164
Thesis on Cloning of Chicken Anemia Virus VP3 Gene in Replicase Based Eukaryotic Vector and Study of Its Apoptic Activity Submitted for the award of DOCTOR OF PHILOSOPHY Degree in Biotechnology By Priyanka Pal UNDER THE SUPERVISION OF Dr. Anant Rai Dr. Kusum Agarwal IBIT, Bareilly Shobhit University, Meerut Faculty of Biological Engineering Shobhit University, Meerut-250110 2012

Cloning of Chicken Anemia Virus VP3 Gene in Replicase ...shodhganga.inflibnet.ac.in/bitstream/10603/40041/1/... · Certificate This is to certify that the thesis entitled “Cloning

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Thesis on

Cloning of Chicken Anemia Virus VP3 Gene in

Replicase Based Eukaryotic Vector and Study of Its

Apoptic Activity

Submitted for the award of

DOCTOR OF PHILOSOPHY

Degree in

Biotechnology

By

Priyanka Pal

UNDER THE SUPERVISION OF

Dr. Anant Rai Dr. Kusum Agarwal

IBIT, Bareilly Shobhit University, Meerut

Faculty of Biological Engineering

Shobhit University, Meerut-250110

2012

TTTTo my parents, my first teacherso my parents, my first teacherso my parents, my first teacherso my parents, my first teachers

Certificate

This is to certify that the thesis entitled “Cloning of Chicken Anemia Virus VP3

gene in replicase based eukaryotic vector and study of its apoptic activity” submitted

by Ms. Priyanka Pal for the award of Degree of Doctor of Philosophy in the Faculty of

Biological Engineering of Shobhit University, Meerut, is a record of authentic work

carried out by her under my supervision.

To the best of my knowledge, the matter embodied in this thesis is the original

work of the candidate and has not been submitted for the award of any other degree or

diploma to any university or institution.

It is further certified that she has worked with me for the required period in the

Faculty of Biological Engineering (Centre for Biotechnology), Shobhit University,

Modipuram, Meerut.

Date: Prof. Kusum Agarwal

(Internal Supervisor)

Dr. Kusum Agarwal

M.Sc., Ph.D. Professor & Coordinator, Centre for Biotechnology,

Faculty of Biological Engineering, Shobhit University, Meerut

E-mail: [email protected]

Institute of Biotechnology & IT 197, Mudia Ahmadnagar, Bareilly-243122 (UP)

Ph: 09219661948, 0581-2602907, Fax: 0581-2602035

www.ibit.org.in [email protected]

Prof. Anant Rai Ph.D. Director

____________________________________________

Certificate

This is to certify that the thesis entitled “Cloning of Chicken Anemia Virus

VP3 gene in replicase based eukaryotic vector and study of its apoptic activity”

submitted by Ms. Priyanka Pal for the award of Degree of Doctor of Philosophy in

the Faculty of Biological Engineering of Shobhit University, Meerut, is a record of

authentic work carried out by her under my supervision.

To the best of my knowledge, the matter embodied in this thesis is the

original work of the candidate and has not been submitted for the award of any

other degree or diploma.

It is further certified that she has worked with me for a period of 43 months in

the Department of Biotechnology of Institute of Biotechnology and IT, Bareilly.

Date: Prof. Anant Rai (External Supervisor)

Declaration

I, hereby, declare that the work presented in this thesis entitled

“Cloning of Chicken Anemia Virus VP3 Gene in Replicase Based

Eukaryotic Vector and Study of its Apoptic Activity” in fulfillment of

the requirements for the award of Degree of Doctor of Philosophy,

submitted in the Faculty of Biological Engineering at Shobhit

University, Modipuiram, Meerut, is an authentic record of my own

research work carried out under the supervision of Prof. Kusum

Agarwal and Prof. Anant Rai.

I also declare that the work embodied in the present thesis

(i) is my original work and has not been copied from any

Journal/thesis/book, and

(ii) has not been submitted by me for any other Degree or

Diploma of any university/ institution.

(Priyanka Pal)

Acknowledgement

First and foremost I would like to give glory to Almighty for his

grace and help in all my endeavors and for bringing me this far in my

academic career.

I would like to express my deep sense of gratitude to Internal-

Supervisor Dr. Kusum Agarwal, Professor & Coordinator, School of

Biotechnology, Shobhit University, Meerut, for her excellent

cooperation and encouragement. I appreciate all her contributions of

time and ideas to make my Ph.D. experience productive and

stimulating. The joy and enthusiasm she has shown was contagious and

motivational, even during tough times in the pursuit of my Ph.D. work.

I would like to express my sincere gratitude to my External-Supervisor

Dr. Anant Rai, Director, Institute of Biotechnology and IT, Bareilly

(former Principal Scientist & Head, Division of Animal Biotechnology,

Indian Veterinary Research Institute, Izatnagar, Bareilly), for guiding

me with attention and care. He has taken pains in going through the

project and made necessary corrections as and when needed.

I express my thanks to Honorable Chancellor Kunwar Shekhar

Vijendra and Honorable Vice Chancellor Prof. R. P. Agarwal, of

Shobhit University, for providing congenial environment for

conducting research in the university.

I gratefully acknowledge the kind and liberal help from Prof.

D.V. Rai, Dean, Faculty of Biological Engineering, Prof. Ranjit Singh,

Former Director, School of Pharmaceutical Sciences, Dr. Rekha Dixit,

Dr. Jayanand and Dr. Manish Kumar Gupta, Shobhit University,

Meerut, for their valuable suggestions and discussions during my work.

I am extremely thankful to Dr. D.V. Rai, Dean, Faculty of

Biological Engineering, Shobhit University, Meerut, and Mrs. Soni

Gangwar, Assistant Professor (Biotechnology), Institute of

Biotechnology & IT, Bareilly for providing assistance in the laboratory.

I thank my fellow lab mates Usha Tiwari, Rajveer Maurya and

Nitin Sharma for the stimulating discussions, I had with them and for

advising me from time to time for all those days we worked together.

Last but not the least my very special and heartfelt gratitude is

due to my parents, brothers and other family members for their love,

support and encouragement without which this work would have been

a distant reality. I also thank my loving, supportive, encouraging, and

patient husband for his faithful support during the final stage of my

Ph.D. work.

(Priyanka Pal)

Research Papers Published / Accepted / Communicated in Refereed Journals/Proceedings

1. Pal, P., Agarwal, K., Rai, A. and Tiwari, U. (2011): Cloning of Chicken

Anemia Virus VP3 Gene in pSinCMV vector. Biotechnology

International, 4, 16-21. (Online www.bti.org.in).

2. Pal, P., Agarwal, K., Rai, A. and Tiwari, U. (2011): Expression of

Recombinant Plasmid Containing Chicken Anemia Virus VP3 Gene in

HeLa Cells. Research journal of Pharmaceutical Science and

Biotechnology, 1, 41-43. (Online www.rjpsb.info).

3. Pal, P., Agarwal, K., Rai, A. and Tiwari, U. (2010): Isolation of

recombinant plasmid containing Chicken Anemia Virus VP3 gene,

accepted for publication in Transaction of physical & life science.

(Accepted).

4. Pal, P., Agarwal, K., Rai, A. and Tiwari, U. (2012): Apoptic activity of

chicken anemia virus VP3 gene cloned in replicase based eukaryotic

vector. Communicated for publication in The journal of microbiology,

biotechnology and food sciences.

Participation in the Conferences/Workshops/Seminars

1. Attended national Seminar-Cum Workshop on “Biomedical Research

and Clinical Applications of Radioisotopes”, Shobhit University,

Meerut, March 31- 01 April 2012.

2. Attended International Conference on “Indian Civilization Through

the Millennia” held at JambooDweep, Hastinapur, March 2-3, 2012.

3. Attended annual Conference on Vijnana Parishad of India and the

Global Society of Mathematical & Allied Sciences, held at School of

Basic and Applied Sciences, Shobhit University, Meerut, March 24-26,

2011. Presented paper entitled “Expression of Recombinant Plasmid

containing Chicken Anemia Virus VP3 Gene in HeLa Cells.”

4. Attended Conference on “Genes & Genomics: Qualitative &

Quantitative Approach” held at School of Biotechnology, Shobhit

University, Meerut, September 11-12, 2010. Presented paper entitled

“Construction of replicase based Eukaryotic pSinCMV vector

containing Chicken Anemia Virus VP3 Gene” and “Preparation of

Recombinant Plasmid pSinCMV containing Human Erythropoietin

Gene”.

5. Attended Workshop on “Nanomaterials: Recent Techniques and

Applications”, Shobhit University, Meerut, March 27, 2010.

Contents

List of Tables (i)

List of Figures (ii)

Abbreviations (iii)

Chapter 1: Introduction 1-5

Chapter 2: Review of Literature 6-47

2.1 Chicken Anemia Virus 6 2.2 Apoptosis 14 2.3 Apoptin 23 2.4 Apoptin from other viruses 43 2.5 DNA Damage Signaling 46

Chapter 3: Material and Methods 48-65

3.1 Materials 48 3.1.1 Vector 48 3.1.2 Gene 48 3.1.3 Primers 48 3.1.4 Host Bacterial Strains 50 3.1.5 Cell Culture 50 3.1.6 Conjugates 50 3.1.7 Hyper-immune Serum 50 3.1.8 Experimental Animals 51 3.1.9 Glasswares, Plastic wares and Chemicals 51 3.1.10 Solutions and Buffers 51

3.2 Methods 51 3.2.1 Revival of the E. coli culture containing

recombinant plasmid with VP3 gene 51

3.2.2 Isolation of plasmid DNA by TELT method 52 3.2.3 Agarose gel electrophoresis 52 3.2.4 Digestion of the recombinant plasmid with

restriction enzyme to release the gene insert using protocol of Sambrook & Russell (2001)

53

3.2.5 DNA extraction from agarose gel 53 3.2.6 Blunting of EcoRI generated VP3 gene 54

staggered ends 3.2.7 Purification of blunted insert 55 3.2.8 Creation of cloning site in vector 55 3.2.9 Dephosphorylation of 5’ ends 56 3.2.10 Purification of dephosphorylated linearized

vector 57

3.2.11 Blunt end ligation of pSin vector and VP3 gene

57

3.2.12 Transformation of E. coli (DH5α) cells with the ligated product

57

3.2.13 Screening of recombinant clones for VP3 gene in right orientation

58

3.2.14 Transfection of HeLa cells 61 3.2.15 Immunoperoxidase test for DNA expression 61 3.2.16 Sodium dodecyl sulfate-polyacrylamide gel

electrophoresis (SDS-PAGE) 62

3.2.17 DNA fragmentation assay by agarose gel electrophoresis

62

3.2.18 Caspase 3 detection assay 63 3.2.19 Annexin V binding assay 64

Chapter 4: Results 66-80

4.1 Cloning of VP3 gene in pSin vector 66 4.2 Expression of VP3 gene in HeLa cells 71 4.3 Apoptic activity of VP3 in cell line 75

Chapter 5: Discussion 81-92

Chapter 6: Summary and Conclusion 93

Chapter 7: Future Scope 94-95

References 96-115

Appendix 116-120

i

List of Tables

Table No. Title Page No.

1 Primers with their sequences 50

2 Reaction mix for blunting 54

3 Reaction mix for StuI digestion 56

4 Reaction mix for dephosphorylation 56

5 Ligation mixture 57

6 Reaction mix for BglII digestion 59

7 Reaction mix for PCR 60

ii

List of Figures

Fig. No. Title Page No.

1 pSin Vector 49

2 Digestion of recombinant plasmid (pTarget.cav.vp3) with restriction enzyme (EcoRI) to release the vp3 gene insert.

68

3 The digestion of pSin.cav.vp3 with BglII yielded three segments of expected size.

69

4 PCR amplification of cav.vp3 gene. 70

5 HeLa cells transfected with pSin.cav.vp3 rplasmidshowing positive IPT test, indicating expression of gene.

72

6 Healthy control HeLa cells with no color reaction.

73

7 VP3 expressed protein analyzed by SDS-PAGE.

74

8 DNA fragmentation assay. 77

9 Caspase positive HeLa cells showing green fluorescence, indicating positive apoptic reaction.

78

10 Caspase positive HeLa cells showing bright green fluorescence, indicating positive apoptic reaction.

79

11 Annexin-V positive HeLa cells showing bright green fluorescence, indicating positive apoptic reaction.

80

iii

Abbreviations

0C Degree Celsius

% Percent

AA Amino Acid

Ab Antibody

CAV Chicken Anemia Virus

DMEM Dulbecco's Modified Eagle Medium

DNA Deoxyribonucleic acid

dNTP Deoxy ribonucleoside triphosphate

EDTA Ethylene DiamineTetraacetic Acid

Etbr Ethidium Bromide

Fig. Figure

CIAP Calf Intestinal Phosphatase

LB Luria Bertani

MW Molecular Weight

NaCl Sodium Chloride

FCS Fetal calf serum

No. Number

PBS Phosphate Buffered Saline

PCR Polymerase Chain Reaction

pH Log10 hydrogen ion concentration

RE Restriction endonuclease

RNA Ribonucleic acid

Rpm Revolution per minute

TAE Tris-acetate-EDTA

TE Tris-EDTA

UV Ultraviolet

v/v Volume by volume

w/v Weight by volume

iv

Units of Measurement

µg Microgram

µl Mirolittre

0C Degree Celsius

bp Base pair

g Gram

hr Hour

hrs Hours

Kb Kilo base

kDa Kilo Dalton

M Molar

mg Milligram

ml Milliliter

mM Millimolar

N Normal

Ng Nanogram

pmol Picomole

sec Second (s)

U Unit (s)

Introduction

Chapter-1

1

Chapter-1

Introduction

Kerr et al. (1972) described distinct morphologic changes of dying

cells and called this phenomenon as apoptosis. The term was coined on

the basis of the fact that release of apoptic bodies by dying cells resembled

with the picture of falling leaves from decidous trees, called in Greek

“apoptosis.”

The apoptotic downfall of a cell resulted in the formation of small

membrane bound entities known as apoptotic bodies. These bodies pinch

off from the dying cell and were consumed by the phagocytic action of

neighboring cells. This engulfment provided means for the dissemination

of the virus without initiating a concomitant host response, which follows

the discharge of progeny into the extracellular solution (Teodoro and

Branton, 1997).

Apoptosis or programmed cell death is a genetically determined

process to destroy cells for maintaining the cellular homeostasis in the

tissue. There is a family of cysteine proteases known as caspases, or

cysteine-aspartic proteases or cysteine-dependent aspartate-directed

proteases that play a vital role in apoptosis (programmed cell death).

Activation of cysteine called caspases plays the main role in the execution

of apoptosis. These activated caspases selectively cut cellular proteins,

which resulted in apoptotic morphology (internucleosomal fragmentation

2

of DNA into 180-200 base pair pieces, shrinkage of the cell as well as

fragmentation of the cell into apoptic bodies) and death of the cell. Two

pathways of caspase activation were reported. The first was through

triggering of cellular death-receptor superfamily and the second was

mitochondrial pathway induced by the changes of the expression of

proto-and anti-apoptic genes in the cell. It leads to the discharge of

cytochrome c and apoptosis inducing factor from mitochondria

(Lesauskaite and Ivanoviene, 2002).

Schmitz et al. (2000); Lauber et al. (2004) defined apoptosis as a

physiological form of cell death characterized via nuclear chromatin

condensation, cytoplasmic shrinking and memberane blebbing. This form

of programmed cell death was mainly induced by cancer treatment

(Kawanishi and Hiraku, 2004; Wesselborg and Lauber, 2005).

Chicken anemia virus, or CAV, is a virus that affects poultry. CAV

causes anemia, bone marrow atrophy, and severe immunosuppression.

Clinical signs of infection of CAV were mainly found in youthful chicks

due to maternal antibodies present mainly in adult chickens (Sommer

and Cardona, 2003).

The chicken anemia virus protein apoptin induced a p53-

independent, Bcl-2-insensitive type of apoptosis in a variety of human

tumor cells. Apoptin induced apoptosis in human transformed and

malignant cells except in normal cells. It was observed that apoptin failed

to induce programmed cell death in normal lymphoid, dermal, epidermal,

and smooth-muscle cells. Long-term expression of apoptin in normal

human fibroblasts showed that apoptin had no toxic or transforming

3

activity in these cells. In normal cells, apoptin was observed

predominantly in the cytoplasm, whereas in transformed and malignant

cells it was located in the nucleus, which recommended that the

localization of apoptin was related to its activity. These properties make

apoptin a potential agent for the treatment of a large number of tumors

(Oorschot et al., 1997).

It was found that a kinase activity present in cancer cells but

negligible in normal cells regulated the activity of apoptin via

phosphorylation. Apoptin can be activated, in normal cells, by transient

transforming signals conferred by ectopically expressed SV40 LT antigen,

which rapidly induced apoptin's phosphorylation, nuclear accumulation

and the ability to induce apoptosis. In normal cells where such signals

were not received, apoptin became aggregated, epitope-shielded and

were eventually despoiled in the cytoplasm. The mechanism of apoptin-

induced apoptosis could lead to the detection of novel tumor-specific

pathways that might be exploitable as anti-cancer drug targets (Rohn et

al., 2004).

It was observed that apaptin induced tumor specific apoptosis,

which was linked with the nuclear localization of the protein in tumor

cells, while in normal human cells, apoptin was detected mostly in the

cytoplasm and did not induce apoptosis. Using a recombinant adenovirus

expressing apoptin, it was found that apoptin induced G(2)-M cell cycle

arrest and chromatin condensation in cancer cells. Adenovirus mediated

apoptin expression also induced G(2)-M arrest in normal cells. In normal

4

cells apoptin was restricted chiefly in the cytoplasm but was also found in

the nucleus of a subset of cells (He et al., 2005).

Wang et al. (2005) also showed that apoptin could not induce

apoptosis in normal cells, but could induce in transformed cells or tumor

cells. The tumor specificity of apoptin was linked to its subcellular

localization. In tumor cells, apoptin migrated to the nuclei, whereas in

non-transformed cells, it remained chiefly within the cytoplasm.

Phosphorylation was accountable for the nuclear localization of apoptin.

Apoptin was phosphorylated in tumor cells, then translocated into the

nuclei, and induced cell apoptosis. Apoptin-induced apoptosis was

independent of functional p53, and colud not be inhibited by

overexpression of Bcl-2 and Bcl-xL, but caspase-3 activation was essential

for apoptin-induced rapid apoptosis. It was possible that apoptin's ability

to bind DNA was closely linked to its aptitude to induce apoptosis.

Liu et al. (2006) showed involvement of sphingolipids in apoptin

induced cell killing. Apoptin activated acid sphingomylinase (ASMase),

which generated ceramide; in turn ceramide acted as a second messenger

and signaled apoptotic response.

The reaction of the cell to DNA damage was a complex procedure,

which included damage signaling, repair, apoptosis or cell death. It was

connected with serious changes in the cell nucleus, which were caused by

post-translational modifications and active relocalizations of proteins as

well as alterations in the expression of many genes. Those changes were

not limited to the sites of DNA damage, but concerned whole cell nucleus,

including its domains: Promyelocytic leukemia (PML) bodies, Cajal

5

bodies and nucleolus. Promyelocytic leukemia (PML) nuclear bodies were

dynamic macromolecular multiprotein complexes that employ and

release a plethora of proteins. A considerable number of PML nuclear

body components play vital role in apoptosis, senescence regulation and

tumor suppression. Cajal bodies were little nuclear organelles with a

number of nuclear functions (Wysokinski et al., 2010).

Keeping the above facts in view, the present study has been designed

with the following objectives:

1. Cloning of chicken anemia virus VP3 gene in pSin mammalian expression vector.

2. In vitro expression of VP3 gene in cultured cells.

3. Study of apoptic activity of apoptin in cell line.

Review of Literature

2.1 Chicken anemia virus

2.2 Apoptosis

2.3 Apoptin

2.4 Apoptin From Other Viruses

2.5 DNA Damage Signaling

Chapter-2

6

Chapter-2

Review of Literature

2.1 Chicken Anemia Virus

Circoviruses are small, non-enveloped, icosahedral viruses that are

circular, single-stranded DNA genomes. Their genomes are also the

smallest possessed by animal viruses. The circovirus family consists of

three members- chicken anemia virus, porcine circovirus and psittacine

beak and feather disease virus, with pigeon circovirus being classified as

an unsure member. Infections with each of the four circoviruses were

linked with potentially fatal diseases in which destruction to lymphoid

tissue and immunosuppression by a virus were common features.

Knowledge with other animal virus families suggested that additional

animal species will be infected by, and yet undiscovered, circoviruses and

that these might display similar tissue tropism and disease-causing

potential. Novel circoviruses might infect other avian species, including

commercial poultry (Todd, 2000).

Chicken anemia virus (CAV) seemed to have a worldwide

distribution. CAV caused a syndrome in chickens which was identified by

increased mortality, anemia associated with atrophy of the hematopoietic

tissues in the bone marrow, subcutaneous and intramuscular

hemorrhages, and atrophy of the lymphoid system. CAV was found to be

7

spread both vertically and horizontally. Vertical transmission occurred

following primary infection of in-lay breeding stock, and resulted in

clinical disease in their progeny around 2 weeks of age. Horizontal spread

was found to result mostly in subclinical disease. Both clinical and

subclinical diseases caused economic loss as indicated by a vaccine (Todd

et al., 1990; McNulty et al., 1991).

Circular double-stranded replication intermediates were

identified in low-molecular-weight DNA of cells of the avian leukemia

virus-induced lymphoblastoid cell line 1104-X-5 infected with CAV.

Linearized CAV DNA was cloned into the vector pIC20H, to characterize

the genome of CAV. A cytopathogenic effect was caused when

transfection of the circularized cloned insert was done into chicken cell

lines, which was arrested when chicken serum with neutralizing

antibodies directed against CAV was added. Chickens inoculated with

CAV collected from cell lines transfected with cloned CAV DNA

developed clinical signs of CAV. The CAV DNA sequence had three

partially overlapping major reading frames coding for putative peptides

of 51.6, 24.0, and 13.6 kDa. The CAV genome probably contained only

one promoter region and only one poly(A) addition signal. Southern blot

analysis by means of oligomers derived from the CAV DNA sequence

showed that infected cells contained double and single-stranded CAV

DNA, while purified virus contained only the minus strand. It was for

the first time that the genome of one of the three known single-stranded

circular DNA viruses had been completely analyzed (Noteborn et al.,

1991).

8

CAV was initially isolated in Japan and the associated disease

chicken infectious anemia was described in 1979. The virus had a world-

wide distribution and was common in intensive poultry raising areas.

CAV was believed to play an important role in several multiple etiology

disease syndromes; hemorrhagic syndrome; hemorrhagic anemia

syndrome, aplastic anemia, gangrenous dermatitis, anemia dermatitis,

hemorrhagic aplastic anemia syndrome and blue wing disease.

Horizontal spread seemed to be less important than vertical

transmission. The most characteristic lesion was yellow fatty bone

marrow and the most consistent finding was thymic atrophy. Thymic

and bone marrow intranuclear inclusion bodies occurred with infection

but were of limited value diagnostically and were very transient which

were seen rarely (Pope et al., 1991).

The CAV genome contained three partially or completely

overlapping genes. It caused cytopathogenic effects which were fatal

in chicken thymocytes and cultured transformed mononuclear cells via

apoptosis. In transformed chicken cells, the synthesis of the VP3 gene

product apoptin mimiced the CAV-induced apoptosis, which was p53

independent. Apoptin induced apoptosis in human tumor cells but, not

in normal cells. These properties suggested that apoptin might have

potential medical applications (Noteborn et al., 1998).

TT virus (TTV) was the only known human virus with single-

stranded circular DNA, with a relationship to CAV of the

family Circoviridae. A new human virus was designated as TTV-like mini

virus (TLMV) was reported, which resembled TTV and CAV. This non-

9

enveloped virus was smaller but had a similar density to TTV, when a

TLMV/TTV-coinfected plasma was analyzed. Full-length sequencing

revealed that the TLMV genome was a circular DNA consisting 2860 nt,

significantly shorter than TTV but longer than CAV. TLMV was observed

to be similar to both TTV and CAV in genomic organization. The

untranslated region of TLMV resembled CAV in that both had direct

repeats, whereas the sequence homology was more evident between

TLMV and TTV. TLMV was an intermediate between the remotely

related TTV and CAV. Since CAV differed much from other circoviruses,

it might better be classified jointly with TTV and TLMV under a new

family (Takahashi et al., 2000).

Aldair et al. (2000) also analysed a disease in young chicks which

was caused by CAV. They also observed generalised lymphoid atrophy,

increased mortality and severe anemia. The virus targeted erythroid and

lymphoid progenitor cells in bone marrow and thymus respectively. The B

cells in the chicken were not susceptible to CAV infection and were not

directly affected by the virus. Severe anemia, and depletion of

granulocytes and thrombocytes was resulted by destruction of erythroid

progenitors in bone marrow. Destruction of precursor T cells resulted in

depletion of mature cytotoxic and helper T cells with consequent effects

on susceptibility to, and enhancement of, the pathogenicity of secondary

infectious agents, and suboptimal antibody responses. It appeared that

apoptin was a feature of the lymphocyte depletion in the thymic cortex,

which might be mediated by one of the non-structural viral proteins, VP3

(apoptin).

10

A serologic survey in unvaccinated broiler parent and broiler

progeny flocks demonstrated seroconversion against chicken infectious

anemia virus in all parent flocks. The presence of CAV antibodies at

slaughter of broilers was positively correlated with the slaughterhouse

condemnation rates. The results showed that CAV infections were highly

prevalent in both broiler parent and broiler flocks. (De Herdt et al., 2001).

CAV was isolated from a chicken flock in Harbin of China. By

polymerase chain reaction and sequence, the genome of the virus was

cloned and analyzed. The genome was made of 2298 base pairs including

three overlapping open reading frames (vp1, vp2, vp3) and a regulative

region (He et al., 2002).

From different geographical regions of India, four CAV isolates

(CAV-A, -B, -E and -P) were recovered and characterized. The nucleotide

and deduced amino acid sequence of the Indian CAV isolates were

associated and compared with other isolates of CAV from different

countries like Europe, Asia, America and Australia. Phylogenetic

analysis of the Indian CAV isolates was done, based on the nucleotide

and deduced amino acid sequences. It was observed that Indian isolates

were genetically evolved from different parts of the world. The study

showed precious information on molecular epidemiological status of

CAV isolates prevalent in India for the first time (Natesan et al., 2006).

A field strain C14 of CAV was isolated from a 14 day-old broiler

flock with growth runting syndromes. Antibody reactions to inactivated

vaccines to avian influenza viruses (AIV) were suppressed in specific

pathogen free (SPF) chickens inoculated with C14 strain CAV at 1 day-

11

old. When chickens were infected with C14 strain and

reticuloendotheliosis viruses, it demonstrated a synergism in

immunosuppression. The viral genomic DNA was amplified by PCR in 3

overlapped fragments and PCR products were cloned into T-vector

plasmid for sequencing. The sequencing results showed that the total

genome of C14 strain CAV was 2298 bp which contained 3 overlapped

open reading frame and 1 non-coding regulation fragment. Its whole

genome had 97.2% - 99.2% of homogeneity to other several published

CAV reference strains (Li et al., 2007).

A total of 12 CAV isolates from different commercial broiler breeder

farms were isolated and characterized. High level of occurrence of vertical

transmission of viral DNA to the progeny was indicated when detection of

CAV positive embryos by the PCR assay was done in the range of 40 to

100% for different farms. By indirect immunoperoxidase staining, CAV

antigen was detected in the thymus and in the bone marrow but not in

spleen, liver, duodenum, ovary and oviduct. Based on partial sequences of

VP1 gene, the 12 CAV isolates were characterized. Six isolates (MF1A,

M3B5, MF3C, NF4A, P24A and P12B) were found to have maximum

homology with previously characterized Malaysian isolate SMSC-1, four

isolates (NF3A, M1B1, PYT4 and PPW4) with isolate BL-5 and the

remaining two (NF1D and NF2C) had maximum homology both with

isolates 3-1 and BL-5. With amino acid profile of 75-I, 97-L, 139-Q and 144-

Q, seven of the isolates were clustered together in cluster I with other

isolates from different geographical places. The observation demonstrated

that CAV was widespread in the studied commercial broiler breeder farms

(Hailemariam et al., 2008).

12

Cancer remains one of the leading causes of death worldwide. By

combining numerous tumor-specific and enhanced targeting signals into

a single modular multifaceted approach, it might be possible in future to

achieve the desired outcome, without any unwanted bystander effects,

the delivery of cytotoxic drugs/DNA directly to the nucleus specifically

within tumor cells. To achieve this, modules such as the tumor-specific

nuclear targeting signal of the chicken anemia virus apoptin protein

represented exciting possibilities (Wagstaff and Jans, 2009).

CAV has gained much importance as an immunosuppressive and

economically important emerging pathogen of poultry. The virus has

been detected in recent years which was isolated from poultry flocks of

India. The first sero-epidemiological investigation of the presence of

CAV infection was reported in poultry flocks of the country. A total of

404 serum samples were collected from chicken flocks of eleven poultry

farms, which contained a total of 0.34 million birds from four Northern

states and were also suspected of having chicken infectious anaemia

(CIA). By using a commercially available enzyme-linked immunosorbent

assay (ELISA) kit, screening of the sera samples was done which revealed

351 serum samples (86.88%) to be positive for CIAV antibodies. A high

prevalence rate of CIAV recorded in the analysis, along with previous

virus detection reports, indicated the extensive distribution of the virus.

So, CIAV should be considered an economically important pathogen of

poultry, affecting poultry industry of India (Bhatt et al., 2011).

Using PCR, the presence of CAV infection genetic variability of

isolated strains based on restriction of VP1 gene by Mbo1 and apoptotic

13

changes in the CAV positive broiler chickens in upper Egypt, were

investigated. The data indicated that infection with CAV induced

apoptosis in lymphoid cells that might affect on cellular immunity. The

findings highlighted the significance of CAV, therefore, focus should be

made on its epidemiology in the upper Egypt and to further develop and

apply reliable diagnostic tools as well as molecular studies so as to

suggest suitable prevention strategies for the economical important avian

pathogen (Mohamed, 2010).

Previously, the CAV putative intergenotypic recombinants have

been reported. The fact was based on the previous classification of CAV

sequences into three genotypes. Together with a variety of computational

recombination detection algorithms, phylogenetic analysis was used to

investigate CAV approximately full genomes. Significant evidence of

intersubtype recombination was detected in the parent-like and two

putative CAV recombinant sequences. The event was shown to occur

between CAV subgroup A1 and A2 sequences in the phylogenetic trees. It

was revealed that intersubtype recombination in CAV genome sequences

played a role in generating genetic diversity within the natural population

of CAV (Eltahir et al., 2011).

Cancer is a complex progressive multistep disorder that results

from the accumulation of genetic and epigenetic abnormalities, which

show the way to the transformation of normal cells into malignant

derivatives. There were few examples of therapies leading to cure the

cancer. Still cancer remained to be one of the largest causes of death

worldwide. The strategies were analysed which were designed to induce

14

tumor-selective death such as the use of oncolytic virus, tumoricidal

proteins (NS1, Eorf4, apoptin, HAMLET (human α-lactalbumin made

lethal to tumor cells)) and activation of signaling pathways which were

involved in tumor surveillance. The potential of the tumor necrosis factor-

related apoptosis-inducing ligand (TRAIL) pathway, was emphasized. An

essential component of the evolutionary developed defense systems that

eradicated malignant cells. The necessity of targeting tumor-initiating cells

(TICs) was discussed to avoid relapse and increased the chances of

complete remission. The emerging concepts that might provide novel

avenues for cancer therapy were also described (Pavet et al., 2011).

2.2 Apoptosis

The transformed cancer cells were more resistant to apoptosis, so

the ability to survive was enhanced in them. They exhibited resistance to

anticancer drugs, they no longer depend on survival signals, and they

could metastasize. Therefore, cancer progressed as the cancer cells

maintained the proliferative supremacy they acquired from their

oncogenes. In other words, when cancer cells became resistant to

apoptosis, they became resistant to treatment, metastasized, and

proliferated destructively (Kataoka and Tsuruo, 1996).

Kerr, (2000) observed a curious form of liver cell death during his

studies of acute liver injury in rats, and his further study led to an increase

in understanding the role of apoptosis in embryogenesis,

spermatogenesis, cancer growth, and tissue remodelling during healing or

functional regression.

15

Defects in the regulation of apoptosis played crucial role in the

pathogenesis and progression of most cancers and leukemias. Apoptosis

defects figured high resistance to radiotherapy, chemotherapy, hormonal

therapy, and immune-based treatments. Apoptosis was caused by

activation of intracellular proteases, known as caspases that were

responsible for the morphologic and biochemical events that

characterized the apoptotic cell. Numerous proteins which regulated the

cell death proteases have been discovered. Those proteins included

proteins belonging to the Bcl-2, inhibitor of apoptosis, caspase-associated

recruitment domain, death effector domain, and death domain families

(Reed, 2004).

Ghobrial et al. (2005) also defined apoptosis, or programmed cell

death as a mechanism by which cells undergo death to control cell

proliferation or in response to DNA damage. The basis for novel targeted

therapies that could induce death in cancer cells or sensitize them to

establish cytotoxic agents and radiation therapy could be provided by the

understanding of apoptosis. These novel agents include those targeting

the extrinsic pathway such as tumor necrosis factor-related apoptosis-

inducing ligand receptor 1, and those targeting the intrinsic Bcl-2 family

pathway such as antisense bcl-2 oligonucleotides. Numerous pathways

and proteins controlled the apoptosis machinery. Examples included were

p53, the nuclear factor kappa B, the pathway of ubiquitin/proteosome

and the pathway of phosphatidylinositol 3 kinase. These could be

targeted by specific modulators like bortezomib, and mammalian target of

rapamycin inhibitors like CCI-779 and RAD 001. Because these pathways

16

might be preferentially altered in tumor cells, there was potential for

selective effect in tumors sparing the normal tissue.

Apoptosis was known to be coupled to multiple signalling

pathways. Identification of vital points in the regulation of apoptosis was

of major interest both for the understanding of control of cell fate and for

the discovery of new pharmacological targets, mainly in oncology.

Indeed, defects in the implementation of apoptosis were known to

participate in tumour initiation and progression as well as in

chemoresistance. The Bcl-2 family members represented essential

intracellular players in the apoptotic mechanism. Those proteins were

either pro or anti-apoptotic, they interacted with each other to control

apoptosis. Inhibiting the heterodimerisation between pro- and anti-

apoptotic members was sufficient to promote apoptosis in mammalian

cells. Small molecules, antagonists or peptidomimetics inhibiting this

heterodimerisation, represented a therapeutic prototype targeting the

apoptotic cascade. They induced cell death by activating directly the

mitochondrial apoptotic trail (Mazars et al., 2005).

Contrary to earlier assumptions, Bcl-2 and Bcl-xL inhibited apoptin-

induced cell death in several tumor cell lines. In contrast, deficit of Bax

conferred resistance, whereas Bax expression sensitized the cells to

apoptin-induced death. Cell death induction by apoptin was linked with

cytochrome c release from mitochondria as well as with caspase-3 and -7

activation. Benzyloxycarbonyl-Val-Ala-Asp-fluoromethyl ketone, a broad

spectrum caspase inhibitor, was very protective against apoptin-induced

cell death. Apoptosis induced by apoptin required Apaf-1, as

17

immortalized Apaf-1-deficient fibroblasts as well as tumor cells devoid of

Apaf-1 were strongly protected. Thus, data revealed that apoptin-induced

apoptosis was not only Bcl-2- and caspase dependent, but also occupied

an Apaf-1 apoptosome-mediated mitochondrial fatality pathway (Burek

et al., 2006).

The molecules that destroyed cancer cells selectively were: apoptin,

viral protein R (VpR), E4orf4 and Brevinin-2R. Apoptin's cancer-selective

toxicity depended on its nuclear localization both nuclear import and

export (Maddika et al., 2006).

Plasmodium falciparum in a subset of patients could guide to cerebral

malaria (CM), a major provider to malaria-associated mortality. Despite

treatment, CM mortality could be as high as 30%, while 10% of survivors

of the disease might experience short and long-term neurological

complications. The pathogenesis of CM was mediated by alterations in

cytokine and chemokine homeostasis. The hypothesis for this study was

that CM-induced changes in inflammatory, apoptotic and angiogenic

factors. Their recognition enabled the development of new prognostic

markers and adjunctive therapies for preventing CM mortalities (Jain et

al., 2008).

Anticancer drug-induced tumor suppression might involve

mechanisms of protection against neoplastic transformation that were

normally dormant in mammalian cells and consisted in a genetic program

implemented during anti-tumoral defense. This defense program

consequenced in the self elimination of cells harboring potentially

dangerous mutations by triggering cell death through apoptosis and/or

18

autophagy or in the implementation of a program that lead to a

permanent growth arrest known as senescence. These responses were

considered vital tumor suppressive mechanisms and their study appeared

to be necessary to develop therapeutical events based on the enhancement

of the different responses. Interesting evidence showed that different

drugs induced senescence or cell death depending on the genetic features

of the tumor cells as well as on the integrity of the relative pathways

(Chiantore et al., 2009).

While investigating the mechanisms underlying cell death during

wound healing processes, the pro-apoptotic effects of basic fibroblast

growth factor (bFGF) on granulation tissue fibroblasts following

pretreatment with transforming growth factor (TGF)-beta1 in vitro were

undiscovered. Fibroblasts that had been treated and originated from the

uninjured dermis did not show apoptosis, indicating that the

mechanisms underlying apoptosis events in fibroblasts that originate

from normal dermal and wound tissues differ. In this process, it was

established that bFGF inhibited Akt phosphorylation at serine 473 and

induced a rapid failure of phosphorylation of focal adhesion kinase

(FAK) at tyrosine 397 in pretreated GF-1 cells, an incident that coincided

with the dissociation of phosphorylated FAK from the focal adhesions.

Therefore, inhibition of survival signals relayed via the disrupted focal

adhesion structures and inactivated Akt following bFGF treatment might

lead to apoptosis in GF-1 cells pretreated with TGF-beta1. Pretreatment

of GF-1 with TGF-beta1 followed by adding of bFGF resulted in

significantly enhanced inhibition of phosphorylation of Akt and FAK

compared to treatment with TGF-beta1 or bFGF alone. The combinatorial

19

treatment also led to proteolysis of FAK and inhibition of FAK and Akt

protein expression in GF-1 cells. These conclusions demonstrated

important role for the two cytokines in apoptosis of granulation tissue

fibroblasts during wound healing. In vivo studies also established a

marked decline in phosphorylation and protein expression of Akt and

FAK in bFGF-injected skin wounds. These consequences led to the

assumption that temporal activation of TGF-beta1 and bFGF at the injury

site promoted apoptosis in granulation tissue fibroblasts, an event that

was vital for the termination of proliferative granulation tissue formation

(Akasaka et al., 2010).

For chronic hyperglycemia, INS-1 cells were cultured for 5 days

with changes of RPMI 1640 medium containing 33 mM glucose every 12

hours. For irregular hyperglycemia, the medium containing 11 mM

glucose was exchanged with the medium containing 33 mM glucose every

12 hours. Apoptosis was assessed by cleaved caspase 3 and TUNEL assay

Hoechst staining. The expression of Mn-SOD and Bcl-2 was measured by

Western blotting and insulin secretory capacity was assessed. In

comparison to the control group, INS-1 cells uncovered to chronic

hyperglycemia and intermittent hyperglycemia showed an increase in

apoptosis. The apoptosis of INS-1 cells uncovered to irregular

hyperglycemia increased considerably more than the apoptosis of INS-1

cells exposed to chronic hyperglycemia. In comparison to the control

group, the insulin secretory capacity in the two hyperglycemic states was

decreased, and more with intermittent hyperglycemia than with chronic

hyperglycemia. The expression of Mn-SOD and Bcl-2 increased more with

chronic hyperglycemia and less with intermittent hyperglycemia.

20

Intermittent hyperglycemia induced a higher degree of apoptosis and

decreased the insulin secretory capacity more in pancreatic beta cells than

chronic hyperglycemia. This activity might be mediated by the anti-

oxidative enzyme Mn-SOD and the anti-apoptotic signal Bcl-2 (Kim et al.,

2010).

Lin et al. (2010) reported that apoptosis might be closely involved

in diabetes-induced embryonic malformations. The occurrence of

apoptosis at an early stage of development, in oocytes and 2-cell embryos

of streptozotocin-induced diabetic mice and nondiabetic mice was

investigated. Reduced number of growing follicles and delayed oocyte

development were seen when diabetic mouse ovarian sections were

stained with hematoxylin and eosin. Annexin V-positive oocytes were

higher in number in diabetic mice than in nondiabetic mice. Quantitative

RT-PCR and immunofluorescence analysis revealed the expression of Bax

and caspase-3 considerably lesser in nondiabetic oocytes and higher in

diabetic oocytes. Annexin V-positive staining was not seen in 2-cell

embryos of diabetic and nondiabetic mice. Bax expression was elevated in

diabetic 2-cell embryos, but caspase-3 expression did not considerably

differ between diabetic and nondiabetic 2-cell embryos. The results

suggested that maternal diabetes might increase oocyte apoptosis by a

Bax-caspase-3 pathway to play a role in embryonic malformations by late

oocyte development. Development of 2-cell embryos might be favorably

affected through Bax-regulated caspase-3 apoptotic pathway.

Sensitizing radioresistant tumours by combining irradiation with

other therapeutics to induce apoptosis have been widely investigated. It

21

was examined whether chicken anemia virus-derived apoptin protein

would have a useful effect on irradiation of radiosensitive SCC61 and

radioresistant SQD9 human head and neck squamous carcinoma cell

lines. In both cell lines, simultaneous exposure to irradiation and apoptin

resulted in mitochondrial cytochrome c discharge and in cleavage of

caspase-3, whereas irradiation alone of SQD9 cells under identical

conditions did not. Only the synchronized treatment of apoptin and

irradiation resulted in increased cell death in comparison with the

irradiation, especially in the radioresistant SQD9 cells, as measured

through colony survival assay. The data was found to show that apoptin

treatment represented an effective way for enhancing radiotherapy of

tumors responding badly to radiotherapy (Schoop et al., 2010).

Silicosis was a chronic lung disease which was characterized by

granulomatous and fibrotic lesions, which occured due to accumulation of

respirable silica mineral particles. Caspase activation played a central role

in the execution of apoptosis. Silica-induced apoptosis of the alveolar

macrophages could potentially favor a proinflammatory state, which took

place in the lungs of silicotic patients. This result in the activation of

caspase prior to induction of the intrinsic and extrinsic apoptosis

pathways. In recent studies, it was indicated that apoptosis might involve

in pulmonary disorders. In addition, caspase could be a key apoptotic

protein that can be used as an effective biomarker for the study of

occupational diseases. It might provide an important link in

understanding the molecular mechanisms of silica-induced lung

pathogenesis (Tumane et al., 2010).

22

Semiconductor nanoparticles conjugated to photosensitizers have

been shown to increase tumor cell death with ionizing radiation. In an

in vitro system, a molecular probe was used to quantify the component of

photodynamic cell-killing. The intracellular distribution of the

nanoparticle conjugate (NC) was determined by the co-localization of

nanoparticles and the lysotracker. TUNEL assay and western blot analysis

were used to measure the induction of apoptosis of the cleaved caspase-3.

As a result, dose-dependent production was observed with 48 nm

nanoparticle conjugate after irradiating with 6 MV x-rays. A high

geometrical coincidence between the fluorescence emission of the

nanoparticle and lysotracker was observed using confocal microscopy.

Apoptosis, as indicated by the cleavage of caspase-3 and TUNEL stain,

was observed in cells treated by both the nanoparticle conjugate and 6 Gy

of radiation but not in cells treated with radiation alone. The cell death

induced by the nanoparticle conjugate in combination with radiation was

consistent with a supra-additive effect to radiation or nanoparticle

conjugate alone-killing. It was mediated by a nanoparticle conjugate-

induced photodynamic therapy mechanism, which was distinctly

different from that for radiation-killing alone. By providing a second

distinct cell-killing mechanism, this nanoparticle conjugate has great

promise as a targeted physical radiosensitizer aimed at overcoming

radioresistant tumor clonogens or/and reducing normal tissue toxicity by

using a lesser ionizing radiation dose (Wang et al., 2010).

Pancratistatin was a natural compound extracted from

Hymenocallis littoralis, which could selectively induced apoptosis by

various pathways. It showed marked effectiveness on cancer cells.

23

Apoptosis was one of the mechanisms, which detached the cells that were

infected with pathogens or with abnormal cell cycle (Patel and Prajapati,

2011).

2.3 Apoptin

Acetylation of p53 was indispensable for its transcriptional activities

and induction of apoptosis upon DNA damage. Chromatin remodelling

protein SMAR1 inhibited p53 acetylation and p53 dependent apoptosis by

repressing p300 expression in reaction to DNA damage. The repression of

p300 expression by SMAR1 was relieved upon treatment with

proteosomal inhibitors MG132 and Lactacystin. SMAR1 interacted with

p53-p300 transcriptional complex and SMAR1 overexpression

antagonized p300 interaction with p53 and suppressed activation of p53

apoptotic targets and p53 regulated miRNA miR-34a. On the contrary,

knockdown of SMAR1 promoted p300 accumulation and p53 acetylation

while ectopic expression of p300 rescued SMAR1 inhibition on p53.

Collectively, these results indicated that SMAR1 was an important player

in p300-p53 regulated DNA damage signalling pathway and could exert

its effect on apoptosis in a transcription independent manner (Sinha et al.,

2012).

Several viral gene products affect apoptosis by interacting directly

with components of the highly conserved biochemical pathway which

regulated cell death. On the one hand it appeared that viruses blocked

apoptosis to prevent premature death of the host cell and so maximized

virus progeny from a lytic infection or facilitate a persistent infection. On

24

the other hand it appeared that a growing number of viruses actively

promoted apoptosis; viruse may perform both functions, the latter being

the culmination of a lytic infection and serving to spread virus progeny

to neighbouring cells while evading host inflammatory responses.

Apoptosis may then contribute to the cytotoxicity associated with virus

infections (O'Brien, 1998). Apoptin also induced apoptosis in human

osteosarcoma cells, despite of whether they expressed wild-type, mutant

p53, or not even p53. Furthermore, the nuclear position of apoptin

appeared to be important for its most favorable induction of apoptosis.

The reality that apoptin can induce p53-independent apoptosis in human

tumor cells made apoptin a potential candidate for treatment of

frequently occurring types of cancer cells that do not contain functional

p53 (Zhuang et al., 1995).

Pietersen et al. (1999) explained the generation and characterization

of an adenovirus vector, AdMLPvp3, for the expression of apoptin.

Infection with AdMLPvp3 of normal rat hepatocytes in cell culture did

not increase the frequency of apoptosis. On the contrary, in the hepatoma

cell lines Hep3b and HepG2, infection with AdMLPvp3, but not with

control vectors, led to a rapid induction of programmed cell death.

Experiments in rats have shown that AdMLPvp3 could be safely

administered by subcutaneous, intraperitoneal or intravenous injection.

Repetitive intravenous doses of AdMLPvp3 were also well tolerated,

showing that the apoptin-expressing virus can be administered without

severe adverse effects. In an introductory experiment, a single

intratumoral injection of AdMLPvp3 into a xenogeneic tumor (HepG2

cells in Balb/Cnu/nu mice) resulted in a significant reduction of tumor

25

growth. Taken together, the data demonstrated that adenovirus vectors

for the expression of the apoptin gene might constitute a powerful tool for

the treatment of solid tumors.

The presence of caspases was an important component of the

apoptotic machinery present in mammalian cells. By means of a specific

antibody, active caspase-3 was observed in cells expressing apoptin and

undergoing apoptosis. Although apoptin activity was not impacted by

CrmA, p35 inhibited apoptin-induced apoptosis, as observed by nuclear

morphology. Cells expressing apoptin and p35 displayed only a slight

change in nuclear morphology. Though, in most of these cells,

cytochrome c was usually released and the mitochondria were not stained

by CMX-Ros, showing a drop in mitochondrial membrane potential.

These observations implied that although the final apoptotic events were

blocked by p35, apoptin activated parts of the upstream apoptotic

pathway that affected mitochondria. Considering together, these data

illustrated that the viral protein apoptin employed cellular apoptotic

factors for induction of apoptosis. Although initiation of upstream

caspases was not required, initiation of caspase-3 and possibly also other

downstream caspases were vital for rapid apoptin-induced apoptosis

(Danen-van Oorschot et al., 2000). Apoptosis was a very common and

the best understood mechanisms of physiological cell death. It results

from the activation, through any of two primary pathways, of site-specific

proteases called caspases. There were many other routes to cell death,

prominently like autophagy and proteasomal degradation of critical

constituents of cells. Most commonly, autophagic or proteasomal

degradation was used to purge massive cytoplasm of very large cells,

26

particularly post-mitotic cells, and these pathways were prominent even

though caspase genes, messages, and pro-enzymes were found in the

cells. These types of cell death were entirely physiological and not just a

default pathway for a defective cell and they were distinct from necrosis

(Lockshin and Zakeri, 2002).

In order to study the antitumor effect of VP3 protein, particularly

its effect against liver carcinoma in vivo, control vector pcDNA3 and

recombinants pcDNA-vp3 having chicken anemia virus vp3 gene were

mixed with murine liver carcinoma cell lines H22 respectively. The

mixture was inserted subcutaneously into Balb/C mice. After some days,

the mice died and the solid tumor weighed. The antitumor efficiency was

examined. The manners of vp3 protein in vivo inducing tumor cell death

were recognised by using TUNEL assay. All the consequences suggested

that the injection of pcDNA-vp3 and H22 mixture resulted in a major

reduction of tumor growth in mice in comparison with the results of

control groups. TUNEL assay showed that vp3 induced apoptosis in vivo.

All these reflected that cav vp3 might be a promising new gene in

reducing the growth rate of tumor cells in liver carcinoma or in other kind

of solid tumors in vivo (Shen et al., 2003).

Through the use of PCR technique, the vp3 gene of CAV was

cloned into the eukaryotic expression vector pcDNA3 to make a

recombinant pcDNA-vp3. Restriction enzyme digestion and sequencing

analysis showed that cav vp3 gene was correctly inserted into the blank

vector pcDNA3. After LipofectAMINE-mediated transfection in vitro with

pcDNA-vp3 and pcDNA3 respectively, the total mRNA was taken out

27

from liver carcinoma cell lines HepG2 and diploid cell line L-02, and RT-

PCR was conducted afterward. The consequences of RT-PCR showed that

vp3 gene was expressed in these two cell lines. At the same time, using in

situ apoptotic finding assay, TUNEL kits, the apoptotic cells were

observed in pcDNA-vp3 transfected HepG2, but not in mock transfected

cell lines. VP3 could cause cell death by apoptosis in cancer cell lines, but

not in diploid cell lines. All the results showed that cav vp3 gene, a

potential therapeutic agent, has the potential to be used for cancer

treatment (Sun et al., 2003).

Apoptin formed distinct, stable multimeric complex which

was remarkably homogeneous and uniform, when it was produced in

bacteria as a recombinant fusion with maltose-binding protein (MBP-

Apoptin). By the use of cytoplasmic microinjection, recombinant MBP-

Apoptin multimers retained the characteristics of the ectopically expressed

wild-type apoptin; viz., the complexes translocated to the nucleus of

tumor cells and induced apoptosis, though they remained in the

cytoplasm of normal, primary cells and exerted no noticeable toxic effect.

In normal cells the MBP-Apoptin formed gradually large, organelle-sized

globular bodies with time postinjection and ultimately lost the ability to be

detected by immunofluorescence analysis. Costaining with an

acidotrophic marker showed that these globular bodies did not

correspond to lysosomes. Immunoprecipitation studies reflected that

MBP-Apoptin remained entirely antibody-accessible despite of buffer

stringency when microinjected into tumor cells. On the contrary, in

normal cells, MBP-Apoptin was only recoverable under stringent lysis

conditions, while under milder conditions they became entirely shielded

28

with time on two epitopes spanning the entire protein. Moreover

biochemical analysis found that the long-term fate of apoptin protein

aggregated in normal cells was their eventual removal. The outcome

showed that the tumor-specific apoptosis-inducing aggregate was

basically sequestered by factors or conditions present in the cytoplasm of

healthy, nontransformed cells. This characteristic should make known

more about the cellular interactions of this viral protein as well as further

improve its safety as a potential tumor-specific therapeutic agent (Zhang

et al., 2003).

A series of biomedical studies on apoptin have been carried out in

human cell systems, which were revealing about the mechanism of CAV-

induced apoptosis in chicken (transformed) cells. Apoptin contained a

bipartite nuclear localization signal, and one domain that resembled a

nuclear export signal. Elucidation of parts of the apoptin-induced

apoptotic pathway exposed unique characteristics: apoptin-induced

apoptosis was independent of the tumor suppressor p53. The anti-

apoptotic protein Bcl-2 did not inhibit but even accelerated apoptin-

induced apoptosis in tumor cells, while over expression of Bcl-2 in normal

cells has no effect on the apoptin activity. Numerous novel proteins were

shown to interact with apoptin in transformed cells (Noteborn, 2004). It

appeared to have innate tumour-specific p53-independent that Bcl-2-

enhanced pro-apoptotic activity, and therefore might be of great value in

the endeavor to achieve specific and efficient removal of cancer cells,

particularly in cases of drug resistance through Bcl-2 overexpression or

loss of p53 function etc. Apoptin has ability to localise specifically in the

nucleus of transformed but not normal cells. The latter ability,

29

importantly, appeared to be integrally related to its tumour-specific pro-

apoptotic action (Oro and Jans, 2004).

Apoptin was generated and cloned into several mammalian

expression vectors. Microinjection or transfection of apoptin cDNA

resulted in its expression, in the cytoplasm with a filamentous pattern.

apoptin entered the nucleus and efficiently induced apoptosis in several

cancer cell lines. For induction of apoptosis, nuclear localization was

shown to be required. Apoptin expression level was found to be an

important determinant of the efficiency of induction of apoptosis.

Expression of apopti or GFP-apoptin cDNA induced apoptosis in some

normal cells. When fused to the HIV-TAT protein transduction domain

and delivered as a protein, TAT-apoptin was transduced efficiently

(>90%) into normal and tumor cells. However, TAT-apoptin remained in

the cytoplasm and did not kill normal 6689 and 1BR3 fibroblasts. In

contrast TAT-apoptin migrated from the cytoplasm to the nucleus of Saos-

2 and HSC-3 cancer cells resulting in apoptosis after 24 hrs. Apoptin was a

powerful apoptosis-inducing protein with a potential for cancer treatment

(Guelsen et al., 2004).

VP3's anti-cancer activity was strongly linked to its ability to

localize more efficiently in the nucleus of cancer and transformed cells

with a tumor cell-specific nuclear targeting signal located at the C-

terminus of the protein. The VP3 tumor cell-specific nuclear targeting

signal was an exciting prospect to improve non-viral-mediated cancer

cell killing. The understanding of the mechanism responsible for VP3

tumor-specific nuclear localization, including its specific

30

phosphorylation, and the implications for the improvement of anti-

cancer treatment have been discussed. It also proposed alternative

strategies to develop tumor cell-specific nuclear targeting signal for anti-

cancer therapies (Alvisi et al., 2006).

Li et al. (2006) constructed a recombinant fowlpox virus

expressing the apoptin protein (vFV-Apoptin) and compared the tumor-

killing activity of the recombinant virus with that of wild-type fowlpox

virus in the human hepatoma cell line HepG2. It was observed that

although cells were somewhat resistant to the basal cytotoxic effect of

wild-type fowlpox virus, infection with vFV-Apoptin caused a

pronounced, additional cytotoxic effect. Also, cell death and disruption of

tumor integrity were noticeable in the vFV-Apoptin-infected cells. They

also tested whether fowlpox virus-mediated expression of apoptin in

tumor cells could stimulate an antitumor effect by injecting aggressive

subcutaneous tumors derived from H22 mouse hepatoma cells in

C57BL/6 mice with vFV-Apoptin. They observed that fowlpox virus-

mediated intratumoral expression of the apoptin gene can induce

protective and therapeutic antitumor effects and significantly increase

survival. Taken together, these data indicated that infection of tumors

with fowlpox virus expressing apoptin inhibited tumor growth, induced

apoptosis and might be an effective cancer cure.

Neoplastic transformation was the consequences of viruses which

could produce viral oncoproteins that drive multiple genetic alterations.

Viral proteins encoded by onco-related viruses such as Epstein-Barr virus

or polyomavirus SV40 were involved in cellular processes, which resulted

31

in the imbalance between proliferation and cell death. Viruses also

generated viral components that could become a tumor-selective

destroyer by sensing the cellular tumorigenic hallmarks. Apoptin protein

which was derived from the avian virus has been proven to induce

tumor-regression in various pre-clinical animal models without showing

noticeable side effects. Studies were described with representative viral

elements that have contributed to the understanding of significant

tumorigenic processes and have conferred an impact on the growth of

novel anti-cancer therapies (Kooistra et al., 2007).

Lee et al. (2007) investigated the antitumor effect of CAV VP3 gene

in canine mammary tumor (CMT) cells and established primary canine

cell lines that arose from epithelial cells of resected CMTs and

nonneoplastic mammary gland epithelial (MGE) cells. Expression vectors

and lentiviral vectors encoding the VP3 gene from a Taiwan-Ilan isolate of

cav were used to deliver the VP3 gene into CMT cells and nonneoplastic

MGE cells. Ectopic gene expression and the pro-apoptotic effect of the

VP3 gene on CMT and nonneoplastic MGE cells by either viral infection

or transfection were evaluated via western blot analysis,

immunofluorescence microscopy and terminal deoxynucleotidyl

transferase-mediated dUTP nick-end labeling analysis. Overexpression of

the enhanced green fluorescent protein with VP3 fusion protein was

detected predominantly in the nuclei of CMT cells. In nonneoplastic MGE

cells, the VP3 protein was found to be localized in cytoplasm. Among the

fusion protein-expressing CMT cells, nearly all underwent characteristic

changes of apoptosis, whereas apoptosis was not detected in fusion

protein-expressing, nonneoplastic MGE cells. Induction of apoptosis by

32

VP3 gene overexpression in CMT cells was linked with the caspase-9

mediated apoptosis pathway. The data indicated that the VP3 gene of

CAV did not induce apoptosis in nonneoplastic canine MGE cells but it

induced apoptosis in malignant CMT cells. The VP3 gene might be a

promising agent on the basis of such tumor cell-specific killing, for the

treatment of malignant mammary gland tumors in dog.

Purification of cav vp3 protein which was expressed in a

prokaryotic expression system was established as histidine-tagged fusion

protein. DNA was extracted from the infected liver of chicken and CAV

particle was obtained. By polymerase chain reaction (PCR), the vp3 gene

was amplified from the extracted DNA. This extracted DNA was then

cloned. The recombinant expression construct (pTrc-VP3) was recognized

by PCR and sequencing analysis. The vp3 expressed protein with a

molecular mass of about 21 kDa was confirmed by Western blotting

analysis with cav-specific antibodies. The in vitro expressed VP3 protein

was purified to near homogeneity by elution from the gel, as elucidated

by sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis.

The purified VP3 protein was recognized by CAV antibodies in a Western

blotting assay. This finding indicated that recombinant VP3 expressed in

the pTrcHis2 vector system can be used as antigen to detect anti-cav

antibodies (Nogueira-Dantas et al., 2007).

Apoptin interacted directly with the promyelocytic leukemia protein

(PML) intumor cells. Apoptin was accumulated in PML nuclear bodies

(NBs), which were involved in apoptosis induction and viral replication.

Apoptin was sumoylated and that a sumoylation-deficient

33

apoptin mutant was no longer recruited to PML-NBs, but localized in

the nuclear matrix. This mutant failed to attach PML, but could still

induce apoptosis as proficiently as wild-type apoptin.

Moreover, apoptin killed PML cells and promyelocytic leukemia cells

with defective PML expression. The results recommended that apoptin

killed tumor cells independently of PML and sumoylation (Janssen et al.,

2007).

In immunocompetent mice, the effects of apoptin expression in

primary oral tumors were induced by the carcinogen 4-Nitroquinoline-1-

oxide. In vivo, a significant amount of primary oral tumor cells

expressing apoptin cells underwent apoptosis, while synthesis of the

LacZ control product did not. Ectopical expression of apoptin in passage

1 cell cultures derived from these oral tumors also resulted in apoptin-

induced. Both in-vitro and in-vivo treated cells underwent apoptosis via

the activation of caspase-3. Apoptin proved to be a potential therapeutic

agent for the cure of head and neck squamous cell carcinoma from the

fact that apoptin induced apoptosis in primary squamous cell carcinoma

cells (Schoop et al., 2008).

Apoptin phosphorylation, nuclear translocation, and apoptosis

colud transiently be induced in normal cells by cotransfecting SV40 large

T oncogene, indicating that apoptin recognizes early stages of oncogenic

transformation. Apoptin appeared to recognize survival signals in cancer

cells which it was able to redirect into cell death impulses. Apoptin targets

included DEDAF, Nur77, Nmi, Hippi, and the potential drug target

APC1. Apoptin-transgenic mice and animal tumor models have revealed

34

apoptin as a safe and proficient antitumor agent, resulting in signifacant

tumor regression. Future antitumor therapies could use apoptin either as

as an early sensor of druggable tumor-specific processes or as a curative

bullet (Backendorf et al., 2008).

It was studied that multiple genes regulated the initiation and

progression of tumor. Survivin belonged to the inhibitor of apoptosis

protein (IAP) family and was overexpressed in most types of human

tumors. The expression of surviving was silenced by microRNA

(miRNA)-mediated RNA interference (RNAi) and the engineered miRNA

vector was also designed to express apoptin gene. The apoptosis and cell

growth were then examined by MTT assay and flow cytometry. The

miRNA-mediated knockdown of survivin in combination with apoptin

overexpression significantly induced apoptosis and inhibited cell growth.

The combined strategy was more effective on inducing apoptosis and

inhibiting cell growth than either survivin downregulation or apoptin

overexpression only. Taken together, the combined strategy offered

potential advantages in control of tumorigenesis. The combined stategy

offer great advantages in control of tumorogenesis and so deserve further

research as a preferred approach in cancer gene therapy (Liu et al., 2008).

In tumor cells, apoptin caused the nuclear accumulation of

survival kinases including Akt and was phosphorylated by CDK2.

Apoptin redirected survival signals into cell death responses. Apoptin

was also find to bind as a multimeric complex to DNA and interacted

with several nuclear targets, such as the anaphase-promoting complex,

resulting in a G2/M phase arrest. The proapoptotic signal of apoptin was

35

then transduced from the nucleus to cytoplasm by Nur77, which

triggered a p53-independent mitochondrial death pathway. Discoveries

of apoptin's mechanism of action that might provide intriguing insights

for the development of novel tumor-selective anticancer drugs were

described (Los et al., 2009).

The abundantly expressed carcinoembryonic antigen (CEA) on

several cancer types was an attractive target for antibody-directed

therapy. CEA was also present in some normal tissues. It was observed

that a dual functioning protein, designated as CAtin exhibited both

specific binding and killing functions, by fusing a tumor-specific

apoptosis-inducing molecular apoptin to C-terminus of an anti-CEA

single-chain disulfide-stabilized Fv antibody (Yan et al., 2010).

The recombinant lentivirus containing the fused gene of SP-TAT-

Apoptin was packaged to infect HepG2 cell and the efficiency of

apoptosis was measured. The eukaryotic expression vector of SP-TAT-

Apoptin fused gene and other packaging plasmids were transfected into

293FT cells by Lipofectamine(TM);2000 reagent. The supernatant of the

cultured 293FT cells was harvested and real time PCR determined the

virus titration. The expression of the fused gene of SP-TAT-Apoptin in

293FT cells infected by the recombinant lentivirus was examined by

immunofluorescence histochemistry method. Flow cytometer was used to

determine the apoptosis rates of the HepG2 cells infected by the

recombinant lentivirus. The 293FT cells infected by the recombinant

lentivirus could express the fused protein SP-TAT-Apoptin. Annexin-V PI

assay showed that SP-TAT-Apoptin carried by the recombinant lentivirus

36

could cause the HepG2 cell apoptosis, and its apoptosis rate was

considerably more than paired control group and SP-TAT-Apoptin

carried by liposomes only. The recombinant lentivirus of SP-TAT-Apoptin

was successfully packaged and it could induce HepG2 cells to apoptosis

(Han et al., 2010).

A lentiviral vector was developed which encoded a green

fluorescent protein-apoptin fusion gene (LV-GFP-AP). It could efficiently

deliver apoptin into hematopoietic cells. Apoptin selectively killed the

human multiple myeloma cell lines MM1.R and MM1.S, and the leukemia

cell lines K562, HL60, U937, KG1, and NB4. The normal CD34(+) cells

were not killed and maintained their differentiation potential in

multilineage colony formation assays. Dexamethasone-resistant MM1.R

cells were found to be more susceptible to apoptin-induced cell death

than the parental matched MM1.S cells. Death susceptibility was

connected with increased phosphorylation and activation of the apoptin

protein in MM1.R cells. Expression array profiling identified differential

kinase profiles between MM1.R and MM1.S cells. Among these kinases,

protein kinase Cβ (PKCβ) was found by immunoprecipitation and in vitro

kinase studies to be a candidate kinase responsible for apoptin

phosphorylation. Apoptin-mediated cell death proceeded during the

upregulation of PKCβ, cleavage of the PKCδ catalytic domain, activation

of caspase-9/3 and downregulation of the MERTK and AKT kinases. The

results elucidated a novel pathway for apoptin activation involving PKCβ

and PKCδ. It was highlighted that the potential of apoptin and its cellular

regulators was used to purge bone marrow in autologous transplantation

for multiple myeloma (Jiang et al., 2010).

37

In mice, therapeutic effect of adenovirus-mediated apoptin gene

transfer combined with ADM and CDDP on hepatocellular carcinoma

was studied. In c57BL/ 6 mice bearing hepatocellular carcinoma, the

changes of tumor volume, tumor inhibition rate, histomorphology and the

side effects were observed after intratumoral injection of adenovirus

containing apoptin gene, ADM and CDDP. After the treatment of seven

days, the mean volume of the tumor in the mice receiving intratumoral

apoptin-containing adenovirus injection combined with ADM and CDDP

reduced considerably as compared with that in mice treated with

adenovirus vehicle and control group. The tumor inhibition rate in the

combined treatment group was 90.13%, significantly higher than that in

the control group. No adverse effect of the treatment was observed in the

course of the experiment. The adenovirus vectors containing apoptin gene

combined with ADM and CDDP might serve a safe treatment of

hepatocellular carcinoma (Liu et al., 2010).

Li et al. (2010) generated a conditional replication-competent

adenovirus (CRCA), designated as Ad-hTERT-E1a-Apoptin, and

investigated the effectiveness of the CRCA, a gene therapy agent for the

clinical trials. The observation that infection with Ad-hTERT-E1a-Apoptin

significantly supperessed the growth of melanoma cells, protecting

normal human epidermal melanocytes from growth inhibition confirmed

cancer cell selective adenoviral replication, apoptosis induction and

growth inhibition of this therapeutic approach. The ability of the

recombinant adenoviruses was evaluated to prolong the survival of the

tumor bearing mice. When treated with Ad-hTERT-E1a-Apoptin, the

subcutaneous primary tumor volume reduction was observed in

38

intratumoral injection group as well as in systemic delivery mice.

Pulmonary metastatic lesions were effectively suppressed by Ad-hTERT-

E1a-Apoptin, in the lung metastasis model. Mice survival was increased

by the treatment of primary and metastatic models with Ad-hTERT-E1a-

Apoptin. The study showed the engineered CRCA induced apoptosis

selectively in various cancer cells without adverse effects on normal cells.

It was also observed that an adenovirus expressing apoptin was more

effective and advocated the potential applications of Ad-hTERT-E1a-

Apoptin in the treatment of neoplastic diseases in upcoming clinical trials.

It was revealed from structural modeling of pDED-HIPPI that R393

of HIPPI was important for such interaction. R393E mutation in pDED-

HIPPI decreased the interaction with the putative promoter of caspase-1

in cells. Expression of caspase-1 was decreased in cells expressing mutant

pDED-HIPPI in comparison to that observed in cells expressing wild

type pDED-HIPPI. Using HIP-1 knocked down cells as well as over

expressing HIP-1 with mutation at its nuclear localization signal and

other deletion mutations, translocation of HIPPI to the nucleus was

mediated by HIP-1 for the increased expression of caspase-1. HIPPI-HIP-

1 heterodimer was detected in cytoplasm as well as in the nucleus and

was associated with transcription complex in cells. Taking together, the

importance of R393 of HIPPI and the role of HIPPI-HIP-1 heterodimer in

the transcription regulation of caspase-1 were shown (Banerjee et al.,

2010).

The development of molecular technologies has allowed the

elucidation of the multiple mutations and dysregulatory effects of

39

pathways leading to tumorogenesis. Specifically targeted agents

represented a major challenge for current research efforts in oncology to

act against these pathways. As conventional anatomically based

pharmacological endpoints might be inappropriate to monitor the tumor

response to these targeted treatments, the recognition and use of more

appropriate, noninvasive pharmacodynamic biomarkers appeared to be

crucial to optimize the design, dosage and schedule of these novel

therapeutic approaches. An aberrant choline phospholipid metabolism

and enhanced flux of glucose derivatives via glycolysis, which sustained

the redirection of mitochondrial ATP to glucose phosphorylation, were

two main hallmarks of cancer cells. The perspectives of present efforts

addressed to recognize enzymes of the phosphatidylcholine cycle as

possible novel targets for anticancer treatment were summarized (Podo

et al., 2011).

Using yeast two-hybrid and immunoprecipitation approaches, it

was revealed that apoptin interacted with heat shock cognate protein 70

(Hsc70). Apoptin induced the translocation of endogenous Hsc70 in vivo,

from the cytoplasm to the nucleus, and both were co-localized in the

nucleus. In addition, apoptin induced Akt phosphorylation, which was

markedly repressed by Hsc70 knockdown, signifying that Hsc70 might

play a critical role in Apoptin-induced Akt phosphorylation. These

observations helped further to understand the molecular mechanism of

apoptin (Chen et al., 2011).

Recent understanding of the molecular pathogenesis of cancer has

led to the development of targeted novel drugs such as small molecule

40

inhibitors, monoclonal antibodies, mimetics, antisense and small

interference RNA-based strategies, among others. These compounds

acted on specific targets that were believed to contribute to the

development and progression of cancers. Whether delivered individually

or combined with chemo- and/or radiotherapy, such novel drugs have

produced significant responses in certain types of cancer. Among the

most successful novel compounds were those which targeted tyrosine

kinase. However, these compounds could cause severe side-effects as

they inhibited pathways such as epidermal growth factor receptor

(EGFR) or platelet-derived growth factor receptor, which were also

important for normal functions in non-transformed cells. The toxicity by

various proteins was independent of tumor type or specific genetic

changes such as p53, pRB or EGFR aberrations. It showed a current

overview of a selection of proteins with preferential toxicity among

cancer cells and which provide an insight into the

possible mechanism of action, tumor specificity and their potential as

novel tumor-specific cancer therapeutics (Argiris et al., 2011).

Resistance to radiation therapy was one of the hallmarks of

hepatocellular carcinoma. To induce apoptosis, sensitizing radioresistant

cancer by combining radiation with other therapeutics was widely

investigated. The earlier study showed that chicken anemia virus-derived

apoptin protein induced the apoptosis of hepatic carcinoma HepG. It was

observed that apoptin sensitized the cells to radiation-induced apoptosis

using a lentivirus-apoptin expression system in hepatic carcinoma HepG2

cells. The combination of apoptin and radiation up-regulated the p53

expression. Thus, apoptin treatment represented a potential method for

41

enhancing the effectiveness of radiotherapy in badly responding

hepatocellular carcinoma (Han et al., 2011).

Apart from viral approaches, some nonviral approaches have also

been used widely for intracellular gene transfer and gene therapy. A

modified wheat histone H4 protein was able to form stable complexes

with plasmid DNA which increased gene delivery efficiency. H4TL has

been used to deliver apoptin gene into a human ovarian carcinoma cell

line HO8910. Increased expression of apoptin at both mRNA and protein

levels was detected after transfection in HO8910 cells, accompanied by

reduced rate of growth of HO8910 cells in vitro and the loss of

mitochondrial membrane potential in these cells. These data showed that

H4TL-mediated transfer of apoptin initiated mitochondrial death

pathway in ovarian cancer cells and suggested a novel therapeutic

strategy for cancer (Wang and Zhang, 2011).

CAV protein apoptin has the intriguing activity of inducing G(2)/M

arrest and apoptosis specifically in cancer cells by a mechanism that was

independent of p53. At the level of localization, the activity of apoptin

was regulated. The study of apoptin was therefore a unique system for

identifying pathways of apoptosis that were able to kill cancer cells

independent of p53. DNA damage response (DDR) signaling was

required to induce apoptin nuclear localization in primary cells.

Induction of DNA damage in combination with apoptin expression was

able to induce apoptosis in primary cells. Chemical or RNA interference

(RNAi) inhibition of DDR signaling by ATM and DNA-dependent

protein kinase (DNA-PK) was sufficient to cause apoptin to localize in

42

the cytoplasm of transformed cells. The nucleocytoplasmic shuttling

activity of apoptin was required for DDR-induced changes in

localization. The nuclear localization of apoptin in primary cells was able

to inhibit the formation of DNA damage foci containing 53BP1. Apoptin

has been shown to bind and inhibited the anaphase-promoting

complex/cyclosome (APC/C). It was observed that apoptin was able to

inhibit formation of DNA damage foci by targeting the APC/C-

associated factor MDC1 for degradation. These results might point to a

novel mechanism of DDR inhibition during viral infection (Kucharski et

al., 2011).

Panigrahi et al. (2012) calculated a 3D structure of apoptin and

through modeling and docking approaches, they showed its interaction

with Bcr-Abl oncoprotein and its downstream signaling components,

which confirmed some of the newly-found interactions by biochemical

methods. Bcr-Abl oncoprotein was unusually expressed in chronic

myelogenous leukaemia (CML). It had several distinct functional domains

in addition to the Abl kinase domain. The calculated structural

coordinates of apoptin was used to study its interaction with the 3D

structure of CML-associated oncoprotein Bcr-Abl. The SH3 and SH2

domains cooperatively played important roles in autoinhibiting its kinase

activity. Adapter molecules such as Grb2 and CrkL interacted with

proline-rich region and activated multiple Bcr-Abl downstream signaling

pathways that contributed to growth and survival. Therefore, the

oncogenic effect of Bcr-Abl could be inhibited by the interaction of small

molecules with these domains. For therapeutic applications, especially for

CML treatment, Bcr-Abl was an attractive target for rational drug design.

43

Some of the interactions that were first predicted in silico were also

biochemically validated. This structure-property relationship of apoptin

might help in unlocking its cancer-selective toxic properties.

The antibacterial peptide cecropin B mutant (ABPS1) gene has a

broad range of antiproliferative and antibacterial properties. To explore

drug combination in human tumor cells, apoptin and ABPS1 eukaryotic

expression vector pIRES2-EGFP-apoptin and pIRES2-EGFP-ABPS1 were

constructed and their expression effect individually and in combinations

were studied in HepG2 and A375 cells. ABPS1 (23.79% in HepG2 and

8.33% inA375 cells). Moreover, the co-expression of apoptin and ABPS1

showed higher apoptotic rates which were 27.66 and 10.33% in HepG2

and A375 cells respectively. The apoptotic rates obtained in HepG2 cells

treated with ABPS1 and apoptin together were closely similar, but, not in

A375 cells. The results of the study indicated that the combination of

ABPS1 and apoptin had synergistic effect in HepG2 and A375 cell lines

(Birame et al., 2012).

2.4 Apoptin from other viruses

Studies with apoptin and E4orf4 have shown that they induced

G2/M arrest by targeting and inhibiting the anaphase-promoting

complex/cyclosome (APC/C). These two viral proteins induced

apoptosis selectively in transformed cells in a p53-independent manner;

thus pathways affected by these proteins were of significant therapeutic

interest. Further study of the underlying mechanism of G2/M arrest and

subsequent apoptosis induced by viral APC/C inhibitors might shed light

on the mechanisms of current cancer therapies and provide the

44

foundation for developing novel therapeutic targets (Heilman et al.,

2005).

A putative Torque teno viruses (TTVs) open reading frame (ORF) in

TTV genotype 1, named TTV apoptosis inducing protein (TAIP), induced,

like apoptin, p53-independent apoptosis in various human

hepatocarcinoma cell lines to a similar level as apoptin. In comparison to

apoptin, TAIP action was less pronounced in several analyzed human

non-hepatocarcinoma-derived cell lines. Detailed sequence analysis has

revealed that the TAIP ORF was conserved within a limited group of the

heterogeneous TTV population. Its N-terminal half, N-TAIP, was well

conserved in a much broader set of TTV isolates. The similarities between

apoptin and TAIP, and their relevance for the development and treatment

of diseases were discussed (De Smith and Noteborn, 2009).

The correlation between Torque teno virus (TTV) and CAV was

examined. Each had a circular single-stranded (ss) DNA genome with

each one of its known open reading frames (ORF) on its antigenomic

strand. The structure was found to be distinct from those of circoviruses.

TTV and CAV had different genomic sizes which were found to be 3.8 kb

and 2.3 kb, respectively. It was seen that the spectrum of the TTV genome

was enormously diverse and the CAV genome was quite narrow. The

relative allocation of ORFs on each frame in these viruses mimiced each

other. Three or more messenger RNA (mRNAs) were generated by

transcription in both of them. In each virus, the structural protein with the

replicase domain was coded by frame 1, and a nonstructural protein with

a phosphatase domain was coded by frame 2. In transformed cells,

45

apoptosis was induced by a protein on frame 3 in each virus (Hino and

Prasetyo, 2009).

Pseudotype baculovirus was developed as an alternative gene

delivery system. It was used as a vector to express apoptin. The resultant

recombinant baculovirus (BV-Apoptin) efficiently expressed the apoptin

protein and induced apoptosis in HepG2 and H22 cells. Studies in vivo

indicated that intratumoral injection of BV-Apoptin into a xenogeneic

tumor which was derived from H22 murine hepatoma cells in C57BL/6

mice, considerably suppressed tumor growth. It also significantly

prolonged the survival of tumor-bearing mice compared to a control

pseudotype baculovirus that expressed EGFP. The results signified that

apoptin which was expressed from the pseudotype baculovirus vector,

had the potential to become a therapeutic agent for the treatment of solid

tumors (Pan et al., 2010).

Sauvage et al. (2011) identified a new virus in a skin swab sample

from a healthy donor which was named human gyrovirus (HGyV)

because of its similarity to CAV, the only previously known member of

the Gyrovirus genus. This virus encoded a homolog of the CAV apoptin, a

protein that selectively induced apoptosis in cancer cells. The genome of

HGyV presented an overall organization similar to that of CAV. By PCR

screening, HGyV was found in 5 of 115 other nonlesional skin specimens

but in 0 of 92 bronchoalveolar lavages or nasopharyngeal aspirates and in

0 of 92 fecal samples. HGyV might also serve as a tool to decipher

pathways of apoptosis in cancer cells.

46

2.5 DNA Damage Signaling

Polycomb group (PcG) proteins were major determinants of cell

identity, epigenetic gene silencing and stem cell pluripotency during

development. The polycomb repressive complex 1, which contained

BMI1, RING1, and RING2, functions as an E3-ubuiquitin ligase. BMI1 and

RING2 were recruited to sites of DNA double-strand breaks (DSBs) where

they contributed to the ubiquitylation of γ-H2AX. In the absence of BMI1,

several proteins dependent on ubiquitin signaling, including 53BP1,

BRCA1, and RAP80, were impaired in recruitment to DSBs. Loss of BMI1

sensitized cells to ionizing radiation to the same extent as loss of RNF8.

An additive increase in radiation sensitivity was revealed by the

simultaneous depletion of both proteins. The data suggested a possible

role for BMI1 in the pathways of DNA damage response (Ismail et al.,

2010).

Using a synthetic lethal RNA interference screen in human cells,

Cyclin-dependent kinase 9 (CDK9) was identified as a component of the

replication stress response. An increase in spontaneous levels of DNA

damage signalling in replicating cells was caused by the loss of CDK9

activity. This activity was restricted to CDK9-cyclin K complexes and was

independent of CDK9-cyclin T complex. CDK9 accumulated on chromatin

in response to replication stress and limited the amount of single-stranded

DNA in cells under stress. Cyclin K and CDK9 interacted with ataxia

telangiectasia, Rad3-related protein and other checkpoint signalling

proteins. These results revealed an important role for CDK9-cyclin K in

checkpoint pathways that maintained genome integrity in response to

replication stress (Yu et al., 2010).

47

Tumor cells were found as more resistant to apoptosis that were

induced by cytotoxic agents, than normal cells. Resistance to apoptosis

induction could be a direct result of mutations in certain tumor-

suppressor genes (p53) or to certain proto-oncogenes (Bcl-2). Therefore,

new cancer therapies were under development to bypass the resistance to

chemo- and radio-therapy of tumors. Apoptin acted independently of

p53, was stimulated by Bcl-2 and was insensitive to BCR-ABL, which

means that apoptin can induce apoptosis in cases where (chemo)-

therapeutic agents fail. The fact that apoptin induced apoptosis in human

tumorigenic cells but not in normal diploid cells, implied that side-effects

of apoptin treatment were expected to be minor. In-vivo results with anti-

tumor therapy based on expression of apoptin indicated that apoptin had

low acute toxicity and was successful as an anti-tumor agent (Pietersen et

al., 2000).

From the above literature it is evident that apoptosis inducing potential of

apoptin can be advantageously utilized in antitumor therapy.

Materials and Methods

3.1 Materials

3.1.1 Vector

3.1.2 Gene

3.1.3 Primers

3.1.4 Host Bacterial Strains

3.1.5 Cell Culture

3.1.6 Conjugates

3.1.7 Hyper-immune Serum

3.1.8 Experimental Animals

3.1.9 Glass wares, Plastic wares and Chemicals

3.1.10 Solutions and Buffers

Chapter-3

3.2 Methods

3.2.1 Revival of the E. coli culture containing recombinant plasmid with VP3 gene

3.2.2 Isolation of plasmid DNA by TELT method

3.2.3 Agarose gel electrophoresis

3.2.4 Digestion of the recombinant plasmid with restriction enzyme to release the gene insert

3.2.5 DNA extraction from agarose gel

3.2.6 Blunting of EcoRI generated VP3 gene staggered ends

3.2.7 Purification of blunted insert

3.2.8 Creation of cloning site in vector

3.2.9 Dephosphorylation of 5’ ends

3.2.10 Purification of dephosphorylated linearized vector

3.2.11 Blunt end ligation of pSin vector and VP3 gene

3.2.12 Transformation of E. coli (DH5α) cells with the ligated product

3.2.13 Screening of recombinant clones for VP3 gene in right orientation

3.2.14 Transfection of HeLa cells

3.2.15 Immunoperoxidase test for DNA expression

3.2.16 Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE)

3.2.17 DNA fragmentation assay by agarose gel electrophoresis

3.2.18 Caspase 3 detection assay

3.2.19 Annexin V binding assay

48

Chapter-3

Material & Methods

3.1 Materials

3.1.1 Vector

The pSin vector (Fig. 1) is derived from an alpha virus (Sindbis

virus). The subgenomic promoter of alpha virus is very strong so that it

makes large number of target mRNA from the sequence downstream to

it. In comparative studies of conventional (nonreplicating) plasmid DNA

vectors and alphavirus DNA-based replicon vectors, the latter generally

produce larger quantity of DNA concentrations than do conventional

vectors.

3.1.2 Gene

The VP3 gene of CAV used in this study was cloned in pTarget

vector in Biotechnology laboratory of IBIT, Bareilly.

3.1.3 Primers

Primers used in the study were synthesized from Chromous

Biotech. (India).

49

pSin Vector

10.8 kb

BssHII SphI StuI ApaI

BGH Poly A

Replicase gene

Fig. 1: pSin Vector

50

Table: 1 Primers with their sequences

S. No. Primer Primer sequence

1. VP3 F 5’ATGAACGCTCTCCAAGAAG3’

2. BGH R 5’TAGAAGGCACAGTCGAGG 3’

3.1.4. Host Bacterial Strains

Escherichia coli (E. coli) DH5α (Proteges, Madison) host strain was

used for transformation with recombinant plasmids.

3.1.5. Cell Culture

HeLa cell lines were obtained from National Centre for Cell

Science (NCCS), Pune. These cell lines were used in the study of

apoptotic activity for recombinant plasmid and was maintained in

DMEM (Gibco, NY) supplemented with 50 µg/ml gentamycin (Amresco,

USA) and 10% fetal calf serum (Hyclone, USA).

3.1.6. Conjugates

Rabbit anti-mouse HRP conjugated antibody was obtained from

Bangalore Genei, Bangalore, (India).

3.1.7. Hyper-immune serum

Anti cav.vp3 hyperimmune sera were raised in mice using

recombinant plasmid by hyper immunization of six mice with

pSin.cav.vp3 plasmid. Fifty µg of plasmid DNA was injected

intramuscularly in lateral region of thigh muscles of each mouse and

51

repeated every week for four weeks consecutively. One week after last

injection, mice were bled through inner canthus of eyes with a capillary

and serum was prepared.

3.1.8. Experimental Animals

Swiss albino mice, 4 weeks old, were procured and kept in animal

room for raising hyper immune sera.

3.1.9. Glasswares, Plasticwares and Chemicals

All the glasswares used were Borosil make. Plasticwares were

from Axygen, Greiner or Eppendorf. All the chemicals used in the study

were procured from Fermentas (UK), Promega, Bangalore Genei

(Bangalore, India), Medox or Qiagen (Germany).

3.1.10. Solutions and buffers

The composition of chemical solutions and buffers used in the study

is given in appendix.

3.2 Methods

3.2.1. Revival of the E. coli culture containing recombinant

plasmid with VP3 gene

The E. coli containing pTarget.cav.vp3 was revived in LB broth

overnight at 370C following the protocol of Sambrook and Russell (2001).

The E. coli containing rplasmid was inoculated by streaking on LB agar

plate containing ampicillin at concentration of 100 µg/ml. The plate was

52

incubated at 370C for 24 hrs. A single colony was picked up from the

plate with the help of a loop and then transferred into 5 ml LB broth

containing 100 µg/ml ampicillin. It was kept in incubator at 370C for 24

hrs.

3.2.2. Isolation of plasmid DNA by TELT method (He et al., 1990).

A 1.5 ml E. coli culture was grown with plasmid of interest in LB

medium containing 100 µg/ml ampicillin for 16 hrs. Cells were pelleted

for 30 sec in microcentrifuge. They were resuspended in 100 µl TELT

solution and an equal volume of 1:1 phenol/chloroform was added. The

sample was vortexed vigorously for 15 sec and spin for 1 minute in

microcentrifuge at 220C. The upper phase of nucleic acid was collected

and mixed with 2 vol of 100% ethanol. After 2 minutes, spin 10 minutes

in microcentrifuge at 40C. The pellet was washed with 1ml of 70%

ethanol, allowed to dry and resuspended in 30 µl TE buffer.

3.2.3. Agarose gel electrophoresis

It was done as per method of Sambrook and Russell, (2001). 1% of

agarose was prepared in 1X TAE buffer and heated till it dissolved. The

solution was cooled to about 450C and then ethidium bromide was

added at a concentration of 0.5 µg/ml. Gel tray was sealed with tape on

its open ends and placed the comb on the tray. Then the above gel was

poured on the sealed tray. It was left to solidify. The comb was removed

and taped. The gel tray was put in electrophoretic tank containing 1X

TAE buffer. The DNA marker and sample DNA were loaded into the

53

slots of submerged gel, using bromophenol dye buffer and then the gel

was run at 50-100 Volts.

3.2.4. Digestion of the recombinant plasmid with restriction

enzyme.

Digestion was done to release the gene insert using protocol of

Sambrook & Russell (2001).

The microfuge tube was labeled to set 50 µl digestion reaction:

DNA (pTarget.cav.vp3)(200 µg/ml) 20 µl

(EcoRI) Enzyme (Promega 12 u/µl) 3 µl

Enzyme buffer (10X) 5 µl

Distilled Water 22 µl

The reaction mixture was mixed by tapping the tube. The tube was

centrifuged at 5000 rpm for 15 sec. Digestion mixture was incubated at

370C for 4 hrs or more in incubator. Agarose gel electrophoresis was run

in 1X TAE buffer using 1.5 % agarose to check restriction digestion.

3.2.5. DNA extraction from agarose gel

The gel extraction of DNA fragment was done using Min Elute gel

extraction kit (Qiagen, Germany) following manufacturers instructions.

Briefly, the agarose gel containing DNA fragment was excised with a

scalpel and weighed in a colorless microfuge tube. Three volumes

of buffer QGH were added to 1 volume of gel and incubated at 50˚C until

gel slice had completely dissolved. One gel volume of isopropanol was

54

added to the melted agarose and mixed by inverting several times. The

sample was applied to the Min Elute column fitted in a collection tube.

The column was centrifuged at 13000 rpm for 1 minute and the flow

through was discarded. 500 µl of QG was applied to the spin column,

centrifuged and the flow through was discarded again. The DNA bound

to the column was washed with 750 µl of buffer PE and the flow through

was discarded. The DNA was eluted from the column in 10 µl nuclease

free water.

3.2.6. Blunting of EcoRI generated VP3 gene staggered ends

For blunting of staggered ends generated by EcoRI enzyme, T4

DNA polymerase (Fermentas) was used. Reaction mixture was prepared

as follows:

Table: 2. Reaction mix for blunting

Component Volume(µl)

VP3 Insert 20

T4 DNA polymerase buffer 10X 5

T4 DNA polymerase (5u/µl) 2

dNTPs mix.(10mM each) 1

Nuclease free water 22

Total 50

The reaction mixture was incubated at 37˚C for 10 minutes.

55

3.2.7. Purification of blunted insert

The blunted VP3 insert was purified by phenol chloroform

precipitation following the protocol of Sambrook and Russell, (2001). In

brief, 50 µl of TE buffer was added to the reaction mixture followed by

addition of equal volume of phenol chloroform isoamyl alcohol (25: 24: 1)

i.e. 100 µl for above reaction mixture. It was centrifuged at 13000 rpm for

10 minutes. The aqueous layer was transformed into a different tube.

Again 50 µl of TE buffer was added to the tube containing phenol

chloroform isoamyl alcohol and centrifuged at 13000 rpm for 10 minutes.

Aqueous layer was transferred to the collection tube. 1/10 volume (5 µl)

of 3M sodium acetate was added to the aqueous collection. Then 2.5

times of 100% ethanol was added to it. The tube was kept at -20˚C

overnight. It was centrifuged at 13000 rpm for 20 minutes. Supernatant

was carefully discarded and pellet was washed in 1ml of 70% ethanol by

centrifuging at 13000 rpm for 10 minutes. The dried pellet was then

dissolved in 12 µl of nuclease free water. The presence of purified

blunted DNA was checked by running 1µl of DNA on 1% agarose gel.

3.2.8. Creation of cloning site in vector

StuI was chosen to create blunt end. A reaction mixture was

prepared with StuI enzyme (Promega 10u/ µl) as per the following

protocol to digest the pSin vector plasmid:

56

Table: 3. Reaction mix for StuI digestion

Component Volume (µl)

StuI enzyme (Promega 10u/ µl) 3

pSin vector DNA (100 µg/ ml) 15

NE Buffer (10X) 5

Nuclease free water 27

Total 50

The reaction mixture was incubated at 37˚C overnight. The lineraized

plasmid was checked on 1% agarose gel electrophoresis. The linearized

plasmid was then extracted from gel.

3.2.9. Dephosphorylation of 5’ ends

The calf intestinal alkaline phosphatase (CIAP) was used to remove

5’ phosphate group from both the ends of linearized plasmid. The

following reaction mixture was prepared in 50 µl volume.

Table: 4. Reaction mix for dephosphorylation

Component Volume(µl)

Linearized vector pSin (100 µg/ ml) 10

CIAP enzyme (Promega 1u/ µl) 1

Buffer (10X) 5

Nuclease free water 34

Total 50

The reaction mixture was incubated at 37˚C for 30 minutes.

57

3.2.10. Purification of dephosphorylated linearized vector

The method used was similar to that used in purification of

blunted insert.

3.2.11. Blunt end ligation of pSin vector and VP3 gene (Sambrook

and Russell, 2001).

A 10 µl reaction mixture was standardized with 30% PEG 8000 for

blunt end ligation. The components were mixed in the following

concentrations:

Table: 5. Ligation mixture

Component Volume (µl)

T4 DNA Ligase ((Promega 1u/ µl) 2

pSin vector (100 µg/ ml) 0.5

Cav.vp3 gene (150 µg/ µl) 4

Ligation buffer (10X) 1

30% PEG 8000 (Amresco) 1.5

Nuclease free water 1

Total 10

The reaction was incubated for overnight at 15˚C.

3.2.12. Transformation of E. coli (DH5α) cells with the ligated

product (Chung et al., 1989).

The single step method of competent cell preparation and

transformation was used.

58

A fresh overnight culture of bacteria was diluted into prewarmed LB

Broth and the cells were incubated at 37˚C in shaking incubator. The cells

were pelleted by centrifugation at 1000xg for 10 minutes at 4˚C,

supernatant removed and resuspended at 1/10th of original in ice cold 1X

TSS. The cell suspension was mixed gently. For transformation, a 0.1 ml

aliquot of cells was pipetted into a cold polypropylene tube containing 1

µl of plasmid DNA and the cell/ DNA suspension was mixed gently. The

cell/ DNA mixture was incubated for 5-7 minutes at 4˚C. A 0.9 ml

aliquot of TSS plus 20 mM glucose was added and the cells were

incubated at 37˚C in shaking incubator at 200 rpm for 1 hr. The above

transformed cells were plated on LB agar plates containing appropriate

antibiotic (ampicillin 50 µg/ ml) and the plates were incubated at 37˚C

for 16 hrs.

3.2.13. Screening of recombinant clones for VP3 gene in right

orientation

Few colonies were picked from the overnight grown

transformants. The individual colonies were inoculated in fresh LB broth

containing ampicillin (50 µg/ ml) and allowed to grow for 18 hrs.

Plasmid was isolated from these colonies by TELT method and checked

on 1.5% agarose gel electrophoresis.

59

3.2.13.1. Restriction digestion of plasmid with BglII enzyme to

check the presence and orientation of CAV-VP3 insert

Isolated plasmids were checked for the presence of insert by

digestion with enzyme BglII. The digestion mixture was prepared by

mixing the components in amounts as mentioned below:

Table: 6. Reaction mix for BglII digestion

Component Volume (µl)

Plasmid DNA (pSin.cav.vp3)( 200 µg/ml) 4

BglII. Enzyme (Fermentas 10 u/µl) 1

Buffer (10X) 1.5

Nuclease free water 8.5

Total 15

The reaction mixture was vortexed and spun; and incubated in a water

bath at 370C overnight. The digested mixture was then electrophoresed

in 1.5% agarose. The released fragments after digestion were compared

against 100 bp marker. One BglII site is present in the vector and one site

in the region of cav.vp3 gene. When the gene is in correct orientation,

three fragments were released viz. 5650 bp, 2180 bp and 3321 bp. The

restriction sites were predicted in pSin.cav.vp3 plasmid using DNASTAR

software.

3.2.13.2. Colony PCR

The presence of gene insert in right orientation in the recombinant

plasmid was confirmed by PCR using VP3 gene specific forward primer

60

and BGH as reverse primer. A part of single colony was picked up and

placed in a microfuge tube containing distilled water. The reaction was

carried as follows:

Table: 7. Reaction mix for PCR

Component Volume (µl)

Autoclaved distilled water 33

r colony (pSin.cav.vp3) 5

Forward primer of VP3 (50 pmole/µl) 2

Reverse primer BGH 2

dNTPs (10 mM) 2

Buffer (10X) 5

Taq DNA polymerase (3 units/ µl) 1

Total 50

The VP3 gene was amplified following initial denaturation at 940C for 5

minutes and 30 cycles of denaturation at 940C for 30 seconds, annealing

at 500C for 50 seconds, and amplification at 720C for 50 seconds and final

extension at 720C for 7 minutes. After amplification, an aliquot of 10

µl was subjected to agarose gel electrophoresis along with 100 bp DNA

molecular weight marker through 1.5% agarose gel at 60 volts for the

analysis of PCR product.

3.2.13.3. Sequencing of cloned chicken anemia virus VP3 gene

The recombinant plasmid selected after above two methods was

sent to Chromous Biotech Ltd. (Bangalore), for sequencing using BGH

reverse primer.

61

3.2.14. Transfection of HeLa cells (Sambrook and Russell 2001).

Cell culture was trypsinised using 0.17% trypsin versenate and

then 4 ml of DMEM containing 10% FCS was added. Then 100 µl of cell

culture suspension was placed in each of 96 well microtitre plate. The

calcium phosphate-DNA coprecipitate was prepared as follows: 50 µl of

2.5M CaCl2 was combined with 10 µg of plasmid DNA and 40 µl distilled

water in a sterile microfuge tube. The calcium phosphate-DNA

suspension was immediately transferred into the wells of 96 well

microtitre plate containing the cell suspension. Transfected cells were

incubated at 370C in a humidified chamber with an atmosphere of 5%

CO2. After 72 hrs of incubation, the cells were assayed for expression of

transfected gene.

3.2.15. Immunoperoxidase test for DNA expression (Gerna et

al., 1976).

The cells were rinsed in 96 well microtitre plate with PBS, pH 7.2.

The cells were fixed with 100 µl of chilled acetone at 40C for 10 minutes

and then air-dried. 10 µl of mouse antiapoptin hyperimmune serum was

added and incubated at 370C for 1 hr. Then they were washed with PBS.

10 µl of Rabbit anti- mouse globulin-HRP conjugate was added and

incubated at 370C for 1 hr. The wells were rinsed three times with PBS.

The preparation was air- dried. Nadi reagent was prepared just before

use as follows: 15 mg alpha naphthol and 22 mg phenylenediamine were

added in 20 ml of phosphate buffer. The solution was then filtered and

few drops of H2O2 were added. 100 µl of Nadi reagent was added in each

62

well. It was allowed to react for 5 minutes. The preparation was rinsed in

PBS. Then it was examined under microscope and photographed.

3.2.16. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis

(SDS-PAGE)

HeLa cells were transfected in 96 well plates with pSin.cav.vp3

rplasmid and pSin vector in separate wells. After 48 hrs of transfection,

cells were processed for SDS-PAGE following the method of Rodriguez

and Tait (1983). The lower resolving gel was 12% and the upper spacking

gel was 4%. The gel was run at 40 mA and stained with Coomassie

Brilliant Blue R250.

3.2.17 DNA fragmentation assay

HeLa cells showing 60% monolayer were transfected as described

earlier. After 48 hrs of transfection, the monolayer was trypsinized and

collected in a 1.5 ml tube. The cells were centrifuged at 3000 rpm for 5

min. After centrifugation, the media was removed and the cells were

resuspended in 200 µl PBS and the genomic DNA was isolated using

protocol of Sambrook and Russell (2001).

The cell suspension was transferred to a microfuge tube containing

600 µl of ice-cold cell lysis buffer. 3 µl of proteinase K solution was added

to the lysate to increase the yield of genomic DNA. It was incubated for

30 minutes at 370C. It was allowed to cool to room temperature and then

3 µl RNase was added. It was incubated for 30 minutes at 370C. The

sample was allowed to cool at room temperature. 200 µl of potassium

63

acetate solution was added and the contents of the tube were mixed by

vortexing vigorously for 20 seconds. The precipitated protein was

pelleted by centrifugation at 10000 rpm for 3 minutes. The supernatant

was transferred to a fresh microfuge tube containing 600 µl of

isopropanol. The solution was well mixed and then the tube was

centrifuged at 10000 rpm for 1 minute to recovere the precipitated DNA.

The supernatant was removed and 600 µl of 70% ethanol was added to

the DNA pellet. The tube was centrifuged at 10000 rpm for 1 minute. The

supernatant was removed carefully and the DNA pellet was allowed to

dry in air for 15 minutes. The DNA pellet was redissolved in 10 µl of

TE pH 7.6. The eluted DNA was stored at -20C.

The genomic DNA collected by above procedure was subjected to 2%

agarose gel electrophoresis. The bands were visualized under UV light

and photographed.

3.2.18. Caspase 3 detection assay

Activation of caspases plays an important role in apoptosis.

Caspase 3 was detected using CaspGLOW Fluoroscein Active caspase -3

Staining kit (Biovision, USA). The assay utilizes the caspase 3 inhibitor,

DEVD-FMK, conjugated to FITC (FITC-DEVD- FMK) as a marker. FITC-

DEVD- FMK is cell permeable, nontoxic and irreversibly binds to

activated caspase 3 in apoptotic cells. The FITC label allows for direct

detection of activated caspases in apoptotic cells by fluorescence

microscopy. The following protocol was followed.

64

Apoptosis was induced by transfecting HeLa cells with

pSin.cav.vp3 recombinant plasmid as described earlier. After 48 hrs of

transfection, the cells were trypsinized and collected in a 1.5 ml

microfuge tube. The cells were washed in cold PBS and then centrifuged

at 3000 rpm for 5 minutes. The cells were resuspended in 300 µl of PBS

and then 1 µl of FITC-DEVD- FMK was added into each tube and

incubated for 0.5 -1 hr at 370C incubator with 50% CO2. The cells were

centrifuged at 3000 rpm for 5 minutes and the supernatant was removed.

The cells were resuspended in 0.5 ml of wash buffer and centrifuged

again. This step was repeated once. Finally, the cells were resuspended in

100 µl of wash buffer, observed under a fluorescent microscope and

photographed.

3.2.19. Annexin V binding assay

This test was mainly done to detect plasma membrane alteration,

which occurs during apoptosis. In normal liver cells, phosphatidylserine

(PS) is located on cytoplasmic surface of cell membrane. However, in

apoptotic cells, PS is translocated from the inner to the outer leaflet of the

plasma membrane, thus exposing PS to the external cellular

environment. The human anticoagulant, annexin V, is a 35-36-kDCa

dependent phospholipid binding protein that has a high affinity for

PS. Labeled Annexin V can identify apoptotic cells by binding to PS

exposed on the outer leaflet. This test was performed using

Vybrant Apoptosis Assay Kit # 2 (Invitrogen, USA). The kit contains

recombinant annexin V conjugated to Alexa Flour 488 dye. The kit also

65

contains a ready-to-use solution of the red fluorescent propidium iodide

(PI) nucleic acid binding dye. PI is impermeant to live cells and apoptotic

cells, but stains dead cells with red fluorescence, binding tightly to the

nucleic acids cells in the cell.

The apoptosis was induced by transfecting HeLa cells with

pSin.cav.vp3 recombinant plasmid as described earlier. After 48 hrs of

transfection, cells were collected by trypsinization in a 1.5 ml microfuge

tube and washed in cold PBS. Then 1X annexin binding buffer was

prepared by mixing 200 µl of 5X annexin binding buffer to 800 µl of

distilled water to a total volume of 1 ml. A working solution of

propidium iodide (PI) (100 µg/ ml) was prepared by diluting 5 µl of the

1 mg/ml PI stock solution in 45 µl of 1 x annexin binding buffer. The

washed cells were recentrifuged and the supernatant was discarded. The

cells were resuspended in 1 x annexin binding buffer. 5 µl of annexin V

conjugate and 1 µl of 100 mg /ml PI working solution were added to 100

µl of cell suspension. The cells were incubated at room temperature for

15 minutes and then washed with 1x Annexin Binding buffer. The cells

were deposited onto slides, examined under fluorescent microscope and

photographed.

Results

4.1 Cloning of VP3 gene in pSin vector

4.2 Expression of VP3 gene in HeLa cells

4.3 Apoptic activity of VP3 in cell line

Chapter-4

66

Chapter-4

Results

4.1 Cloning of VP3 gene in pSin vector

E. coli cells containing pTarget.cav.vp3 were grown on LB broth

containing ampicillin (100 µg/ml) and plasmid DNA was isolated and

run on agarose gel electrophoresis which yielded good amount of DNA.

The cav.vp3 gene insert was released from pTarget vector by

digesting with EcoRI enzyme (Fig. 2). Two different bands were seen.

Smaller band was of insert DNA and the larger band was of vactor DNA.

The size of smaller band i.e. vp3 gene was 372 bp and the size of larger

band i.e. pTarget vector was 5670 bp. Smaller band (insert) was cut and

ligated to pSin mammalian expression vector. The recombinant DNA was

designated as pSin.cav.vp3. This was transformed in E. coli DH5 alpha

competent cells using single step method of competent cell preparation

and transformation. In this method, bacteria do not require a heat shock

for optimal uptake of plasmid DNA. The transformed cells were plated on

LB agar plates containing ampicillin 50 µg/ ml and the plates were

incubated at 37˚C for 16 hrs. After 16 hrs incubation at 370C, the plates

were found to contain several colonies from which ten colonies were

selected randomly and grown in LB broth containing ampicillin. The

67

recombinant pSin.cav.vp3 plasmids were isolated from these colonies by

TELT method.

The restriction sites predicted in pSin.cav.vp3 plasmid using

DNASTAR software showed that one BglII site is present in the pSin

vector and one site in the region of cav.vp3 gene. When the gene was in

correct orientation, three fragments were expected viz. 5650 bp, 2180 bp

and 3321 bp. The digestion of pSin.cav.vp3 with BglII yielded three

fragments of expected size (Fig. 3).

Colony PCR using VP3 gene specific forward primer and BGH as

reverse primer yielded 372 bp PCR product showing that the insert was in

right orientation (Fig. 4). The recombinant plasmid selected after above

two methods was sequenced using BGH reverse primer and sequence

confirmed that the gene was in right orientation and the sequence was

also homologous to VP3 gene.

68

M 1

Fig. 2 : Digestion of recombinant plasmid (pTarget.cav.vp3) with restriction enzyme (EcoRI) to release the vp3 gene insert.

Lane M : 100 bp marker

Lane 1 : pTarget vector 5670 bp and cav.vp3 gene 372 bp.

372 bp

5670 bp

69

M 1

Fig. 3: The digestion of pSin.cav.vp3 with BglII yielded three fragments of

expected size.

5650 bp

4.0 kb

3.0 kb

2.0 kb

5.0 kb

3321 bp

2180 bp

Fig. 4 :

Lane M :

Lane1 :

70

M 1

: PCR amplification of cav.vp3 gene.

: 1 kb ladder

: CAV-VP3 PCR product

372 bp

71

4.2. Expression of VP3 gene in HeLa cells

The cav.vp3 gene cloned in pSin verctor was analysed for its

expression in Hela cells using Immunoperoxidase Test (IPT) and SDS-

PAGE.

The transfection was done into the wells of 96 well microtitre plate.

The transfected cells were incubated at 370C in a humidified chamber

with an atmosphere of 5% CO2. After 72 hrs of incubation, the cells were

assayed for expression of transfected gene. The expression was seen by

Immunoperoxidase Test (IPT), also known as Immunoenzyme technique

(IET). In this test, there occurs a purple/ brown color precipitation

reaction. Immunoperoxidase test of pSin.cav.vp3 plasmid transfected

HeLa cells revealed brown color in transfected cells (Fig. 5), which

confirmed the expression of cav.vp3 gene. However, control showed no

such color (Fig. 6). It was also found that more than 80% of the cells in a

microscopic view showed expression.

The SDS-PAGE (Sodium dodecyl sulphate-polyacrylamide gel

electrophoresis) analysis of expressed protein showed a specific band of

size 13 kDa (Fig. 7). Expression of vp3 in HeLa cells by the gene construct

pSin.cav.vp3 was established and the gene construct was used for

evaluating the apoptic potential of vp3 in HeLa cells.

72

Fig. 5: HeLa cells transfected with pSin.cav.vp3 rplasmid showing positive

IPT test, indicating expression of gene, magnification 100x.

73

Fig. 6: Healthy control HeLa cells showing no color reaction, magnification 100x.

74

Fig. 7 : VP3 expressed protein analyzed by SDS-PAGE.

Lane M : Protein molecular weight marker

Lane 1 : 13 kDa expressed protein

75

4.3 Apoptic activity of VP3 in cell line

The apoptosis inducing potential of CAV-VP3 (apoptin) was

evaluated in HeLa cells by the following three assays:

i. DNA laddering assay

ii. Caspase detection assay

iii. Annexin-V-binding assay

DNA laddering assay

In DNA laddering assay, the cleavage of chromosomal DNA occurs

into the oligonucleosomal size fragments, which is an integral part of

apoptosis. When HeLa cells, transfected with rplasmid, were analysed

after 72 hrs, the bands merged and nucleosomal laddering was detected

on agarose gel electrophoresis (Fig. 8) while control cells showed no such

laddering pattern.

Caspase detection assay

The caspases are a family and are one of the main executers of

apoptotic process. These proteins breakdown or cleave key cellular

components that are required for normal; cellular function including

structural proteins in cytoskeleton and nuclear proteins such as DNA

repair enzymes. The caspases can also activate other degrative enzymes

such as DNAses which begin to cleave the DNA in the nucleus. In caspase

detection assay, when HeLa cells transfected with rplasmid, were analysed

after 72 hrs, caspase 3 positive cells showed green fluorescence (Fig. 9 &

Fig. 10). However, control cells showed no such fluorescence.

76

Annexin-V-binding assay

Annexin V was used as a probe in this assay, to detect the cells that

have expressed phosphatidylserine on the cell surface, a feature found

in apoptosis. Phosphatidyl serine at the outer membrane surface of a cell

is a universal process occurring during early apoptosis. Using the

Annexin-V affinity assay, the apoptic cells in suspension was

determined in a fast, simple and sensitive way. Annexin V staining was

specific to apoptic cells and background staining was low in unaffected

cells. In Annexin-V-binding assay, when HeLa cells transfected with

rplasmid were analysed after 72 hrs, apoptic cells showed bright green

fluorescence while control showed yellow fluorescence (Fig. 11).

All these three assays revealed that CAV-VP3 showed good apoptic

activity in cultured cells. It was also observed that majority of the cells

showed apoptic activity.

77

1 2

Fig. 8 : DNA fragmentation assay.

Lane 1 : Control showing no laddering pattern

Lane 2 : DNA laddering pattern of DNA from pSin.cav.vp3 transfected

HeLa cells.

78

Fig. 9: Caspase positive HeLa cells showing green fluorescence, indicating

positive apoptic reaction, magnification 100x.

79

Fig. 10: Caspase positive HeLa cells showing green fluorescence, indicating positive apoptic reaction, magnification 10x.

80

Fig. 11: Annexin-V positive HeLa cells showing bright green fluorescence, indicating positive apoptic reaction, magnification 100x

Discussion

Chapter-5

81

Chapter-5

Discussion

The CAV genome has been shown to incorporate three overlapping

open reading frames (ORFs) which code for three viral proteins

designated as VP1, VP2 and VP3, all of which were expressed in cells

infected with CAV (Noteborn and Koch, 1995).

CAV causes severe depletion of some cell types, such as

lymphocytes, by the induction of apoptosis (Noteborn et al., 2004). It

causes cytopathogenic effects in chicken thymocytes and cultured

transformed mononuclear cells through apoptosis. Early after infection of

chicken mononuclear cells, the CAV-encoded protein VP3 exhibited a

finely granular distribution within the nucleus. At a later stage after

infection, VP3 forms aggregated. At this point, the cell became apoptotic

and the cellular DNA was fragmented and condensed. By immunogold

electron microscopy, VP3 was shown to be linked with apoptotic

structures. In vitro, expression of VP3 induced apoptosis in chicken

lymphoblastoid T cells and myeloid cells, which were susceptible to CAV

infection (Noteborn et al., 1994).

The present study, as already explained, was undertaken to clone

the cav.vp3 gene in pSin vector, in vitro expression of VP3 in cultured cells

and the study of apoptotic activity of apoptin (VP3 protein) in cell line.

82

The VP3 gene of CAV was successfully cloned in pSin vector. pSin is

derived from Sindbis virus, which is one of the members of Alphavirus

genus. Several members of the Alphavirus genus, first Sindbis virus

(Bredenbeek et al., 1993; Schlesinger, 1993; Xiong et al. 1989), later

Semliki Forest virus (Berglund et al., 1993; Lilgestorm 1994; Lilgestorm

and Garoff 1991) and other alphavirus members (Davis et al. 1995; Davis

et al., 1989), have received considerable attention for use as virus-based

expression vectors. Typically, these vectors are constructed in a format

known as a replicon, due to the self-amplifying nature of the vector RNA

(Xiong et al., 1989). We have used pSin vector due to the following

reasons. There are many properties which make alphavirus vectors a

desirable alternative to other virus-derived vector systems being

developed. These properties include potential high-level expression of up

to 108 molecules of heterologous protein per cell (Xiong et al., 1989),

infection of nondividing cells and a broad host range (Strauss and Strauss

1994). In addition, replication occurs entirely in the cytoplasm of the

infected cell as RNA molecule, without a DNA intermediate. This is in

contrast to retrovirus and adeno-associated virus vectors, which enter the

nucleus and usually integrate into the host genome for initiation of vector

activity (Jolly 1994; Miller et al., 1993; Samulski et al., 1989). Thus,

retrovirus-associated and adeno-associated virus-derived vectors have

application for long-term expression of foreign proteins, while the

alphavirus vectors are likely better suited for short-term high-level

expression. Furlov et al. (1996) explained the advantages of alphavirus,

which include a broad range of susceptible host cells, high levels of

cytoplasmic RNA and protein expression without splicing the facile

83

construction and manipulation of recombinant RNA molecules using full-

length cDNA clones from which infectious RNA transcripts can be

generated by in vitro transcription. In review of the above, we have seen

the expression of cav.vp3 gene cloned in pSin vector and approximately

more than 80% of cells were found to express the protein. Moreover, it has

the advantage that less amount of DNA is needed to be delivered for

therapeutic purpose.

Earlier cav.vp3 gene was cloned in different vectors by different

authors. Pietersen et al. (1999) used adenovirus vector, AdMLPvp3, for the

expression of apoptin and demonstrated that adenovirus vectors for the

expression of the apoptin gene may constitute a powerful tool for the

treatment of solid tumors. Shen et al. (2003) showed the antitumor effect

of VP3 protein, especially its effect against liver carcinoma in vivo. The

recombinants pcDNA-vp3 containing chicken anemia virus vp3 gene, and

control vector pcDNA3 were mixed with murine liver carcinoma cell lines

H22 respectively. They found that cav vp3 might be a potential new gene

in reducing the growth rate of tumor cells in liver carcinoma or in other

kind of solid tumors in vivo. Nogueira-Dantas et al. (2007) indicated that

recombinant VP3 expressed in the pTrcHis2 vector system can be used as

antigen to detect anti-cav antibodies. Li et al. (2010) generated a

conditional replication-competent adenovirus (CRCA), designated Ad-

hTERT-E1a-Apoptin, and investigated the effectiveness of the CRCA a

gene therapy agent for further clinical trials and observed that infection

with Ad-hTERT-E1a-Apoptin significantly inhibited growth of the

melanoma cells. Jiang et al. (2010) developed lentiviral vector, encoding a

green fluorescent protein-apoptin fusion gene (LV-GFP-AP) that can

84

efficiently deliver apoptin into hematopoietic cells. Saxena et al. (2012)

used pcDNA vector to clone the vp3 gene and observed its oncolytic

property. Birame et al. (2012) constructed eukaryotic expression vector

pIRES2-EGFP-apoptin and pIRES2-EGFP-ABPS1 and observed their

expression effects individually and in combinations with HepG2 and A375

cells. The apoptotic rates obtained in HepG2 cells, treated with apoptin

and ABPS1 together, were closely similar, but, not in A375 cells. Therefore,

the results of their study showed that the combination of apoptin and

ABPS1 has synergistic effect in HepG2 and A375 cell lines.

In the present study, the expression of VP3 gene in HeLa cells was

confirmed by Immunoperoxidase Test (IPT) and SDS-PAGE and thus

indicating that apoptotic action seen in HeLa cells was due to the action of

VP3 protein, which was confirmed by in vitro studies. The

Immunoperoxidase test is easy to detect expressed protein in cell culture.

This test was found to be simple and reliable. It was easy to carry out, it

did not require any special technical equipment and moreover all reagents

were easy to obtain. In SDS-PAGE, the protein of 13kDa was observed.

Earlier expression of vp3 gene was reported by different workers using

different methods like Indirect Immunofluorescence technique (Natesan et

al., 2006). The CAV DNA sequence had three partially overlapping major

reading frames coding for putative peptides of 51.6, 24.0, and 13.6 kDa

(Noteborn et al., 1991). The cell lines MDCC-MSBI and LSCC-HD11 were

transfected with CAV plasmid DNA, the CAV infected cells were anlysed

using immunoblotting and immunoperoxidse assay and the size of VP3

protein was found to be 16 kDa, which is a strong inducer of apoptosis

(Noteborn et al., 1994). Nogueira-Dantas et al. (2007) showed that the in

85

vitro expressed VP3 protein was purified to near homogeneity by elution

from the gel, as judged by sodium dodecyl sulfate-polyacrylamide gel

electrophoresis analysis. The purified VP3 protein with a molecular mass

of approximately 21kDa was confirmed by Western blotting analysis.

Wang et al. (2009) constructed a prokaryotic expression vector

for apoptin and prepared polyclonal antibody of apoptin. The

apoptin protein expression induced by IPTG was analyzed by SDS-PAGE.

Han et al. (2010) showed that the expression of the fused gene of SP-TAT-

Apoptin in 293FT cells infected by the recombinant lentivirus was

examined by immunofluorescence histochemistry method. The

recombinant lentivirus of SP-TAT-apoptin was successfully packaged and

it can induce HepG2 cells to apoptosis. Cao et al. (2010) reported that

Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand (TRAIL) and

VP3 genes were cloned into a pBudCE4.1 vector and delivered by

attenuated Salmonella typhimurium into gastric cancer cells, and

their expression and antitumor effects in nude mice were monitored by

Western blot and fluorescence microscopy. Delivery of TRAIL and VP3

genes by attenuated S. typhimurium could significantly inhibit the growth

of gastric cancer cells in vitro and in vivo. Li et al. (2012) investigated the

pro-apoptotic effect of 7, 3'-dimethoxy hesperetin (DMHP) on

synovial apoptosis in vivo. Bcl-2, Bax mRNA and protein expressions in

synovium were determined by western blot and it was found that DMHP

treatment on adjuvant arthritis (AA) rats significantly decreased the

protein ratio of Bcl-2/Bax in synovium. The 16 kDa His-TAT-Apoptin and

the 42 kDa GST-TAT-Apoptin were detected using monoclonal anti-His

antibody and anti-GST antibody respectively (Lee et al., 2012). Saxena et

86

al. (2012) showed expression of VP3 in HeLa cells, transfected with

pcDNA.cav.vp3. Expression was confirmed by flow cytometry, IFAT.

IFAT in cells transfected with pcDNA.cav.vp3 using polyclonal sera

showing specific binding of antibody to the expressed 13.4 kDa vp3

protein when compared to the control untransfected cells. The size of

apoptin protein and apoptin like protein from other viruses, varied due to

size of the recombinant construct used by these workers. In certain cases,

His tag was used in construct, which increased the total size of expressed

protein.

In our study, induction of apoptosis in HeLa cells was analysed by

employing DNA laddering assay, caspase 3 detection assay and Annexin-

V-binding assay. In DNA laddering assay, nucleosomal laddering was

detected on 2% agarose gel electrophoresis. The bands were found to be

merged and a laddering pattern was observed. The nucleosomal laddering

is a hallmark of apoptosis. Li et al. (2006) detected DNA ladder to

investigate the pro-apoptotic effect of 7, 3'-dimethoxy hesperetin (DMHP)

on synovial apoptosis in vivo. Typical DNA ladder formation was found

in DNA extraction of synovium from DMHP treated groups. Cipak et al.

(2007) investigated anticancer activity of newly synthesized 2-

phenoxymethyl-3H-quinazolin-4-one (PMQ). Apoptosis was analysed

using internucleosomal DNA fragmentation assay. DNA ladder formation

assay indicated that PMQ actively induced apoptosis of cells. Ahmed et al.

(2008) investigated whether endosulfan, an organochlorine pesticide was

able to deplete glutathione (GSH) and induce apoptosis in human

peripheral blood mononuclear cells (PBMC) in vitro. Apoptotic cell death

was determined by DNA fragmentation assays. Significant ladder

87

formation was observed at higher concentration, which was indicative of

apoptotic cell death. Mohamed (2010) detected DNA ladder assay with

0.9% agarose. He observed DNA laddering pattern with multiple bands of

180-200 bp and its multiplication that is considered a characteristic feature

of apoptosis. Kucharski et al. (2011) demonstrated that apoptin was able

to inhibit formation of DNA damage foci by targeting the APC/C-

associated factor MDC1 for degradation. They suggested that these results

may point to a novel mechanism of DNA damage response inhibition

during viral infection. Jiang et al. (2012) reported that HepG2 cells were

incubated with tectorigenin at different concentrations, and their viability

was assessed by 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium

bromide assay. Apoptosis was detected by morphological observation of

nuclear change and agarose gel electrophoresis of DNA ladder.

Tectorigenin at a concentration of 20 mg/L greatly inhibited the viability

of HepG2 cells and induced the condensation of chromatin and

fragmentation of nuclei. Kumar et al. (2012) showed that Ochratoxin A

(OTA) and citrinin (CIT) were nephrotoxic mycotoxins produced mainly

by fungal species Aspergillus ochraceus and Penicillium citrinum,

respectively, which have been found to occur together in various food and

feed commodities. Both OTA and CIT were evaluated for their potential to

induce apoptosis by flow cytometry and gel electrophoresis. The

concentration of MDA was found significantly higher in OTA and

combination-treated groups. OTA and combination-treated groups

revealed more apoptotic cells in flow cytometry when compared with the

CIT-treated group.Characteristic DNA fragmentation, as evidenced

by ladder pattern in electrophoresis appeared in the toxin-treated groups.

88

Steigerova et al. (2012) surveyed the effects of brassinosteroids (BR) on

prostate cancer cells using DNA ladder assays. They observed BRs

inhibited cell growth.

In caspase detection assay, when HeLa cells transfected with

rplasmid were analysed after 72 hrs, caspase 3 positive cells showed green

fluorescence. However, control cells showed no such fluorescence.

Activation of caspase cascade plays a central role in execution of apoptosis

by cleaving a large number of proteins. Danen-van Oorschot et al. (2000)

demonstrated that activation of upstream caspases was not required,

activation of caspase-3 and possibly also other downstream caspases was

essential for rapid apoptin-induced apoptosis. Lesauskaite and

Ivanoviene (2002) showed that activation of caspases plays a major role in

the execution of apoptosis. These activated caspases selectively cleave

cellular proteins, which results in apoptotic morphology. Reed (2004)

showed that numerous proteins that regulate the cell death proteases have

been discovered, including proteins belonging to the Bcl-2, inhibitor of

apoptosis, caspase-associated recruitment domain, death domain, and

death effector domain families. Wang et al. (2005) showed that apoptin-

induced apoptosis does not depend on functional p53, and can't be

inhibited by overexpression of Bcl-2 and Bcl-xL, but caspase-3 activation

was necessary for apoptin-induced rapid apoptosis. Burek et al. (2006)

showed that cell death induction by apoptin was associated with

cytochrome c release from mitochondria as well as with caspase-3 and -7

activation and observed that apoptin-induced apoptosis was caspase

dependent. Schoop et al. (2008) observed that the active form of caspase 3

was present only in apoptin positive cells having an apoptic morphology.

89

Lin. et al. (2010) showed that maternal diabetes might increase oocyte

apoptosis by a Bax-caspase-3 pathway to play a role in embryonic

malformations by delayed oocyte development. Tumane et al. (2010)

observed that silica-induced apoptosis of the alveolar macrophages could

potentially favor a proinflammatory state, occurring in the lungs of

silicotic patients, resulting in the activation of caspase prior to induction of

the intrinsic and extrinsic apoptosis pathways. In addition, caspase could

be a key apoptotic protein that can be used as an effective biomarker for

the study of occupational diseases. It may provide an important link in

understanding the molecular mechanisms of silica-induced lung

pathogenesis. Han et al. (2011) did combination therapy with radiation

and apoptin which dramatically induced mitochondrial cytochrome c

release and the cleavage of caspases -9, -3 and -7 and showed that apoptin

treatment represented a potential method for enhancing the effectiveness

of radiotherapy in poorly responding hepatocellular carcinoma. It is

evident that the caspases are a family that are one of the main executers of

apoptotic process. It was seen that our work is in conformity with other

workers.

In our study, the Annexin-V-binding assay revealed apoptic cells

showeing bright green fluorescence while control cells showed faint green

fluorescence. In double labeling experiments, using Annexin V binding

and counterstaining with the supervital DNA dye Hoechst 33342,

Koopman et al. (1994) in their experiment showed that cells showing

chromatin condensation were annexin v positive. Martin et al. (1995)

showed that annexin-V binds specifically to phosphatidyl serine in a

calcium-dsependent manner and has been used to stain dying cells that

90

expose phosphatidyl serine on their cell surface Van Engeland et al.

(1996) showed that it was feasible to use the bivariate PI/annexin v

analysis for adhering cell in culture. Eijende et al. (1997) used annexin v

assay for the detection of apoptotic cells in situ, biotin- labeled annexin v

was injected into mice. In this way, apoptin could be detected in

developing mouse embryo. So, apoptin can be effectively used for in situ

purpose. Van Engeland et al. (1998) demonstrated that Annexin V

interacted strongly and specifically with phosphatidyl serine (PS) and

could be used to detect apoptosis by targeting for the loss of plasma

membrane asymmetry. Kekre et al. (2005) showed that Annexin-V assay

was carried out at several time-points in order to monitor phosphatidyl

serine flipping to the outer leaflet of the plasma membrane, which is a

characteristic apoptotic event. Annexin-V staining is specific to

apoptotic cells, and background staining is low in unaffected cells Lin. et

al. (2010) showed that apoptosis might be closely involved in diabetes-

induced embryonic malformations. Diabetic mouse ovarian sections

stained with hematoxylin and eosin showed reduced number of growing

follicles and delayed oocyte development. Annexin V-positive oocytes

were higher in number in diabetic mice than in non-diabetic mice. Han et

al. (2010) revealed that Annexin-V PI assay showed that SP-TAT-Apoptin,

carried by the recombinant lentivirus, could cause the HepG2 cell

apoptosis and its apoptosis rate was significantly more than paired control

group and SP-TAT-Apoptin carried by liposomes only. The recombinant

lentivirus of SP-TAT-Apoptin is successfully packaged and it can induce

HepG2 cells to apoptosis. Jiang et al. (2012) showed that HepG2 cells were

incubated with tectorigenin at different concentrations, and their viability

91

was assessed by 3-(4, 5-dimethylthiazol-2-yl)-2, 5-diphenyltetrazolium

bromide assay. Apoptosis was detected by Annexin V-EGFP and

propidium iodide staining. Tectorigenin induced apoptosis of HepG2 cells

mainly via mitochondrial-mediated pathway, and produced a slight

cytotoxicity to L02 cells. It was seen that our work is in conformity with

other workers.

Van der Eb et al. (2002) described that multiple AdMLP.apoptin

injections with apoptin gave an evenly distribution throughout the tumor.

Peng et al. (2007) demonstrated that systemic delivery of Asor-apoptinin

mice bearing hepatocarcinomas resulted distribution of apoptin in both

hepatocarcinoma cells and normal liver cells. The hepatocarcinomas showed

clear regression, while normal hepatocytes were not affected. These

remarkable observations can be apotential breakthrough for the current

encountered dilemma of lack of delivery of therapeutic genes in targeted

cells. Further studies with this attractive vehicle linked to apoptin could

allow to choose the most suitable vector for treatment of squamous cell

carcinoma.

Apoptosis is a complex, genetically-determined process involved in

the growth and maintenance of homeostasis in multicellular organisms.

Deregulation of apoptosis has been implicated in a number of diseases,

including cancer and autoimmune disease. Thus, the examination of

apoptotic regulation has evoked considerable interest. Many apoptotic

proteins have been shown to be post-translationally modulated, such as

by protein cleavage, protein-protein interaction, translocation and various

post-translational modifications, which fall specifically within the range of

proteomic analysis. Recently, current proteomic technologies have

92

achieved significant advances and have accelerated research in functional

and chemical proteomics, which have been applied to the field of

apoptosis research and have the potential to be a driving force for the

field (Wang et al., 2011).

The constructed recombinant plasmid has replicase gene which

produces large amount of apoptin protein and thus only small amount of

DNA will be needed to be injected for therapeutic agent. The DNA can be

stored at room temperature. The present study also clearly showed the

apoptotic activity of VP3 protein in HeLa cell line and further studies are

required to exploit its therapeutic potential.

Summary and Conclusion

Chapter-6

93

Chapter-6

Summary and Conclusion

CAV is one of the smallest avian viruses; it is 23–25 nm in size,

icosahedral in shape and non-enveloped, having a 2.3 kb, circular, single-

stranded DNA genome. The genome encodes three viral proteins (vp1,

vp2 and vp3) that are transcribed from a single major transcript of 2.0 kb.

It is believed that the cav genome replicates through the rolling-circle

model. In the present study the vp3 gene was cloned in pSin mammalian

expression vector and the resultant recombinant plasmid pSin.cav.vp3

was then characterized by restriction digestion with BglII and sequencing.

Expression of recombinant clone was confirmed by transfection in HeLa

cell line by Immunoperoxidase test and SDS-PAGE.

The apoptosis inducing potential of cav.vp3 (apoptin) was studied

in cultured HeLa cell line. The present study confirmed that vp3 induced

apoptosis in HeLa cells, which was confirmed by demonstrating the

characteristics changes of apoptosis viz. nuclear condensation, DNA

fragmentation by DNA laddering assay, plasma membrane alteration by

annexin-V binding assy. Caspase 3 was also detected after 48 h of

transfection of Hela cell line by pSin.cav.vp3. Further in vivo trials may be

undertaken to know its therapeutic potential.

Future Scope

Chapter-7

94

Chapter-7

Future Scope

Tumors are commonly prevalent in Indian population. Apoptin

induces apoptosis/ destruction of tumor cells in human transformed and

malignant cells but not in normal cells. The apoptin (VP3 protein) derived

from chicken anemia virus is a potential agent for the treatment of a large

number of tumors. The VP3 protein (apoptin) induces apoptosis in

chicken mononuclear cells. It has attracted great attention, because it

specifically kills tumor cells while leaving normal cells unharmed. Other

tumorogenic agents show resistance when injected in body but apoptin

does not show any such type of resistance. In normal cells, apoptin resides

in the cytoplasm, whereas in cancerous cells it translocates into the

nucleus. Animal tumor models have revealed apoptin as a safe and

efficient antitumor agent, resulting in significant tumor regression. Future

antitumor therapies could use apoptin as a potential therapeutic drug.

The constructed recombinant plasmid has replicase gene which produces

large amount of apoptin protein and thus only small amount of DNA will

be needed to be injected for therapeutic agent. The DNA can be stored at

room temperature.

Future antitumor therapies could use apoptin either as a therapeutic

drug or as an early sensor of druggable tumor-specific processes

(Backendorf et al. 2008). For improved treatments to be developed, the

95

ability to target tumor cells selectively is essential, allowing a greater dose

of therapeutic agent to be delivered without adversely affecting non-

malignant tissue and further studies are required to exploit its therapeutic

potential.

References

96

References

Ahmed, T., Tripathi, A.K., Ahmed, R.S., Das, S., Suke, S.G., Pathak,

R., Chakraboti, A. and Banerjee, B.D. (2008): Endosulfan-induced

apoptosis and glutathione depletion in human peripheral blood

mononuclear cells: Attenuation by N-acetylcysteine. J Biochem Mol

Toxicol, 22: 299-304.

Akasaka, Y., Ono, I., Kamiya, T., Ishikawa, Y., Kinoshita, T., Ishiguro,

S., Yokoo, T., Imaizumi, R., Inomata, N., Fujita, K., Akishima-

Fukasawa, Y., Uzuki, M., Ito, K. and Ishii, T. (2010): The mechanisms

underlying fibroblast apoptosis regulated by growth factors during

wound healing. J Pathol, 221: 285-299.

Aldair B.M. (2000): Immunopathogenesis of chicken anemia virus

infection. Dev Comp Immunol, 24: 247-255.

Alvisi, G., Poon, I.K. and Jans, D.A. (2006): Tumor-specific nuclear

targeting: promises for anti-cancer therapy? Drug Resist Updat, 9: 40-50.

Argiris, K., Panethymitaki, C. and Tavassoli, M. (2011): Naturally

occurring, tumor-specific, therapeutic proteins. Exp Biol Med, 236: 524-

536.

Backendorf, C., Visser, A.E., De Boer, A.G., Zimmerman, R., Visser,

M., Voskamp, P., Zhang, Y.H. and Noteborn, M.H.M. (2008): Apoptin:

97

therapeutic potential of an early sensor of carcinogenic transformation.

Ann Rev Pharmacol Toxicol, 48: 143-169.

Banerjee, M., Datta, M., Majumder, P., Mukhopadhyay,

D. and Bhattacharyya, N.P. (2010): Transcription regulation of caspase-1

by R393 of HIPPI and its molecular partner HIP-1. Nucleic Acids Res, 38:

878–892.

Berglund, P., Sjoberg, M., Atkins, G.J., Sheahan, B.J., Garoff, H. and

Liljestrom, P. (1993): Semliki Forest virus expression system: production

of conditionally infectious recombinant particles. Bio/Technology, 11: 916–

920.

Bhatt P., Shukla, S.K., Mahendran, M., Dhama, K., Chawak, M.M. and

Kataria, J.M. (2011): Prevalence of chicken infectious anaemia virus

(CIAV) in commercial poultry flocks of northern India: a serological

survey. Transbound Emerg Dis, 58: 458-460.

Birame, B.M., Jigui, W., Fuxian, Y., Shuang, W., Li, Y.D., Jiazeng, S.,

Zhili, L., Bao, Y. and Weiquan, L. (2012): Co-expression of apoptin (VP3)

and antibacterial peptide cecropin B mutant (ABPS1) genes induce higher

rate of apoptosis in HepG2 and A375 cell lines. African J Biotech, 11:

10981-10988.

Bredenbeek, P.J., Frolov, I., Rice, C.M. and Schlesinger, S. (1993):

Sindbis virus expression vectors: packaging of RNA replicons by using

defective helper RNAs. J. Virol., 67: 6439–6446.

98

Burek, M., Maddika, S., Burek, C.J., Daniel, P.T., Schulze-Osthoff, K.

and Los, M. (2006): Apoptin-induced cell death is modulated by Bcl-2

family members and is Apaf-1 dependent. Oncogene, 25: 2213-2222.

Cao, H.D., Yang, Y.X., Lü, L., Liu, S.N., Wang, P.L., Tao, X.H., Wang, L.J.

and Xiang, T.X. (2010): Attenuated Salmonella typhimurium carrying

TRAIL and VP3 genes inhibits the growth of gastric cancer cells in vitro

and in vivo. Tumori, 96: 296-303.

Chen, K., Luo, Z., Tang, J. and Zheng, S.J. (2011): A critical role of heat

shock cognate protein 70 in apoptin-induced phosphorylation of Akt.

Biochem Biophys Res Commun, 409: 200-204.

Chiantore, M.V., Vannucchi, S., Mangino, G., Percario, Z.A., Affabris,

E., Fiorucci, G., Romeo, G. (2009): Senescence and cell death pathways

and their role in cancer therapeutic outcome. Curr Med Chem, 16: 287-300.

Chung, C, T., Niemela, S, L. and Miller, R, H. (1989): One step

preparation of competent Escherichia coli: Transformation and storage of

bacterial cells in the same solution. Proc. Natl. Acad. Sci, 86: 2172-2175.

Cipak, L., Repicky, A. and Jantova, S. (2007): Growth inhibition and

apoptosis induced by 2-phenoxymethyl-3H-quinazolin-4-one in HL-60

leukemia cells. Exp Oncol, 29: 13-17.

Danen-van Oorschot, A.A.A.M., Van der Eb, A.J., and Noteborn,

M.H.M. (2000): The chicken anemia virus-derived protein Apoptin

requires activation of caspases for induction of apoptosis in human

tumor cells. J. Virol, 74: 7072-7078.

99

Davis, N.L., Brown, K.W., Caley, I.J., Swanstrom, R.I. and Johnston,

R.E. (1995): Protection against influenza in mice by vaccination with a

Venezuelan equine encephalitis virus vector expressing the HA protein.

J. Cell. Biochem. Suppl., 19: 310.

Davis, N.L., Willis, L.V., Smith, J.F. and Johnston, R.E. (1989): In vitro

synthesis of infectious Venezuelan equine encephalitis virus RNA from a

cDNA clone: analysis of a viable deletion mutant. Virology, 171: 189–204.

De Herdt, P., Van den Bosch, G., Ducatelle, R., Uyttebroek, E. and

Schrier, C. (2001): Epidemiology and significance of chicken infectious

anemia virus infections in broilers and broiler parents under non

vaccinated European circumstances. Avian Dis, 45: 706-708.

De Smit, M.H. and Noteborn, M.H. (2009): Apoptosis-inducing proteins

in chicken anemia virus and TT virus. Curr Top Microbiol Immunol, 331:

131-149.

Eijnde, S.M., Boshart, L., Reutelingsperger, C.P., Zeeuw, C.L. and

Vermeij-Keers, C. (1997): Phosphatidyl serine plasma asymmetry in

vivo: A pancellular phenomenon which alters during apoptosis. Cell

Death Diff, 4: 311-316.

Eltahir, M.Y., Qian, K., Jin, W. and Qin, A. (2011): Analysis of chicken

anemia virus genome: evidence of intersubtype recombination. Virol J, 8:

512.

100

Furlov, I., Hoffman, T.A., Pragai, B.M., Dryga, S.A., Huang, H.V.,

Schlesinger, S. and Rice, C.M. (1996): Aiphavirus-based expression

vectors: Strategies and applications. Proc. Natl. Acad. Sci., 93: 11371-11377.

Gerna, G., McCloud, C. and Chambers, R. W. (1976):

Immunoperoxidase technique for detection of antibodies to human

cytomegalovirus. J Clinical Microbiol, 3: 364-372.

Ghobrial, I.M., Witzig, T.E. and Adjei, A.A. (2005): Targeting apoptosis

pathways in cancer therapy. CA Cancer J Clin, 55: 178-194.

Guelsen, L., Paterson, H., Gäken, J., Meyers, M., Farzaneh, F. and

Tavassoli, M. (2004): TAT-apoptin is efficiently delivered and induces

apoptosis in cancer cells. Oncogene, 23: 1153–1165.

Hailemariam, Z., Omar, R.A., Hair-Bejo, M. and Giap, C.T. (2008):

Detection and characterization of chicken anemia virus from commercial

broiler breeder chickens. Virol J, 5: 128.

Han, S.X., Zhao, J., Ma, J.L., Huang, C., Lv, Y., Ou, W. and Jia, X. (2010):

The effect of the fused gene of SP-TAT-Apoptin transfected by lentivirus

on HepG2 cells. Chinese J Cell Mol Immunol, 26: 301-302.

Han, S.X., Zhu, Q., Ma, J.L., Lv, Y., Zhao, J., Huang, C., Jia, X., Ou, W.

and Guo, H.T. (2011): Apoptin sensitizes radiation-induced cell death via

classic mitochondrial, caspase and p53-dependent signaling in HepG2

cells. Mol Med Report, 4: 59-63.

101

He, C., Ding, N., Li, J. and Li, Y. (2002): Gene cloning and sequencing of

chicken anemia virus (CAV) isolated from Harbin. Wei Sheng Wu Xue

Bao, 42: 436-441.

He, Wilde and Kaderbhai, M.A. (1990): A simple single step procedure

for small scale preparation of Escherichia Coli plasmid. Nucleic Acids Res,

18: 1660.

He, X, Zhang, Q., Liu, Y. and He, P. (2005): Apoptin induces chromatin

condensation in normal cells. Virus Genes, 31: 49-55.

Heilman, D.W., Green, M.R. and Teodoro, J.G. (2005): The anaphase

promoting complex: a critical target for viral proteins and anti-cancer

drugs. Cell Cycle, 4: 560-563.

Hino, S. and Prasetyo, A.A. (2009): Relationship of Torque teno virus to

chicken anemia virus. Curr Top Microbiol Immunol, 331: 117-130.

Ismail, I.H., Andrin, C., McDonald, D. and Hendzel, M.J. (2010): BMI1-

mediated histone ubiquitylation promotes DNA double-strand break

repair. J Cell Biol, 191: 45-60.

Jain, V., Armah, B.H., Tongren, E.J., Ned, M.R., Wilson, O.N.,

Crawford, S., Joel, K.P., Singh, P.M., Nagpal, C.A., Dash, A.P.,

Udhayakumar, V., Singh, N. and Stiles, K.J. (2008): Plasma IP-10,

apoptotic and angiogenic factors associated with fatal cerebral malaria in

India. Malaria Journal, 7: 83.

Janssen, K., Hofmann, T.G., Jans, D.A., Hay, R.T., Schulze-Osthoff K.

and Fischer U. (2007): Apoptin is modified by SUMO conjugation and

102

targeted to promyelocytic leukemia protein nuclear bodies. Oncogene, 26:

1557-1566.

Jiang, C.P., Ding, H., Shi, D.H., Wang, Y.R., Li, E.G. and Wu, J.H.

(2012): Pro-apoptotic effects of tectorigenin on human hepatocellular

carcinoma HepG2 cells. World J Gastroenterol, 18: 1753-1764.

Jiang, J., Cole, D., Westwood, N., Macpherson. L., Farzaneh, F., Mufti,

G., Tavassoli, M. and Gäken J. (2010): Crucial roles for protein kinase C

isoforms in tumor- specific killing by apoptin. Cancer Res, 70: 7242-7252.

Jolly, D. (1994): Viral vector systems for gene therapy. Cancer Gene Ther.,

1: 51–64.

Kataoka, S. and Tsuruo, T. (1996): Physician Education: Apoptosis.

Oncologist, 1: 399-401.

Kawanishi, S. and Hiraku, Y. (2004): Amplification of anti-cancer drug

induced DNA damage and apoptosis by DNA binding compounds. Curr,

Med. Chem. Antican. Agent, 4: 415-419.

Kekre N Griffin, C., McNulty, J. and Pandey, S. (2005): Pancratistatin

causes early activation of caspase-3 and the flipping of phosphatidyl

serine followed by rapid apoptosis specifically in human lymphoma

cells. Cancer Chemother Pharmacol, 56: 29-38.

Kerr, J (2000): Apoptosis. Med. J. Aust, 173: 616–617.

103

Kerr, J.F., Wyllie, A.H. and Currie, A.R. (1972): Apoptosis: a basic

biological phenomenon with wide ranging implications in tissue kinetics.

Br. J. Cancer, 26: 239-257.

Kim, M.K., Jung, H.S., Yoon, C.S., Ko, J.H., Jun, H.J., Kim, T.K., Kwon,

M.J., Lee, S.H., Ko, K.S., Rhee, B.D. and Park, J.H. (2010): The effect of

glucose fluctuation on apoptosis and function of INS-1 pancreatic beta

cells. Korean Diabetes J, 34: 47-54.

Kooistra, K., Zhang, Y.H. and Noteborn, M.H. (2007): Viral elements

sense tumorigenic processes: approaching selective cancer therapy. Mini

Rev Med Chem, 11: 1155-1165.

Koopman, G., Reutelingsperger, C.P., Kuijten, G.A., Keehnen, R, M.,

Pals, S.T. and Van Oers M.H. (1994): Annexin V for flow cytometric

detection of phosphatidyl expression on B cells undergoing apoptosis.

Blood, 84: 1415-1420.

Kucharski, T.J., Gamache, I., Gjoerup, O. and Teodoro, J.G. (2011):

DNA damage response signaling triggers nuclear localization of the

chicken anemia virus protein apoptin. J Virol, 85: 12638-12649.

Kumar, M., Dwivedi, P., Sharma, A.K., Sankar, M., Patil, R.D.

and Singh, N.D. (2012): Apoptosis and lipid peroxidation in ochratoxin

A- and citrinin-induced nephrotoxicity in rabbits. Toxicol Ind Health,

[Epub ahead of print]

Lauber, K., B lumenthal, S.G., Waibel, M. and Wesselborg, S. (2004):

Clearance of apoptotic cells: getting rid of corpses. Mol. Cell, 14: 277-287.

104

Leasauskaite V and Ivanoviene L, (2002): Programmed cell death:

molecular mechanism and detection. Medicina, 38: 869-875.

Lee, J.J., Chen, P.B., Yang, S.H., Cheng, C.H., Chueh, L.L., Pang, V.F.,

Hsiao, M. andLin, C.T. (2007): Effect of the VP3 gene of chicken anemia

virus on canine mammary tumor cells. Am J Vet Res, 68: 411-422.

Lee, M.S., Sun, F.C., Huang, C.H., Lien, Y.Y., Feng, S.H., Lai, G.H., Lee,

M.S., Chao, J., Chen, H.J., Tzen, J.T.C. and Cheng, H.Y. (2012): Efficient

production of an engineered apoptin from chicken anemia virus in a

recombinant E.coli for tumor therapeutic applications. BMC Biotech, 12:

doi: 10. 1186/1472-6750-12-27.

Li, R., Cai, L., Xie, X.F., Peng, L., Wu, T.N. and Li, J. (2012): 7,3'-

dimethoxy hesperetin inhibits inflammation by inducing

synovial apoptosis in rats with adjuvant-induced arthritis.

Immunopharmacol Immunotoxicol. [Epub ahead of print]

Li, X., Jin, N., Mi, Z., Lian, H., Sun, L., Li, X. and Zheng, H. (2006):

Antitumor effects of a recombinant fowlpox virus expressing apoptin in

vivo and in vitro. Int J Cancer, 119: 2948-2957.

Li, X., Liu, Y., Wen, Z., Li, C., Lu, H., Tian, M., Jin, K., Sun, L., Gao, P.,

Yang, E., Xu, X., Kan, S., Wang, Z., Wang, Y. and Jin, N. (2010): Potent

anti-tumor effects of a dual specific oncolytic adenovirus expressing

apoptin in vitro and in vivo. Mol Cancer, 20: 9-10.

105

Li, Y.P. and Cui, Z.Z (2007): Pathogenicity and genomic sequence

comparison of a chicken infectious anemia virus field isolate. Wei Sheng

Wu Xue Bao, 47: 894-898.

Liljestrom, P. (1994): Alphavirus expression systems. Curr. Opin.

Biotechnol., 5: 495–500.

Liljestrom, P. and Garoff, H. (1991): A new generation of animal cell

expression vectors based on the Semliki Forest virus replicon.

Bio/Technology, 9: 1356–1361.

Lin, S., Liu, K., Li, W., Zhou, X. and Huang, T. (2010): Maternal diabetes

increases apoptosis in mice oocytes, not 2-cell embryos. Endocrine, 37:

460-466.

Liu, Q., Fu, H., Xing, R., Tie, Y., Zhu, J., Sun, Z., and Zheng, X. (2008):

Survivin knockdown combined with apoptin overexpression inhibits

cells growth significantly. Cancer Biol. Ther, 7: 1061-1062.

Liu, Q., Luo, Y.S., Geng, L., Wang, W., Lai, Z.and Li, F. (2010):

Therapeutic effect of adenovirus-mediated apoptin gene transfer

combined with ADM and CDDP on hepatocellular carcinoma in mice. J.

of Southern Medical Univ, 30: 538-540.

Liu, X., Zaiden, Y. H., Elojeimy, S., Holman, D. H., El-Zawahry, A. M.,

Guo, G., Bielawska, A., Bielawski, J., Szule, Z., Rubinchik, S., Dong, J.,

Keane, T.E., Tavassoli, M., Hannum, Y.A. and Norris, J.S. (2006):

Involvement of sphingolipids in apoptin induced cell killing. Mol. Ther,

14: 627-636.

106

Lockshin, R.A. and Zakeri, Z. (2002): Caspase-independent cell death.

Curr Opin Cell Biol, 14: 727-733.

Los, M., Panigrahi, S., Rashedi, I., Mandal, S., Stetefeld, J., Essmann, F.,

and Schulze–Osthoff, K. (2009): Apoptin, a tumor–selective killer.

Biochem, Biophys. Acta, 8: 1335-1342.

Maddika, S., Mendoza, F.J., Hauff, K., Zamzow, C.R., Paranjothy, T.

and Los, M. (2006): Cancer-selective therapy of the future: apoptin and

its mechanism of action. Cancer Biol Ther, 5: 10-19.

Martin, J.S., Reutelingsperger, C.P.M., McGahon, A.J., Rader, J.A., Van

Schie, R.C.A., Laface, D.M. and Green, D.R. (1995): Early redistribution

of plasma membrane phosphatidylserine is a general feature of apoptosis

regardless of the initiating stimulus: inhibition by overexpression of Bcl-2

and Abl. J Exp Med, 182: 1545–1556.

Mazars, A., Geneste, O. and Hickman, J. (2005): The Bcl-2 family of

proteins as drug targets. J Soc Biol, 199: 253-265.

McNulty, M.S. (1991): Chicken anaemia agent: a review. Avian Pathol, 20:

187-203.

Miller, A.D., Miller, D.G., Garcia, J.V. and Lynch, C.M. (1993): Use of

retroviral vectors for gene transfer and expression. Methods Enzymol., 217:

581–599.

Mohamed, M.A. (2010): Chicken Infectious Anemia Status in

Commercial Broiler Chickens Flocks in Assiut-upper Egypt: Occurrence,

107

Molecular Analysis Using PCR-RFLP and Apoptosis Effect on Affected

Tissues. International J Poultry Science, 9: 591-598.

Natesan, S., Kataria, J.M., Dhama, K., Bhardwaj, N. and Sylvester, A.

(2006): Anti-neoplastic effect of chicken anemia virus VP3 protein

(apoptin) in Rous sarcoma virus-induced tumours in chicken. J Gen Virol,

10: 2933-2940.

Natesan, S., Kataria, J.M., Dhama, K., Rahul, S. and Baradhwaj, N.

(2006): Biological and molecular characterization of chicken anaemia

virus isolates of Indian origin. Virus Research, 118: 78–86.

Nogueira-Dantas, E.O., Ferreira, A.J., Astolfi-Ferreira, C.S. and

Brentano, L. (2007): Cloning and expression of chicken anemia virus VP3

protein in Escherichia coli. Comp Immunol Microbiol Infect Dis, 30: 133-142.

Noteborn M.H. (2004): Chicken anemia virus induced apoptosis:

underlying molecular mechanisms. Vet. Microbiol, 98: 89-94.

Noteborn, M.H., Todd, D., Verschueren, C.A., de Gauw, H.W., Curran,

W.L., Veldkamp, S., Douglas, A.J., McNulty, M.S., Van der EB, A.J. and

Koch G. (1994): A single chicken anemia virus protein induces apoptosis.

J Virol, 68: 346-351.

Noteborn, M.H., de Boer, G.F., van Roozelaar, D.J., Karreman, C.,

Kranenburg, O., Vos, J.G., Jeurissen, S.H., Hoeben, R.C., Zantema,

A. and Koch, G. (1991): Characterization of cloned chicken anemia virus

DNA that contains all elements for the infectious replication cycle. J.

Virol, 65: 3131 – 3139.

108

Noteborn, M.H.M. and Koch, G (1995): Chicken anaemia virus infection:

Molecular basis of pathogenicity, Avian Pathology, 24: 11-31.

Noteborn, M.H.M., Danen-van Oorschot, A.A.A.M. and Van der Eb,

A.J. (1998): Chicken anemia virus: Induction of apoptosis by a single

protein of a single-stranded DNA virus. Seminars in Virology, 8: 497-504.

O'Brien, V. (1998): Viruses and apoptosis. Gen. Virol, 79: 1833-1845.

Oorschot, A.A.A.M., Fisher, D.F., Gimbergen, J.M., Klien, B., Zuang,

S.M., Falkenburg, J.H.F., Backendorf, C., Quax, P.H.A., Van Der Eb,

A.J., and Noteborn M.H.M. (1997): Apoptin induces apoptosis in human

transformed and malignant cells but not in normal cells. Leiden

University, 94: 5843-5847.

Oro, C. and Jans, D.A. (2004): The tumour specific proapoptotic factor

apoptin Vp3 from chicken anaemia virus. Curr Drug Targets, 5: 179-190.

Pan, Y., Fang, L., Fan, H., Luo, R., Zhao, Q., Chen, H. and Xiao, S. (2010):

Antitumor effects of a recombinant pseudotype baculovirus expressing

apoptin in vitro and in vivo. Int J Cancer, 126: 2741-2751.

Panigrahi, S., Stetefeld, J., Jangamreddy, J.R., Mandal, S., Mandal, S.K.

and Los, M. (2012): Modeling of molecular interaction between apoptin,

BCR-Abl and CrkL--an alternative approach to conventional rational

drug design. PLoS One, 7: 28395.

Patel, R.M. and Prajapati, R.N. (2011): Pancratistatin, an apoptic inducer:

New horizon for targeted therapy in cancer. Biotechnol Mol Biol Rev, 6: 59-

68.

109

Pavet, V., Portal, M.M., Moulin, J.C., Herbrecht, R. and Gronemeyer, H.

(2011): Towards novel paradigms for cancer therapy. Oncogene, 30: 1-20.

Peng, D.J., Sun, J., Wang, Y.Z., Tian, J., Zhang, Y.H., Noteborn, M.H.

and Qu, S. (2007): Inhibition of hepato- carcinoma by systemic delivery

of Apoptin gene via the hepatic asialoglycoprotein receptor. Cancer Gene

Ther, 14: 66-73.

Pietersen, A. and Noteborn, H.M (2000): Apoptin. Adv Exp Med Biol, 465:

153-161.

Pietersen, A.M., van der Eb, M.M., Rademaker, H.J., van den

Wollenberg, D.J., Rabelink, M.J., Kuppen, P.J., van Dierendonck, J.H.,

van Ormondt, H., Masman, D., Van de Velde, C.J., Van der Eb, A.J.,

Hoeben, R.C. and Noteborn, M.H. (1999): Specific tumor-cell killing

with adenovirus vectors containing the apoptin gene. Gene Ther, 6: 882-

892.

Podo, F., Canevari, S., Canese, R., Pisanu, M.E., Ricci, A. and Iorio, E.

(2011): MR evaluation of response to targeted treatment in cancer cells.

NMR Biomed, 24: 648-672.

Pope, C.R. (1991): Chicken anemia agent. Vet. Immunopathol, 30: 51-65.

Reed, J.C. (2004): Apoptosis mechanisms: implications for cancer drng

discovery. Oncology, 18: 11-20.

Rodriguez, R.L. and Tait, R.C. (1983): Recombinant DNA techniques. An

introduction. Addison-Wesley Publ. Co. London.

110

Rohn, J.L. and Noteborn, M.H. (2004): The viral death effector apoptin

reveals tumor- specific processes. Apoptosis, 9: 315-322.

Sambrook, J. and Russell, D.W. (2001): Molecular cloning. A Laboratory

Manual, 3rd ed., Cold Spring Harbor Laboratory Press, edd Spring

Harbor, New York.

Samulski, R.J., Chang, L.S. and Shenk, T. (1989): Helper-free stocks of

recombinant adeno-associated viruses: normal integration does not

require viral gene expression. J. Virol., 63: 3822–3828.

Sauvage, V., Cheval, J., Foulongne, V., Gouilh, M.A., Pariente, K.,

Manuguerra, J.C., Richardson, J., Dereure, O., Lecuit, M., Burguiere, A.,

Caro, V. and Eloit, M. (2011): Identification of the first human gyrovirus,

a virus related to chicken anemia virus. J Virol, 85: 7948-7950.

Saxena, S., Kumar, G.R., Singh, P., Chaturvedi, U., Saxena, L., Kumar,

R., Sahoo, A.P., Doley, J., Rajmani, Kumar, A., Kumar, S. and Tiwari,

A. K. (2012): Prokaryotic expression of chicken infectious anaemia

apoptin protein and characterization of its polyclonal antibodies. Indian J

Exp Biol, 50: 325-331.

Schlesinger, S. (1993): Alphaviruses—vectors for the expression of

heterologous genes. Trends Biotechnol., 11: 18–22.

Schmitz, I., Kirchoff, S. and Krammer, P.H. (2000): Regulation of death

receptor mediator apoptosis pathways. Int. J. Biochem. Cell. Biol, 32: 1123-

1136.

111

Schoop, R.A, Baatenburg de Jong, R.J. and Noteborn, M.H (2008):

Apoptin induces apoptosis in an oral cancer mouse model. Cancer Biol

Ther, 7: 1368-1373.

Schoop, R.A., Verdegaal, E.M., de Jong, R.J. and Noteborn, M.H. (2010):

Apoptin enhances radiation-induced cell death in poorly responding

head and neck squamous cell carcinoma cells. Basic Clin Pharmacol

Toxicol, 106: 130-134.

Shen, Z., Wang, Y., Zong, Y. and Qu, S. (2003): Experimental study on

the antitumor effect of chicken anemia virus vp3 gene against liver

carcinoma in vivo. J Huazhong Univ Sci Technolog Med Sci, 23: 105-107, 115

Sinha, S., Malonia, F.K., Mittal, S.P., Mathai, J., Pal, J.K. and

Chattopadhyay, S. (2012): Chromatin remodelling protein SMAR1

inhibits p53 independent transactivation by regulating acetyl transferase

p300. Int. J Biochem Cell Biol, 44: 46-52.

Sommer, F. and Cardona, C. (2003): Chicken anemia virus in broilers:

dynamics of the infection in two commercial broiler flocks. Avian Dis, 47:

1466–1473.

Steigerova, J., Rarova, L., Oklesťkova, J., Krizova, K., Levkova,

M., Svachova, M., Kolar, Z. and Strnad, M. (2012): Mechanisms of

natural brassinosteroid-induced apoptosis of prostate cancer cells. Food

Chem Toxicol, [Epub ahead of print]

Strauss, J.H. and Strauss, E.G. (1994): The alphaviruses: gene expression,

replication, and evolution. Microbiol. Rev., 58: 491–562.

112

Sun, J., Wang, Y., Zong, Y. and Qu, S. (2003): Cloning of chicken anemia

virus vp3 gene and apoptosis inductive effect of vp3 gene in vitro. J

Huazhong Univ Sci Technolog Med Sci, 23: 329-331, 334.

Takahashi, K., Iwasa, Y., Hijikata, M. and Mishiro, S. (2000):

Identification of a new human DNA virus (TTV-like mini virus, TLMV)

intermediately related to TT virus and chicken anemia virus. Archives

Virol, 145: 979–993.

Teodoro, J.G. and Branton, P.E. (1997): Regulation of apoptosis by viral

gene products. J. Virol, 71: 1739-1746.

Todd, D. (2000): Circoviruses: immunosuppressive threats to avian

species: a review. Avian Pathol, 29: 373-394.

Todd, D., Creelan, J.L., Mackie, D.P., Rixon, F. and McNulty, M.S.

(1990): Purification and biochemical characterization of chicken anemia

agent. J Gen Virol, 71: 819-823.

Tumane, R.G., Pingle, S.K., Jawade, A.A. and Nath, N.N. (2010): An

overview of caspase: Apoptotic protein for silicosis. Indian J Occup

Environ Med, 14: 31-38.

Van der Eb, M.M., Pietersen, A.M., Speetjens, F.M., Kuppen, P.J., Van

De Velde, C.J., Noteborn, M.H. and Hoeben, R.C. (2002): Gene therapy

with apoptin induces regression of xenografted human hepato- mas.

Cancer Gene Ther, 9: 53-61.

VanEngeland, M., Nieland, L.J.W., Ramaekers, F.C.S., Schutte, B.

and Reutelingsperger, C.P.M. (1998): AnnexinV-Affinity Assay: A

113

review on an apoptosis detection system based on phosphatidyl

serine exposure. Cytometry, 31: 1-9.

VanEngeland, M., Ramaekers, F.C.S., Schutte, B., Reutelingsperger,

C.P.M. (1996): A novel assay to measure loss of plasma membrane

asymmetry during apoptosis of adherent cells in culture. Cytometry,

24: 131–139.

Wagstaff, K.M. and Jans, D.A. (2009): Nuclear drug delivery to target

tumour cells. Eur J Pharmacol, 625: 174-180.

Wang, C. and Zhang, Y. (2011): Apoptin gene transfer via modified

wheat histone H4 facilitates apoptosis of human ovarian cancer cells.

Cancer Biotherapy & Radiopharmaceuticals, 26: 121-126.

Wang, J.S., Zhang, M.X., Duan, X.Y., Mo, F., Wang, Q.Y. and Yang, G.X.

(2009): Construction of a prokaryotic expression vector for apoptin and

preparation of polyclonal antibody of apoptin. J Southern Med Univ, 29:

1405-1407.

Wang, L., Yang, W., Read, P., Lamer, J. and Sheng, K. (2010): Tumor cell

apoptosis induced by nanoparticle conjugate in combination with

radiation therapy. Nanotechnology, 21: 475103.

Wang, Q.M., and He, F.C (2005): Molecular mechanism of specific

induction of apoptosis in tumor cells by apoptin. Ai Zheng, 24: 509-512.

Wesselborg, S. and Lauber, K. (2005): Mechanism of anti-cancer drug

action. In Apoptotic Pathways As Target For Novel Therapies in Cancer and

114

Other Diseases (ed. M. Los and S. B. Gibson), New York: Springer

Academic Press, pp 31-56.

Wysokinski, D., Błasiak, J. and Wozniak, K (2010): Involvement of

nuclear domains in the cellular response to DNA damage. Postepy

Biochem, 56: 328-340.

Xiong, C., Levis, R., Shen, P., Schlesinger, S., Rice, C.M. and Huang,

H.V. (1989): Sindbis virus: an efficient, broad host range vector for gene

expression in animal cells. Science, 243: 1188–1191.

Yan, L., Xiangwei, M., Xiao, L., Peng, G., Chang, L., Mingyao,

T., Encheng, Y., Xiaohong, X., Peng, J., Shifu, K., Zhongmei, W.

and Ningyi, J. (2010): Construction, expression and characterization of a

dual cancer-specific fusion protein targeting carcinoembryonic antigen in

intestinal carcinomas. Protein Expr Purif, 69: 120-125.

Yu, D.S., Zhao, R., Hsu, E.L., Cayer, J., Ye, F., Guo, Y., Shyr, Y. and

Cortez, D (2010): Cyclin-dependent kinase 9-cyclin K functions in the

replication stress response. EMBO Rep, 11: 876-882.

Zhang, Y.H., Leliveld, S.R., Kooistra, K., Molenaar, C., Rohn, J.L.,

Tanke, H.J., Abrahams, J.P. and Noteborn, M.H.M. (2003): Recombinant

Apoptin multimers kill tumor cells but are nontoxic and epitope-shielded

in a normal-cell-specific fashion. Exp Cell Res, 289: 36-46.

Zhuang, S.M., Shvarts, A., van Ormondt, H., Jochemsen, A.G., Van der

Eb, A.J. and Noteborn, M.H. (1995): Apoptin, a protein derived from

115

chicken anemia virus, induces p53-independent apoptosis in human

osteosarcoma cells. Cancer Res, 55: 486-489.

Appendix

116

Appendix

A. PLASMID DNA ISOLATION BY TELT METHOD

TELT Solution

Tris Cl, pH 8.0 50 mM

Na2 EDTA 62.5 mM

LiCl 2.5 M

Triton X-100 4% v/v

Phenol/Chloroform 1:1 v/v

B. REAGENTS USED IN AGAROSE GEL LECTROPHORESIS

1. Tris-acetate-EDTA (TAE) buffer 50X

Tris base 24.2 g

Glacial acetic acid 57.1 ml

0.5 M EDTA (pH 8.0) 100 ml

Distilled water was added to make final volume up to 100 ml. A

working solution of 1X was used.

2. Ethidium bromide stock solution (10 mg/ ml)

Ethidium bromide 100 mg

Distilled water 10 ml

The solution was mixed and stored at 40C. A concentration of

0.5 mg/ ml was used in preparing agarose gel.

117

3. Loading Dye (6X)

Bromophenol blue 0.25% (w/v)

Xylene cyanol 0.25% (w/v)

Sucrose 40% (w/v)

C. REAGENTS USED FOR BACTERIOLOGICAL PROCEDURES

1. LB (Luria-Bertani) broth

Bacto tryptone 10 g

Bacto yeast extract 5 g

NaCl 10 g

Distilled water 950 ml

Adjust the pH to 7.0 with 5N NaOH.

Volume was made up to 1 litre and sterilized by autoclaving at 15

psi for 20 minutes and the solution was stored at 40C.

2. LB (Luria-Bertani) Agar Medium

1.5% Agar in LB medium.

Sterilized by autoclaving at 15 psi for 20 minutes and stored at 40C.

3. SOB Medium

Bacto tryptone 20 g

Bacto yeast extract 5 g

NaCl 0.5 g

Nuclease free water 950 ml

118

Dissolve by shaking. Add 10 ml of a 250 mM solution of KCl.

Adjust the pH to 7.0 with 5N NaOH and make volume up to 1000 ml.

Sterilize by autoclaving. Just before use, add 5 ml of a sterile 2M MgCl2.

4. 1X Transformation & Storage Solution (TSS)

LB broth 95 ml

PEG (8000) (w/v) 10 g

MgSO4 (50 mM final conc.) 1.23 g

DMSO 5 ml

DMSO was filtered through 0.22 µ filter and added after

autoclaving of other ingredients at 1210C, 15 psi for 15 minutes.

5. Ampicillin (50 mg/ ml)

Ampicillin 250 mg

Distilled water 5 ml

The mixture was sterilized by filtration through filter and stored at

200C.

D. REAGENTS USED IN CELL CULTURE

1.) PBS

Sodium Chloride (NaCl) 8.00 g

Potassium Chloride (KCl) 0.20 g

Dissodium hydrdrogen phosphate

(Na2HPO42H2O) 1.15 g

119

Distilled water (up to) 1000 ml

The solution was sterilized at 15 psi for 20 minutes and stored at 200C.

2.) DMEM (Gibco)

DMEM powder 13.5 g

with high glucose

With L-glutamine 300 mg

Fungizone 1 ml

Gentamycin 1.2 ml

Sodium bicarbonate 1.5 g

Distilled water (up to) 1000 ml

Distilled water was added to make final volume up to 1000 ml.

Sterilized by filtration.

3. Growth Medium

DMEM 90 ml

FCS 10 ml

4. Trypsin Versene Solution (0.17%)

NaCl 10.0 g

KCl 0.250 g

Na2HPO4 1.9 g

KH2PO4 0.250 g

Trypsin 1.7 g

Versene 1.4 g

0.4% Phenol red 1.0 µl

120

Distilled water was added to make up to 1000 ml

E. REAGENTS USED IN DNA FRAGMENTATION ASSAY

1. Cell lysis buffer

Tris (ph 8.0) 10 mM

EDTA (ph 8.0) 1 mM

SDS 1%

2. Potassium acetate solution

5 M Potassium acetate 60 ml

Glacial acetic acid 11.5 ml

H2O 28.5 ml

This solution is stored at room temperature.

1

Isolation of recombinant plasmid containing Chicken Anemia

Virus VP3 gene.

AUTHORS:

PRIYANKA PAL: Near old chandmari tiwari temple, Street no. 4,

Damodar puram, Subhash nagar, Bareilly. (U.P.)

E-mail: [email protected]. Contact No:

9411217514

PROF. KUSUM AGARWAL: E-mail: [email protected]. Contact No.:

9456089155

PROF. ANANT RAI: E-mail: [email protected]. Contact No.: 9219661948

USHA TIWARI: E-mail: [email protected]. Contact No.:

9917632096

RAJVEER MAURYA: Contact No.: 9457257179

SUBMITTED BY

PRIYANKA PAL

RESEARCH SCHOLAR, BIOTECHNOLOGY

SHOBHIT UNIVERSITY, MEERUT, (U. P.)

2

Isolation of recombinant plasmid containing Chicken Anemia

Virus VP3 gene.

ABSTRACT

Programmed cell death plays critical roles in a wide variety of physiological

processes during fetal development and in adult tissues. Defects in apoptotic cell

death regulation contribute to many diseases, including disorders where cell

accumulation occurs (cancer, restenosis). Knowledge of the molecular mechanisms of

apoptosis is providing insights into the causes of multiple pathologies where aberrant

cell death regulation occurs and is beginning to provide new approaches to the

treatment of human diseases Reed [1]. Apoptin, a small protein from chicken anemia

virus, has attracted great attention, because it specifically kills tumor cells while

leaving normal cells unharmed. In normal cells, apoptin resides in the cytoplasm,

whereas in cancerous cells it translocates into the nucleus Los [2]. For isolation of

plasmid DNA, the procedure described here (originally presented by He [3] for small

isolation of plasmid DNA can also be readily extended for large scale preparation.

Plasmid DNA is obtained from E.coli grown on plates as colonies or in liquid

cultures. The merit of the approach is that it is extremely reliable and rapid- requiring

no more than 20 min of simple and economical operations for a preparation. The final

plasmid DNA preparations are of a purity and quality usable for most biological

application. The yield should be about 30 µg plasmid DNA per preparation. It only

need one (a single) tube per sample!

KEYWORDS

Chicken Anemia Virus, VP3 gene, Apoptin, Apoptosis, Cancer therapy.

3

INTRODUCTION

All human cells have a genetic program that upon activation will cause cell

death, named apoptosis. Cancer cells can grow due to unbalances in proliferation, cell

cycle regulation and their apoptosis machinery: genomic mutations resulting in non-

functional pro-apoptosis proteins or over-expression of anti-apoptosis proteins form

the basis of tumor formation. Surprisingly, lessons learned from viruses show that

cancer cannot be regarded simply as the opposite of apoptosis. For instance,

adenovirus can only transform cells when both its anti- and pro-apoptotic proteins are

produced. Oncolytic viruses are known to replicate selectively in tumor cells resulting

in cell death. Proteins derived from viruses, i.e. chicken anemia virus (CAV)-derived

apoptosis-inducing protein (apoptin), adenovirus early region 4 open reading frame

(E4orf4) and parvovirus-H1 derived non-structural protein 1 (NS1), the human alpha-

lactalbumin made lethal to tumor cells (HAMLET), which is present in human milk or

the human cytokines melanoma differentiation-associated gene-7 (mda-7) and tumor

necrosis factor-related apoptosis-inducing ligand (TRAIL) have all the ability to

induce tumor-selective apoptosis. The tumor-selective apoptosis-inducing proteins

seem to interact with transforming survival processes, which can become redirected

by these proteins into cell death. Transformation-related processes have been

identified, which seem to be crucial for the tumor-selectively killing activity of these

proteins. For instance, the transformation-related protein phosphatase 2A (PP2A)

plays a role in the induction of tumor-selective apoptosis. The proteins mda-7, TRAIL

and HAMLET are already successfully tested in first clinical trials. Proteins harboring

tumor-selective apoptosis characteristics represent, therefore, a therapeutic potential

and a tool for unraveling tumor-related processes. Fundamental molecular and

(pre)clinical therapeutic studies of the various tumor-selective apoptosis-inducing

proteins apoptin, E4orf4, HAMLET, mda-7, NS1, TRAIL and related proteins will be

discussed Noteborn [4].

Caspases belong to a class of cysteine proteases which function as critical

effectors in cellular processes such as apoptosis and inflammation by cleaving

substrates immediately after unique tetrapeptide sites. With hundreds of reported

substrates and many more expected to be discovered, the elucidation of the caspase

4

degradome will be an important milestone in the study of these proteases in human

health and disease. Several computational methods for predicting caspase cleavage

sites have been developed recently for identifying potential substrates. However, as

most of these methods are based primarily on the detection of the tetrapeptide

cleavage sites - a factor necessary but not sufficient for predicting in vivo substrate

cleavage - prediction outcomes will inevitably include many false positives. In this

paper, its shown that structural factors such as the presence of disorder and solvent

exposure in the vicinity of the cleavage site are important and can be used to enhance

results from cleavage site prediction. The multi-factor model augments existing

methods and complements experimental efforts to define the caspase degradome on

the systems-wide basis Wee [5]. Specificity is a prerequisite for systemic gene

therapy of hepatocarcinoma. In vitro, the tumor-specific viral death effector Apoptin

selectively induces apoptosis in malignant hepatic cells. Intratumoral treatment of

xenografted subcutaneous hepatomas with Apoptin results in tumor regression. Here,

a systemic delivery vehicle containing the Apoptin gene linked to asialoglycoprotein

(Asor), which targets asialoglycoprotein receptor (ASGPR) present only on the

surface of hepatocytes. In vitro, the protein-DNA complex Asor-Apoptin induced

apoptosis in HepG2 hepatocarcinoma cells but not in normal L-02 hepatocytes. Non-

hepatocyte-derived tumorigenic human A549 cells lacking the membrane ASGPR

were not affected by Asor-Apoptin. In vivo systemic delivery of Asor-Apoptin via the

tail vein into mice bearing in situ hepatocarcinoma resulted in specific and efficient

distribution of Apoptin in both hepatocarcinoma cells and normal hepatocytes. Five

days after injection of Asor-Apoptin, the in situ hepatocarcinomas showed significant

signs of regression, whereas the surrounding normal hepatocytes did not. Systemically

delivered Asor-LacZ expressing non-apoptotic LacZ gene did not inhibit tumor

growth. Our data reveal that systemic delivery of Asor-Apoptin specifically induces

apoptosis in malignant hepatocytes and thus constitutes a powerful and safe

therapeutics against hepatocarcinomas Peng [6]. The initiation and progression of

tumor is regulated by multiple genes. Survivin belongs to the inhibitor of apoptosis

protein (IAP) family and is overexpressed in most types of human tumors. Apoptin,

originally identified from chicken anemia virus (CAV), can specifically induce

apoptosis of human tumor cells rather than normal cells. In this study, survivin

5

expression was silenced by microRNA (miRNA)-mediated RNA interference (RNAi);

meanwhile, the engineered miRNA vector was also designed to express apoptin gene.

The apoptosis and cell growth were then examined by flow cytometry and MTT

assay. The miRNA-mediated knockdown of survivin in combination with apoptin

overexpression significantly induced apoptosis and inhibited cell growth. Importantly,

the combined strategy was more effective on inducing apoptosis and inhibiting cell

growth than either survivin downregulation or apoptin overexpression alone. Taken

together, the combined strategy offers potential advantages in control of

tumorigenesis, and thus deserves further research as a preferred approach in cancer

gene therapy Liu [7].

MATERIALS AND METHODS

MATERIALS

Host Bacterial strains

Escherichia coli (E.coli) DH5α (Proteges, Madison) host strain was used for isolation

of recombinant plasmid.

Chemicals, glasswares and plastic wares

All chemicals used in this study were either Analar or molecular biology grade from

Sigma (MO), Promega (Madison). Plasticwares and other consumables were from

Axygen, TRP, Nunc, Gneiner, Tarson, Corning or Borosil.

METHODS

Revival of the E.coli culture containing recombinant plasmid with

VP3 gene

Growth of rE.coli cells:

Preparation of LB broth :

2.5gm LB broth in 100ml d/w, sterilized by autoclaving at 15 lbs psi (1210C) for 15

min.

Preparation of LB Agar:

4.0 gm LB Agar powder in 100 ml d/w sterilized by autoclaving as above.

LB Agar Plate: Prepared LB agar plate by dispersing LB agar, cooled to 420C and

added ampicillin to provide concentration of 100µg/ml. Add given rE.coli culture by

streaking method.

6

For preparing E.coli culture in LB broth containing 100µg/ml ampicillin, inoculated

200µl culture in 100 ml broth and kept it in incubator at room temperature (370C) for

24 hours.

Glass test tubes with plastic caps are suitable. Place the tubes at a suitably inclined

angle to achieve good agitation.

Isolation of plasmid DNA containing VP3 gene

Isolation of plasmid DNA miniprep by TELT method-

TELT Solution

50 mM Tris.Cl, pH 8.0

62.5 mM Na2EDTA

2.5 M LiCl

4% (v/v) Triton X-100

1:1 (v/v) phenol/chloroform

Note: All steps are performed at room temperature.

Procedure:

Grown a 1.5 ml E.coli culture with plasmid of interest in LB medium containing

100 mg/ml ampicillin for 16h. Pellet cells 30 sec in micro centrifuge. Resuspend in

100µl TELT solution and add an equal volume of 1:1 phenol/chloroform. Vortex

vigorously 15 sec and spin 1 min in microcentrifuge at 220C. Collect the upper phase

of nucleic acid and mix with 2 vol of 100% ethanol. After 2 minutes, spin 10 minutes

in microcentrifuge at 40C. Wash pellet with 1ml of 70% ethanol, dry under vacuum,

and resuspend in 30µl TE buffer. The pellet may be located as a smeer on one side of

the tube. Dissolve the DNA on a shaker.

Plasmid DNA was isolated and run on agarose gel electrophoresis which yielded

good amount of DNA. Contaminating RNA may interfere with detection of DNA

fragments on the agarose gel; it can be destroyed by adding 1 µl of a 10 mg/ml RNase

soloution (DNase-free) to the digestion mixture.

Agarose gel electrophoresis:

Prepared loading Buffer: TAE (50 X for 100 ml).

Tris Base 24.2 gm

Glatial Acetic Acid 5.71 gm

0.5 M EDTA (pH 8.0) 10 ml

7

Distilled water was added to make the final volume up to 100 ml. A working solution

of 1X was used.

Preparation of loading dye: Bromophenol blue 25% and Sucrose 40% in d/w and mix

it.

Procedure:

Prepared 1% agarose in 1X TAE buffer and heated till it dissolved. Cooled the

solution to about 600C and then added ethidium bromide at a concentration of 0.5

µg/ml. Gel tray was sealed with tape on its open ends and placed the comb on the tray.

Then pour the above gel on the sealed tray. Leave it to solidify. Remove the comb and

tape. The gel tray was put in electrophoretic tank containing 1X TAE buffer. The

DNA marker and sample DNA were loaded into the slots of submerged gel, using

bromophenol dye buffer and run at for about 70-80 volts.

Further inoculated the culture for maintenance.

Preservation of the culture:

It is used for back up and for research work.

It is preserved in glycerol:

Glycerol 10 ml

Autoclave it.

Mixed 200µl glycerol with 800µl of culture and stored at -200C.

RESULTS

The plates containing recombinant E.coli culture were found to contain colonies,

from which ten discrete colonies were selected randomly and grown in LB broth

containing ampicillin and the recombinant pTARGET CAV-VP3 plasmids were

isolated. The DNA Bands were visualized under UV light and photographed.

DISCUSSION

Chicken anemia virus (CAV), etiological agent of a highly immunosuppressive

disease of yong chicken, i.e. chicken infectious anemia (CIV), is economically

important avian pathogen worldwide. It belongs to Gyrovirus genus of the

Circoviridae family having circular single stranded DNA genome of 2.3 kb in size.

The disease is characterized by severe anemia, aplasia of bone marrow and

generalized lymphoid atropy with concomitant immunosuppression. Among the viral

proteins of Chicken anemia virus (CAV), VP3 has the death inducing abilities and so

8

it was renamed as apoptin. Apoptin can induce apoptois in cell lines derived from a

great variety of human tumors.

REFERENCES

Reed, J.C. (2000). Mechanisms of apoptosis, Am. J. Pathol. 157, 1415-1430.

Los, M., Panigrahi, S., Rashedi, I., Mandal, S., Stetefeld, J., Essmann, F., and

Schulze-Osthoff, K. (2009). Apoptin, a tumor-selective killer, Biochem.

Biophys. Acta. 8, 1335-42.

He, M., Wilde, A. and Kaderbhai, M.A. (1990). A simple single step procedure for

small scale preparation of Escherichia Coli plasmid, Nucleic Acids Res., 18,

1660.

Noteborn, H. (2009). Proteins selectively killing tumor cells, Eur. J. Pharmacol. 625,

165-73.

Wee, L.J., Tong, J.C., Tan, T.W. and Ranganathan, S. (2009). A multi-factor model

for caspase degradome prediction, BMC Genomics. 10, Supplements 3.

Peng, D.J., Sun, J., Wan, Y.Z., Tian, J., Zhang, Y.H., Noteborn, M.H. and Qu, S.

(2007). Inhibition of hepatocarcinoma by systemic delivery of Apoptin gene

via the hepatic asialoglycoprotein receptor, Cancer Gene Therapy. 14, 66-73.

Liu, Q., Fu, H., Xing, R., Tie, Y., Zhu, J., Sun, Z., and Zheng, X. (2008). Survivin

knockdown combined with apoptin overexpression inhibits cells growth

significantly, Cancer Biol. Ther. 7, 1053-60.