryonic antigen-related cell adhesion molecule<br />

1 (biliary glycoprotein)<br />

serpin peptidase inhibitor, clade E (nexin, plasminogen<br />

activator inhibitor type 1), member 2<br />

AL541302 SERPINE2<br />

131 AI762552 HNRPDL Heterogeneous nuclear ribonucleoprotein D-like<br />

132 AI130920 PABPN1 poly(A) binding protein, nuclear 1<br />

133 AW052179 COL4A5 collagen, type IV, alpha 5 (Alport syndrome)<br />

134 AI572079 SNAI2 snail homolog 2 (Drosophila)<br />

135 AA083478 TRIM22 tripartite motif-containing 22<br />

136 AL050154 KIAA1462 KIAA1462<br />

137<br />

Heterogeneous nuclear ribonucleoprotein D (AU-rich<br />

element RNA binding protein 1, 37kDa)<br />

W74620 HNRNPD<br />

138 AU155515 RPL37A ribosomal protein L37a<br />

139 AW103422 PCBP2 poly(rC) binding protein 2<br />

140 AA554945 --- Transcribed locus<br />

141 AL037167 RABL4 RAB, member of RAS oncogene family-like 4<br />

142 AI912583 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

143<br />

ATP synthase, H+ transporting, mitochondrial F1<br />

BG232034 ATP5C1 complex, gamma polypeptide 1<br />

144 X90579 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

145 X90579 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

146<br />

AI962377<br />

LOC729222<br />

/// PPFIBP1<br />

PTPRF interacting protein, binding protein 1 (liprin beta<br />

1) /// similar to mKIAA1230 protein<br />

147 NM_006417 IFI44 interferon-induced protein 44<br />

148 M23699 SAA1 /// serum amyloid A1 /// serum amyloid A2<br />

102


CHAPTER 3<br />

RESULTS<br />

SAA2<br />

149 AU134977 TncRNA Trophoblast-derived noncoding RNA<br />

150 AF070569 C17orf91 chromosome 17 open reading frame 91<br />

151 AL080207 ABCA12 ATP-binding cassette, sub-family A (ABC1), member 12<br />

152 X83493 FAS Fas (TNF receptor superfamily, member 6)<br />

153 AC004770 FADS3 fatty acid desaturase 3<br />

154 Z70519 FAS Fas (TNF receptor superfamily, member 6)<br />

155 AF107846 GNAS GNAS complex locus<br />

156<br />

collagen, type VII, alpha 1 (epidermolysis bullosa,<br />

dystrophic, dominant and recessive)<br />

L23982 COL7A1<br />

157 AJ276888 MDM2 Mdm2 p53 binding protein homolog (mouse)<br />

158 BE888744 IFIT2 interferon-induced protein with tetratricopeptide repeats 2<br />

159 BE930512 MGC5370 Hypothetical protein MGC5370<br />

160 NM_000064 C3 complement component 3<br />

161 NM_022772 EPS8L2 EPS8-like 2<br />

162 NM_014454 SESN1 sestrin 1<br />

163<br />

THSD1 /// thrombospondin, type I, domain containing 1 ///<br />

NM_018676 THSD1P thrombospondin, type I, domain containing 1 pseudogene<br />

164 NM_004669 CLIC3 chloride intracellular channel 3<br />

165 NM_016321 RHCG Rh family, C glycoprotein<br />

166 NM_015900 PLA1A phospholipase A1 member A<br />

167 NM_022470 ZMAT3 zinc finger, matrin type 3<br />

168 NM_024728 C7orf10 chromosome 7 open reading frame 10<br />

169 NM_017938 FAM70A family with sequence similarity 70, member A<br />

170 NM_017986 GPR172B G protein-coupled receptor 172B<br />

171 AL136861 CRISPLD2 cysteine-rich secretory protein LCCL domain containing 2<br />

172 AF133207 HSPB8 heat shock 22kDa protein 8<br />

173 AL537457 NEFL neurofilament, light polypeptide 68kDa<br />

174 BF055311 NEFL neurofilament, light polypeptide 68kDa<br />

175<br />

HNRNPA1<br />

///<br />

LOC728844<br />

heterogeneous nuclear ribonucleoprotein A1 ///<br />

hypothetical LOC728844<br />

AW450929<br />

176 BF131886 SESN2 sestrin 2<br />

177 AB056106 ABI3BP ABI gene family, member 3 (NESH) binding protein<br />

178 AL136680 GBP3 guanylate binding protein 3<br />

179 AL136826 RGMA RGM domain family, member A<br />

180 AF160477 PVRL4 poliovirus receptor-related 4<br />

181<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

AF113016 MALAT1<br />

182 BF062193 CYorf15B chromosome Y open reading frame 15B<br />

183 AF329088 OSAP ovary-specific acidic protein<br />

184 AF229179 TMEM27 transmembrane protein 27<br />

185<br />

AF132202<br />

MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

103


CHAPTER 3<br />

RESULTS<br />

186 BC006355 PLCD4 phospholipase C, delta 4<br />

187<br />

AI446756 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

188<br />

AF001540 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

189(non-protein coding)<br />

189 BE675516 TncRNA trophoblast-derived noncoding RNA<br />

190 AI042152 TncRNA trophoblast-derived noncoding RNA<br />

191<br />

BG534952 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

192<br />

AW005982 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

193 AL133001 SULF2 sulfatase 2<br />

194 AI952357 MGC5370 hypothetical protein MGC5370<br />

195 AI355441 --- CDNA FLJ26120 fis, clone SYN00419<br />

196 N21643 --- CDNA FLJ39585 fis, clone SKMUS2006633<br />

197 N21426 SYTL2 synaptotagmin-like 2<br />

198<br />

AI601101<br />

FAM84A ///<br />

LOC653602<br />

family with sequence similarity 84, member A ///<br />

hypothetical LOC653602<br />

199 AW006123 FBXO32 F-box protein 32<br />

200 AL039862 FAM84B family with sequence similarity 84, member B<br />

201 AW341649 TP53INP1 tumor protein p53 inducible nuclear protein 1<br />

202 N21030 NRBP2 nuclear receptor binding protein 2<br />

203 AB033041 VANGL2 vang-like 2 (van gogh, Drosophila)<br />

204 N32834 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

205 AV682252 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

206 BE217880 IL7R interleukin 7 receptor<br />

207 BE967311 MCC mutated in colorectal cancers<br />

208 AI565067 KIAA1324 KIAA1324<br />

209 AI188653 MXD1 MAX dimerization protein 1<br />

210<br />

AW117717 AHSA2<br />

AHA1, activator of heat shock 90kDa protein ATPase<br />

homolog 2 (yeast)<br />

211 AI446414 KITLG KIT ligand<br />

212 AL513673 CYGB cytoglobin<br />

213<br />

AI986239 AHSA2<br />

AHA1, activator of heat shock 90kDa protein ATPase<br />

homolog 2 (yeast)<br />

214<br />

W80468 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

215 AA131041 IFIT2 interferon-induced protein with tetratricopeptide repeats 2<br />

216 AI571166 --- CDNA FLJ39306 fis, clone OCBBF2013123<br />

217 AU155361 TncRNA trophoblast-derived noncoding RNA<br />

218<br />

AA524669<br />

LOC1001315<br />

64 hypothetical protein LOC100131564<br />

219 AI343467 --- CDNA FLJ11041 fis, clone PLACE1004405<br />

220 N36085 --- CDNA clone IMAGE:5261213<br />

104


CHAPTER 3<br />

RESULTS<br />

221<br />

LOC643167 RNA binding motif protein 39 /// similar to RNA-binding<br />

BE466173 /// RBM39 region (RNP1, RRM) containing 2<br />

222 BE673226 C15orf52 chromosome 15 open reading frame 52<br />

223 AK024898 --- CDNA: FLJ21245 fis, clone COL01184<br />

224 AI459194 EGR1 Early growth response 1<br />

225 AA632295 STON2 stonin 2<br />

226<br />

AL037917 MALAT1<br />

227 AI817145 PDCD5 Programmed cell death 5<br />

228 BF591483 NTN1 netrin 1<br />

229 AB046848 NOPE neighbor of Punc E11<br />

230<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

Transcribed locus, strongly similar to NP_005768.1 RNA<br />

binding motif protein 6 [Homo sapiens]<br />

AI041522 ---<br />

231 AI953847 RNF144B ring finger 144B<br />

232 BE645119 MGC13057 hypothetical protein MGC13057<br />

233 BE048571 MGC16121 hypothetical protein MGC16121<br />

234 AI927692 --- CDNA FLJ41270 fis, clone BRAMY2036387<br />

235 AA741307 SAMD9 sterile alpha motif domain containing 9<br />

236<br />

Metastasis associated lung adenocarcinoma transcript 1<br />

AI475544 MALAT1 (non-protein coding)<br />

237 AI809749 ORMDL1 ORM1-like 1 (S. cerevisiae)<br />

238 AW071793 MXD1 MAX dimerization protein 1<br />

239<br />

AW007289 ADAMTS7<br />

ADAM metallopeptidase with thrombospondin type 1<br />

motif, 7<br />

240 AI659800 C13orf31 chromosome 13 open reading frame 31<br />

241 AI810764 --- Transcribed locus<br />

242 AA058832 TMEM178 transmembrane protein 178<br />

243 AI559300 SPATA18 spermatogenesis associated 18 homolog (rat)<br />

244 AA741058 ATG16L2 ATG16 autophagy related 16-like 2 (S. cerevisiae)<br />

245 AI075407 IFIT3 interferon-induced protein with tetratricopeptide repeats 3<br />

246 AA760738 --- CDNA FLJ42331 fis, clone TSTOM2000588<br />

247 BG252802 --- Transcribed locus<br />

248 AA902480 MGC5370 hypothetical protein MGC5370<br />

249 AA234096 MGC16121 hypothetical protein MGC16121<br />

250 AI625235 C20orf199 chromosome 20 open reading frame 199<br />

251<br />

Phosphoribosylglycinamide formyltransferase,<br />

phosphoribosylglycinamide synthetase,<br />

phosphoribosylaminoimidazole synthetase<br />

AI207338 GART<br />

252 AI139993 --- Transcribed locus<br />

253<br />

PRP38 pre-mRNA processing factor 38 (yeast) domain<br />

N32872 PRPF38B containing B<br />

254 BE326919 --- Transcribed locus<br />

255 BF510429 CLINT1 clathrin interactor 1<br />

256 AL042483 SPATA18 spermatogenesis associated 18 homolog (rat)<br />

105


CHAPTER 3<br />

RESULTS<br />

257 AI912194 --- Transcribed locus<br />

258<br />

AK024931 GALNT10<br />

UDP-N-acetyl-alpha-D-galactosamine:polypeptide N-<br />

acetylgalactosaminyltransferase 10 (GalNAc-T10)<br />

259 AI698574 TTLL6 tubulin tyrosine ligase-like family, member 6<br />

260 AW262311 --- ---<br />

261<br />

metastasis associated lung adenocarcinoma transcript 1<br />

NM_014086 MALAT1 (non-protein coding)<br />

262 N90870 --- CDNA FLJ38472 fis, clone FEBRA2022148<br />

263 AL137763 GRHL3 grainyhead-like 3 (Drosophila)<br />

264 BE927772 SFRS3 Splicing factor, arginine/serine-rich 3<br />

265 AV703311 SCARB1 scavenger receptor class B, member 1<br />

266 AK000106 --- CDNA FLJ20099 fis, clone COL04544<br />

267 AU156721 PAPPA pregnancy-associated plasma protein A, pappalysin 1<br />

268 AB046817 SYTL2 synaptotagmin-like 2<br />

269 AU155094 ANXA1 Annexin A1<br />

270 AK021804 --- CDNA FLJ38472 fis, clone FEBRA2022148<br />

271 AL034418 SULF2 sulfatase 2<br />

272 AV699657 TncRNA trophoblast-derived noncoding RNA<br />

273 AA828371 VPS13C Vacuolar protein sorting 13 homolog C (S. cerevisiae)<br />

274 AW292765 --- CDNA FLJ42786 fis, clone BRAWH3006761<br />

275 AI270356 --- CDNA FLJ31593 fis, clone NT2RI2002481<br />

276 AI040887 ARHGEF7 Rho guanine nucleotide exchange factor (GEF) 7<br />

277 BF689253 SPOCD1 SPOC domain containing 1<br />

278 AA889628 HNRNPC heterogeneous nuclear ribonucleoprotein C (C1/C2)<br />

279 AL036662 --- ---<br />

280 AA703280 --- ---<br />

281 AW591809 --- CDNA FLJ35091 fis, clone PLACE6005786<br />

282 AI693336 FOXA1 Forkhead box A1<br />

283 AI359676 LOC727770 similar to ankyrin repeat domain 20 family, member A1<br />

284 AI479082 FLJ41484 hypothetical LOC650669<br />

285 AV659198 TncRNA trophoblast-derived noncoding RNA<br />

286 AI307802 ANKRD29 ankyrin repeat domain 29<br />

287 AI521618 TPM1 Tropomyosin 1 (alpha)<br />

288 AI422414 --- Transcribed locus<br />

289 AA876179 --- Transcribed locus<br />

290 AI924046 --- Transcribed locus<br />

291 BF512162 C3orf55 chromosome 3 open reading frame 55<br />

292 AI342246 --- Transcribed locus<br />

293 BF970185 --- Transcribed locus<br />

294 AI743261 FAM98A Family with sequence similarity 98, member A<br />

295 AI686890 --- Transcribed locus<br />

296 BE274992 RNF144B ring finger 144B<br />

297 BF061543 --- Transcribed locus<br />

298 AW172407 --- CDNA FLJ37304 fis, clone BRAMY2016070<br />

106


CHAPTER 3<br />

RESULTS<br />

299 BF244402 FBXO32 F-box protein 32<br />

300 AW962458 --- Transcribed locus<br />

301 AA019641 --- Transcribed locus<br />

302 BG260086 XDH xanthine dehydrogenase<br />

303 AW973232 RNF12 Ring finger protein 12<br />

304 AW970888 --- CDNA clone IMAGE:5314281<br />

305 AI078033 --- Transcribed locus<br />

306<br />

Transcribed locus, strongly similar to XP_001160512.1<br />

PREDICTED: hypothetical protein [Pan troglodytes]<br />

AI291804 ---<br />

307 AI458949 IFNGR1 interferon gamma receptor 1<br />

308 W84667 --- ---<br />

309 AI498610 --- ---<br />

310 AA883831 --- Transcribed locus<br />

311 BF224218 --- Transcribed locus<br />

312 M15330 IL1B interleukin 1, beta<br />

Table 3.1.2B List of Down-regulated Genes: Control Vs 5X2 Gy<br />

S.No Public ID Gene Symbol Gene Title<br />

1 NM_020380 CASC5 cancer susceptibility candidate 5<br />

2 NM_144508 CASC5 cancer susceptibility candidate 5<br />

3<br />

TAF5 RNA polymerase II, TATA box binding protein<br />

(TBP)-associated factor, 100kDa<br />

NM_139052 TAF5<br />

4 NM_001453 FOXC1 forkhead box C1<br />

5 BC032247 RBL1 retinoblastoma-like 1 (p107)<br />

6 BC008680 ST6GAL2 ST6 beta-galactosamide alpha-2,6-sialyltranferase 2<br />

7<br />

sterile alpha motif and leucine zipper containing kinase<br />

AZK<br />

AF465843 ZAK<br />

8 BC026969 WDR67 WD repeat domain 67<br />

9 BC043596 FANCB Fanconi anemia, complementation group B<br />

10 AI917078 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

11 BC041481 FLJ35848 hypothetical protein FLJ35848<br />

12 BC040700 --- Homo sapiens, clone IMAGE:5580334, mRNA<br />

13 AK096748 ZNF519 zinc finger protein 519<br />

14 BC033560 --- CDNA clone IMAGE:4824158<br />

15 NM_006623 PHGDH phosphoglycerate dehydrogenase<br />

16 NM_006739 MCM5 minichromosome maintenance complex component 5<br />

17<br />

18<br />

BE550486<br />

SLC2A3<br />

solute carrier family 2 (facilitated glucose transporter),<br />

member 3<br />

solute carrier family 2 (facilitated glucose transporter),<br />

member 3<br />

NM_006931 SLC2A3<br />

19 BC000192 DHFR dihydrofolate reductase<br />

20 NM_015180 SYNE2 spectrin repeat containing, nuclear envelope 2<br />

107


CHAPTER 3<br />

RESULTS<br />

21 NM_001254 CDC6 cell division cycle 6 homolog (S. cerevisiae)<br />

22<br />

NM_001262 CDKN2C<br />

cyclin-dependent kinase inhibitor 2C (p18, inhibits<br />

CDK4)<br />

23 BF305380 GTSE1 G-2 and S-phase expressed 1<br />

24 NM_004523 KIF11 kinesin family member 11<br />

25<br />

ARHGAP11<br />

NM_014783 A<br />

Rho GTPase activating protein 11A<br />

26 NM_003686 EXO1 exonuclease 1<br />

27 NM_004856 KIF23 kinesin family member 23<br />

28 NM_007086 WDHD1 WD repeat and HMG-box DNA binding protein 1<br />

29 U17105 CCNF cyclin F<br />

30 NM_014264 PLK4 polo-like kinase 4 (Drosophila)<br />

31 AF154830 CPS1 carbamoyl-phosphate synthetase 1, mitochondrial<br />

32 NM_002828 PTPN2 protein tyrosine phosphatase, non-receptor type 2<br />

33 NM_001809 CENPA centromere protein A<br />

34 NM_001673 ASNS asparagine synthetase<br />

35 NM_004153 ORC1L origin recognition complex, subunit 1-like (yeast)<br />

36 NM_013296 GPSM2 G-protein signaling modulator 2 (AGS3-like, C. elegans)<br />

37 AL365505 RBL1 retinoblastoma-like 1 (p107)<br />

38<br />

NM_021992<br />

MGC39900<br />

/// TMSL8 thymosin-like 8 /// thymosin beta15b<br />

39 NM_001274 CHEK1 CHK1 checkpoint homolog (S. pombe)<br />

40 NM_002692 POLE2 polymerase (DNA directed), epsilon 2 (p59 subunit)<br />

41 NM_014645 CEP135 centrosomal protein 135kDa<br />

42 NM_014875 KIF14 kinesin family member 14<br />

43 NM_002530 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

44 NM_015653 RIBC2 RIB43A domain with coiled-coils 2<br />

45 NM_005196 CENPF centromere protein F, 350/400ka (mitosin)<br />

46 U30872 CENPF centromere protein F, 350/400ka (mitosin)<br />

47 AI373676 SMC3 structural maintenance of chromosomes 3<br />

48<br />

AW157094 ID4<br />

inhibitor of DNA binding 4, dominant negative helixloop-helix<br />

protein<br />

49 BG054916 PTCH1 patched homolog 1 (Drosophila)<br />

50<br />

AF225416 SPC25<br />

SPC25, NDC80 kinetochore complex component,<br />

homolog (S. cerevisiae)<br />

51<br />

AF016535 ABCB1<br />

ATP-binding cassette, sub-family B (MDR/TAP), member<br />

1<br />

52<br />

AF016535<br />

ABCB1 ///<br />

ABCB4<br />

ATP-binding cassette, sub-family B (MDR/TAP), member<br />

1 /// ATP-binding cassette, sub-family B (MDR/TAP),<br />

member 4<br />

53 Z25425 NEK2 NIMA (never in mitosis gene a)-related kinase 2<br />

54 BC005832 KIAA0101 KIAA0101<br />

55 BC005995 GINS4 GINS complex subunit 4 (Sld5 homolog)<br />

56 U17074 CDKN2C cyclin-dependent kinase inhibitor 2C (p18, inhibits<br />

108


CHAPTER 3<br />

RESULTS<br />

CDK4)<br />

57 AU152107 MKI67 antigen identified by monoclonal antibody Ki-67<br />

58 AU132185 MKI67 antigen identified by monoclonal antibody Ki-67<br />

59 BF001806 MKI67 antigen identified by monoclonal antibody Ki-67<br />

60 AU147044 MKI67 antigen identified by monoclonal antibody Ki-67<br />

61 AI695017 MTUS1 mitochondrial tumor suppressor 1<br />

62 BE552421 MTUS1 mitochondrial tumor suppressor 1<br />

63 AL096842 MTUS1 mitochondrial tumor suppressor 1<br />

64 D38553 NCAPH non-SMC condensin I complex, subunit H<br />

65 BE045993 OIP5 Opa interacting protein 5<br />

66 T87225 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

67 X95152 BRCA2 breast cancer 2, early onset<br />

68 AL109696 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

69 BF973178 GTSE1 G-2 and S-phase expressed 1<br />

70<br />

solute carrier family 2 (facilitated glucose transporter),<br />

member 3 /// solute carrier family 2 (facilitated glucose<br />

transporter), member 14<br />

SLC2A14 ///<br />

AL110298 SLC2A3<br />

71 AA807529 MCM5 minichromosome maintenance complex component 5<br />

72 AJ251844 MYST3 MYST histone acetyltransferase (monocytic leukemia) 3<br />

73 NM_022346 NCAPG non-SMC condensin I complex, subunit G<br />

74 NM_014109 ATAD2 ATPase family, AAA domain containing 2<br />

75 AF118887 VAV3 vav 3 guanine nucleotide exchange factor<br />

76 NM_012177 FBXO5 F-box protein 5<br />

77 NM_020242 KIF15 kinesin family member 15<br />

78<br />

excision repair cross-complementing rodent repair<br />

NM_017669 ERCC6L deficiency, complementation group 6-like<br />

79 NM_018365 MNS1 meiosis-specific nuclear structural 1<br />

80<br />

NM_018123 ASPM<br />

asp (abnormal spindle) homolog, microcephaly associated<br />

(Drosophila)<br />

81 NM_024680 E2F8 E2F transcription factor 8<br />

82 NM_022488 ATG3 ATG3 autophagy related 3 homolog (S. cerevisiae)<br />

83 NM_018518 MCM10 minichromosome maintenance complex component 10<br />

84 BC001651 CDCA8 cell division cycle associated 8<br />

85 BC003186 GINS2 GINS complex subunit 2 (Psf2 homolog)<br />

86 AF360549 BRIP1 BRCA1 interacting protein C-terminal helicase 1<br />

87 AA886335 CALML4 calmodulin-like 4<br />

88 AW195581 GPSM2 G-protein signaling modulator 2 (AGS3-like, C. elegans)<br />

89 AI859865 MCM4 minichromosome maintenance complex component 4<br />

90<br />

solute carrier family 2 (facilitated glucose transporter),<br />

member 3 /// solute carrier family 2 (facilitated glucose<br />

transporter), member 14<br />

SLC2A14 ///<br />

AA778684 SLC2A3<br />

91 BF684446 AXIN2 axin 2 (conductin, axil)<br />

92 AI925583 ATAD2 ATPase family, AAA domain containing 2<br />

93 AB042719 MCM10 minichromosome maintenance complex component 10<br />

109


CHAPTER 3<br />

RESULTS<br />

94 BC002493 TCF19 transcription factor 19 (SC1)<br />

95 AF177340 CLDN2 claudin 2<br />

96 AL136840 MCM10 minichromosome maintenance complex component 10<br />

97 AY028916 MND1 meiotic nuclear divisions 1 homolog (S. cerevisiae)<br />

98 AV686514 EMP2 epithelial membrane protein 2<br />

99 AL512758 CPLX2 complexin 2<br />

100 BE504653 RIF1 RAP1 interacting factor homolog (yeast)<br />

101 AW003459 SLFN11 schlafen family member 11<br />

102<br />

103<br />

AK024288<br />

regulatory factor X, 2 (influences HLA class II<br />

expression)<br />

RFX2<br />

LOC1001281<br />

91 hypothetical protein LOC100128191<br />

H28597<br />

104 BF248364 CASC5 cancer susceptibility candidate 5<br />

105 AW004016 ST6GAL2 ST6 beta-galactosamide alpha-2,6-sialyltranferase 2<br />

106 AI140305 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

107 BF059556 C18orf54 chromosome 18 open reading frame 54<br />

108 N62196 ZNF367 zinc finger protein 367<br />

109 AW572379 --- CDNA clone IMAGE:5263455<br />

110<br />

hCG_203303<br />

9 hypothetical protein BC006136<br />

AA805700<br />

111 AI638593 C15orf42 chromosome 15 open reading frame 42<br />

112 AA041523 C6orf176 chromosome 6 open reading frame 176<br />

113 BE858063 LOC730631 Hypothetical LOC730631<br />

114 AK024292 SGOL1 Shugoshin-like 1 (S. pombe)<br />

115 BE670098 USP37 ubiquitin specific peptidase 37<br />

116<br />

asp (abnormal spindle) homolog, microcephaly associated<br />

(Drosophila)<br />

AK001380 ASPM<br />

117 AI953354 --- CDNA: FLJ20913 fis, clone ADSE00630<br />

118 AK026197 FBXO5 F-box protein 5<br />

119 AL138828 FAM54A family with sequence similarity 54, member A<br />

120 BG170335 ECT2 epithelial cell transforming sequence 2 oncogene<br />

121 AI868315 PHF17 PHD finger protein 17<br />

122 BF974463 CXorf56 chromosome X open reading frame 56<br />

123 AA740849 ESCO2 establishment of cohesion 1 homolog 2 (S. cerevisiae)<br />

124 AA938184 --- Transcribed locus<br />

125 AW183154 KIF14 kinesin family member 14<br />

126 AW269645 ECT2 Epithelial cell transforming sequence 2 oncogene<br />

127 T90771 SYNC1 Syncoilin, intermediate filament 1<br />

128 BF447038 SP8 Sp8 transcription factor<br />

129 BE218107 EPHA5 EPH receptor A5<br />

130 BE620598 LOC201725 hypothetical protein LOC201725<br />

131 AA224205 CHEK1 CHK1 checkpoint homolog (S. pombe)<br />

132 AA573088 LOC730631 Hypothetical LOC730631<br />

133 AI860012 GAS2L3 Growth arrest-specific 2 like 3<br />

110


CHAPTER 3<br />

RESULTS<br />

134 AL045793 METT10D methyltransferase 10 domain containing<br />

135 R54683 MCM6 minichromosome maintenance complex component 6<br />

136 AI220472 --- Transcribed locus<br />

137 AI928037 RUNDC3B RUN domain containing 3B<br />

138 AA019836 C18orf54 Chromosome 18 open reading frame 54<br />

139 BF666241 RIF1 RAP1 interacting factor homolog (yeast)<br />

140 AI733194 --- Transcribed locus<br />

141 BF516341 --- Transcribed locus<br />

142 AI360832 --- Homo sapiens, clone IMAGE:3660074, mRNA<br />

143 AI684761 SYNE2 spectrin repeat containing, nuclear envelope 2<br />

144 AI924134 --- Transcribed locus<br />

145 AW300380 SYNE1 spectrin repeat containing, nuclear envelope 1<br />

146<br />

sema domain, immunoglobulin domain (Ig), short basic<br />

domain, secreted, (semaphorin) 3A<br />

BF215018 SEMA3A<br />

147 BG283921 C18orf54 chromosome 18 open reading frame 54<br />

148 AI144299 DHFR dihydrofolate reductase<br />

149 AW025529 CALML4 calmodulin-like 4<br />

Table 3.1.2C List of Up-regulated genes: 2Gy Vs 5X2Gy<br />

S.No Public ID Gene Symbol Gene Title<br />

1 M21121 CCL5 chemokine (C-C motif) ligand 5<br />

2 NM_005764 PDZK1IP1 PDZK1 interacting protein 1<br />

3 AF043341 CCL5 chemokine (C-C motif) ligand 5<br />

4 BC039509 LOC643401 hypothetical protein LOC643401<br />

5<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

BE708432 MALAT1<br />

6 BC043411 --- Homo sapiens, clone IMAGE:6155889, mRNA<br />

7 AL832227 BCL8 B-cell CLL/lymphoma 8<br />

8 AL832227 BCL8 B-cell CLL/lymphoma 8<br />

9 AF086134 --- Full length insert cDNA clone ZA88B06<br />

10<br />

serpin peptidase inhibitor, clade E (nexin, plasminogen<br />

activator inhibitor type 1), member 1<br />

BC020765 SERPINE1<br />

11 BC023629 PDLIM7 PDZ and LIM domain 7 (enigma)<br />

12 NM_001613 ACTA2 actin, alpha 2, smooth muscle, aorta<br />

13 AF010314 ENC1 ectodermal-neural cortex (with BTB-like domain)<br />

14 BG491844 JUN jun oncogene<br />

15 NM_002228 JUN jun oncogene<br />

16 NM_002276 KRT19 keratin 19<br />

17<br />

myxovirus (influenza virus) resistance 1, interferoninducible<br />

protein p78 (mouse)<br />

NM_002462 MX1<br />

18 NM_005562 LAMC2 laminin, gamma 2<br />

19 BC002666 GBP1 guanylate binding protein 1, interferon-inducible, 67kDa<br />

111


CHAPTER 3<br />

RESULTS<br />

20 NM_001710 CFB complement factor B<br />

21 AW117498 FOXO1 forkhead box O1<br />

22 NM_004120 GBP2 guanylate binding protein 2, interferon-inducible<br />

23<br />

angiotensinogen (serpin peptidase inhibitor, clade A,<br />

member 8)<br />

NM_000029 AGT<br />

24 NM_001548 IFIT1 interferon-induced protein with tetratricopeptide repeats 1<br />

25 NM_001345 DGKA diacylglycerol kinase, alpha 80kDa<br />

26 NM_005103 FEZ1 fasciculation and elongation protein zeta 1 (zygin I)<br />

27<br />

carbohydrate (N-acetylglucosamine-6-O) sulfotransferase<br />

2<br />

NM_004267 CHST2<br />

28 NM_004585 RARRES3 retinoic acid receptor responder (tazarotene induced) 3<br />

29 NM_006851 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

30 NM_000698 ALOX5 arachidonate 5-lipoxygenase<br />

31<br />

chemokine (C-X-C motif) ligand 1 (melanoma growth<br />

stimulating activity, alpha)<br />

NM_001511 CXCL1<br />

32 NM_004961 GABRE gamma-aminobutyric acid (GABA) A receptor, epsilon<br />

33 NM_002985 CCL5 chemokine (C-C motif) ligand 5<br />

34 NM_001549 IFIT3 interferon-induced protein with tetratricopeptide repeats 3<br />

35 NM_000203 IDUA iduronidase, alpha-L-<br />

36 NM_000576 IL1B interleukin 1, beta<br />

37 AB046692 AOX1 aldehyde oxidase 1<br />

38 NM_001159 AOX1 aldehyde oxidase 1<br />

39 NM_000600 IL6 interleukin 6 (interferon, beta 2)<br />

40 AF026303 SULT1C2 sulfotransferase family, cytosolic, 1C, member 2<br />

41 NM_005101 ISG15 ISG15 ubiquitin-like modifier<br />

42 NM_005658 TRAF1 TNF receptor-associated factor 1<br />

43 NM_001785 CDA cytidine deaminase<br />

44 NM_003733 OASL 2'-5'-oligoadenylate synthetase-like<br />

45 NM_000777 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

46 NM_002185 IL7R interleukin 7 receptor<br />

47 NM_016352 CPA4 carboxypeptidase A4<br />

48<br />

GABBR1 ///<br />

NM_006398 UBD<br />

49 NM_001086 AADAC arylacetamide deacetylase (esterase)<br />

50 NM_002652 PIP prolactin-induced protein<br />

51 NM_015364 LY96 lymphocyte antigen 96<br />

gamma-aminobutyric acid (GABA) B receptor, 1 ///<br />

ubiquitin D<br />

52<br />

egf-like module containing, mucin-like, hormone<br />

NM_001974 EMR1 receptor-like 1<br />

53 NM_001531 MR1 major histocompatibility complex, class I-related<br />

54 NM_013314 BLNK B-cell linker<br />

55<br />

SAA1 ///<br />

NM_030754 SAA2 serum amyloid A1 /// serum amyloid A2<br />

56 AF204231 GOLGA8A golgi autoantigen, golgin subfamily a, 8A<br />

57 BC001743 C7orf44 chromosome 7 open reading frame 44<br />

112


CHAPTER 3<br />

RESULTS<br />

58 AF237813 ABAT 4-aminobutyrate aminotransferase<br />

59 AF237813 ABAT 4-aminobutyrate aminotransferase<br />

60<br />

X16354 CEACAM1<br />

carcinoembryonic antigen-related cell adhesion molecule<br />

1 (biliary glycoprotein)<br />

61<br />

M87507 CASP1<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

62 M15329 IL1A interleukin 1, alpha<br />

63 AF010446 MR1 major histocompatibility complex, class I-related<br />

64 M11734 CSF2 colony stimulating factor 2 (granulocyte-macrophage)<br />

65 M13436 INHBA inhibin, beta A<br />

66 AF063612 OASL 2'-5'-oligoadenylate synthetase-like<br />

67 AF064771 DGKA diacylglycerol kinase, alpha 80kDa<br />

68<br />

U13699 CASP1<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

69<br />

U13700 CASP1<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

70<br />

AF119873 SERPINA1<br />

serpin peptidase inhibitor, clade A (alpha-1 antiproteinase,<br />

antitrypsin), member 1<br />

71 AF186255 SULT1C2 sulfotransferase family, cytosolic, 1C, member 2<br />

72 AF043337 IL8 interleukin 8<br />

73<br />

M76742 CEACAM1<br />

carcinoembryonic antigen-related cell adhesion molecule<br />

1 (biliary glycoprotein)<br />

74 D86983 PXDN peroxidasin homolog (Drosophila)<br />

75<br />

AL541302 SERPINE2<br />

serpin peptidase inhibitor, clade E (nexin, plasminogen<br />

activator inhibitor type 1), member 2<br />

76 AW052179 COL4A5 collagen, type IV, alpha 5 (Alport syndrome)<br />

77 AA083478 TRIM22 tripartite motif-containing 22<br />

78 AA554945 --- Transcribed locus<br />

79 AI337069 RSAD2 radical S-adenosyl methionine domain containing 2<br />

80 AI912583 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

81 X90579 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

82 X90579 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

83 NM_006417 IFI44 interferon-induced protein 44<br />

84<br />

SAA1 ///<br />

M23699 SAA2 serum amyloid A1 /// serum amyloid A2<br />

85 AK000923 SENP3 SUMO1/sentrin/SMT3 specific peptidase 3<br />

86 W46388 SOD2 superoxide dismutase 2, mitochondrial<br />

87 AL080207 ABCA12 ATP-binding cassette, sub-family A (ABC1), member 12<br />

88 BE888744 IFIT2 interferon-induced protein with tetratricopeptide repeats 2<br />

89 NM_000064 C3 complement component 3<br />

90 NM_004669 CLIC3 chloride intracellular channel 3<br />

91 NM_015900 PLA1A phospholipase A1 member A<br />

92 AL136861 CRISPLD2 cysteine-rich secretory protein LCCL domain containing 2<br />

93 AF133207 HSPB8 heat shock 22kDa protein 8<br />

113


CHAPTER 3<br />

RESULTS<br />

94<br />

oligonucleotide/oligosaccharide-binding fold containing<br />

AU157541 OBFC2A 2A<br />

95 AK024680 --- CDNA: FLJ21027 fis, clone CAE07110<br />

96 AB056106 ABI3BP ABI gene family, member 3 (NESH) binding protein<br />

97 AL136680 GBP3 guanylate binding protein 3<br />

98 AF228422 C15orf48 chromosome 15 open reading frame 48<br />

99<br />

AA827878 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

100<br />

AF113016 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

101 AF229179 TMEM27 transmembrane protein 27<br />

102<br />

AF132202 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

103<br />

AI446756 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

104<br />

AF001540 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

105<br />

BG534952 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

106<br />

AW005982 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

107 AI355441 --- CDNA FLJ26120 fis, clone SYN00419<br />

108 N21643 --- CDNA FLJ39585 fis, clone SKMUS2006633<br />

109 N21426 SYTL2 synaptotagmin-like 2<br />

110 AW006123 FBXO32 F-box protein 32<br />

111 N32834 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

112 AV682252 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

113 BE967311 MCC mutated in colorectal cancers<br />

114 AL513673 CYGB cytoglobin<br />

115<br />

W80468 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

116 AA131041 IFIT2 interferon-induced protein with tetratricopeptide repeats 2<br />

117 AU155361 TncRNA trophoblast-derived noncoding RNA<br />

118<br />

AA524669<br />

LOC1001315<br />

64 hypothetical protein LOC100131564<br />

119 AI343467 --- CDNA FLJ11041 fis, clone PLACE1004405<br />

120 BE673226 C15orf52 chromosome 15 open reading frame 52<br />

121 AI459194 EGR1 Early growth response 1<br />

122<br />

AL037917 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

123 BF591483 NTN1 netrin 1<br />

124 AI953847 RNF144B ring finger 144B<br />

125 BE048571 MGC16121 hypothetical protein MGC16121<br />

126 AA741307 SAMD9 sterile alpha motif domain containing 9<br />

114


CHAPTER 3<br />

RESULTS<br />

127 AI809749 ORMDL1 ORM1-like 1 (S. cerevisiae)<br />

128 AI810764 --- Transcribed locus<br />

129 AA058832 TMEM178 transmembrane protein 178<br />

130 AI559300 SPATA18 spermatogenesis associated 18 homolog (rat)<br />

131 AI075407 IFIT3 interferon-induced protein with tetratricopeptide repeats 3<br />

132 AA760738 --- CDNA FLJ42331 fis, clone TSTOM2000588<br />

133 BG252802 --- Transcribed locus<br />

134 AA234096 MGC16121 hypothetical protein MGC16121<br />

135 AI625235 C20orf199 chromosome 20 open reading frame 199<br />

136 BE669858 SAMD9L sterile alpha motif domain containing 9-like<br />

137<br />

Phosphoribosylglycinamide formyltransferase,<br />

phosphoribosylglycinamide synthetase,<br />

phosphoribosylaminoimidazole synthetase<br />

AI207338 GART<br />

138 BF510429 CLINT1 clathrin interactor 1<br />

139 AL042483 SPATA18 spermatogenesis associated 18 homolog (rat)<br />

140 AW003173 STC1 Stanniocalcin 1<br />

141 AW014593 GBP1 guanylate binding protein 1, interferon-inducible, 67kDa<br />

142 AW262311 --- ---<br />

143<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

NM_014086 MALAT1<br />

144 AK000106 --- CDNA FLJ20099 fis, clone COL04544<br />

145 AU144005 --- CDNA FLJ11397 fis, clone HEMBA1000622<br />

146 AB046817 SYTL2 synaptotagmin-like 2<br />

147 AU155094 ANXA1 Annexin A1<br />

148<br />

oligonucleotide/oligosaccharide-binding fold containing<br />

2A<br />

AV734843 OBFC2A<br />

149 AK021804 --- CDNA FLJ38472 fis, clone FEBRA2022148<br />

150 AV699657 TncRNA trophoblast-derived noncoding RNA<br />

151 AW292765 --- CDNA FLJ42786 fis, clone BRAWH3006761<br />

152 BF689253 SPOCD1 SPOC domain containing 1<br />

153 AL036662 --- ---<br />

154 H09245 WWC1 WW and C2 domain containing 1<br />

155 AV659198 TncRNA trophoblast-derived noncoding RNA<br />

156 AI521618 TPM1 Tropomyosin 1 (alpha)<br />

157 AI342246 --- Transcribed locus<br />

158 AW954199 --- Transcribed locus<br />

159 BF061543 --- Transcribed locus<br />

160 AW172407 --- CDNA FLJ37304 fis, clone BRAMY2016070<br />

161 BF244402 FBXO32 F-box protein 32<br />

162 AW962458 --- Transcribed locus<br />

163 AA019641 --- Transcribed locus<br />

164 BG260086 XDH xanthine dehydrogenase<br />

165 BE877420 --- Transcribed locus<br />

166 AI078033 --- Transcribed locus<br />

115


CHAPTER 3<br />

RESULTS<br />

167 BG026159 --- CDNA FLJ42225 fis, clone THYMU2040427<br />

168 AI986192 SERPINB9 serpin peptidase inhibitor, clade B (ovalbumin), member 9<br />

169 AI458949 IFNGR1 interferon gamma receptor 1<br />

170 AI498610 --- ---<br />

171 BF224218 --- Transcribed locus<br />

172 M15330 IL1B interleukin 1, beta<br />

Table 3.1.2D List of Down-regulated genes: 2Gy Vs 5X2Gy<br />

S.No Public ID Gene Symbol Gene Title<br />

1 BC032247 RBL1 retinoblastoma-like 1 (p107)<br />

2 BC008680 ST6GAL2 ST6 beta-galactosamide alpha-2,6-sialyltranferase 2<br />

3 BC026969 WDR67 WD repeat domain 67<br />

4 BC043596 FANCB Fanconi anemia, complementation group B<br />

5 AI917078 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

6 BG387892 RBL1 retinoblastoma-like 1 (p107)<br />

7 AK021715 --- CDNA FLJ11653 fis, clone HEMBA1004538<br />

8 BC040700 --- Homo sapiens, clone IMAGE:5580334, mRNA<br />

9 AK096748 ZNF519 zinc finger protein 519<br />

10 NM_006623 PHGDH phosphoglycerate dehydrogenase<br />

11 NM_006739 MCM5 minichromosome maintenance complex component 5<br />

12 NM_015180 SYNE2 spectrin repeat containing, nuclear envelope 2<br />

13<br />

cyclin-dependent kinase inhibitor 2C (p18, inhibits<br />

CDK4)<br />

NM_001262 CDKN2C<br />

14 BE535746 REEP1 receptor accessory protein 1<br />

15 NM_003609 HIRIP3 HIRA interacting protein 3<br />

16 NM_003686 EXO1 exonuclease 1<br />

17 NM_007086 WDHD1 WD repeat and HMG-box DNA binding protein 1<br />

18 NM_001673 ASNS asparagine synthetase<br />

19<br />

MGC39900<br />

NM_021992 /// TMSL8 thymosin-like 8 /// thymosin beta15b<br />

20 NM_002692 POLE2 polymerase (DNA directed), epsilon 2 (p59 subunit)<br />

21 NM_015653 RIBC2 RIB43A domain with coiled-coils 2<br />

22<br />

SPC25, NDC80 kinetochore complex component,<br />

AF225416 SPC25 homolog (S. cerevisiae)<br />

23 Z25425 NEK2 NIMA (never in mitosis gene a)-related kinase 2<br />

24 BC005995 GINS4 GINS complex subunit 4 (Sld5 homolog)<br />

25 AU152107 MKI67 antigen identified by monoclonal antibody Ki-67<br />

26 AU132185 MKI67 antigen identified by monoclonal antibody Ki-67<br />

27 BF001806 MKI67 antigen identified by monoclonal antibody Ki-67<br />

28 AI695017 MTUS1 mitochondrial tumor suppressor 1<br />

29 AL096842 MTUS1 mitochondrial tumor suppressor 1<br />

30 T87225 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

116


CHAPTER 3<br />

RESULTS<br />

31 BE615699 UNC84A unc-84 homolog A (C. elegans)<br />

32 S76476 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

33 AL109696 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

34<br />

defective in sister chromatid cohesion 1 homolog (S.<br />

cerevisiae)<br />

NM_024094 DSCC1<br />

35 NM_018365 MNS1 meiosis-specific nuclear structural 1<br />

36 NM_024749 VASH2 vasohibin 2<br />

37 NM_024680 E2F8 E2F transcription factor 8<br />

38 BC003186 GINS2 GINS complex subunit 2 (Psf2 homolog)<br />

39 AA886335 CALML4 calmodulin-like 4<br />

40 BF684446 AXIN2 axin 2 (conductin, axil)<br />

41 AB042719 MCM10 minichromosome maintenance complex component 10<br />

42 BC002493 TCF19 transcription factor 19 (SC1)<br />

43 AF177340 CLDN2 claudin 2<br />

44 AY028916 MND1 meiotic nuclear divisions 1 homolog (S. cerevisiae)<br />

45 AV686514 EMP2 epithelial membrane protein 2<br />

46 AL512758 CPLX2 complexin 2<br />

47 AI092930 WDR20 WD repeat domain 20<br />

48 AW004016 ST6GAL2 ST6 beta-galactosamide alpha-2,6-sialyltranferase 2<br />

49 AA406603 CLCC1 Chloride channel CLIC-like 1<br />

50 AI823645 PRIMA1 proline rich membrane anchor 1<br />

51 AK024292 SGOL1 Shugoshin-like 1 (S. pombe)<br />

52 BE670098 USP37 ubiquitin specific peptidase 37<br />

53 AI953354 --- CDNA: FLJ20913 fis, clone ADSE00630<br />

54 AA938184 --- Transcribed locus<br />

55 AW269645 ECT2 Epithelial cell transforming sequence 2 oncogene<br />

56 AL045793 METT10D methyltransferase 10 domain containing<br />

57 R54683 MCM6 minichromosome maintenance complex component 6<br />

58 AA019836 C18orf54 Chromosome 18 open reading frame 54<br />

59 AI924134 --- Transcribed locus<br />

60 BG283921 C18orf54 chromosome 18 open reading frame 54<br />

61 AI144299 DHFR dihydrofolate reductase<br />

62 AW025529 CALML4 calmodulin-like 4<br />

Table 3.1.2E List of Up-regulated genes: 10Gy Vs 5X2Gy<br />

S.No Public ID Gene Symbol Gene Title<br />

1 M21121 CCL5 chemokine (C-C motif) ligand 5<br />

2 NM_153838 GPR115 G protein-coupled receptor 115<br />

3 NM_005764 PDZK1IP1 PDZK1 interacting protein 1<br />

4 AF043341 CCL5 chemokine (C-C motif) ligand 5<br />

5 CA428769 RTN4 reticulon 4<br />

6 AW079553 --- CDNA FLJ41020 fis, clone UTERU2019163<br />

117


CHAPTER 3<br />

RESULTS<br />

7 AA580691 RBM25 RNA binding motif protein 25<br />

8 AK055254 --- CDNA FLJ30692 fis, clone FCBBF2000677<br />

9 BQ944989 STRAP Serine/threonine kinase receptor associated protein<br />

10<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

BE708432 MALAT1<br />

11 AL832227 BCL8 B-cell CLL/lymphoma 8<br />

12 AL832227 BCL8 B-cell CLL/lymphoma 8<br />

13 AF086134 --- Full length insert cDNA clone ZA88B06<br />

14 BI868311 ARF1 ADP-ribosylation factor 1<br />

15 AL832409 TGIF1 TGFB-induced factor homeobox 1<br />

16<br />

serpin peptidase inhibitor, clade E (nexin, plasminogen<br />

activator inhibitor type 1), member 1<br />

BC020765 SERPINE1<br />

17 BC023629 PDLIM7 PDZ and LIM domain 7 (enigma)<br />

18 BF111821 WSB1 WD repeat and SOCS box-containing 1<br />

19 AF010314 ENC1 ectodermal-neural cortex (with BTB-like domain)<br />

20 NM_003633 ENC1 ectodermal-neural cortex (with BTB-like domain)<br />

21 BG491844 JUN jun oncogene<br />

22 NM_002228 JUN jun oncogene<br />

23 NM_002276 KRT19 keratin 19<br />

24 NM_005562 LAMC2 laminin, gamma 2<br />

25 BC002666 GBP1 guanylate binding protein 1, interferon-inducible, 67kDa<br />

26 NM_001710 CFB complement factor B<br />

27 NM_006290 TNFAIP3 tumor necrosis factor, alpha-induced protein 3<br />

28 NM_001548 IFIT1 interferon-induced protein with tetratricopeptide repeats 1<br />

29 AF097493 GLS glutaminase<br />

30 NM_014905 GLS glutaminase<br />

31 NM_001345 DGKA diacylglycerol kinase, alpha 80kDa<br />

32<br />

carbohydrate (N-acetylglucosamine-6-O) sulfotransferase<br />

NM_004267 CHST2 2<br />

33 NM_004585 RARRES3 retinoic acid receptor responder (tazarotene induced) 3<br />

34 NM_006851 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

35<br />

NM_001511 CXCL1<br />

chemokine (C-X-C motif) ligand 1 (melanoma growth<br />

stimulating activity, alpha)<br />

36 NM_004961 GABRE gamma-aminobutyric acid (GABA) A receptor, epsilon<br />

37 NM_003155 STC1 stanniocalcin 1<br />

38 NM_002985 CCL5 chemokine (C-C motif) ligand 5<br />

39 NM_001549 IFIT3 interferon-induced protein with tetratricopeptide repeats 3<br />

40 NM_006994 BTN3A3 butyrophilin, subfamily 3, member A3<br />

41 NM_006307 SRPX sushi-repeat-containing protein, X-linked<br />

42 NM_000203 IDUA iduronidase, alpha-L-<br />

43 NM_000576 IL1B interleukin 1, beta<br />

44 AB046692 AOX1 aldehyde oxidase 1<br />

45 NM_001159 AOX1 aldehyde oxidase 1<br />

46 NM_000600 IL6 interleukin 6 (interferon, beta 2)<br />

118


CHAPTER 3<br />

RESULTS<br />

47 AF026303 SULT1C2 sulfotransferase family, cytosolic, 1C, member 2<br />

48 NM_005101 ISG15 ISG15 ubiquitin-like modifier<br />

49 NM_005658 TRAF1 TNF receptor-associated factor 1<br />

50 NM_001785 CDA cytidine deaminase<br />

51 NM_000756 CRH corticotropin releasing hormone<br />

52 NM_003733 OASL 2'-5'-oligoadenylate synthetase-like<br />

53 NM_000777 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

54 NM_002185 IL7R interleukin 7 receptor<br />

55 NM_016352 CPA4 carboxypeptidase A4<br />

56 NM_005393 PLXNB3 plexin B3<br />

57 NM_001086 AADAC arylacetamide deacetylase (esterase)<br />

58 NM_002852 PTX3 pentraxin-related gene, rapidly induced by IL-1 beta<br />

59 NM_002652 PIP prolactin-induced protein<br />

60 NM_015364 LY96 lymphocyte antigen 96<br />

61<br />

FAM36A /// homeobox A7 /// family with sequence similarity 36,<br />

NM_006896 HOXA7 member A<br />

62<br />

egf-like module containing, mucin-like, hormone<br />

NM_001974 EMR1 receptor-like 1<br />

63 NM_013314 BLNK B-cell linker<br />

64 NM_002090 CXCL3 chemokine (C-X-C motif) ligand 3<br />

65<br />

SAA1 ///<br />

NM_030754 SAA2 serum amyloid A1 /// serum amyloid A2<br />

66 AF204231 GOLGA8A golgi autoantigen, golgin subfamily a, 8A<br />

67 BC001743 C7orf44 chromosome 7 open reading frame 44<br />

68 AF237813 ABAT 4-aminobutyrate aminotransferase<br />

69 AF237813 ABAT 4-aminobutyrate aminotransferase<br />

70<br />

X16354 CEACAM1<br />

carcinoembryonic antigen-related cell adhesion molecule<br />

1 (biliary glycoprotein)<br />

71 M57731 CXCL2 chemokine (C-X-C motif) ligand 2<br />

72<br />

M87507 CASP1<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

73 D26121 SF1 splicing factor 1<br />

74 M11734 CSF2 colony stimulating factor 2 (granulocyte-macrophage)<br />

75<br />

AF164622<br />

GOLGA8A<br />

///<br />

GOLGA8B<br />

golgi autoantigen, golgin subfamily a, 8A /// golgi<br />

autoantigen, golgin subfamily a, 8B<br />

76 M13436 INHBA inhibin, beta A<br />

77 AF119863 MEG3 maternally expressed 3<br />

78 AF063612 OASL 2'-5'-oligoadenylate synthetase-like<br />

79 AF064771 DGKA diacylglycerol kinase, alpha 80kDa<br />

80<br />

U13699 CASP1<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

81<br />

U13700 CASP1<br />

caspase 1, apoptosis-related cysteine peptidase<br />

(interleukin 1, beta, convertase)<br />

119


CHAPTER 3<br />

RESULTS<br />

82 AF186255 SULT1C2 sulfotransferase family, cytosolic, 1C, member 2<br />

83 AF043337 IL8 interleukin 8<br />

84 U20489 PTPRO protein tyrosine phosphatase, receptor type, O<br />

85<br />

carcinoembryonic antigen-related cell adhesion molecule<br />

1 (biliary glycoprotein)<br />

M76742 CEACAM1<br />

86 D86983 PXDN peroxidasin homolog (Drosophila)<br />

87 AW052179 COL4A5 collagen, type IV, alpha 5 (Alport syndrome)<br />

88 BE327172 JUN Jun oncogene<br />

89 AA083478 TRIM22 tripartite motif-containing 22<br />

90 AW103422 PCBP2 poly(rC) binding protein 2<br />

91 AA554945 --- Transcribed locus<br />

92 AI912583 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

93 X90579 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

94 X90579 CYP3A5 cytochrome P450, family 3, subfamily A, polypeptide 5<br />

95<br />

LOC729222 PTPRF interacting protein, binding protein 1 (liprin beta<br />

AI962377 /// PPFIBP1 1) /// similar to mKIAA1230 protein<br />

96 NM_006417 IFI44 interferon-induced protein 44<br />

97<br />

SAA1 ///<br />

M23699 SAA2 serum amyloid A1 /// serum amyloid A2<br />

98 U79273 EIF4A1 Eukaryotic translation initiation factor 4A, isoform 1<br />

99 AL080207 ABCA12 ATP-binding cassette, sub-family A (ABC1), member 12<br />

100 BE888744 IFIT2 interferon-induced protein with tetratricopeptide repeats 2<br />

101<br />

aldo-keto reductase family 1, member C1 (dihydrodiol<br />

dehydrogenase 1; 20-alpha (3-alpha)-hydroxysteroid<br />

dehydrogenase)<br />

BF508244 AKR1C1<br />

102 NM_000064 C3 complement component 3<br />

103 NM_021730 NLRP1 NLR family, pyrin domain containing 1<br />

104 NM_017734 PALMD palmdelphin<br />

105 NM_004669 CLIC3 chloride intracellular channel 3<br />

106 NM_016321 RHCG Rh family, C glycoprotein<br />

107 NM_015900 PLA1A phospholipase A1 member A<br />

108 NM_018194 HHAT hedgehog acyltransferase<br />

109 AL136861 CRISPLD2 cysteine-rich secretory protein LCCL domain containing 2<br />

110<br />

oligonucleotide/oligosaccharide-binding fold containing<br />

AU157541 OBFC2A 2A<br />

111 AK024680 --- CDNA: FLJ21027 fis, clone CAE07110<br />

112<br />

BE646573 NFKBIZ<br />

nuclear factor of kappa light polypeptide gene enhancer in<br />

B-cells inhibitor, zeta<br />

113<br />

AB037925 NFKBIZ<br />

nuclear factor of kappa light polypeptide gene enhancer in<br />

B-cells inhibitor, zeta<br />

114 AB056106 ABI3BP ABI gene family, member 3 (NESH) binding protein<br />

115 AL136680 GBP3 guanylate binding protein 3<br />

116<br />

AF113016 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

120


CHAPTER 3<br />

RESULTS<br />

117 BF062193 CYorf15B chromosome Y open reading frame 15B<br />

118 AF229179 TMEM27 transmembrane protein 27<br />

119 AF317392 BCOR BCL6 co-repressor<br />

120<br />

AF132202 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

121 BC006355 PLCD4 phospholipase C, delta 4<br />

122<br />

AI446756 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

123<br />

AF001540 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

124 AI042152 TncRNA trophoblast-derived noncoding RNA<br />

125<br />

BG534952 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

126<br />

AW005982 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

127 AI355441 --- CDNA FLJ26120 fis, clone SYN00419<br />

128 N21643 --- CDNA FLJ39585 fis, clone SKMUS2006633<br />

129 N21426 SYTL2 synaptotagmin-like 2<br />

130 AW006123 FBXO32 F-box protein 32<br />

131 N21030 NRBP2 nuclear receptor binding protein 2<br />

132 N32834 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

133 AV682252 GLIPR1 GLI pathogenesis-related 1 (glioma)<br />

134 BE217880 IL7R interleukin 7 receptor<br />

135 BE967311 MCC mutated in colorectal cancers<br />

136 BE966604 SAMD9L sterile alpha motif domain containing 9-like<br />

137 AL513673 CYGB cytoglobin<br />

138<br />

AI986239 AHSA2<br />

AHA1, activator of heat shock 90kDa protein ATPase<br />

homolog 2 (yeast)<br />

139<br />

W80468 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

140 AA131041 IFIT2 interferon-induced protein with tetratricopeptide repeats 2<br />

141 AU155361 TncRNA trophoblast-derived noncoding RNA<br />

142<br />

AA524669<br />

LOC1001315<br />

64 hypothetical protein LOC100131564<br />

143 AI343467 --- CDNA FLJ11041 fis, clone PLACE1004405<br />

144 BE673226 C15orf52 chromosome 15 open reading frame 52<br />

145 AI459194 EGR1 Early growth response 1<br />

146 AV728999 MGC16121 hypothetical protein MGC16121<br />

147<br />

AL037917 MALAT1<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

148 BF591483 NTN1 netrin 1<br />

149<br />

AI041522 ---<br />

Transcribed locus, strongly similar to NP_005768.1 RNA<br />

binding motif protein 6 [Homo sapiens]<br />

150 BE048571 MGC16121 hypothetical protein MGC16121<br />

121


CHAPTER 3<br />

RESULTS<br />

151 AA741307 SAMD9 sterile alpha motif domain containing 9<br />

152<br />

AI475544 MALAT1<br />

Metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

153 AI809749 ORMDL1 ORM1-like 1 (S. cerevisiae)<br />

154 AI810764 --- Transcribed locus<br />

155 AA058832 TMEM178 transmembrane protein 178<br />

156 AI075407 IFIT3 interferon-induced protein with tetratricopeptide repeats 3<br />

157 BG252802 --- Transcribed locus<br />

158 AA234096 MGC16121 hypothetical protein MGC16121<br />

159 AI129626 DCLK1 Doublecortin-like kinase 1<br />

160 AI625235 C20orf199 chromosome 20 open reading frame 199<br />

161<br />

Phosphoribosylglycinamide formyltransferase,<br />

phosphoribosylglycinamide synthetase,<br />

phosphoribosylaminoimidazole synthetase<br />

AI207338 GART<br />

162 BF510429 CLINT1 clathrin interactor 1<br />

163 AL042483 SPATA18 spermatogenesis associated 18 homolog (rat)<br />

164 AW003173 STC1 Stanniocalcin 1<br />

165<br />

UDP-N-acetyl-alpha-D-galactosamine:polypeptide N-<br />

acetylgalactosaminyltransferase 10 (GalNAc-T10)<br />

AK024931 GALNT10<br />

166 AI698574 TTLL6 tubulin tyrosine ligase-like family, member 6<br />

167 AI038059 DIO2 deiodinase, iodothyronine, type II<br />

168 AW014593 GBP1 guanylate binding protein 1, interferon-inducible, 67kDa<br />

169 AW262311 --- ---<br />

170<br />

metastasis associated lung adenocarcinoma transcript 1<br />

(non-protein coding)<br />

NM_014086 MALAT1<br />

171 N90870 --- CDNA FLJ38472 fis, clone FEBRA2022148<br />

172 AV703311 SCARB1 scavenger receptor class B, member 1<br />

173 AK000106 --- CDNA FLJ20099 fis, clone COL04544<br />

174 AU156721 PAPPA pregnancy-associated plasma protein A, pappalysin 1<br />

175 AB046817 SYTL2 synaptotagmin-like 2<br />

176 AU155094 ANXA1 Annexin A1<br />

177<br />

oligonucleotide/oligosaccharide-binding fold containing<br />

AV734843 OBFC2A 2A<br />

178 AK021804 --- CDNA FLJ38472 fis, clone FEBRA2022148<br />

179 AK026869 TMEM43 Transmembrane protein 43<br />

180<br />

AK025151<br />

hCG_200366<br />

3 hCG2003663<br />

181 AV699657 TncRNA trophoblast-derived noncoding RNA<br />

182 AI972661 --- CDNA FLJ39413 fis, clone PLACE6015729<br />

183 BF591288 --- Transcribed locus<br />

184 AW292765 --- CDNA FLJ42786 fis, clone BRAWH3006761<br />

185 BF689253 SPOCD1 SPOC domain containing 1<br />

186<br />

AV713062<br />

MAP3K8<br />

Mitogen-activated protein kinase kinase kinase 8 ///<br />

CDNA clone IMAGE:4689481<br />

122


CHAPTER 3<br />

RESULTS<br />

187 AL036662 --- ---<br />

188 AW591809 --- CDNA FLJ35091 fis, clone PLACE6005786<br />

189 BE222109 --- Transcribed locus<br />

190 AI693336 FOXA1 Forkhead box A1<br />

191 AV659198 TncRNA trophoblast-derived noncoding RNA<br />

192 AW015521 CAT Catalase<br />

193 AI521618 TPM1 Tropomyosin 1 (alpha)<br />

194 AA644178 IRS1 Insulin receptor substrate 1<br />

195 AI342246 --- Transcribed locus<br />

196 AW963634 --- Transcribed locus<br />

197 AW954199 --- Transcribed locus<br />

198 BF061543 --- Transcribed locus<br />

199 AW962458 --- Transcribed locus<br />

200 AA019641 --- Transcribed locus<br />

201 BG260086 XDH xanthine dehydrogenase<br />

202 BE877420 --- Transcribed locus<br />

203 AW973232 RNF12 Ring finger protein 12<br />

204 AI078033 --- Transcribed locus<br />

205 AI791820 --- Transcribed locus<br />

206<br />

Transcribed locus, strongly similar to XP_001160512.1<br />

PREDICTED: hypothetical protein [Pan troglodytes]<br />

AI291804 ---<br />

207 AI458949 IFNGR1 interferon gamma receptor 1<br />

208 AI498610 --- ---<br />

209 AA167167 --- ---<br />

210 AW292830 --- Transcribed locus<br />

211 M15330 IL1B interleukin 1, beta<br />

Table 3.1.2F List of Down-regulated genes of 10Gy Vs 5X2Gy<br />

S.No Public ID Gene Symbol Gene Title<br />

1 BC032247 RBL1 retinoblastoma-like 1 (p107)<br />

2 BC008680 ST6GAL2 ST6 beta-galactosamide alpha-2,6-sialyltranferase 2<br />

3 BC026969 WDR67 WD repeat domain 67<br />

4 BC040700 --- Homo sapiens, clone IMAGE:5580334, mRNA<br />

5 AK096748 ZNF519 zinc finger protein 519<br />

6 NM_006623 PHGDH phosphoglycerate dehydrogenase<br />

7 BC000192 DHFR dihydrofolate reductase<br />

8 NM_000791 DHFR dihydrofolate reductase<br />

9 NM_015180 SYNE2 spectrin repeat containing, nuclear envelope 2<br />

10 NM_002358 MAD2L1 MAD2 mitotic arrest deficient-like 1 (yeast)<br />

11 U77949 CDC6 cell division cycle 6 homolog (S. cerevisiae)<br />

12 NM_001254 CDC6 cell division cycle 6 homolog (S. cerevisiae)<br />

13 NM_001262 CDKN2C cyclin-dependent kinase inhibitor 2C (p18, inhibits<br />

123


CHAPTER 3<br />

RESULTS<br />

CDK4)<br />

14 NM_003686 EXO1 exonuclease 1<br />

15 AW772140 WDHD1 WD repeat and HMG-box DNA binding protein 1<br />

16 NM_007086 WDHD1 WD repeat and HMG-box DNA binding protein 1<br />

17 NM_002828 PTPN2 protein tyrosine phosphatase, non-receptor type 2<br />

18 NM_001673 ASNS asparagine synthetase<br />

19 AL365505 RBL1 retinoblastoma-like 1 (p107)<br />

20<br />

MGC39900<br />

/// TMSL8 thymosin-like 8 /// thymosin beta15b<br />

NM_021992<br />

21 NM_024908 WDR76 WD repeat domain 76<br />

22 NM_002692 POLE2 polymerase (DNA directed), epsilon 2 (p59 subunit)<br />

23 NM_015653 RIBC2 RIB43A domain with coiled-coils 2<br />

24<br />

acidic (leucine-rich) nuclear phosphoprotein 32 family,<br />

member E<br />

NM_030920 ANP32E<br />

25 BC005832 KIAA0101 KIAA0101<br />

26 BC005995 GINS4 GINS complex subunit 4 (Sld5 homolog)<br />

27 AU132185 MKI67 antigen identified by monoclonal antibody Ki-67<br />

28 BF001806 MKI67 antigen identified by monoclonal antibody Ki-67<br />

29 AI695017 MTUS1 mitochondrial tumor suppressor 1<br />

30 AL096842 MTUS1 mitochondrial tumor suppressor 1<br />

31 T87225 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

32<br />

PDS5, regulator of cohesion maintenance, homolog A (S.<br />

cerevisiae)<br />

AW991219 PDS5A<br />

33 BE615699 UNC84A unc-84 homolog A (C. elegans)<br />

34 X95152 BRCA2 breast cancer 2, early onset<br />

35 AL109696 NTRK3 neurotrophic tyrosine kinase, receptor, type 3<br />

36<br />

37<br />

AA905286<br />

DLEU2 ///<br />

DLEU2L<br />

deleted in lymphocytic leukemia, 2 /// deleted in<br />

lymphocytic leukemia 2-like<br />

defective in sister chromatid cohesion 1 homolog (S.<br />

cerevisiae)<br />

NM_024094 DSCC1<br />

38 NM_018365 MNS1 meiosis-specific nuclear structural 1<br />

39 NM_024680 E2F8 E2F transcription factor 8<br />

40 NM_022488 ATG3 ATG3 autophagy related 3 homolog (S. cerevisiae)<br />

41 NM_018518 MCM10 minichromosome maintenance complex component 10<br />

42 AF360549 BRIP1 BRCA1 interacting protein C-terminal helicase 1<br />

43 AA886335 CALML4 calmodulin-like 4<br />

44 AI859865 MCM4 minichromosome maintenance complex component 4<br />

45 AI859865 MCM4 minichromosome maintenance complex component 4<br />

46 AI925583 ATAD2 ATPase family, AAA domain containing 2<br />

47 BC005400 CENPK centromere protein K<br />

48 AB042719 MCM10 minichromosome maintenance complex component 10<br />

49 BC002493 TCF19 transcription factor 19 (SC1)<br />

50 AY028916 MND1 meiotic nuclear divisions 1 homolog (S. cerevisiae)<br />

51 BC001104 NUPL1 nucleoporin like 1<br />

124


CHAPTER 3<br />

RESULTS<br />

52 AV686514 EMP2 epithelial membrane protein 2<br />

53 AL512758 CPLX2 complexin 2<br />

54 AI889959 HELLS helicase, lymphoid-specific<br />

55 AI341146 E2F7 E2F transcription factor 7<br />

56 AW004016 ST6GAL2 ST6 beta-galactosamide alpha-2,6-sialyltranferase 2<br />

57 AI754871 SPATA5 spermatogenesis associated 5<br />

58 BF059556 C18orf54 chromosome 18 open reading frame 54<br />

59 AI823645 PRIMA1 proline rich membrane anchor 1<br />

60 AW014743 --- Transcribed locus<br />

61 AI654224 ALDH1L2 Aldehyde dehydrogenase 1 family, member L2<br />

62 BE670098 USP37 ubiquitin specific peptidase 37<br />

63 AL117589 KIF26A kinesin family member 26A<br />

64 AI961235 VASH2 vasohibin 2<br />

65 BF974463 CXorf56 chromosome X open reading frame 56<br />

66 AA938184 --- Transcribed locus<br />

67 AA224205 CHEK1 CHK1 checkpoint homolog (S. pombe)<br />

68 AL045793 METT10D methyltransferase 10 domain containing<br />

69 R54683 MCM6 minichromosome maintenance complex component 6<br />

70 AI220472 --- Transcribed locus<br />

71 AA019836 C18orf54 Chromosome 18 open reading frame 54<br />

72 BF516341 --- Transcribed locus<br />

73 AW340004 NARG2 NMDA receptor regulated 2<br />

74 AI360832 --- Homo sapiens, clone IMAGE:3660074, mRNA<br />

75 AI024029 ZNF789 Zinc finger protein 789<br />

76 BG283921 C18orf54 chromosome 18 open reading frame 54<br />

77 AI144299 DHFR dihydrofolate reductase<br />

78 AW025529 CALML4 calmodulin-like 4<br />

3.1.3 Validation of Microarray data.<br />

In order to verify the microarray data, semi-quantitative RT-PCR was performed for four genes<br />

from control and fractionated group of A549 cells. Four genes were CDKN1A (p21),<br />

GADD45α, Toll-like receptor 3 (TLR3) and Ferredoxin reductase (FDXR). These results<br />

corresponded very well with those for the microarray data for all four genes, strongly supporting<br />

the reliability and rationale of the strategy (Fig. 3.1.3A & B).<br />

125


CHAPTER 3<br />

RESULTS<br />

Fig. 3.1.3 Validation of microarray data by semi-quantitative RT-PCR. (A) Fold changes of<br />

genes as determined by microarray analysis. (B) Fold changes of genes as determined by RT-<br />

PCR.<br />

3.1.4 Activation of repair related genes<br />

Microarray data had shown that DNA repair pathways were up-regulated in A549 cells that had<br />

been exposed to fractionated irradiation; it was of interest to look at the activation of genes<br />

involved in DNA repair pathways. When compared to the controls there was a significant upregulation<br />

of ATM, DNA-PK, Rad52 and MLH1 genes in A549 cells that had been exposed to<br />

fractionated irradiation (Fig.3.1.4, A, B, C and D, Lane3). However, in A549 cells that had been<br />

126


CHAPTER 3<br />

RESULTS<br />

exposed to 10 Gy acute dose there was marginal up-regulation of ATM, DNA-PK and Rad52<br />

genes although it was significantly lower than the fractionated group (Fig.3.1.4, A, B, C and D,<br />

Lane4). In A549 cells that had been exposed to 2 Gy acute dose there was marginal upregulation<br />

of Rad52 and MLH1 genes but these were also significantly lower than the<br />

fractionated group(Fig.3.1.4, A, B, C and D, Lane2).<br />

3.1.5 Efficient Repair of DNA<br />

It has been established recently that H2AX at the DNA double strand break (DSB) sites are<br />

immediately phosphorylated upon irradiation, and the phosphorylated H2AX (γ-H2AX) can be<br />

visualized in situ by immunostaining with a γ-H2AX specific antibody [232, 233]. γ-H2AX<br />

induction can be measured quantitatively at physiological doses, and the numbers of residual γ-<br />

H2AX foci can be used to estimate the kinetics of DSB rejoining. This has become the gold<br />

standard for the detection of DSB [233, 234].<br />

Repair kinetics was determined by counting the γ-H2AX foci at 15 min and 4h after irradiation.<br />

As expected maximum numbers of foci were seen at 15 min in A549 cells exposed to 10 Gy<br />

followed by cells exposed to fractionated irradiation and very few foci were seen in cells<br />

exposed to 2 Gy acute dose. By 4h the number of foci in cells exposed to fractionated irradiation<br />

had been reduced to almost control but in the cells exposed to 10 Gy the number of foci was still<br />

very high, indicating that there was efficient repair of DNA in A549 cells that had been exposed<br />

to fractionated irradiation (Fig. 3.1.5).<br />

127


CHAPTER 3<br />

RESULTS<br />

3.1.6 Radiation induced foci of DNA damage response proteins ATM and BRCA1<br />

ATM, an important DNA damage response protein to ionizing radiation and a part of irradiation<br />

induced foci (IRIF), is mainly involved in cell cycle arrest and repair. It gets activated after DNA<br />

double strand break by phosphorylation at ser 1981 and activates many key cellular proteins that<br />

include H2AX, BRCA1, Chk1/2 and p53. Phospho ATM foci and intensity were looked at 4<br />

hours after irradiation. All the cells either exposed to 10 Gy (23±3 foci per cell) or fractionated<br />

irradiation (25±2.8 foci per cell) showed ATM foci but only 10% of cells exposed to 2 Gy<br />

showed ATM foci and the number of foci per cell was 2±0.08 (Fig. 3.1.6A). The intensity of<br />

phosphorylation of ATM was quantified by ImageJ software, was significantly higher in the<br />

fractionated group as compared to any of the treatment groups (control, 2Gy or 10Gy )<br />

BRCA1, which is phosphorylated by ATM, is a part of IRIF and shares many downstream<br />

substrates of ATM. 4 hours after irradiation, A549 cells that had been exposed to 2 Gy acute<br />

dose did not show any BRCA1 foci where as 55 % of cells that had been exposed to fractionated<br />

irradiation showed BRCA1 foci and the number of foci per cells was 24±4. 22% of cells that had<br />

been exposed to 10 Gy acute showed BRCA1 foci and the number of foci per cell was 15±2.1<br />

(Fig. 3.1.6B). The intensity of phosphorylation of BRCA1 as determined by ImageJ software<br />

was significantly higher in the fractionated group as compared to any of the treatment groups<br />

(control, 2Gy or 10Gy).<br />

128


CHAPTER 3<br />

RESULTS<br />

Fig.3.1.4 Gene expression of ATM, DNA-PK, Rad52 and MLH1 in A549 cells irradiated with<br />

different doses of γ-radiation. Total RNA from A549 cells was isolated and reverse transcribed<br />

4h after irradiation. RT-PCR analysis of ATM, DNA-PK, Rad52 and MLH1 genes was carried<br />

out as described in materials and methods. β-Actin gene expression in each group was used as an<br />

internal control. Ratio of intensities of (A) ATM, (B) DNA-PK, (C) Rad52 and (D) MLH1 band<br />

to that of respective β-actin band as quantified from gel pictures are shown above each gel<br />

picture. Key;Lane 1, control unirradiated A549 cell line. Lane 2, 2 Gy acute dose (single dose<br />

that was delivered in each fraction) of 60 Co γ-irradiation Lane 3 2 Gy of 60 Co γ-irradiation daily<br />

over a period of 5 days, Lane 4, 10 Gy acute (which is the sum total of all the fraction) 60 Co γ-<br />

irradiation. Data represents means ± SE of three independent experiments; significantly different<br />

from unirradiated controls: *P < 0.05, **P < 0.01.<br />

129


CHAPTER 3<br />

RESULTS<br />

Fig. 3.1.5 Repair kinetics as determined by γ-H2AX foci in different treatment group of A549<br />

cells at 15 min and 4h after irradiation. Cells were fixed 15 minutes and 4 hours after irradiation<br />

in 4% paraformaldehyde, permeabilized in 0.25% Triton X-100 and labeled with specific<br />

antibodies as described in “Materials and Methods”. Each phospho-site antibody was indirectly<br />

labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with<br />

ProLong Gold antifede with DAPI (blue). The encircled area roughly represents cell nuclei as<br />

seen by DAPI staining. All images were captured using Carl Zeiss confocal microscope with the<br />

same exposure time. Representative image of three independent experiments is shown here; at<br />

least 100 cells per experiment were analyzed from three independent experiments.<br />

130


CHAPTER 3<br />

RESULTS<br />

Fig. 3.1.6 Radiation induced foci of DNA damage response proteins (A) ATM and (B) BRCA1<br />

in different treatment group of A549 cells 4h after irradiation. Cells were fixed in 4%<br />

paraformaldehyde 4h after irradiation, permeabilized in 0.25% Triton X-100 and labeled with<br />

specific antibodies as described in “Materials and Methods”. Each phospho-site antibody was<br />

indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted<br />

with ProLong Gold antifede with DAPI (blue). All images were captured using Carl Zeiss<br />

confocal microscope with the same exposure time. Representative images of three independent<br />

experiments are shown here; at least 100 cells per experiment were analyzed from three<br />

independent experiments.<br />

131


CHAPTER 3<br />

RESULTS<br />

3.1.7 Translocation of phospho-p53<br />

Microarray data had shown that p53 signaling pathway was up-regualted in A549 cells that had<br />

been exposed to fractionated irradiation; it was of interest to look at the phosphorylation of p53<br />

and its translocation to the nucleus of A549 cells. 18 hours after irradiation, phospho-p53 was<br />

found to be translocated into the nucleus of A549 cells exposed to fractionated irradiation (Fig.<br />

3.1.7A). Phospho-p53 was not found to be translocated into the nucleus of A549 cells that had<br />

been exposed to 2Gy acute dose. However, in A549 cells that had been exposed to 10 Gy acute<br />

dose, there was marginally higher phosphorylation of p53 although it was significantly lower<br />

than the fractionated group (Fig. 3.1.7A). The intensity of phosphorylation was significantly<br />

higher in the cells exposed to fractionated irradiation. Phosphorylation of p53 was further<br />

confirmed by western blot. (Fig. 3.1.7B, Lane3).<br />

3.1.8 Response of A549 and MCF-7 cells exposed to fractionated irradiation.<br />

Having established that among the three cell lines the lung carcinoma cell line A549 was found<br />

to be relatively more radioresistant if the 10 Gy dose was delivered as a fractionated regimen and<br />

MCF-7 cell line more radiosensitive. It was of interest to look at mechanism of such response. 4<br />

h after irradiation, there was a significant up-regulation of ATM, DNA-PK, Rad52 and MLH1<br />

genes in A549 cells but not in MCF-7 cells (Fig. 3.1.8B, Lane3). ATM and BRCA1<br />

phosphorylation was found to be significantly higher 4 h after irradiation in A549 cells but not in<br />

MCF-7 cells (Fig. 3.1.8C). 18 h after irradiation, phospho-p53 was found to be translocated into<br />

the nucleus of A549 cells but not in MCF-7 cells (Fig. 3.1.8C).There was efficient repair of DNA<br />

by 4h in A549 cells as determined by counting γ-H2AX but not in MCF-7 cells (Fig. 3.1.8D).<br />

132


CHAPTER 3<br />

RESULTS<br />

Fig. 3.1.7 Translocation of phospho-p53 into the nucleus of A549 cells 18h after irradiation. (A)<br />

Representative image of three independent experiments showing phosphorylation and<br />

translocation of p53 into nucleus of A549 cells. Each phospho-site antibody was indirectly<br />

labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with<br />

ProLong Gold antifede with DAPI (blue). All images were captured using Carl Zeiss confocal<br />

microscope with the same exposure time. At least 100 cells per experiment were analyzed from<br />

three independent experiments. (B) Representative immunoblot image of three independent<br />

experiments depicting phosphorylation of p53 in different treatment group. Graph represents<br />

phospho-p53 band intensities (divided with the respective loading control intensities and<br />

normalized). Key Lane 1, control unirradiated A549 cell line. Lane 2, 2 Gy acute dose (single<br />

dose that was delivered in each fraction) of 60 Co γ-irradiation Lane 3 2 Gy of 60 Co γ-irradiation<br />

daily over a period of 5 days, Lane 4, 10 Gy acute (which is the sum total of all the fraction)<br />

60 Co γ-irradiation.<br />

133


CHAPTER 3<br />

RESULTS<br />

Fig. 3.1.8 Response of A549 and MCF-7 cells exposed to fractionated irradiation. (A)<br />

Clonogenic Cell Survival of A549 and MCF-7 cells exposed to fractionated irradiation. Data<br />

represents means ± SE of three independent experiments; **P < 0.01 compared with MCF-7 (B)<br />

Representative image of three independent experiments showing gene expression of ATM,<br />

DNA-PK, Rad52 and MLH1 in A549 and MCF-7 cells exposed to fractionated irradiation. β-<br />

Actin gene expression in each group was used as an internal control. (C) Phosphorylation of<br />

ATM, BRCA1 and p53 in A549 and MCF-7 cells exposed to fractionated irradiation. (D) Repair<br />

kinetics as determined by γ-H2AX foci in A549 and MCF-7 cells exposed to fractionated<br />

irradiation at 15 min and 4h after irradiation. Each phospho-site antibody was indirectly labeled<br />

with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong<br />

Gold antifede with DAPI (blue). The encircled area roughly represents cell nuclei as seen by<br />

DAPI staining. All images were captured using Carl Zeiss confocal microscope with the same<br />

exposure time. Representative images of three independent experiments are shown here; at least<br />

100 cells per experiment were analyzed from three independent experiments.<br />

134


CHAPTER 3<br />

RESULTS<br />

3.1.9 Stabilization of DNA-PK mRNA<br />

mRNA stability plays a major role in gene expression in mammalian cells, affecting the rates at<br />

which mRNAs disappear following transcriptional repression and accumulate following<br />

transcriptional induction. An array of exogenous factors, from hormones to viruses to radiation,<br />

influence mRNA halflives in ways that are incompletely understood. It was of interest to look for<br />

the stabilization of different mRNA in A549 cells exposed to fractionated irradiation.<br />

The activation of DNA-PK was accompanied by the stabilization of its mRNA in A549 cells that<br />

had been exposed to fractionated irradiation (Fig. 3.1.9). The ATM and Rad 52 mRNA were not<br />

stabilized indicating a specific and selective stabilization of DNA-PK mRNA. Very little is<br />

known about the signaling pathways regulating mRNA turnover in contrast to current<br />

understanding of how signaling pathways regulate transcription. Recently, however, several<br />

studies have provided increasing support for the notion that mRNA stability is regulated through<br />

signal transduction pathways involving phosphorylation events. It was of interest to look for the<br />

role of signal transduction pathway in the regulation of mRNA stabilization of DNA-PK. DNA-<br />

PK mRNA stabilization was found to be regulated by the p38 MAP Kinase pathway but not by<br />

ERK or JNK MAP Kinase pathway (Fig. 3.1.9).<br />

135


CHAPTER 3<br />

RESULTS<br />

y<br />

Relative expression of DNA-PK<br />

30000<br />

25000<br />

20000<br />

15000<br />

10000<br />

5000<br />

0<br />

2h+Act 2h+Act+In 4h+Act 4h+Act+In 6h+Act 6h+Act+In 8h+Act 8h+Act+In<br />

Time (h) 2 2 4 4 6 6 8 8<br />

Actinomycin (5mg/ml) + + + + + + + +<br />

p38 MAPK inhibitor (1µM) - + - + - + - +<br />

DNA-PK<br />

Fig.3.1.9 Stabilization of DNA-PK mRNA in A549, that had been exposed to 2 Gy of 60 Co γ-<br />

irradiation daily over a period of 5 days and its stabilization is control by p38 MAP Kinase.<br />

3.1.10 Role of Rad52 and MLH1<br />

Since there was an intense activation of Rad52 and MLH1 in A549 cells exposed to fractionated<br />

irradiation, it was of interest to look for their role in survival of A549 cells. MLH1 and Rad52<br />

gene expression knockdown was carried out in A549 cells using MLH1 and Rad52 shRNA<br />

plasmids. Stably transfected cells were selected and the transfection efficiency was found to be<br />

85% for MLH1 gene and 80% for Rad52 gene.<br />

There was 48±2.8% survival of A549 cells exposed to fractionated irradiation. The survival of<br />

A549 cells that had been transfected with MLH1 shRNA plasmid and exposed to fractionated<br />

irradiation was found to be 40±1.8% where as the survival of A549 cells that had been<br />

transfected with Rad52 shRNA plasmid and exposed to fractionated irradiation was found to be<br />

136


CHAPTER 3<br />

RESULTS<br />

23±1.5% (Fig. 3.1.10, A and B). The survival of cells transfected with control shRNA plasmid<br />

was the same as unirradiated control cells.<br />

Fig.3.1.10 Effect of Rad52 and MLH1 on the growth and radioresistance of A549 cells which<br />

had been exposed to 5 fractions of 2Gy radiation. A549 cells were transfected with either MLH1<br />

or Rad52 ShRNA as described in “Materials and Methods”. (A) Clonogenic survival assay of<br />

A549 cells. Key; control, control unirradiated A549 cells. 5X2Gy, A549 cells exposed to 2 Gy<br />

of 60Co γ-irradiation daily over a period of 5 days. 5X2Gy + M, A549 cells transfected with<br />

MLH1 ShRNA and exposed to 2 Gy of 60 Co γ-irradiation daily over a period of 5 days. 5X2Gy +<br />

R, A549 cells transfected with Rad52 ShRNA and exposed to 2 Gy of 60 Co γ-irradiation daily<br />

over a period of 5 days. (B) Representative image of three independent experiments. Data<br />

represents means ± SE of three independent experiments; *P < 0.05, **P < 0.01 compared with<br />

5X2Gy treatment group.<br />

137


CHAPTER 3<br />

RESULTS<br />

3.2 Proton beam induced signaling in mammalian cells and its comparison with Gamma<br />

irradiation.<br />

Having established that among the three cell lines, A549 was the most radioresistant, the<br />

variance in signaling pattern and effectiveness of cell killing with the change in LET was looked<br />

at. In this study, A549 cells were irradiated with either 2Gy proton beam or 2Gy γ-radiation and<br />

ensuing activation of few important components involved in DNA repair, cell cycle arrest, and<br />

apoptosis pathways were followed to elucidate the mechanisms behind observed cellular<br />

response. Despite the variability in effectiveness of these two radiations, equal doses of each<br />

were chosen to ensure that each cell is hit by proton beam and no bystander effects are<br />

manifested. Essentially all biological differences between proton and gamma radiations arise<br />

from differences in their ionization tracks. Equitoxic doses were not preferred because the<br />

probability of unhit cells would increase. How this qualitative difference of radiation is translated<br />

into activation of crucial intracellular pathways was the focus of the study.<br />

3.2.1 Cell Survival Assay<br />

There was only 23±2 % of A549 cells were survived that had been exposed to 2Gy of proton<br />

beam where as 56±2.2 % cells were survived that had been exposed to 2Gy of γ-radiation as<br />

observed with clonogenic assay (Fig.3.2.1). This result clearly showed that proton beam is more<br />

cytotoxic than gamma radiation.<br />

3.2.2 Differential modulation of Gene expression following proton beam and gamma irradiation<br />

Expression of ATM, DNA-PK and Chk2 genes were significantly upregulated after 4h of<br />

irradiation only in A549 cells that had been exposed to proton beam (2Gy) (Fig.3.2.2 A, C),<br />

138


CHAPTER 3<br />

RESULTS<br />

whereas only the ATR gene was significantly upregulated after 4h of irradiation in A549 cells<br />

that had been exposed to γ-radiation (2Gy) (Fig.3.2.2 B). There was no change in the gene<br />

expression of Chk1 in A549 cells that had been exposed to either γ-radiation or proton beam<br />

(2Gy) (Fig.3.2.2 C, D). Independent experiments were carried out for gamma and proton beam<br />

irradiated cells, so the control group of gamma and proton irradiated cells may have different<br />

transcript level. However, the data were normalized for the respective controls.<br />

3.2.3 Activation of signaling components involved in DNA damage sensing and repair following<br />

proton beam irradiation.<br />

Since ataxia telangiectasia mutated (ATM) protein plays a central role in DNA-double strand<br />

break (DSBs) sensing and repair and determines the cellular response in eukaryotes. It was of<br />

interest to look for its phosphorylation and its substrates at 4 hours after irradiation. There was<br />

significantly higher phosphorylation of ATM at serine 1981, H2AX at γ-position, Chk2 at<br />

threonine 68 and p53 at serine 15 in A549 cells that had been exposed to 2Gy of proton beam,<br />

but their phosphorylation was not found to be increased 4 hours after 2Gy of γ-radiation as<br />

compared to the control (Fig.3.2.3).<br />

3.2.4 Expression of Apoptosis related genes<br />

Since there was significantly higher phosphorylation of p53 in A549 cells that had been exposed<br />

to 2Gy proton beam (Fig.3.2.3), it was our interest to look for the expression of apoptosis related<br />

gene like Bax and Bcl-2.<br />

139


CHAPTER 3<br />

RESULTS<br />

There was significant upregulation of pro-apoptotic gene, Bax, 12 h after irradiation in A549<br />

cells that had been exposed to 2Gy proton beam as compared to the unirradiated control. In γ-<br />

irradiated (2Gy) cells there was no significant upregulation of the Bax gene (p


CHAPTER 3<br />

RESULTS<br />

Fig.3.2.2 Gene expression of ATM, DNA-PK, ATR, Chk1 and Chk2 in A 549 cells irradiated<br />

with 2Gy Proton Beam or 2Gy γ-radiation at 2, 3 or 4h after irradiation. Total RNA from A 549<br />

cells was isolated and reverse transcribed. RT-PCR analysis of ATM, DNA-PK, ATR, Chk1 and<br />

Chk2 genes was carried out as described in materials and methods. PCR products were resolved<br />

on 1.5% agarose gels containing ethidium bromide. β-Actin gene expression in each group was<br />

used as an internal control. Ratio of intensities of (A & B) ATM, DNA-PK or ATR, (C & D)<br />

Chk1 and Chk2 band to that of respective β-actin band as quantified from gel pictures are shown<br />

above each gel picture. Con means unirradiated control. Data represents means ± SE of three<br />

independent experiments; significantly different from unirradiated controls: *P < 0.05, **P <<br />

0.01.<br />

141


CHAPTER 3<br />

RESULTS<br />

Fig.3.2.3 Immunofluorescence staining and Image analysis of γ-H2AX, phospho-ATM, phospho-Chk2<br />

and phospho-p53 4h after irradiation in A 549 cells irradiated with 2Gy Proton Beam or 2Gy γ-radiation.<br />

Cells were fixed in 4% paraformaldehyde, permeabilized in 0.25% Triton X-100 and labeled with specific<br />

antibodies as described in “Materials and Methods”. Each phospho-site antibody was indirectly labeled<br />

with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold<br />

antifede with DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the same<br />

exposure time. Comparison of phosphorylation between 2Gy Proton beam and 2Gy γ-radiation of (A)<br />

H2AX, (B) ATM, (C) Chk2, (D) p53. (E) Graph represents relative phosphorylation of H2AX, ATM,<br />

Chk2 and p53 as determined by ImageJ software. At least 100 cells per experiment were analyzed from<br />

three independent experiments. Data represents means ± SE of three independent experiments;<br />

significantly different from unirradiated controls: *P < 0.05, **P < 0.01.<br />

142


CHAPTER 3<br />

RESULTS<br />

Fig.3.2.4 Gene expression of Bax and Bcl-2 12h after irradiation in A 549 cells irradiated with<br />

2Gy Proton Beam or 2Gy γ-radiation. Total RNA from A 549 cells was isolated and reverse<br />

transcribed. RT-PCR analysis of Bax and Bcl-2 genes was carried out as described in materials<br />

and methods. PCR products were resolved on 1.5% agarose gels containing ethidium bromide. β-<br />

Actin gene expression in each group was used as an internal control. Ratio of intensities of Bax<br />

and Bcl-2 band to that of respective β-actin band as quantified from gel pictures are shown above<br />

each gel picture. Con means unirradiated control, H means proton. Data represents means ± SE<br />

of three independent experiments; significantly different from unirradiated controls: *P < 0.05,<br />

**P < 0.01.<br />

143


CHAPTER 3<br />

RESULTS<br />

3.3 High LET irradiation induced signaling in mammalian cells and its comparison with<br />

Gamma irradiation.<br />

Having established the signaling differences between proton beam and gamma ray in lung<br />

carcinoma A549 cells, the variance in signaling pattern with higher LET (Carbon and Oxygen)<br />

was studied. Although enough gross evidence exists of the activation of signaling pathways after<br />

irradiation, the differences between gamma and High LET have not been studied in detail. The<br />

two seem to activate the same pathways but must diverge at some point so that the end point can<br />

be different. Many studies on signaling pathways after low and high LET radiations have found<br />

the activation to be similar except in the intensity [214] but since the end result is different, there<br />

must be a divergence of pathways at some stage. It is this stage of signaling, that is different, that<br />

will be crucial to designing therapies that target specific molecules or pathways. The aim of the<br />

present study was to look for these subtle differences that are of immense consequences.<br />

3.3.1 Carbon beam induced signaling in mammalian cells and its comparison with Gamma<br />

irradiation.<br />

Carbon beams are high linear energy transfer (LET) radiation characterized by higher relative<br />

biological effectiveness than low LET radiation. The aim of the current study was to determine<br />

the signaling differences between γ and carbon ion-irradiation. A549 cells were irradiated with<br />

1Gy carbon or γ-radiation and the activation of DNA damage response signaling was studied.<br />

3.3.1.1 Carbon ions were three times more cytotoxic than gamma<br />

To determine the sensitivity of cells to similar dose of high and low LET radiation, their ability<br />

to form colonies after irradiation was observed. Irradiation of exponentially growing A549 cells<br />

144


CHAPTER 3<br />

RESULTS<br />

with equal doses of gamma and carbon (290Mev/micron) revealed that at same dose carbon ions<br />

were much more cytotoxic than gamma rays. At 1Gy, the percentage survival for gamma and<br />

carbon was 92.3±6.8% and 24.5±1.22% respectively. As against 1Gy of carbon ions, 20% cell<br />

survival was obtained at 3Gy of gamma. Thus, carbon ions were found to be three times more<br />

toxic than gamma rays. At 2Gy percentage survivals were 60.6±4% and 4.8±0.24 % for gamma<br />

and carbon respectively (Fig.1).<br />

Fig.3.3.1.1 Clonogenic Cell Survival of A549 cells irradiated with 1 and 2Gy Carbon Beam or γ-<br />

radiation. Data represents means ± SD of three independent experiments.<br />

145


CHAPTER 3<br />

RESULTS<br />

3.3.1.2 Number of initial γ-H2AX foci formed for both radiation types of radiation.<br />

Since DNA double strand breaks are the primary toxic lesions induced by radiation, the number<br />

of double strand breaks were scored by examining the foci of phosphorylated H2AX in irradiated<br />

cells. After 15 minutes of either radiation, γ-H2AX foci, visualized as bright spots, were present<br />

in all the cells. The average numbers of foci after 15min of 1Gy gamma irradiation were<br />

15.8±4.8, while after 1Gy carbon irradiation the foci were 19±4.7 (Fig.3.3.1.2). Thus unlike the<br />

cell survival, not much difference was observed in number of γ-H2AX foci after irradiation with<br />

equal doses of carbon and gamma. The foci were also counted 4 hours after irradiation. 57% of<br />

gamma irradiated cells still showed γ-H2AX foci and the average foci per cell had reduced to<br />

only 2.9±2.28. Carbon ion irradiated cells also showed a significant reduction in the number of γ-<br />

H2Ax foci (2±2) per cell that were visible in only 22% of cells (Fig. Fig.3.3.1.2A, B). When total<br />

intensity of γ-H2AX rather than the foci was estimated by ImageJ software in the same images, it<br />

corroborated well with the number of γ-H2AX foci in gamma irradiated cells at 15min and 4hr<br />

after irradiation (Fig. Fig.3.3.1.2C). However, in case of carbon irradiation, the total intensity of<br />

γ-H2AX at 4 hours was yet high (Fig.3.3.1.2C).<br />

Since initial response as observed from γ-H2AX foci at 15 minutes was similar with both type of<br />

radiation, irradiation induced foci of downstream signaling components was followed at 4 hours<br />

to highlight the signaling differences at that time period.<br />

146


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.1.2 Formation of γ-H2AX foci after exposure to 1Gy gamma or carbon irradiation in<br />

A549 cells.(A) Representative image showing γ-H2AX foci (green) in cells fixed 15 minutes and<br />

4 hours after irradiation. The encircled area roughly represents cell nuclei as seen by DAPI<br />

staining. All images were captured using Carl Zeiss confocal microscope with the same exposure<br />

time (B) Graph represents average foci per cell, percentage of cells showing the foci is marked<br />

above the bars. (C) Graph represents relative intensity of γ-H2AX as determined by ImageJ<br />

software. At least 100 cells per experiment were analyzed from three independent experiments.<br />

147


CHAPTER 3<br />

RESULTS<br />

3.3.1.3 ATM foci in gamma and carbon irradiated cells<br />

ATM, an important DNA damage response protein to ionizing radiation and a part of irradiation<br />

induced foci (IRIF), is mainly involved in cell cycle arrest and repair. It gets activated after DNA<br />

dsb by phosphorylation at ser 1981 and activates many key cellular proteins that include H2AX,<br />

BRCA1, Chk1/2 and p53. Phospho ATM foci and intensity were scored at 4 hours after gamma<br />

or carbon irradiation. At 4 hours after gamma irradiation, 88% cells showed pATM foci with the<br />

average foci per cell was 6.5±4.4, as against 25% control cells (3.4±1.67) (Fig.3.3.1.3A, B).<br />

Like γ-H2AX foci, small number of ATM foci 4 hours after gamma irradiation indicates that by<br />

4 hours most of the repair have taken place. However, at 4 hours after carbon ion irradiation only<br />

52 % of the cells showed the foci but the average pATM foci per cell (21±10) was much higher<br />

as compared to 4 hours after gamma (Fig.3.3.1.3A, B). Total intensity levels of phospho ATM<br />

matched with the number of foci observed (Fig.3.3.1.3C).<br />

3.3.1.4 ATR foci in gamma and carbon irradiated cells<br />

ATR, another important DNA damage sensor, which is known to respond late to the IR induced<br />

damage, was also looked for its activation at 4 hours. After gamma irradiation, 53% of the cells<br />

showed ATR foci (2.8±1.64), as against no foci in any of the control-unirradiated cells. On the<br />

other hand, 4 hours after carbon ion irradiation although only 10% of cells showed the foci, the<br />

average foci per cell (7±4.5) were thrice that of gamma irradiated cells (Fig.3.3.1.4A, B). Total<br />

intensity levels of phospho ATR matched with the number of foci observed (Fig.3.3.1.4C).<br />

148


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.1.3 Phosphorylation of ATM at 4h after exposure to 1Gy gamma or carbon irradiation in<br />

A549 cells. (A) Representative image showing p-ATM foci 4 hours after irradiation. Each<br />

phospho-ATM antibody was indirectly labeled with Molecular Probe 488 secondary antibody<br />

(green) and cells were mounted with ProLong Gold antifede with DAPI (blue). All images were<br />

captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph<br />

represents average numbers of foci per cell, percentage of cells showing the foci is marked above<br />

the bars. (C) Graph represents relative intensity of p-ATM as determined by ImageJ software. At<br />

least 100 cells per experiment were analyzed from three independent experiments. Data<br />

represents means ± SD of three independent experiments.<br />

149


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.1.4 Phosphorylation of ATR at 4h after exposure to 1Gy gamma or carbon irradiation in<br />

A549 cells. (A) Representative image showing p-ATR foci 4 hours after irradiation. Each<br />

phospho-ATR antibody was indirectly labeled with Molecular Probe 488 secondary antibody<br />

(green) and cells were mounted with ProLong Gold antifede with DAPI (blue). All images were<br />

captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph<br />

represents average numbers of foci per cell, percentage of cells showing the foci is marked above<br />

the bars. (C) Graph represents relative intensity of p-ATR as determined by ImageJ software. At<br />

least 100 cells per experiment were analyzed from three independent experiments. Data<br />

represents means ± SD of three independent experiments.<br />

3.3.1.5 BRCA1 foci<br />

Another important protein involved in DSB repair is the tumor suppressor protein BRCA1 which<br />

is phosphorylated by ATM and is also a part of IRIF. 4 hours after gamma irradiation, 55% cells<br />

displayed BRCA1 foci (13±8) while 100% of carbon irradiated cells displayed BRCA1 foci<br />

150


CHAPTER 3<br />

RESULTS<br />

(7.5±0.64) (Fig. 3.3.1.5) and the foci, although less in number were larger in size than gamma<br />

induced foci. Presence of BRCA1 foci in all cells irradiated after carbon irradiation indicates that<br />

BRCA1 might be the important player in response to complex DNA damage induced by high<br />

LET radiation.<br />

Fig.3.3.1.5 Phosphorylation of BRCA1 at 4h after exposure to 1Gy gamma or carbon irradiation in A549<br />

cells. (A) Representative image showing p-BRCA1 foci 4 hours after irradiation. Each phospho-BRCA1<br />

antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were<br />

mounted with ProLong Gold antifede with DAPI (blue). All images were captured using Carl Zeiss<br />

confocal microscope with the same exposure time. (B) Graph represents average numbers of foci per cell,<br />

percentage of cells showing the foci is marked above the bars. (C) Graph represents relative intensity of<br />

p-BRCA1 as determined by ImageJ software. At least 100 cells per experiment were analyzed from three<br />

independent experiments. Data represents means ± SD of three independent experiments.<br />

151


CHAPTER 3<br />

RESULTS<br />

3.3.1.6 Activation of downstream effectors; Chk1 and Chk2<br />

Activation of ATM, ATR and BRCA1 after DNA damage leads to the activation of further<br />

downstream components including Chk1 and Chk2. After activation in the nuclei, Chk1 and<br />

Chk2 move to cytoplasm. Phospho Chk1, the major target of ATR, was found to increase in<br />

nuclei of carbon-irradiated cells but not in gamma (Fig.6 A). Marginal activation of Chk2, the<br />

preferred substrate of ATM, was observed in the nucleus in all the cells irrespective of radiation<br />

quality (Fig.6 B). Many pChk2 foci were observed outside the nuclei of carbon irradiated cells<br />

unlike gamma-irradiated cells (Fig.6 C).<br />

3.3.1.7 Activation of intracellular MAPKs, ERK and JNK<br />

ERK1/2 showed significant activation 1 hour after gamma irradiation and thereafter the increase<br />

was persistent till 4 hours (Fig.3.3.1.7a), while JNK did not show any activation after gamma<br />

irradiation (Fig.3.3.1.7b). Interestingly, after carbon irradiation the activation pattern of these<br />

kinases was exact opposite. ERK was not activated at all after carbon irradiation (Fig.3.3.1.7a)<br />

while JNK showed marginal activation at 15 minutes that remained so till 1 hour (Fig.3.3.1.7b).<br />

3.3.1.8 Apoptosis<br />

At 4 hours after irradiation, morphological features of DAPI stained apoptotic nuclei were scored<br />

and many apoptotic nuclei were observed in cells irradiated with carbon ion but not in gamma<br />

irradiated cells (Fig. 3.3.1.8 A, B).<br />

Upregulation of Bax expression was also observed as early as 2 hours after carbon ion irradiation<br />

indicating apoptotic tendency of the cells (Fig. 3.3.1.8C). Gamma irradiated cells did not show<br />

any upregulation of Bax.<br />

152


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.1.6 Phosphorylation of Chk1 and Chk2 at 4h after exposure to 1Gy gamma or carbon<br />

irradiation in A549 cells. (A) Graph represents relative phosphorylation of Chk1 as determined<br />

by ImageJ software (B) Representative image showing phosphorylation and Chk2 foci (marked<br />

by “arrow”). The encircled area roughly represents cell nuclei as seen by DAPI staining. All<br />

images were captured using Carl Zeiss confocal microscope with the same exposure time (C)<br />

Graph represents relative phosphorylation of Chk2 as determined by ImageJ software. At least<br />

100 cells per experiment were analyzed from three independent experiments. Data represents<br />

means ± SD of three independent experiments.<br />

153


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.1.7a Phosphorylation of ERK (pERK) in A549 cells. Cells were lysed at different time<br />

periods (hours) after exposure to 1Gy dose of gamma or carbon ion irradiation, immunoblotted<br />

and detected using anti- phospho ERK. One representative image of three independent<br />

experiments is shown here (A). Graph represents pERK band intensities (divided with the<br />

respective loading control intensities and normalized) plotted against time (B).<br />

154


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.1.7b Phosphorylation of JNK (pJNK) in A549 cells. Cells were lysed at different time<br />

periods (hours) after exposure to 1Gy dose of gamma or carbon ion irradiation, immunoblotted<br />

and detected using anti- phospho JNK. One representative image of three independent<br />

experiments is shown here (A). Graph represents band intensities of pJNK (divided with the<br />

respective loading control intensities and normalized) plotted against time (B).<br />

155


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.1.8 Apoptosis in carbon irradiated A549 cells. (A), Apoptotic nuclei of A549 cells were<br />

visualized and quantified by using DAPI staining, 4 hours after carbon or gamma irradiation<br />

(1Gy). (B), Graph showing the percentage of cells with apoptotic nuclei. At least 100 cells per<br />

experiment were analyzed from three independent experiments. (C), Representative immunoblot<br />

image depicting upregulation of Bax, observed at various time periods (2, 3, 4 and 6 hours) after<br />

carbon irradiation (1Gy).<br />

156


CHAPTER 3<br />

RESULTS<br />

3.3.2 Oxygen beam induced signaling in mammalian cells and its comparison with Gamma<br />

irradiation.<br />

Having established the signaling differences between carbon beam and gamma irradiation in<br />

A549 cells, the variance in signaling pattern with oxygen beam was studied. Oxygen beams are<br />

high linear energy transfer (LET) radiation characterized by higher relative biological<br />

effectiveness than low LET radiation. The aim of the current study was to determine the<br />

signaling differences between γ- and oxygen ion-irradiation. Activation of various signaling<br />

molecules was looked in A549 lung adenocarcinoma cells irradiated with 2Gy oxygen, 2Gy or<br />

6Gy γ-radiation.<br />

3.3.2.1. Cytotoxicity of oxygen ions Vs gamma<br />

To determine the sensitivity of cells to similar dose of high and low LET radiation, their ability<br />

to form colonies after irradiation was observed. Irradiation of exponentially growing A549 cells<br />

with equal doses of gamma and oxygen beam (614 keV/ µm) revealed that at same dose oxygen<br />

ions were much more cytotoxic than gamma rays as assessed by the clonogenic cell survival<br />

assay. The calculated RBE (relative biological effectiveness) for A549 cells irradiated with 1Gy<br />

oxygen ions particles relative to γ rays was found to be nearly 3 at 20% survival (Fig. 3.3.2.1).<br />

The calculated D 0 value from the plot curve for A549 cells was found to be 0.6Gy for oxygen<br />

ions and 2.5Gy for γ rays.<br />

157


CHAPTER 3<br />

RESULTS<br />

Fig. 3.3.2.1 Clonogenic Cell Survival of A549 cells irradiated with 1, 2 or 3Gy Oxygen beam or<br />

γ-radiation. Data represents means ± SD of three independent experiments.<br />

3.3.2.2. Numbers of initial γ-H2AX foci formed for both types of radiation<br />

Since DNA double strand breaks are the primary toxic lesions induced by radiation, the number<br />

of double strand breaks were scored by examining the foci of phosphorylated H2AX in irradiated<br />

cells. After 15 minutes of either radiation, γ-H2AX foci, visualized as bright spots, were present<br />

158


CHAPTER 3<br />

RESULTS<br />

in all the cells. The average numbers of foci after 15min of 2Gy gamma irradiation were 24±5,<br />

while after 2Gy oxygen irradiation; the foci were 28±4.7 (Fig. 3.3.2.2 A & C). Thus unlike the<br />

cell survival, the difference in induction of γ-H2AX foci 15 min after 2Gy of either radiation was<br />

not as profound. Activation of signaling molecules at dose of 2Gy of gamma radiation was<br />

studied because the initial response manifested as double strand breaks, as indicated by γ-H2AX<br />

foci, was found to be almost the same at 2Gy of either radiation. Since the RBE of oxygen ions<br />

was nearly 3, it would be more logical to compare 2Gy of oxygen ion with 6Gy of gamma ray.<br />

The average numbers of foci after 15min of 6Gy gamma irradiation were 35±3.5 (Fig. 3.3.2.2 A<br />

& C). Thus, 6Gy gamma ray induced 1.25 times more foci than 2Gy oxygen ions.<br />

The γ-H2AX foci were also counted 4 hours after irradiation to compare the repair status at that<br />

time. 53% of 2Gy gamma irradiated cells still showed γ-H2AX foci but the average foci per cell<br />

had reduced to only 4.3±2.2, where as 100% of 6Gy gamma irradiated cells showed γ-H2AX<br />

foci and the average foci per cell had reduced to only 8±2.5. In oxygen ion irradiated cells,<br />

although bright green γ-H2AX intensity was observed till 4 hours in 60% of cells, there was a<br />

significant reduction in appearance of distinct foci like structures (10±2) (Fig. 3.3.2.2 B & C).<br />

Disappearance of γ-H2AX foci indicates DNA repair/rejoining. However, from the survival data<br />

of A549 cells it was clear that in oxygen irradiated cells the repair of DNA had not taken place.<br />

The result was so intriguing that it was essential to measure the total intensity of γ-H2AX in<br />

irradiated cells. Instead of counting number of foci, the intensity of γ-H2AX in cells showing the<br />

foci was estimated by ImageJ software in the same images. In case of 2Gy or 6Gy gamma<br />

irradiation, both γ-H2AX foci and its total intensity levels had reduced at 4 hours (Fig. 3.3.2.2<br />

D). However, in case of oxygen irradiation, the total intensity of γ-H2AX at 4 hours was yet very<br />

high (Fig. 3.3.2.2D) although foci were not properly visible. This could indicate diffusion of γ-<br />

159


CHAPTER 3<br />

RESULTS<br />

H2AX foci rather than its disappearance due to actual dephosphorylation of γ-H2AX. Presence<br />

of γ-H2AX even 4 hours after oxygen ion irradiation suggests that DNA had not been repaired.<br />

Fig. 3.3.2.2 Phosphorylation of H2AX (γ-H2AX) at different time points after exposure to 2Gy<br />

and 6Gy gamma or 2Gy oxygen irradiation in A549 cells. Representative image showing γ-<br />

H2AX foci (A) 15 minutes and (B) 4 hours after irradiation. Each phospho-H2AX antibody was<br />

indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted<br />

with ProLong Gold antifede with DAPI (blue). All images were captured using Carl Zeiss<br />

confocal microscope with the same exposure time. (C) Graph represents average numbers of foci<br />

per cell, percentage of cells showing the foci is marked above the bars. (D) Graph represents<br />

relative intensity of γ-H2AX as determined by ImageJ software. At least 100 cells per<br />

experiment were analyzed from three independent experiments. Data represents means ± SD of<br />

three independent experiments; significantly different from unirradiated controls: *P < 0.05, **P<br />

< 0.01.<br />

160


CHAPTER 3<br />

RESULTS<br />

3.3.2.3 Radiation induced foci of DNA damage response proteins ATM<br />

ATM is an important DNA damage response protein to ionizing radiation and a part of<br />

irradiation induced foci (IRIF). It gets activated after DNA dsb by phosphorylation at ser 1981<br />

and activates many key cellular proteins that include H2AX, BRCA1, Chk1/2 and p53 [235,<br />

236]. Phospho ATM foci and intensity were looked at 4 hours after gamma or oxygen<br />

irradiation. At 4 hours after 2Gy gamma irradiation, most of the cells (93%) showed ATM foci<br />

(5.9±3.7) as against 25% unirradiated cells (3.4±1.6) (Fig.3.3.2.3). Like γ-H2AX foci, lesser<br />

number of ATM foci 4 hours after 2Gy gamma irradiation indicates that by 4 hours most of the<br />

repair due to damage by low LET irradiation would have taken place. 4 hours after 6Gy gamma<br />

irradiation, 100% of the cells showed the foci and average ATM foci per cell were 14.5±6.05.<br />

However, at 4 hours after oxygen ion irradiation, 100% of the cells still showed the foci and<br />

average ATM foci per cell (18.6±8.7) were higher than gamma irradiation. Total intensity levels<br />

of phospho ATM matched with the number of foci observed (Fig.3.3.2.3).<br />

3.3.2.4 Radiation induced foci of DNA damage response proteins ATR<br />

ATR, another PI3 kinase family member, which is known to respond late to the IR induced<br />

damage, was also looked for its activation at 4 hours. After gamma irradiation, average foci per<br />

cell were very less (2.7±0.95 for 2Gy and 4±1 for 6Gy) and that too were visible in only half of<br />

the irradiated cells. (Fig.3.3.2.4). On the other hand, although only 36% of oxygen ion irradiated<br />

cells showed ATR foci, the average foci per cell were (24±14) at least six times more than that of<br />

gamma irradiated cells. Total intensity levels of phospho ATR matched with the number of foci<br />

observed (Fig.3.3.2.4).<br />

161


CHAPTER 3<br />

RESULTS<br />

3.3.2.5. Activation of downstream substrates of ATR and ATM: Chk1, Chk2 and p53<br />

There was significantly higher phosphorylation of p53 at serine 15 in A549 cells that had been<br />

exposed to 2Gy of oxygen beam, but not after 2Gy γ-radiation (Fig.3.3.2.5). Cells exposed to<br />

6Gy γ-radiation showed marginally higher phosphorylation of p53, but it was significantly lower<br />

than 2Gy oxygen irradiated cells (Fig.3.3.2.5).<br />

Chk2, which is activated by ATM and shares its substrates including p53, was found to be<br />

significantly activated in oxygen irradiated cells. Activation of Chk2 is critical in regulating cell<br />

cycle arrest and DSB repair. Presence of pChk2 foci in nuclei of oxygen irradiated cells but not<br />

gamma indicates Chk2 as an active player in high LET radiation induced responses<br />

(Fig.3.3.2.6A, B). Immunofluorescence of phospho Chk1, which is a major target of ATR, was<br />

also found to increase in nuclei of oxygen-irradiated cells but not after 2Gy γ-radiation<br />

(Fig.3.3.2.6C). Cells exposed to 6Gy γ-radiation showed marginally higher phosphorylation of<br />

Chk1, but it was significantly lower than 2Gy oxygen irradiated cells (Fig.3.3.2.6C).<br />

3.3.2.6 Apoptosis<br />

At 18 hours after irradiation, morphological features of DAPI stained apoptotic nuclei were<br />

scored and many apoptotic nuclei were observed in cells irradiated with 2Gy oxygen ion and<br />

6Gy gamma but not in 2Gy gamma irradiated cells (Fig.3.3.2.7). Cells exposed to 2Gy oxygen<br />

ion showed significantly higher apoptotic nuclei than 6Gy gamma (Fig.3.3.2.7).<br />

162


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.2.3 Phosphorylation of ATM at 4h after exposure to 2Gy and 6Gy gamma or 2Gy<br />

oxygen irradiation in A549 cells. (A) Representative image showing p-ATM foci 4 hours after<br />

irradiation. Each phospho-ATM antibody was indirectly labeled with Molecular Probe 488<br />

secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI<br />

(blue). All images were captured using Carl Zeiss confocal microscope with the same exposure<br />

time. (B) Graph represents average numbers of foci per cell, percentage of cells showing the foci<br />

is marked above the bars. (C) Graph represents relative intensity of p-ATM as determined by<br />

ImageJ software. At least 100 cells per experiment were analyzed from three independent<br />

experiments. Data represents means ± SD of three independent experiments; significantly<br />

different from unirradiated controls: *P < 0.05, **P < 0.01.<br />

163


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.2.4 Phosphorylation of ATR at 4h after exposure to 2Gy and 6Gy gamma or 2Gy oxygen<br />

irradiation in A549 cells. (A) Representative image showing p-ATR foci 4 hours after<br />

irradiation. Each phospho-ATR antibody was indirectly labeled with Molecular Probe 488<br />

secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI<br />

(blue). All images were captured using Carl Zeiss confocal microscope with the same exposure<br />

time. (B) Graph represents average numbers of foci per cell, percentage of cells showing the foci<br />

is marked above the bars. (C) Graph represents relative intensity of p-ATR as determined by<br />

ImageJ software. At least 100 cells per experiment were analyzed from three independent<br />

experiments. Data represents means ± SD of three independent experiments; significantly<br />

different from unirradiated controls: *P < 0.05, **P < 0.01. Significantly different from 6Gy<br />

gamma: ## P < 0.01.<br />

164


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.2.5 Phosphorylation of TP53 at 4h after exposure to 2Gy and 6Gy gamma or 2Gy oxygen<br />

irradiation in A549 cells. (A) Representative image showing p-TP53 4 hours after irradiation.<br />

Each phospho-TP53 antibody was indirectly labeled with Molecular Probe 488 secondary<br />

antibody (green) and cells were mounted with ProLong Gold antifede with DAPI (blue). All<br />

images were captured using Carl Zeiss confocal microscope with the same exposure time. (B)<br />

Graph represents relative intensity of p-TP53 as determined by ImageJ software. At least 100<br />

cells per experiment were analyzed from three independent experiments. Data represents means<br />

± SD of three independent experiments; significantly different from unirradiated controls: *P <<br />

0.05, **P < 0.01. Significantly different from 6Gy gamma: # P < 0.05.<br />

165


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.2.6 Phosphorylation of Chk2 at 4h after exposure to 2Gy and 6Gy gamma or 2Gy oxygen<br />

irradiation in A549 cells. (A) Representative image showing p-Chk2 4 hours after irradiation.<br />

Each phospho-Chk2 antibody was indirectly labeled with Molecular Probe 488 secondary<br />

antibody (green) and cells were mounted with ProLong Gold antifede with DAPI (blue). All<br />

images were captured using Carl Zeiss confocal microscope with the same exposure time. (B)<br />

Graph represents relative intensity of p-Chk2 as determined by ImageJ software. (C) Graph<br />

represents relative intensity of p-Chk1 as determined by ImageJ software At least 100 cells per<br />

experiment were analyzed from three independent experiments. Data represents means ± SD of<br />

three independent experiments; significantly different from unirradiated controls: *P < 0.05, **P<br />

< 0.01. Significantly different from 6Gy gamma: # P < 0.05.<br />

166


CHAPTER 3<br />

RESULTS<br />

Fig.3.3.2.7 Apoptosis in oxygen irradiated A549 cells. Apoptotic nuclei of A549 cells were<br />

visualized and quantified by using DAPI staining, 18 hours after 2Gy and 6Gy gamma or 2Gy<br />

oxygen irradiation. The percentage of cells with apoptotic nuclei are represented as graph. At<br />

least 100 cells per experiment were analyzed from three independent experiments. Data<br />

represents means ± SD of three independent experiments; significantly different from<br />

unirradiated controls: *P < 0.05, **P < 0.01. Significantly different from 6Gy gamma: # P < 0.05.<br />

167


CHAPTER 3<br />

RESULTS<br />

3.4 Radiation induced bystander signaling in mammalian cells.<br />

Radiation induced bystander effect has generated a lot of interest in recent times wherein the<br />

effect of irradiating the target cell on the surrounding unirradiated bystander cells have been<br />

studied. Having established that the response of the signaling factors varied with dose, dose<br />

fractionation, time, microenvironment and radiation quality (LET), it was of interest to<br />

investigate this emerging phenomenon. An attempt has been made to investigate the underlying<br />

mechanism involved and also to determine what component of the signaling system could be<br />

altered by bystander effect and contribution of the bystander effect in the survival of the<br />

neighbouring cells. To understand bystander signaling a number of studies have been made<br />

between the same cell lines. However, not much is known whether bystander effect is extendable<br />

to dissimilar cells, particularly those cells that communicate with each other under physiological<br />

situations. The present study describes the bystander effects of radiation between similar cells<br />

and dissimilar cells in vitro. Gap junction independent signaling mechanism was looked into<br />

using human A549 cells where they were exposed to 2Gy of 60 Co γ irradiation, the medium<br />

isolated and added to fresh cells. Bystander signaling was also studied between U937 cells<br />

(human monocyte) and Lung carcinoma A549 cells. Before designing the experiment there were<br />

two aspects that had to be taken care of, one was the likelihood that the replacement of the<br />

medium itself caused transient expression of certain signaling factors and second that the<br />

components of the medium, which include proteins like growth factors, etc. which are<br />

unavoidably exposed to radiation, contribute to the changes in the bystander cell. The<br />

experimental design, therefore, consisted of the following sets: (a) control unirradiated cells, (b)<br />

γ irradiated cells, (c) A549 cells receiving medium from unirradiated A549 cells or U937 cells or<br />

168


CHAPTER 3<br />

RESULTS<br />

PMA stimulated U937 cells (d) A549 cells receiving medium from γ irradiated A549 cells or<br />

irradiated U937 cells or PMA stimulated and irradiated U937 cells and (e) A549 cells receiving γ<br />

irradiated medium.<br />

3.4.1 Bystander effect in similar cells (A549 Vs A549)<br />

Irradiated A549 cells as well as those cultured in the presence of medium from γ-irradiated A<br />

549 cells showed decrease in survival, increase in γ-H2AX and pATM foci (Fig.3.4.1). However,<br />

the cells which had received only irradiated medium or medium from unirradiated control cells<br />

did not show decrease in survival or increase in γ-H2AX and pATM foci as compared to that in<br />

unirradiated control.<br />

3.4.2 Bystander effect in dissimilar cells (A549 Vs U937)<br />

Having confirmed the existence of bystander signaling between similar cells (A549 Vs A549),<br />

the work was extended to determine if bystander effects can be demonstrated between different<br />

cell types (A549 Vs U937). Bystander signaling was also found to exist between U937 cells<br />

(human monocyte) and Lung carcinoma A549 cells. PMA stimulated and irradiated U937 cells<br />

induced bystander response in unirradiated A549 cells. The latter showed a decrease in survival,<br />

increase in γ-H2AX and pATM foci (Fig.3.4.2). The unstimulated and/or irradiated U937 cells<br />

did not induce bystander effect in unirradiated A549 cells.<br />

169


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.1. Irradiated A549 cells as well as those cultured in the presence of medium from γ-<br />

irradiated A 549 cells (Bystander) showed decrease in survival, increase in γ-H2AX and pATM<br />

foci<br />

170


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.2. Existance of cross bystander effect between Human Lung carcinoma A549 cells and<br />

Human Monocyte U937 cells. Key: Lane 1, control unirradiated A549 cells; Lane 2, unirradiated<br />

A549 cells receiving medium from PMA-stimulated (32nM) and γ irradiated U937 cells; Lane 3,<br />

unirradiated A549 cells receiving medium from γ irradiated U937 cells without PMAstimulation;<br />

Lane 4, unirradiated A549 cells receiving medium from PMA-stimulated U937<br />

cells; Lane 5, unirradiated A549 cells receiving medium from unirradiated U937 cells without<br />

PMA-stimulation.<br />

171


CHAPTER 3<br />

RESULTS<br />

Having established the bystander effect in human cell lines, the mechanism of such a bystander<br />

effect was investigated in mouse cell line.<br />

The experimental design, consisted of the following sets: (a) control unirradiated cells, (b) γ<br />

irradiated cells, (c) EL-4 cells receiving medium from unirradiated EL-4 cells or RAW 264.7<br />

cells or LPS stimulated RAW 264.7 cells (d) EL-4 cells receiving medium from γ irradiated EL-<br />

4 cells or irradiated RAW 264.7 cells or LPS stimulated and irradiated RAW 264.7 cells and (e)<br />

EL-4 cells receiving γ irradiated medium. The cells (EL-4 and RAW 264.7) were also divided<br />

into two groups; one was treated with 20 µM cPTIO (NO scavenger) and the other with 1 mmol<br />

L-NAME (NOS inhibitor). L-NAME was washed off after 30 min so that it is not present in the<br />

bystander cells during incubation.<br />

3.4.3 Bystander effect in similar cells (EL-4 Vs EL-4)<br />

3.4.3.1 Gene expression of NF-κB (p65), iNOS, p53 and p21 in irradiated and bystander cells<br />

The relative gene expression in EL-4 cells cultured for 2 h in presence of different conditioned<br />

media was studied. Actin-β gene expression in each group was used as internal control.<br />

Expression of NF-κB (p65) and iNOS genes were significantly upregulated in irradiated as well<br />

as in bystander cells as compared to the unirradiated control (Fig.3.4.3.1 A and B, Lane, 2 and<br />

3), but their expression was not altered in the cells that had received only irradiated medium or<br />

medium from unirradiated control cells (Fig.3.4.3.1A and B, Lane, 4 and 5). Similarly, the<br />

expression of p53 and p21 genes were found to be significantly upregulated in irradiated and<br />

bystander cells as compared to the control (Fig.3.4.3.1 C and D, Lane, 2 and 3). However, in the<br />

cells cultured in presence of irradiated medium there was a marginal up-regulation of p53 and no<br />

172


CHAPTER 3<br />

RESULTS<br />

up-regulation of p21 (Fig.3.4.3.1 C and D, Lane, 4). The expression of p53 gene in these cells<br />

was significantly lower than irradiated cells as well as bystander cells. The cells cultured in the<br />

presence of medium derived from unirradiated cells did not show up-regulation either p53 or p21<br />

gene expression (Fig.3.4.3.1 C and D, Lane 5).<br />

3.4.3.2 NO production in irradiated and bystander cells<br />

Since the expressions of iNOS gene was significantly upregulated in irradiated and bystander<br />

cells (Fig.3.4.3.1B, Lane 2 and 3), it was of interest to look for NO production in these cells. NO<br />

production was found to be significantly enhanced in irradiated cells and bystander cells as<br />

compared to that in unirradiated control cells (Fig. 3, Lane 2 and 3). The NO production in the<br />

cells cultured in presence of only irradiated medium or medium derived from unirradiated cells<br />

was found to be similar to that in unirradiated cells (Fig.3.4.3.2, Lane 4 and 5).<br />

3.4.3.3 DNA damage in irradiated and bystander cells<br />

Damage to cellular DNA was expressed as tail length (Fig.3.4.3.3 A), per cent DNA in tail<br />

(Fig.3.4.3.3 B), Olive tail moment (Fig.3.4.3.3 C) and Tail moment (Fig.3.4.3.3 D). The DNA<br />

damage was significantly higher in irradiated EL-4 cells as compared to that in unirradiated cells<br />

(Fig.3.4.3.3, A, B, C, D, Lane 2). The cells cultured in the presence of medium derived from<br />

irradiated cells also showed a significantly higher DNA damage as compared to the unirradiated<br />

control cells (Fig.3.4.3.3, A, B, C, D, Lane 3). However, the cells that had received only<br />

irradiated medium or medium from unirradiated control cells did not show any increase in DNA<br />

damage as compared to that in unirradiated control (Fig.3.4.3.3, A, B, C, D, Lane 4 and 5).<br />

173


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.3.1 Gene expression of iNOS, p21, p53 and NF-kB (p65) in EL-4 cells cultured for 2 h in<br />

presence of different conditioned media. Total RNA from EL-4 cells was isolated and reverse<br />

transcribed. RT-PCR analysis of iNOS, p21, p53 and NF-kB (p65) genes was carried out as<br />

described in materials and methods. PCR products were resolved on 1.5% agarose gels<br />

containing ethidium bromide. Actin-β gene expression in each group was used as an internal<br />

control. Ratio of intensities of (A) NF-kB (p65), (B) iNOS, (C) p53 and (D) p21band to that of<br />

respective actin- β band as quantified from gel pictures are shown above each gel picture. Key:<br />

Lane 1, control unirradiated EL-4 cells; Lane 2, γ irradiated EL-4 cells; Lane 3, unirradiated<br />

EL-4 cells receiving medium from γ irradiated EL-4 cells; Lane 4, unirradiated EL-4 cells<br />

receiving medium irradiated with γ rays in the absence of cells; Lane 5, unirradiated EL-4 cells<br />

receiving medium from unirradiated EL-4 cells. Data represents means ± SE of three<br />

independent experiments; significantly different from unirradiated controls: *P < 0.05, **P <<br />

0.01.<br />

174


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.3.2 NO production from EL-4 cells cultured for 24 h in the presence of different<br />

conditioned media. The culture supernatant from each group of EL-4 cells was used for<br />

-<br />

determination of NO 2 with Griess reagent. Key: Lane 1, control unirradiated EL-4 cells; Lane 2,<br />

γ irradiated EL-4 cells; Lane 3, unirradiated EL-4 cells receiving medium from γ irradiated EL-4<br />

cells; Lane 4, unirradiated EL-4 cells receiving medium irradiated with γ rays in the absence of<br />

cells; Lane 5, unirradiated EL-4 cells receiving medium from unirradiated EL-4 cells. Data<br />

represents means ± SE of three independent experiments; significantly different from<br />

unirradiated controls: *P < 0.05, **P < 0.01.<br />

175


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.3.3 DNA damage in EL-4 cells cultured for 2 h in presence of different conditioned<br />

media was measured using single cell gel electrophoresis (‘comet assay’), (A) Tail length, (B) %<br />

DNA in tail, (C) Olive tail moment, and (D) Tail moment. Key: Lane 1, control unirradiated EL-<br />

4 cells; Lane 2, γ irradiated EL-4 cells; Lane 3, unirradiated EL-4 cells receiving medium from γ<br />

irradiated EL-4 cells; Lane 4, unirradiated EL-4 cells receiving medium irradiated with γ rays in<br />

the absence of cells; Lane 5, unirradiated EL-4 cells receiving medium from unirradiated EL-4<br />

cells; Lane 6, unirradiated EL-4 cells receiving fresh medium. Data represents means ± SE of<br />

three independent experiments; significantly different from unirradiated controls: *P < 0.05, **P<br />

< 0.01.<br />

176


CHAPTER 3<br />

RESULTS<br />

3.4.3.4 Apoptotic DNA fragmentation in irradiated and bystander cells (EL-4 Vs EL-4)<br />

There is a clear DNA ladder formation (a hallmark of apoptosis) in irradiated and bystander EL-<br />

4 cells. But the extent of apoptosis was found to be more in irradiated cells as compared to that in<br />

bystander cells (Fig. 4, Lane 2 and 3). There was no increase in apoptosis in the cells that had<br />

received only irradiated medium (Fig. 3.4.3.4a, Lane 4) or medium from unirradiated control<br />

cells (Fig. 3.4.3.4a, Lane 5) as compared to that in unirradiated cells (Fig. 3.4.3.4a, Lane 1).<br />

Apoptosis in irradiated and bystander EL-4 cells was further confirmed by annexin V/PI assay<br />

(Fig. 3.4.3.4b).<br />

Fig. 3.4.3.4a DNA fragmentation assay of EL-4 cells. EL-4 cells cultured for 24 h in the<br />

presence of different conditioned media. DNAs of each group of cells were isolated as described<br />

in materials and methods section and electrophoresed on 1.8% Agarose gel containing ethidium<br />

bromide. Key: Lane M, DNA molecular weight marker; Lane 1, control unirradiated EL-4 cells;<br />

Lane 2, γ irradiated EL-4 cells (5Gy); Lane 3, unirradiated EL-4 cells receiving medium from γ<br />

irradiated EL-4 cells; Lane 4, unirradiated EL-4 cells receiving medium irradiated with γ rays in<br />

the absence of cells; Lane 5, unirradiated EL-4 cells receiving medium from unirradiated EL-4<br />

cells. Data shown from representative experiment of three independent experiments.<br />

177


CHAPTER 3<br />

RESULTS<br />

Fig. 3.4.3.4b Apoptosis assay of Control, Irradiated and Bystander EL-4 cells by Annexin-PI<br />

staining. EL-4 cells cultured for 18 h in the presence of different conditioned media. The EL-4<br />

cells were stained using an annexin V/PI-staining kit. Characterization and quantification of<br />

apoptosis was done by confocal microscopy. Key: A, Annexin-V; B, PI; C, DAPI; D, Merge<br />

image of annexin-v, PI and DAPI. Data represents means ± SE of three independent<br />

experiments; significantly different from unirradiated controls: **P < 0.01.<br />

178


CHAPTER 3<br />

RESULTS<br />

3.4.4 Bystander effect in dissimilar cells (EL-4 Vs RAW 264.7)<br />

Having confirmed the existence of bystander signaling between similar cells (EL-4 Vs EL-4), the<br />

work was extended to determine if bystander effects can be demonstrated between different cell<br />

type (RAW 264.7 and EL-4) and whether the state of stimulation of macrophages (RAW 264.7)<br />

affects bystander response. The expression of NF-κB as well as iNOS genes were significantly<br />

upregulated in unirradiated EL-4 cells that had received medium from LPS stimulated and<br />

irradiated RAW 264.7 cells (Fig. 3.4.4a, Compare Lane, 2 and Lane 1); but the expression of<br />

NF-κB was not altered in the cells that had received medium from either irradiated RAW 264.7<br />

cells or unstimulated RAW 264.7 cells (Fig.3.4.4a, Compare Lane, 3 and 5 Vs Lane 1), however,<br />

the expression of iNOS was down regulated in the above groups. The EL-4 cells cultured in the<br />

presence of medium derived from LPS stimulated RAW 264.7 cells showed only marginal<br />

upregulation of iNOS gene but not NF-κB gene (3.4.4a, Compare Lane 4 and Lane 1).<br />

DNA damage in the above groups of EL-4 cells cultured 2 h after exposure to different<br />

conditioned media was significantly higher in unirradiated EL-4 cells that had received medium<br />

from LPS stimulated and irradiated RAW 264.7 cells (Fig.3.4.4b, Compare Lane, 2 and Lane 1).<br />

However, the cells that had received medium from irradiated RAW 264.7 cells, LPS stimulated<br />

RAW 264.7 cells or unstimulated RAW 264.7 cells did not show any increase in DNA damage<br />

(Fig.3.4.4b, Lane, 3, 4 and 5).<br />

179


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.4a Gene expression of NF-kB (p65) and iNOS gene in EL-4 cells which had received<br />

medium from irradiated RAW 264.7 cells. EL-4 cells cultured for 2 h in presence of different<br />

conditioned media. β-Actin gene expression in each group was used as an internal control. Ratio<br />

of intensities of NF-kB (p65) and iNOS band to that of respective β-actin band as quantified<br />

from gel pictures are shown above each gel picture. Key: Lane 1, control EL-4 cells; Lane 2,<br />

EL-4 cells receiving medium from LPS-stimulated (500ng/ml) and γ irradiated RAW 264.7 cells;<br />

Lane 3, EL-4 cells receiving medium from γ irradiated RAW 264.7 cells without LPSstimulation;<br />

Lane 4, EL-4 cells receiving medium from LPS-stimulated RAW 264.7 cells; Lane<br />

5, EL-4 cells receiving medium from unirradiated RAW 264.7 cells. Data represents means ± SE<br />

of three independent experiments; significantly different from unirradiated controls: *P < 0.05,<br />

**P < 0.01.<br />

180


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.4b DNA damage in EL-4 cells cultured for 2 h in presence of different conditioned<br />

media was measured using single cell gel electrophoresis (‘comet assay’) and expressed as tail<br />

length. Key: Lane 1, control EL-4 cells; Lane 2, EL-4 cells receiving medium from LPSstimulated<br />

(500ng/ml) and γ irradiated RAW 264.7 cells; Lane 3, EL-4 cells receiving medium<br />

from γ irradiated RAW 264.7 cells without LPS-stimulation; Lane 4, EL-4 cells receiving<br />

medium from LPS-stimulated RAW 264.7 cells; Lane 5, EL-4 cells receiving medium from<br />

unirradiated RAW 264.7 cells. Data represents means ± SE of three independent<br />

3.4.5 Effect of L-NAME and cPTIO on Bystander response.<br />

Since the expressions of NF-κB and iNOS gene were significantly upregulated in irradiated as<br />

well as in bystander cells (Fig.3.4.5a, Lane 2 and 3), it was of interest to determine which<br />

181


CHAPTER 3<br />

RESULTS<br />

signaling factors were involved in bystander response. The response was looked at in the<br />

presence of cPTIO or L-NAME. L-NAME was washed out before irradiation so that the<br />

bystander cells are not exposed to L-NAME and iNOS is inhibited only in the irradiated cells.<br />

The addition of cPTIO, which scavenges the NO released in the medium, did not affect the<br />

expression of iNOS gene if the bystanding cell was of same origin, although NF-κB was<br />

somewhat inhibited (Fig.3.4.5a, Compare Lane 5 and Lane 3). The addition of L-NAME, which<br />

inhibits the iNOS in the irradiated cells, completely inhibited the bystander response in the EL-4<br />

cells (Fig.3.4.5a, Compare Lane 4 and Lane 3). However both L-NAME and cPTIO inhibited<br />

the expression of NF-κB and iNOS in the bystander EL-4 cells if the medium was from LPS<br />

stimulated and irradiated RAW 264.7 cells (Fig.3.4.5a, Lane 6 and 7) i.e. the cell type was<br />

different, indicating that the bystander response was dependent on the cells type and the state of<br />

stimulation of the irradiated cells.<br />

The DNA damage was significantly higher in irradiated as well as in bystander cells (Fig.3.4.5b,<br />

Lane 2 and 3) as compared to the unirradiated control. However, the unirradiated EL-4 cells that<br />

had received medium either from L-NAME treated and irradiated EL-4 cells or from LPS<br />

stimulated, L-NAME treated and irradiated RAW 264.7 cells did not show any increase in DNA<br />

damage as compared to that in unirradiated control (Fig.3.4.5b, Lane 4 and 6). Similarly the<br />

unirradiated EL-4 cells that had received medium either from cPTIO treated and irradiated EL-4<br />

cells or from LPS stimulated, cPTIO treated and irradiated RAW 264.7 cells did not show any<br />

increase in DNA damage as compared to that in unirradiated control (Fig.3.4.5b, Lane 5 and 7).<br />

Neither L-NAME nor cPTIO itself had any effect on DNA damage in EL-4 cells (Fig.3.4.5b,<br />

Lane 8 and 9).<br />

182


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.5a Gene expression of NF-kB (p65) and iNOS gene in EL-4 cells which had received<br />

medium from irradiated cells. EL-4 cells cultured for 2 h in presence of different conditioned<br />

media. β-Actin gene expression in each group was used as an internal control. Ratio of intensities<br />

of NF-kB (p65) and iNOS band to that of respective β-actin band as quantified from gel pictures<br />

are shown above each gel picture. Key: Lane 1, control EL-4 cells; Lane 2, γ irradiated EL-4<br />

cells; Lane 3, EL-4 cells receiving medium from γ irradiated EL-4 cells; Lane 4, EL-4 cells<br />

receiving medium from L-NAME treated and γ irradiated EL-4 cells; Lane 5, EL-4 cells<br />

receiving medium from cPTIO treated and γ irradiated EL-4 cells; Lane 6, EL-4 cells receiving<br />

medium from LPS-stimulated (500ng/ml), L-NAME treated and γ irradiated RAW 264.7 cells;<br />

Lane 7, EL-4 cells receiving medium from LPS-stimulated (500ng/ml), cPTIO treated and γ<br />

irradiated RAW 264.7 cells. Data represents means ± SE of three independent experiments;<br />

significantly different from unirradiated controls: *P < 0.05, **P < 0.01.<br />

183


CHAPTER 3<br />

RESULTS<br />

Fig.3.4.5b DNA damage in EL-4 cells cultured for 2 h in presence of different conditioned<br />

media was measured using single cell gel electrophoresis (‘comet assay’) and expressed as tail<br />

length. Key: Lane 1, control EL-4 cells; Lane 2, γ irradiated EL-4 cells; Lane 3, EL-4 cells<br />

receiving medium from γ irradiated EL-4 cells; Lane 4, EL-4 cells receiving medium from L-<br />

NAME treated and γ irradiated EL-4 cells; Lane 5, EL-4 cells receiving medium from cPTIO<br />

treated and γ irradiated EL-4 cells; Lane 6, EL-4 cells receiving medium from LPS-stimulated<br />

(500ng/ml), L-NAME treated and γ irradiated RAW 264.7 cells; Lane 7, EL-4 cells receiving<br />

medium from LPS-stimulated (500ng/ml), cPTIO treated and γ irradiated RAW 264.7 cells;<br />

Lane 8, EL-4 cells treated with L-NAME; Lane 9, EL-4 cells treated with cPTIO. Data<br />

represents means ± SE of three independent experiments; significantly different from<br />

unirradiated controls: *P < 0.05, **P < 0.01.<br />

184


CHAPTER 4<br />

DISCUSSION<br />

DISCUSSION<br />

185


CHAPTER 4<br />

DISCUSSION<br />

Ionising radiation (IR) has been used for nearly a century to treat human cancers. The objective<br />

of radiotherapy is to deliver a lethal dose to cancer cells but attenuate the toxic effects of IR on<br />

adjacent normal tissue. In order to achieve this, radiotherapy is given as fractionated dose<br />

ranging from 2-4 Gy per fraction. Therefore, the manner in which cell senses and respond to<br />

radiation damage is critical for the outcome of radiotherapy. The molecular pathways that are<br />

triggered by acute dose of irradiation will be different from fractionated dose of irradiation. This<br />

will ultimately determine the multiple possible responses, whether they be DNA repair, cell<br />

cycle arrest, cell death or cell adaptive response. Moreover, these pathways provide molecular<br />

explanation of why different cell types differ in their response to radiation. Some cell types are<br />

relatively more radioresistant than others. The reason for the radioresistance could, therefore be<br />

efficient DNA repair, but cannot be the sole reason for it. The contribution of signaling to these<br />

phenomena could be enormous. Radioresistant cells and cells that become radioresistant after<br />

radiation treatment need newer modalities of irradiation like heavy ions because the relative<br />

biological effectiveness (RBE) increases with an increase in the LET of the incident radiation.<br />

The effectiveness of Ionising radiation is because it induces DNA damage, which generates a<br />

complex cascade of events leading to cell cycle arrest, transcriptional and post-transcriptional<br />

activation of a subset of genes including those associated with DNA repair and triggering<br />

apoptosis. Probably the most dangerous of all the types of DNA damage are double strand breaks<br />

(DSBs) and their repair is complex. Moreover, the degree of lesion complexity increases with<br />

increasing LET. The cellular genotoxic response depends on the type of DNA damage which in<br />

turn can evoke unique cellular response.<br />

The signaling events seen in totality may also have contribution from the surrounding<br />

microenvironment that is of necessity exposed to radiation. The cells nearby which are not<br />

186


CHAPTER 4<br />

DISCUSSION<br />

directly exposed to the radiation but are juxtaposed to the irradiated cells are also affected<br />

because of signals transduced from the ‘hit’ cell. The classical paradigm of radiation biology<br />

asserts that all radiation effects on cells, tissues and organisms are due to the direct action of<br />

radiation. However, there has been a recent growth of interest in the indirect actions of radiation<br />

including the radiation-induced adaptive response, the bystander effect, low-dose<br />

hypersensitivity, and genomic instability, which are specific modes of stress exhibited in<br />

response to low-dose/low-dose rate radiation.<br />

4.1 Fractionated irradiation induced signaling and repair in mammalian cells.<br />

It is known that cellular response to ionising radiation is a complex phenomenon where the dose,<br />

dose rate and its fractionation play an equally important role in deciding the fate of the cell.<br />

Extensive work has been done and published on the response of the cell to single doses of<br />

radiation which are reflected as lesions in the DNA and non DNA targets. Among the latter are<br />

the signaling proteins that act to signal that the DNA is damaged as well as to regulate processes<br />

such as cell cycle progression and DNA repair. These pathways are also intricately linked with<br />

the intrinsic radioresistance of the cell [237, 238]. In recent years these signaling proteins are<br />

being exploited as targets to enhance tumor radiosensitivity [238-240]. The targets have been<br />

chosen based on their activation following a single dose of radiation. Since signaling, as it is now<br />

known, is not a linear activation of a pathway but a complex network where both positive and<br />

negative signals are activated simultaneously following exposure to stress , it is quite likely that<br />

the signaling factors that are being targeted may not be activated in a cancer cell that has<br />

undergone radiotherapy. The study of the ultimate radioresistant phenotype is undoubtedly a<br />

very important aspect but the immediate response of the signaling factors to repeated doses of<br />

187


CHAPTER 4<br />

DISCUSSION<br />

radiation are yet unknown. More information about the response of these signaling factors at<br />

clinically relevant doses during and after a fractionated regimen is needed to understand their<br />

role.<br />

In the present study human lung adenocarcinoma A549 cells were found to be more<br />

radioresistant if 10 Gy dose was delivered as fractionated dose (Fig.3.1.1). Although<br />

fractionation is supposed to play an important role in inducing an adaptive response and deciding<br />

the fate of cells, the mechanism by which it does so can be many and innumerable factors could<br />

be responsible. In microarray analysis, most of the pathways which were up-regulated in A549<br />

cells exposed to fractionated irradiation in all comparisons were associated with the survival or<br />

repair of the A549 cells. This indicated that fractionated irradiation could induce up-regulation of<br />

survival/repair related pathways in cancer cells.<br />

Previous reports concerning radioresistant non–small-cell lung cancer and HeLa cells<br />

indicated that cell cycle signaling pathways, mismatch repair, homologous recombination were<br />

up-regulated [241, 242]. In this study, some of these genes were previously known to be<br />

associated with responsiveness to radiation, such as p21 and GADD45α, but others were novel.<br />

These novel genes can be expected to be involved in radioresistance, but the precise function of<br />

each gene remains unclear; so further study was necessary to clarify the nature of associations.<br />

Microarray data had shown that p53 signaling pathway was up-regualted in A549 cells<br />

that had been exposed to fractionated irradiation; it was of interest to look at the phosphorylation<br />

of p53 and its translocation to the nucleus of A549 cells. Lung adenocarcinoma A549 cells<br />

typically express wild-type p53 protein [243] and would be expected to be sensitive to the DNAdamaging<br />

agents used for cancer therapy; however, these cells are radioresistant if the radiation<br />

188


CHAPTER 4<br />

DISCUSSION<br />

dose is delivered as fractionated irradiation. p53 classically considered to be a tumor suppressor<br />

protein, but in this study the results are contradictory. The fact that A549 cells express wild-type<br />

p53 protein, it was unable to suppress the proliferation of A549 cells that had been exposed to<br />

fractionated irradiation.<br />

Wild-type p53 function involves two steps. First, upstream signaling pathways stabilize and<br />

activate the p53 protein in response to DNA damage or other forms of cellular stress. In the<br />

second step, activated p53 binds to regulatory sequences of its target genes and activates their<br />

expression [244]. Either step may be defective in lung adenocarcinoma A549 cells. Defective<br />

activation of wild-type p53 after DNA damage could be the mechanism for suppression of p53<br />

function in A549 cells. The response of p53 to DNA damage involves an increase in p53 protein<br />

levels due to stabilization of the protein and an increase in p53 functional activity [245, 246]. It is<br />

quite likely that there is no defect in regulation of p53 stabilization in response to DNA damage<br />

because p53 was stabilized in response to fractionated irradiation in A549 cell line examined, and<br />

the steady-state levels of p53 were very high (Fig. 3.1.7A, B & C). The regulation of the<br />

functional activity of p53 seems to be defective. These findings are consistent with previous<br />

reports showing the function of wild-type p53 protein is apparently compromised in many cancer<br />

cells [247, 248]. Moreover, translocation of phospho-p53 into the nucleus could lead to cell cycle<br />

arrest, which will allow the cells to repair the DNA damage.<br />

Since DNA repair pathways were up-regulated in A549 cells that had been exposed to<br />

fractionated irradiation, it was of interest to look at the activation of genes and proteins involved<br />

in DNA repair pathways. Results of this study indicated that there is intense activation of repair<br />

genes and protein in A549 cells exposed to fractionated irradiation (Fig.3.1.4 and Fig.3.1.6).<br />

189


CHAPTER 4<br />

DISCUSSION<br />

These results on DNA-PK are consistent with earlier works where authors have also shown the<br />

role of DNA-PK in the radioresistance of lung carcinoma cells [249, 250]. However, these<br />

authors have looked at the response only after single dose of irradiation and it is logical to expect<br />

these pathways to get activated. In this study it was shown that DNA-PK is also involved in<br />

radioresistance if the cells are subjected to fractionated irradiation but the reason for the<br />

development of radioresistance still remains an enigma.<br />

ATM and BRCA1 have been regarded as primary regulators of homologous<br />

recombination repair (HRR) [251]. In this study an intense activation of ATM gene and ATM<br />

foci were seen in all the cells exposed to fractionated irradiation and most of the cells showed<br />

BRCA1 foci (Fig.3.1.4 and Fig.3.1.6) indicating the fact that HRR pathway was activated in<br />

A549 cells and therefore, these proteins could be involved in HRR which can take advantage of<br />

the other strand that may be intact. ATM and BRCA1 reside in macromolecular complex and<br />

form foci after irradiation. These results are in agreement with earlier works where authors have<br />

also shown the role ATM in radioresistance of cancer cells [252, 253]. However, these authors<br />

have studied using single dose of irradiation.<br />

The repair of DNA at 4h in A549 cells exposed to fractionated irradiation but not in cells<br />

that had been exposed to 10 Gy acute dose, indicating the fact that fractionated irradiation<br />

induced better repair capability of the cells. Moreover, the DNA repair pathways were activated<br />

in A549 cells exposed to fractionated irradiation allow these cells to repair its damage DNA<br />

faster than 10 Gy acute dose.<br />

190


CHAPTER 4<br />

DISCUSSION<br />

A clear cause and effect relationship was established between DNA damage signaling,<br />

gene expression and efficient DNA repair and was evident in the form of cell survival in A549<br />

cells that had been exposed to fractionated irradiation.<br />

It has been reported that A549 cells are relatively more radioresistant than MCF-7 cells if<br />

radiation dose was delivered as fractionated regimen [254]. But the mechanism of such response<br />

has not been studied. To best of my knowledge this is the first study explaining the mechanism<br />

of such response as activation of repair pathway, efficient DNA repair and translocation of<br />

phospho-p53 into the nucleus of A549 cells exposed to fractionated irradiation and not in MCF-7<br />

cells exposed to fractionated irradiation (Fig. 3.1.8).<br />

The level of mRNA expression is known to be modulated through transcriptional and<br />

posttranscriptional control mechanisms. An important posttranscriptional control is exerted on<br />

mRNA stability, which varies considerably from one mRNA species to another and can be<br />

modulated by various stimuli such as radiation [255-257].How these stimuli influence mRNA<br />

halflives in ways that are incompletely understood. Although much has been researched about<br />

radioresistance and DNA damage, surprisingly there are absolutely no reports on the stabilization<br />

of mRNA of crucial DNA damage and maintenance of proteins. It is quite likely that mRNA<br />

stabilization could contribute to the development of radio resistance. The expression of DNA-PK<br />

was partly due to up-regulation of DNA-PK transcriptional activity to some degree, DNA-PK<br />

mRNA stabilization played important roles in maintaining high DNA-PK expression level in this<br />

study (Fig. 3.1.9). The ATM and Rad52 mRNA were not stabilized indicating a specific and<br />

selective stabilization of DNA-PK mRNA. The DNA-PK mRNA stabilization was mediated by<br />

p38 MAP kinase signaling pathways but not by ERK or JNK MAP Kinase pathway (Fig. 3.1.9).<br />

191


CHAPTER 4<br />

DISCUSSION<br />

This result indicates that there is existence of cross talk between DNA repair pathway and p38<br />

MAP Kinase pathway. The p38 MAP kinase pathway plays an important role in the posttranscriptional<br />

regulation of many genes. p38 has been found to regulate both the translation and<br />

the stability of inflammatory mRNAs. The mRNAs regulated by p38 share common AU-rich<br />

elements (ARE) present in their 3’-untranslated regions. AREs act as mRNA instability<br />

determinants but also confer stabilization of the mRNA by the p38 pathway. In recent years,<br />

AREs have shown to be binding sites for numerous proteins [258]. How p38 MAP Kinase<br />

pathway regulates DNA-PK mRNA stability remains to be investigated.<br />

Numerous reports have implicated DNA repair genes as being mainly responsible for the<br />

development of radioresistance due to fractionated irradiation [249, 250, 252, 253]. Processing of<br />

DNA double strand breaks (DSB) in eukaryotes is most likely achieved by multiple pathways<br />

including homologous recombination. Although Rad52 has been shown to be important in DSB<br />

repair in yeast, its role in mammalian cells has not been demonstrated. In this study, silencing of<br />

Rad52 gene in A549 cells by transfection followed by exposure to fractionated irradiation<br />

showed increased sensitivity of these cells to fractionated irradiation. This observation suggests<br />

that homologous recombination repair mediated by Rad52 is involved in double strand break<br />

repair in A549 cells.<br />

In summary, these results indicated that A549 cells were relatively more radioresistant if<br />

these cells were exposed to fractionated irradiation. The reason for radioresistance could be the<br />

activation of DNA repair pathways and Rad52 gene is an important factor modulating<br />

radioresistance in A549 cells.<br />

192


CHAPTER 4<br />

DISCUSSION<br />

4.2 Proton beam induced signaling in mammalian cells and its comparison with Gamma<br />

irradiation.<br />

After establishing that A549 cells were relatively more radioresistant if these cells were exposed<br />

to fractionated irradiation, the variance in signaling pattern and effectiveness of cell killing with<br />

the change in LET was studies in A549 cells.<br />

More than 2 fold decrease in survival of the A549 cells as assessed by the clonogenic cell<br />

survival assay was indicative of the fact that the proton beam is more efficient in killing the<br />

tumor cells and may be a more preferred mode of treatment (Fig.3.2.1). The mechanisms of cell<br />

death caused by proton treatment and the signaling pathways that ensue have not yet been<br />

investigated in detail although proton beam therapy is in use for many cancers [259-264].<br />

The massive activation of DNA-PK and ATM after proton irradiation when compared<br />

with gamma (Fig.3.2.2 A &B) was more or less expected and may be due to the fact that DNA<br />

damage activates Phosphatidylinositol 3-kinase-like kinases (DNA-PK, ATM), which then<br />

amplify and channel the signal by activating downstream kinases (Chk1 and Chk2). These, in<br />

turn, phosphorylate target proteins such as p53, Cdc25A and Cdc25C [11] to delay the cell cycle,<br />

a process called the DNA damage checkpoint. In addition to checkpoint control, the downstream<br />

kinases also modulate DNA repair and trigger apoptosis [13]. The lack of activation of ATR after<br />

proton irradiation was however intriguing and was found to be reflected no activation of Chk1<br />

(Fig.3.2.2).<br />

Phosphorylation of H2AX even at 4 hours in A549 cells that had been exposed to proton<br />

beam indicates the presence of large number of double strand breaks at that time indicating very<br />

193


CHAPTER 4<br />

DISCUSSION<br />

slow repair. DNA repair pathways activated by charged particle radiation might be different from<br />

those studied after gamma irradiation.<br />

Likewise, there was significantly higher phosphorylation of ATM at serine 1981 in A549 cells<br />

that had been exposed to proton beam (Fig.3.2.3), indicating the fact that despite significant<br />

activation of ATM (4 hours) much of the damage still persisted. The activated/phosphorylated<br />

ATM would further phosphorylate various downstream target proteins such as H2AX, Chk2, p53<br />

etc. leading to opening of the chromatin and recruitment of other proteins involved in DNA<br />

repair or cell cycle arrest.<br />

Phosphorylation of Chk2 indicated that small component of ATM activation is also diverted to<br />

cell cycle arrest pathways such as Chk2, CDC25 phosphatase and CDKs. These could also be<br />

involved in activation of other proteins such as BRCA1 and p53. Phosphorylation of p53 at<br />

serine 15 was then observed in unirradiated and irradiated A549 cells at 2Gy of proton beam<br />

(Fig.3.2.3). These results indicated cell cycle arrest in A549 cells exposed proton beam.<br />

Whether DNA-PK signals DSBs to the checkpoint machinery remains controversial, but the<br />

emerging consensus is that it does not [22]. DNA-PK may, however, be involved in signaling<br />

DNA damage to the apoptosis machinery [24].<br />

The degree of radiation-induced apoptosis has been shown to correlate with the p53 wild-type<br />

status [265]. In addition, apoptosis is induced when wild-type p53 is transfected into certain cell<br />

lines lacking p53 [266]. This indicates that p53 not only plays a role in regulating the progression<br />

through the cell cycle, it can also induce apoptosis in cells. p53 is not an essential component of<br />

the machinery that carries out apoptosis; however, it appears to be an activator of apoptosis.<br />

194


CHAPTER 4<br />

DISCUSSION<br />

Although the exact means by which p53 activates apoptosis is unclear, evidence has shown that<br />

p53 mediates apoptosis by way of transcription-independent and transcription-dependent<br />

mechanisms [267, 268]. p53 is known to regulate the expression of several proteins involved in<br />

the apoptotic pathway and also interacts with BAX, BCL-XL, and BCL-2 to exert a direct<br />

apoptotic effect at the level of the mitochondria [111, 269, 270].<br />

Increased phosphorylation of p53, upregulation of Bax and down-regulation of Bcl-2 in proton<br />

beam irradiated A549 cells as compared to γ-irradiated cells (Fig.3.2.3 and Fig.3.2.4) indicate the<br />

activation of apoptotic pathways.<br />

The noteworthy finding of this study is the biphasic activation of the sensor proteins, ATM and<br />

DNA-PK and no activation of ATR by proton irradiation and the significant activation of Chk2<br />

even at the gene level only in the proton beam irradiated cells (Fig.3.2.2). Such biphasic response<br />

after irradiation with proton or other heavy ions have been observed and may be characteristics<br />

of particle irradiation [271]. Unlike gamma irradiation, the biphasic induction of ATM and<br />

DNA-PK following proton beam irradiation could be responsible for the activation of apoptotic<br />

machinery, however, further studies in this direction will reveal the importance of such biphasic<br />

response.<br />

The use of proton beam therapy has definite advantage over gamma therapy. The mechanism of<br />

apoptosis will give the clinician a handle to manipulate therapy for enhanced cell killing.<br />

195


CHAPTER 4<br />

DISCUSSION<br />

4.3 High LET irradiation induced signaling in mammalian cells and its comparison with<br />

Gamma irradiation.<br />

Although many studies on signaling pathways after low and high LET radiations have found the<br />

activation to be similar except in the intensity [214] but since the end result of the two<br />

irradiations is different, there must be a divergence of pathways at some stage. It is this stage of<br />

signaling, that is different, that will be crucial to designing therapies that target specific<br />

molecules or pathways<br />

4.3.1 Carbon beam induced signaling in mammalian cells and its comparison with Gamma<br />

irradiation.<br />

Although 1Gy dose of carbon ions was three times more toxic to the cells than gamma<br />

rays, both produced same number of γ-H2AX foci, indicators of DNA double strand breaks, the<br />

primary lesions induced by radiation (Fig.3.3.1.1 and Fig.3.3.1.2). This discrepancy between<br />

observed cytotoxicity and estimated γ-H2AX foci after high LET radiation has also been<br />

reported previously [272]. The observed decrease in number of γ-H2AX foci 4 hours after<br />

gamma irradiation indicated that DNA damage induced by low LET was repaired faster than<br />

high LET radiation and is supported by nearly 100% survival. However, the decrease in γ-H2AX<br />

foci after Carbon ion irradiation was definitely not indicative of repair because the cell survival<br />

data contradicts the fact (Fig.3.3.1.2). Since total intensity of γ-H2AX was yet high at 4 hours in<br />

carbon irradiated cells, it indicated diffused staining of γ-H2AX rather than dephosphorylation. It<br />

may represent sites of increased DNA damage arising from misrepair or incomplete repair [273,<br />

274].<br />

196


CHAPTER 4<br />

DISCUSSION<br />

Similar number of γ-H2AX foci at 15 minutes after irradiation irrespective of the radiation<br />

quality indicated that initial events might not differ much with the radiation quality. Activation<br />

of other molecules including ATM, ATR, BRCA1, Chk1, Chk2 and Bax was therefore followed<br />

only at later time point of 4 hours. Moreover, since by this time most of the damage would be<br />

repaired, comparing the activation of these molecules at this time could highlight signaling<br />

differences with respect to clustered damage.<br />

ATM and ATR, the initial DNA damage sensors, are critical regulators of cell cycle checkpoints<br />

and repair. ATM has particularly been shown to respond to IR induced damage. Presence of few<br />

ATM foci in ~90% of the gamma irradiated cells 4 hours after irradiation indicates the repair of<br />

the damage in most cells. However, in carbon-irradiated cells, although the percentage of cells<br />

showing the pATM foci (55%) was less, the average number of foci/cell was more (Fig.3.3.1.3).<br />

Similar trend was also observed with ATR foci with respect to radiation quality (Fig.3.3.1.4).<br />

Higher percentage of gamma irradiated cells showing the ATM/ATR foci 4 hours after<br />

irradiation as compared to carbon irradiated cells indicates more homogenous nature of gamma<br />

radiation induced responses while presence of lesser average ATM/ATR foci per gamma<br />

irradiated cell as against carbon irradiation indicates efficient repair after gamma. This<br />

observation is in agreement with the previous studies where higher activation of repair molecules<br />

had been observed few hours after heavy ion irradiation as compared to low LET irradiation<br />

[214, 275] and has been attributed to unrepaired DNA that keeps the damage sensors in activated<br />

state.<br />

The formation of BRCA1 foci was unlike ATM/ATR and very interesting. BRCA1 exists in a<br />

complex containing ATM and other DSB repair proteins and both ATM and BRCA1 have been<br />

197


CHAPTER 4<br />

DISCUSSION<br />

regarded as primary regulators of HRR [251]. Presence of BRCA1 foci in ~55% (Fig.3.3.1.5) of<br />

the gamma irradiated cells is in agreement with the previous studies where BRCA1 foci have<br />

been found to be activated only in cells in S, G2/M phases of the cycle [275] and therefore is<br />

thought to play a role in HRR which can take the advantage of the other strand that may be<br />

intact. Unlike after gamma irradiation, presence of BRCA1 foci in all of the carbon irradiated<br />

cells was quite intriguing. Moreover, the downstream substrates, Chk1 (Fig.3.3.1.6A) and Chk2<br />

(Fig.3.3.1.6C) also showed relatively higher activation in carbon irradiated cells as compared to<br />

gamma. Activation of all these repair components with high LET radiation at first indicates<br />

persistent and even more intense activation of repair pathways. However, presence of ATM foci<br />

in only half of carbon irradiated cells while BRCA1 foci in all cells suggest that BRCA1 is<br />

associated with different proteins and may not require activated ATM. The larger size of BRCA1<br />

foci after Carbon irradiation also indicated that macromolecular complexes containing BRCA1<br />

might be different after two types of radiation treatments. Number of Chk2 foci, another<br />

important regulator of HRR, also did not seem to parallel ATM foci. This, along with the fact<br />

that the survival of the cells is only 20% (Fig.3.3.1.1) and many apoptotic nuclei are present at 4<br />

hours (Fig.3.3.1.8) indicates that these macromolecular complexes containing BRCA1 may be<br />

majorly involved in apoptotic responses after a complex damage to the DNA. Recent studies<br />

suggest multifunctional nature of BRCA1 in maintaining genome integrity. It has been shown to<br />

play an important role in apoptosis and cell cycle control and does so by associating in distinct<br />

protein complexes [276].<br />

The cytoplasmic MAPK pathways, ERK and JNK, that feed into and are fed upon by DNA<br />

damage repair signaling also play an equally important role in deciding the fate of the irradiated<br />

cell. Reactive oxygen species generated after gamma irradiation have been thought to be one of<br />

198


CHAPTER 4<br />

DISCUSSION<br />

the mechanism for early activation of these kinases [277]. ATM has been shown to enhance<br />

repair via activation of ERK pathway [251] and inhibition of JNK [278]. Observed activation of<br />

ERK (Fig.3.3.1.7a) but not JNK (Fig.3.3.1.7b) in gamma irradiated cells indicated the dominance<br />

of pro-survival pathways.<br />

Since, with high LET radiation yield of ROS is very low because of the radical-radical<br />

recombination, the early activation of ERK was not expected. However, complete absence of<br />

ERK activation even hours after high LET radiation suggests that repair pathways are neither<br />

feeding into nor being fed upon by intracellular cytoprotective signaling. However, the proapoptotic<br />

JNK pathway was found to be activated after high LET radiation. BRCA1 induced<br />

apoptotic responses have also been shown to be via activation of JNK [279]. Thus, after carbon<br />

irradiation, DNA damage induced activation of repair components are being fed into by<br />

cytotoxic signaling rather than cytoprotective responses and was evident in the form of early<br />

induction of apoptosis (Fig.3.3.1.8). Since, cellular decisions are summation of all the activated<br />

pathways, the final response after high LET radiation tips towards death.<br />

In summary, these results suggest that the subtle differences leading to different outcomes due to<br />

radiation quality seem to lie in different macromolecular complexes of crucial DNA damage<br />

response proteins and activation of other intracellular pathways. Analysis of functionality of<br />

these macromolecular complexes as a whole rather than individual proteins and holistic study<br />

involving intracellular survival pathways could help in revealing the mechanistic details of<br />

complex DNA damage responses.<br />

199


CHAPTER 4<br />

DISCUSSION<br />

4.3.2 Oxygen beam induced signaling in mammalian cells and its comparison with Gamma<br />

irradiation.<br />

The calculated RBE (relative biological effectiveness) for A549 cells irradiated with 1Gy oxygen<br />

ions particles relative to γ rays was found to be nearly 3 at 20% survival. Thus, oxygen ions<br />

were found to be three times more toxic than gamma rays (Fig. 3.3.2.1). These results therefore<br />

further support previous studies where high LET oxygen beam have been found to be much more<br />

efficient in killing the tumor cells and hence may be a preferred mode of treatment for<br />

radioresistant tumors. However, mechanisms of cell death caused by oxygen treatment and the<br />

signaling pathways that ensue have not yet been investigated in detail although heavy ion beam<br />

therapy is in use for many cancers [280-283]. Since DNA double strand breaks are the primary<br />

toxic lesions induced by radiation, the number of dsbs were scored by examining the foci of<br />

phosphorylated H2AX in irradiated cells. γ-H2AX foci detected in cells irradiated with 2Gy<br />

oxygen ion is contradictory to the fact that number of ion that hits on nucleus and number of γ-<br />

H2AX foci formed agrees well in heavy-ion irradiated cells, because heavy-ion densely deposits<br />

energy along their trajectories. Calculating from LET of oxygen ion, it is estimated that the<br />

number of ion hit on each cell nucleus by irradiation of 2Gy oxygen ions was approximately 3<br />

particles. However, nearly 10 times of foci were detected in the cells irradiated with 2Gy of<br />

oxygen ions. This discrepancy between observed γ-H2AX foci and estimated γ-H2AX foci after<br />

high LET radiation has also been reported previously [214]. Moreover, 3 particles of oxygen<br />

may induce more than 3 double strand break in the DNA which will ultimately decide the<br />

number of γ-H2AX foci.<br />

200


CHAPTER 4<br />

DISCUSSION<br />

Thus unlike the cell survival, the difference in induction of γ-H2AX foci 15 min after 2Gy of<br />

either radiation was not as profound (Fig. 3.3.2.2). Similar number of γ-H2AX at 15 minutes<br />

after irradiation irrespective of the radiation quality indicated that initial events might not differ<br />

much with the radiation quality. Activation of other molecules involved in DNA damage<br />

response were therefore followed only at later time point of 4 hours. Moreover, by this time low<br />

LET induced damage is mostly repaired and the activation of these factors is reduced to control<br />

levels [284] thereby sealing the fate of the cell. Comparing the activation of these molecules at<br />

this time could therefore give important insight into differences in damage response signaling<br />

with respect to radiation quality. Phosphorylation of H2AX even at 4 hours (Fig.3.3.2.2B & D)<br />

indicates the presence of large number of double strand breaks at that time indicating very slow<br />

repair in A549 cells exposed to oxygen beam.<br />

Persistence of significant pATM foci till 4 hours after oxygen ion irradiation (Fig.3.3.2.3)<br />

indicates the presence of unrepaired DNA, which is still keeping the damage sensor protein<br />

ATM in activated form. The slower disappearance of pATM foci from cells, unlike the γ-H2AX<br />

foci (that was seen in 60% of oxygen irradiated cells at 4 hours), could reflect their different<br />

roles in the repair process with γ-H2AX foci playing a primary role in activating downstream<br />

pathways and initiating DNA repair while ATM foci remaining till broken DNA ends are<br />

present. Since the cell survival after 2Gy oxygen irradiation was only 11%, ATM might be<br />

activating downstream apoptotic responses.<br />

The massive activation of ATR after oxygen irradiation when compared with gamma<br />

(Fig.3.3.2.4) in 36% of the cells indicated its specific role in high LET induced complex damage.<br />

The downstream component Chk2, which is activated by ATM and shares its substrates like p53,<br />

201


CHAPTER 4<br />

DISCUSSION<br />

was found to be significantly activated in oxygen irradiated cells. Activation of Chk2 is critical<br />

in regulating cell cycle arrest and DSB repair. Presence of pChk2 foci in nuclei of oxygen<br />

irradiated cells but not gamma indicated Chk2 was an active player in high LET radiation<br />

induced responses (Fig.3.3.2.6A, B). Both ATM and Chk2 have been regarded as primary<br />

regulators of homologous recombination repair (HRR) [103, 251]. Activation of these kinases<br />

after oxygen irradiation indicated involvement of HRR pathway for repairing complex damage<br />

caused by high LET. It is also reasonable to suppose that cells would prefer to repair a complex<br />

damage by activating the machinery of HRR, which can take the advantage of the other strand<br />

that may be intact. Contribution of HRR after high LET radiation had also been shown by<br />

previous studies [212]. Immunofluorescence of phospho Chk1, which is a major target of ATR,<br />

was also found to increase in nuclei of oxygen-irradiated cells (Fig.3.3.2.6C). Activation of Chk2<br />

and Chk1 after oxygen irradiation also indicates cell cycle arrest.<br />

Activation of p53 in oxygen irradiated cells (Fig.3.3.2.5) along with observance of apoptotic<br />

cells (Fig.3.3.2.7) indicated p53 dependent apoptosis in these cells. Despite significant activation<br />

of ATM, ATR, Chk1 and Chk2 in A549 cells exposed to 2Gy oxygen beam, DNA was not<br />

repaired as indicated by the survival data. This further point to the involvement of these kinases<br />

in activation of apoptotic machinery after high LET irradiation.<br />

From a critical analysis of all the results of the present study, it was evident that in<br />

general, high LET radiations cause more severe genomic damage as compared to low LET<br />

radiations. The noteworthy finding of this study is the activation of the sensor proteins, ATM and<br />

ATR by oxygen irradiation and the significant activation of Chk1, Chk2 and p53 only in the<br />

oxygen beam irradiated cells. Since the signaling pathways activated by oxygen beam are<br />

202


CHAPTER 4<br />

DISCUSSION<br />

different from gamma irradiation, signaling component and intensity of activation will be<br />

different by both types of irradiation. These kinases play an important role in repair, cell cycle<br />

arrest and apoptosis. However, only 11% survival after high LET radiation suggests that most of<br />

the damage could not be repaired and that the persistent activation of these kinases may be a<br />

signature for activation of apoptotic responses. Further studies in this direction will reveal the<br />

importance of such response.<br />

4.4 Radiation induced bystander signaling in mammalian cells.<br />

The signaling events seen in totality in the foregoing work may have contribution from the<br />

surrounding microenvironment that is of necessity exposed to radiation. The cells nearby which<br />

are not directly exposed to the radiation but are juxtaposed to the irradiated cells are also affected<br />

because of signals transduced from the ‘hit’ cell. The classical paradigm of radiation biology<br />

asserts that all radiation effects on cells, tissues and organisms are due to the direct action of<br />

radiation. However, there has been a recent growth of interest in the indirect actions of radiation<br />

including the radiation-induced adaptive response, the bystander effect, low-dose<br />

hypersensitivity, and genomic instability, which are specific modes of stress exhibited in<br />

response to low-dose/low-dose rate radiation.<br />

Radiation-induced bystander signals (whether observed through use of medium harvested from<br />

irradiated cells, or through observation of non-hit cells in contact with irradiated cells), appear to<br />

coordinate tissue or population responses so that cells not directly exposed or traversed by<br />

radiation show responses, which may be part of higher order homeostatic control. Bystander<br />

effects appear to be the result of a generalized stress response in tissues or cells. The signals may<br />

be produced by all exposed cells but the response may require a quorum in order to be expressed.<br />

203


CHAPTER 4<br />

DISCUSSION<br />

The major response involving low LET radiation exposure discussed in the existing literature is a<br />

death response, which has many characteristics of apoptosis but may be detected in cell lines<br />

without p53 expression [285].<br />

Two main models have emerged concerning the mechanisms of bystander mediated responses,<br />

mainly based on the way the experiments were performed and the cell type selected for<br />

investigation. They are communication via gap junction intercellular communication (GJIC) and<br />

communication via media-borne or transmissible factors [286-289] (Fig. 1.6).<br />

The present study was carried out to investigate what genes and proteins are activated in the<br />

bystanding cell and whether the state of activation of the irradiated cell alters the bystander<br />

response. The results revealed that there was significant decrease in the survival of bystander<br />

A549 cells and significant increase in γ-H2AX and pATM foci but only in those cells that had<br />

received medium from irradiated cells (Fig.3.4.1). Since the cells treated only with irradiated<br />

medium did not show any change in survival or γ-H2AX and pATM foci the signals which<br />

caused the damage in the bystander cells must be coming from the irradiated cells themselves.<br />

Having confirmed the existence of bystander effect between similar cells (A549 Vs<br />

A549), the work was extended to look for the existence of cross bystander effect between A549<br />

and U937 cells i.e. lung carcinoma and monocytes. There was a significant decrease in the<br />

survival of A549 cells that had received medium from PMA stimulated and irradiated U937 cells<br />

(Fig.3.4.2) and expectedly the γ-H2AX and pATM foci were also significantly higher in these<br />

cells (Fig.3.4.2). A549 cells that had received medium from either irradiated or PMA stimulated<br />

U937 cells did not show bystander response, indicating the fact that activation of U937 was prerequisite<br />

for the induction of cross bystander response. The PMA treatment of monocytes<br />

204


CHAPTER 4<br />

DISCUSSION<br />

upregulates many genes in the monocyte hence their response to irradiation will be different<br />

from unstimulated monocyte.<br />

Having confirmed that bystander effect contributed the survival of the neighbouring cells, the<br />

lung carcinoma A549 cells, the mechanism of bystander effect and the factor involved in the<br />

bystander effect was studied in mouse cell line.<br />

The present results revealed a significant increase in DNA damage in the bystander cells but only<br />

in those cells that had received medium from irradiated cells (Fig.3.4.3.3). Since the cells treated<br />

only with irradiated medium did not show any DNA damage or NO production, the signals<br />

which caused the damage in the bystander cells must be coming from the irradiated cells<br />

themselves. The expression of cell cycle related genes (p53 and p21) was significantly<br />

upregulated in bystander similar cells (Fig.3.4.3.1C and D, Lane 3). These results are in<br />

agreement with the work on bystander signaling [290, 291]. The first report describing the<br />

induction of sister chromatid exchanges (SCE) was examined in Chinese hamster ovary (CHO)<br />

cells. Since CHO cells contain mutant p53, therefore, p53 could not be involved in the bystander<br />

induction of SCE [220]. If the p53 of the bystander cell is mutated there should be less DNA<br />

repair and in fact more bystander effect can be seen when assessed in terms of DNA damage.<br />

A clear cause and effect relationship was established between DNA damage, gene expression<br />

and cell survival and was evident in the form of increase apoptosis in bystander cells (Fig.<br />

3.4.3.4a, Lane 3 and Fig. 3.4.3.4b).<br />

An increased expression of NF-κB gene (p65) could lead to the activation of iNOS gene since<br />

the promoter of the iNOS gene contains binding sites for NF-κB. NF-κB is known to be<br />

205


CHAPTER 4<br />

DISCUSSION<br />

activated by several oxidative stresses including ionization radiation [292-297]. As a<br />

consequence of iNOS gene upregulation and increased production of NO, many forms of<br />

reactive nitrogen species (RNS) would be formed leading to covalent modification of proteins<br />

and the expression of various other genes [298].<br />

Having confirmed the existence of bystander effect between similar cells (EL-4 Vs EL-4), the<br />

work was extended to look for the existence of cross bystander effect between EL-4 and RAW<br />

264.7 cells i.e. lymphocytes and macrophages. There was a significant upregulation of NF-κB<br />

and iNOS genes in EL-4 cells that had received medium from LPS stimulated and irradiated<br />

RAW 264.7 cells (Fig. 3.4.4a, Lane, 2) and expectedly the DNA damage was also significantly<br />

higher in these cells (Fig. 3.4.4b, Lane, 2). A noteworthy finding of this study is the fact that the<br />

unirradiated EL-4 cells that had received medium from LPS stimulated RAW 264.7 cells also<br />

showed upregulation of iNOS gene but not NF-κB gene (Fig. 3.4.4a, Lane, 4). The LPS<br />

treatment of macrophage upregulates many genes in the macrophage hence their response to<br />

irradiation will be different from unstimulated macrophage. Another noteworthy finding of this<br />

study is the fact that the addition of medium from non-irradiated RAW 264.7 cells resulted in<br />

significantly decrease in iNOS gene expression in EL-4 cells relative to the control (Fig. 3.4.4a,<br />

Lane, 5). Non-irradiated macrophages may release some factor which is responsible for down<br />

regulation of iNOS in EL-4 cells. Further studies in this direction will reveal the nature of the<br />

factor. It is clear from the above results that effective cross bystander effects exists between these<br />

two cell lines only when the RAW 264.7 cells are both stimulated and irradiated. Other studies<br />

have looked at only the clonogenic survival of dissimilar bystander cells [216].<br />

206


CHAPTER 4<br />

DISCUSSION<br />

Numerous studies have implicated extracellular secreted signaling proteins in mediating the<br />

bystander effect [289], but the strongest contenders have been reactive oxygen and nitrogen<br />

species [299]. To investigate what factors are involved in radiation induced bystander effect, the<br />

cells were treated with either L-NAME (NOS inhibitor) or cPTIO (NO scavenger). The effect of<br />

cPTIO and L-NAME in bystander response between similar cell (EL-4 Vs EL-4) was different in<br />

the fact that treatment with L-NAME reduces the bystander induction of NF-κB as well as iNOS<br />

gene expression in EL-4 cells receiving medium from irradiated EL-4 cells (Fig. 3.4.5a, Lane 4)<br />

but the treatment with cPTIO reduces the bystander induction of only NF-κB gene expression<br />

and not iNOS gene expression (Fig. 3.4.5a, Lane 5). The above results demonstrate that the<br />

activated NOS in the irradiated cell is essential for gene expression in the bystanding cell.<br />

The addition of either L-NAME or cPTIO reduces the bystander induction of NF-κB as well as<br />

iNOS gene expression in EL-4 cells receiving medium from LPS stimulated and irradiated RAW<br />

264.7 cells (Fig. 3.4.5a, Lane 6 and 7), indicating that both activation of iNOS in the irradiated<br />

cell and NO production play a role in cross bystander effect.<br />

The treatment with L-NAME or cPTIO also reduces the bystander induction of DNA damage in<br />

EL-4 cells receiving medium from irradiated EL-4 cells or LPS stimulated and irradiated RAW<br />

264.7 cells (Fig. 3.4.5b, Lane 4, 5, 6 and 7), demonstrating that NO contributes to the DNA<br />

damage in bystander EL-4 cells. NO is generated endogeneously from L-arginine by inducible<br />

NO synthases [300, 301] that can be stimulated by radiation in mammalian cells [302]. It is<br />

believed that NO itself does not induce DNA strand breaks, although one of its oxidation<br />

product, peroxynitrite is toxic and can cause DNA damage by both attacking deoxyribose and<br />

direct oxidation of purine and pyrimidine [303]. Radiation induced NO has the possibility of<br />

207


CHAPTER 4<br />

DISCUSSION<br />

attacking nearby cells within their diffusion distance of less than 0.5mm [304, 305]. However,<br />

because of their having very short half-lives, they may not be direct contributors to the cellular<br />

damage in the bystander population. This is confirmed by the fact that inclusion of cPTIO in the<br />

medium only marginally inhibits the bystander response in similar cells (Fig. 3.4.5a, Lane 5).<br />

Contrarily, it has also been reported that NO is involved in the bystander responses caused by the<br />

conditioned medium harvested from irradiated cells [306]. Therefore, some long-lived bioactive<br />

factors downstream of NO are most likely involved in this bystander response.<br />

4.5 Summary and Conclusion<br />

Summary:<br />

The irradiation of cells with fractionated doses led to a signaling response that was<br />

different from a single dose of irradiation. A549 cells were found to be relatively more<br />

radioresistant if the 10Gy dose was delivered as a fractionated regimen. Microarray<br />

analysis showed upregulation of DNA repair and cell cycle arrest genes in the cells<br />

exposed to fractionated irradiation. There was intense activation of DNA repair pathway<br />

(DNA-PK, ATM, Rad52, MLH1 and BRCA1), efficient DNA repair and phospho-p53<br />

was found to be translocated to the nucleus of A549 cells exposed to fractionated<br />

irradiation. MCF-7 cells responded differently in fractionated regimen. Silencing of the<br />

Rad52 gene in fractionated group of A549 cells made the cells radiosensitive. The<br />

reasons for the radioresistance of the A549 cells lay in the repair pathway, the Rad52, the<br />

208


CHAPTER 4<br />

DISCUSSION<br />

inhibition of which could revert the cells to radiosensitivity leading to a decrease in<br />

survival.<br />

Proton beam was found to be more cytotoxic than γ-radiation. Proton beam irradiated<br />

cells showed phosphorylation of H2AX, ATM, Chk2 and p53. The mechanism of<br />

excessive cell killing in proton beam irradiated cells was found to be upregulation of Bax<br />

and down-regulation of Bcl-2. The noteworthy finding of this study is the biphasic<br />

activation of the sensor proteins, ATM and DNA-PK and no activation of ATR by proton<br />

irradiation.<br />

Carbon beam was found to be three times more cytotoxic than γ-radiation despite the fact<br />

that the numbers of γ-H2AX foci were same. Percentage of cells showing ATM/ATR foci<br />

were more with gamma however number of foci per cell were more in case of carbon<br />

irradiation. Large BRCA1 foci were found in all carbon irradiated cells unlike gamma<br />

irradiated cells and prosurvival ERK pathway was activated after gamma irradiation but<br />

not carbon. The noteworthy finding of this study is the early phase apoptosis induction by<br />

carbon ions. Despite activation of same repair molecules, differences in low and high<br />

LET damage responses are due to distinct macromolecular complexes of repair proteins<br />

such as ATM, BRCA1 etc rather than their individual activation and the activation of<br />

cytoplasmic pathways such as ERK.<br />

Oxygen beam was found to be three times more cytotoxic than γ-radiation. By 4h there<br />

was efficient repair of DNA in A549 cells exposed to 2Gy or 6Gy gamma radiation but<br />

not in cells exposed to 2Gy oxygen beam as determined by γ-H2AX counting. Number of<br />

ATM foci was found to be significantly higher in cells exposed to 2Gy oxygen beam.<br />

209


CHAPTER 4<br />

DISCUSSION<br />

Percentage of cells showing ATR foci were more with gamma however number of foci<br />

per cell were more in case of oxygen beam. Oxygen beam irradiated cells showed<br />

phosphorylation of Chk1, Chk2 and p53. Many apoptotic nuclei were seen by DAPI<br />

staining in cells exposed to oxygen beam. The noteworthy finding of this study is the<br />

activation of the sensor proteins, ATM and ATR by oxygen irradiation and the significant<br />

activation of Chk1, Chk2 and p53 only in the oxygen beam irradiated cells.<br />

The current study is a clear evidence of radiation induced bystander effect being involved<br />

in adaptive responses in the bystander cells. Moreover, it also points towards the potential<br />

involvement of signaling molecules released into the medium by the irradiated cells<br />

which elicit a response in the bystander cells. Also there may be potential involvement of<br />

stable free radicals that are produced in the medium as a result of irradiation, in eliciting a<br />

response from the bystander cells. The mechanism of which may be via NO generation.<br />

Conclusion:<br />

Radiation induced signal transduction, as understood presently, is an activation of the<br />

existing network, where both positive and negative signals converge and these are interpreted as<br />

cell survival and cell death. In spite of the fact that mere phosphorylation by kinases seem to be<br />

the basic theme, it is crucial as based on this life or death decisions are taken. Hence enormous<br />

cross checking of the ultimate signal must exist. The lung carcinoma cell line A549, a highly<br />

radioresistant cell line could be effectively made radiosensitive by inhibiting Rad52, one of the<br />

components of its repair pathway. The survival of the cells decreased significantly after doing so.<br />

210


CHAPTER 4<br />

DISCUSSION<br />

The survival of the A549 cell line could also be significantly decreased by heavy ion<br />

irradiation. The mechanism of which seems to be a lack of repair, despite the activation of some<br />

of the component of the repair pathway. Besides the microenvironment of the cells and the LET<br />

of the radiation, even transfer of medium from irradiated cells to non-irradiated cells can alter the<br />

pattern of signaling in a cell.<br />

211


CHAPTER 5<br />

REFERENCES<br />

REFERENCES<br />

212


CHAPTER 5<br />

REFERENCES<br />

[1] Khan, F. The physics of radiation therapy. Baltimore, MD: Williams and Wilkins;<br />

1994.<br />

[2] Alpen, E. L. Radiation Biophysics. Prentice Hall, New Jersey; 1990.<br />

[3] Little, J. B. Cellular effects of ionizing radiation. N Engl J Med 278:369-376<br />

concl; 1968.<br />

[4] Prise, K. M.; Schettino, G.; Folkard, M.; Held, K. D. New insights on cell death<br />

from radiation exposure. Lancet Oncol 6:520-528; 2005.<br />

[5] Szostak, M. J.; Kyprianou, N. Radiation-induced apoptosis: predictive and<br />

therapeutic significance in radiotherapy of prostate cancer (review). Oncol Rep 7:699-<br />

706; 2000.<br />

[6] Aviles, A.; Fernandez, R.; Calva, A.; Neri, N.; Huerta-Guzman, J.; Nambo, M. J.<br />

Radiotherapy versus combined therapy in early stages with bulky disease aggressive<br />

malignant lymphoma. Hematology 8:7-10; 2003.<br />

[7] Zhang, M.; Zhang, Z.; Garmestani, K.; Schultz, J.; Axworthy, D. B.; Goldman, C.<br />

K.; Brechbiel, M. W.; Carrasquillo, J. A.; Waldmann, T. A. Pretarget radiotherapy with<br />

an anti-CD25 antibody-streptavidin fusion protein was effective in therapy of<br />

leukemia/lymphoma xenografts. Proc Natl Acad Sci U S A 100:1891-1895; 2003.<br />

[8] Thames, H. D.; Bentzen, S. M.; Turesson, I.; Overgaard, M.; Van den Bogaert,<br />

W. Time-dose factors in radiotherapy: a review of the human data. Radiother Oncol<br />

19:219-235; 1990.<br />

[9] Caspari, T. How to activate p53. Curr Biol 10:R315-317; 2000.<br />

213


CHAPTER 5<br />

REFERENCES<br />

[10] Mailand, N.; Falck, J.; Lukas, C.; Syljuasen, R. G.; Welcker, M.; Bartek, J.;<br />

Lukas, J. Rapid destruction of human Cdc25A in response to DNA damage. Science<br />

288:1425-1429; 2000.<br />

[11] Lakin, N. D.; Jackson, S. P. Regulation of p53 in response to DNA damage.<br />

Oncogene 18:7644-7655; 1999.<br />

[12] Lowndes, N. F.; Murguia, J. R. Sensing and responding to DNA damage. Curr<br />

Opin Genet Dev 10:17-25; 2000.<br />

[13] Rich, T.; Allen, R. L.; Wyllie, A. H. Defying death after DNA damage. Nature<br />

407:777-783; 2000.<br />

[14] Hartley, K. O.; Gell, D.; Smith, G. C.; Zhang, H.; Divecha, N.; Connelly, M. A.;<br />

Admon, A.; Lees-Miller, S. P.; Anderson, C. W.; Jackson, S. P. DNA-dependent protein<br />

kinase catalytic subunit: a relative of phosphatidylinositol 3-kinase and the ataxia<br />

telangiectasia gene product. Cell 82:849-856; 1995.<br />

[15] Smith, G. C.; Jackson, S. P. The DNA-dependent protein kinase. Genes Dev<br />

13:916-934; 1999.<br />

[16] Dvir, A.; Stein, L. Y.; Calore, B. L.; Dynan, W. S. Purification and<br />

characterization of a template-associated protein kinase that phosphorylates RNA<br />

polymerase II. J Biol Chem 268:10440-10447; 1993.<br />

[17] Gottlieb, T. M.; Jackson, S. P. The DNA-dependent protein kinase: requirement<br />

for DNA ends and association with Ku antigen. Cell 72:131-142; 1993.<br />

[18] Hammarsten, O.; Chu, G. DNA-dependent protein kinase: DNA binding and<br />

activation in the absence of Ku. Proc Natl Acad Sci U S A 95:525-530; 1998.<br />

214


CHAPTER 5<br />

REFERENCES<br />

[19] Woo, R. A.; McLure, K. G.; Lees-Miller, S. P.; Rancourt, D. E.; Lee, P. W. DNAdependent<br />

protein kinase acts upstream of p53 in response to DNA damage. Nature<br />

394:700-704; 1998.<br />

[20] Shieh, S. Y.; Ikeda, M.; Taya, Y.; Prives, C. DNA damage-induced<br />

phosphorylation of p53 alleviates inhibition by MDM2. Cell 91:325-334; 1997.<br />

[21] Rathmell, W. K.; Kaufmann, W. K.; Hurt, J. C.; Byrd, L. L.; Chu, G. DNAdependent<br />

protein kinase is not required for accumulation of p53 or cell cycle arrest after<br />

DNA damage. Cancer Res 57:68-74; 1997.<br />

[22] Burma, S.; Kurimasa, A.; Xie, G.; Taya, Y.; Araki, R.; Abe, M.; Crissman, H. A.;<br />

Ouyang, H.; Li, G. C.; Chen, D. J. DNA-dependent protein kinase-independent activation<br />

of p53 in response to DNA damage. J Biol Chem 274:17139-17143; 1999.<br />

[23] Jhappan, C.; Yusufzai, T. M.; Anderson, S.; Anver, M. R.; Merlino, G. The p53<br />

response to DNA damage in vivo is independent of DNA-dependent protein kinase. Mol<br />

Cell Biol 20:4075-4083; 2000.<br />

[24] Wang, S.; Guo, M.; Ouyang, H.; Li, X.; Cordon-Cardo, C.; Kurimasa, A.; Chen,<br />

D. J.; Fuks, Z.; Ling, C. C.; Li, G. C. The catalytic subunit of DNA-dependent protein<br />

kinase selectively regulates p53-dependent apoptosis but not cell-cycle arrest. Proc Natl<br />

Acad Sci U S A 97:1584-1588; 2000.<br />

[25] Kurimasa, A.; Kumano, S.; Boubnov, N. V.; Story, M. D.; Tung, C. S.; Peterson,<br />

S. R.; Chen, D. J. Requirement for the kinase activity of human DNA-dependent protein<br />

kinase catalytic subunit in DNA strand break rejoining. Mol Cell Biol 19:3877-3884;<br />

1999.<br />

215


CHAPTER 5<br />

REFERENCES<br />

[26] Zou, L.; Elledge, S. J. Sensing DNA damage through ATRIP recognition of RPAssDNA<br />

complexes. Science 300:1542-1548; 2003.<br />

[27] You, Z.; Kong, L.; Newport, J. The role of single-stranded DNA and polymerase<br />

alpha in establishing the ATR, Hus1 DNA replication checkpoint. J Biol Chem<br />

277:27088-27093; 2002.<br />

[28] Abraham, R. T. Cell cycle checkpoint signaling through the ATM and ATR<br />

kinases. Genes Dev 15:2177-2196; 2001.<br />

[29] Shiloh, Y. ATM and related protein kinases: safeguarding genome integrity. Nat<br />

Rev Cancer 3:155-168; 2003.<br />

[30] Brown, E. J.; Baltimore, D. Essential and dispensable roles of ATR in cell cycle<br />

arrest and genome maintenance. Genes Dev 17:615-628; 2003.<br />

[31] Brown, E. J.; Baltimore, D. ATR disruption leads to chromosomal fragmentation<br />

and early embryonic lethality. Genes Dev 14:397-402; 2000.<br />

[32] Hekmat-Nejad, M.; You, Z.; Yee, M. C.; Newport, J. W.; Cimprich, K. A.<br />

Xenopus ATR is a replication-dependent chromatin-binding protein required for the<br />

DNA replication checkpoint. Curr Biol 10:1565-1573; 2000.<br />

[33] Dart, D. A.; Adams, K. E.; Akerman, I.; Lakin, N. D. Recruitment of the cell<br />

cycle checkpoint kinase ATR to chromatin during S-phase. J Biol Chem 279:16433-<br />

16440; 2004.<br />

[34] Shechter, D.; Costanzo, V.; Gautier, J. ATR and ATM regulate the timing of<br />

DNA replication origin firing. Nat Cell Biol 6:648-655; 2004.<br />

216


CHAPTER 5<br />

REFERENCES<br />

[35] Nghiem, P.; Park, P. K.; Kim, Y.; Vaziri, C.; Schreiber, S. L. ATR inhibition<br />

selectively sensitizes G1 checkpoint-deficient cells to lethal premature chromatin<br />

condensation. Proc Natl Acad Sci U S A 98:9092-9097; 2001.<br />

[36] Cha, R. S.; Kleckner, N. ATR homolog Mec1 promotes fork progression, thus<br />

averting breaks in replication slow zones. Science 297:602-606; 2002.<br />

[37] O'Driscoll, M.; Ruiz-Perez, V. L.; Woods, C. G.; Jeggo, P. A.; Goodship, J. A. A<br />

splicing mutation affecting expression of ataxia-telangiectasia and Rad3-related protein<br />

(ATR) results in Seckel syndrome. Nat Genet 33:497-501; 2003.<br />

[38] Tibbetts, R. S.; Cortez, D.; Brumbaugh, K. M.; Scully, R.; Livingston, D.;<br />

Elledge, S. J.; Abraham, R. T. Functional interactions between BRCA1 and the<br />

checkpoint kinase ATR during genotoxic stress. Genes Dev 14:2989-3002; 2000.<br />

[39] Dodson, G. E.; Shi, Y.; Tibbetts, R. S. DNA replication defects, spontaneous<br />

DNA damage, and ATM-dependent checkpoint activation in replication protein A-<br />

deficient cells. J Biol Chem 279:34010-34014; 2004.<br />

[40] Bomgarden, R. D.; Yean, D.; Yee, M. C.; Cimprich, K. A. A novel protein<br />

activity mediates DNA binding of an ATR-ATRIP complex. J Biol Chem 279:13346-<br />

13353; 2004.<br />

[41] Stiff, T.; Reis, C.; Alderton, G. K.; Woodbine, L.; O'Driscoll, M.; Jeggo, P. A.<br />

Nbs1 is required for ATR-dependent phosphorylation events. Embo J 24:199-208; 2005.<br />

[42] Busino, L.; Donzelli, M.; Chiesa, M.; Guardavaccaro, D.; Ganoth, D.; Dorrello,<br />

N. V.; Hershko, A.; Pagano, M.; Draetta, G. F. Degradation of Cdc25A by beta-TrCP<br />

during S phase and in response to DNA damage. Nature 426:87-91; 2003.<br />

217


CHAPTER 5<br />

REFERENCES<br />

[43] Jin, J.; Shirogane, T.; Xu, L.; Nalepa, G.; Qin, J.; Elledge, S. J.; Harper, J. W.<br />

SCFbeta-TRCP links Chk1 signaling to degradation of the Cdc25A protein phosphatase.<br />

Genes Dev 17:3062-3074; 2003.<br />

[44] Heffernan, T. P.; Simpson, D. A.; Frank, A. R.; Heinloth, A. N.; Paules, R. S.;<br />

Cordeiro-Stone, M.; Kaufmann, W. K. An ATR- and Chk1-dependent S checkpoint<br />

inhibits replicon initiation following UVC-induced DNA damage. Mol Cell Biol<br />

22:8552-8561; 2002.<br />

[45] Bartek, J.; Lukas, C.; Lukas, J. Checking on DNA damage in S phase. Nat Rev<br />

Mol Cell Biol 5:792-804; 2004.<br />

[46] Andreassen, P. R.; D'Andrea, A. D.; Taniguchi, T. ATR couples FANCD2<br />

monoubiquitination to the DNA-damage response. Genes Dev 18:1958-1963; 2004.<br />

[47] Wang, X.; D'Andrea, A. D. The interplay of Fanconi anemia proteins in the DNA<br />

damage response. DNA Repair (Amst) 3:1063-1069; 2004.<br />

[48] Venkitaraman, A. R. Tracing the network connecting BRCA and Fanconi<br />

anaemia proteins. Nat Rev Cancer 4:266-276; 2004.<br />

[49] Pichierri, P.; Rosselli, F. The DNA crosslink-induced S-phase checkpoint depends<br />

on ATR-CHK1 and ATR-NBS1-FANCD2 pathways. Embo J 23:1178-1187; 2004.<br />

[50] Tibbetts, R. S.; Brumbaugh, K. M.; Williams, J. M.; Sarkaria, J. N.; Cliby, W. A.;<br />

Shieh, S. Y.; Taya, Y.; Prives, C.; Abraham, R. T. A role for ATR in the DNA damageinduced<br />

phosphorylation of p53. Genes Dev 13:152-157; 1999.<br />

[51] Lakin, N. D.; Hann, B. C.; Jackson, S. P. The ataxia-telangiectasia related protein<br />

ATR mediates DNA-dependent phosphorylation of p53. Oncogene 18:3989-3995; 1999.<br />

218


CHAPTER 5<br />

REFERENCES<br />

[52] McKay, B. C.; Ljungman, M.; Rainbow, A. J. Potential roles for p53 in nucleotide<br />

excision repair. Carcinogenesis 20:1389-1396; 1999.<br />

[53] Sengupta, S.; Harris, C. C. p53: traffic cop at the crossroads of DNA repair and<br />

recombination. Nat Rev Mol Cell Biol 6:44-55; 2005.<br />

[54] Sorensen, C. S.; Hansen, L. T.; Dziegielewski, J.; Syljuasen, R. G.; Lundin, C.;<br />

Bartek, J.; Helleday, T. The cell-cycle checkpoint kinase Chk1 is required for mammalian<br />

homologous recombination repair. Nat Cell Biol 7:195-201; 2005.<br />

[55] Shinozaki, T.; Nota, A.; Taya, Y.; Okamoto, K. Functional role of Mdm2<br />

phosphorylation by ATR in attenuation of p53 nuclear export. Oncogene 22:8870-8880;<br />

2003.<br />

[56] Shiloh, Y. ATM: ready, set, go. Cell Cycle 2:116-117; 2003.<br />

[57] Lavin, M. F.; Shiloh, Y. The genetic defect in ataxia-telangiectasia. Annu Rev<br />

Immunol 15:177-202; 1997.<br />

[58] Savitsky, K.; Sfez, S.; Tagle, D. A.; Ziv, Y.; Sartiel, A.; Collins, F. S.; Shiloh, Y.;<br />

Rotman, G. The complete sequence of the coding region of the ATM gene reveals<br />

similarity to cell cycle regulators in different species. Hum Mol Genet 4:2025-2032;<br />

1995.<br />

[59] Savitsky, K.; Bar-Shira, A.; Gilad, S.; Rotman, G.; Ziv, Y.; Vanagaite, L.; Tagle,<br />

D. A.; Smith, S.; Uziel, T.; Sfez, S.; Ashkenazi, M.; Pecker, I.; Frydman, M.; Harnik, R.;<br />

Patanjali, S. R.; Simmons, A.; Clines, G. A.; Sartiel, A.; Gatti, R. A.; Chessa, L.; Sanal,<br />

O.; Lavin, M. F.; Jaspers, N. G.; Taylor, A. M.; Arlett, C. F.; Miki, T.; Weissman, S. M.;<br />

Lovett, M.; Collins, F. S.; Shiloh, Y. A single ataxia telangiectasia gene with a product<br />

similar to PI-3 kinase. Science 268:1749-1753; 1995.<br />

219


CHAPTER 5<br />

REFERENCES<br />

[60] Goodarzi, A. A.; Block, W. D.; Lees-Miller, S. P. The role of ATM and ATR in<br />

DNA damage-induced cell cycle control. Prog Cell Cycle Res 5:393-411; 2003.<br />

[61] Manning, G.; Whyte, D. B.; Martinez, R.; Hunter, T.; Sudarsanam, S. The protein<br />

kinase complement of the human genome. Science 298:1912-1934; 2002.<br />

[62] Kim, S. T.; Lim, D. S.; Canman, C. E.; Kastan, M. B. Substrate specificities and<br />

identification of putative substrates of ATM kinase family members. J Biol Chem<br />

274:37538-37543; 1999.<br />

[63] O'Neill, T.; Dwyer, A. J.; Ziv, Y.; Chan, D. W.; Lees-Miller, S. P.; Abraham, R.<br />

H.; Lai, J. H.; Hill, D.; Shiloh, Y.; Cantley, L. C.; Rathbun, G. A. Utilization of oriented<br />

peptide libraries to identify substrate motifs selected by ATM. J Biol Chem 275:22719-<br />

22727; 2000.<br />

[64] Bosotti, R.; Isacchi, A.; Sonnhammer, E. L. FAT: a novel domain in PIK-related<br />

kinases. Trends Biochem Sci 25:225-227; 2000.<br />

[65] Bakkenist, C. J.; Kastan, M. B. DNA damage activates ATM through<br />

intermolecular autophosphorylation and dimer dissociation. Nature 421:499-506; 2003.<br />

[66] Perry, J.; Kleckner, N. The ATRs, ATMs, and TORs are giant HEAT repeat<br />

proteins. Cell 112:151-155; 2003.<br />

[67] Llorca, O.; Rivera-Calzada, A.; Grantham, J.; Willison, K. R. Electron<br />

microscopy and 3D reconstructions reveal that human ATM kinase uses an arm-like<br />

domain to clamp around double-stranded DNA. Oncogene 22:3867-3874; 2003.<br />

[68] Leuther, K. K.; Hammarsten, O.; Kornberg, R. D.; Chu, G. Structure of DNAdependent<br />

protein kinase: implications for its regulation by DNA. Embo J 18:1114-1123;<br />

1999.<br />

220


CHAPTER 5<br />

REFERENCES<br />

[69] Cary, R. B.; Peterson, S. R.; Wang, J.; Bear, D. G.; Bradbury, E. M.; Chen, D. J.<br />

DNA looping by Ku and the DNA-dependent protein kinase. Proc Natl Acad Sci U S A<br />

94:4267-4272; 1997.<br />

[70] Smith, G. C.; Cary, R. B.; Lakin, N. D.; Hann, B. C.; Teo, S. H.; Chen, D. J.;<br />

Jackson, S. P. Purification and DNA binding properties of the ataxia-telangiectasia gene<br />

product ATM. Proc Natl Acad Sci U S A 96:11134-11139; 1999.<br />

[71] Merkle, D.; Douglas, P.; Moorhead, G. B.; Leonenko, Z.; Yu, Y.; Cramb, D.;<br />

Bazett-Jones, D. P.; Lees-Miller, S. P. The DNA-dependent protein kinase interacts with<br />

DNA to form a protein-DNA complex that is disrupted by phosphorylation. Biochemistry<br />

41:12706-12714; 2002.<br />

[72] Yaneva, M.; Kowalewski, T.; Lieber, M. R. Interaction of DNA-dependent<br />

protein kinase with DNA and with Ku: biochemical and atomic-force microscopy studies.<br />

Embo J 16:5098-5112; 1997.<br />

[73] Banin, S.; Moyal, L.; Shieh, S.; Taya, Y.; Anderson, C. W.; Chessa, L.;<br />

Smorodinsky, N. I.; Prives, C.; Reiss, Y.; Shiloh, Y.; Ziv, Y. Enhanced phosphorylation<br />

of p53 by ATM in response to DNA damage. Science 281:1674-1677; 1998.<br />

[74] Lakin, N. D.; Weber, P.; Stankovic, T.; Rottinghaus, S. T.; Taylor, A. M.;<br />

Jackson, S. P. Analysis of the ATM protein in wild-type and ataxia telangiectasia cells.<br />

Oncogene 13:2707-2716; 1996.<br />

[75] Canman, C. E.; Lim, D. S.; Cimprich, K. A.; Taya, Y.; Tamai, K.; Sakaguchi, K.;<br />

Appella, E.; Kastan, M. B.; Siliciano, J. D. Activation of the ATM kinase by ionizing<br />

radiation and phosphorylation of p53. Science 281:1677-1679; 1998.<br />

221


CHAPTER 5<br />

REFERENCES<br />

[76] Chan, D. W.; Gately, D. P.; Urban, S.; Galloway, A. M.; Lees-Miller, S. P.; Yen,<br />

T.; Allalunis-Turner, J. Lack of correlation between ATM protein expression and tumour<br />

cell radiosensitivity. Int J Radiat Biol 74:217-224; 1998.<br />

[77] Kozlov, S.; Gueven, N.; Keating, K.; Ramsay, J.; Lavin, M. F. ATP activates<br />

ataxia-telangiectasia mutated (ATM) in vitro. Importance of autophosphorylation. J Biol<br />

Chem 278:9309-9317; 2003.<br />

[78] Chan, D. W.; Son, S. C.; Block, W.; Ye, R.; Khanna, K. K.; Wold, M. S.;<br />

Douglas, P.; Goodarzi, A. A.; Pelley, J.; Taya, Y.; Lavin, M. F.; Lees-Miller, S. P.<br />

Purification and characterization of ATM from human placenta. A manganese-dependent,<br />

wortmannin-sensitive serine/threonine protein kinase. J Biol Chem 275:7803-7810; 2000.<br />

[79] Suzuki, K.; Kodama, S.; Watanabe, M. Recruitment of ATM protein to double<br />

strand DNA irradiated with ionizing radiation. J Biol Chem 274:25571-25575; 1999.<br />

[80] Andegeko, Y.; Moyal, L.; Mittelman, L.; Tsarfaty, I.; Shiloh, Y.; Rotman, G.<br />

Nuclear retention of ATM at sites of DNA double strand breaks. J Biol Chem 276:38224-<br />

38230; 2001.<br />

[81] Petrini, J. H.; Stracker, T. H. The cellular response to DNA double-strand breaks:<br />

defining the sensors and mediators. Trends Cell Biol 13:458-462; 2003.<br />

[82] van den Bosch, M.; Bree, R. T.; Lowndes, N. F. The MRN complex: coordinating<br />

and mediating the response to broken chromosomes. EMBO Rep 4:844-849; 2003.<br />

[83] Stewart, G. S.; Maser, R. S.; Stankovic, T.; Bressan, D. A.; Kaplan, M. I.; Jaspers,<br />

N. G.; Raams, A.; Byrd, P. J.; Petrini, J. H.; Taylor, A. M. The DNA double-strand break<br />

repair gene hMRE11 is mutated in individuals with an ataxia-telangiectasia-like disorder.<br />

Cell 99:577-587; 1999.<br />

222


CHAPTER 5<br />

REFERENCES<br />

[84] Uziel, T.; Lerenthal, Y.; Moyal, L.; Andegeko, Y.; Mittelman, L.; Shiloh, Y.<br />

Requirement of the MRN complex for ATM activation by DNA damage. Embo J<br />

22:5612-5621; 2003.<br />

[85] Carson, C. T.; Schwartz, R. A.; Stracker, T. H.; Lilley, C. E.; Lee, D. V.;<br />

Weitzman, M. D. The Mre11 complex is required for ATM activation and the G2/M<br />

checkpoint. Embo J 22:6610-6620; 2003.<br />

[86] Lukas, C.; Falck, J.; Bartkova, J.; Bartek, J.; Lukas, J. Distinct spatiotemporal<br />

dynamics of mammalian checkpoint regulators induced by DNA damage. Nat Cell Biol<br />

5:255-260; 2003.<br />

[87] Rogakou, E. P.; Boon, C.; Redon, C.; Bonner, W. M. Megabase chromatin<br />

domains involved in DNA double-strand breaks in vivo. J Cell Biol 146:905-916; 1999.<br />

[88] Goldberg, M.; Stucki, M.; Falck, J.; D'Amours, D.; Rahman, D.; Pappin, D.;<br />

Bartek, J.; Jackson, S. P. MDC1 is required for the intra-S-phase DNA damage<br />

checkpoint. Nature 421:952-956; 2003.<br />

[89] Mochan, T. A.; Venere, M.; DiTullio, R. A., Jr.; Halazonetis, T. D. 53BP1 and<br />

NFBD1/MDC1-Nbs1 function in parallel interacting pathways activating ataxiatelangiectasia<br />

mutated (ATM) in response to DNA damage. Cancer Res 63:8586-8591;<br />

2003.<br />

[90] Foray, N.; Marot, D.; Gabriel, A.; Randrianarison, V.; Carr, A. M.; Perricaudet,<br />

M.; Ashworth, A.; Jeggo, P. A subset of ATM- and ATR-dependent phosphorylation<br />

events requires the BRCA1 protein. Embo J 22:2860-2871; 2003.<br />

223


CHAPTER 5<br />

REFERENCES<br />

[91] Kim, S. T.; Xu, B.; Kastan, M. B. Involvement of the cohesin protein, Smc1, in<br />

Atm-dependent and independent responses to DNA damage. Genes Dev 16:560-570;<br />

2002.<br />

[92] Yarden, R. I.; Pardo-Reoyo, S.; Sgagias, M.; Cowan, K. H.; Brody, L. C. BRCA1<br />

regulates the G2/M checkpoint by activating Chk1 kinase upon DNA damage. Nat Genet<br />

30:285-289; 2002.<br />

[93] Gatei, M.; Zhou, B. B.; Hobson, K.; Scott, S.; Young, D.; Khanna, K. K. Ataxia<br />

telangiectasia mutated (ATM) kinase and ATM and Rad3 related kinase mediate<br />

phosphorylation of Brca1 at distinct and overlapping sites. In vivo assessment using<br />

phospho-specific antibodies. J Biol Chem 276:17276-17280; 2001.<br />

[94] Wang, Y.; Cortez, D.; Yazdi, P.; Neff, N.; Elledge, S. J.; Qin, J. BASC, a super<br />

complex of BRCA1-associated proteins involved in the recognition and repair of aberrant<br />

DNA structures. Genes Dev 14:927-939; 2000.<br />

[95] Yu, X.; Chini, C. C.; He, M.; Mer, G.; Chen, J. The BRCT domain is a phosphoprotein<br />

binding domain. Science 302:639-642; 2003.<br />

[96] Manke, I. A.; Lowery, D. M.; Nguyen, A.; Yaffe, M. B. BRCT repeats as<br />

phosphopeptide-binding modules involved in protein targeting. Science 302:636-639;<br />

2003.<br />

[97] Rodriguez, M.; Yu, X.; Chen, J.; Songyang, Z. Phosphopeptide binding<br />

specificities of BRCA1 COOH-terminal (BRCT) domains. J Biol Chem 278:52914-<br />

52918; 2003.<br />

224


CHAPTER 5<br />

REFERENCES<br />

[98] Liu, Y.; Virshup, D. M.; White, R. L.; Hsu, L. C. Regulation of BRCA1<br />

phosphorylation by interaction with protein phosphatase 1alpha. Cancer Res 62:6357-<br />

6361; 2002.<br />

[99] Li, N.; Banin, S.; Ouyang, H.; Li, G. C.; Courtois, G.; Shiloh, Y.; Karin, M.;<br />

Rotman, G. ATM is required for IkappaB kinase (IKKk) activation in response to DNA<br />

double strand breaks. J Biol Chem 276:8898-8903; 2001.<br />

[100] Sapkota, G. P.; Deak, M.; Kieloch, A.; Morrice, N.; Goodarzi, A. A.; Smythe, C.;<br />

Shiloh, Y.; Lees-Miller, S. P.; Alessi, D. R. Ionizing radiation induces ataxia<br />

telangiectasia mutated kinase (ATM)-mediated phosphorylation of LKB1/STK11 at Thr-<br />

366. Biochem J 368:507-516; 2002.<br />

[101] Baskaran, R.; Wood, L. D.; Whitaker, L. L.; Canman, C. E.; Morgan, S. E.; Xu,<br />

Y.; Barlow, C.; Baltimore, D.; Wynshaw-Boris, A.; Kastan, M. B.; Wang, J. Y. Ataxia<br />

telangiectasia mutant protein activates c-Abl tyrosine kinase in response to ionizing<br />

radiation. Nature 387:516-519; 1997.<br />

[102] Gatei, M.; Sloper, K.; Sorensen, C.; Syljuasen, R.; Falck, J.; Hobson, K.; Savage,<br />

K.; Lukas, J.; Zhou, B. B.; Bartek, J.; Khanna, K. K. Ataxia-telangiectasia-mutated<br />

(ATM) and NBS1-dependent phosphorylation of Chk1 on Ser-317 in response to ionizing<br />

radiation. J Biol Chem 278:14806-14811; 2003.<br />

[103] Zhang, J.; Willers, H.; Feng, Z.; <strong>Ghosh</strong>, J. C.; Kim, S.; Weaver, D. T.; Chung, J.<br />

H.; Powell, S. N.; Xia, F. Chk2 phosphorylation of BRCA1 regulates DNA double-strand<br />

break repair. Mol Cell Biol 24:708-718; 2004.<br />

225


CHAPTER 5<br />

REFERENCES<br />

[104] Lee, J. S.; Collins, K. M.; Brown, A. L.; Lee, C. H.; Chung, J. H. hCds1-mediated<br />

phosphorylation of BRCA1 regulates the DNA damage response. Nature 404:201-204;<br />

2000.<br />

[105] Chen, G.; Yuan, S. S.; Liu, W.; Xu, Y.; Trujillo, K.; Song, B.; Cong, F.; Goff, S.<br />

P.; Wu, Y.; Arlinghaus, R.; Baltimore, D.; Gasser, P. J.; Park, M. S.; Sung, P.; Lee, E. Y.<br />

Radiation-induced assembly of Rad51 and Rad52 recombination complex requires ATM<br />

and c-Abl. J Biol Chem 274:12748-12752; 1999.<br />

[106] Foray, N.; Marot, D.; Randrianarison, V.; Venezia, N. D.; Picard, D.; Perricaudet,<br />

M.; Favaudon, V.; Jeggo, P. Constitutive association of BRCA1 and c-Abl and its ATMdependent<br />

disruption after irradiation. Mol Cell Biol 22:4020-4032; 2002.<br />

[107] Powell, S. N.; Kachnic, L. A. Roles of BRCA1 and BRCA2 in homologous<br />

recombination, DNA replication fidelity and the cellular response to ionizing radiation.<br />

Oncogene 22:5784-5791; 2003.<br />

[108] Villunger, A.; Michalak, E. M.; Coultas, L.; Mullauer, F.; Bock, G.;<br />

Ausserlechner, M. J.; Adams, J. M.; Strasser, A. p53- and drug-induced apoptotic<br />

responses mediated by BH3-only proteins puma and noxa. Science 302:1036-1038; 2003.<br />

[109] Siliciano, J. D.; Canman, C. E.; Taya, Y.; Sakaguchi, K.; Appella, E.; Kastan, M.<br />

B. DNA damage induces phosphorylation of the amino terminus of p53. Genes Dev<br />

11:3471-3481; 1997.<br />

[110] Saito, S.; Goodarzi, A. A.; Higashimoto, Y.; Noda, Y.; Lees-Miller, S. P.;<br />

Appella, E.; Anderson, C. W. ATM mediates phosphorylation at multiple p53 sites,<br />

including Ser(46), in response to ionizing radiation. J Biol Chem 277:12491-12494;<br />

2002.<br />

226


CHAPTER 5<br />

REFERENCES<br />

[111] Fei, P.; El-Deiry, W. S. P53 and radiation responses. Oncogene 22:5774-5783;<br />

2003.<br />

[112] el-Deiry, W. S.; Tokino, T.; Velculescu, V. E.; Levy, D. B.; Parsons, R.; Trent, J.<br />

M.; Lin, D.; Mercer, W. E.; Kinzler, K. W.; Vogelstein, B. WAF1, a potential mediator<br />

of p53 tumor suppression. Cell 75:817-825; 1993.<br />

[113] Noda, A.; Ning, Y.; Venable, S. F.; Pereira-Smith, O. M.; Smith, J. R. Cloning of<br />

senescent cell-derived inhibitors of DNA synthesis using an expression screen. Exp Cell<br />

Res 211:90-98; 1994.<br />

[114] Sherr, C. J.; Roberts, J. M. CDK inhibitors: positive and negative regulators of<br />

G1-phase progression. Genes Dev 13:1501-1512; 1999.<br />

[115] LaBaer, J.; Garrett, M. D.; Stevenson, L. F.; Slingerland, J. M.; Sandhu, C.; Chou,<br />

H. S.; Fattaey, A.; Harlow, E. New functional activities for the p21 family of CDK<br />

inhibitors. Genes Dev 11:847-862; 1997.<br />

[116] Harper, J. W.; Elledge, S. J.; Keyomarsi, K.; Dynlacht, B.; Tsai, L. H.; Zhang, P.;<br />

Dobrowolski, S.; Bai, C.; Connell-Crowley, L.; Swindell, E.; et al. Inhibition of cyclindependent<br />

kinases by p21. Mol Biol Cell 6:387-400; 1995.<br />

[117] Smits, V. A.; Klompmaker, R.; Vallenius, T.; Rijksen, G.; Makela, T. P.;<br />

Medema, R. H. p21 inhibits Thr161 phosphorylation of Cdc2 to enforce the G2 DNA<br />

damage checkpoint. J Biol Chem 275:30638-30643; 2000.<br />

[118] Roninson, I. B. Oncogenic functions of tumour suppressor p21(Waf1/Cip1/Sdi1):<br />

association with cell senescence and tumour-promoting activities of stromal fibroblasts.<br />

Cancer Lett 179:1-14; 2002.<br />

227


CHAPTER 5<br />

REFERENCES<br />

[119] Jascur, T.; Brickner, H.; Salles-Passador, I.; Barbier, V.; El Khissiin, A.; Smith,<br />

B.; Fotedar, R.; Fotedar, A. Regulation of p21(WAF1/CIP1) stability by WISp39, a<br />

Hsp90 binding TPR protein. Mol Cell 17:237-249; 2005.<br />

[120] Waldman, T.; Lengauer, C.; Kinzler, K. W.; Vogelstein, B. Uncoupling of S<br />

phase and mitosis induced by anticancer agents in cells lacking p21. Nature 381:713-716;<br />

1996.<br />

[121] Wood, R. D.; Mitchell, M.; Sgouros, J.; Lindahl, T. Human DNA repair genes.<br />

Science 291:1284-1289; 2001.<br />

[122] Sonoda, E.; Sasaki, M. S.; Morrison, C.; Yamaguchi-Iwai, Y.; Takata, M.;<br />

Takeda, S. Sister chromatid exchanges are mediated by homologous recombination in<br />

vertebrate cells. Mol Cell Biol 19:5166-5169; 1999.<br />

[123] Moynahan, M. E.; Chiu, J. W.; Koller, B. H.; Jasin, M. Brca1 controls homologydirected<br />

DNA repair. Mol Cell 4:511-518; 1999.<br />

[124] Jimenez, G. S.; Bryntesson, F.; Torres-Arzayus, M. I.; Priestley, A.; Beeche, M.;<br />

Saito, S.; Sakaguchi, K.; Appella, E.; Jeggo, P. A.; Taccioli, G. E.; Wahl, G. M.; Hubank,<br />

M. DNA-dependent protein kinase is not required for the p53-dependent response to<br />

DNA damage. Nature 400:81-83; 1999.<br />

[125] Modesti, M.; Hesse, J. E.; Gellert, M. DNA binding of Xrcc4 protein is associated<br />

with V(D)J recombination but not with stimulation of DNA ligase IV activity. Embo J<br />

18:2008-2018; 1999.<br />

[126] Singleton, B. K.; Torres-Arzayus, M. I.; Rottinghaus, S. T.; Taccioli, G. E.;<br />

Jeggo, P. A. The C terminus of Ku80 activates the DNA-dependent protein kinase<br />

catalytic subunit. Mol Cell Biol 19:3267-3277; 1999.<br />

228


CHAPTER 5<br />

REFERENCES<br />

[127] Frank, K. M.; Sekiguchi, J. M.; Seidl, K. J.; Swat, W.; Rathbun, G. A.; Cheng, H.<br />

L.; Davidson, L.; Kangaloo, L.; Alt, F. W. Late embryonic lethality and impaired V(D)J<br />

recombination in mice lacking DNA ligase IV. Nature 396:173-177; 1998.<br />

[128] Barnes, D. E.; Stamp, G.; Rosewell, I.; Denzel, A.; Lindahl, T. Targeted<br />

disruption of the gene encoding DNA ligase IV leads to lethality in embryonic mice.<br />

Curr Biol 8:1395-1398; 1998.<br />

[129] Gao, Y.; Sun, Y.; Frank, K. M.; Dikkes, P.; Fujiwara, Y.; Seidl, K. J.; Sekiguchi,<br />

J. M.; Rathbun, G. A.; Swat, W.; Wang, J.; Bronson, R. T.; Malynn, B. A.; Bryans, M.;<br />

Zhu, C.; Chaudhuri, J.; Davidson, L.; Ferrini, R.; Stamato, T.; Orkin, S. H.; Greenberg,<br />

M. E.; Alt, F. W. A critical role for DNA end-joining proteins in both lymphogenesis and<br />

neurogenesis. Cell 95:891-902; 1998.<br />

[130] Riballo, E.; Critchlow, S. E.; Teo, S. H.; Doherty, A. J.; Priestley, A.; Broughton,<br />

B.; Kysela, B.; Beamish, H.; Plowman, N.; Arlett, C. F.; Lehmann, A. R.; Jackson, S. P.;<br />

Jeggo, P. A. Identification of a defect in DNA ligase IV in a radiosensitive leukaemia<br />

patient. Curr Biol 9:699-702; 1999.<br />

[131] Baumann, P.; West, S. C. Role of the human RAD51 protein in homologous<br />

recombination and double-stranded-break repair. Trends Biochem Sci 23:247-251; 1998.<br />

[132] Lee, S. E.; Mitchell, R. A.; Cheng, A.; Hendrickson, E. A. Evidence for DNA-<br />

PK-dependent and -independent DNA double-strand break repair pathways in<br />

mammalian cells as a function of the cell cycle. Mol Cell Biol 17:1425-1433; 1997.<br />

[133] New, J. H.; Sugiyama, T.; Zaitseva, E.; Kowalczykowski, S. C. Rad52 protein<br />

stimulates DNA strand exchange by Rad51 and replication protein A. Nature 391:407-<br />

410; 1998.<br />

229


CHAPTER 5<br />

REFERENCES<br />

[134] Benson, F. E.; Baumann, P.; West, S. C. Synergistic actions of Rad51 and Rad52<br />

in recombination and DNA repair. Nature 391:401-404; 1998.<br />

[135] Van Dyck, E.; Stasiak, A. Z.; Stasiak, A.; West, S. C. Binding of double-strand<br />

breaks in DNA by human Rad52 protein. Nature 398:728-731; 1999.<br />

[136] Lim, D. S.; Hasty, P. A mutation in mouse rad51 results in an early embryonic<br />

lethal that is suppressed by a mutation in p53. Mol Cell Biol 16:7133-7143; 1996.<br />

[137] Rijkers, T.; Van Den Ouweland, J.; Morolli, B.; Rolink, A. G.; Baarends, W. M.;<br />

Van Sloun, P. P.; Lohman, P. H.; Pastink, A. Targeted inactivation of mouse RAD52<br />

reduces homologous recombination but not resistance to ionizing radiation. Mol Cell Biol<br />

18:6423-6429; 1998.<br />

[138] Essers, J.; Hendriks, R. W.; Swagemakers, S. M.; Troelstra, C.; de Wit, J.;<br />

Bootsma, D.; Hoeijmakers, J. H.; Kanaar, R. Disruption of mouse RAD54 reduces<br />

ionizing radiation resistance and homologous recombination. Cell 89:195-204; 1997.<br />

[139] Tan, T. L.; Essers, J.; Citterio, E.; Swagemakers, S. M.; de Wit, J.; Benson, F. E.;<br />

Hoeijmakers, J. H.; Kanaar, R. Mouse Rad54 affects DNA conformation and DNAdamage-induced<br />

Rad51 foci formation. Curr Biol 9:325-328; 1999.<br />

[140] Liu, N.; Lamerdin, J. E.; Tebbs, R. S.; Schild, D.; Tucker, J. D.; Shen, M. R.;<br />

Brookman, K. W.; Siciliano, M. J.; Walter, C. A.; Fan, W.; Narayana, L. S.; Zhou, Z. Q.;<br />

Adamson, A. W.; Sorensen, K. J.; Chen, D. J.; Jones, N. J.; Thompson, L. H. XRCC2 and<br />

XRCC3, new human Rad51-family members, promote chromosome stability and protect<br />

against DNA cross-links and other damages. Mol Cell 1:783-793; 1998.<br />

230


CHAPTER 5<br />

REFERENCES<br />

[141] Pierce, A. J.; Johnson, R. D.; Thompson, L. H.; Jasin, M. XRCC3 promotes<br />

homology-directed repair of DNA damage in mammalian cells. Genes Dev 13:2633-<br />

2638; 1999.<br />

[142] Callebaut, I.; Mornon, J. P. From BRCA1 to RAP1: a widespread BRCT module<br />

closely associated with DNA repair. FEBS Lett 400:25-30; 1997.<br />

[143] Gowen, L. C.; Avrutskaya, A. V.; Latour, A. M.; Koller, B. H.; Leadon, S. A.<br />

BRCA1 required for transcription-coupled repair of oxidative DNA damage (Retracted<br />

article. See vol 300, pg 1657, June 13 2003). Science 281:1009-1012; 1998.<br />

[144] Cortez, D.; Wang, Y.; Qin, J.; Elledge, S. J. Requirement of ATM-dependent<br />

phosphorylation of brca1 in the DNA damage response to double-strand breaks. Science<br />

286:1162-1166; 1999.<br />

[145] Chen, J. J.; Silver, D. P.; Walpita, D.; Cantor, S. B.; Gazdar, A. F.; Tomlinson,<br />

G.; Couch, F. J.; Weber, B. L.; Ashley, T.; Livingston, D. M.; Scully, R. Stable<br />

interaction between the products of the BRCA1 and BRCA2 tumor suppressor genes in<br />

mitotic and meiotic cells. Molecular Cell 2:317-328; 1998.<br />

[146] Patel, K. J.; Yu, V. P. C. C.; Lee, H. S.; Corcoran, A.; Thistlethwaite, F. C.;<br />

Evans, M. J.; Colledge, W. H.; Friedman, L. S.; Ponder, B. A. J.; Venkitaraman, A. R.<br />

Involvement of Brca2 in DNA repair. Molecular Cell 1:347-357; 1998.<br />

[147] Zhong, Q.; Chen, C. F.; Li, S.; Chen, Y. M.; Wang, C. C.; Xiao, J.; Chen, P. L.;<br />

Sharp, Z. D.; Lee, W. H. Association of BRCA1 with the hRad50-hMre11-p95 complex<br />

and the DNA damage response. Science 285:747-750; 1999.<br />

231


CHAPTER 5<br />

REFERENCES<br />

[148] Cressman, V. L.; Backlund, D. C.; Avrutskaya, A. V.; Leadon, S. A.; Godfrey,<br />

V.; Koller, B. H. Growth retardation, DNA repair defects, and lack of spermatogenesis in<br />

BRCA1-deficient mice. Mol Cell Biol 19:7061-7075; 1999.<br />

[149] Shafman, T.; Khanna, K. K.; Kedar, P.; Spring, K.; Kozlov, S.; Yen, T.; Hobson,<br />

K.; Gatei, M.; Zhang, N.; Watters, D.; Egerton, M.; Shiloh, Y.; Kharbanda, S.; Kufe, D.;<br />

Lavin, M. F. Interaction between ATM protein and c-Abl in response to DNA damage.<br />

Nature 387:520-523; 1997.<br />

[150] Winkler, G. S.; Vermeulen, W.; Coin, F.; Egly, J. M.; Hoeijmakers, J. H.; Weeda,<br />

G. Affinity purification of human DNA repair/transcription factor TFIIH using epitopetagged<br />

xeroderma pigmentosum B protein. J Biol Chem 273:1092-1098; 1998.<br />

[151] Balajee, A. S.; Bohr, V. A. Genomic heterogeneity of nucleotide excision repair.<br />

Gene 250:15-30; 2000.<br />

[152] Wood, R. D. Nucleotide excision repair in mammalian cells. J Biol Chem<br />

272:23465-23468; 1997.<br />

[153] Evans, A. R.; Limp-Foster, M.; Kelley, M. R. Going APE over ref-1. Mutat Res-<br />

DNA Repair 461:83-108; 2000.<br />

[154] Hardman, R. A.; Afshari, C. A.; Barrett, J. C. Involvement of mammalian MLH1<br />

in the apoptotic response to peroxide-induced oxidative stress. Cancer Res 61:1392-<br />

1397; 2001.<br />

[155] Ward, J. F. DNA damage as the cause of ionizing radiation-induced gene<br />

activation. Radiat Res 138:S85-88; 1994.<br />

232


CHAPTER 5<br />

REFERENCES<br />

[156] Roots, R.; Okada, S. Protection of DNA molecules of cultured mammalian cells<br />

from radiation-induced single-strand scissions by various alcohols and SH compounds.<br />

Int J Radiat Biol Relat Stud Phys Chem Med 21:329-342; 1972.<br />

[157] Ibuki, Y.; Goto, R. Enhancement of NO production from resident peritoneal<br />

macrophages by in vitro gamma-irradiation and its relationship to reactive oxygen<br />

intermediates. Free Radic Biol Med 22:1029-1035; 1997.<br />

[158] Beckman, J. S.; Koppenol, W. H. Nitric oxide, superoxide, and peroxynitrite: the<br />

good, the bad, and ugly. Am J Physiol 271:C1424-1437; 1996.<br />

[159] Clutton, S. M.; Townsend, K. M.; Walker, C.; Ansell, J. D.; Wright, E. G.<br />

Radiation-induced genomic instability and persisting oxidative stress in primary bone<br />

marrow cultures. Carcinogenesis 17:1633-1639; 1996.<br />

[160] Wu, L. J.; Randers-Pehrson, G.; Xu, A.; Waldren, C. A.; Geard, C. R.; Yu, Z.;<br />

Hei, T. K. Targeted cytoplasmic irradiation with alpha particles induces mutations in<br />

mammalian cells. Proc Natl Acad Sci U S A 96:4959-4964; 1999.<br />

[161] Rossman, T. G.; Goncharova, E. I. Spontaneous mutagenesis in mammalian cells<br />

is caused mainly by oxidative events and can be blocked by antioxidants and<br />

metallothionein. Mutat Res 402:103-110; 1998.<br />

[162] Narayanan, P. K.; Goodwin, E. H.; Lehnert, B. E. Alpha particles initiate<br />

biological production of superoxide anions and hydrogen peroxide in human cells.<br />

Cancer Res 57:3963-3971; 1997.<br />

[163] Morales, A.; Miranda, M.; Sanchez-Reyes, A.; Biete, A.; Fernandez-Checa, J. C.<br />

Oxidative damage of mitochondrial and nuclear DNA induced by ionizing radiation in<br />

human hepatoblastoma cells. Int J Radiat Oncol Biol Phys 42:191-203; 1998.<br />

233


CHAPTER 5<br />

REFERENCES<br />

[164] Lander, H. M.; Ogiste, J. S.; Teng, K. K.; Novogrodsky, A. p21ras as a common<br />

signaling target of reactive free radicals and cellular redox stress. J Biol Chem<br />

270:21195-21198; 1995.<br />

[165] Bae, Y. S.; Kang, S. W.; Seo, M. S.; Baines, I. C.; Tekle, E.; Chock, P. B.; Rhee,<br />

S. G. Epidermal growth factor (EGF)-induced generation of hydrogen peroxide. Role in<br />

EGF receptor-mediated tyrosine phosphorylation. J Biol Chem 272:217-221; 1997.<br />

[166] Lee, S. R.; Kwon, K. S.; Kim, S. R.; Rhee, S. G. Reversible inactivation of<br />

protein-tyrosine phosphatase 1B in A431 cells stimulated with epidermal growth factor. J<br />

Biol Chem 273:15366-15372; 1998.<br />

[167] Lander, H. M.; Ogiste, J. S.; Pearce, S. F.; Levi, R.; Novogrodsky, A. Nitric<br />

oxide-stimulated guanine nucleotide exchange on p21ras. J Biol Chem 270:7017-7020;<br />

1995.<br />

[168] Ritov, V. B.; Banni, S.; Yalowich, J. C.; Day, B. W.; Claycamp, H. G.; Corongiu,<br />

F. P.; Kagan, V. E. Non-random peroxidation of different classes of membrane<br />

phospholipids in live cells detected by metabolically integrated cis-parinaric acid.<br />

Biochim Biophys Acta 1283:127-140; 1996.<br />

[169] Schmidt-Ullrich, R. K.; Valerie, K.; Fogleman, P. B.; Walters, J. Radiationinduced<br />

autophosphorylation of epidermal growth factor receptor in human malignant<br />

mammary and squamous epithelial cells. Radiat Res 145:81-85; 1996.<br />

[170] Rosette, C.; Karin, M. Ultraviolet light and osmotic stress: activation of the JNK<br />

cascade through multiple growth factor and cytokine receptors. Science 274:1194-1197;<br />

1996.<br />

234


CHAPTER 5<br />

REFERENCES<br />

[171] Xia, Z.; Dickens, M.; Raingeaud, J.; Davis, R. J.; Greenberg, M. E. Opposing<br />

effects of ERK and JNK-p38 MAP kinases on apoptosis. Science 270:1326-1331; 1995.<br />

[172] Schmidt-Ullrich, R. K.; Mikkelsen, R. B.; Dent, P.; Todd, D. G.; Valerie, K.;<br />

Kavanagh, B. D.; Contessa, J. N.; Rorrer, W. K.; Chen, P. B. Radiation-induced<br />

proliferation of the human A431 squamous carcinoma cells is dependent on EGFR<br />

tyrosine phosphorylation. Oncogene 15:1191-1197; 1997.<br />

[173] Contessa, J. N.; Reardon, D. B.; Todd, D.; Dent, P.; Mikkelsen, R. B.; Valerie, K.;<br />

Bowers, G. D.; Schmidt-Ullrich, R. K. The inducible expression of dominant-negative<br />

epidermal growth factor receptor-CD533 results in radiosensitization of human mammary<br />

carcinoma cells. Clin Cancer Res 5:405-411; 1999.<br />

[174] Baselga, J.; Norton, L.; Albanell, J.; Kim, Y. M.; Mendelsohn, J. Recombinant<br />

humanized anti-HER2 antibody (Herceptin) enhances the antitumor activity of paclitaxel<br />

and doxorubicin against HER2/neu overexpressing human breast cancer xenografts.<br />

Cancer Res 58:2825-2831; 1998.<br />

[175] Reardon, D. B.; Contessa, J. N.; Mikkelsen, R. B.; Valerie, K.; Amir, C.; Dent, P.;<br />

Schmidt-Ullrich, R. K. Dominant negative EGFR-CD533 and inhibition of MAPK<br />

modify JNK1 activation and enhance radiation toxicity of human mammary carcinoma<br />

cells. Oncogene 18:4756-4766; 1999.<br />

[176] Kashles, O.; Yarden, Y.; Fischer, R.; Ullrich, A.; Schlessinger, J. A dominant<br />

negative mutation suppresses the function of normal epidermal growth factor receptors<br />

by heterodimerization. Mol Cell Biol 11:1454-1463; 1991.<br />

[177] Todd, D. G.; Mikkelsen, R. B.; Rorrer, W. K.; Valerie, K.; Schmidt-Ullrich, R. K.<br />

Ionizing radiation stimulates existing signal transduction pathways involving the<br />

235


CHAPTER 5<br />

REFERENCES<br />

activation of epidermal growth factor receptor and ERBB-3, and changes of intracellular<br />

calcium in A431 human squamous carcinoma cells. J Recept Signal Transduct Res<br />

19:885-908; 1999.<br />

[178] Mason, C. S.; Springer, C. J.; Cooper, R. G.; Superti-Furga, G.; Marshall, C. J.;<br />

Marais, R. Serine and tyrosine phosphorylations cooperate in Raf-1, but not B-Raf<br />

activation. Embo J 18:2137-2148; 1999.<br />

[179] Todd, D. G.; Mikkelsen, R. B. Ionizing radiation induces a transient increase in<br />

cytosolic free [Ca2+] in human epithelial tumor cells. Cancer Res 54:5224-5230; 1994.<br />

[180] Dent, P.; Reardon, D. B.; Morrison, D. K.; Sturgill, T. W. Regulation of Raf-1<br />

and Raf-1 mutants by Ras-dependent and Ras-independent mechanisms in vitro. Mol Cell<br />

Biol 15:4125-4135; 1995.<br />

[181] Le Gallic, L.; Sgouras, D.; Beal, G., Jr.; Mavrothalassitis, G. Transcriptional<br />

repressor ERF is a Ras/mitogen-activated protein kinase target that regulates cellular<br />

proliferation. Mol Cell Biol 19:4121-4133; 1999.<br />

[182] Kuroki, M.; O'Flaherty, J. T. Extracellular signal-regulated protein kinase (ERK)-<br />

dependent and ERK-independent pathways target STAT3 on serine-727 in human<br />

neutrophils stimulated by chemotactic factors and cytokines. Biochem J 341 ( Pt 3):691-<br />

696; 1999.<br />

[183] Navarro, P.; Valverde, A. M.; Benito, M.; Lorenzo, M. Activated Ha-ras induces<br />

apoptosis by association with phosphorylated Bcl-2 in a mitogen-activated protein<br />

kinase-independent manner. J Biol Chem 274:18857-18863; 1999.<br />

[184] Carter, S.; Auer, K. L.; Reardon, D. B.; Birrer, M.; Fisher, P. B.; Valerie, K.;<br />

Schmidt-Ullrich, R.; Mikkelsen, R.; Dent, P. Inhibition of the mitogen activated protein<br />

236


CHAPTER 5<br />

REFERENCES<br />

(MAP) kinase cascade potentiates cell killing by low dose ionizing radiation in A431<br />

human squamous carcinoma cells. Oncogene 16:2787-2796; 1998.<br />

[185] Abbott, D. W.; Holt, J. T. Mitogen-activated protein kinase kinase 2 activation is<br />

essential for progression through the G2/M checkpoint arrest in cells exposed to ionizing<br />

radiation. J Biol Chem 274:2732-2742; 1999.<br />

[186] Bernhard, E. J.; McKenna, W. G.; Hamilton, A. D.; Sebti, S. M.; Qian, Y.; Wu, J.<br />

M.; Muschel, R. J. Inhibiting Ras prenylation increases the radiosensitivity of human<br />

tumor cell lines with activating mutations of ras oncogenes. Cancer Res 58:1754-1761;<br />

1998.<br />

[187] Gokhale, P. C.; McRae, D.; Monia, B. P.; Bagg, A.; Rahman, A.; Dritschilo, A.;<br />

Kasid, U. Antisense raf oligodeoxyribonucleotide is a radiosensitizer in vivo. Antisense<br />

Nucleic Acid Drug Dev 9:191-201; 1999.<br />

[188] Tombes, R. M.; Auer, K. L.; Mikkelsen, R.; Valerie, K.; Wymann, M. P.;<br />

Marshall, C. J.; McMahon, M.; Dent, P. The mitogen-activated protein (MAP) kinase<br />

cascade can either stimulate or inhibit DNA synthesis in primary cultures of rat<br />

hepatocytes depending upon whether its activation is acute/phasic or chronic. Biochem J<br />

330 ( Pt 3):1451-1460; 1998.<br />

[189] Auer, K. L.; Park, J. S.; Seth, P.; Coffey, R. J.; Darlington, G.; Abo, A.;<br />

McMahon, M.; Depinho, R. A.; Fisher, P. B.; Dent, P. Prolonged activation of the<br />

mitogen-activated protein kinase pathway promotes DNA synthesis in primary<br />

hepatocytes from p21Cip-1/WAF1-null mice, but not in hepatocytes from p16INK4a-null<br />

mice. Biochem J 336 ( Pt 3):551-560; 1998.<br />

237


CHAPTER 5<br />

REFERENCES<br />

[190] Sewing, A.; Wiseman, B.; Lloyd, A. C.; Land, H. High-intensity Raf signal<br />

causes cell cycle arrest mediated by p21Cip1. Mol Cell Biol 17:5588-5597; 1997.<br />

[191] Woods, D.; Parry, D.; Cherwinski, H.; Bosch, E.; Lees, E.; McMahon, M. Rafinduced<br />

proliferation or cell cycle arrest is determined by the level of Raf activity with<br />

arrest mediated by p21Cip1. Mol Cell Biol 17:5598-5611; 1997.<br />

[192] Fiddes, R. J.; Janes, P. W.; Sivertsen, S. P.; Sutherland, R. L.; Musgrove, E. A.;<br />

Daly, R. J. Inhibition of the MAP kinase cascade blocks heregulin-induced cell cycle<br />

progression in T-47D human breast cancer cells. Oncogene 16:2803-2813; 1998.<br />

[193] Macleod, K. F.; Sherry, N.; Hannon, G.; Beach, D.; Tokino, T.; Kinzler, K.;<br />

Vogelstein, B.; Jacks, T. p53-dependent and independent expression of p21 during cell<br />

growth, differentiation, and DNA damage. Genes Dev 9:935-944; 1995.<br />

[194] Poon, R. Y.; Chau, M. S.; Yamashita, K.; Hunter, T. The role of Cdc2 feedback<br />

loop control in the DNA damage checkpoint in mammalian cells. Cancer Res 57:5168-<br />

5178; 1997.<br />

[195] Leach, S. D.; Scatena, C. D.; Keefer, C. J.; Goodman, H. A.; Song, S. Y.; Yang,<br />

L.; Pietenpol, J. A. Negative regulation of Wee1 expression and Cdc2 phosphorylation<br />

during p53-mediated growth arrest and apoptosis. Cancer Res 58:3231-3236; 1998.<br />

[196] Pap, M.; Cooper, G. M. Role of glycogen synthase kinase-3 in the<br />

phosphatidylinositol 3-Kinase/Akt cell survival pathway. J Biol Chem 273:19929-19932;<br />

1998.<br />

[197] Shaw, M.; Cohen, P.; Alessi, D. R. Further evidence that the inhibition of<br />

glycogen synthase kinase-3beta by IGF-1 is mediated by PDK1/PKB-induced<br />

238


CHAPTER 5<br />

REFERENCES<br />

phosphorylation of Ser-9 and not by dephosphorylation of Tyr-216. FEBS Lett 416:307-<br />

311; 1997.<br />

[198] Magnusson, C.; Vaux, D. L. Signalling by CD95 and TNF receptors: not only life<br />

and death. Immunol Cell Biol 77:41-46; 1999.<br />

[199] Davis, R. J. Signal transduction by the c-Jun N-terminal kinase. Biochem Soc<br />

Symp 64:1-12; 1999.<br />

[200] Verheij, M.; Bose, R.; Lin, X. H.; Yao, B.; Jarvis, W. D.; Grant, S.; Birrer, M. J.;<br />

Szabo, E.; Zon, L. I.; Kyriakis, J. M.; Haimovitz-Friedman, A.; Fuks, Z.; Kolesnick, R.<br />

N. Requirement for ceramide-initiated SAPK/JNK signalling in stress-induced apoptosis.<br />

Nature 380:75-79; 1996.<br />

[201] Chmura, S. J.; Mauceri, H. J.; Advani, S.; Heimann, R.; Beckett, M. A.;<br />

Nodzenski, E.; Quintans, J.; Kufe, D. W.; Weichselbaum, R. R. Decreasing the apoptotic<br />

threshold of tumor cells through protein kinase C inhibition and sphingomyelinase<br />

activation increases tumor killing by ionizing radiation. Cancer Res 57:4340-4347; 1997.<br />

[202] Haimovitz-Friedman, A. Radiation-induced signal transduction and stress<br />

response. Radiat Res 150:S102-108; 1998.<br />

[203] Auer, K. L.; Contessa, J.; Brenz-Verca, S.; Pirola, L.; Rusconi, S.; Cooper, G.;<br />

Abo, A.; Wymann, M. P.; Davis, R. J.; Birrer, M.; Dent, P. The<br />

Ras/Rac1/Cdc42/SEK/JNK/c-Jun cascade is a key pathway by which agonists stimulate<br />

DNA synthesis in primary cultures of rat hepatocytes. Mol Biol Cell 9:561-573; 1998.<br />

[204] Potapova, O.; Haghighi, A.; Bost, F.; Liu, C.; Birrer, M. J.; Gjerset, R.; Mercola,<br />

D. The Jun kinase/stress-activated protein kinase pathway functions to regulate DNA<br />

239


CHAPTER 5<br />

REFERENCES<br />

repair and inhibition of the pathway sensitizes tumor cells to cisplatin. J Biol Chem<br />

272:14041-14044; 1997.<br />

[205] Prise, K. M.; Ahnstrom, G.; Belli, M.; Carlsson, J.; Frankenberg, D.; Kiefer, J.;<br />

Lobrich, M.; Michael, B. D.; Nygren, J.; Simone, G.; Stenerlow, B. A review of dsb<br />

induction data for varying quality radiations. Int J Radiat Biol 74:173-184; 1998.<br />

[206] Tilly, N.; Fernandez-Varea, J. M.; Grusell, E.; Brahme, A. Comparison of Monte<br />

Carlo calculated electron slowing-down spectra generated by 60Co gamma-rays,<br />

electrons, protons and light ions. Phys Med Biol 47:1303-1319; 2002.<br />

[207] Kagawa, Y.; Shimazu, T.; Gordon, A. J.; Fukunishi, N.; Inabe, N.; Suzuki, M.;<br />

Hirano, M.; Kato, T.; Watanabe, M.; Hanaoka, F.; Yatagai, F. Complex hprt deletion<br />

events are recovered after exposure of human lymphoblastoid cells to high-LET carbon<br />

and neon ion beams. Mutagenesis 14:199-205; 1999.<br />

[208] Masumura, K.; Kuniya, K.; Kurobe, T.; Fukuoka, M.; Yatagai, F.; Nohmi, T.<br />

Heavy-ion-induced mutations in the gpt delta transgenic mouse: comparison of mutation<br />

spectra induced by heavy-ion, X-ray, and gamma-ray radiation. Environ Mol Mutagen<br />

40:207-215; 2002.<br />

[209] Singleton, B. K.; Griffin, C. S.; Thacker, J. Clustered DNA damage leads to<br />

complex genetic changes in irradiated human cells. Cancer Res 62:6263-6269; 2002.<br />

[210] Albertini, R. J.; Clark, L. S.; Nicklas, J. A.; O'Neill, J. P.; Hui, T. E.; Jostes, R.<br />

Radiation quality affects the efficiency of induction and the molecular spectrum of HPRT<br />

mutations in human T cells. Radiat Res 148:S76-86; 1997.<br />

240


CHAPTER 5<br />

REFERENCES<br />

[211] Anderson, J. A.; Harper, J. V.; Cucinotta, F. A.; O'Neill, P. Participation of DNA-<br />

PKcs in DSB repair after exposure to high- and low-LET radiation. Radiat Res 174:195-<br />

205; 2010.<br />

[212] Zafar, F.; Seidler, S. B.; Kronenberg, A.; Schild, D.; Wiese, C. Homologous<br />

recombination contributes to the repair of DNA double-strand breaks induced by highenergy<br />

iron ions. Radiat Res 173:27-39; 2010.<br />

[213] Harper, J. V.; Anderson, J. A.; O'Neill, P. Radiation induced DNA DSBs:<br />

Contribution from stalled replication forks? DNA Repair (Amst) 9:907-913; 2010.<br />

[214] Asaithamby, A.; Uematsu, N.; Chatterjee, A.; Story, M. D.; Burma, S.; Chen, D.<br />

J. Repair of HZE-particle-induced DNA double-strand breaks in normal human<br />

fibroblasts. Radiat Res 169:437-446; 2008.<br />

[215] Hall, E. J. Radiobiology for the radiobiologist. J.B. Lippincott Company,<br />

Philadelphia.; 1994.<br />

[216] Mothersill, C.; Seymour, C. Medium from irradiated human epithelial cells but<br />

not human fibroblasts reduces the clonogenic survival of unirradiated cells. International<br />

Journal of Radiation Biology 71:421-427; 1997.<br />

[217] Mothersill, C.; Seymour, C. B. Cell-cell contact during gamma irradiation is not<br />

required to induce a bystander effect in normal human keratinocytes: evidence for release<br />

during irradiation of a signal controlling survival into the medium. Radiat Res 149:256-<br />

262; 1998.<br />

[218] Seymour, C.; Mothersill, C. Cell communication and the "bystander effect".<br />

Radiat Res 151:505-506; 1999.<br />

241


CHAPTER 5<br />

REFERENCES<br />

[219] Seymour, C. B.; Mothersill, C. Relative contribution of bystander and targeted<br />

cell killing to the low-dose region of the radiation dose-response curve. Radiat Res<br />

153:508-511; 2000.<br />

[220] Nagasawa, H.; Little, J. B. Induction of sister chromatid exchanges by extremely<br />

low doses of alpha-particles. Cancer Res 52:6394-6396; 1992.<br />

[221] Folkard, M.; Schettino, G.; Vojnovic, B.; Gilchrist, S.; Michette, A. G.;<br />

Pfauntsch, S. J.; Prise, K. M.; Michael, B. D. A focused ultrasoft x-ray microbeam for<br />

targeting cells individually with submicrometer accuracy. Radiat Res 156:796-804; 2001.<br />

[222] Randers-Pehrson, G.; Geard, C. R.; Johnson, G.; Elliston, C. D.; Brenner, D. J.<br />

The Columbia University single-ion microbeam. Radiat Res 156:210-214; 2001.<br />

[223] Ballarini, F.; Biaggi, M.; Ottolenghi, A.; Sapora, O. Cellular communication and<br />

bystander effects: a critical review for modelling low-dose radiation action. Mutat Res<br />

501:1-12; 2002.<br />

[224] Ziegler, J. F. The Stopping and Range of Ions in Matter, http://www.srim.org/.<br />

[225] Ziegler, J. F.; Bierserk, J. P.; Littmark, U. Stopping and ranges of ions in matter.<br />

Pergamon Press, New York; 1985.<br />

[226] Kraft, G. Tumor therapy with heavy charged particles. Prog Part Nucl Phys<br />

45:S473-S544; 2000.<br />

[227] Rasband, W. S.; Image, J. U. S. <strong>National</strong> Institutes of Health, Bethesda, USA.<br />

http://rsb.info.nih.gov/ij/. 1997-2006.<br />

[228] Maurya, D. K.; Salvi, V. P.; Nair, C. K. Radiation protection of DNA by ferulic<br />

acid under in vitro and in vivo conditions. Mol Cell Biochem 280:209-217; 2005.<br />

242


CHAPTER 5<br />

REFERENCES<br />

[229] Konca, K.; Lankoff, A.; Banasik, A.; Lisowska, H.; Kuszewski, T.; Gozdz, S.;<br />

Koza, Z.; Wojcik, A. A cross-platform public domain PC image-analysis program for the<br />

comet assay. Mutat Res 534:15-20; 2003.<br />

[230] Green, L. C.; Wagner, D. A.; Glogowski, J.; Skipper, P. L.; Wishnok, J. S.;<br />

Tannenbaum, S. R. Analysis of nitrate, nitrite, and [15N]nitrate in biological fluids. Anal<br />

Biochem 126:131-138; 1982.<br />

[231] Jeon, S. H.; Kang, M. G.; Kim, Y. H.; Jin, Y. H.; Lee, C.; Chung, H. Y.; Kwon,<br />

H.; Park, S. D.; Seong, R. H. A new mouse gene, SRG3, related to the SWI3 of<br />

Saccharomyces cerevisiae, is required for apoptosis induced by glucocorticoids in a<br />

thymoma cell line. J Exp Med 185:1827-1836; 1997.<br />

[232] Rogakou, E. P.; Pilch, D. R.; Orr, A. H.; Ivanova, V. S.; Bonner, W. M. DNA<br />

double-stranded breaks induce histone H2AX phosphorylation on serine 139. J Biol<br />

Chem 273:5858-5868; 1998.<br />

[233] Olive, P. L.; Banath, J. P. Phosphorylation of histone H2AX as a measure of<br />

radiosensitivity. Int J Radiat Oncol Biol Phys 58:331-335; 2004.<br />

[234] Fernandez-Capetillo, O.; Lee, A.; Nussenzweig, M.; Nussenzweig, A. H2AX: the<br />

histone guardian of the genome. DNA Repair (Amst) 3:959-967; 2004.<br />

[235] Kastan, M. B.; Lim, D. S. The many substrates and functions of ATM. Nat Rev<br />

Mol Cell Biol 1:179-186; 2000.<br />

[236] Kurz, E. U.; Lees-Miller, S. P. DNA damage-induced activation of ATM and<br />

ATM-dependent signaling pathways. DNA Repair (Amst) 3:889-900; 2004.<br />

[237] Jameel, J. K.; Rao, V. S.; Cawkwell, L.; Drew, P. J. Radioresistance in carcinoma<br />

of the breast. Breast 13:452-460; 2004.<br />

243


CHAPTER 5<br />

REFERENCES<br />

[238] Belka, C.; Jendrossek, V.; Pruschy, M.; Vink, S.; Verheij, M.; Budach, W.<br />

Apoptosis-modulating agents in combination with radiotherapy-current status and<br />

outlook. Int J Radiat Oncol Biol Phys 58:542-554; 2004.<br />

[239] Mackay, H. J.; Twelves, C. J. Protein kinase C: a target for anticancer drugs?<br />

Endocr Relat Cancer 10:389-396; 2003.<br />

[240] Osaki, M.; Oshimura, M.; Ito, H. PI3K-Akt pathway: its functions and alterations<br />

in human cancer. Apoptosis 9:667-676; 2004.<br />

[241] Lee, Y. S.; Oh, J. H.; Yoon, S.; Kwon, M. S.; Song, C. W.; Kim, K. H.; Cho, M.<br />

J.; Mollah, M. L.; Je, Y. J.; Kim, Y. D.; Kim, C. D.; Lee, J. H. Differential gene<br />

expression profiles of radioresistant non-small-cell lung cancer cell lines established by<br />

fractionated irradiation: tumor protein p53-inducible protein 3 confers sensitivity to<br />

ionizing radiation. Int J Radiat Oncol Biol Phys 77:858-866; 2010.<br />

[242] Qing, Y.; Yang, X. Q.; Zhong, Z. Y.; Lei, X.; Xie, J. Y.; Li, M. X.; Xiang, D. B.;<br />

Li, Z. P.; Yang, Z. Z.; Wang, G.; Wang, D. Microarray analysis of DNA damage repair<br />

gene expression profiles in cervical cancer cells radioresistant to 252Cf neutron and X-<br />

rays. BMC Cancer 10:71; 2010.<br />

[243] Lehman, T. A.; Bennett, W. P.; Metcalf, R. A.; Welsh, J. A.; Ecker, J.; Modali, R.<br />

V.; Ullrich, S.; Romano, J. W.; Appella, E.; Testa, J. R.; et al. p53 mutations, ras<br />

mutations, and p53-heat shock 70 protein complexes in human lung carcinoma cell lines.<br />

Cancer Res 51:4090-4096; 1991.<br />

[244] Levine, A. J. p53, the cellular gatekeeper for growth and division. Cell 88:323-<br />

331; 1997.<br />

244


CHAPTER 5<br />

REFERENCES<br />

[245] Kastan, M. B.; Onyekwere, O.; Sidransky, D.; Vogelstein, B.; Craig, R. W.<br />

Participation of p53 protein in the cellular response to DNA damage. Cancer Res<br />

51:6304-6311; 1991.<br />

[246] Chernov, M. V.; Stark, G. R. The p53 activation and apoptosis induced by DNA<br />

damage are reversibly inhibited by salicylate. Oncogene 14:2503-2510; 1997.<br />

[247] Shu, H. K.; Kim, M. M.; Chen, P.; Furman, F.; Julin, C. M.; Israel, M. A. The<br />

intrinsic radioresistance of glioblastoma-derived cell lines is associated with a failure of<br />

p53 to induce p21(BAX) expression. Proc Natl Acad Sci U S A 95:14453-14458; 1998.<br />

[248] Satyamoorthy, K.; Chehab, N. H.; Waterman, M. J.; Lien, M. C.; El-Deiry, W. S.;<br />

Herlyn, M.; Halazonetis, T. D. Aberrant regulation and function of wild-type p53 in<br />

radioresistant melanoma cells. Cell Growth Differ 11:467-474; 2000.<br />

[249] Lees-Miller, S. P.; Godbout, R.; Chan, D. W.; Weinfeld, M.; Day, R. S., 3rd;<br />

Barron, G. M.; Allalunis-Turner, J. Absence of p350 subunit of DNA-activated protein<br />

kinase from a radiosensitive human cell line. Science 267:1183-1185; 1995.<br />

[250] Sirzen, F.; Nilsson, A.; Zhivotovsky, B.; Lewensohn, R. DNA-dependent protein<br />

kinase content and activity in lung carcinoma cell lines: correlation with intrinsic<br />

radiosensitivity. Eur J Cancer 35:111-116; 1999.<br />

[251] Golding, S. E.; Rosenberg, E.; Khalil, A.; McEwen, A.; Holmes, M.; Neill, S.;<br />

Povirk, L. F.; Valerie, K. Double strand break repair by homologous recombination is<br />

regulated by cell cycle-independent signaling via ATM in human glioma cells. J Biol<br />

Chem 279:15402-15410; 2004.<br />

245


CHAPTER 5<br />

REFERENCES<br />

[252] Tribius, S.; Pidel, A.; Casper, D. ATM protein expression correlates with<br />

radioresistance in primary glioblastoma cells in culture. Int J Radiat Oncol Biol Phys<br />

50:511-523; 2001.<br />

[253] Collis, S. J.; Swartz, M. J.; Nelson, W. G.; DeWeese, T. L. Enhanced radiation<br />

and chemotherapy-mediated cell killing of human cancer cells by small inhibitory RNA<br />

silencing of DNA repair factors. Cancer Res 63:1550-1554; 2003.<br />

[254] Matthews, J. H.; Meeker, B. E.; Chapman, J. D. Response of human tumor cell<br />

lines in vitro to fractionated irradiation. Int J Radiat Oncol Biol Phys 16:133-138; 1989.<br />

[255] Mitchell, P.; Tollervey, D. mRNA stability in eukaryotes. Curr Opin Genet Dev<br />

10:193-198; 2000.<br />

[256] Ross, J. mRNA stability in mammalian cells. Microbiol Rev 59:423-450; 1995.<br />

[257] Wilusz, C. J.; Wormington, M.; Peltz, S. W. The cap-to-tail guide to mRNA<br />

turnover. Nat Rev Mol Cell Biol 2:237-246; 2001.<br />

[258] Dean, J. L.; Sully, G.; Clark, A. R.; Saklatvala, J. The involvement of AU-rich<br />

element-binding proteins in p38 mitogen-activated protein kinase pathway-mediated<br />

mRNA stabilisation. Cell Signal 16:1113-1121; 2004.<br />

[259] Krengli, M.; Hug, E. B.; Adams, J. A.; Smith, A. R.; Tarbell, N. J.; Munzenrider,<br />

J. E. Proton radiation therapy for retinoblastoma: comparison of various intraocular<br />

tumor locations and beam arrangements. Int J Radiat Oncol Biol Phys 61:583-593; 2005.<br />

[260] St Clair, W. H.; Adams, J. A.; Bues, M.; Fullerton, B. C.; La Shell, S.; Kooy, H.<br />

M.; Loeffler, J. S.; Tarbell, N. J. Advantage of protons compared to conventional X-ray<br />

or IMRT in the treatment of a pediatric patient with medulloblastoma. Int J Radiat Oncol<br />

Biol Phys 58:727-734; 2004.<br />

246


CHAPTER 5<br />

REFERENCES<br />

[261] Miralbell, R.; Crowell, C.; Suit, H. D. Potential improvement of three dimension<br />

treatment planning and proton therapy in the outcome of maxillary sinus cancer. Int J<br />

Radiat Oncol Biol Phys 22:305-310; 1992.<br />

[262] Jones, B.; Dale, R. G.; Carabe-Fernandez, A. Charged particle therapy for cancer:<br />

the inheritance of the Cavendish scientists? Appl Radiat Isot 67:371-377; 2009.<br />

[263] Chang, J. Y.; Zhang, X.; Wang, X.; Kang, Y.; Riley, B.; Bilton, S.; Mohan, R.;<br />

Komaki, R.; Cox, J. D. Significant reduction of normal tissue dose by proton<br />

radiotherapy compared with three-dimensional conformal or intensity-modulated<br />

radiation therapy in Stage I or Stage III non-small-cell lung cancer. Int J Radiat Oncol<br />

Biol Phys 65:1087-1096; 2006.<br />

[264] Fukumitsu, N.; Sugahara, S.; Nakayama, H.; Fukuda, K.; Mizumoto, M.; Abei,<br />

M.; Shoda, J.; Thono, E.; Tsuboi, K.; Tokuuye, K. A prospective study of<br />

hypofractionated proton beam therapy for patients with hepatocellular carcinoma. Int J<br />

Radiat Oncol Biol Phys 74:831-836; 2009.<br />

[265] Fei, P.; Bernhard, E. J.; El-Deiry, W. S. Tissue-specific induction of p53 targets in<br />

vivo. Cancer Res 62:7316-7327; 2002.<br />

[266] Yonish-Rouach, E.; Resnitzky, D.; Lotem, J.; Sachs, L.; Kimchi, A.; Oren, M.<br />

Wild-type p53 induces apoptosis of myeloid leukaemic cells that is inhibited by<br />

interleukin-6. Nature 352:345-347; 1991.<br />

[267] Sabbatini, P.; Lin, J.; Levine, A. J.; White, E. Essential role for p53-mediated<br />

transcription in E1A-induced apoptosis. Genes Dev 9:2184-2192; 1995.<br />

[268] Caelles, C.; Helmberg, A.; Karin, M. p53-dependent apoptosis in the absence of<br />

transcriptional activation of p53-target genes. Nature 370:220-223; 1994.<br />

247


CHAPTER 5<br />

REFERENCES<br />

[269] Polyak, K.; Xia, Y.; Zweier, J. L.; Kinzler, K. W.; Vogelstein, B. A model for<br />

p53-induced apoptosis. Nature 389:300-305; 1997.<br />

[270] Mihara, M.; Erster, S.; Zaika, A.; Petrenko, O.; Chittenden, T.; Pancoska, P.;<br />

Moll, U. M. p53 has a direct apoptogenic role at the mitochondria. Molecular Cell<br />

11:577-590; 2003.<br />

[271] Mitra, A. K.; Bhat, N.; Sarma, A.; Krishna, M. Alteration in the expression of<br />

signaling parameters following carbon ion irradiation. Mol Cell Biochem 276:169-173;<br />

2005.<br />

[272] Prise, K. M.; Pinto, M.; Newman, H. C.; Michael, B. D. A review of studies of<br />

ionizing radiation-induced double-strand break clustering. Radiat Res 156:572-576;<br />

2001.<br />

[273] Foster, E. R.; Downs, J. A. Histone H2A phosphorylation in DNA double-strand<br />

break repair. Febs J 272:3231-3240; 2005.<br />

[274] Lowndes, N. F.; Toh, G. W. DNA repair: the importance of phosphorylating<br />

histone H2AX. Curr Biol 15:R99-R102; 2005.<br />

[275] Karlsson, K. H.; Stenerlow, B. Focus formation of DNA repair proteins in normal<br />

and repair-deficient cells irradiated with high-LET ions. Radiat Res 161:517-527; 2004.<br />

[276] Deng, C. X. BRCA1: cell cycle checkpoint, genetic instability, DNA damage<br />

response and cancer evolution. Nucleic Acids Res 34:1416-1426; 2006.<br />

[277] Valerie, K.; Yacoub, A.; Hagan, M. P.; Curiel, D. T.; Fisher, P. B.; Grant, S.;<br />

Dent, P. Radiation-induced cell signaling: inside-out and outside-in. Mol Cancer Ther<br />

6:789-801; 2007.<br />

248


CHAPTER 5<br />

REFERENCES<br />

[278] Weizman, N.; Shiloh, Y.; Barzilai, A. Contribution of the Atm protein to<br />

maintaining cellular homeostasis evidenced by continuous activation of the AP-1<br />

pathway in Atm-deficient brains. J Biol Chem 278:6741-6747; 2003.<br />

[279] Lafarge, S.; Sylvain, V.; Ferrara, M.; Bignon, Y. J. Inhibition of BRCA1 leads to<br />

increased chemoresistance to microtubule-interfering agents, an effect that involves the<br />

JNK pathway. Oncogene 20:6597-6606; 2001.<br />

[280] Miyamoto, T.; Yamamoto, N.; Nishimura, H.; Koto, M.; Tsujii, H.; Mizoe, J. E.;<br />

Kamada, T.; Kato, H.; Yamada, S.; Morita, S.; Yoshikawa, K.; Kandatsu, S.; Fujisawa, T.<br />

Carbon ion radiotherapy for stage I non-small cell lung cancer. Radiother Oncol 66:127-<br />

140; 2003.<br />

[281] Ogata, T.; Teshima, T.; Kagawa, K.; Hishikawa, Y.; Takahashi, Y.; Kawaguchi,<br />

A.; Suzumoto, Y.; Nojima, K.; Furusawa, Y.; Matsuura, N. Particle irradiation suppresses<br />

metastatic potential of cancer cells. Cancer Res 65:113-120; 2005.<br />

[282] Schulz-Ertner, D.; Nikoghosyan, A.; Thilmann, C.; Haberer, T.; Jakel, O.; Karger,<br />

C.; Kraft, G.; Wannenmacher, M.; Debus, J. Results of carbon ion radiotherapy in 152<br />

patients. Int J Radiat Oncol Biol Phys 58:631-640; 2004.<br />

[283] Tsuji, H.; Ishikawa, H.; Yanagi, T.; Hirasawa, N.; Kamada, T.; Mizoe, J. E.;<br />

Kanai, T.; Tsujii, H.; Ohnishi, Y. Carbon-ion radiotherapy for locally advanced or<br />

unfavorably located choroidal melanoma: a Phase I/II dose-escalation study. Int J Radiat<br />

Oncol Biol Phys 67:857-862; 2007.<br />

[284] Wang, H.; Zeng, Z. C.; Bui, T. A.; DiBiase, S. J.; Qin, W.; Xia, F.; Powell, S. N.;<br />

Iliakis, G. Nonhomologous end-joining of ionizing radiation-induced DNA double-<br />

249


CHAPTER 5<br />

REFERENCES<br />

stranded breaks in human tumor cells deficient in BRCA1 or BRCA2. Cancer Res<br />

61:270-277; 2001.<br />

[285] Mothersill, C.; Seymour, C. Radiation-induced bystander effects: are they good,<br />

bad or both? Med Confl Surviv 21:101-110; 2005.<br />

[286] Mothersill, C.; Seymour, C. Radiation-induced bystander effects, carcinogenesis<br />

and models. Oncogene 22:7028-7033; 2003.<br />

[287] Hall, E. J.; Hei, T. K. Genomic instability and bystander effects induced by high-<br />

LET radiation. Oncogene 22:7034-7042; 2003.<br />

[288] Prise, K. M.; Folkard, M.; Michael, B. D. Bystander responses induced by low<br />

LET radiation. Oncogene 22:7043-7049; 2003.<br />

[289] Azzam, E. I.; de Toledo, S. M.; Little, J. B. Oxidative metabolism, gap junctions<br />

and the ionizing radiation-induced bystander effect. Oncogene 22:7050-7057; 2003.<br />

[290] Azzam, E. I.; De Toledo, S. M.; Spitz, D. R.; Little, J. B. Oxidative metabolism<br />

modulates signal transduction and micronucleus formation in bystander cells from alphaparticle-irradiated<br />

normal human fibroblast cultures. Cancer Res 62:5436-5442; 2002.<br />

[291] Azzam, E. I.; de Toledo, S. M.; Gooding, T.; Little, J. B. Intercellular<br />

communication is involved in the bystander regulation of gene expression in human cells<br />

exposed to very low fluences of alpha particles. Radiat Res 150:497-504; 1998.<br />

[292] Brach, M. A.; Hass, R.; Sherman, M. L.; Gunji, H.; Weichselbaum, R.; Kufe, D.<br />

Ionizing radiation induces expression and binding activity of the nuclear factor kappa B.<br />

J Clin Invest 88:691-695; 1991.<br />

[293] Ginn-Pease, M. E.; Whisler, R. L. Redox signals and NF-kappaB activation in T<br />

cells. Free Radic Biol Med 25:346-361; 1998.<br />

250


CHAPTER 5<br />

REFERENCES<br />

[294] Zhou, H.; Ivanov, V. N.; Lien, Y. C.; Davidson, M.; Hei, T. K. Mitochondrial<br />

function and nuclear factor-kappaB-mediated signaling in radiation-induced bystander<br />

effects. Cancer Res 68:2233-2240; 2008.<br />

[295] Li, N.; Karin, M. Ionizing radiation and short wavelength UV activate NF-kappaB<br />

through two distinct mechanisms. Proc Natl Acad Sci U S A 95:13012-13017; 1998.<br />

[296] Lowenstein, C. J.; Alley, E. W.; Raval, P.; Snowman, A. M.; Snyder, S. H.;<br />

Russell, S. W.; Murphy, W. J. Macrophage nitric oxide synthase gene: two upstream<br />

regions mediate induction by interferon gamma and lipopolysaccharide. Proc Natl Acad<br />

Sci U S A 90:9730-9734; 1993.<br />

[297] Xie, Q. W.; Whisnant, R.; Nathan, C. Promoter of the mouse gene encoding<br />

calcium-independent nitric oxide synthase confers inducibility by interferon gamma and<br />

bacterial lipopolysaccharide. J Exp Med 177:1779-1784; 1993.<br />

[298] Martinez-Ruiz, A.; Lamas, S. Signalling by NO-induced protein S-nitrosylation<br />

and S-glutathionylation: Convergences and divergences. Cardiovasc Res 75:220-228;<br />

2007.<br />

[299] Kadhim, M. A.; Moore, S. R.; Goodwin, E. H. Interrelationships amongst<br />

radiation-induced genomic instability, bystander effects, and the adaptive response. Mutat<br />

Res 568:21-32; 2004.<br />

[300] Maragos, C. M.; Morley, D.; Wink, D. A.; Dunams, T. M.; Saavedra, J. E.;<br />

Hoffman, A.; Bove, A. A.; Isaac, L.; Hrabie, J. A.; Keefer, L. K. Complexes of No with<br />

Nucleophiles as Agents for the Controlled Biological Release of Nitric-Oxide -<br />

Vasorelaxant Effects. J Med Chem 34:3242-3247; 1991.<br />

251


CHAPTER 5<br />

REFERENCES<br />

[301] Nathan, C. Nitric-Oxide as a Secretory Product of Mammalian-Cells. Faseb J<br />

6:3051-3064; 1992.<br />

[302] Matsumoto, H.; Hayashi, S.; Hatashita, M.; Shioura, H.; Ohtsubo, T.; Kitai, R.;<br />

Ohnishi, T.; Yukawa, O.; Furusawa, Y.; Kano, E. Induction of radioresistance to<br />

accelerated carbon-ion beams in recipient cells by nitric oxide excreted from irradiated<br />

donor cells of human glioblastoma. Int J Radiat Biol 76:1649-1657; 2000.<br />

[303] Tamir, S.; Tannenbaum, S. R. The role of nitric oxide (NO center dot) in the<br />

carcinogenic process. Bba-Rev Cancer 1288:F31-F36; 1996.<br />

[304] Lancaster, J. R. Diffusion of free nitric oxide. Method Enzymol 268:31-50; 1996.<br />

[305] Saran, M.; Bors, W. Signaling by O-2(-Center-Dot) and No-Center-Dot - How<br />

Far Can Either Radical, or Any Specific Reaction-Product, Transmit a Message under in-<br />

Vivo Conditions. Chem-Biol Interact 90:35-45; 1994.<br />

[306] Shao, C.; Stewart, V.; Folkard, M.; Michael, B. A.; Prise, K. M. Nitric oxidemediated<br />

signaling in the bystander response of individually targeted glioma cells.<br />

Cancer Research 63:8437-8442; 2003.<br />

252


PUBLICATIONS AND PRESENTATIONS<br />

Publications<br />

1. <strong>Ghosh</strong>, S.; Krishna, M. Role of Rad52 in fractionated irradiation induced signaling in<br />

A549 lung adenocarcinoma cells. Mutation Research/Fundamental and Molecular<br />

Mechanisms of Mutagenesis; 2012 Jan 3;729(1-2):61-72.<br />

2. <strong>Ghosh</strong>, S.; Narang, H.; Sarma, A.; Krishna, M. DNA damage response signaling in lung<br />

adenocarcinoma A549 cells following gamma and carbon beam irradiation. Mutation<br />

Research/Fundamental and Molecular Mechanisms of Mutagenesis; 2011 Nov 1;716(1-<br />

2):10-9.<br />

3. <strong>Ghosh</strong>, S.; Narang, H.; Sarma, A.; Kaur, H.; Krishna, M. Activation of DNA damage<br />

response signaling in lung adenocarcinoma A549 cells following oxygen beam<br />

irradiation. Mutat Res 723:190-198; 2011.<br />

4. <strong>Ghosh</strong>, S.; Bhat, N. N.; Santra, S.; Thomas, R. G.; Gupta, S. K.; Choudhury, R. K.;<br />

Krishna, M. Low energy proton beam induces efficient cell killing in A549 lung<br />

adenocarcinoma cells. Cancer Invest 28:615-622; 2010.<br />

5. <strong>Ghosh</strong>, S.; Maurya, D. K.; Krishna, M. Role of iNOS in bystander signaling between<br />

macrophages and lymphoma cells. Int J Radiat Oncol Biol Phys 72:1567-1574; 2008.<br />

253


Symposia<br />

1. <strong>Ghosh</strong> S and Krishna M. Role of Nitric Oxide in Radiation induced Bystander Signaling.<br />

Presented at International Conference on Advances in Free Radical Research: Natural<br />

products, antioxidant and radioprotector, CSM Medical University, Lucknow. March 19-<br />

21, 2009<br />

2. <strong>Ghosh</strong> S and Krishna M. Mechanism of relative radioresistance of different cell lines.<br />

Presented at DAE-BRNS 4 th<br />

Life Science Symposium on Recent Advances in<br />

Immunomodulation in Stress and Cancer, BARC, Mumbai. December 22-24, 2008<br />

3. <strong>Ghosh</strong> S and Krishna M. Radiation Induced Bystander Effect: Similar and dissimilar<br />

cells. Presented at International Conference on Free Radical & Natural Products in<br />

Health, University of Rajasthan, Jaipur, 14-16 th February, 2008.<br />

4. <strong>Ghosh</strong> S, Maurya DK and Krishna M. Molecular Mechanism of Bystander effect in<br />

similar cells: Lack of Bystander effect in dissimilar cells. Presented at DAE-BRNS<br />

Symposium 2007 on DNA damage, Repair and their Implications, BARC, Mumbai, 5-7,<br />

December, 2007.<br />

254


Mutation Research 729 (2012) 61–72<br />

Contents lists available at SciVerse ScienceDirect<br />

Mutation Research/Fundamental and Molecular<br />

Mechanisms of Mutagenesis<br />

jo ur n al hom ep a ge: www.elsevier.com/locate/molmut<br />

C om mun i ty a ddress: www.elsevier.com/locate/mutres<br />

Role of Rad52 in fractionated irradiation induced signaling in A549 lung<br />

adenocarcinoma cells<br />

<strong>Somnath</strong> <strong>Ghosh</strong>, Malini Krishna ∗<br />

Radiation Biology and Health Sciences Division, <strong>Bhabha</strong> Atomic Research Centre, Trombay, Mumbai 400085, India<br />

a r t i c l e i n f o<br />

Article history:<br />

Received 16 June 2011<br />

Received in revised form<br />

22 September 2011<br />

Accepted 27 September 2011<br />

Available online 4 October 2011<br />

Keywords:<br />

Fractionated irradiation<br />

ATM<br />

BRCA1<br />

DNA repair<br />

Rad52<br />

a b s t r a c t<br />

The effect of fractionated doses of -irradiation (2 Gy per fraction over 5 days), as delivered in cancer<br />

radiotherapy, was compared with acute doses of 10 and 2 Gy, in A549 cells. A549 cells were found to<br />

be relatively more radioresistant if the 10 Gy dose was delivered as a fractionated regimen. Microarray<br />

analysis showed upregulation of DNA repair and cell cycle arrest genes in the cells exposed to fractionated<br />

irradiation. There was intense activation of DNA repair pathway-associated genes (DNA-PK, ATM, Rad52,<br />

MLH1 and BRCA1), efficient DNA repair and phospho-p53 was found to be translocated to the nucleus of<br />

A549 cells exposed to fractionated irradiation. MCF-7 cells responded differently in fractionated regimen.<br />

Silencing of the Rad52 gene in fractionated group of A549 cells made the cells radiosensitive. The above<br />

result indicated increased radioresistance in A549 cells due to the activation of Rad52 gene.<br />

© 2011 Elsevier B.V. All rights reserved.<br />

1. Introduction<br />

Cellular response to ionizing radiation is a complex phenomenon<br />

where the dose, dose rate and fractionation play an<br />

equally important role in deciding the fate of the cell. Chronic exposure<br />

of cells to ionizing radiation induces an adaptive response<br />

that results in the increased tolerance to subsequent cytotoxicity<br />

caused by the same [1–4]. Evidence from a number of studies has<br />

indicated that a potential cause of radiation treatment failure may<br />

be that multifraction irradiation selects a population of radiation<br />

resistant cells from which regrowth of tumor progresses [5–7]. A<br />

better understanding of how resistance is promoted at the molecular<br />

level can form the basis for novel treatment strategies in the<br />

future. The targets have been chosen based on their activation following<br />

a single dose of radiation. However, more information about<br />

the response of these signaling factors at clinically relevant doses<br />

following a fractionated regimen is needed to understand their role,<br />

if any, in the development of a radioresistant phenotype.<br />

Mammalian cells have evolved a number of repair systems to<br />

deal with the DNA double-strand breaks (DSBs) induced by exposure<br />

to ionizing radiation [8]. The two major types of double-strand<br />

break repair are homologous recombination and nonhomologous<br />

∗ Corresponding author. Tel.: +91 22 2559 0416; fax: +91 22 2550 5151.<br />

E-mail addresses: ghosh.barc@gmail.com (S. <strong>Ghosh</strong>), malinik00@gmail.com<br />

(M. Krishna).<br />

end-joining. The relative contribution of each of these types of<br />

repair is controversial [9–11].<br />

The phosphatidyl-inositol kinase-related protein ATM is the<br />

most proximal signal transducer initiating cell cycle changes after<br />

the DNA damage induced by ionizing radiation [12]. Likewise,<br />

the rapid induction of ATM serine/threonine protein kinase activity<br />

after ionizing radiation has also suggested that ATM acts at<br />

an early stage of signal transduction in mammalian cells [13,14].<br />

Mammalian ATM is a member of a family of protein kinases that<br />

include ATM-Rad3-related (ATR), DNA-dependent protein kinase,<br />

and FRAP, which are related because they have a similar carboxyterminal<br />

kinase domain [15,16]. Recently, Bakkenist and Kastan<br />

[17] showed that, ATM activation may result from changes in the<br />

structure of chromatin brought about by intermolecular autophosphorylation<br />

and ATM dimer dissociation. Once dissociated, ATM<br />

can then potentially phosphorylate numerous downstream targets,<br />

including p53, MDM2, CHK2, NBS1, RAD9, and BRCA1. ATM acts<br />

specifically in the cellular responses to ionizing radiation and DNA<br />

DSBs, rather than having a more general function in DNA repair.<br />

It resides in a complex with BRCA1 and phosphorylates BRCA1<br />

in a region that contains clusters of serine-glutamine residues<br />

[18]. Phosphorylation of this domain appears to be functionally<br />

important because a mutated BRCA1 protein that lacks these key<br />

phosphorylation sites is unable to rescue the radiation hypersensitivity<br />

of BRCA1-deficient cell lines [18]. Cells deficient in BRCA1<br />

show genetic instability, defective G2/M checkpoint control, and<br />

reduced homologous recombination [19,20]<br />

0027-5107/$ – see front matter © 2011 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.mrfmmm.2011.09.007


62 S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72<br />

Repair of DNA damage is critical for the maintenance of genome<br />

integrity and cell survival. Living organisms have developed different<br />

pathways of DNA repair to deal with various types of<br />

DNA damage. In Saccharomyces cerevisiae, three epistasis groups<br />

of DNA-damage-repair genes have been identified [21,22]. The<br />

Rad52 epistasis group, which is mainly responsible for doublestrand<br />

break (DSB) repair contains several genes: Rad50–Rad57,<br />

MRE11, and XRS2. Among these genes, mutations in Rad51, Rad52,<br />

and Rad54 cause the most severe and pleiotropic defects [23–25].<br />

Yeast strains lacking a functional Rad52 gene are extremely X-<br />

ray sensitive and deficient in mitotic and meiotic recombination<br />

[26]. Mutations in different regions of Rad52 often result in different<br />

phenotypes [27]. Shinohara et al. proposed that the product of<br />

the Rad52 gene is not required for the initiation of recombination,<br />

but is essential for an intermediate stage following the formation<br />

of DSBs but before the appearance of stable recombinants<br />

[28]. Recently, homologs of the Rad52 gene were found in several<br />

eukaryotic organisms. Sequence analysis has revealed that the N-<br />

terminal amino acid sequence of Rad52 protein is highly conserved<br />

while the C-terminal region is less conserved [29–32]. Rad52 protein<br />

interacts through its C-terminal domain with the N-terminal<br />

domain of Rad51 protein in a species specific manner [33–35]. It<br />

was reported that the over expression of human Rad52 HsRad52<br />

conferred enhanced resistance to -rays and induced homologous<br />

intrachromosomal recombination in cultured monkey cells [36]<br />

MLH1, another protein involved in Mismatch repair (MMR), has<br />

also been suggested to be involved in the DSB repair pathway [37].<br />

MLH1 modulates error prone NHEJ by inhibiting the annealing of<br />

DNA ends containing noncomplementary base pairs [38]. Furthermore,<br />

the influence of MLH1 on IR-induced autosomal mutations<br />

has been studied in a mouse Mlh1−/− kidney cell line. A high frequency<br />

of IR induced mutations was found in Mlh1−/− cells due<br />

to increased crossover events of mitotic recombination, suggesting<br />

that MLH1, or MMR may serve as a modulator in mitotic HR<br />

repair [39]. Previous study showed that the expression of MLH1 was<br />

induced by IR, and the loss of MLH1 expression not only elevated<br />

cell cycle progression but also increased the yield of IR-induced<br />

chromosomal translocations, suggesting that this gene, involved<br />

mostly in MMR, may play a role in DSB repair and cell cycle regulation<br />

[40]<br />

The present study was conducted to look at how repair pathways<br />

respond to different irradiation regimens with aim to identify<br />

markers of radioresistance which may serve as future targets for<br />

modulation to enhance the efficacy of radiotherapy.<br />

2. Materials and methods<br />

2.1. Cell culture<br />

Human Lung Adenocarcinoma A549 cells and human breast carcinoma MCF-7<br />

cells were grown in DMEM (Sigma, USA), supplemented with 10% fetal calf serum<br />

(Sigma). Cells were maintained at 37 ◦ C in humidified atmosphere with 5% CO 2.<br />

2.2. Irradiation of cells<br />

Before designing the experiment, there were two aspects that had to be taken<br />

care of. One the stochastic effects of radiation and second the deterministic effects of<br />

radiation. Radiation dose as well as confluency of cells could influence the stochastic<br />

or deterministic effects of radiation. Exposure of cells (≥2 Gy) with confluency more<br />

than 80%, deterministic effects will be more and stochastic effects will be minimal.<br />

At this confluency of cells and radiation dose, it is expected that all the cells will<br />

get threshold number of hits. Deterministic rather than stochastic factors explain<br />

most of the variation in radio-responsiveness following fractionated irradiation of<br />

2 Gy per fraction [41]. A549 and MCF-7 cells with 80% confluency were exposed to<br />

different doses of irradiation using Gamma Cell 220 irradiator (Atomic Energy of<br />

Canada Ltd.), at a dose rate of 3.5 Gy/min; the first set was irradiated with acute<br />

dose of 2 Gy or 10 Gy. The second set was exposed to 5 fraction of 2 Gy each over a<br />

period of five days, after irradiation cells were trypsinized and seeded, so that cells<br />

will reach 80% confluency by next day.<br />

2.3. Clonogenic cell survival assay<br />

After irradiation, clonogenic assay was done as described earlier [42]. Absolute<br />

plating efficiency of A549 cells at 0 Gy (sham irradiated control) was 32.3 ± 2.4%.<br />

Absolute plating efficiency of MCF-7 cells at 0 Gy (sham irradiated control) was<br />

45.2 ± 2.4%.<br />

2.4. Immunofluorescence staining<br />

Immunofluorescence staining was done as described earlier [42]. The primary<br />

antibodies used were mouse anti-pATM (S1981), rabbit anti--H2AX (S139), rabbit<br />

anti-pBRCA1 (Ser1524) and mouse anti-phospho-p53 (S15) (Cell Signaling). Secondary<br />

antibodies used were Alexa Fluor 488 goat anti-rabbit IgG and Alexa Fluor<br />

488 goat anti-mouse IgG (Molecular Probe, USA).<br />

2.5. Image analysis using ImageJ software<br />

The captured images were analyzed for relative quantification of phosphorylation<br />

using ImageJ software [43]. DAPI stained nucleus of each cell was selected and<br />

the same frame was used for relative quantification of phosphorylation using ImageJ<br />

software as described earlier [42]. All the foci were counted manually; at least 100<br />

cells per experiment were analyzed from three independent experiments.<br />

2.6. Western Blotting<br />

Immunoblotting was done as described previously [4]. Primary antibody used<br />

was mouse anti-phospho-p53 (S15) (Cell Signaling). The bound HRP labeled secondary<br />

antibody was then detected using BM Chemiluminiscence Western Blotting<br />

Kit (Roche Molecular Bio chemicals, Germany). The band intensity was quantified by<br />

Image J software. The total intensity of bands obtained in ponceau staining was used<br />

as a protein loading and transfer control. All other band intensities were divided by<br />

the intensity of their respective ponceau blot.<br />

2.7. Microarray<br />

For microarray analysis, 4 h after irradiation was selected as the time point<br />

to monitor the early response of cells to irradiation and to identify differentially<br />

expressed early genes that mediate cellular events such as DNA repair and apoptosis.<br />

Two independent series of experiments were performed with Human Genome<br />

U-133 Plus 2.0 microarrays containing 54,675 probe sets (Affymetrix, Santa Clara,<br />

CA). Microarray analysis was carried out at Oscimum Biosolution, Hyderabad on paid<br />

basis. Total RNA was isolated from ∼3 × 10 6 A549 cells with RNeasy Mini Kits (Qiagen)<br />

including digestion with RNase-free DNase I, and its quantity and integrity were<br />

checked spectrophotometrically and by electrophoresis in 1% agarose gels. Materials<br />

and methods for microarrays were from Affymetrix (Santa Clara); double-stranded<br />

cDNA was prepared with the GeneChip Expression 3 ′ -Amplification One-Cycle cDNA<br />

Synthesis Kit, cleaned using the Sample Cleanup Module, and biotinylated cRNA<br />

was synthesized with GeneChip Expression 3 ′ -Amplification Reagents for IVT Labeling,<br />

cleaned on RNA Sample Cleanup columns, and fragmented at 94 ◦ C for 35 min<br />

in Fragmentation Buffer. The biotinylated cRNA was hybridized first to a control<br />

Test3 microarray to evaluate its quality and then to a Human Genome U-133<br />

Plus 2.0 microarrays containing 54,675 probe sets, using GeneChip Expression 3 ′ -<br />

Amplification Reagents. Microarrays were stained with streptavidin–phycoerythrin<br />

conjugate and scanned in a Gene Array G2500A scanner (Agilent) and the expression<br />

data on human Affymetrix chip was generated. Since the numbers of samples<br />

in each treatment type were sparse, we used a bivariate simulation approach to<br />

identify differentially expressed (DE) genes for the specified threshold fold change<br />

(FC). The DE candidates across each comparison were further evaluated for their<br />

biological relevance through Gene Ontology and Pathways studies. The statistical<br />

analysis was carried out using R package and the biological analysis was carried out<br />

using Genowiz TM software.<br />

For comparative evaluation, an un-irradiated control cell was used. The expression<br />

data on Affymetrix Human 133AB chip was obtained for each sample. The<br />

primary objective of the study was to obtain differentially expressed (DE) probe<br />

sets (genes) across various comparisons. Also, the interest was to study the clustering<br />

of genes and samples and the functional relevance of the differentially expressed<br />

genes.<br />

Upon identifying the DE genes, the next interest was to study the biological<br />

relevance of selected marker genes through GO and Pathway analysis.<br />

The data analysis process involves Robust Multichip Average (RMA) normalization,<br />

quality check analysis, differential expression analysis and the gene enrichment<br />

analysis.<br />

The correlation between the expression values for samples was studied through<br />

Pearson’s coefficient. A pairwise correlation analysis was performed to generate<br />

a correlation matrix indicating how the samples are related with each other. The<br />

coefficient tells about the similarity of expression trend between the two arrays.<br />

A coefficient value close to 1.0 indicates the linear expression between the arrays.<br />

The minimum correlation between the samples was 0.987, which was above the<br />

acceptable criterion of 0.8.


S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72 63<br />

Table 1<br />

The sequences of gene specific primers and their annealing temperature (T m) used<br />

in RT-PCR experiments.<br />

Gene Primer sequence T m ( ◦ C)<br />

-Actin F 5 ′ -TGGAATCCTGTGGCATCCATGAAA-3 ′ 55<br />

R 5 ′ -TAAAACGCAGCTCAGTAACAGTCCG-3 ′<br />

ATM F 5 ′ -GGACAGTGGAGGCACAAAAT-3 ′ 57<br />

R 5 ′ -GTGTCGAAGACAGCTGGTGA-3 ′<br />

DNA-PK F 5 ′ -TGCCAATCCAGCAGTCATTA-3 ′ 64<br />

R 5 ′ -GGTCCATCAGGCACTTCACT-3 ′<br />

Rad52 F 5 ′ -AGTTTTGGGAATGCACTTGG-3 ′ 50<br />

R 5 ′ -TCGGCAGCTGTTGTATCTTG-3 ′<br />

MLH1 F 5 ′ -GAGGTGAATTGGGACGAAGA-3 ′ 52<br />

R 5 ′ -TCCAGGAGTTTGGAATGGAG-3 ′<br />

CDKN1A (p21) F 5 ′ -CAGCAGAGGAAGACCATGTG-3 ′ 59<br />

R 5 ′ -GGCGTTTGGAGTGGTAGAAA-3 ′<br />

GADD45 F 5 ′ -TGCGAGAACGACATCAACAT-3 ′ 58<br />

R 5 ′ -TCCCGGCAAAAACAAATAAG-3 ′<br />

TLR3 F 5 ′ -GCCTCTTCGTAACTTGACCA-3 ′ 57<br />

R 5 ′ -AAGGATGTGGAGGTGAGACA-3 ′<br />

FDXR F 5 ′ -AGAGAACGGACATCACGAAG-3 ′ 57<br />

R 5 ′ -GTCCTGGAGACCCAAGAAAT-3 ′<br />

F, forward and R, reverse.<br />

Percentage Survival<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0Gy 2Gy 5X2Gy 10Gy<br />

Radiation Dose<br />

Fig. 1. Clonogenic cell survival of A549 cells irradiated with 2 Gy, 5 fractions of 2 Gy<br />

or 10 Gy -radiation. Data represent means ± SE of three independent experiments;<br />

*P < 0.05, **P < 0.01 compared with 10 Gy acute dose.<br />

**<br />

2.8. Semiquantitative reverse transcriptional polymerase chain reaction (RT-PCR)<br />

Total RNA from 1 × 10 6 cells was extracted 4 h after irradiation and RT-PCR was<br />

carried out as described earlier [44]. Primer sequence and annealing temperature of<br />

specific primers are given in Table 1.<br />

2.9. Transfection with ShRNA<br />

MLH1, Rad52 and Control shRNA plasmids were procured from Santa cruz<br />

biotechnology (Cat No. sc-35943-SH, sc-37399-SH and sc-108060 respectively). In<br />

a six well tissue culture plate, A549 cells were grown to 50–70% confluency in<br />

antibiotic-free normal growth medium supplemented with FBS and transfection<br />

was carried out as per supplier’s instruction. 48 h post-transfection, medium was<br />

aspirated and replaced with fresh medium containing puromycin (2 g/ml) for<br />

selection of stably transfected cells. Semi-quantitative RT-PCR was performed to<br />

monitor MLH1 and Rad52 gene expression knockdown using gene specific primers<br />

(Table 1).<br />

2.10. Statistical analysis<br />

The data were imported to excel work sheets, and graphs were made using<br />

Origin version 5.0. One-way ANOVA with Tukey–Kramer Multiple Comparisons as<br />

post-test for P < 0.05 used to study the significant level. Data were insignificant at<br />

P > 0.05. Each point represented as mean ± SE (standard error of the mean).<br />

3. Results<br />

3.1. Clonogenic cell survival<br />

To determine the sensitivity of cells to different doses of radiation,<br />

their ability to form colonies after irradiation was observed.<br />

There was 60 ± 2.5% survival of A549 cells that had been exposed<br />

to 2 Gy acute dose where as there was 48 ± 2.8% of cells survived<br />

that had been exposed to 5 fractions of 2 Gy radiations and 4 ± 0.5%<br />

of cells survived that had been exposed to 10 Gy acute dose (Fig. 1).<br />

3.2. Human genome U-133 Plus 2.0 microarray data analysis<br />

Global expression analysis of 54,675 probe set using Human<br />

Genome U-133 Plus 2.0 microarray was performed in A549 cells.<br />

On comparison of control vs 5X2 Gy (henceforth called fractionated<br />

dose), 461 genes were identified as differentially expressed;<br />

312 genes up-regulated (defined as a 2.0-fold or greater increase,<br />

Supplementary data) and 149 down-regulated (defined as a 2.0-<br />

fold or greater decrease, Supplementary data). Up-regulated genes<br />

were associated with p53 signaling pathway, cytokine–cytokine<br />

receptor interaction, hematopoietic cell lineage, Toll-like receptor<br />

signaling pathway, Jak-STAT signaling pathway, MAPK signaling<br />

pathway, cell-cycle check point, B cell receptor signaling pathway.<br />

Down-regulated genes were associated with TGF-beta signaling<br />

pathway and tyrosine metabolism.<br />

In the second group (2 Gy vs 5X2 Gy), 234 genes were identified<br />

as differentially expressed; 172 genes as up-regulated<br />

(Supplementary data) and 62 as down-regulated (Supplementary<br />

data). Genes involved in cytokine–cytokine receptor interaction,<br />

toll-like receptor signaling pathway, hematopoietic cell lineage,<br />

Jak-STAT signaling pathway, MAPK signaling pathway were upregulated.<br />

Gene involved in folate biosynthesis, glycine, serine and<br />

threonine metabolisms were down-regulated.<br />

In the 10 Gy vs fractionated group, 289 genes were identified<br />

as differentially expressed; 211 genes as up-regulated<br />

(Supplementary data) and 78 as down-regulated (Supplementary<br />

data). Genes involved in cytokine–cytokine receptor interaction,<br />

toll-like receptor signaling pathway, cell cycle, DNA replication,<br />

mismatch repair, homologous recombination were up-regulated.<br />

Genes involved in regulation of autophagy, base excision repair,<br />

folate biosynthesis were down-regulated.<br />

Cluster analysis of the expression of these genes showed that<br />

the fractionated group differed most in gene expression compared<br />

to the other three groups (control, 2 Gy or 10 Gy) (Supplementary<br />

data, TCS2 means 2 Gy treatment group, TCS3 means fractionated<br />

group, TCS4 means 10 Gy treatment group).<br />

3.3. Validation of microarray data<br />

In order to verify the microarray data, semi-quantitative RT-PCR<br />

was performed for four genes from control and fractionated group<br />

of A549 cells. These results corresponded very well with those for<br />

the microarray data for all four genes, strongly supporting the reliability<br />

and rationale of our strategy (Fig. 2A and B).<br />

3.4. Activation of repair related genes<br />

When compared to the controls there was a significant upregulation<br />

of ATM, DNA-PK, Rad52 and MLH1 genes in A549 cells<br />

that had been exposed to fractionated irradiation (Fig. 3A–D, Lane<br />

3). However, in A549 cells that had been exposed to 10 Gy acute<br />

dose there was marginal up-regulation of ATM, DNA-PK and Rad52<br />

genes although it was significantly lower than the fractionated<br />

group (Fig. 3A–D, Lane 4). In A549 cells which had been exposed<br />

to 2 Gy acute dose there was marginal up-regulation of Rad52 and


64 S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72<br />

(A)<br />

rray)<br />

Fold Change (Microar<br />

(B)<br />

Fold Change (RT-PC CR)<br />

3.6<br />

3.4<br />

Control<br />

3.2<br />

5X2Gy<br />

3.0<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

CDK1A(p21) GADD45A TLR3 FDXR<br />

Genes<br />

3.6<br />

3.4<br />

3.2<br />

Control<br />

3.0<br />

5X2Gy<br />

2.8<br />

2.6<br />

2.4<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

CDK1A(p21) GADD45A TLR3 FDXR<br />

Genes<br />

Fig. 2. Validation of microarray data by semi-quantitative RT-PCR. (A) Fold changes<br />

of genes as determined by microarray analysis. (B) Fold changes of genes as determined<br />

by RT-PCR. Four genes from control and 5X2 Gy treatment group of A549 cells<br />

were selected and performed three sets of semi-quantitative RT-PCR as described<br />

in Section 2. Four genes were CDKN1A (p21), GADD45, Toll-like receptor 3 (TLR3)<br />

and Ferredoxin reductase (FDXR).<br />

MLH1 genes but these were also significantly lower than the fractionated<br />

group (Fig. 3A–D, Lane 2).<br />

3.5. Efficient repair of DNA<br />

Repair kinetics was determined by counting the -H2AX foci<br />

at 15 min and 4 h after irradiation. As expected maximum numbers<br />

of foci were seen at 15 min in A549 cells exposed to 10 Gy<br />

followed by cells exposed to fractionated irradiation and very few<br />

foci were seen in cells exposed to 2 Gy acute dose. By 4 h the number<br />

of foci in cells exposed to fractionated irradiation had been reduced<br />

to almost control but in the cells exposed to 10 Gy the number of<br />

foci was still very high, indicating that there was efficient repair of<br />

DNA in A549 cells that had been exposed to fractionated irradiation<br />

(Fig. 4).<br />

3.6. Radiation induced foci of DNA damage response proteins<br />

ATM and BRCA1<br />

ATM, an important DNA damage response protein to ionizing<br />

radiation and a part of irradiation induced foci (IRIF), is mainly<br />

involved in cell cycle arrest and repair. It gets activated after DNA<br />

DSB by phosphorylation at ser 1981 and activates many key cellular<br />

proteins that include H2AX, BRCA1, Chk1/2 and p53. Phospho<br />

ATM foci and intensity were looked at 4 h after irradiation. All the<br />

cells either exposed to 10 Gy (23 ± 3 foci per cell) or fractionated<br />

irradiation (25 ± 2.8 foci per cell) showed ATM foci but only 10%<br />

of cells exposed to 2 Gy showed ATM foci and the number of foci<br />

per cell was 2 ± 0.08 (Fig. 5A). The intensity of phosphorylation of<br />

ATM was quantified by ImageJ software, was significantly higher in<br />

the fractionated group as compared to any of the treatment groups<br />

(control, 2 Gy or 10 Gy) (Supplementary data, Fig. 5C).<br />

BRCA1, which is phosphorylated by ATM, is a part of IRIF and<br />

shares many downstream substrates of ATM. 4 h after irradiation,<br />

A549 cells that had been exposed to 2 Gy acute dose did not show<br />

any BRCA1 foci where as 55% of cells that had been exposed to fractionated<br />

irradiation showed BRCA1 foci and the number of foci per<br />

cell was 24 ± 4. 22% of cells that had been exposed to 10 Gy acute<br />

showed BRCA1 foci and the number of foci per cell was 15 ± 2.1<br />

(Fig. 5B). The intensity of phosphorylation of BRCA1 as determined<br />

by ImageJ software was significantly higher in the fractionated<br />

group as compared to any of the treatment groups (control, 2 Gy<br />

or 10 Gy) (Supplementary data, Fig. 5D).<br />

3.7. Translocation of phospho-p53<br />

18 h after irradiation, phospho-p53 was found to be translocated<br />

into the nucleus of A549 cells exposed to fractionated irradiation<br />

(Fig. 6A). The intensity of phosphorylation was significantly higher<br />

in the cells exposed to fractionated irradiation (Supplementary<br />

data, Fig. 6C). Phosphorylation of p53 was further confirmed by<br />

Western Blot (Fig. 6B, Lane 3).<br />

3.8. Response of A549 and MCF-7 cells exposed to fractionated<br />

irradiation<br />

There was 48 ± 2.8% survival of A549 cells exposed to fractionated<br />

irradiation where as only 10 ± 1.2% of MCF-7 cells survived<br />

exposed to fractionated irradiation (Fig. 7A). 4 h after irradiation,<br />

there was a significant up-regulation of ATM, DNA-PK, Rad52 and<br />

MLH1 genes in A549 cells but not in MCF-7 cells (Fig. 7B, Lane<br />

3). ATM and BRCA1 phosphorylation was found to be significantly<br />

higher 4 h after irradiation in A549 cells but not in MCF-7 cells<br />

(Fig. 7C). 18 h after irradiation, phospho-p53 was found to be<br />

translocated into the nucleus of A549 cells but not in MCF-7 cells<br />

(Fig. 7C). There was efficient repair of DNA by 4 h in A549 cells as<br />

determined by counting -H2AX but not in MCF-7 cells (Fig. 7D).<br />

3.9. Role of Rad52 and MLH1<br />

Since there was an intense activation of Rad52 and MLH1 in<br />

A549 cells exposed to fractionated irradiation, it was of interest to<br />

look for their role in survival of A549 cells. MLH1 and Rad52 gene<br />

expression knockdown was carried out in A549 cells using MLH1<br />

and Rad52 shRNA plasmids. Stably transfected cells were selected<br />

and the transfection efficiency was found to be 85% for MLH1 gene<br />

and 80% for Rad52 gene (Fig. 8A).<br />

There was 48 ± 2.8% survival of A549 cells exposed to fractionated<br />

irradiation. The survival of A549 cells that had been transfected<br />

with MLH1 shRNA plasmid and exposed to fractionated irradiation<br />

was found to be 40 ± 1.8% where as the survival of A549 cells that<br />

had been transfected with Rad52 shRNA plasmid and exposed to<br />

fractionated irradiation was found to be 23 ± 1.5% (Fig. 8B and C).<br />

The survival of cells transfected with control shRNA plasmid was<br />

the same as unirradiated control cells (data not shown).<br />

4. Discussion<br />

Chronic exposure of cells to IR has been reported to induce<br />

an adaptive response that results in an enhanced tolerance to the<br />

cytotoxicity of subsequent doses of IR [1–4]. There could be many<br />

possible mechanisms for the induction of the adaptive response<br />

which would eventually determine the cell fate. In the present


S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72 65<br />

Fig. 3. Gene expression of ATM, DNA-PK, Rad52 and MLH1 in A549 cells irradiated with different doses of -radiation. Total RNA from A549 cells was isolated and reverse<br />

transcribed 4 h after irradiation. RT-PCR analysis of ATM, DNA-PK, Rad52 and MLH1 genes was carried out as described in Section 2. PCR products were resolved on 1.5%<br />

agarose gels containing ethidium bromide. -Actin gene expression in each group was used as an internal control. Ratio of intensities of (A) ATM, (B) DNA-PK, (C) Rad52 and<br />

(D) MLH1 band to that of respective -actin band as quantified from gel pictures are shown above each gel picture. Key: Lane 1, control unirradiated A549 cell line; Lane 2,<br />

2 Gy acute dose (single dose that was delivered in each fraction) of 60 Co -irradiation; Lane 3, 2 Gy of 60 Co -irradiation daily over a period of 5 days; Lane 4, 10 Gy acute<br />

(which is the sum total of all the fraction) 60 Co -irradiation. Data represent means ± SE of three independent experiments; significantly different from unirradiated controls:<br />

*P < 0.05, **P < 0.01.<br />

study human lung adenocarcinoma A549 cells were found to be<br />

more radioresistant if 10 Gy dose was delivered as 5 fractions of<br />

2 Gy dose. Although fractionation is supposed to play an important<br />

role in inducing an adaptive response and deciding the fate of cells,<br />

the mechanism by which it does so can be many and innumerable<br />

factors could be responsible. DNA microarray analysis is a powerful<br />

tool for obtaining comprehensive information concerning expression<br />

of thousands of genes in cancer cells [45] DNA microarray<br />

analysis revealed variable expression of many genes. Up-regulated<br />

genes in A549 cells exposed to fractionated irradiation were associated<br />

with p53 signaling pathway, cytokine-cytokine receptor<br />

interaction, hematopoietic cell lineage, toll-like receptor signaling<br />

pathway, Jak-STAT signaling pathway, MAPK signaling pathway,<br />

cell-cycle check point, B cell receptor signaling pathway, DNA<br />

replication, mismatch repair, homologous recombination. Toll-like<br />

receptor signaling pathway is responsible for survival and proliferation<br />

of the cells [46]. The Janus kinase-signal transducer and<br />

activator of transcription (JAK-STAT) pathway mediate signaling<br />

by cytokines, which control survival, proliferation and differentiation<br />

of several cell types [47]. MAPK signaling pathways are<br />

known to be involved in various processes of growth, differentiation<br />

and cell death [48]. Most of the pathways which were<br />

up-regulated in A549 cells exposed to fractionated irradiation in<br />

all comparisons were associated with the survival or repair of<br />

the A549 cells. This indicates that fractionated irradiation could<br />

induce up-regulation of survival/repair related pathways in cancer<br />

cells.<br />

To the authors’ knowledge, this is the first report indicating<br />

the differential gene expression profiles of A549 cells<br />

exposed to fractionated irradiation. Previous reports concerning


66 S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72<br />

Fig. 4. Repair kinetics as determined by -H2AX foci in different treatment group of A549 cells at 15 min and 4 h after irradiation. Cells were fixed 15 min and 4 h after<br />

irradiation in 4% paraformaldehyde, permeabilized in 0.25% Triton X-100 and labeled with specific antibodies as described in Section 2. Each phospho-site antibody was<br />

indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI (blue). The encircled area roughly<br />

represents cell nuclei as seen by DAPI staining. All images were captured using Carl Zeiss confocal microscope with the same exposure time. Representative image of three<br />

independent experiments is shown here; at least 100 cells per experiment were analyzed from three independent experiments. (For interpretation of the references to color<br />

in this figure legend, the reader is referred to the web version of the article.)<br />

radioresistant non-small-cell lung cancer and HeLa cells<br />

also indicated that cell cycle signaling pathways, mismatch<br />

repair, homologous recombination were up-regulated<br />

[49,50]. In the current study, some of these genes were<br />

previously known to be associated with responsiveness<br />

to radiation, such as p21 and GADD45˛, but others were<br />

novel. These novel genes can be expected to be involved in<br />

radioresistance, but the precise function of each gene remains<br />

unclear; so further study is necessary to clarify the nature of<br />

associations.<br />

Microarray data had shown that p53 signaling pathway was<br />

up-regulated in A549 cells that had been exposed to fractionated


S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72 67<br />

Fig. 5. Radiation induced foci of DNA damage response proteins. (A) ATM and (B) BRCA1 in different treatment groups of A549 cells 4 h after irradiation. Cells were fixed in 4%<br />

paraformaldehyde 4 h after irradiation, permeabilized in 0.25% Triton X-100 and labeled with specific antibodies as described in Section 2. Each phospho-site antibody was<br />

indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI (blue). All images were captured using<br />

Carl Zeiss confocal microscope with the same exposure time. Representative images of three independent experiments are shown here; at least 100 cells per experiment<br />

were analyzed from three independent experiments. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of the article.)


68 S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72<br />

Fig. 6. Translocation of phospho-p53 into the nucleus of A549 cells 18 h after irradiation. (A) Representative image of three independent experiments showing phosphorylation<br />

and translocation of p53 into nucleus of A549 cells. Each phospho-site antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were<br />

mounted with ProLong Gold antifede with DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. At least 100 cells per<br />

experiment were analyzed from three independent experiments. (B) Representative immunoblot image of three independent experiments depicting phosphorylation of p53<br />

in different treatment group. Graph represents phospho-p53 band intensities (divided with the respective loading control intensities and normalized). Key: Lane 1, control<br />

unirradiated A549 cell line; Lane 2, 2 Gy acute dose (single dose that was delivered in each fraction) of 60 Co -irradiation; Lane 3, 2 Gy of 60 Co -irradiation daily over a period<br />

of 5 days; Lane 4, 10 Gy acute (which is the sum total of all the fraction) 60 Co -irradiation. Data represent means ± SE of three independent experiments. (For interpretation<br />

of the references to color in this figure legend, the reader is referred to the web version of the article.)


S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72 69<br />

Fig. 7. Response of A549 and MCF-7 cells exposed to fractionated irradiation. (A) Clonogenic cell survival of A549 and MCF-7 cells exposed to fractionated irradiation.<br />

Data represent means ± SE of three independent experiments; **P < 0.01 compared with MCF-7. (B) Representative image of three independent experiments showing gene<br />

expression of ATM, DNA-PK, Rad52 and MLH1 in A549 and MCF-7 cells exposed to fractionated irradiation. -Actin gene expression in each group was used as an internal<br />

control. (C) Phosphorylation of ATM, BRCA1 and p53 in A549 and MCF-7 cells exposed to fractionated irradiation. (D) Repair kinetics as determined by -H2AX foci in A549<br />

and MCF-7 cells exposed to fractionated irradiation at 15 min and 4 h after irradiation. Each phospho-site antibody was indirectly labeled with Molecular Probe 488 secondary<br />

antibody (green) and cells were mounted with ProLong Gold antifede with DAPI (blue). The encircled area roughly represents cell nuclei as seen by DAPI staining. All the<br />

images were taken using a 100× objective lens except the images of p53 phosphorylation. The images of p53 phosphorylation were taken using 40× objective lens for A549<br />

cells and MCF-7 cells. All images were captured using Carl Zeiss confocal microscope with the same exposure time. Representative images of three independent experiments<br />

are shown here; at least 100 cells per experiment were analyzed from three independent experiments. (For interpretation of the references to color in this figure legend, the<br />

reader is referred to the web version of the article.)


70 S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72<br />

Fig. 8. Effect of inhibition of Rad52 and MLH1 on the growth and radioresistance of<br />

A549 cells which had been exposed to 5 fractions of 2 Gy radiation. A549 cells were<br />

transfected with either MLH1 or Rad52 ShRNA as described in Section 2. (A) Transfection<br />

efficiency as determined by semi-quantitative RT-PCR. Gene expression of<br />

Rad52 and MLH1 in A549 cells transfected with Control ShRNA, Rad52 ShRNA or<br />

MLH1 ShRNA. Total RNA from A549 cells was isolated and reverse transcribed 7<br />

days after transfection. RT-PCR analysis of Rad52 and MLH1 genes were carried out<br />

as described in Section 2. PCR products were resolved on 1.5% agarose gels containing<br />

ethidium bromide. -Actin gene expression in each group was used as an internal<br />

control. Ratio of intensities of Rad52 or MLH1 band to that of respective -actin band<br />

as quantified from gel pictures are shown above each gel picture. Key: Rad52, Gene<br />

expression of Rad52 to that of -actin in A549 cells transfected with control ShRNA;<br />

−ve Rad52, gene expression of Rad52 to that of -actin in A549 cells transfected<br />

with Rad52 ShRNA; MLH1, gene expression of MHL1 to that of -actin in A549 cells<br />

transfected with control ShRNA; −ve MLH1, gene expression of MLH1 to that of -<br />

actin in A549 cells transfected with MLH1 ShRNA. (B) Clonogenic survival assay of<br />

A549 cells. Key: control, control unirradiated A549 cells; 5X2 Gy, A549 cells exposed<br />

to 2 Gy of 60Co -irradiation daily over a period of 5 days; 5X2 Gy + M, A549 cells<br />

transfected with MLH1 ShRNA and exposed to 2 Gy of 60 Co -irradiation daily over a<br />

period of 5 days; 5X2 Gy + R, A549 cells transfected with Rad52 ShRNA and exposed<br />

to 2 Gy of 60 Co -irradiation daily over a period of 5 days. (C) Representative image<br />

of three independent experiments. Data represent means ± SE of three independent<br />

experiments; *P < 0.05, **P < 0.01 compared with 5X2 Gy treatment group.<br />

irradiation; it was our interest to look at the phosphorylation of<br />

p53 and its translocation to the nucleus of A549 cells. Lung adenocarcinoma<br />

A549 cells typically express wild-type p53 protein<br />

[51] and would be expected to be sensitive to the DNA-damaging<br />

agents used for cancer therapy; however, these cells are radioresistant<br />

if the radiation dose is delivered as fractionated irradiation.<br />

p53 classically considered to be a tumor suppressor protein, but in<br />

this study the results are contradictory. Even though the p53 levels<br />

are up-regulated following fractionated doses (Fig. 6A and B), the<br />

apoptotic function of the p53 appears to be compromised. p53 is<br />

considered to be a classic “gatekeeper” of cellular fate [52,53]. p53<br />

is activated in response to genotoxic stress such as radiation. And<br />

once activated, it initiates cell cycle arrest, senescence or apoptosis<br />

via pathways involving the transactivation of p53 target genes<br />

[52,53]. There was no change in the expression of pro apoptotic<br />

targets of p53 such as Bax, Perp, Puma and Noxa in the cells that<br />

had been exposed to fractionated irradiation as compared to control<br />

cells (Microarray data in Supplementary Files). However, there<br />

was higher expression of cell cycle target genes of p53 such as p21<br />

and GADD45 in the cells that had been exposed to fractionated<br />

irradiation as compared to control cells (Fig. 2).<br />

Since DNA repair pathways were up-regulated in A549 cells that<br />

had been exposed to fractionated irradiation, it was of interest to<br />

look at the activation of genes and proteins involved in DNA repair<br />

pathways. Our results indicate that there was intense activation of<br />

repair genes and protein in A549 cells exposed to fractionated irradiation<br />

(Figs. 3 and 5). Our results on DNA-PK are consistent with<br />

earlier works where authors have also shown the role of DNA-PK in<br />

the radioresistance of lung carcinoma cells [54,55]. However, these<br />

authors have looked at the response only after single dose of irradiation<br />

and it is logical to expect these pathways to get activated.<br />

In this study we reiterate the fact that DNA-PK is also involved in<br />

radioresistance if the cells are subjected to fractionated irradiation<br />

but the reason for the development of radioresistance still remains<br />

an enigma.<br />

ATM and BRCA1 have been regarded as primary regulators of<br />

homologous recombination repair (HRR) [56]. In this study there<br />

was intense activation of ATM gene and ATM foci were seen in all<br />

the cells exposed to fractionated irradiation and most of the cells<br />

showed BRCA1 foci (Figs. 3 and 5) indicating the fact that HRR pathway<br />

was activated in A549 cells and therefore, these proteins could<br />

be involved in HRR which can take advantage of the other strand<br />

that may be intact. ATM and BRCA1 reside in macromolecular complex<br />

and form foci after irradiation. Our results are in agreement<br />

with earlier works where authors have also shown the role ATM<br />

in radioresistance of cancer cells [57,58]. However, these authors<br />

have studied using single dose of irradiation.<br />

By 4 h there was an efficient repair of DNA in A549 cells exposed<br />

to fractionated irradiation but not in cells that had been exposed to<br />

10 Gy acute dose as determined by the disappearance of -H2AX<br />

foci (Fig. 4). It has been established recently that H2AX at the DNA<br />

DSB sites are immediately phosphorylated upon irradiation, and<br />

the phosphorylated H2AX (-H2AX) can be visualized in situ by<br />

immunostaining with a -H2AX specific antibody [59,60]. -H2AX<br />

induction can be measured quantitatively at physiological doses,<br />

and the numbers of residual -H2AX foci can be used to estimate<br />

the kinetics of DSB rejoining. This has become the gold standard for<br />

the detection of DSB [60,61].<br />

A clear cause and effect relationship was established between<br />

DNA damage signaling, gene expression and efficient DNA repair<br />

and was evident in the form of cell survival in A549 cells that had<br />

been exposed to fractionated irradiation.<br />

It has been reported that A549 cells are relatively more<br />

radioresistant than MCF-7 cells if radiation dose was delivered as<br />

fractionated regimen [62]. But the mechanism of such response has<br />

not been studied. To the authors’ knowledge this is the first report


S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72 71<br />

explaining the mechanism of such response as activation of repair<br />

pathway, efficient DNA repair and translocation of phospho-p53<br />

into the nucleus of A549 cells exposed to fractionated irradiation<br />

and not in MCF-7 cells exposed to fractionated irradiation (Fig. 7).<br />

Numerous reports have implicated DNA repair genes as being<br />

mainly responsible for the development of radioresistance due to<br />

fractionated irradiation [54,55,57,58]. To investigate what factors<br />

are involved in fractionated irradiation induced radioresistance in<br />

A549 cells, the cells were transfected with either Rad52 or MLH1<br />

shRNA plasmid and exposed to 5 fractions of 2 Gy. The survival of<br />

A549 cells that had been transfected with MLH1 shRNA plasmid<br />

and exposed to fractionated irradiation was found to be 40 ± 1.8%<br />

where as the survival of A549 cells which had been transfected with<br />

Rad52 shRNA plasmid and exposed to fractionated irradiation was<br />

23 ± 1.5% (Fig. 8). The above results demonstrate that the Rad52<br />

gene is one of the factors responsible for the increased radioresistance<br />

in A549 cells if 10 Gy dose was delivered as fractionated<br />

irradiation. These observations strongly argue for the functional<br />

importance of Rad52 in DSB repair in lung adenocarcinoma A549<br />

cells.<br />

5. Conclusions<br />

In summary, our results indicated that A549 cells were relatively<br />

more radioresistant if these cells were exposed to fractionated irradiation.<br />

The reason for radioresistance could be the activation of<br />

DNA repair pathways and Rad52 gene is an important factor modulating<br />

radioresistance in A549 cells. This study contributes basic<br />

information about some of the genes involved in radioresistance<br />

and may eventually contribute to clinical investigations of prediction<br />

markers and prognostic factors, as well as the development of<br />

radiosensitive molecules for therapeutic use.<br />

Conflict of interest statement<br />

The authors declare that there are no conflicts of interest.<br />

Acknowledgments<br />

Authors would like to thank Mr. Manjoor Ali and Mr. Paresh<br />

Khadilkar, RB&HSD, BARC for their help in Confocal Microscopy and<br />

Lab work respectively.<br />

Appendix A. Supplementary data<br />

Supplementary data associated with this article can be found, in<br />

the online version, at doi:10.1016/j.mrfmmm.2011.09.007.<br />

References<br />

[1] J. Russell, T.E. Wheldon, P. Stanton, A radioresistant variant derived from a<br />

human neuroblastoma cell line is less prone to radiation-induced apoptosis,<br />

Cancer Res. 55 (1995) 4915–4921.<br />

[2] W.K. Dahlberg, E.I. Azzam, Y. Yu, J.B. Little, Response of human tumor cells of<br />

varying radiosensitivity and radiocurability to fractionated irradiation, Cancer<br />

Res. 59 (1999) 5365–5369.<br />

[3] A.G. Pearce, T.M. Segura, A.C. Rintala, N.D. Rintala-Maki, H. Lee, The generation<br />

and characterization of a radiation-resistant model system to study radioresistance<br />

in human breast cancer cells, Radiat. Res. 156 (2001) 739–750.<br />

[4] A.K. Mitra, M. Krishna, Fractionated and acute irradiation induced signaling in<br />

a murine tumor, J. Cell. Biochem. 101 (2007) 745–752.<br />

[5] R.R. Weichselbaum, M.A. Beckett, W. Dahlberg, A. Dritschilo, Heterogeneity of<br />

radiation response of a parent human epidermoid carcinoma cell line and four<br />

clones, Int. J. Radiat. Oncol. Biol. Phys. 14 (1988) 907–912.<br />

[6] M.C. Joiner, Evidence for induced radioresistance from survival and other end<br />

points: an introduction, Radiat. Res. 138 (1994) S5–S8.<br />

[7] S. Henness, M.W. Davey, R.M. Harvie, R.A. Davey, Fractionated irradiation of H69<br />

small-cell lung cancer cells causes stable radiation and drug resistance with<br />

increased MRP1, MRP2, and topoisomerase IIalpha expression, Int. J. Radiat.<br />

Oncol. Biol. Phys. 54 (2002) 895–902.<br />

[8] K. Valerie, L.F. Povirk, Regulation and mechanisms of mammalian doublestrand<br />

break repair, Oncogene 22 (2003) 5792–5812.<br />

[9] R. Kanaar, J.H. Hoeijmakers, D.C. van Gent, Molecular mechanisms of DNA double<br />

strand break repair, Trends Cell Biol. 8 (1998) 483–489.<br />

[10] J.E. Haber, Partners and pathwaysrepairing a double-strand break, Trends<br />

Genet. 16 (2000) 259–264.<br />

[11] R.D. Johnson, M. Jasin, Double-strand-break-induced homologous recombination<br />

in mammalian cells, Biochem. Soc. Trans. 29 (2001) 196–201.<br />

[12] T. Samuel, H.O. Weber, J.O. Funk, Linking DNA damage to cell cycle checkpoints,<br />

Cell Cycle 1 (2002) 162–168.<br />

[13] S. Banin, L. Moyal, S. Shieh, Y. Taya, C.W. Anderson, L. Chessa, N.I. Smorodinsky,<br />

C. Prives, Y. Reiss, Y. Shiloh, Y. Ziv, Enhanced phosphorylation of p53 by ATM<br />

in response to DNA damage, Science 281 (1998) 1674–1677.<br />

[14] C.E. Canman, D.S. Lim, K.A. Cimprich, Y. Taya, K. Tamai, K. Sakaguchi, E. Appella,<br />

M.B. Kastan, J.D. Siliciano, Activation of the ATM kinase by ionizing radiation<br />

and phosphorylation of p53, Science 281 (1998) 1677–1679.<br />

[15] C.H. Westphal, Cell-cycle signaling: Atm displays its many talents, Curr. Biol. 7<br />

(1997) R789–R792.<br />

[16] M.F. Hoekstra, Responses to DNA damage and regulation of cell cycle checkpoints<br />

by the ATM protein kinase family, Curr. Opin. Genet. Dev. 7 (1997)<br />

170–175.<br />

[17] C.J. Bakkenist, M.B. Kastan, DNA damage activates ATM through intermolecular<br />

autophosphorylation and dimer dissociation, Nature 421 (2003) 499–506.<br />

[18] D. Cortez, Y. Wang, J. Qin, S.J. Elledge, Requirement of ATM-dependent phosphorylation<br />

of brca1 in the DNA damage response to double-strand breaks,<br />

Science 286 (1999) 1162–1166.<br />

[19] X. Xu, Z. Weaver, S.P. Linke, C. Li, J. Gotay, X.W. Wang, C.C. Harris, T. Ried, C.X.<br />

Deng, Centrosome amplification and a defective G2-M cell cycle checkpoint<br />

induce genetic instability in BRCA1 exon 11 isoform-deficient cells, Mol. Cell 3<br />

(1999) 389–395.<br />

[20] M.E. Moynahan, J.W. Chiu, B.H. Koller, M. Jasin, Brca1 controls homologydirected<br />

DNA repair, Mol. Cell 4 (1999) 511–518.<br />

[21] E. Friedberg, W. Siede, A.J. Cooper, The Molecular and Cellular Biology of the<br />

Yeast Saccharomyces, Cold Spring Harbor Laboratory Press, Plainview, NY,<br />

1991.<br />

[22] J. Game, Radiation-sensitive Mutants and Repair in Yeast, Springer-Verlag, New<br />

York, 1983.<br />

[23] M. Ajimura, S.H. Leem, H. Ogawa, Identification of new genes required<br />

for meiotic recombination in Saccharomyces cerevisiae, Genetics 133 (1993)<br />

51–66.<br />

[24] E.L. Ivanov, V.G. Korolev, F. Fabre, XRS2, a DNA repair gene of Saccharomyces<br />

cerevisiae, is needed for meiotic recombination, Genetics 132 (1992) 651–664.<br />

[25] T. Petes, R.E. Malone, I.S. Symington, Recombination in Yeast, Cold Spring Harbor<br />

Laboratory Press, New York, 1991.<br />

[26] M.A. Resnick, Genetic control of radiation sensitivity in Saccharomyces cerevisiae,<br />

Genetics 62 (1969) 519–531.<br />

[27] K.L. Boundy-Mills, D.M. Livingston, A Saccharomyces cerevisiae RAD52 allele<br />

expressing a C-terminal truncation protein: activities and intragenic complementation<br />

of missense mutations, Genetics 133 (1993) 39–49.<br />

[28] A. Shinohara, H. Ogawa, T. Ogawa, Rad51 protein involved in repair and recombination<br />

in S. cerevisiae is a RecA-like protein, Cell 69 (1992) 457–470.<br />

[29] O.Y. Bezzubova, H. Schmidt, K. Ostermann, W.D. Heyer, J.M. Buerstedde,<br />

Identification of a chicken RAD52 homologue suggests conservation of the<br />

RAD52 recombination pathway throughout the evolution of higher eukaryotes,<br />

Nucleic Acids Res. 21 (1993) 5945–5949.<br />

[30] D.F. Muris, K. Vreeken, A.M. Carr, B.C. Broughton, A.R. Lehmann, P.H. Lohman, A.<br />

Pastink, Cloning the RAD51 homologue of Schizosaccharomyces pombe, Nucleic<br />

Acids Res. 21 (1993) 4586–4591.<br />

[31] C. Bendixen, I. Sunjevaric, R. Bauchwitz, R. Rothstein, Identification of a mouse<br />

homologue of the Saccharomyces cerevisiae recombination and repair gene,<br />

RAD52, Genomics 23 (1994) 300–303.<br />

[32] Z. Shen, K. Denison, R. Lobb, J.M. Gatewood, D.J. Chen, The human and mouse<br />

homologs of the yeast RAD52 gene: cDNA cloning, sequence analysis, assignment<br />

to human chromosome 12p12.2-p13, and mRNA expression in mouse<br />

tissues, Genomics 25 (1995) 199–206.<br />

[33] G.T Milne, D.T. Weaver, Dominant negative alleles of RAD52 reveal a DNA<br />

repair/recombination complex including Rad51 and Rad52, Genes Dev. 7 (1993)<br />

1755–1765.<br />

[34] J.W. Donovan, G.T. Milne, D.T. Weaver, Homotypic and heterotypic protein<br />

associations control Rad51 function in double-strand break repair, Genes Dev.<br />

8 (1994) 2552–2562.<br />

[35] Z. Shen, K.G. Cloud, D.J. Chen, M.S. Park, Specific interactions between the<br />

human RAD51 and RAD52 proteins, J. Biol. Chem. 271 (1996) 148–152.<br />

[36] M.S. Park, Expression of human RAD52 confers resistance to ionizing radiation<br />

in mammalian cells, J. Biol. Chem. 270 (1995) 15467–15470.<br />

[37] Y. Habraken, O. Jolois, J. Piette, Differential involvement of the<br />

hMRE11/hRAD50/NBS1 complex, BRCA1 and MLH1 in NF-kappaB activation<br />

by camptothecin and X-ray, Oncogene 22 (2003) 6090–6099.<br />

[38] L.A. Bannister, B.C. Waldman, A.S. Waldman, Modulation of error-prone doublestrand<br />

break repair in mammalian chromosomes by DNA mismatch repair<br />

protein Mlh1, DNA Repair (Amst) 3 (2004) 465–474.<br />

[39] Q. Wang, O.N. Ponomareva, M. Lasarev, M.S. Turker, High frequency induction<br />

of mitotic recombination by ionizing radiation in Mlh1 null mouse cells, Mutat.<br />

Res. 594 (2006) 189–198.<br />

[40] Y. Zhang, L.H. Rohde, K. Emami, D. Hammond, R. Casey, S.K. Mehta, A.S. Jeevarajan,<br />

D.L. Pierson, H. Wu, Suppressed expression of non-DSB repair genes inhibits


72 S. <strong>Ghosh</strong>, M. Krishna / Mutation Research 729 (2012) 61–72<br />

gamma-radiation-induced cytogenetic repair and cell cycle arrest, DNA Repair<br />

(Amst) 7 (2008) 1835–1845.<br />

[41] A. Safwat, S.M. Bentzen, I. Turesson, J.H. Hendry, Deterministic rather than<br />

stochastic factors explain most of the variation in the expression of skin telangiectasia<br />

after radiotherapy, Int. J. Radiat. Oncol. Biol. Phys. 52 (2002) 198–204.<br />

[42] S. <strong>Ghosh</strong>, N.N. Bhat, S. Santra, R.G. Thomas, S.K. Gupta, R.K. Choudhury, M.<br />

Krishna, Low energy proton beam induces efficient cell killing in A549 lung<br />

adenocarcinoma cells, Cancer Invest. 28 (2010) 615–622.<br />

[43] W.S. Rasband, J.U.S. Image, <strong>National</strong> Institutes of Health, Bethesda, USA.<br />

http://rsb.info.nih.gov/ij/, 1997–2006.<br />

[44] S. <strong>Ghosh</strong>, D.K. Maurya, M. Krishna, Role of iNOS in bystander signaling between<br />

macrophages and lymphoma cells, Int. J. Radiat. Oncol. Biol. Phys. 72 (2008)<br />

1567–1574.<br />

[45] S.M. Bentzen, Preventing or reducing late side effects of radiation therapy:<br />

radiobiology meets molecular pathology, Nat. Rev. Cancer 6 (2006) 702–713.<br />

[46] F.J. Barrat, R.L. Coffman, Development of TLR inhibitors for the treatment of<br />

autoimmune diseases, Immunol. Rev. 223 (2008) 271–283.<br />

[47] S.N. Constantinescu, M. Girardot, C. Pecquet, Mining for JAK-STAT mutations in<br />

cancer, Trends Biochem. Sci. 33 (2008) 122–131.<br />

[48] M. Krishna, H. Narang, The complexity of mitogen-activated protein kinases<br />

(MAPKs) made simple, Cell. Mol. Life Sci. 65 (2008) 3525–3544.<br />

[49] Y.S. Lee, J.H. Oh, S. Yoon, M.S. Kwon, C.W. Song, K.H. Kim, M.J. Cho, M.L. Mollah,<br />

Y.J. Je, Y.D. Kim, C.D. Kim, J.H. Lee, Differential gene expression profiles<br />

of radioresistant non-small-cell lung cancer cell lines established by fractionated<br />

irradiation: tumor protein p53-inducible protein 3 confers sensitivity to<br />

ionizing radiation, Int. J. Radiat. Oncol. Biol. Phys. 77 (2010) 858–866.<br />

[50] Y. Qing, X.Q. Yang, Z.Y. Zhong, X. Lei, J.Y. Xie, M.X. Li, D.B. Xiang, Z.P. Li, Z.Z. Yang,<br />

G. Wang, D. Wang, Microarray analysis of DNA damage repair gene expression<br />

profiles in cervical cancer cells radioresistant to 252Cf neutron and X-rays, BMC<br />

Cancer 10 (2010) 71.<br />

[51] T.A. Lehman, W.P. Bennett, R.A. Metcalf, J.A. Welsh, J. Ecker, R.V. Modali, S.<br />

Ullrich, J.W. Romano, E. Appella, J.R. Testa, et al., p53 mutations, ras mutations,<br />

and p53-heat shock 70 protein complexes in human lung carcinoma cell lines,<br />

Cancer Res. 51 (1991) 4090–4096.<br />

[52] P. May, E. May, Twenty years of p53 research: structural and functional aspects<br />

of the p53 protein, Oncogene 18 (1999) 7621–7636.<br />

[53] B. Vogelstein, D. Lane, A.J. Levine, Surfing the p53 network, Nature 408 (2000)<br />

307–310.<br />

[54] S.P. Lees-Miller, R. Godbout, D.W. Chan, M. Weinfeld, R.S. Day, G.M. 3rd,<br />

J. Barron, Allalunis-Turner, Absence of p350 subunit of DNA-activated protein<br />

kinase from a radiosensitive human cell line, Science 267 (1995)<br />

1183–1185.<br />

[55] F. Sirzen, A. Nilsson, B. Zhivotovsky, R. Lewensohn, DNA-dependent protein<br />

kinase content and activity in lung carcinoma cell lines: correlation with intrinsic<br />

radiosensitivity, Eur. J. Cancer 35 (1999) 111–116.<br />

[56] S.E. Golding, E. Rosenberg, A. Khalil, A. McEwen, M. Holmes, S. Neill, L.F. Povirk,<br />

K. Valerie, Double strand break repair by homologous recombination is regulated<br />

by cell cycle-independent signaling via ATM in human glioma cells, J. Biol.<br />

Chem. 279 (2004) 15402–15410.<br />

[57] S. Tribius, A. Pidel, D. Casper, ATM protein expression correlates with radioresistance<br />

in primary glioblastoma cells in culture, Int. J. Radiat. Oncol. Biol. Phys.<br />

50 (2001) 511–523.<br />

[58] S.J. Collis, M.J. Swartz, W.G. Nelson, T.L. DeWeese, Enhanced radiation and<br />

chemotherapy-mediated cell killing of human cancer cells by small inhibitory<br />

RNA silencing of DNA repair factors, Cancer Res. 63 (2003) 1550–1554.<br />

[59] E.P. Rogakou, D.R. Pilch, A.H. Orr, V.S. Ivanova, W.M. Bonner, DNA doublestranded<br />

breaks induce histone H2AX phosphorylation on serine 139, J. Biol.<br />

Chem. 273 (1998) 5858–5868.<br />

[60] P.L. Olive, J.P. Banath, Phosphorylation of histone H2AX as a measure of<br />

radiosensitivity, Int. J. Radiat. Oncol. Biol. Phys. 58 (2004) 331–335.<br />

[61] O. Fernandez-Capetillo, A. Lee, M. Nussenzweig, A. Nussenzweig, H2AX: the<br />

histone guardian of the genome, DNA Repair (Amst) 3 (2004) 959–967.<br />

[62] J.H. Matthews, B.E. Meeker, J.D. Chapman, Response of human tumor cell lines<br />

in vitro to fractionated irradiation, Int. J. Radiat. Oncol. Biol. Phys. 16 (1989)<br />

133–138.


Mutation Research 716 (2011) 10–19<br />

Contents lists available at ScienceDirect<br />

Mutation Research/Fundamental and Molecular<br />

Mechanisms of Mutagenesis<br />

jou rnal h omepa g e: www.elsevier.com/locate/molmut<br />

Co mm unit y ad d ress: www.elsevier.com/locate/mutres<br />

DNA damage response signaling in lung adenocarcinoma A549 cells following<br />

gamma and carbon beam irradiation<br />

<strong>Somnath</strong> <strong>Ghosh</strong> a , Himanshi Narang a,∗ , Asitikantha Sarma b , Malini Krishna a<br />

a Radiation Biology and Health Sciences Division, <strong>Bhabha</strong> Atomic Research Centre, Trombay, Mumbai 400 085, India<br />

b Radiation Biology Laboratory, Inter University Accelerator Centre, Aruna Asaf Ali Marg, New Delhi 110 067, India<br />

a r t i c l e i n f o<br />

Article history:<br />

Received 24 March 2011<br />

Received in revised form 22 July 2011<br />

Accepted 26 July 2011<br />

Available online 3 August 2011<br />

Key words:<br />

Carbon<br />

-Rays<br />

-H2AX<br />

ATM<br />

BRCA1<br />

a b s t r a c t<br />

Carbon beams (5.16 MeV/u, LET = 290 keV/m) are high linear energy transfer (LET) radiation characterized<br />

by higher relative biological effectiveness than low LET radiation. The aim of the current study was<br />

to determine the signaling differences between -rays and carbon ion-irradiation. A549 cells were irradiated<br />

with 1 Gy carbon or -rays. Carbon beam was found to be three times more cytotoxic than -rays<br />

despite the fact that the numbers of -H2AX foci were same. Percentage of cells showing ATM/ATR foci<br />

were more with -rays however number of foci per cell were more in case of carbon irradiation. Large<br />

BRCA1 foci were found in all carbon irradiated cells unlike -rays irradiated cells and prosurvival ERK<br />

pathway was activated after -rays irradiation but not carbon. The noteworthy finding of this study is<br />

the early phase apoptosis induction by carbon ions. In the present study in A549 lung adenocarcinoma,<br />

authors conclude that despite activation of same repair molecules such as ATM and BRCA1, differences in<br />

low and high LET damage responses might be due to their distinct macromolecular complexes rather than<br />

their individual activation and the activation of cytoplasmic pathways such as ERK, whether it applies to<br />

all the cell lines need to be further explored.<br />

© 2011 Elsevier B.V. All rights reserved.<br />

1. Introduction<br />

Biological consequences of high-energy charged particle exposure<br />

are of crucial importance in the clinic, space exploration and<br />

risk assessments [1]. Track structure of charged-particle radiation<br />

is known to be a critical determinant of its biological effectiveness<br />

[2–4]. Energy deposition by HZE ions is highly heterogeneous,<br />

with a localized contribution along the trajectory of every particle<br />

and lateral diffusion of energetic electrons (i.e., -rays, the target<br />

atom electrons ionized by the incident HZE ion and emitted at high<br />

energy) many microns from the path of the ions. These particles<br />

are therefore densely ionizing (high-LET) along the primary track<br />

(e.g., the track followed by the incident heavy ion); and also they<br />

have a low-LET component because of the high-energy electrons<br />

ejected by ions as they traverse tissue. Biophysical models have<br />

shown that energy-deposition events by high-LET radiation produce<br />

different DNA lesions, including complex DNA breaks, and<br />

that qualitative difference between high-LET radiation and low-<br />

LET radiation affects both the induction and repair of DNA damage<br />

[4–8]. Because of the effective cell killing many clinicians now<br />

favor charged particle beam therapy over conventional photons<br />

for tumors that are radio resistant and when the survival of nearby<br />

∗ Corresponding author. Tel.: +91 22 2559 5310; fax: +91 22 2550 5151.<br />

E-mail address: himinarang@gmail.com (H. Narang).<br />

cells cannot be compromised e.g. in treatment of lung cancer [9,10].<br />

Although charged particle beams like carbon etc. are currently<br />

being used for the treatment, the signaling pathways activated and<br />

their variance from -rays irradiation is not well understood.<br />

DNA, the main target for radiation induced cell killing, upon<br />

damage, triggers a signaling network that senses different types<br />

of damage and coordinates responses which include activation of<br />

transcription, cell cycle control, apoptosis, senescence, and DNA<br />

repair processes [11]. Hundreds of proteins are recruited in time<br />

dependent manner in the vicinity of a DSB leading to formation<br />

of a signaling-repair complex which can be visualized as foci and<br />

are therefore termed as radiation induced foci (RIF). One of the<br />

components of RIF is -H2AX that is immediately phosphorylated<br />

upon irradiation, can be visualized in situ by immunostaining with<br />

a -H2AX specific antibody and measured quantitatively [12–14].<br />

Mathematical methods has been developed to analyze flow cytometry<br />

data to describe the kinetics of DNA damage response protein<br />

activation after exposure to X-rays and high energy Iron nuclei [15].<br />

Other proteins like ATM (ataxia telangiectasia mutated), ATR (ATM<br />

and Rad3-related) and BRCA1, also function in complexes that can<br />

be seen as foci. There are reports that ATM responds differently<br />

to DSBs induced by high-LET and low-LET radiations [16,17]. The<br />

signal is transduced downstream by proteins such as Chk1/Chk2<br />

and p53 that trigger off DNA repair, cycle arrest or apoptosis. The<br />

cytoplasmic MAPK pathways, namely ERK and JNK, have also been<br />

linked to the DNA damage response and ATM-mediated signaling<br />

0027-5107/$ – see front matter © 2011 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.mrfmmm.2011.07.015


S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19 11<br />

events [18]. ERK1/2 signaling has been shown to be a positive regulator<br />

of homologous recombination repair (HRR) [19] and has also<br />

been shown to be activated by ATM [20]. JNK on the other hand has<br />

been shown to be inhibited by ATM [21] and plays a role in radiation<br />

induced apoptosis. These cytoplasmic signaling pathways coordinate<br />

with DNA damage activated pathways and summation of all<br />

the pathways decides the ultimate fate of the cell [18].<br />

Many studies on signaling pathways after low and high LET radiations<br />

have found the activation to be similar except in the intensity<br />

[22] but since the end result is different, there must be a divergence<br />

of pathways at some stage. It is this stage of signaling, that is different,<br />

that will be crucial to designing therapies that target specific<br />

molecules or pathways. The aim of the present study was to look for<br />

these subtle differences that are of important consequences. A549,<br />

a lung carcinoma cell line was chosen as lung cancer is one of the<br />

most commonly diagnosed types of cancer with the highest death<br />

rates [23] and high LET ions pose a therapeutic advantage over<br />

conventional photons for treatment of non-small-cell lung cancer<br />

(NSCLC) [10].<br />

2. Methods<br />

2.1. Cell culture<br />

Human Lung Adenocarcinoma, A549 cells were cultured and seeded as described<br />

earlier [24].<br />

2.2. Irradiation of cells<br />

For the experiments plateau phase A549 cells were irradiated with 60 Co -rays<br />

and 12 C ions from 15 UD Pelletron accelerator at the Inter University Accelerator<br />

Centre, New Delhi. During carbon irradiation, 35 mm petri plates [NUNC] with cellular<br />

monolayers were kept perpendicular to the particle beam coming out of the<br />

exit window. The petri dishes were kept in the serum free medium in a plexiglass<br />

magazine during the irradiation. The computer controlled irradiation system is such<br />

that during the actual irradiation the dish was removed from medium. Control cells<br />

were sham irradiated in similar way.<br />

At the cell surface the energy of C ions was 62 MeV [5.16 MeV/u, LET = 290 keV/<br />

m]. Both the energy and LET were calculated using the Monte Carlo Code SRIM<br />

code [25]. The corresponding fluence for the dose of 1 Gy was 2.2 × 10 6 particles<br />

cm −2 , thus each cell nuclei can be expected to be physically traversed at least once<br />

by carbon ion.<br />

Fluence was calculated using the relation given below, approximating the density<br />

of cellular matrix as 1.0:<br />

Dose [Gy]=1.6 × 10 −9 ×LET (keV/m) ×particle fluence (cm −2 ) × 1 (cm3 /g)<br />

For gamma irradiation, cells were exposed to 1, 2 and 3 Gy of 60 Co -rays in<br />

in-house facility Gamma Cell 220 [AECL, Canada] at dose rate of 3 Gy/min and same<br />

experimental conditions were kept as far as possible. In all experiments, controls<br />

were sham irradiated.<br />

2.3. Clonogenic cell survival assay<br />

2.6. Image analysis using ImageJ software<br />

The captured images were analyzed for relative quantification of phosphorylation<br />

using ImageJ software [27]. DAPI stained nucleus of each cell was selected and<br />

the same frame was used for relative quantification of phosphorylation using ImageJ<br />

software as described earlier [24]. All the foci were counted manually; at least 100<br />

cells per experiment were analyzed from three independent experiments.<br />

3. Results<br />

3.1. Carbon ions were three times more cytotoxic than gamma<br />

To determine the sensitivity of cells to similar dose of high and<br />

low LET radiations, their ability to form colonies after irradiation<br />

was observed. Irradiation of plateau phase A549 cells with equal<br />

doses of -rays and carbon (290 keV/m) revealed that at same<br />

dose carbon ions were much more cytotoxic than -rays. The calculated<br />

RBE (relative biological effectiveness) of carbon ions relative<br />

to rays for 20% survival of A549 cells was found to be nearly 3.<br />

Thus, carbon ions were found to be three times more toxic than -<br />

rays. At 2 Gy percentage survivals were 60.6 ± 4% and 4.8 ± 0.24%<br />

for -rays and carbon respectively (Fig. 1).<br />

3.2. Number of initial -H2AX foci formed was almost same for<br />

both radiation type<br />

Since DNA double strand breaks are the primary toxic lesions<br />

induced by radiation, the number of DSBs was scored by examining<br />

the foci of phosphorylated H2AX in irradiated cells. After 15 min<br />

of either radiation, -H2AX foci, visualized as bright spots, were<br />

present in all the cells. The average numbers of foci after 15 min<br />

of 1 Gy -rays irradiation were 15.8 ± 4.8, while after 1 Gy carbon<br />

irradiation the foci were 19 ± 4.7 (Fig. 2). Thus unlike the cell survival,<br />

not much difference was observed in number of -H2AX foci<br />

after irradiation with equal doses of carbon and -rays. However,<br />

the size of -H2AX foci in carbon irradiated samples was larger than<br />

-rays irradiated samples. The foci were also counted 4 h after irradiation.<br />

57% of -rays irradiated cells still showed -H2AX foci and<br />

the average foci per cell had reduced to only 2.9 ± 2.28. Carbon ion<br />

irradiated cells also showed a significant reduction in the number<br />

of -H2Ax foci (2 ± 2) per cell that were visible in only 22% of cells<br />

(Fig. 2B and C). When total intensity of -H2AX rather than the foci<br />

was estimated by ImageJ software in the same images, it corroborated<br />

well with the number of -H2AX foci in -ray irradiated cells<br />

at 15 min and 4 h after irradiation (Fig. 2D). However, in case of car-<br />

1<br />

After irradiation, clonogenic assay was done as described earlier [24]. Absolute<br />

plating efficiency at 0 Gy was 32.3 ± 2.4%.<br />

2.4. Western blotting<br />

Unirradiated and irradiated cells were lysed at different time periods and<br />

immunoblotted, as described previously [26]. Blots were probed with anti-Bax, antiphospho<br />

ERK (Thr/Tyr) and anti-phospho JNK (Thr/Tyr). The bound HRP labeled<br />

secondary antibody was detected by Chemiluminiscence (Roche Molecular Biochemicals,<br />

Germany). The total intensity of bands obtained in ponceau staining was<br />

used as a protein loading and transfer control. All band intensities were divided by<br />

the intensity of their respective ponceau blot.<br />

Surviving Fraction<br />

0.1<br />

0.01<br />

γ ray<br />

Carbon beam<br />

2.5. Immunofluorescence staining<br />

Unirradiated and irradiated cells were fixed at different time periods and<br />

immunofluorescence staining was done as described earlier [24]. The primary antibodies<br />

used were mouse anti-pATM (S1981), rabbit anti--H2AX (S139), rabbit<br />

anti-pChk1 (Ser296), rabbit anti-pChk2 (Thr68), rabbit anti-pBRCA1 (Ser1524) and<br />

rabbit anti-pATR (Ser428) (Cell Signaling). Secondary antibodies used were Alexa<br />

Fluor 488 goat anti-rabbit IgG and Alexa Fluor 488 goat anti-mouse IgG (Molecular<br />

Probe, USA).<br />

1E-3<br />

0<br />

1<br />

Dose (Gy)<br />

Fig. 1. Clonogenic cell survival of A549 cells irradiated with 1 and 2 Gy carbon beam<br />

or -rays. Data represents means ± SD of three independent experiments.<br />

2<br />

3


12 S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19<br />

Fig. 2. Phosphorylation of H2AX (-H2AX) at different time points after exposure to 1 Gy -rays or carbon irradiation in A549 cells. Representative images showing -H2AX<br />

foci (A) 15 min and (B) 4 h after irradiation. Each phospho-H2AX antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted<br />

with ProLong Gold antifede with DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. (C) Graph represents average<br />

numbers of foci per cell, percentage of cells showing the foci is marked above the bars. (D) Graph represents relative intensity of -H2AX as determined by ImageJ software.<br />

At least 100 cells per experiment were analyzed from three independent experiments. Data represents means ± SD of three independent experiments. (For interpretation of<br />

the references to color in this figure legend, the reader is referred to the web version of the article.)<br />

bon irradiation, the total intensity of -H2AX at 4 h was yet high<br />

(Fig. 2D).<br />

Since initial response as observed from -H2AX foci at 15 min<br />

was similar with both types of radiation, irradiation induced foci of<br />

downstream signaling components was followed at 4 h to highlight<br />

the signaling differences at that time period.<br />

3.3. ATM/ATR foci in gamma and carbon irradiated cells<br />

ATM, an important DNA damage response protein to ionizing<br />

radiation and a part of radiation induced foci (RIF), is mainly<br />

involved in cell cycle arrest and repair. In response to DSBs, ATM is<br />

activated and phosphorylates key players in various branches of the


S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19 13<br />

Fig. 3. Phosphorylation of ATM at 4 h after exposure to 1 Gy -rays or carbon irradiation in A549 cells. (A) Representative image showing p-ATM foci 4 h after irradiation.<br />

Each phospho-ATM antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI<br />

(blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph represents average numbers of foci per cell, percentage of<br />

cells showing the foci is marked above the bars. (C) Graph represents relative intensity of p-ATM as determined by ImageJ software. At least 100 cells per experiment were<br />

analyzed from three independent experiments. Data represents means ± SD of three independent experiments. (For interpretation of the references to color in this figure<br />

legend, the reader is referred to the web version of the article.)<br />

DNA damage response network that include H2AX, BRCA1, Chk1/2<br />

and p53. Phospho ATM foci and intensity were scored at 4 h after -<br />

rays or carbon irradiation. At 4 h after gamma irradiation, 88% cells<br />

showed that pATM foci with the average foci per cell was 6.5 ± 4.4,<br />

as against 25% control cells (3.4 ± 1.67) (Fig. 3). Like -H2AX foci,<br />

small number of ATM foci 4 h after -rays irradiation indicates that<br />

by 4 h most of the repair has taken place. However, at 4 h after carbon<br />

ion irradiation only 52% of the cells showed the foci but the<br />

average pATM foci per cell (21 ± 10) was much higher as compared<br />

to 4 h after -rays (Fig. 3). Total intensity levels of phospho ATM<br />

matched with the number of foci observed (Fig. 3C).<br />

ATR, another important DNA damage sensor, which is known to<br />

respond late to the IR induced damage, was also looked for its activation<br />

at 4 h. After -rays irradiation, 53% of the cells showed ATR<br />

foci (2.8 ± 1.64), as against no foci in any of the control-unirradiated<br />

cells. On the other hand, 4 h after carbon ion irradiation although<br />

only 10% of cells showed the foci, the average foci per cell (7 ± 4.5)<br />

were thrice that of -rays irradiated cells (Fig. 4). Total intensity<br />

levels of phospho ATR matched with the number of foci observed<br />

(Fig. 4C).<br />

3.4. BRCA1 foci<br />

Another important protein involved in DSB repair is the tumor<br />

suppressor protein BRCA1 which is phosphorylated by ATM and is<br />

also a part of RIF. 4 h after -rays irradiation, 55% cells displayed<br />

BRCA1 foci (13 ± 8) while 100% of carbon irradiated cells displayed<br />

BRCA1 foci (7.5 ± 0.64) (Fig. 5) and the foci, although less in number<br />

were larger in size than -ray induced foci. The intensity of<br />

phosphorylation of BRCA1 as determined by ImageJ software was<br />

higher in carbon irradiated cells (Fig. 5C). Presence of BRCA1 foci<br />

in all cells irradiated after carbon irradiation indicates that BRCA1<br />

might be the important player in response to complex DNA damage<br />

induced by high LET radiation.<br />

3.5. Activation of downstream effectors, Chk1 and Chk2<br />

Activation of ATM, ATR and BRCA1 after DNA damage leads to<br />

the activation of further downstream components including Chk1<br />

and Chk2. After activation in the nuclei, Chk1 and Chk2 move to<br />

cytoplasm. Phospho Chk1, the major target of ATR, was found to<br />

increase in nuclei of carbon-irradiated cells but not in -rays irradiated<br />

cells (Fig. 6A). Marginal activation of Chk2, the preferred<br />

substrate of ATM, was observed in the nucleus in all the cells irrespective<br />

of radiation quality (Fig. 6B).<br />

3.6. Activation of intracellular MAPKs, ERK and JNK<br />

ERK1/2 showed significant activation 1 h after -rays irradiation<br />

and thereafter the increase was persistent till 4 h (Fig. 7A), while


14 S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19<br />

Fig. 4. Phosphorylation of ATR at 4 h after exposure to 1 Gy -rays or carbon irradiation in A549 cells. (A) Representative image showing p-ATR foci 4 h after irradiation. Each<br />

phospho-ATR antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI (blue). All<br />

images were captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph represents average numbers of foci per cell, percentage of cells showing<br />

the foci is marked above the bars. (C) Graph represents relative intensity of p-ATR as determined by ImageJ software. At least 100 cells per experiment were analyzed from<br />

three independent experiments. Data represents means ± SD of three independent experiments. (For interpretation of the references to color in this figure legend, the reader<br />

is referred to the web version of the article.)<br />

JNK did not show any activation after -rays irradiation (Fig. 7B).<br />

Interestingly, after carbon irradiation the activation pattern of these<br />

kinases was exact opposite. ERK was not activated at all after carbon<br />

irradiation (Fig. 7A) while JNK showed marginal activation at<br />

15 min that remained so till 1 h (Fig. 7B). The expression of total<br />

ERK did not change during the same period in the cells that had<br />

been exposed to either -rays or carbon beam.<br />

3.7. Apoptosis<br />

At 4 h after irradiation, morphological features of DAPI stained<br />

apoptotic nuclei were scored and many apoptotic nuclei were<br />

observed in cells irradiated with carbon ion but not in -rays irradiated<br />

cells (Fig. 8A and B).<br />

Upregulation of Bax expression was also observed as early as<br />

2 h after carbon ion irradiation indicating apoptotic tendency of the<br />

cells (Fig. 8C). -Rays irradiated cells did not show any upregulation<br />

of Bax (data not shown).<br />

4. Discussion<br />

Although 1 Gy dose of carbon ions was three times more toxic to<br />

the cells than -rays, both produced same number of -H2AX foci,<br />

indicators of DNA double strand breaks. This discrepancy between<br />

observed cytotoxicity and estimated -H2AX foci after high LET<br />

radiation has also been reported previously [28]. Since our aim<br />

was to find out signaling differences arising from DNA damaged<br />

by either of the radiations, we chose 1 Gy of both radiations rather<br />

than their equitoxic doses. The observed decrease in number of -<br />

H2AX foci 4 h after -rays irradiation indicates repair of damage<br />

and is supported by nearly 100% survival. However, the decrease in<br />

-H2AX foci after carbon ion irradiation was definitely not indicative<br />

of repair because the cell survival data contradicts the fact<br />

(Fig. 2). Foci formation is a biochemical kinetic process involving<br />

both formation and loss of foci [29], there is never an exact one-toone<br />

correspondence between the number of DSB and foci [30]. In<br />

addition, multiple DSB may be visible as a single focus due to DNA<br />

supercoiling [31]. For heavy ions, clustering of damage including<br />

DSB along the trajectory of the ion leads to further complexities<br />

in relating foci counts to damage numbers [14]. However, since<br />

total intensity of -H2AX was yet high at 4 h in carbon irradiated<br />

cells, indicating its phosphorylated state, it may represent sites of<br />

increased DNA damage after carbon irradiation, arising from misrepair<br />

or incomplete repair [32,33].<br />

Similar number of -H2AX foci at 15 min after both high and low<br />

LET radiations indicated that initial events might not differ much<br />

with the radiation quality. Activation of other molecules including<br />

ATM, ATR, BRCA1, Chk1, Chk2 and Bax was therefore followed


S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19 15<br />

Fig. 5. Phosphorylation of BRCA1 at 4 h after exposure to 1 Gy -rays or carbon irradiation in A549 cells. (A) Representative image showing p-BRCA1 foci 4 h after irradiation.<br />

Each phospho-BRCA1 antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI<br />

(blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph represents average numbers of foci per cell, percentage of<br />

cells showing the foci is marked above the bars. (C) Graph represents relative intensity of p-BRCA1 as determined by ImageJ software. At least 100 cells per experiment were<br />

analyzed from three independent experiments. Data represents means ± SD of three independent experiments. (For interpretation of the references to color in this figure<br />

legend, the reader is referred to the web version of the article.)<br />

only at later time point of 4 h as comparing the activation of these<br />

molecules at this time could highlight signaling differences with<br />

respect to clustered damage.<br />

ATM and ATR, the initial DNA damage sensors, are critical regulators<br />

of cell cycle checkpoints and repair. ATM has particularly<br />

been shown to respond to IR induced damage. Presence of few<br />

ATM foci in -rays irradiated cells 4 h after irradiation indicates the<br />

repair of the damage in most cells. However, in carbon irradiated<br />

cells, although the percentage of cells showing the pATM foci (55%)<br />

was less, the average number of foci/cell was more than -rays irradiated<br />

cells (Fig. 3). Similar trend was also observed with ATR foci<br />

with respect to radiation quality (Fig. 4). Higher percentage of -<br />

rays irradiated cells showing the ATM/ATR foci 4 h after irradiation<br />

as compared to carbon irradiated cells indicates more homogenous<br />

nature of -rays induced responses while presence of lesser average<br />

ATM/ATR foci per -rays irradiated cell as against carbon irradiation<br />

indicates efficient repair after -rays irradiation. This observation<br />

is in agreement with the previous studies where higher activation<br />

of repair molecules had been observed few hours after heavy ion<br />

irradiation as compared to low LET irradiation [22,34] and has been<br />

attributed to unrepaired DNA that keeps the damage sensors in activated<br />

state. However, these studies have shown that size of the foci<br />

and the number of foci at 24 h is different between high and low<br />

LET radiations.<br />

The formation of BRCA1 foci was unlike ATM/ATR and very interesting.<br />

BRCA1 exists in a complex containing ATM and other DSB<br />

repair proteins and both ATM and BRCA1 have been regarded as primary<br />

regulators of HRR [35]. Presence of BRCA1 foci in ∼55% (Fig. 5)<br />

of the -rays irradiated cells is in agreement with the previous studies<br />

where BRCA1 foci have been found to be activated only in cells<br />

in S, G2/M phases of the cycle [34] and therefore is thought to play<br />

a role in HRR which can take the advantage of the other strand that<br />

may be intact. Unlike after -rays irradiation, presence of BRCA1<br />

foci in all of the carbon irradiated cells was quite intriguing. Moreover,<br />

the downstream substrates, Chk1 (Fig. 6A) and Chk2 (Fig. 6B)<br />

also showed relatively higher activation in carbon irradiated cells.<br />

Despite higher activation of all these repair components in carbon<br />

irradiated cells, their DNA was still unrepaired as indicated by -<br />

H2AX staining and survival data. The authors therefore, believe that<br />

activated repair components might have different functionalities<br />

after low and high LET radiations and that this functionality might<br />

be imparted through association of proteins in different macromolecular<br />

complexes. This was supported by larger size of -H2AX<br />

and BRCA1 foci observed after carbon irradiation. This result is<br />

in accordance with previous studies where authors had observed<br />

larger foci of DNA damage response proteins in the cells that had<br />

been exposed to high LET radiation [14,36]. This means that foci<br />

from -rays in essentially have a distinct meaning from those from


16 S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19<br />

Fig. 6. Phosphorylation of Chk1 and Chk2 at 4 h after exposure to 1 Gy -rays or carbon irradiation in A549 cells. (A) Representative image showing p-Chk1 4 h after<br />

irradiation. Graph represents relative phosphorylation of Chk1 as determined by ImageJ software. (B) Representative image showing p-Chk2 4 h after irradiation. Graph<br />

represents relative phosphorylation of Chk2 as determined by ImageJ software. Each phospho-Chk1 or phospho-Chk2 antibody was indirectly labeled with Molecular Probe<br />

488 secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the<br />

same exposure time. At least 100 cells per experiment were analyzed from three independent experiments. Data represents means ± SD of three independent experiments.<br />

(For interpretation of the references to color in this figure legend, the reader is referred to the web version of the article.)<br />

high LET ions. Presence of ATM foci in only half of carbon irradiated<br />

cells while BRCA1 foci in all the cells also suggest that BRCA1<br />

is associated with different proteins and may not require activated<br />

ATM. Further study on co-localization experiment for BRCA1 with<br />

ATM by immunofluorescence would provide better understanding<br />

of repair responses elicited by low and high LET radiations. Number<br />

of Chk2 foci, another important regulator of HRR, also did not<br />

seem to parallel ATM foci. This, along with the fact that the survival<br />

of the cells is only 20% (Fig. 1) and many apoptotic nuclei are<br />

present at 4 h (Fig. 8) indicates that these macromolecular complexes<br />

containing BRCA1 may be involved in apoptotic responses<br />

after a complex damage to the DNA. Recent studies suggest multifunctional<br />

nature of BRCA1 in maintaining genome integrity. It has<br />

been shown to play an important role in apoptosis and cell cycle<br />

control and does so by associating in distinct protein complexes<br />

[37]. However, further studies on BRCA1 and apoptosis again need<br />

to be experimentally addressed.<br />

The cytoplasmic MAPK pathways, ERK and JNK, that feed into<br />

and are fed upon by DNA damage repair signaling also play an<br />

equally important role in deciding the fate of the irradiated cell.<br />

Reactive oxygen species generated after -rays irradiation has been<br />

thought to be one of the mechanisms for early activation of these<br />

kinases [18]. ATM has been shown to enhance repair via activation<br />

of ERK pathway [35] and inhibition of JNK [21]. Observed activation<br />

of ERK (Fig. 7A) but not JNK (Fig. 7B) in -rays irradiated cells<br />

indicated the dominance of pro-survival pathways.


S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19 17<br />

Fig. 7. Phosphorylation of ERK (pERK) and JNK (pJNK) in A549 cells at different time periods (h) after exposure to 1 Gy dose of -rays or carbon ion irradiation. Treated and<br />

untreated cells were immunoblotted and detected using anti-phospho ERK (panel A) or anti phospho JNK (panel B). One representative image of three independent experiments<br />

is shown here. Bar graph represents the relative pERK (panel A) and pJNK (panel B) band intensities plotted against time. The intensities of pERK/pJNK were obtained after<br />

dividing them with the respective loading control intensities and have been normalized for comparison. Data represents means ± SEM of three independent experiments.<br />

Fig. 8. Apoptosis in carbon irradiated A549 cells. (A) Apoptotic nuclei of A549 cells were visualized and quantified by using DAPI staining, 4 h after carbon or -rays<br />

irradiation (1 Gy). (B) Graph showing the percentage of cells with apoptotic nuclei. At least 100 cells per experiment were analyzed from three independent experiments. (C)<br />

Representative immunoblot image depicting upregulation of Bax, observed at various time periods (2, 3, 4 and 6 h) after carbon irradiation (1 Gy).


18 S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19<br />

Since, with high LET radiation yield of ROS is very low as compared<br />

to low LET radiation [38–40], the early activation of ERK was<br />

not expected with high LET radiation. However, complete absence<br />

of ERK activation even hours after high LET radiation suggests that<br />

repair pathways are neither feeding into nor being fed upon by<br />

intracellular cytoprotective signaling. However, the pro-apoptotic<br />

JNK pathway was found to be activated after high LET radiation.<br />

BRCA1 induced apoptotic responses have also been shown to be<br />

via activation of JNK [41]. Thus, after carbon irradiation, DNA damage<br />

induced activation of repair components are being fed into<br />

by cytotoxic signaling rather than cytoprotective responses and<br />

was evident in the form of early induction of apoptosis (Fig. 8).<br />

Since, cellular decisions are summation of all the activated pathways,<br />

the final response after high LET radiation tips towards<br />

death.<br />

5. Conclusions<br />

In summary, our results suggest that the subtle differences<br />

leading to different outcomes due to radiation quality seem to<br />

lie in different macromolecular complexes of crucial DNA damage<br />

response proteins and activation of other intracellular pathways<br />

in A549 lung adenocarcinoma cell line. Analysis of functionality of<br />

these macromolecular complexes as a whole rather than individual<br />

proteins and holistic study involving intracellular survival pathways<br />

could help in revealing the mechanistic details of complex<br />

DNA damage responses.<br />

Conflict of interest<br />

The authors declare that there are no conflicts of interest.<br />

Acknowledgments<br />

This work was funded by Board of Research in Nuclear Sciences,<br />

Department of Atomic Energy [DAE], Government of India, through<br />

a project sanctioned to one of the authors, A. Sarma, bearing sanction<br />

number 2007/37/37/BRNS. All the authors are thankful to the<br />

Director, IUAC, New Delhi for providing radiation facility for the<br />

work. We would also like to extend our thanks to all the members<br />

of Pelletron group of IUAC and Harminder Kaur, Radiation Biology<br />

Laboratory, IUAC for their sincere help during irradiation. Authors<br />

would like to thank Mr. Manjoor Ali and Mr. Paresh Khadilkar,<br />

RB&HSD, BARC for their help in Confocal Microscopy and Lab work<br />

respectively.<br />

References<br />

[1] F.A. Cucinotta, M. Durante, Cancer risk from exposure to galactic cosmic rays:<br />

implications for space exploration by human beings, Lancet Oncol. 7 (2006)<br />

431–435.<br />

[2] F.A. Cucinotta, H. Wu, M.R. Shavers, K. George, Radiation dosimetry and biophysical<br />

models of space radiation effects, Gravit. Space Biol. Bull. 16 (2003)<br />

11–18.<br />

[3] D.T. Goodhead, The initial physical damage produced by ionizing radiations,<br />

Int. J. Radiat. Biol. 56 (1989) 623–634.<br />

[4] F.A. Cucinotta, H. Nikjoo, D.T. Goodhead, Model for radial dependence of frequency<br />

distributions for energy imparted in nanometer volumes from HZE<br />

particles, Radiat. Res. 153 (2000) 459–468.<br />

[5] E.A. Blakely, A. Kronenberg, Heavy-ion radiobiology: new approaches to delineate<br />

mechanisms underlying enhanced biological effectiveness, Radiat. Res.<br />

150 (1998) S126–S145.<br />

[6] F.A. Cucinotta, R. Katz, J.W. Wilson, Radial distribution of electron spectra from<br />

high-energy ions, Radiat. Environ. Biophys. 37 (1998) 259–265.<br />

[7] B.M. Sutherland, P.V. Bennett, O. Sidorkina, J. Laval, Clustered DNA damages<br />

induced in isolated DNA and in human cells by low doses of ionizing radiation,<br />

Proc. Natl. Acad. Sci. U.S.A. 97 (2000) 103–108.<br />

[8] B. Rydberg, B. Cooper, P.K. Cooper, W.R. Holley, A. Chatterjee, Dose-dependent<br />

misrejoining of radiation-induced DNA double-strand breaks in human fibroblasts:<br />

experimental and theoretical study for high- and low-LET radiation,<br />

Radiat. Res. 163 (2005) 526–534.<br />

[9] D. Schulz-Ertner, A. Nikoghosyan, C. Thilmann, T. Haberer, O. Jakel, C. Karger, G.<br />

Kraft, M. Wannenmacher, J. Debus, Results of carbon ion radiotherapy in 152<br />

patients, Int. J. Radiat. Oncol. Biol. Phys. 58 (2004) 631–640.<br />

[10] T. Miyamoto, N. Yamamoto, H. Nishimura, M. Koto, H. Tsujii, J.E. Mizoe, T.<br />

Kamada, H. Kato, S. Yamada, S. Morita, K. Yoshikawa, S. Kandatsu, T. Fujisawa,<br />

Carbon ion radiotherapy for stage I non-small cell lung cancer, Radiother. Oncol.<br />

66 (2003) 127–140.<br />

[11] S.P. Jackson, Sensing and repairing DNA double-strand breaks, Carcinogenesis<br />

23 (2002) 687–696.<br />

[12] E.P. Rogakou, D.R. Pilch, A.H. Orr, V.S. Ivanova, W.M. Bonner, DNA doublestranded<br />

breaks induce histone H2AX phosphorylation on serine 139, J. Biol.<br />

Chem. 273 (1998) 5858–5868.<br />

[13] O. Fernandez-Capetillo, A. Lee, M. Nussenzweig, A. Nussenzweig, H2AX: the<br />

histone guardian of the genome, DNA Repair (Amst.) 3 (2004) 959–967.<br />

[14] N. Desai, E. Davis, P. O’Neill, M. Durante, F.A. Cucinotta, H. Wu, Immunofluorescence<br />

detection of clustered gamma-H2AX foci induced by HZE-particle<br />

radiation, Radiat. Res. 164 (2005) 518–522.<br />

[15] L.J. Chappell, M.K. Whalen, S. Gurai, A. Ponomarev, F.A. Cucinotta, J.M. Pluth,<br />

Analysis of flow cytometry DNA damage response protein activation kinetics<br />

after exposure to X rays and high-energy iron nuclei, Radiat. Res. 174 (2010)<br />

691–702.<br />

[16] M.K. Whalen, S.K. Gurai, H. Zahed-Kargaran, J.M. Pluth, Specific ATM-mediated<br />

phosphorylation dependent on radiation quality, Radiat. Res. 170 (2008)<br />

353–364.<br />

[17] R. Ugenskiene, K. Prise, M. Folkard, J. Lekki, Z. Stachura, M. Zazula, J. Stachura,<br />

Dose response and kinetics of foci disappearance following exposure to highand<br />

low-LET ionizing radiation, Int. J. Radiat. Biol. 85 (2009) 872–882.<br />

[18] K. Valerie, A. Yacoub, M.P. Hagan, D.T. Curiel, P.B. Fisher, S. Grant, P. Dent,<br />

Radiation-induced cell signaling: inside-out and outside-in, Mol. Cancer Ther.<br />

6 (2007) 789–801.<br />

[19] S.E. Golding, E. Rosenberg, S. Neill, P. Dent, L.F. Povirk, K. Valerie, Extracellular<br />

signal-related kinase positively regulates ataxia telangiectasia mutated,<br />

homologous recombination repair, and the DNA damage response, Cancer Res.<br />

67 (2007) 1046–1053.<br />

[20] K.E. Keating, N. Gueven, D. Watters, H.P. Rodemann, M.F. Lavin, Transcriptional<br />

downregulation of ATM by EGF is defective in ataxia-telangiectasia cells<br />

expressing mutant protein, Oncogene 20 (2001) 4281–4290.<br />

[21] N. Weizman, Y. Shiloh, A. Barzilai, Contribution of the ATM protein to<br />

maintaining cellular homeostasis evidenced by continuous activation of<br />

the AP-1 pathway in ATM-deficient brains, J. Biol. Chem. 278 (2003)<br />

6741–6747.<br />

[22] A. Asaithamby, N. Uematsu, A. Chatterjee, M.D. Story, S. Burma, D.J. Chen, Repair<br />

of HZE-particle-induced DNA double-strand breaks in normal human fibroblasts,<br />

Radiat. Res. 169 (2008) 437–446.<br />

[23] A. Jemal, R. Siegel, E. Ward, Y. Hao, J. Xu, T. Murray, M.J. Thun, Cancer statistics,<br />

CA Cancer J. Clin. 58 (2008) 71–96.<br />

[24] S. <strong>Ghosh</strong>, N.N. Bhat, S. Santra, R.G. Thomas, S.K. Gupta, R.K. Choudhury, M.<br />

Krishna, Low energy proton beam induces efficient cell killing in A549 lung<br />

adenocarcinoma cells, Cancer Invest. 28 (2010) 615–622.<br />

[25] J.F. Ziegler, M.D. Ziegler, J.P. Biersack, SRIM – The stopping and range of ions in<br />

matter, Nucl. Instrum. Meth. B 268 (2010) 1818–1823.<br />

[26] H. Narang, M. Krishna, Effect of nitric oxide donor and gamma irradiation on<br />

MAPK signaling in murine peritoneal macrophages, J. Cell. Biochem. 103 (2008)<br />

576–587.<br />

[27] W.S. Rasband, ImageJ, U.S. <strong>National</strong> Institutes of Health, Bethesda, USA,<br />

1997–2006, http://rsb.info.nih.gov/ij/.<br />

[28] K.M. Prise, M. Pinto, H.C. Newman, B.D. Michael, A review of studies of ionizing<br />

radiation-induced double-strand break clustering, Radiat. Res. 156 (2001)<br />

572–576.<br />

[29] F.A. Cucinotta, J.M. Pluth, J.A. Anderson, J.V. Harper, P. O’Neill, Biochemical<br />

kinetics model of DSB repair and induction of gamma-H2AX foci by nonhomologous<br />

end joining, Radiat. Res. 169 (2008) 214–222.<br />

[30] S. <strong>Ghosh</strong>, H. Narang, A. Sarma, H. Kaur, M. Krishna, Activation of DNA damage<br />

response signaling in lung adenocarcinoma A549 cells following oxygen beam<br />

irradiation, Mutat. Res. 723 (2011) 190–198.<br />

[31] A.L. Ponomarev, S.V. Costes, F.A. Cucinotta, Stochastic properties of radiationinduced<br />

DSB: DSB distributions in large scale chromatin loops the HPRT gene<br />

and within the visible volumes of DNA repair foci, Int. J. Radiat. Biol. 84 (2008)<br />

916–929.<br />

[32] E.R. Foster, J.A. Downs, Histone H2A phosphorylation in DNA double-strand<br />

break repair, FEBS J. 272 (2005) 3231–3240.<br />

[33] N.F. Lowndes, G.W. Toh, DNA repair: the importance of phosphorylating histone<br />

H2AX, Curr. Biol. 15 (2005) R99–R102.<br />

[34] K.H. Karlsson, B. Stenerlow, Focus formation of DNA repair proteins in normal<br />

and repair-deficient cells irradiated with high-LET ions, Radiat. Res. 161 (2004)<br />

517–527.<br />

[35] S.E. Golding, E. Rosenberg, A. Khalil, A. McEwen, M. Holmes, S. Neill, L.F. Povirk,<br />

K. Valerie, Double strand break repair by homologous recombination is regulated<br />

by cell cycle-independent signaling via ATM in human glioma cells, J. Biol.<br />

Chem. 279 (2004) 15402–15410.<br />

[36] A.L. Ponomarev, J. Huff, F.A. Cucinotta, The analysis of the densely populated<br />

patterns of radiation-induced foci by a stochastic, Monte Carlo model of DNA<br />

double-strand breaks induction by heavy ions, Int. J. Radiat. Biol. 86 (2010)<br />

507–515.<br />

[37] C.X. Deng, BRCA1: cell cycle checkpoint genetic instability, DNA damage<br />

response and cancer evolution, Nucleic Acids Res. 34 (2006) 1416–1426.


S. <strong>Ghosh</strong> et al. / Mutation Research 716 (2011) 10–19 19<br />

[38] A. Kuppermann, Diffusion kinetics in radiation chemistry: an assessment, in:<br />

R.D. Cooper, R.W. Wood (Eds.), Physical Mechanisms in Radiation Biology,<br />

Springfield, Virginia, 1974, pp. 155–183.<br />

[39] D.E. Watt, Absolute biological effectiveness of neutrons and photons, Radiat.<br />

Prot. Dosim. 23 (1988) 63–67.<br />

[40] M. Spotheim-Maurizot, M. Charlier, R. Sabattier, DNA radiolysis by fast neutrons,<br />

Int. J. Radiat. Biol. 57 (1990) 301–313.<br />

[41] S. Lafarge, V. Sylvain, M. Ferrara, Y.J. Bignon, Inhibition of BRCA1 leads to<br />

increased chemoresistance to microtubule-interfering agents, an effect that<br />

involves the JNK pathway, Oncogene 20 (2001) 6597–6606.


Mutation Research 723 (2011) 190–198<br />

Contents lists available at ScienceDirect<br />

Mutation Research/Genetic Toxicology and<br />

Environmental Mutagenesis<br />

j o ur nal homep ag e: www.elsevier.com/locate/gentox<br />

Co mm unit y add re ss: www.elsevier.com/locate/mutres<br />

Activation of DNA damage response signaling in lung adenocarcinoma A549 cells<br />

following oxygen beam irradiation<br />

<strong>Somnath</strong> <strong>Ghosh</strong> a,∗ , Himanshi Narang a , Asiti Sarma b , Harminder Kaur b , Malini Krishna a<br />

a Radiation Biology and Health Science Division, <strong>Bhabha</strong> Atomic Research Centre, Trombay, Mumbai 400085, India<br />

b Radiation Biology Laboratory, Inter University Accelerator Centre, Aruna Asaf Ali Marg, New Delhi 110 067, India<br />

a r t i c l e i n f o<br />

Article history:<br />

Received 1 December 2010<br />

Received in revised form 5 April 2011<br />

Accepted 9 May 2011<br />

Available online 14 May 2011<br />

Keywords:<br />

Oxygen<br />

Gamma<br />

-H2AX<br />

ATM<br />

ATR<br />

a b s t r a c t<br />

Oxygen beams are high linear energy transfer (LET) radiation characterized by higher relative biological<br />

effectiveness than low LET radiation. The aim of the current study was to determine the signaling differences<br />

between - and oxygen ion-irradiation. Activation of various signaling molecules was looked in<br />

A549 lung adenocarcinoma cells irradiated with 2 Gy oxygen, 2 Gy or 6 Gy -radiation. Oxygen beam was<br />

found to be three times more cytotoxic than -radiation. By 4 h there was efficient repair of DNA in A549<br />

cells exposed to 2 Gy or 6 Gy gamma radiation but not in cells exposed to 2 Gy oxygen beam as determined<br />

by -H2AX counting. Number of ATM foci was found to be significantly higher in cells exposed<br />

to 2 Gy oxygen beam. Percentage of cells showing ATR foci were more with gamma however number of<br />

foci per cell were more in case of oxygen beam. Oxygen beam irradiated cells showed phosphorylation of<br />

Chk1, Chk2 and p53. Many apoptotic nuclei were seen by DAPI staining in cells exposed to oxygen beam.<br />

The noteworthy finding of this study is the activation of the sensor proteins, ATM and ATR by oxygen<br />

irradiation and the significant activation of Chk1, Chk2 and p53 only in the oxygen beam irradiated cells.<br />

© 2011 Elsevier B.V. All rights reserved.<br />

1. Introduction<br />

Exposure of mammalian cells to ionizing radiation leads to the<br />

activation of several signaling pathways that result in apoptosis.<br />

Although, the end point, apoptosis is effectively activated by conventional<br />

X-ray treatment, the development of radio resistance<br />

following therapy remains a major cause for concern. Clinicians<br />

are now favoring charged particle therapy to avoid the issue [1–4].<br />

However, the mechanism of cell killing by oxygen beam is not yet<br />

well delineated.<br />

High LET radiations like oxygen have an increased biological<br />

effectiveness compared to X-rays for gene mutation, genomic instability,<br />

and carcinogenesis [5]. This is because charged particle tracks<br />

show a high density of ionizing events along the center of particle<br />

path, resulting in spatially localized energy deposition within<br />

the cellular target. This high relative biological effectiveness has<br />

been imparted to difficult to repair clustered DNA damage that is<br />

generated in cells after high LET irradiation.<br />

Lung cancer is a frequently occurring, mostly lethal disease in<br />

all countries world wide [6]. It is one of the most commonly diagnosed<br />

types of cancer and has the highest death rates [7,8]. Overall,<br />

fewer than 10% of people with primary lung cancer are alive 5<br />

∗ Corresponding author. Tel.: +91 22 2559 0415; fax: +91 22 2550 5151.<br />

E-mail addresses: somnath@barc.gov.in, ghosh.barc@gmail.com (S. <strong>Ghosh</strong>).<br />

years after diagnosis. Depending on tumor size, location and histology,<br />

several treatment options are available, including surgery and<br />

radiotherapy (RT), both often combined with chemotherapy [9]. A<br />

major problem remains local tumor control [10,11]. Therefore, new<br />

ways to deliver radiotherapy beyond photons have been sought<br />

for, including oxygen and carbon ions [12]. Furthermore, treatment<br />

with charged particle radiation has several potential advantages<br />

over treatment with gamma or X-ray radiation such as the Bragg<br />

peak enables precise localization of the radiation dose, an inverted<br />

depth-dose distribution, a higher relative biological effectiveness<br />

(RBE), and a lower cellular capability for repair of radiation injury.<br />

Because of their superior dose-distribution, a therapeutic gain can<br />

be expected with charged particles [1–4]. Although charged particle<br />

therapies are currently being used for the treatment of many<br />

cancers, the mechanism of cell killing and its variance from gamma<br />

irradiation is not well understood. Understanding the specific biological<br />

effects of charged particle radiation on cancer cells could also<br />

provide valuable insights for the design of novel therapeutic applications<br />

for the treatment of cancers which are resistant to many<br />

types of therapies. Moreover, therapies can be designed only if the<br />

signaling pathway is understood.<br />

ATM (ataxia telangiectasia mutated), a protein kinase, is the<br />

major mediator of DSB responses. The kinase is activated by<br />

autophosphorylation when DSBs are formed and it phosphorylates<br />

a number of proteins involved in the repair and damage signaling<br />

pathways after exposure to ionizing radiation [13,14]. Many<br />

1383-5718/$ – see front matter © 2011 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.mrgentox.2011.05.002


S. <strong>Ghosh</strong> et al. / Mutation Research 723 (2011) 190–198 191<br />

proteins such as ATM, ATR (ATM and Rad3-related) and H2AX<br />

are part of irradiation-induced foci (IRIF), which are at the core of<br />

DNA damage signaling apparatus [15–17]. These foci contain hundreds<br />

to thousands of proteins and are believed to represent sites<br />

with ongoing repair or to be an indication of a checkpoint mechanism.<br />

All these proteins form foci at different time points after<br />

irradiation. In this study we have looked at the foci at 4 h after irradiation.<br />

This time point was chosen since by this time most of the<br />

repairable damage is repaired and hence the foci formation is no<br />

longer observed if DNA repair has been completed.<br />

In this study, A549 cells were irradiated with either 2 Gy oxygen<br />

beam or 2 Gy and 6 Gy -radiation and ensuing activation of few<br />

important components involved in DNA repair, cell cycle arrest, and<br />

apoptosis pathways were followed to elucidate the mechanisms<br />

behind observed cellular response.<br />

n<br />

Fractio<br />

viving F<br />

Surv<br />

1<br />

0.1<br />

0.01<br />

γ ray<br />

Oxygen<br />

2. Materials and methods<br />

2.1. Cell culture<br />

Human A549 cells were cultured in Dulbecco’s modified eagles medium (Sigma)<br />

supplemented with 10% fetal calf serum (Sigma). Cells were maintained at 37 ◦ C in<br />

humidified atmosphere with 5% CO 2. Cells were seeded at density of 5 × 10 5 cells in<br />

small 35 mm petri plates [NUNC] 24 h before the irradiation.<br />

1E-3<br />

0<br />

1<br />

2<br />

Dose (Gy)<br />

3<br />

2.2. Irradiation of cells<br />

For the experiments, exponentially growing A549 cells were irradiated with<br />

60 Co gamma rays and 16 O ions from 15 UD Pelletron accelerator at the Inter University<br />

Accelerator Centre, New Delhi. During oxygen irradiation, 35 mm petri plates<br />

[NUNC] with cellular monolayers were kept perpendicular to the particle beam coming<br />

out of the exit window. The petri dishes were kept in the serum free medium<br />

in a plexiglass magazine during the irradiation. The computer controlled irradiation<br />

system is such that during the actual irradiation the dish was removed from<br />

medium.<br />

At the cell surface the energy of O ions was 55.8 MeV [LET = 614 keV/m]. Both<br />

the energy and LET were calculated using the Monte Carlo Code SRIM code [18,19].<br />

The corresponding fluence for the dose of 2 Gy was 2.02 × 10 6 particles cm −2 , thus<br />

each cell nuclei can be expected to be physically traversed at least once by oxygen<br />

ion.<br />

Fluence was calculated using the relation given below [20], approximating the<br />

density of cellular matrix as 1.0:<br />

Dose [Gy] = 1.6 × 10 −9 × LET (keV/m) × particle fluence (cm −2 ) × 1 (cm3 /g)<br />

For gamma irradiation, cells were exposed to 1–3 and 6 Gy of 60 Co radiation in<br />

in-house facility Gamma Cell 220 [AECL, Canada] at dose rate of 3 Gy/min and same<br />

experimental conditions were kept as far as possible. In all experiments, controls<br />

were sham irradiated.<br />

Fig. 1. Clonogenic cell survival of A549 cells irradiated with 1, 2 or 3 Gy Oxygen beam<br />

or -radiation. Data represents means ± SD of three independent experiments.<br />

rabbit anti-pChk1 (Ser296), rabbit anti-pChk2 (Thr68) and mouse anti-phosphop53<br />

(S15) (all primary antibodies were from Cell Signaling). Secondary antibodies<br />

used were Alexa Fluor 488 goat anti-rabbit IgG and Alexa Fluor 488 goat anti-mouse<br />

IgG (Molecular Probe, USA).<br />

2.5. Image analysis using ImageJ software<br />

The captured images were analyzed for relative quantification of phosphorylation<br />

using ImageJ software [22]. DAPI stained nucleus of each cell was selected and<br />

the same frame was used for relative quantification of phosphorylation using ImageJ<br />

software as described earlier [21]. All the foci were counted manually; at least 100<br />

cells per experiment were analyzed from three independent experiments.<br />

2.6. Statistical analysis<br />

The data were imported to excel work sheets, and graphs were made using<br />

Origin version 5.0. One-way ANOVA with Tukey–Kramer multiple comparisons as<br />

post-test for P < 0.05 used to study the significance level. Data were insignificant at<br />

P > 0.05. Each point represented as mean ± S.D. (standard deviation).<br />

2.3. Clonogenic cell survival assay<br />

After irradiation, clonogenic assay was done as described earlier [21]. Briefly,<br />

after treatment, cells were seeded out in appropriate dilutions (500 cells/60 mm<br />

petriplate) to form colonies in 14 days. Colonies were fixed with glutaraldehyde<br />

(6.0%, v/v), stained with crystal violet (0.5%, w/v) and counted using a stereomicroscope.<br />

Colonies containing more than 50 cells were scored as survivors. Absolute<br />

plating efficiency at 0 Gy (sham irradiated control) was 32.3 ± 2.4%.<br />

2.4. Immunofluorescence staining<br />

Unirradiated and irradiated cells were fixed at different time periods and<br />

immunofluorescence staining was done as described earlier [21]. Briefly, cells were<br />

grown in cover slips which were kept in 35 mm petri plates and irradiated as mentioned<br />

above. Cell layer were washed at different time period after irradiation in PBS<br />

and fixed for 20 min in 4% paraformaldehyde at room temperature. Afterwards, the<br />

cells were washed twice in PBS. For immunofluorescence staining, cells were permeabilized<br />

for 3 min in 0.25% Triton X-100 in PBS, washed two times in PBS and blocked<br />

for 1 h with 5% BSA in PBS. Antibodies were diluted (1:200) in 1% BSA in PBS. Cells<br />

were incubated with primary antibodies for 1 h 30 min at room temperature, washed<br />

three times in PBS and incubated with secondary antibodies for 1 h at room temperature.<br />

Finally, cells were rinsed and mounted with ProLong Gold antifede with DAPI<br />

mounting media (Molecular Probe, USA). Images were captured using Carl Zeiss confocal<br />

microscope. Acquisition settings were optimized to obtain maximal signal in<br />

immunostained cells with minimal background. The primary antibodies used were<br />

mouse anti-pATM (S1981), rabbit anti-pATR (Ser428), rabbit anti--H2AX (S139),<br />

3. Results and discussion<br />

3.1. Oxygen ions were three times more cytotoxic than gamma<br />

To determine the sensitivity of cells to similar dose of high and<br />

low LET radiation, their ability to form colonies after irradiation<br />

was observed. Irradiation of exponentially growing A549 cells with<br />

equal doses of gamma and oxygen beam (614 keV/m) revealed<br />

that at same dose oxygen ions were much more cytotoxic than<br />

gamma rays as assessed by the clonogenic cell survival assay. The<br />

calculated RBE (relative biological effectiveness) for A549 cells irradiated<br />

with 1 Gy oxygen ions particles relative to rays was found<br />

to be nearly 3 at 20% survival. Thus, oxygen ions were found to be<br />

three times more toxic than gamma rays (Fig. 1). The calculated<br />

D 0 value from the plot curve for A549 cells was found to be 0.6 Gy<br />

for oxygen ions and 2.5 Gy for rays. Our results therefore further<br />

support previous studies where high LET oxygen beam have been<br />

found to be much more efficient in killing the tumor cells and hence<br />

may be a preferred mode of treatment for radio resistant tumors.<br />

However, mechanisms of cell death caused by oxygen treatment<br />

and the signaling pathways that ensue have not yet been investi-


192 S. <strong>Ghosh</strong> et al. / Mutation Research 723 (2011) 190–198<br />

Fig. 2. Phosphorylation of H2AX (-H2AX) at different time points after exposure to 2 Gy and 6 Gy gamma or 2 Gy oxygen irradiation in A549 cells. Representative image<br />

showing -H2AX foci (A) 15 min and (B) 4 h after irradiation. Each phospho-H2AX antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and<br />

cells were mounted with ProLong Gold antifede with DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. (C) Graph<br />

represents average numbers of foci per cell, percentage of cells showing the foci is marked above the bars. (D) Graph represents relative intensity of -H2AX as determined<br />

by ImageJ software. At least 100 cells per experiment were analyzed from three independent experiments. Data represents means ± SD of three independent experiments;<br />

significantly different from unirradiated controls: *P < 0.05, **P < 0.01. (For interpretation of the references to colour in this figure legend, the reader is referred to the web<br />

version of this article.)<br />

gated in detail although heavy ion beam therapy is in use for many<br />

cancers [1–4].<br />

3.2. Numbers of initial -H2AX foci formed were almost same for<br />

both types of radiation<br />

Since DNA double strand breaks are the primary toxic lesions<br />

induced by radiation, the number of DSBs were scored by examining<br />

the foci of phosphorylated H2AX in irradiated cells. It has been<br />

established recently that H2AX at the DNA DSB sites is immediately<br />

phosphorylated upon irradiation, and the phosphorylated H2AX<br />

(-H2AX) can be visualized in situ by immunostaining with a -<br />

H2AX specific antibody [23,24]. -H2AX induction can be measured<br />

quantitatively at physiological doses, and the numbers of residual<br />

-H2AX foci can be used to estimate the kinetics of DSB rejoining.<br />

This has become the gold standard for the detection of DSB [16,23].<br />

After 15 min of either radiation, -H2AX foci, visualized as bright<br />

spots, were present in all the cells. The average numbers of foci after<br />

15 min of 2 Gy gamma irradiation were 24 ± 5, while after 2 Gy oxygen<br />

irradiation; the foci were 28 ± 4.7 (Fig. 2(A and C)). Our result<br />

on -H2AX foci detected in cells irradiated with 2 Gy oxygen ion is<br />

contradictory to the fact that number of ion that hits on nucleus and<br />

number of -H2AX foci formed agrees well in heavy-ion irradiated<br />

cells, because heavy-ion densely deposits energy along their trajectories.<br />

Calculating from LET of oxygen ion, it is estimated that the<br />

number of ion hit on each cell nucleus by irradiation of 2 Gy oxygen<br />

ions was approximately 3 particles. However, nearly 10 times of foci<br />

were detected in the cells irradiated with 2 Gy of oxygen ions. This<br />

discrepancy between observed -H2AX foci and estimated -H2AX<br />

foci after high LET radiation has also been reported previously [25].<br />

Thus unlike the cell survival, the difference in induction of -H2AX<br />

foci 15 min after 2 Gy of either radiation was not as profound. We<br />

looked at activation of signaling molecules at dose of 2 Gy of gamma<br />

radiation because the initial response manifested as double strand<br />

breaks, as indicated by -H2AX foci was found to be almost same<br />

at 2 Gy of either radiation. Since the RBE of oxygen ions was nearly<br />

3, it would be more logical to compare 2 Gy of oxygen ion with 6 Gy<br />

of gamma ray. The average numbers of foci after 15 min of 6 Gy<br />

gamma irradiation were 35 ± 3.5. Thus, 6 Gy gamma ray induced<br />

1.25 times more foci than 2 Gy oxygen ions. The cell irradiated with


S. <strong>Ghosh</strong> et al. / Mutation Research 723 (2011) 190–198 193<br />

Fig. 3. Phosphorylation of ATM at 4 h after exposure to 2 Gy and 6 Gy gamma or 2 Gy oxygen irradiation in A549 cells. (A) Representative image showing p-ATM foci 4 h<br />

after irradiation. Each phospho-ATM antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold<br />

antifede with DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph represents average numbers of foci per<br />

cell, percentage of cells showing the foci is marked above the bars. (C) Graph represents relative intensity of p-ATM as determined by ImageJ software. At least 100 cells per<br />

experiment were analyzed from three independent experiments. Data represents means ± SD of three independent experiments; significantly different from unirradiated<br />

controls: *P < 0.05, **P < 0.01. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)<br />

6 Gy of gamma-ray exhibited more -H2AX foci than 2 Gy gammaray<br />

sample. However the increased ratio is only 1.45, and far from<br />

3. This result is contradictory to the fact that induction of DNA double<br />

strand break is physical process, not biological process. Thus<br />

it is expected to increase in proportion with irradiated physical<br />

dose if the source of radiation is same. We have used immunofluorescent<br />

staining assay for -H2AX, one of the limitation of this<br />

assay is the exposures of cells higher than 4 Gy result in foci saturation,<br />

reducing the useful range of the assay to doses less than 4 Gy.<br />

However, major advantage of this assay, it is more sensitive than<br />

other methods for detecting -H2AX [26]. Many downstream signaling<br />

molecules are known to get activated from these irradiation<br />

induced -H2AX foci [27].<br />

The -H2AX foci were also counted 4 h after irradiation to compare<br />

the repair status at that time with respect to radiation quality.<br />

53% of 2 Gy gamma irradiated cells still showed -H2AX foci but the<br />

average foci per cell had reduced to only 4.3 ± 2.2, where as 100% of<br />

6 Gy gamma irradiated cells showed -H2AX foci and the average<br />

foci per cell had reduced to only 8 ± 2.5. In oxygen ion irradiated<br />

cells, although bright green -H2AX intensity was observed till 4 h<br />

in 60% of cells, there was a significant reduction in appearance of<br />

distinct foci like structures (10 ± 2) (Fig. 2(B and C)). Disappearance<br />

of -H2AX foci indicates DNA repair/rejoining. Instead of counting<br />

number of foci, the intensity of -H2AX in cells showing the foci was<br />

estimated by ImageJ software in the same images. In case of 2 Gy<br />

or 6 Gy gamma irradiation, both -H2AX foci and its total intensity<br />

levels had reduced at 4 h (Fig. 2(D)). However, in case of oxygen<br />

irradiation, the total intensity of -H2AX at 4 h was yet very high<br />

(Fig. 2(D)) though foci were not properly visible. This could indicate<br />

diffusion of -H2AX foci rather than its disappearance due to actual<br />

dephosphorylation of -H2AX. Presence of -H2AX even 4 h after<br />

oxygen ion irradiation suggests that DNA had not been repaired.<br />

Similar number of -H2AX at 15 min after irradiation irrespective<br />

of the radiation quality indicated that initial events might<br />

not differ much with the radiation quality. Activation of other<br />

molecules involved in DNA damage response was therefore fol-


194 S. <strong>Ghosh</strong> et al. / Mutation Research 723 (2011) 190–198<br />

Fig. 4. Phosphorylation of ATR at 4 h after exposure to 2 Gy and 6 Gy gamma or 2 Gy oxygen irradiation in A549 cells. (A) Representative image showing p-ATR foci 4 h after<br />

irradiation. Each phospho-ATR antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold antifede with<br />

DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph represents average numbers of foci per cell, percentage<br />

of cells showing the foci is marked above the bars. (C) Graph represents relative intensity of p-ATR as determined by ImageJ software. At least 100 cells per experiment were<br />

analyzed from three independent experiments. Data represents means ± SD of three independent experiments; significantly different from unirradiated controls: *P < 0.05,<br />

**P < 0.01. Significantly different from 6 Gy gamma: ## P < 0.01. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of<br />

this article.)<br />

lowed only at later time point of 4 h. Moreover, by this time low<br />

LET induced damage is mostly repaired and the activation of these<br />

factors is reduced to control levels [28] thereby sealing the fate of<br />

the cell. Comparing the activation of these molecules at this time<br />

could therefore give important insight into differences in damage<br />

response signaling with respect to radiation quality. Phosphorylation<br />

of H2AX even at 4 h indicates the presence of large number<br />

of double strand breaks at that time indicating very slow repair in<br />

A549 cells exposed to oxygen beam.<br />

3.3. Radiation induced foci of DNA damage response proteins<br />

ATM and ATR<br />

ATM is an important DNA damage response protein to ionizing<br />

radiation and a part of irradiation induced foci (IRIF). It gets activated<br />

after DNA DSB by phosphorylation at ser 1981 and activates<br />

many key cellular proteins that include H2AX, BRCA1, Chk1/2 and<br />

p53 [14,29]. Phospho ATM foci and intensity were looked at 4 h<br />

after gamma or oxygen irradiation. At 4 h after 2 Gy gamma irradiation,<br />

most of the cells (93%) cells showed ATM foci (5.9 ± 3.7)<br />

as against 25% unirradiated cells (3.4 ± 1.6) (Fig. 3). Like -H2AX<br />

foci, lesser number of ATM foci 4 h after 2 Gy gamma irradiation<br />

indicates that by 4 h most of the repair due to damage by low LET<br />

irradiation would have taken place. 4 h after 6 Gy gamma irradiation,<br />

100% of the cells showed the foci and average ATM foci per<br />

cell were 14.5 ± 6.05. However, at 4 h after oxygen ion irradiation,<br />

100% of the cells still showed the foci and average ATM foci per cell<br />

(18.6 ± 8.7) were higher than gamma irradiation. Total intensity<br />

levels of phospho ATM matched with the number of foci observed<br />

(Fig. 3). ATM is major mediator of DSB response and is involved<br />

in cell cycle arrest, repair and apoptosis. Persistence of significant<br />

pATM foci till 4 h after oxygen ion irradiation indicates the presence<br />

of unrepaired DNA, which is still keeping the damage sensor<br />

protein ATM in activated form. The slower disappearance of pATM


S. <strong>Ghosh</strong> et al. / Mutation Research 723 (2011) 190–198 195<br />

Fig. 5. Phosphorylation of TP53 at 4 h after exposure to 2 Gy and 6 Gy gamma or 2 Gy oxygen irradiation in A549 cells. (A) Representative image showing p-TP53 4 h after<br />

irradiation. Each phospho-TP53 antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold antifede<br />

with DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph represents relative intensity of p-TP53 as determined<br />

by ImageJ software. At least 100 cells per experiment were analyzed from three independent experiments. Data represents means ± SD of three independent experiments;<br />

significantly different from unirradiated controls: *P < 0.05, **P < 0.01. Significantly different from 6 Gy gamma: # P < 0.05. (For interpretation of the references to colour in this<br />

figure legend, the reader is referred to the web version of this article.)<br />

foci from cells, unlike the -H2AX foci (that was seen in 60% of<br />

oxygen irradiated cells at 4 h), could reflect their different roles in<br />

the repair process with -H2AX foci playing a primary role in activating<br />

downstream pathways and initiating DNA repair while ATM<br />

foci remaining till broken DNA ends are present. Since the cell survival<br />

after 2 Gy oxygen irradiation was only 11%, ATM might be<br />

activating downstream apoptotic responses.<br />

ATR, another PI3 kinase family member, which is known to<br />

respond late to the IR induced damage, was also looked for its<br />

activation at 4 h. After gamma irradiation, average foci per cell<br />

were very less (2.7 ± 0.95 for 2 Gy and 4 ± 1 for 6 Gy) and that<br />

too were visible in only half of the irradiated cells (Fig. 4). On<br />

the other hand, although only 36% of oxygen ion irradiated cells<br />

showed ATR foci, the average foci per cell were (24 ± 14) at least<br />

six times more than that of gamma irradiated cells. Total intensity<br />

levels of phospho ATR matched with the number of foci<br />

observed (Fig. 4). The massive activation of ATR after oxygen<br />

irradiation when compared with gamma (Fig. 4) in 36% of the


196 S. <strong>Ghosh</strong> et al. / Mutation Research 723 (2011) 190–198<br />

Fig. 6. Phosphorylation of Chk2 at 4 h after exposure to 2 Gy and 6 Gy gamma or 2 Gy oxygen irradiation in A549 cells. (A) Representative image showing p-Chk2 4 h<br />

after irradiation. Each phospho-Chk2 antibody was indirectly labeled with Molecular Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold<br />

antifede with DAPI (blue). All images were captured using Carl Zeiss confocal microscope with the same exposure time. (B) Graph represents relative intensity of p-Chk2<br />

as determined by ImageJ software. (C) Graph represents relative intensity of p-Chk1 as determined by ImageJ software At least 100 cells per experiment were analyzed<br />

from three independent experiments. Data represents means ± SD of three independent experiments; significantly different from unirradiated controls: *P < 0.05, **P < 0.01.<br />

Significantly different from 6 Gy gamma: # P < 0.05. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)<br />

cells indicates its specific role in high LET induced complex damage.<br />

3.4. Activation of downstream substrates of ATR and ATM: Chk1,<br />

Chk2 and p53<br />

Increased activation of ATM and ATR in irradiated cells led<br />

us to investigate the activation of further downstream components<br />

including p53, Chk1 and Chk2 which are actual effectors<br />

of the cell cycle arrest or apoptosis [30,31]. There was significantly<br />

higher phosphorylation of p53 at serine 15 in A549<br />

cells that had been exposed to 2 Gy of oxygen beam, but<br />

not after 2 Gy -radiation (Fig. 5). Cells exposed to 6 Gy -<br />

radiation showed marginally higher phosphorylation of p53,<br />

but it was significantly lower than 2 Gy oxygen irradiated<br />

cells (Fig. 5). This along with survival data supports the activation<br />

of p53 dependent apoptotic responses after high LET<br />

radiation.


S. <strong>Ghosh</strong> et al. / Mutation Research 723 (2011) 190–198 197<br />

Fig. 7. Apoptosis in oxygen irradiated A549 cells. Apoptotic nuclei of A549 cells were visualized and quantified by using DAPI staining, 18 h after 2 Gy and 6 Gy gamma or<br />

2 Gy oxygen irradiation. The percentage of cells with apoptotic nuclei are represented as graph. At least 100 cells per experiment were analyzed from three independent<br />

experiments. Data represents means ± SD of three independent experiments; significantly different from unirradiated controls: *P < 0.05, **P < 0.01. Significantly different<br />

from 6 Gy gamma: # P < 0.05.<br />

Chk2, which is activated by ATM and shares its substrates<br />

including p53, was found to be significantly activated in oxygen<br />

irradiated cells. Activation of Chk2 is critical in regulating cell cycle<br />

arrest and DSB repair. Presence of pChk2 foci in nuclei of oxygen<br />

irradiated cells but not gamma indicates Chk2 as an active player<br />

in high LET radiation induced responses (Fig. 6(A and B)). Both ATM<br />

and Chk2 have been regarded as primary regulators of homologous<br />

recombination repair (HRR) [32,33]. Activation of these kinases<br />

after oxygen irradiation indicates involvement of HRR pathway for<br />

repairing complex damage caused by high LET. It is also reasonable<br />

to suppose that cells would prefer to repair a complex damage<br />

by activating the machinery of HRR, which can take the advantage<br />

of the other strand that may be intact. Contribution of HRR after<br />

high LET radiation had also been shown by previous studies [34].<br />

Immunofluorescence of phospho Chk1, which is a major target of<br />

ATR, was also found to increase in nuclei of oxygen-irradiated cells<br />

but not after 2 Gy -radiation (Fig. 6(C)). Cells exposed to 6 Gy -<br />

radiation showed marginally higher phosphorylation of Chk1, but it<br />

was significantly lower than 2 Gy oxygen irradiated cells (Fig. 6(C)).<br />

Activation of Chk2 and Chk1 after oxygen irradiation also indicates<br />

cell cycle arrest.<br />

3.5. Apoptosis<br />

At 18 h after irradiation, morphological features of DAPI stained<br />

apoptotic nuclei were scored and many apoptotic nuclei were<br />

observed in cells irradiated with 2 Gy oxygen ion and 6 Gy gamma<br />

but not in 2 Gy gamma irradiated cells (Fig. 7). Cells exposed to<br />

2 Gy oxygen ion showed significantly higher apoptotic nuclei than<br />

6 Gy gamma (Fig. 7). The degree of radiation-induced apoptosis has<br />

been shown to correlate with the p53 wild-type status [35]. In addition,<br />

apoptosis is induced when wild-type p53 is transfected into<br />

certain cell lines lacking p53 [36]. Activation of p53 in oxygen irradiated<br />

cells along with observance of apoptotic cells indicates p53<br />

dependent apoptosis in these cells. Despite significant activation<br />

of ATM, ATR, Chk1 and Chk2 in A549 cells exposed to 2 Gy oxygen<br />

beam, DNA was not repaired. This further point to the involvement<br />

of these kinases in activation of apoptotic machinery after high LET<br />

irradiation.<br />

4. Conclusion<br />

From a critical analysis of all the results of the present study,<br />

it will be evident that in general, high LET radiations cause more<br />

severe genomic damage as compared to low LET radiations. The<br />

noteworthy finding of this study is the activation of the sensor<br />

proteins, ATM and ATR by oxygen irradiation and the significant<br />

activation of Chk1, Chk2 and p53 only in the oxygen beam irradiated<br />

cells. These kinases play an important role in repair, cell cycle<br />

arrest and apoptosis. However, only 11% survival after high LET<br />

radiation suggests that most of the damage could not be repaired<br />

and that the persistent activation of these kinases may be a signa-


198 S. <strong>Ghosh</strong> et al. / Mutation Research 723 (2011) 190–198<br />

ture for activation of apoptotic responses. Further studies in this<br />

direction will reveal the importance of such response.<br />

The use of oxygen beam therapy has definite advantage over<br />

gamma therapy and the understanding of the mechanism of apoptosis<br />

will give the clinician a handle to manipulate therapy for<br />

enhanced cell killing.<br />

Conflict of interest<br />

The authors declare that there are no conflicts of interest.<br />

Acknowledgements<br />

This work was funded by Board of Research in Nuclear Sciences,<br />

Department of Atomic Energy [DAE], Government of India,<br />

through a project sanctioned to one of the authors, AS, bearing sanction<br />

number 2007/37/37/BRNS. All the authors are thankful to the<br />

Director, IUAC, New Delhi for providing radiation facility for the<br />

work. We would also like to extend our thanks to all the members<br />

of Pelletron group of IUAC for their sincere help during irradiation.<br />

Authors would like to thank Mr. Manjoor Ali and Mr. Paresh<br />

Khadilkar, RB&HSD, BARC for their help in Confocal Microscopy and<br />

Lab work, respectively.<br />

References<br />

[1] T. Miyamoto, N. Yamamoto, H. Nishimura, M. Koto, H. Tsujii, J.E. Mizoe, T.<br />

Kamada, H. Kato, S. Yamada, S. Morita, K. Yoshikawa, S. Kandatsu, T. Fujisawa,<br />

Carbon ion radiotherapy for stage I non-small cell lung cancer, Radiother. Oncol.<br />

66 (2003) 127–140.<br />

[2] T. Ogata, T. Teshima, K. Kagawa, Y. Hishikawa, Y. Takahashi, A. Kawaguchi,<br />

Y. Suzumoto, K. Nojima, Y. Furusawa, N. Matsuura, Particle irradiation<br />

suppresses metastatic potential of cancer cells, Cancer Res. 65 (2005)<br />

113–120.<br />

[3] D. Schulz-Ertner, A. Nikoghosyan, C. Thilmann, T. Haberer, O. Jakel, C. Karger, G.<br />

Kraft, M. Wannenmacher, J. Debus, Results of carbon ion radiotherapy in 152<br />

patients, Int. J. Radiat. Oncol. Biol. Phys. 58 (2004) 631–640.<br />

[4] H. Tsuji, H. Ishikawa, T. Yanagi, N. Hirasawa, T. Kamada, J.E. Mizoe, T. Kanai, H.<br />

Tsujii, Y. Ohnishi, Carbon-ion radiotherapy for locally advanced or unfavorably<br />

located choroidal melanoma: a phase I/II dose-escalation study, Int. J. Radiat.<br />

Oncol. Biol. Phys. 67 (2007) 857–862.<br />

[5] M. Durante, F.A. Cucinotta, Heavy ion carcinogenesis and human space exploration,<br />

Nat. Rev. Cancer 8 (2008) 465–472.<br />

[6] K. Shibuya, C.D. Mathers, C. Boschi-Pinto, A.D. Lopez, C.J. Murray, Global and<br />

regional estimates of cancer mortality and incidence by site. II. Results for the<br />

global burden of disease 2000, BMC Cancer 2 (2002) 37.<br />

[7] A. Jemal, R. Siegel, E. Ward, Y. Hao, J. Xu, T. Murray, M.J. Thun, Cancer statistics2008,<br />

CA Cancer J. Clin. 58 (2008) 71–96.<br />

[8] J. Ferlay, P. Autier, M. Boniol, M. Heanue, M. Colombet, P. Boyle, Estimates of<br />

the cancer incidence and mortality in Europe in 2006, Ann. Oncol. 18 (2007)<br />

581–592.<br />

[9] J. Jassem, The role of radiotherapy in lung cancer: where is the evidence?<br />

Radiother. Oncol. 83 (2007) 203–213.<br />

[10] Y.S. Kim, S.M. Yoon, E.K. Choi, B.Y. Yi, J.H. Kim, S.D. Ahn, S.W. Lee, S.S. Shin, J.S.<br />

Lee, C. Suh, S.W. Kim, D.S. Kim, W.S. Kim, H.J. Park, C.I. Park, Phase II study of<br />

radiotherapy with three-dimensional conformal boost concurrent with paclitaxel<br />

and cisplatin for Stage IIIB non-small-cell lung cancer, Int. J. Radiat. Oncol.<br />

Biol. Phys. 62 (2005) 76–81.<br />

[11] D. Farray, N. Mirkovic, K.S. Albain, Multimodality therapy for stage III nonsmall-cell<br />

lung cancer, J. Clin. Oncol. 23 (2005) 3257–3269.<br />

[12] M. Koto, T. Miyamoto, N. Yamamoto, H. Nishimura, S. Yamada, H. Tsujii, Local<br />

control and recurrence of stage I non-small cell lung cancer after carbon ion<br />

radiotherapy, Radiother. Oncol. 71 (2004) 147–156.<br />

[13] T.K. Pandita, H.B. Lieberman, D.S. Lim, S. Dhar, W. Zheng, Y. Taya, M.B. Kastan,<br />

Ionizing radiation activates the ATM kinase throughout the cell cycle, Oncogene<br />

19 (2000) 1386–1391.<br />

[14] M.B. Kastan, D.S. Lim, The many substrates and functions of ATM, Nat. Rev. Mol.<br />

Cell. Biol. 1 (2000) 179–186.<br />

[15] R.S. Maser, K.J. Monsen, B.E. Nelms, J.H. Petrini, hMre11 and hRad50 nuclear foci<br />

are induced during the normal cellular response to DNA double-strand breaks,<br />

Mol. Cell. Biol. 17 (1997) 6087–6096.<br />

[16] O. Fernandez-Capetillo, A. Lee, M. Nussenzweig, A. Nussenzweig, H2AX: the<br />

histone guardian of the genome, DNA Repair (Amst.) 3 (2004) 959–967.<br />

[17] S. Bekker-Jensen, C. Lukas, R. Kitagawa, F. Melander, M.B. Kastan, J. Bartek, J.<br />

Lukas, Spatial organization of the mammalian genome surveillance machinery<br />

in response to DNA strand breaks, J. Cell. Biol. 173 (2006) 195–206.<br />

[18] J.F. Ziegler, The Stopping and Range of Ions in Matter, http://www.srim.org/.<br />

[19] J.F. Ziegler, J.P. Bierserk, U. Littmark, Stopping and Ranges of Ions in Matter,<br />

Pergamon Press, New York, 1985.<br />

[20] G. Kraft, Tumor therapy with heavy charged particles, Prog. Part. Nucl. Phys. 45<br />

(2000) S473–S544.<br />

[21] S. <strong>Ghosh</strong>, N.N. Bhat, S. Santra, R.G. Thomas, S.K. Gupta, R.K. Choudhury, M.<br />

Krishna, Low energy proton beam induces efficient cell killing in A549 lung<br />

adenocarcinoma cells, Cancer Invest. 28 (2010) 615–622.<br />

[22] W.S. Rasband, J.U.S. Image, <strong>National</strong> Institutes of Health, Bethesda, USA,<br />

1997–2006, http://rsb.info.nih.gov/ij/.<br />

[23] E.P. Rogakou, D.R. Pilch, A.H. Orr, V.S. Ivanova, W.M. Bonner, DNA doublestranded<br />

breaks induce histone H2AX phosphorylation on serine 139, J. Biol.<br />

Chem. 273 (1998) 5858–5868.<br />

[24] P.L. Olive, J.P. Banath, Phosphorylation of histone H2AX as a measure of<br />

radiosensitivity, Int. J. Radiat. Oncol. Biol. Phys. 58 (2004) 331–335.<br />

[25] A. Asaithamby, N. Uematsu, A. Chatterjee, M.D. Story, S. Burma, D.J. Chen, Repair<br />

of HZE-particle-induced DNA double-strand breaks in normal human fibroblasts,<br />

Radiat. Res. 169 (2008) 437–446.<br />

[26] D. Avondoglio, T. Scott, W.J. Kil, M. Sproull, P.J. Tofilon, K. Camphausen, High<br />

throughput evaluation of gamma-H2AX, Radiat. Oncol. 4 (2009) 31.<br />

[27] T.T. Paull, E.P. Rogakou, V. Yamazaki, C.U. Kirchgessner, M. Gellert, W.M. Bonner,<br />

A critical role for histone H2AX in recruitment of repair factors to nuclear foci<br />

after DNA damage, Curr. Biol. 10 (2000) 886–895.<br />

[28] H. Wang, Z.C. Zeng, T.A. Bui, S.J. DiBiase, W. Qin, F. Xia, S.N. Powell, G. Iliakis, Nonhomologous<br />

end-joining of ionizing radiation-induced DNA double-stranded<br />

breaks in human tumor cells deficient in BRCA1 or BRCA2, Cancer Res. 61 (2001)<br />

270–277.<br />

[29] E.U. Kurz, S.P. Lees-Miller, DNA damage-induced activation of ATM and ATMdependent<br />

signaling pathways, DNA Repair (Amst.) 3 (2004) 889–900.<br />

[30] N.D. Lakin, S.P. Jackson, Regulation of p53 in response to DNA damage, Oncogene<br />

18 (1999) 7644–7655.<br />

[31] T. Rich, R.L. Allen, A.H. Wyllie, Defying death after DNA damage, Nature 407<br />

(2000) 777–783.<br />

[32] J. Zhang, H. Willers, Z. Feng, J.C. <strong>Ghosh</strong>, S. Kim, D.T. Weaver, J.H. Chung, S.N.<br />

Powell, F. Xia, Chk2 phosphorylation of BRCA1 regulates DNA double-strand<br />

break repair, Mol. Cell. Biol. 24 (2004) 708–718.<br />

[33] S.E. Golding, E. Rosenberg, A. Khalil, A. McEwen, M. Holmes, S. Neill, L.F. Povirk,<br />

K. Valerie, Double strand break repair by homologous recombination is regulated<br />

by cell cycle-independent signaling via ATM in human glioma cells, J. Biol.<br />

Chem. 279 (2004) 15402–15410.<br />

[34] F. Zafar, S.B. Seidler, A. Kronenberg, D. Schild, C. Wiese, Homologous recombination<br />

contributes to the repair of DNA double-strand breaks induced by<br />

high-energy iron ions, Radiat. Res. 173 (2010) 27–39.<br />

[35] P. Fei, E.J. Bernhard, W.S. El-Deiry, Tissue-specific induction of p53 targets in<br />

vivo, Cancer Res. 62 (2002) 7316–7327.<br />

[36] E. Yonish-Rouach, D. Resnitzky, J. Lotem, L. Sachs, A. Kimchi, M. Oren, Wildtype<br />

p53 induces apoptosis of myeloid leukaemic cells that is inhibited by<br />

interleukin-6, Nature 352 (1991) 345–347.


Cancer Investigation, 28:615–622, 2010<br />

ISSN: 0735-7907 print / 1532-4192 online<br />

Copyright c○ Informa Healthcare USA, Inc.<br />

DOI: 10.3109/07357901003630991<br />

ORIGINAL ARTICLE<br />

Cellular and Molecular Biology<br />

Low Energy Proton Beam Induces Efficient Cell<br />

Killing in A549 Lung Adenocarcinoma Cells<br />

<strong>Somnath</strong> <strong>Ghosh</strong>, 1 Nagesh N. Bhat, 2 S. Santra, 3 R. G. Thomas, 3 S.K. Gupta, 3 R.K. Choudhury, 3 and Malini Krishna 1<br />

Radiation Biology and Health Sciences Division, 1 Radiological Physics and Advisory Division, 2 Nuclear Physics Division, <strong>Bhabha</strong> Atomic<br />

Research Centre, Trombay, Mumbai, India 3<br />

Cancer Invest Downloaded from informahealthcare.com by University of Chicago Library on 06/10/11<br />

For personal use only.<br />

ABSTRACT<br />

The aim of the current study was to determine the signaling differences between γ- and<br />

proton beam-irradiations. A549 lung adenocarcinoma cells were irradiated with 2 Gy proton<br />

beam or γ-radiation. Proton beam was found to be more cytotoxic than γ-radiation. Proton<br />

beam-irradiated cells showed phosphorylation of H2AX, ATM, Chk2, and p53. The mechanism<br />

of excessive cell killing in proton beam-irradiated cells was found to be upregulation of Bax<br />

and downregulation of Bcl-2. The noteworthy finding of this study is the biphasic activation of<br />

the sensor proteins, ATM, and DNA-PK and no activation of ATR by proton irradiation.<br />

INTRODUCTION<br />

Exposureof mammalian cells to ionizing radiation leads to<br />

the activation of several signaling pathways that result in apoptosis.<br />

Although, the end point, apoptosis is effectively activated<br />

by conventional X-ray treatment, the development of radioresistance<br />

following therapy remains a major cause for concern.<br />

Clinicians are now favoring proton beam therapy to avoid the<br />

issue (1–6). However, the mechanism of cell killing by proton<br />

beam is not yet well delineated.<br />

Lung cancer is a frequently occurring, mostly lethal disease<br />

in all countries worldwide (7). It is one of the most commonly<br />

diagnosed types of cancer and has the highest death rate (8, 9).<br />

Overall, fewer than 10% of people with primary lung cancer<br />

are alive 5 years after diagnosis. Depending on tumor size, location,<br />

and histology, several treatment options are available,<br />

including surgery and radiotherapy (RT), both often combined<br />

with chemotherapy (10). A major problem remains local tumor<br />

control (11, 12). Therefore, new ways to deliver radiotherapy<br />

beyond photons have been sought for, including protons (13, 14)<br />

Keywords Proton beam, ATM, p53, Bax, Bcl-2<br />

Correspondence to:<br />

Malini Krishna<br />

Radiation Biology and Health Sciences Division<br />

<strong>Bhabha</strong> Atomic Research Centre<br />

Trombay, Mumbai<br />

India, 400085<br />

email: malini@barc.gov.in; malinik00@gmail.com<br />

and carbon ions (15). Furthermore, treatment with charged particle<br />

radiation has several potential advantages over treatment<br />

with γ or X-ray radiation such as the Bragg peak enables precise<br />

localization of the radiation dose, an inverted depth-dose<br />

distribution, a higher relative biological effectiveness (RBE),<br />

and a lower cellular capability for repair of radiation injury.<br />

Because of their superior dose distribution, a therapeutic gain<br />

can be expected with charged particles (1–4). Although proton<br />

beams are currently being used for the treatment of many cancers<br />

(1–6), the mechanism of cell killing and its variance from<br />

γ -irradiation is not well understood. Understanding the specific<br />

biological effects of charged particle radiation on cancer cells<br />

could also provide valuable insights for the design of novel<br />

therapeutic applications for the treatment of cancers, which are<br />

resistant to many types of therapies. Moreover, therapies can be<br />

designed only if the signaling pathway is understood.<br />

A cytological manifestation of nuclear activity in response<br />

to ionizing radiation (IR) is the formation of the socalled<br />

IR-induced foci (IRIF) (16). IRIFs are dynamic, microscopically<br />

discernible structures containing thousands of<br />

copies of proteins, including γ H2AX, ataxia telangiectasia<br />

mutated (ATM), BRCA1, 53BP1, MDC1, RAD51, and the<br />

MRE11/RAD50/NBS1 (MRN) complex, which accumulate in<br />

the vicinity of a DNA double-strand breaks (DSB) (17, 18). It is<br />

being appreciated lately that DNA repair pathways activated by<br />

charged particle radiation might be different from those studied<br />

after γ -irradiation.<br />

When exposed to ionizing radiation, eukaryotic cells also<br />

activate checkpoint pathways to delay the progression of the<br />

615


Cancer Invest Downloaded from informahealthcare.com by University of Chicago Library on 06/10/11<br />

For personal use only.<br />

cell cycle. In case of charged particle irradiation a prolonged to the primary beam, after the gold foil, served as a monitor<br />

cell cycle arrest (19) and a slower rejoining of DSB have been detector. The ratio of the monitor detector counts to that of the<br />

reported to occur (20).<br />

flux measured using SSB detector at the sample position was<br />

Apart from DNA repair pathways, ionizing radiation leads to measured by multiple trials and the calibration factor was obtained.<br />

Monitor detector counts and the measured ratio were<br />

activation of intracellular signaling pathways that play an important<br />

role in cellular decisions of death or survival. Although used for delivering the required fluence to the samples. Flux<br />

a fair amount of literature exists for the activation of signaling was calculated to be 10 8 and the fluence was kept such that each<br />

pathway after charged particle irradiation, there has been no cell is hit by at least 100 proton particles ensuring that bystander<br />

systematic comparison of photon and proton beam irradiation. effect is kept to minimum. Fluence was then used to calculate<br />

In this study, A549 cells were irradiated with either 2 Gy dose with the help of specific ionization curve constructed for<br />

proton beam- or 2 Gy γ -radiation and ensuing activation of few the given energy distribution and composition of the cell line.<br />

important components involved in DNA repair, cell cycle arrest, The average energy of the proton beam used in this study was<br />

and apoptosis pathways were followed to elucidate the mechanisms<br />

behind observed cellular response. Despite the variability 12.5 keV/µm within the cell layer. Initial portion of Bragg’s<br />

measured to be 3.2 MeV and corresponding estimated LET is<br />

in effectiveness of these two radiations, equal doses of each were curve was used for dose estimation since the cell layer used for<br />

chosen to ensure that each cell is hit by proton beam and no bystander<br />

effects are manifested. Essentially all biological differ-<br />

10 µm, which helped to avoid steep increase in LET within the<br />

irradiation was in monolayer geometry with thickness less than<br />

ences between proton and γ -radiations arise from differences sample.<br />

in their ionization tracks. Equitoxic doses were not preferred The cells were grown in monolayer geometry on 35 mm<br />

because the probability of unhit cells would increase. How this petri dishes. Media was removed and petri dishes were mounted<br />

qualitative difference of radiation is translated into activation of on the exit window vertically for the irradiation, after which<br />

crucial intracellular pathways was the focus of the study. Only they were resupplemented with the medium. On an average<br />

one cell line was chosen in this study because the response of samples were irradiated for 1 min to get the required fluence. The<br />

different cell lines is different to proton irradiation (21). cells were exposed to 2 Gy dose with the help of monitor detector<br />

counts.<br />

MATERIALS AND METHODS<br />

Cell culture<br />

Clonogenic cell survival assay<br />

After treatment, cells were seeded out in appropriate dilutions<br />

(500 cells/60 mm petriplate) to form colonies in 14 days.<br />

Human A549 cells were cultured in Dulbecco’s modified eagles<br />

medium (Sigma) supplemented with 10% fetal calf serum<br />

(Sigma). Cells were maintained at 37 ◦ Colonies were fixed with glutaraldehyde (6.0% v/v), stained<br />

C in humidified atmosphere<br />

with 5% CO 2 . Cells were seeded at density of 5 × 10 5 with crystal violet (0.5% w/v) and counted using a stereomicroscope.<br />

Colonies containing more than 50 cells were scored as<br />

cells in small 35 mm petri dishes 24 hr before the irradiation.<br />

survivors. Absolute plating efficiency at 0 Gy (sham irradiated<br />

control) was 32.3 ± 2.4%<br />

Treatment of cells<br />

Irradiation<br />

Semiquantitative reverse transcriptional<br />

For γ -irradiation, cells were exposed to 2 Gy of 60 Co γ -<br />

polymerase chain reaction (RT-PCR)<br />

radiation in in-house facility Gamma Cell 220 at dose rate of<br />

3 Gy/min.<br />

Total RNA from A549 cells (1 × 10 6 cells) was extracted after<br />

Proton beam irradiation was carried out using the Radiation required time periods of irradiation and RT-PCR was carried out<br />

Biology beam line of in-house 4 MeV Folded Tandem Ion Accelerator<br />

(FOTIA) facility. The primary proton beam from the was extracted after required time periods of 1, 2, 3, and 4 hr of<br />

as described earlier (22). Breifly, total RNA from A549 cells<br />

FOTIA was collimated using an adjustable slit to reduce the irradiation and eluted in 30 µL RNAase free water using RNA<br />

fluence and then diffused using a gold foil. The diffused beam tissue kit (Roche). Equal amount of RNA in each group was<br />

was channeled to the exit window made of 20 µm titanium foil reverse transcribed using cMaster RT kit (Eppendorf). Equal<br />

of 3-cm diameter to get uniformly distributed irradiation area. amount (2 µL) of cDNA in each group was used for specific<br />

The samples were thus irradiated under normal atmospheric amplification of β-actin, ATM, DNA-PK, ATR, Chk1, Chk2,<br />

pressure at 24 ◦ C. The target to be irradiated was positioned at a Bcl-2, or Bax gene using gene specific primers (Table 1). The<br />

distance of 11 mm from the exit window, that being the closest PCR condition were 94 ◦ C, 5 min initial denaturation followed<br />

possible place to mount the target. Before the irradiation, a silicon<br />

surface barrier (SSB) detector was positioned at the same annealing temperature of specific primers given in Table 1),<br />

by 30 cycles of 94 ◦ C 45 s, 55–64 ◦ C 45 s (depending on the<br />

position where samples were irradiated and the beam energy 72 ◦ C 45 s and final extension at 72 ◦ C for 10 min. Equal amount<br />

as well as the uniformity was measured. Another SSB detector of each PCR product (10 µL) was run on 1.5% agarose gel<br />

placed inside the scattering chamber at a forward angle of 40 ◦ containing ethidium bromide in tris borate EDTA buffer at 60 V.<br />

616 S. <strong>Ghosh</strong> et al.


Cancer Invest Downloaded from informahealthcare.com by University of Chicago Library on 06/10/11<br />

For personal use only.<br />

Table 1. The Sequences of Gene Specific Primers and Their<br />

Annealing Temperature (Tm) Used in RT-PCR Experiments<br />

Gene Primer Sequence Tm ( ◦ C)<br />

β-actin F 5 ′ -TGGAATCCTGTGGCATCCATGAAA-3 ′ 55<br />

R5 ′ -TAAAACGCAGCTCAGTAACAGTCCG-3 ′<br />

ATM F 5 ′ -GGACAGTGGAGGCACAAAAT-3 ′ 57<br />

R5 ′ -GTGTCGAAGACAGCTGGTGA-3 ′<br />

ATR F 5 ′ -CTCGCTGAACTGTACGTGGA-3 ′ 57<br />

R5 ′ -GCATAGCTCGACCATGGATT-3 ′<br />

DNA-PK F 5 ′ -TGCCAATCCAGCAGTCATTA-3 ′ 64<br />

R5 ′ -GGTCCATCAGGCACTTCACT-3 ′<br />

Chk2 F 5 ′ -GGCTTCAGGATGAAGACATGA-3 ′ 64<br />

R5 ′ -CACAACACAGCAGCACACAC-3 ′<br />

Chk1 F 5 ′ -TGTCAGAGTCTCCCAGTGGA-3 ′ 59<br />

R5 ′ -AGGGGCTGGTATCCCATAAG-3 ′<br />

Bax F 5 ′ -TCCCCCCGAGAGGTCTTT-3 ′ 56<br />

R5 ′ -CGGCCCCAGTTGAAGTTG-3 ′<br />

Bcl-2 F 5 ′ -GAGGATTGTGGCCTTCTTTG-3 ′ 59<br />

R5 ′ -ACAGTTCCACAAAGGCATCC-3 ′<br />

F indicates forward and R indicates reverse.<br />

The bands in the gel were visualized under UV lamp and relative<br />

intensities were quantified using Gel doc software (Syngene).<br />

Immunofluorescence staining<br />

Cells were grown in cover slips that were kept in 35 mm<br />

petri dishes and irradiated as mentioned above. Cell layers were<br />

washed 4 h after irradiation in PBS and fixed for 20 min in 4%<br />

paraformaldehyde at room temperature. Afterward, the cells<br />

were washed twice in PBS. For immunofluorescence staining,<br />

cells were permeabilized for 3 min in 0.25% Triton X-100 in<br />

PBS, washed two times in PBS, and blocked for 1 hr with 5%<br />

BSA in PBS. Antibodies were diluted (1:200) in 1% BSA in<br />

PBS. Cells were incubated with primary antibodies for 1 hr<br />

30 min at room temperature, washed three times in PBS, and<br />

incubated with secondary antibodies for 1 hr at room temperature.<br />

Finally, cells were rinsed and mounted with ProLong<br />

Gold antifede with DAPI mounting media (Molecular Probe,<br />

USA). Images were captured using Carl Zeiss confocal microscope.<br />

Acquisition settings were optimized to obtain maximal<br />

signal in immunostained cells with minimal background. The<br />

primary antibodies used were mouse anti-pATM (S1981), rabbit<br />

anti-γ -H2AX (S139), rabbit anti-pChk2 (Thr68), and mouse<br />

anti-phospho-p53 (S15) (Cell Signaling). Secondary antibodies<br />

used were Alexa Fluor 488 goat antirabbit IgG and Alexa Fluor<br />

488 goat antimouse IgG (Molecular Probe, USA).<br />

Image analysis using ImageJ software<br />

The captured images were analyzed for relative quantification<br />

of phosphorylation using ImageJ software (23). A 25 × 25<br />

pixel box was positioned over the fluorescent image of each cell,<br />

and the average intensity within the selection (the sum of the<br />

intensities of all the pixels in the selection divided by the number<br />

of pixels) was measured. At least 100 cells per experiment<br />

were analyzed from three independent experiments.<br />

Statistical analysis<br />

The data were imported to excel worksheets, and graphs<br />

were made using Origin version 5.0. One-way ANOVA with<br />

Tukey–Kramer Multiple Comparisons as post-test for p < .05<br />

used to study the significant level. Data were insignificant at<br />

p > .05. Each point represented as mean ± S.E. (standard error<br />

of the mean).<br />

RESULTS<br />

Cell survival assay<br />

There were only 23 ± 2% of A549 cells that survived which<br />

had been exposed to 2 Gy of proton beam where as 56 ± 2.2%<br />

cells were survived which had been exposed to 2 Gy of γ -<br />

radiation as observed with clonogenic assay(Figure 1). This<br />

result clearly showed that proton beam is more cytotoxic than<br />

γ -radiation.<br />

Differential modulation of gene expression<br />

following proton beam and γ-irradiation<br />

Expression of ATM, DNA-PK, and Chk2 genes were significantly<br />

upregulated after 4 hr of irradiation only in A549<br />

cells which had been exposed to proton beam (2 Gy) [Figures<br />

2(A) and (C)], whereas only the ATR gene was significantly<br />

upregulated after 4 hr of irradiation in A549 cells which had<br />

been exposed to γ -radiation (2 Gy) [Figure 2(B)]. There was<br />

no change in the gene expression of Chk1 in A549 cells which<br />

had been exposed to either γ -radiation or proton beam (2 Gy)<br />

[Figures 2(C), and (D)]. Independent experiments were carried<br />

out for γ - and proton beam-irradiated cells, so the control group<br />

of γ - and proton-irradiated cells may have different transcript<br />

level. However, the data were normalized for the respective controls.<br />

Activation of signaling components involved in<br />

DNA damage sensing and repair following<br />

proton beam irradiation<br />

SinceATM protein plays a central role in DNA double-strand<br />

breaks (DSBs) sensing and repair and determines the cellular<br />

Figure 1. Clonogenic Cell Survival of A549 cells irradiated with 2<br />

Gy proton beam or 2 Gy γ-radiation. Data shown from representative<br />

experiment of three independent experiments.<br />

Efficient Cell killing by Proton Beam 617


Cancer Invest Downloaded from informahealthcare.com by University of Chicago Library on 06/10/11<br />

For personal use only.<br />

Figure 2. Gene expression of ATM, DNA-PK, ATR, Chk1, and Chk2 in A549 cells irradiated with 2 Gy proton beam or 2 Gy γ-radiation at 2, 3,<br />

or 4 hr after irradiation. Total RNA from A549 cells was isolated and reverse transcribed. RT-PCR analysis of ATM, DNA-PK, ATR, Chk1, and<br />

Chk2 genes was carried out as described in materials and methods. PCR products were resolved on 1.5% agarose gels containing ethidium<br />

bromide. β-actin gene expression in each group was used as an internal control. Ratio of intensities of (A and B) ATM, DNA-PK, or ATR and (C<br />

and D) Chk1 and Chk2 band to that of respective β-actin band as quantified from gel pictures are shown above each gel picture. Con means<br />

unirradiated control. Data represents means ± SE of three independent experiments, significantly different from unirradiated controls. ∗ p < .05,<br />

∗∗ p < .01.<br />

response in eukaryotes (24). It was of interest to look for its<br />

phosphorylation and its substrates at 4 hr after irradiation. There<br />

was significantly higher phosphorylation of ATM at serine 1981,<br />

H2AX at γ -position, Chk2 at threonine 68, and p53 at serine 15<br />

in A549 cells which had been exposed to 2 Gy of proton beam,<br />

but their phosphorylation was not found to be increased 4 hr<br />

after 2 Gy of γ -radiation as compared to the control (Figure 3).<br />

Expression of apoptosis-related genes<br />

Since there was significantly higher phosphorylation of p53<br />

in A549 cells which had been exposed to 2 Gy proton beam<br />

(Figure 3), it was our interest to look for the expression of<br />

apoptosis-related gene like Bax and Bcl-2.<br />

In our study, there was significant upregulation of proapoptotic<br />

gene, Bax, 12 hr after irradiation in A549 cells, which had<br />

been exposed to 2 Gy proton beam as compared to the unirradiated<br />

control. In γ -irradiated (2 Gy) cells there was no significant<br />

upregulation of the Bax gene (p < .05) (Figure 4).<br />

On the other hand there was significant downregulation of<br />

antiapoptotic gene, Bcl-2, 12 hr after irradiation in A549 cells,<br />

which had been exposed to 2 Gy proton beam. However, following<br />

γ -irradiation there was no significant downregulation of<br />

Bcl-2 gene. (p < .05) (Figure 4).<br />

618 S. <strong>Ghosh</strong> et al.


Cancer Invest Downloaded from informahealthcare.com by University of Chicago Library on 06/10/11<br />

For personal use only.<br />

Figure 3. Immunofluorescence staining and Image analysis of γ-H2AX, phospho-ATM, phospho-Chk2, and phospho-p53 4 hr after irradiation in<br />

A549 cells irradiated with 2 Gy proton beam or 2 Gy γ-radiation. Cells were fixed in 4% paraformaldehyde, permeabilized in 0.25% Triton X-100,<br />

and labeled with specific antibodies as described in “Materials and Methods.” Each phospho-site antibody was indirectly labeled with Molecular<br />

Probe 488 secondary antibody (green) and cells were mounted with ProLong Gold antifede with DAPI (blue). All images were captured using<br />

Carl Zeiss confocal microscope with the same exposure time. (A)–(D) shows the comparison of phosphorylation between 2 Gy proton beam<br />

and 2 Gy γ-radiation of [(A) H2AX, (B) ATM, (C) Chk2, (D) p53]. (E) Graph represents relative phosphorylation of H2AX, ATM, Chk2, and p53 as<br />

determined by ImageJ software. At least 100 cells per experiment were analyzed from three independent experiments. Data represents means<br />

± SE of three independent experiments, significantly different from unirradiated controls. ∗ p < .05, ∗∗ p < .01.<br />

Efficient Cell killing by Proton Beam 619


Cancer Invest Downloaded from informahealthcare.com by University of Chicago Library on 06/10/11<br />

For personal use only.<br />

Figure 4. Gene expression of Bax and Bcl-2 12 hr after irradiation<br />

in A549 cells irradiated with 2 Gy proton beam or 2 Gy γ-radiation.<br />

Total RNA from A549 cells was isolated and reverse transcribed.<br />

RT-PCR analysis of Bax and Bcl-2 genes was carried out as described<br />

in materials and methods. PCR products were resolved on<br />

1.5% agarose gels containing ethidium bromide. β-actin gene expression<br />

in each group was used as an internal control. Ratio of<br />

intensities of Bax and Bcl-2 band to that of respective β-actin band<br />

as quantified from gel pictures are shown above each gel picture.<br />

Con means unirradiated control, H means proton. Data represents<br />

means ± SE of three independent experiments, significantly different<br />

from unirradiated controls. ∗ p < .05, ∗∗ p < .01.<br />

DISCUSSION<br />

More than two-fold decrease in survival of the A549 cells<br />

as assessed by the clonogenic cell survival assay is indicative<br />

of the fact that the proton beam is more efficient in killing the<br />

tumor cells and may be a more preferred mode of treatment.<br />

The mechanisms of cell death caused by proton treatment and<br />

the signaling pathways that ensue have not yet been investigated<br />

in detail although proton beam therapy is in use for many<br />

cancers (1–6).<br />

The massive activation of DNA-PK and ATM after proton<br />

irradiation when compared with γ [Figures 2, (A) & (B)] was<br />

more or less expected and may be due to the fact that DNA damage<br />

activates phosphatidylinositol 3-kinase-like kinases (DNA-<br />

PK, ATM), which then amplify and channel the signal by activating<br />

downstream kinases (Chk1 and Chk2). These, in turn,<br />

phosphorylate target proteins such as p53, Cdc25A, and Cdc25C<br />

(25) to delay the cell cycle, a process called the DNA damage<br />

checkpoint. In addition to checkpoint control, the downstream<br />

kinases also modulate DNA repair and trigger apoptosis (26).<br />

The lack of activation of ATR after proton irradiation was however<br />

intriguing and was found to be reflecting no activation of<br />

Chk1 (Figure 2).<br />

The phosphorylation of H2AX was found to significantly<br />

increase 4 hr after proton beam irradiation (Figure 3). This<br />

time period of 4 hr was chosen for looking at the activation of<br />

important components involved in repair because by 4 hr after<br />

γ -irradiation most of the repair is complete and the activation of<br />

these factors is reduced to control levels (27). Phosphorylation<br />

of H2AX even at 4 hr indicates the presence of large number<br />

of DSBs at that time indicating very slow repair. It has been<br />

established recently that H2AX at the DNA DSB sites is immediately<br />

phosphorylated upon irradiation, and the phosphorylated<br />

H2AX (γ -H2AX) can be visualized in situ by immunostaining<br />

with a γ -H2AX specific antibody (28, 29). γ -H2AX induction<br />

can be measured quantitatively at physiological doses, and the<br />

numbers of residual γ -H2AX foci can be used to estimate the<br />

kinetics of DSB rejoining. This has become the gold standard<br />

for the detection of DSB (17, 28).<br />

Since ATM protein plays a central DNA DSBs sensing and<br />

repair and determines the cellular response in eukaryotes it was<br />

of interest to look for its phosphorylation at 4 hr after irradiation.<br />

There was significantly higher phosphorylation of ATM at<br />

serine 1981 in A549 cells, which had been exposed to 2 Gy of<br />

proton beam (Figure 3). This indicates that despite significant<br />

activation of ATM till 4 hr after irradiation much of the damage<br />

still persisted as indicated by the presence of significant levels<br />

of γ -H2AX. The activated/phosphorylated ATM further phosphorylates<br />

various downstream target proteins such as H2AX,<br />

Chk2, p53 leading to opening of the chromatin and recruitment<br />

of other proteins involved in DNA repair or cell cycle arrest.<br />

Chk2 which is a another substrate of ATM and is involved in<br />

cell cycle arrest was also found to be phosphorylated at threonine<br />

68 in A549 cells which had been exposed to 2 Gy of proton as<br />

compared to controls (Figure 3), however the activation was not<br />

as intense as observed with ATM or H2AX.<br />

This indicates that small component of ATM activation is also<br />

diverted to cell cycle arrest pathways involving Chk2, CDC25<br />

phosphatase, and CDKs. Chk2 could also be involved in activation<br />

of other proteins such as BRCA1 and p53. Phosphorylation<br />

of p53 at serine 15 was then observed in unirradiated and irradiated<br />

A549 cells at 2 Gy of proton beam (Figure 3).<br />

Whether DNA-PK signals DSBs to the checkpoint machinery<br />

remains controversial, but the emerging consensus is that it does<br />

not (30). DNA-PK may, however, be involved in signalling DNA<br />

damage to the apoptosis machinery (31).<br />

The degree of radiation-induced apoptosis has been shown<br />

to correlate with the p53 wild-type status (32). In addition,<br />

apoptosis is induced when wild-type p53 is transfected into<br />

certain cell lines lacking p53 (33). This indicates that p53 not<br />

only plays a role in regulating the progression through the cell<br />

cycle, it can also induce apoptosis in cells. p53 is not an essential<br />

component of the machinery that carries out apoptosis; however,<br />

620 S. <strong>Ghosh</strong> et al.


Cancer Invest Downloaded from informahealthcare.com by University of Chicago Library on 06/10/11<br />

For personal use only.<br />

it appears to be an activator of apoptosis. Although the exact<br />

means by which p53 activates apoptosis is unclear, evidence<br />

has shown that p53 mediates apoptosis by way of transcriptionindependent<br />

and transcription-dependent mechanisms (34, 35).<br />

p53 is known to regulate the expression of several proteins<br />

involved in the apoptotic pathway and also interacts with BAX,<br />

BCL-XL, and Bcl-2 to exert a direct apoptotic effect at the level<br />

of the mitochondria (36–38).<br />

Increased phosphorylation of p53, upregulation of Bax and<br />

downregulation of Bcl-2 in proton beam irradiated A549 cells<br />

as compared to γ -irradiated cells (Figures 3 and 4) indicate the<br />

activation of apoptotic pathways.<br />

The noteworthy finding of this study is the biphasic activation<br />

of the sensor proteins, ATM, and DNA-PK and no activation<br />

of ATR by proton irradiation and the significant activation<br />

of Chk2 even at the gene level only in the proton beamirradiated<br />

cells (Figure 2). Such biphasic response after irradiation<br />

with proton or other heavy ions has been observed by us<br />

earlier and may be characteristics of particle irradiation (39).<br />

Unlike γ -irradiation, the biphasic induction of ATM and DNA-<br />

PK following proton beam-irradiation could be responsible for<br />

the activation of apoptotic machinery, however, further studies<br />

in this direction will reveal the importance of such biphasic<br />

response.<br />

The use of proton beam therapy has definite advantage over<br />

γ therapy. The mechanism of apoptosis will give the clinician a<br />

handle to manipulate therapy for enhanced cell killing.<br />

DECLARATION OF INTEREST<br />

The authors report no conflicts of interest. The authors alone<br />

are responsible for the content and writing of the paper.<br />

REFERENCES<br />

1. Krengli, M.; Hug, E.B.; Adams, J.A.; Smith, A.R.; Tarbell, N.J.;<br />

Munzenrider, J.E. Proton radiation therapy for retinoblastoma:<br />

comparison of various intraocular tumor locations and beam arrangements.<br />

Int J Radiat Oncol Biol Phys 2005, 61, 583–593.<br />

2. St Clair, W.H.; Adams, J.A.; Bues, M.; Fullerton, B.C.; La Shell,<br />

S.; Kooy, H.M.; Loeffler, J.S.; Tarbell, N.J. Advantage of protons<br />

compared to conventional X-ray or IMRT in the treatment of a<br />

pediatric patient with medulloblastoma. Int J Radiat Oncol Biol<br />

Phys 2004, 58, 727–734.<br />

3. Miralbell, R.; Crowell, C.; Suit, H.D. Potential improvement of three<br />

dimension treatment planning and proton therapy in the outcome<br />

of maxillary sinus cancer. Int J Radiat Oncol Biol Phys 1992, 22,<br />

305–310.<br />

4. Jones, B.; Dale, R.G.; Cárabe-Fernández, A. Charged particle<br />

therapy for cancer: the inheritance of the Cavendish scientists?<br />

Appl Radiat Isot 2009, 67, 371–377.<br />

5. Chang, J.Y.; Zhang, X.; Wang, X.; Kang, Y.; Riley, B.; Bilton, S.;<br />

Mohan, R.; Komaki, R.; Cox, J.D. Significant reduction of normal tissue<br />

dose by proton radiotherapy compared with three-dimensional<br />

conformal or intensity-modulated radiation therapy in Stage I or<br />

Stage III non-small-cell lung cancer. Int J Radiat Oncol Biol Phys<br />

2006, 65, 1087–1096.<br />

6. Fukumitsu, N.; Sugahara, S.; Nakayama, H.; Fukuda, K.; Mizumoto,<br />

M.; Abei, M.; Shoda, J.; Thono, E.; Tsuboi, K.; Tokuuye, K. A<br />

prospective study of hypofractionated proton beam therapy for patients<br />

with hepatocellular carcinoma. Int J Radiat Oncol Biol Phys<br />

2009, 74, 831–836.<br />

7. Shibuya, K.; Mathers, C.D.; Boschi-Pinto, C.; Lopez, A.D.; Murray,<br />

C.J. Global and regional estimates of cancer mortality and incidence<br />

by site: II. Results for the global burden of disease 2000.<br />

BMC Cancer 2002, 2, 37.<br />

8. Jemal, A.; Siegel, R.; Ward, E.; Hao, Y.; Xu, J.; Murray, T.; Thun,<br />

M.J. Cancer statistics, 2008. CA Cancer J Clin 2008, 58, 71–96.<br />

9. Ferlay, J.; Autier, P.; Boniol, M.; Heanue, M.; Colombet, M.; Boyle,<br />

P. Estimates of the cancer incidence and mortality in Europe in<br />

2006. Ann Oncol 2007, 18, 581–592.<br />

10. Jassem J. The role of radiotherapy in lung cancer: where is the<br />

evidence? Radiother Oncol 2007, 83, 203–213.<br />

11. Kim, Y.S.; Yoon, S.M.; Choi, E.K.; Yi, B.Y.; Kim, J.H.; Ahn, S.D.;<br />

Lee, S.W.; Shin, S.S.; Lee, J.S.; Suh, C.; Kim, S.W.; Kim, D.S.;<br />

Kim, W.S.; Park, H.J.; Park, C.I. Phase II study of radiotherapy<br />

with three-dimensional conformal boost concurrent with paclitaxel<br />

and cisplatin for Stage IIIB non-small-cell lung cancer. Int J Radiat<br />

Oncol Biol Phys 2005, 62, 76–81.<br />

12. Farray, D.; Mirkovic, N.; Albain, K.S. Multimodality therapy for stage<br />

III non-small-cell lung cancer. J Clin Oncol 2005, 23, 3257–3269.<br />

13. Bush, D.A.; Slater, J.D.; Bonnet, R.; Cheek, G.A.; Dunbar, R.D.;<br />

Moyers, M.; Slater, J.M. Proton-beam radiotherapy for early-stage<br />

lung cancer. Chest 1999, 116, 1313–1319.<br />

14. Shioyama, Y.; Tokuuye, K.; Okumura, T.; Kagei, K.; Sugahara, S.;<br />

Ohara, K.; Akine, Y.; Ishikawa, S.; Satoh, H.; Sekizawa, K. Clinical<br />

evaluation of proton radiotherapy for non-small-cell lung cancer.<br />

Int J Radiat Oncol Biol Phys 2003; 56: 7–13.<br />

15. Koto, M.; Miyamoto, T.; Yamamoto, N.; Nishimura, H.; Yamada, S.;<br />

Tsujii, H. Local control and recurrence of stage I non-small cell<br />

lung cancer after carbon ion radiotherapy. Radiother Oncol 2004,<br />

71, 147–156.<br />

16. Maser, R.S.; Monsen, K.J.; Nelms, B.E.; Petrini, J.H. hMre11 and<br />

hRad50 nuclear foci are induced during the normal cellular response<br />

to DNA double-strand breaks. Mol Cell Biol 1997, 17,<br />

6087–6096.<br />

17. Fernandez-Capetillo, O.; Lee, A.; Nussenzweig, M.; Nussenzweig,<br />

A. H2AX: the histone guardian of the genome. DNA Repair 2004,<br />

3, 959–967.<br />

18. Bekker-Jensen, S.; Lukas, C.; Kitagawa, R.; Melander, F.; Kastan,<br />

M.B.; Bartek, J.; Lukas, J. Spatial organization of the mammalian<br />

genome surveillance machinery in response to DNA strand breaks.<br />

J Cell Biol 2006, 173, 195–206.<br />

19. Goto, S.; Watanabe, M.; Yatagai, F. Delayed cell cycle progression<br />

in human lymphoblastoid cells after exposure to high-LET radiation<br />

correlates with extremely localized DNA damage. Radiat Res<br />

2002, 158, 678–686.<br />

20. Stenerlow, B.; Hoglund, E.; Carlsson, J.; Blomquist, E. Rejoining<br />

of DNA fragments produced by radiations of different linear energy<br />

transfer. Int J Radiat Biol 2000, 76, 549–557.<br />

21. Lee, K.B.; Lee, J.S.; Park, J.W.; Huh, T.L.; Lee, Y.M. Low energy<br />

proton beam induces tumor cell apoptosis through reactive oxygen<br />

species and activation of caspases. Exp Mol Med 2008, 40,<br />

118–129<br />

22. <strong>Ghosh</strong>, S.; Maurya, D.K.; Krishna, M. Role of iNOS in bystander<br />

signaling between macrophages and lymphoma cells. Int J Radiat<br />

Oncol Boil Phy 2008, 72, 1567–1574<br />

23. Rasband, W.S. and Image, J.U.S. <strong>National</strong> Institutes of Health,<br />

Bethesda, USA, http://rsb.info.nih.gov/ij/ (1997–2006).<br />

24. Khanna, K.K.; Lavin, M.F.; Jackson, S.P.; Mulhern, T.D. ATM, a<br />

central controller of cellular responses to DNA damage. Cell Death<br />

Differ 2001, 8(11), 1052–1065.<br />

25. Lakin, N.D.; Jackson, S.P. Regulation of p53 in response to DNA<br />

damage. Oncogene 1999, 18, 7644–7655.<br />

26. Rich, T.; Allen, R.L.; Wyllie, A.H. Defying death after DNA damage.<br />

Nature 2000, 407, 777–783.<br />

Efficient Cell killing by Proton Beam 621


Cancer Invest Downloaded from informahealthcare.com by University of Chicago Library on 06/10/11<br />

For personal use only.<br />

27. Wang, H.; Zeng, Z.C.; Bui, T.A.; DiBiase, S.J.; Qin, W.; Xia,<br />

F.; Powell, S.N.; Iliakis, G. Nonhomologous end-joining of ionizing<br />

radiation-induced DNA double-stranded breaks in human tumor<br />

cells deficient in BRCA1 or BRCA2. Cancer Res 2001, 61,<br />

270–277<br />

28. Rogakou, E.P.; Pilch, D.R.; Orr, A.H.; Ivanova, V.S.; Bonner,<br />

W.M. DNA double-stranded breaks induce histone H2AX phosphorylation<br />

on serine 139. J Biol Chem 1998, 273, 5858–<br />

5868.<br />

29. Olive, P.L.; Banath, J.P. Phosphorylation of histone H2AX as a<br />

measure of radiosensitivity. Int J Radiat Oncol Biol Phys 2004, 58,<br />

331–335.<br />

30. Burma, S.; Kurimasa, A.; Xie, G.; Taya, Y.; Araki, R.; Abe, M.;<br />

Crissman, H.A.; Ouyang, H.; Li, G.C.; Chen, D.J. DNA-dependent<br />

protein kinase-independent activation of p53 in response to DNA<br />

damage. J Biol Chem 1999, 274, 17139–17143.<br />

31. Wang, S.; Guo, M.; Ouyang, H.; Li, X.; Cordon-Cardo, C.; Kurimasa,<br />

A.; Chen, D.J.; Fuks, Z.; Ling, C.C.; Li, G.C. The catalytic<br />

subunit of DNA-dependent protein kinase selectively regulates<br />

p53-dependent apoptosis but not cell-cycle arrest. Proc Natl Acad<br />

Sci USA 2000, 97, 1584–1588.<br />

32. Fei, P.; Bernhard, E.J.; El-Deiry, W.S. Tissue-specific induction of<br />

p53 targets in vivo. Cancer Res 2002, 62, 7316–7327.<br />

33. Yonish-Rouach, E.; Resnitzky, D.; Lotem, J.; Sachs, L.; Kimchi, A.;<br />

Oren, M. Wild-type p53 induces apoptosis of myeloid leukaemic<br />

cells that is inhibited by interleukin-6. Nature 1991, 352, 345–347.<br />

34. Sabbatini, P.; Lin, J.; Levine, A.J.; White, E. Essential role for<br />

p53-mediated transcription in E1A-induced apoptosis. Genes Dev<br />

1995, 9, 2184–2192.<br />

35. Caelles, C.; Helmberg, A.; Karin, M. p53-dependent apoptosis in<br />

the absence of transcriptional activation of p53-target genes. Nature<br />

1994, 370, 220–223.<br />

36. Polyak, K.; Xia, Y.; Zweier, J.L.; Kinzler, K.W.; Vogelstein, B. A<br />

model for p53-induced apoptosis. Nature 1997, 389, 300–305.<br />

37. Mihara, M.; Erster, S.; Zaika, A.; Petrenko, O.; Chittenden, T.; Pancoska,<br />

P.; Moll, U.M. p53 has a direct apoptogenic role at the<br />

mitochondria. Mol Cell 2003, 11, 577–590.<br />

38. Fei, P.; El-Deiry, W.S. p53 and radiation responses. Oncogene<br />

2003, 22, 5774–5783.<br />

39. Mitra, A.K.; Bhat, N.; Sarma, A.; Krishna, M. Alteration in the expression<br />

of signaling parameters following carbon ion irradiation.<br />

Mol Cell Biochem 2005, 276(1–2), 169–173.<br />

622 S. <strong>Ghosh</strong> et al.


doi:10.1016/j.ijrobp.2008.08.006<br />

Int. J. Radiation Oncology Biol. Phys., Vol. 72, No. 5, pp. 1567–1574, 2008<br />

Copyright Ó 2008 Elsevier Inc.<br />

Printed in the USA. All rights reserved<br />

0360-3016/08/$–see front matter<br />

BIOLOGY CONTRIBUTION<br />

ROLE OF iNOS IN BYSTANDER SIGNALING BETWEEN MACROPHAGES AND<br />

LYMPHOMA CELLS<br />

SOMNATH GHOSH, M.SC.,* DHARMENDRA KUMAR MAURYA, PH.D.,* AND MALINI KRISHNA, PH.D.*<br />

*Radiation Biology and Health Science Division, <strong>Bhabha</strong> Atomic Research Centre, Mumbai, India<br />

Purpose: The present report describes the bystander effects of radiation between similar and dissimilar cells and<br />

the role of iNOS in such communication.<br />

Materials and Methods: EL-4 and RAW 264.7 cells were exposed to 5 Gy g-irradiation. The medium from irradiated<br />

cells was transferred to unirradiated cells.<br />

Results: Irradiated EL-4 cells as well as those cultured in the presence of medium from g-irradiated EL-4 cells<br />

showed an upregulation of NF-kB, iNOS, p53, and p21/waf1 genes. The directly irradiated and the bystander<br />

EL-4 cells showed an increase in DNA damage, apoptosis, and NO production. Bystander signaling was also found<br />

to exist between RAW 264.7 (macrophage) and EL-4 (lymphoma) cells. Unstimulated or irradiated RAW 264.7<br />

cells did not induce bystander effect in unirradiated EL-4 cells, but LPS stimulated and irradiated RAW 264.7 cells<br />

induced an upregulation of NF-kB and iNOS genes and increased the DNA damage in bystander EL-4 cells. Treatment<br />

of EL-4 or RAW 264.7 cells with L-NAME significantly reduced the induction of gene expression and DNA<br />

damage in the bystander EL-4 cells, whereas treatment with cPTIO only partially reduced the induction of gene<br />

expression and DNA damage in the bystander EL-4 cells.<br />

Conclusions: It was concluded that active iNOS in the irradiated cells was essential for bystander response.<br />

Ó 2008 Elsevier Inc.<br />

Radiation, Bystander signaling, Nuclear factor-kB, iNOS, p53 and p21.<br />

INTRODUCTION<br />

It is now apparent that the target for biologic effects of ionizing<br />

radiation is not solely the irradiated cell, but also the surrounding<br />

cells and tissues. Preliminary evidence indicates<br />

that similar signal transduction pathways are activated in directly<br />

irradiated cells and the bystander cells and contribute to<br />

the induction of bystander DNA damage (1). Many genes<br />

have been shown to be activated in bystander cells (2), but<br />

evidence for genes that are involved in the signaling pathways<br />

is lacking.<br />

Several factors involving soluble growth factors (3, 4), oxidative<br />

metabolites (5), and those that pass through gap junction<br />

(5–7) have been implicated. Among the various<br />

compounds that could transmit the signal from irradiated<br />

cells to bystander cells, nitric oxide (NO), by virtue of its diffusible<br />

nature and long life in vivo is a promising candidate<br />

(8–10). The accumulation of inducible nitric oxide synthase<br />

(iNOS) and the increased activity of constitutive NOS have<br />

been known to be an early signaling event induced by radiation.<br />

Gamma irradiation–induced DNA damage is known to<br />

enhance NO production in lipopolysaccharide (LPS)-stimulated<br />

macrophages (11). It is likely that stimulated and unstimulated<br />

macrophage behave differently where bystander<br />

response is concerned.<br />

Although abscopal effects have been deemed responsible<br />

for retarding the growth of a similar tumors located at a distant<br />

site in the mice (12), their contribution in transferring signals<br />

between dissimilar cells have not been extensively examined.<br />

To validate the presence of abscopal effects in vitro, we have<br />

used two dissimilar cell lines viz. EL-4 cells, a lymphoma cell<br />

line and RAW 264.7 cells, a macrophage cell line. Because<br />

there is extensive intercellular communications between dissimilar<br />

cells during immune responses, the leukocytes and<br />

lymphocytes were deemed suitable candidates to study<br />

bystander effects.<br />

The present report looks at the mechanism of bystander response<br />

in similar cells (EL-4 Vs EL-4 cells) and whether the<br />

state of activation of RAW 264.7 cells affects the bystander<br />

response in dissimilar cells (EL-4 Vs RAW 264.7).<br />

METHODS AND MATERIALS<br />

Cell culture, reagents, and irradiation schedule<br />

Mouse lymphoma cell line, EL-4 and mouse macrophage RAW<br />

264.7 cells were grown in RPMI 1640 (Sigma, St. Louis, Mo),<br />

Reprint requests to: Dr. Malini Krishna, Ph.D., <strong>Bhabha</strong> Atomic<br />

Research Centre, Radiation Biology and Health Science Division,<br />

Trombay, Mumbai, Mumbai, Maharashtra 400085, India. Tel:<br />

(+91) 22-25593868; Fax: (+91) 22-2550 5151; E-mail: malini@<br />

barc.gov.in<br />

1567<br />

Conflict of interest: none<br />

Received April 25, 2008, and in revised form July 30, 2008.<br />

Accepted for publication Aug 2, 2008.


1568 I. J. Radiation Oncology d Biology d Physics Volume 72, Number 5, 2008<br />

supplemented with 10% fetal calf serum (Sigma, St. Louis, Mo).<br />

Cells were maintained at 37 C in humidified atmosphere with 5%<br />

CO 2 .<br />

EL-4 and RAW 264.7 cells were resuspended in RPMI medium at<br />

a density of 1 10 6 cells/mL and exposed to 5 Gy g irradiation using<br />

Gamma Cell 220 irradiator (Atomic Energy of Canada Ltd), at<br />

a dose rate of 4.74 Gy/min. Medium from irradiated cells was transferred<br />

to unirradiated cells (1 h postirradiation). Some group of<br />

RAW 264.7 cells were stimulated with 500 ng/mL of bacterial<br />

lipopolysaccharide (LPS, Sigma, St. Louis, Mo) for 3 h and then<br />

the medium was replaced before irradiation.<br />

Before designing the experiment, there were two aspects that had<br />

to be taken care of. One was the likelihood that the replacement of<br />

the medium itself caused transient expression of certain signaling<br />

factors and second that the components of the medium, which include<br />

proteins such as growth factors, which are unavoidably exposed<br />

to radiation, contribute to the changes in the bystander cell.<br />

The experimental design, therefore, consisted of the following<br />

sets: (a) control unirradiated cells, (b) g-irradiated cells, (c) EL-4<br />

cells receiving medium from unirradiated EL-4 cells or RAW<br />

264.7 cells or LPS-stimulated RAW 264.7 cells, (d) EL-4 cells receiving<br />

medium from g-irradiated EL-4 cells or irradiated RAW<br />

264.7 cells or LPS-stimulated and irradiated RAW 264.7 cells,<br />

and (e) EL-4 cells receiving g-irradiated medium. The cells (EL-4<br />

and RAW 264.7) were also divided into two groups; one was treated<br />

with 20 mM cPTIO (NO scavenger; Molecular Probe, Invitrogen,<br />

Carlsbad, CA) and the other with 1 mmol L-NAME (NOS inhibitor;<br />

Sigma, St. Louis, MO). L-NAME was washed off after 30 min so<br />

that it is not present in the bystander cells during incubation.<br />

Measurement of DNA damage<br />

DNA damage in all the groups of EL-4 cells after exposure to different<br />

conditioned media was measured after 2 h by alkaline single<br />

cell gel electrophoresis (comet assay) (13).<br />

Semiquantitative reverse transcriptional polymerase chain<br />

reaction<br />

Total RNA from all the groups of EL-4 cells (1 10 6 cells) was<br />

extracted 2 h after exposure to different conditioned media and eluted<br />

in 30 mL RNAase free water using RNA tissue kit (Roche, Mannheim,<br />

Germany). Equal amount of RNA in each group was reverse transcribed<br />

using cMaster RT kit (Eppendorf, Hamburg, Germany).<br />

Equal amount (2 mL) of cDNA in each group was used for specific<br />

amplification of b-actin, NF-kB (p65), iNOS, p21, or p53 gene using<br />

gene specific primers (Table 1). The polymerase chain reaction (PCR)<br />

condition were 94 C, 5 min initial denaturation followed by 30 cycles<br />

of 94 C 45 s, 55–62 C 45 s (depending on the annealing temperature<br />

of specific primers given in Table 1), 72 C 45 s, and final extension at<br />

72 C for 10 min. Equal amount of each PCR product (10 mL) was run<br />

on 1.5% agarose gel containing ethidium bromide in tris borate<br />

EDTA buffer at 60 V. The bands in the gel were visualized under<br />

an ultraviolet lamp and relative intensities were quantified using<br />

Gel doc software (Syngene, Nuffied Road, Cambridge, UK).<br />

NO production in EL-4 cells<br />

All the groups of EL-4 cells after exposure to different conditioned<br />

media were cultured for 24 h. The production of NO in the<br />

culture supernatants was measured by assaying NO 2 – using colorimetric<br />

Griess reaction (14).<br />

Table 1. The sequences of gene specific primers and their<br />

annealing temperature (Tm) used in reverse transcriptasepolymerase<br />

chain reaction experiments<br />

Gene<br />

Primer sequence<br />

Apoptosis in EL-4 cells<br />

Apoptosis in control unirradiated cells, cells exposed to radiation,<br />

and the cells exposed to different conditioned media as mentioned<br />

previously was measured by DNA fragmentation assay (15).<br />

The characterization and quantification of apoptosis was further<br />

confirmed by confocal microscopy (Carl Zeiss, LSM 510 META,<br />

Carl Zeiss Micro Imaging GMBH, Jena, Germany) using an annexin<br />

V/PI-staining kit (Roche, Mannheim, Germany) according to the<br />

manufacturer’s instructions. Five fields of cells for each group<br />

(1,000 cells) were counted to generate apoptotic index. Annexin-V<br />

binds to phosphatidylserine is a marker of early apoptosis<br />

and binding of PI indicates necrosis.<br />

Statistical analysis<br />

The data were imported to Excel work sheets, and graphs were<br />

made using Origin version 5.0. One-way analysis of variance with<br />

Tukey-Kramer multiple comparisons as posttest for p < 0.05 used<br />

to study the significant level. Data were insignificant at p > 0.05.<br />

Each point represented as mean standard error of the mean.<br />

RESULTS<br />

Tm<br />

( C)<br />

55<br />

b-Actin Forward 5 0 -<br />

TGGAATCCTGTGGCATCCATGAAAC-3 0<br />

Reverse 5 0 -<br />

TAAAACGCAGCTCAGTAACAGTCCG-3 0<br />

iNOS Forward 5 0 -CTCTGACAGCCCAGAGTTCC-3 0 62<br />

Reverse 5 0 -AGGCAAAGGAGGAGAAGGAG-3 0<br />

p21 Forward 5 0 -GCCTTAGCCCTCACTCTGTG-3 0 58<br />

Reverse 5 0 -GGTTGGGAGGGGCTTAAATA-3 0<br />

p53 Forward 5 0 -CGGGTGGAAGGAAATTTGTA-3 0 58<br />

Reverse 5 0 -CTTCTGTACGGCGGTCTCTC -3 0<br />

p65 Forward 5 0 -ACAACTGAGCCCATGCTGAT-3 0 50<br />

Reverse 5 0 -GAGAGGTCCATGTCCGCAAT-3 0<br />

Bystander effect in similar cells (EL-4 Vs EL-4)<br />

The expression of NF-kB (p65), iNOS, p53, and p21 genes<br />

were significantly upregulated in irradiated as well as in bystander<br />

cells (Fig. 1 A–D, Lanes 2, 3). However, in the cells<br />

cultured in the presence of only irradiated medium, there was<br />

a marginal upregulation of p53 but no upregulation of p21<br />

(data not shown). The expression of p53 gene in these cells<br />

was significantly lower than irradiated cells as well as bystander<br />

cells.<br />

Because the expressions of iNOS gene was significantly<br />

upregulated in irradiated and bystander cells (Fig. 1B, Lanes<br />

2, 3), it was of interest to look for NO production in these<br />

cells and expectedly NO production was also found to be significantly<br />

enhanced in irradiated and bystander cells (Fig. 2,<br />

Lanes 2, 3).<br />

Damage to cellular DNA, expressed as tail length<br />

(Fig. 3A), percent DNA in tail (Fig. 3B), Olive tail moment<br />

(Fig. 3C) and tail moment (Fig. 3D) was significantly higher<br />

in irradiated EL-4 cells and bystander cells as compared with<br />

the unirradiated control cells (Fig. 3A–D, Lanes 2, 3).


Bystander effect: role of iNOS d S. GHOSH et al. 1569<br />

Fig. 1. Gene expression of inducible nitric oxide synthase (iNOS),<br />

p21, p53, and NF-kB (p65) in EL-4 cells cultured for 2 h in presence<br />

of different conditioned media. Total RNA from EL-4 cells was isolated<br />

and reverse transcribed. Reverse transcriptase-polymerase<br />

chain reaction (PCR) analysis of iNOS, p21, p53, and NF-kB<br />

(p65) genes was carried out as described in materials and methods.<br />

PCR products were resolved on 1.5% agarose gels containing ethidium<br />

bromide. b-Actin gene expression in each group was used as an<br />

internal control. Ratio of intensities of (A) NF-kB (p65), (B) iNOS,<br />

(C) p53, and (D) p21band to that of respective b-actin band as quantified<br />

from gel pictures are shown above each gel picture. Key: Lane<br />

1, control unirradiated EL-4 cells; Lane 2, g irradiated EL-4 cells;<br />

Lane 3, unirradiated EL-4 cells receiving medium from g-irradiated<br />

EL-4 cells. Data represent means SE of three independent experiments;<br />

significantly different from unirradiated controls: *p < 0.05,<br />

**p < 0.01.<br />

There was a clear DNA ladder formation (a hallmark of apoptosis)<br />

in irradiated and bystander EL-4 cells. Apoptosis<br />

was found to be higher in irradiated cells than bystander cells<br />

(Fig. 4, Lanes 2, 3). Apoptosis in irradiated and bystander<br />

EL-4 cells was further confirmed by annexin V/PI assay<br />

(Fig. 5).<br />

EL-4 cells that had received only irradiated medium or<br />

medium from unirradiated control cells did not show any<br />

upregulation of NF-kB and iNOS or NO production or<br />

DNA damage or apoptosis (data not shown).<br />

Bystander effect in dissimilar cells (EL-4 Vs RAW 264.7)<br />

Having confirmed the existence of bystander signaling between<br />

similar cells (EL-4 Vs EL-4), the work was extended to<br />

determine if bystander effects can be demonstrated between<br />

different cell type (RAW 264.7 and EL-4) and whether the<br />

Fig. 2. Nitric oxide (NO) production from EL-4 cells cultured for 24<br />

h in the presence of different conditioned media. The culture supernatant<br />

from each group of EL-4 cells was used for determination of<br />

NO 2 – with Griess reagent. Key: Lane 1, control unirradiated EL-4<br />

cells; Lane 2, g irradiated EL-4 cells; Lane 3, unirradiated EL-4 cells<br />

receiving medium from g irradiated EL-4 cells. Data represent<br />

means SE of three independent experiments; significantly different<br />

from unirradiated controls: *p < 0.05, **p < 0.01.<br />

state of stimulation of macrophages (RAW 264.7) affects bystander<br />

response. The expression of NF-kB was not altered in<br />

EL-4 cells that had received medium from either unstimulated<br />

RAW 264.7 cells or irradiated RAW 264.7 cells, but<br />

the expression of iNOS was downregulated (Fig. 6, Lanes<br />

3, 5). In EL-4 cells that had received medium from LPS stimulated<br />

RAW 264.7 cells there was an upregulation of iNOS<br />

gene (Fig. 6, Lane 4). The expression of both the genes<br />

(NF-kB and iNOS) was significantly upregulated in EL-4<br />

cells that had received medium from LPS stimulated and<br />

irradiated RAW 264.7 cells (Fig. 6, Lane 2).<br />

The EL-4 cells that had received medium from unstimulated<br />

RAW 264.7 cells or LPS-stimulated RAW 264.7 cells<br />

or irradiated RAW 264.7 cells did not show any increase in<br />

DNA damage (Fig. 7, Lanes 3–5). But there was significantly<br />

higher DNA damage in EL-4 cells that had received medium<br />

from LPS stimulated and irradiated RAW 264.7 cells<br />

(p < 0.01) (Fig. 7, Lane, 2).<br />

Effect of L-NAME and cPTIO on Bystander response<br />

Because the expressions of NF-kB and iNOS gene were<br />

significantly upregulated in irradiated and bystander cells<br />

(Fig. 8, Lanes 2, 3), it was of interest to determine what


1570 I. J. Radiation Oncology d Biology d Physics Volume 72, Number 5, 2008<br />

Fig. 3. DNA damage in EL-4 cells cultured for 2 h in presence of<br />

different conditioned media was measured using single-cell gel electrophoresis<br />

(‘‘comet assay’’). (A) Tail length, (B) % DNA in tail, (C)<br />

olive tail moment, and (D) tail moment. Key: Lane 1, control unirradiated<br />

EL-4 cells; Lane 2, g-irradiated EL-4 cells; Lane 3, unirradiated<br />

EL-4 cells receiving medium from g-irradiated EL-4 cells. Data<br />

represent means SE of three independent experiments; significantly<br />

different from unirradiated controls: *p < 0.05, **p < 0.01.<br />

Fig. 4. DNA fragmentation assay of EL-4 cells. EL-4 cells cultured<br />

for 24 h in the presence of different conditioned media. DNAs of<br />

each group of cells were isolated as described in Methods and Materials<br />

section and electrophoresed on 1.8% agarose gel containing<br />

ethidium bromide. Key: Lane M, DNA molecular weight marker;<br />

Lane 1, control unirradiated EL-4 cells; Lane 2, g irradiated EL-4<br />

cells; Lane 3, unirradiated EL-4 cells receiving medium from g-<br />

irradiated EL-4 cells. Data shown from representative experiment<br />

of three independent experiments.<br />

transmits the signal to the bystanding cells. The response was<br />

looked at in the presence of cPTIO or L-NAME. L-NAME<br />

was washed out before irradiation so that the bystander cells<br />

are not exposed to L-NAME, and iNOS is inhibited only in<br />

the irradiated cells. The addition of cPTIO, which scavenges<br />

the NO released in the medium, did not affect the expression<br />

of iNOS gene if the bystanding cell was of same origin, although<br />

NF-kB was somewhat inhibited (Fig. 8, compare<br />

Lanes 5 and 3). The addition of L-NAME, which inhibits<br />

the iNOS in the irradiated cells, completely inhibited the bystander<br />

response in the EL-4 cells (Fig. 8, compare Lanes 4<br />

and 3).<br />

However, both L-NAME and cPTIO inhibited the expression<br />

of NF-kB and iNOS in the bystander EL-4 cells if the<br />

medium was from LPS stimulated and irradiated RAW<br />

264.7 cells (Fig. 8, Lanes 6, 7) (i.e., the cell type was different),<br />

indicating that the bystander response was dependent on<br />

the cells type and the state of stimulation of the irradiated<br />

cells.<br />

DNA damage was significantly higher in irradiated and<br />

bystander cells (Fig. 9, Lanes 2, 3). However, the EL-4 cells<br />

that had received medium either from L-NAME–treated and<br />

irradiated EL-4 cells or from LPS-stimulated, L-NAMEtreated<br />

and irradiated RAW 264.7 cells did not show any increase<br />

in DNA damage (Fig. 9, Lanes 4, 6). Similarly, the EL-<br />

4 cells that had received medium either from cPTIO-treated<br />

and irradiated EL-4 cells or from LPS-stimulated, cPTIOtreated<br />

and irradiated RAW 264.7 cells did not show any increase<br />

in DNA damage (Fig. 9, Lanes 5, 7). Neither L-NAME<br />

nor cPTIO itself had any effect on DNA damage in EL-4 cells<br />

(Fig. 9, Lanes 8, 9).<br />

DISCUSSION<br />

In an attempt to understand bystander signaling, several<br />

studies have been made between the same cell lines (1, 2,<br />

16). However, it is not known whether bystander effect is<br />

extendable to dissimilar cells, particularly those cells that<br />

communicate with each other under physiologic situations<br />

like the macrophages and lymphocytes. In immunology,<br />

there is an immense cross-talk between the lymphocytes<br />

and the macrophages. This could make them more prone to<br />

bystander signaling.


Bystander effect: role of iNOS d S. GHOSH et al. 1571<br />

Fig. 5. Apoptosis assay of control, irradiated, and bystander EL-4 cells by Annexin-PI staining. EL-4 cells cultured for 18 h<br />

in the presence of different conditioned media. The EL-4 cells were stained using an annexin V/PI-staining kit. Characterization<br />

and quantification of apoptosis was done by confocal microscopy. Key: (A) Annexin-V; (B) PI; (C) DAPI; (D)<br />

Merge image of annexin-v, PI, and DAPI. Data represent means SE of three independent experiments; significantly<br />

different from unirradiated controls: **p < 0.01.<br />

There are numerous reports on the activation of signaling<br />

factors in the bystanding cells (1,2). However, not much is<br />

known about the genes that are activated in the bystanding<br />

cells in response to these signals. In our earlier work, we<br />

had shown that NF-kB and p21 levels were elevated in<br />

K562 cells (16). The present study was carried out to investigate<br />

which genes are activated in the bystanding cell and<br />

whether the state of activation of the irradiated cell alters<br />

the bystander response. The study was restricted only to the<br />

genes because, from our experience, even the change of irradiated<br />

medium leads to an activation of NF-kB (16). Additionally,<br />

minor stresses are known to alter the state of<br />

phosphorylation of signaling proteins. Our present results revealed<br />

a significant increase in DNA damage in the bystander<br />

cells, but only in those cells that had received medium from<br />

irradiated cells (Fig. 3). Because the cells treated only with<br />

irradiated medium did not show any DNA damage or NO<br />

production (results not shown), the signals that caused the<br />

damage in the bystander cells must be coming from the irradiated<br />

cells themselves. The expression of cell cycle–related<br />

genes (p53 and p21) was significantly upregulated in bystander<br />

similar cells (Fig. 1C, D, Lane 3). Our results are in<br />

agreement with the work on bystander signaling (1,2). The<br />

first report describing the induction of sister chromatid exchanges<br />

was examined in Chinese hamster ovary cells. Because<br />

Chinese hamster ovary cells contain mutant p53, p53<br />

could not be involved in the bystander induction of sister<br />

chromatid exchanges (17). If the p53 of the bystander cell<br />

is mutated, there should be less DNA repair and in fact<br />

more bystander effect can be seen when assessed in terms<br />

of DNA damage.<br />

A clear cause and effect relationship was established between<br />

DNA damage, gene expression, and cell survival and<br />

was evident in the form of increase apoptosis in bystander<br />

cells (Fig. 4, Lane 3, Fig. 5).<br />

An increased expression of NF-kB gene (p65) could lead<br />

to the activation of iNOS gene because the promoter of the<br />

iNOS gene contains binding sites for NF-kB. NF-kB is<br />

known to be activated by several oxidative stresses including<br />

ionization radiation (18–23). As a consequence of iNOS gene


1572 I. J. Radiation Oncology d Biology d Physics Volume 72, Number 5, 2008<br />

Fig. 6. Gene expression of NF-kB (p65) and iNOS gene in EL-4<br />

cells that had received medium from irradiated RAW 264.7 cells.<br />

EL-4 cells cultured for 2 h in presence of different conditioned media.<br />

b-Actin gene expression in each group was used as an internal<br />

control. Ratio of intensities of NF-kB (p65) and iNOS band to that of<br />

respective b-actin band as quantified from gel pictures are shown<br />

above each gel picture. Key: Lane 1, control EL-4 cells; Lane 2,<br />

EL-4 cells receiving medium from lipopolysaccharide (LPS)-stimulated<br />

(500 ng/mL) and g-irradiated RAW 264.7 cells; Lane 3, EL-4<br />

cells receiving medium from g-irradiated RAW 264.7 cells without<br />

LPS stimulation; Lane 4, EL-4 cells receiving medium from LPSstimulated<br />

RAW 264.7 cells; Lane 5, EL-4 cells receiving medium<br />

from unirradiated RAW 264.7 cells. Data represent means SE of<br />

three independent experiments; significantly different from unirradiated<br />

controls: *p < 0.05, **p < 0.01.<br />

upregulation and increased production of NO, many forms of<br />

reactive nitrogen species would be formed leading to covalent<br />

modification of proteins and the expression of various<br />

other genes (24).<br />

Having confirmed the existence of bystander effect between<br />

similar cells (EL-4 Vs EL-4), the work was extended<br />

to look for the existence of cross-bystander effect between<br />

EL-4 and RAW 264.7 cells (i.e., lymphocytes and macrophages).<br />

There was a significant upregulation of NF-kB and<br />

iNOS genes in EL-4 cells that had received medium from<br />

LPS stimulated and irradiated RAW 264.7 cells (Fig. 6,<br />

Lane 2) and expectedly the DNA damage was also significantly<br />

higher in these cells (Fig. 7, Lane 2). A noteworthy<br />

finding of this study is that the unirradiated EL-4 cells that<br />

had received medium from LPS-stimulated RAW 264.7 cells<br />

also showed upregulation of iNOS gene but not NF-kB gene<br />

(Fig. 6, Lane 4). The LPS treatment of macrophage upregulates<br />

many genes in the macrophage hence their response to<br />

irradiation will be different from unstimulated macrophage.<br />

Another noteworthy finding of this study is that the addition<br />

of medium from nonirradiated RAW 264.7 cells resulted in<br />

significantly decrease in iNOS gene expression in EL-4 cells<br />

Fig. 7. DNA damage in EL-4 cells cultured for 2 h in presence of<br />

different conditioned media was measured using single-cell gel electrophoresis<br />

(‘‘comet assay’’) and expressed as tail length. Key: Lane<br />

1, control EL-4 cells; Lane 2, EL-4 cells receiving medium from<br />

LPS-stimulated (500 ng/mL) and g-irradiated RAW 264.7 cells;<br />

Lane 3, EL-4 cells receiving medium from g irradiated RAW<br />

264.7 cells without lipopolysaccharide stimulation; Lane 4, EL-4<br />

cells receiving medium from LPS-stimulated RAW 264.7 cells;<br />

Lane 5, EL-4 cells receiving medium from unirradiated RAW<br />

264.7 cells. Data represent means SE of three independent; significantly<br />

different from unirradiated controls: *p < 0.01.<br />

relative to the control (Fig. 6, Lane 5). Nonirradiated macrophages<br />

may release some factor that is responsible for down<br />

regulation of iNOS in EL-4 cells. Further studies in this direction<br />

will reveal the nature of the factor. It is clear from these<br />

results that effective cross-bystander effects exists between<br />

these two cell lines only when the RAW 264.7 cells are<br />

both stimulated and irradiated. Other studies have looked at<br />

only the clonogenic survival of dissimilar bystander cells (25).<br />

Numerous studies have implicated extracellular secreted<br />

signaling proteins in mediating the bystander effect (5), but<br />

the strongest contenders have been reactive oxygen and nitrogen<br />

species (26). To investigate what factors are involved in<br />

radiation induced bystander effect, the cells were treated with<br />

either L-NAME (NOS inhibitor) or cPTIO (NO scavenger).<br />

The effect of cPTIO and L-NAME in bystander response<br />

between similar cell (EL-4 Vs EL-4) was different in that<br />

treatment with L-NAME reduces the bystander induction of<br />

NF-kB as well as iNOS gene expression in EL-4 cells receiving<br />

medium from irradiated EL-4 cells (Fig. 8, Lane 4), but<br />

the treatment with cPTIO reduces the bystander induction<br />

of only NF-kB gene expression and not iNOS gene expression<br />

(Fig. 8, Lane 5). These results demonstrate that the activated<br />

NOS in the irradiated cell is essential for gene<br />

expression in the bystanding cell.<br />

The addition of either L-NAME or cPTIO reduces the<br />

bystander induction of NF-kB as well as iNOS gene expression<br />

in EL-4 cells receiving medium from LPS stimulated<br />

and irradiated RAW 264.7 cells (Fig. 8, Lanes 6, 7), indicating<br />

that both activation of iNOS in the irradiated cell and NO<br />

production play a role in cross-bystander effect.


Bystander effect: role of iNOS d S. GHOSH et al. 1573<br />

Fig. 8. Gene expression of NF-kB (p65) and iNOS gene in EL-4<br />

cells that had received medium from irradiated cells. EL-4 cells cultured<br />

for 2 h in presence of different conditioned media. b-Actin<br />

gene expression in each group was used as an internal control. Ratio<br />

of intensities of NF-kB (p65) and inducible nitric oxide synthase<br />

band to that of respective b-actin band as quantified from gel pictures<br />

are shown above each gel picture. Key: Lane 1, control EL-<br />

4 cells; Lane 2, g-irradiated EL-4 cells; Lane 3, EL-4 cells receiving<br />

medium from g-irradiated EL-4 cells; Lane 4, EL-4 cells receiving<br />

medium from L-NAME treated and g-irradiated EL-4 cells; Lane 5,<br />

EL-4 cells receiving medium from cPTIO-treated and g-irradiated<br />

EL-4 cells; Lane 6, EL-4 cells receiving medium from LPS-stimulated<br />

(500 ng/mL), L-NAME treated and g-irradiated RAW 264.7<br />

cells; Lane 7, EL-4 cells receiving medium from lipopolysaccharide-stimulated<br />

(500 ng/mL), cPTIO-treated and g-irradiated<br />

RAW 264.7 cells. Data represent means SE of three independent<br />

experiments; significantly different from unirradiated controls:<br />

*p < 0.05, **p < 0.01.<br />

The treatment with L-NAME or cPTIO also reduces the<br />

bystander induction of DNA damage in EL-4 cells receiving<br />

medium from irradiated EL-4 cells or LPS-stimulated and<br />

LPS-irradiated RAW 264.7 cells (Fig. 9, Lanes 4–7), demonstrating<br />

that NO contributes to the DNA damage in bystander<br />

EL-4 cells. NO is generated endogenously from L-arginine<br />

by inducible NO synthases (27, 28) that can be stimulated<br />

by radiation in mammalian cells (8). It is believed that NO itself<br />

does not induce DNA strand breaks, although one of its<br />

oxidation product, peroxynitrite, is toxic and can cause DNA<br />

Fig. 9. DNA damage in EL-4 cells cultured for 2 h in presence of<br />

different conditioned media was measured using single-cell gel electrophoresis<br />

(‘‘comet assay’’) and expressed as tail length. Key: Lane<br />

1, control EL-4 cells; Lane 2, g-irradiated EL-4 cells; Lane 3, EL-4<br />

cells receiving medium from g-irradiated EL-4 cells; Lane 4, EL-4<br />

cells receiving medium from L-NAME treated and g-irradiated EL-<br />

4 cells; Lane 5, EL-4 cells receiving medium from cPTIO treated<br />

and g-irradiated EL-4 cells; Lane 6, EL-4 cells receiving medium<br />

from lipopolysaccharide (LPS)-stimulated (500 ng/mL), L-<br />

NAME-treated, and g-irradiated RAW 264.7 cells; Lane 7, EL-4<br />

cells receiving medium from LPS-stimulated (500 ng/mL),<br />

cPTIO-treated, and g-irradiated RAW 264.7 cells; Lane 8, EL-4<br />

cells treated with L-NAME; Lane 9, EL-4 cells treated with cPTIO.<br />

Data represent means SE of three independent experiments;<br />

significantly different from unirradiated controls: *p < 0.05,<br />

**p < 0.01.<br />

damage by both attacking deoxyribose and direct oxidation<br />

of purine and pyrimidine (29). Radiation-induced NO has<br />

the possibility of attacking nearby cells within their diffusion<br />

distance of less than 0.5 mm (30, 31). However, because of<br />

their having very short half-lives, they may not be direct contributors<br />

to the cellular damage in the bystander population.<br />

This is confirmed by the fact that inclusion of cPTIO in the<br />

medium only marginally inhibits the bystander response in<br />

similar cells (Fig. 8, Lane 5). Contrarily, it has also been reported<br />

that NO is involved in the bystander responses caused<br />

by the conditioned medium harvested from irradiated cells<br />

(32). Therefore some long-lived bioactive factors downstream<br />

of NO are most likely involved in this bystander<br />

response.<br />

REFERENCES<br />

1. Azzam EI, De Toledo SM, Spitz DR, et al. Oxidative metabolism<br />

modulates signal transduction and micronucleus formation<br />

in bystander cells from alphaparticle- irradiated normal human<br />

fibroblast cultures. Cancer Res 2002;62:5436–5442.<br />

2. Azzam EI, De Toledo SM, Gooding T, et al. Intercellular communication<br />

is involved in the bystander regulation of gene expression<br />

in human cells exposed to very low fluences of<br />

a particles. Radiat Res 1998;150:497–504.<br />

3. Mothersill C, Seymour C. Radiation-induced bystander effects,<br />

carcinogenesis and models. Oncogene 2003;22:7028–7033.<br />

4. Mothersill C, Seymour C. Radiation-induced bystander effects:<br />

Are they good, bad or both? Med Confl Surviv 2005;21:<br />

101–110.<br />

5. Azzam EI, De Toledo SM, Little JB. Oxidative metabolism, gap<br />

junctions and the ionizing radiation-induced bystander effect.<br />

Oncogene 2003;22:7050–7057.


1574 I. J. Radiation Oncology d Biology d Physics Volume 72, Number 5, 2008<br />

6. Hall EJ, Hei TK. Genomic instability and bystander effects induced<br />

by high-LET radiation. Oncogene 2003;22:7034–7042.<br />

7. Prise KM, Folkard M, Michael BD. Bystander responses induced<br />

by low LET radiation. Oncogene 2003;22:7043–7049.<br />

8. Matsumoto H, Hayashi S, Hatashita M, et al. Induction of radioresistance<br />

to accelerated carbon-ion beams in recipient cells by<br />

nitric oxide excreted from irradiated donor cells of human glioblastoma.<br />

Int J Radiat Biol 2000;76:1649–1657.<br />

9. Matsumoto H, Hayashi S, Hatashita M, et al. Induction of radioresistance<br />

by a nitric oxide-mediated bystander effect. Radiat<br />

Res 2001;155:387–396.<br />

10. Shao C, Furusawa Y, Aoki M, et al. Nitric oxide-mediated bystander<br />

effect induced by heavy-ions in human salivary gland<br />

tumour cells. Int J Radiat Biol 2002;78:837–844.<br />

11. Iluki Y, Mizuno S, Goto R. g-irradiation-induced DNA damage<br />

enhance NO production via NF-kB activation in RAW 264.7<br />

cells. Biochim Biophys Acta 2003;1593:159–167.<br />

12. Camphausen K, Moses MA, Menard C, et al. Radiation abscopal<br />

antitumor effect is mediated through p53. Cancer Res 2003;<br />

63:1990–1993.<br />

13. Maurya DK, Adhikari S, Nair CK, et al. DNA protective properties<br />

of vanillin against gamma-radiation under different conditions:<br />

possible mechanisms. Mutat Res 2007;634:69–80.<br />

14. Green LC, Wagner DA, Glogowski J, et al. Analysis of nitrate,<br />

nitrite and [15N] nitrate in biological fluids. Anal Biochem<br />

1982;126:131–138.<br />

15. Jeon SH, Kang MG, Kim YH, et al. A new mouse gene, SRG3,<br />

related to the SWI3 of Saccharomyces cerevisiae, is required for<br />

apoptosis induced by glucocorticoids in a thymoma cell line.<br />

J Exp Med 1997;185:1827–1836.<br />

16. Mitra AK, Krishna M. Radiation-induced bystander effect: activation<br />

of signaling molecules in K562 erythroleukemia cells.<br />

J Cell Biochem 2007;100:991–997.<br />

17. Nagasawa H, Little JB. Induction of sister chromatid exchanges<br />

by extremely low doses of a-particles. Cancer Res 1992;52:<br />

6394–6396.<br />

18. Brach MA, Hass R, Sherman ML, et al. Ionizing radiation induces<br />

expression and binding activity of the nuclear factor<br />

kB. J Clin Invest 1991;88:691–695.<br />

19. Ginn-Pease ME, Whisler RL. Redox signals and NF-kB activation<br />

in T cells. Free Radic Biol Med 1998;25:346–361.<br />

20. Zhou H, Ivanov VN, Lien YC, et al. Mitochondrial function and<br />

nuclear factor-kappa-mediated signaling in radiation induced<br />

bystander effects. Cancer Res 2008;68:2233–2240.<br />

21. Li N, Karin M. Ionizing radiation and short wavelength UV activate<br />

NF-kB through two distinct mechanisms. Proc Natl Acad<br />

Sci U S A 1998;95:13012–13017.<br />

22. Lowenstein CJ, Alley EW, Raval P, et al. Macrophage nitric oxide<br />

synthase gene: two upstream regions mediate induction by<br />

interferon g and lipopolysaccharide. Proc Natl Acad Sci U S A<br />

1993;90:9730–9734.<br />

23. Xie QW, Whisnant R, Nathan C. Promoter of the mouse gene<br />

encoding calcium-independent nitric oxide synthase confers inducibility<br />

by interferon g and bacterial lipopolysaccharide.<br />

J Exp Med 1993;177:1779–1784.<br />

24. Martinez-Ruiz A, Lamas S. Signalling by NO-induced protein<br />

S-nitrosylation and S-glutathionylation: Convergences and divergences.<br />

Cardiovasc Res 2007;75:220–228.<br />

25. Mothersill C, Seymour C. Medium from irradiated human epithelial<br />

cells but not human fibroblasts reduces the clonogenic<br />

survival of unirradiated cells. Int J Radiat Biol 1997;71:<br />

421–427.<br />

26. Kadhim MA, Moore SR, Goodwin EH. Interrelationships<br />

amongst radiation-induced genomic instability, bystander effects,<br />

and the adaptive response. Mutat Res 2004;568:21–32.<br />

27. Maragos CM, Morley D, Wink DA, et al. Complexes of NO<br />

with nucleophiles as agents for the controlled biological release<br />

of nitric oxide: Vasorelaxant effects. J Med Chem 1991;34:<br />

3242–3247.<br />

28. Nathan C. Nitric oxide as a secretory product of mammalian<br />

cells. FASEB J 1992;6:3051–3064.<br />

29. Tamir S, Tannenbaum SR. The role of nitric oxide (NO) in the<br />

carcinogenic process. Biochim Biophys Acta 1996;1288:<br />

F31–F36.<br />

30. Lancaster JR Jr. Diffusion of free nitric oxide. Methods Enzymol<br />

1996;268:31–50.<br />

31. Saran M, Bors W. Signaling by O2- and NO: How far can either<br />

radical, or any specific reaction product, transmit a message under<br />

in vivo conditions? Chem Biol Interact 1994;90:35–45.<br />

32. Shao C, Stewart V, Folkard M, et al. Nitric oxide-mediated signaling<br />

in the bystander response of individually targeted glioma<br />

cells. Cancer Res 2003;63:8437–8442.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!