sequence analysis, pathogenicity and cytokine gene

184
Sequence analysis, pathogenicity and cytokine gene expression patterns associated with fowl adenovirus infection by Helena Grgić A Thesis presented to The University of Guelph In partial fulfillment of requirements for the degree of Doctor of Philosophy in Pathobiology Guelph, Ontario, Canada © Helena Grgić, May, 2012

Upload: others

Post on 05-Nov-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Sequence analysis, pathogenicity and cytokine gene

Sequence analysis, pathogenicity and cytokine gene expression patterns

associated with fowl adenovirus infection

by

Helena Grgić

A Thesis presented to

The University of Guelph

In partial fulfillment of requirements for the degree of

Doctor of Philosophy in

Pathobiology

Guelph, Ontario, Canada © Helena Grgić, May, 2012

Page 2: Sequence analysis, pathogenicity and cytokine gene

ABSTRACT

SEQUENCE ANALYSIS, PATHOGENICITY AND CYTOKINE GENE

EXPRESSION PATTERNS ASSOCIATED WITH FOWL ADENOVIRUS

INFECTION

Helena Grgić Advisor: University of Guelph, 2012 Dr. Éva Nagy

The family Adenoviridae consists of five genera, including the genus

Aviadenovirus, which infects avian species. The genus Aviadenovirus currently

comprises five fowl (Fowl adenovirus A-E), one falcon (Falcon adenovirus A), and one

goose (Goose adenovirus) adenovirus species. Fowl adenoviruses (FAdVs) have a

worldwide distribution. Some are associated with diseases such as inclusion body

hepatitis (IBH), while FAdV species C serotype 4 (FAdV-4) has been associated with

hydropericardium-hepatitis syndrome (HHS).

In this study, the complete nucleotide sequence of fowl adenovirus serotype 8

(FAdV-8) was determined. The full genome was 44,055 nucleotides (nt) in length, with

an organization similar to that of the FAdV-1 and FAdV-9 genomes. No regions

homologous to early regions E1, E3, and E4 of mastadenoviruses were recognized

Page 3: Sequence analysis, pathogenicity and cytokine gene

Pathogenicity of FAdV-8 and FAdV-4 were studied in specific-pathogen-free

chickens following oral and intramuscular inoculations. Pathogenicity was determined on

the basis of clinical signs and gross and histological lesions. Additionally, virus shedding

and viral genome copy numbers in liver, cecal tonsil, and bursa of Fabricius were

determined.

The role of interleukins (IL) in the pathogenicity of and immune response to

FAdVs is unknown. Therefore, in a chicken experiment, interferon-γ, IL-10, IL-18, and

IL-8 gene expression was evaluated following FAdV-8 and FAdV-4 infection. Cytokine

gene expression was examined in the liver, spleen, and cecal tonsils. This study explored

the ability of fowl adenoviruses to subvert the host cell’s secretion of cytokines in

response to infection as an important viral mechanism for immune evasion during

infection.

Variations in virulence of FAdVs are likely to be determined by the fiber alone as

shown by Pallister et al. (1996). Therefore, we compared and analyzed the nt and amino

acid (aa) sequences of the fiber gene of pathogenic and non-pathogenic FAdVs

representing species groups D (FAdV-11) and E (FAdV-8). According to our data,

virulence might not be associated only with sequence of the fiber gene.

This work is a continuation of our efforts towards better understanding of the

molecular biology of FAdVs and the pathogenesis of the disease, with an emphasis on the

role of interleukins, an unknown area.

Page 4: Sequence analysis, pathogenicity and cytokine gene

iv

ACKNOWLEDGEMENTS

I wish to express my sincere appreciation and thanks to my advisor Dr. Ѐva

Nagy, for her support and guidance through the past several years and for keeping on

pushing me whenever I hit a bump in my research. I am also very thankful to my

committee members Drs. Annette Nassuth and Sarah Wootton for the many thoughtful

comments, ideas and suggestions that have greatly improved this thesis.

Above all, I would like to acknowledge the invaluable support of Dr. Bruce

Hunter, my mentor and friend, who taught me everything I know about poultry pathology

and histology. Bruce was always approachable, always generous with his time, always

ready to help and encourage. I will miss him greatly for his knowledge and expertise, but

even more for his kind and gentle nature.

I thank to Dr. Shayan Sharif for his constructive criticism and comments on my

cytokine research and for the willingness to let me use his lab.

Many of the current and past Virology lab members have become dear friends

over the years and I will miss all of them, including the rare lab drama. Sheila, Rita, Juan,

Andres, Bryan, Rasha, Li, Betty-Anne, thank for brightening many of my days in lab.

Most of all I would like to thank my family, my husband Zvonimir and our

daughter Korina: I can’t thank you enough for your endless support and patience through

Page 5: Sequence analysis, pathogenicity and cytokine gene

v

this process and for being the best two distractions from work one could wish for. I could

not have done this without you.

Page 6: Sequence analysis, pathogenicity and cytokine gene

vi

TABLE OF CONTENTS

ABSTRACT ....................................................................................................................... ii

ACKNOWLEDGEMENTS ............................................................................................ iv

LIST OF FIGURES ......................................................................................................... ix

LIST OF ABBREVIATIONS .......................................................................................... x

CHAPTER 1 ...................................................................................................................... 1

LITERATURE REVIEW ....................................................................................................... 1 1. Diseases caused by avian adenoviruses .................................................................. 1

1.1. Inclusion body hepatitis ................................................................................... 1 1.1.1. Disease ...................................................................................................... 1 1.1.2. Clinical signs and pathogenesis ................................................................ 1 1.1.3. Gross lesions and histopathology.............................................................. 2 1.1.4. Virus excretion and transmission .............................................................. 4

1.2. Hydropericardium-hepatitis syndrome ............................................................ 5 1.2.1. Disease ...................................................................................................... 5 1.2.2. Clinical signs ............................................................................................. 6 1.2.3. Gross lesion and histopathology ............................................................... 7 1.2.4. Virus excretion and transmission .............................................................. 8

1.3. Other diseases caused by avian adenoviruses .................................................. 8 1.4. Diseases in birds other than poultry ............................................................... 10 1.5. Diagnosis of aviadenovirus infection............................................................. 12 1.6. Control ........................................................................................................... 13

2. Adenoviruses ......................................................................................................... 14 2.1. Taxonomy ...................................................................................................... 14 2.2. Characteristics of adenoviruses ...................................................................... 15

2.2.1. Structure and physicochemical composition of the adenovirus .............. 15 2.2.2. Proteins ................................................................................................... 16 2.2.3. Genome ................................................................................................... 20

2.3. Adenovirus replication cycle ......................................................................... 24 2.3.1. Attachment and penetration .................................................................... 24 2.3.2. Activation of early viral genes ................................................................ 27 2.3.3. DNA replication ...................................................................................... 28 2.3.4. Late gene expression ............................................................................... 29 2.3.5. Assembly and release .............................................................................. 29

3. Avian immune response ........................................................................................ 30 3.1. Innate immune response ................................................................................ 32

3.1.1. Cytokines ................................................................................................ 37 3.2. Adaptive immunity ........................................................................................ 40

HYPOTHESIS AND OBJECTIVES ....................................................................................... 43

Page 7: Sequence analysis, pathogenicity and cytokine gene

vii

CHAPTER 2. Pathogenicity and complete genome sequence of a fowl adenovirus serotype 8 isolate ............................................................................................................. 46

Abstract ............................................................................................................. 47 1. Introduction ............................................................................................... 49 2. Materials and methods .............................................................................. 51 3. Results ....................................................................................................... 54 4. Discussion ................................................................................................. 59

CHAPTER 3. Key cytokines activated during infection of chickens with a serotype 8 fowl adenovirus ............................................................................................................... 71

Abstract ............................................................................................................. 72 1. Introduction .............................................................................................. 74 2. Materials and methods ............................................................................. 76 3. Results ...................................................................................................... 79 4. Discussion ................................................................................................ 81

CHAPTER 4. Pathogenicity and cytokine gene expression patterns of a serotype 4 fowl adenovirus isolate, a vaccine candidate ................................................................ 97

Abstract ............................................................................................................. 98 1. Introduction ............................................................................................ 100 2. Material and methods ............................................................................. 101 3. Results .................................................................................................... 106 4. Discussion .............................................................................................. 109

CHAPTER 5. Molecular characterization of the fiber gene of fowl adenovirus serotypes 8 and 11 ......................................................................................................... 123

Abstract ........................................................................................................... 124

GENERAL DISCUSSION ........................................................................................... 141

References ...................................................................................................................... 148

Page 8: Sequence analysis, pathogenicity and cytokine gene

viii

LIST OF TABLES Table 2. 1. Percent identities/similarities of selected FAdV-8 proteins and their

homologues ............................................................................................................... 65 Table 2. 2. Viral genome copy numbers in tissues of chickens inoculated with FAdV-8

................................................................................................................................... 66 Table 2. 3. Virus titers in the feces of chickens inoculated with FAdV-8 ....................... 67 Supplementary Table 2. 1. Position of 46 predicted genes of the FAdV-8 genome ...... 68 Table 4. 1. Viral genome copy numbers in tissues of chickens inoculated with FAdV-4

................................................................................................................................. 115 Table 4. 2. Virus titers in the feces of chickens inoculated with FAdV-4 ..................... 116 Table 5. 1. Description of the isolates. ........................................................................... 133 Table 5. 2. Percent identities/similarities between FAdV-8 and FAdV-11 fiber proteins

and their homologues. ............................................................................................. 134 Table 5. 3. Differences in amino acids between FAdV-8 isolates. ................................ 135

Page 9: Sequence analysis, pathogenicity and cytokine gene

ix

LIST OF FIGURES

Figure 1. 1. Schematic transection of the adenovirus particle ......................................... 23 Figure 1. 2. Reproductive cycle of human adenovirus type 2. ........................................ 26 Figure 2. 1. Schematic representation of FAdV-8, FAdV-9 and FAdV-1 genomes,

members of the genus Aviadenovirus ....................................................................... 69 Figure 2. 2. Antibody response to viral proteins in chickens inoculated with FAdV-8 by

the oral and intramuscular routes and mock infected as measured by S/P ratios ..... 70

Figure 3. 1. Cytokine mRNA expression in CH-SAH following FAdV-8 infection ....... 89 Figure 3. 2. Cytokine mRNA expression in spleen of chickens infected with FAdV-8. . 91 Figure 3. 3. Cytokine mRNA expression in cecal tonsil of chickens infected with FAdV-

8................................................................................................................................. 93 Figure 3. 4. Cytokine mRNA expression in liver of chickens infected with FAdV-8 ..... 95 Figure 4. 1. Antibody response to viral proteins in chickens inoculated with FAdV-4 by

the oral and intramuscular routes and mock infected as measured by S/P ratios. Panel A. Serum samples were tested against homologous virus, FAdV-4. Panel B. Serum samples were tested against heterologous virus, FAdV-9 ........................... 117

Figure 4. 2. Cytokine mRNA expression in cecal tonsil of chickens infected with FAdV-4............................................................................................................................... 119

Figure 4. 3. Cytokine mRNA expression in liver of chickens infected with FAdV-4 ... 120 Figure 4. 4. Cytokine mRNA expression in spleen of chickens infected with FAdV-4..

................................................................................................................................. 122 Figure 5. 1. Amino acid sequence alignment of the tail sequence of FAdV-8 and FAdV-

11 isolates made by CLUSTALW .......................................................................... 136 Figure 5. 2. Regions of FAdV-8 and FAdV-11 fibers ................................................... 137 Figure 5. 3. Supplementary data. Amino acid alignment of FAdV-8 fibers derived from

flocks with and without signs of IBH ..................................................................... 139 Figure 5. 4. Supplementary data. Amino acid alignment of FAdV-11 fibers derived from

flocks with and without signs of IBH ..................................................................... 140

Page 10: Sequence analysis, pathogenicity and cytokine gene

x

LIST OF ABBREVIATIONS vßs Alphavbeta integrins AAdV Avian adenoviruses AASV Avian adenovirus splenomegaly virus AAV Adeno-associated virus Ab Antibody ADP Adenovirus death protein AGID Agar gel immunodiffusion APC Antigen presenting cells CAdV Canine adenovirus CAR Coxsackie and adenovirus receptor CEK Chicken embryo kidney CEL Chicken embryo liver CELO Chicken embryo lethal orphan CIAV Chicken infectious anemia virus CK Chicken kidney CMV Cytomegalovirus CPE Cytopathic effect CSIF Cytokine-synthesis inhibitory factor DAdV Duck adenovirus DBP DNA-binding protein DC Dendritic cells DNA Deoxyribonucleic acid ds double-stranded d.p.i. days post infection EDS Egg drop syndrome EDSV Egg drop syndrome virus EID Embryonic development ELISA Enzyme linked immunosorbent assay EM Electron microscopy ER Endoplasmic reticulum FaAdV Falcon adenovirus FAdV Fowl adenovirus GALT Gut associated lymphoid tissue GoAdV Goose adenovirus HAdV Human adenovirus HEV Hemorrhagic enteritis virus HHS Hydropericardium-hepatitis syndrome HIV-1 Human immunodeficiency virus 1 h.p.i. hour post infection HPS Hydropericardium syndrome HPV Human papilloma virus hr hour HSP Heat-shock protein

Page 11: Sequence analysis, pathogenicity and cytokine gene

xi

IBDV Infectious bursal disease virus IBH Inclusion body hepatitis IBV Infectious bronchitis virus Ig Immunoglobulin IFN Interferon ITRs Inverted terminal repeats LRR Leucine rich repeat motif MALT Mucosal associated lymphoid tissue MDV Marek’s disease virus ml Milliliter MHC Major histocompatibility complex MMTV Mouse mammary tumor virus MSB-1 Marek’s disease lymphoblastoid cell NF-1 Nuclear factor I NK Natural killer NLRP Nod-like receptor NOD Nucleotide-binding oligomerization domain nt Nucleotide NTR N-terminal region Oct-1 Octamer-binding protein ORF Open reading frame PAMPs Pathogen associated molecular patterns PiAdV Pigeon adenovirus Pol Polymerase PRRs Pattern recognition receptors pTP Pre-terminal protein RE Restriction enzyme RGD Arg-Gly-Asp motif RFLP Restriction fragment length polymorphism RNA Ribonucleic acide ROS Ractive oxygen species RSV Respiratory syncytial virus RT Reverse transcription SDS Sodium dodecyl sulfate SPF Specific pathogen free ss single-stranded TAdV Turkey adenovirus TCID50 50% tissue culture infective dose TLR Toll-like receptors TP Terminal protein TR Tandem repeat wt Wild type

Page 12: Sequence analysis, pathogenicity and cytokine gene

1

CHAPTER 1

Literature Review 1. Diseases caused by avian adenoviruses 1.1. Inclusion body hepatitis 1.1.1. Disease

Inclusion body hepatitis (IBH) was first described by Helmboldt and Frazier

(1963). The syndrome is characterized by sudden death and massive liver necrosis and

the disease is mainly diagnosed in broiler chickens. Birds are usually found dead, but

occasionally are seen in an extremely depressed condition shortly before death. Death

occurs within a few hours after initial signs. Outbreaks are usually observed at two to

three weeks of age and last up to three weeks, with total losses as high as 8%. In the last

two decades, IBH has been reported in broiler flocks in almost every country around the

world (Philippe et al., 2005; Gomis et al., 2006; Goodwin et al., 1993; Young et al.,

1972; Saiffudin and Wilks, 1991; Erny et al., 1991). Although adenoviruses are

associated with numerous clinical conditions, their primary role in disease is still not

understood. Most outbreaks are described in broiler flocks, but outbreaks in pullets and

replacement layers, respectively are reported by Bickford (1974) and Hoffmann et al.

(1975).

1.1.2. Clinical signs and pathogenesis

Mortality is closely related to the pathogenicity of the virus, secondary infection with

other pathogens, and susceptibility of the chickens. In Canada, the most commonly

isolated serotypes are 2, 7, 8, and 11 (Ojkic et al., 2008b). From the available evidence, it

Page 13: Sequence analysis, pathogenicity and cytokine gene

2

seems that most of the strains of fowl adenoviruses (FAdVs) follow the same pattern of

infection. Following initial multiplication, viraemia occurs and results in virus spread to

virtually all organs. Viraemia is detected from 24 hours post inoculation (h.p.i.), with two

peaks, one at day 2 and the second at day 7 post infection (p.i.) (Saifuddin and Wilks,

1991). The virus has been detected in pancreas, bursa of Fabricius, esophagus, trachea,

spleen, proventriculus, bone marrow, thymus, and kidney. The main sites of virus

replication are the respiratory and alimentary systems, including the caeca (Saiffudin and

Wilks, 1991). In chickens, immunosuppression caused by other pathogens can influence

the clinical severity of adenoviral infection (Adair and Fitzgerald, 2008). It is apparent

that impairment of the bursa of Fabricius by infectious bursal disease virus (IBDV)

facilitates adenoviruses in producing IBH (Fadly et al., 1975). The ability of IBDV to

enhance the pathogenesis of IBH leads to the conclusion that field cases of IBH must be

preceded by a stress on the humoral immune response (Fadly et al., 1975). Another

commonly known immunosuppressant is chicken infectious anemia virus (CIAV), proven

to be a potent immunosuppressive agent for very young unprotected chicks and therefore

increasing their susceptibility to secondary infections in general (Adair and Fitzgerald,

2008). Beside viruses, stress is one of the most common causes of immunosuppression in

poultry. It is believed that the negative economic impact of IBH is usually enhanced in

stressed birds (Adair and Fitzgerald, 2008).

1.1.3. Gross lesions and histopathology

At gross examination, there is marked anemia, icterus of the skin and

subcutaneous fat deposits, haemorrhages in various organs, and pale inactive bone

Page 14: Sequence analysis, pathogenicity and cytokine gene

3

marrow. The main lesion is in the liver, which is swollen and light brown to yellow,

occasionally with haemorrhages (Adair and Fitzgerald, 2008).

At histopathological examination, large, round, eosinophilic inclusion bodies with

clear halos are present in hepatocytes (Grimes et al., 1977a; Itakura et al., 1974).

However, some researchers reported that basophilic inclusions may occasionally be seen

(Grimes et al., 1977b; McCracken et al., 1976). Small focal haemorrhages and nephrosis

are reported in the kidney (Howell et al., 1970). Moreover, the first signs of infection can

be seen in hepatocytes between 16 and 20 h.p.i. and consist of a ring or a partial ring of

medium electron density (Sharpless and Jungherr, 1961). Similar structures have been

observed in cells infected with human adenovirus (HAdV) type 12 (Martinez-Palomo et

al., 1967). Early morphological lesions of infected hepatocytes are observed in the

nucleus (nuclear and nucleolar hypertrophy) 8 hr after virus inoculation (Martinez-

Palomo et al., 1967). At 18 h.p.i., irregular patches of dense material in the nucleoplasm

are reported, differentiated into type I and II inclusions depending on their structure.

Observed inclusions increase rapidly in size and virus particles start to appear between 20

and 30 h.p.i. In general, viral particles tend to be located in the vicinity of type II

inclusions. In the following step, the nuclei of infected cells show large round foci (type

III inclusions), which are larger than the other two types of inclusions. Type III inclusions

gradually fill up most of the nucleus. At 30 h.p.i., the proportions of all three types of

inclusions increase and viral particles develop into crystalline formations (Martinez-

Palomo et al., 1967). By 36 h.p.i., viral crystalline formations are commonly seen and a

dense reticulated substance appears, designated as type IV inclusions. Disorganization of

the normal architecture of the nucleus in most cells is observed 40 h.p.i.. During the final

Page 15: Sequence analysis, pathogenicity and cytokine gene

4

stage of infection, the nucleus is enlarged, the nuclear membrane is distorted, and virus

particles are arranged in crystalline formations which increase in size until disruption of

the nucleus. Lysis of infected cells occurs at the 3rd day after viral inoculation (Martinez-

Palomo et al., 1967).

Cytopathological examination of the gastrointestinal tracts of 1-day-old chicks

infected with avian adenovirus strain 93 shows that antigen is present only in the

epithelial cells on the surface of the villi. Approximately 50% of the infected cells are

detected in the proximal third of the villus, while 25% occured in the middle and another

25% in the distal third (Clemmer and Ichinose, 1968). Both columnar and goblet cells are

infected. The most prominent change is hypertrophy of the epithelial cells with

enlargement of the nuclei. The nuclei are oval with margination of chromatin along the

interior surface of the nuclear membrane (Clemmer and Ichinose, 1968).

1.1.4. Virus excretion and transmission

FAdVs can be transmitted horizontally and vertically. Horizontal spread of the

virus occurs mainly via the feces, tracheal and nasal mucosa, and semen. FAdVs are most

commonly transmitted by the oral-fecal route (Adair and Fitzgerald, 2008; Grgic et al.,

2006).Various studies clearly showed that excretion in the feces, and to a lesser extent in

naso-oral secretions, are the major routes for virus spread. Clemmer and Ichinose (1968)

recorded peak fecal titers of 104.7 50% tissue culture infective doses (TCID50) per g at 4

d.p.i. and excretion continued up to 14 d.p.i. Virus isolation from layer replacements is

most successful when birds are 8 to 14 weeks old and then decreases with increasing age

Page 16: Sequence analysis, pathogenicity and cytokine gene

5

until 34 weeks when virus can no longer be isolated (Cowen et al., 1977). FAdVs are

isolated from asymptomatic chickens as well, and the majority of these viruses induce

cytopathic effect (CPE) by the second passage in chicken kidney cell culture (Cowen et

al., 1977). Also, it is not uncommon to isolate multiple serotypes from one bird,

suggesting that there is little cross protection (Cowen et al., 1977). Therefore, it is not

unexpected to have birds that excrete one serotype in spite of high levels of neutralizing

antibody to another serotype (Cowen et al., 1977). The superiority of cloacal swabs over

tracheal swabs for adenovirus isolation is confirmed by Cowen et al. (1977). Vertical

transmission is considered an important factor in the epidemiology of IBH (McFerran and

Smyth, 2000). Although contradictory reports have been published over the years about

the role and model of vertical transmission of FAdVs (Cowen et al., 1978; Reece et al.,

1985), a recent publication confirmed vertical transmission and establishment of latent

infection with FAdVs (Grgic et al., 2006).

1.2. Hydropericardium-hepatitis syndrome 1.2.1. Disease

Fowl adenovirus C serotype 4 plays a major role in the aetiology of a broiler

disease called infectious hydropericardium-hepatitis syndrome (HHS), an emerging

disease of broiler chickens initially reported from Angara Goth near Karachi, Pakistan

(Khawaja et al., 1988). In addition to India and Pakistan, outbreaks have now also been

reported in the Middle East, Japan, Mexico, Peru, Ecuador and Chile (Abe et al., 1998;

Hess et al., 1999; Toro et al., 1999).

Initially it was hypothesized that HHS is caused by toxicity or by a nutritional

disorder. Investigated causative factors were mycotoxins, phycotoxins, and chemicals

Page 17: Sequence analysis, pathogenicity and cytokine gene

6

such as chloride, polychlorinated biphenyls, and chlordane, which have all been

associated with hydropericardium (Jaffery, 1988). However, all attempts to reproduce the

disease with feed samples from infected farms failed. Therefore, an infectious agent was

suggested as the probable cause (Abdul-Aziz and Al-Attar, 1991). Based on successful

isolation of an agent from the liver, heart, and kidneys, and on transmission of disease

after subcutaneous inoculation of infected liver homogenate, the aetiological agent was

tentatively designated as an adenovirus (Khawaja et al., 1988). Subsequent studies

demonstrated typical icosahedral adenovirus particles in the purified liver extracts by

negative-staining electron microscopy. Additionally, basophilic intranuclear inclusion

bodies in the hepatocytes supported the view that the disease is caused by adenovirus

(Mazaheri et al., 1998).

1.2.2. Clinical signs

The disease is prevalent in 3- to 7-week-old broiler chickens. In field outbreaks,

the affected birds may not exhibit signs other than a high mortality, up to 75% (Jaffery,

1988). However, in the terminal stage of the disease, some birds might become dull,

depressed, huddle in corners with ruffled feathers, and exhibit a characteristic posture,

with the chest and beak resting on the ground (Anjum, 1990). The disease is observed in

healthy broilers showing a good growth rate, suggesting that rapid growth rate may be a

factor that predisposes broilers to HHS (Shafique and Shakoori, 1994). Rare outbreaks

are also recorded in broiler breeder flocks (32 weeks old) and commercial layers (17

weeks old), with up to 8% mortality (Anjum, 1990). Many researchers have reported

equal susceptibility of different broiler strains in both field and in experimental cases.

Page 18: Sequence analysis, pathogenicity and cytokine gene

7

The Hubbard strain of broilers is reported as the most susceptible, followed by Indian

River and Lohmann strains (Khan et al., 1995).

1.2.3. Gross lesion and histopathology

The predominant gross lesion is hydropericardium, characterized by the

accumulation of a jelly-like fluid in the pericardial sac. The heart appears misshapen and

flabby. The pericardial sac has a balloon-like appearance and may contain up to 20 ml of

clear straw-colored fluid (Kumar et al., 1997). Other changes are a pale, swollen, and

friable liver with multifocal necrosis, congested lungs, and pale, swollen kidneys

containing urates in the tubules and ureters, and yellowish discoloration and petechial

haemorrhages in pericardial fat (Kumar et al., 1997; Nakamura et al., 1999).

At histopathological examination, the most consistent findings in the liver are

small multifocal areas of necrosis and mononuclear cell infiltration, with basophilic

intranuclear inclusion bodies in hepatocytes surrounded by a clear halo or filling the

entire nucleus (Nakamura et al., 1999). Microscopic changes in the heart are

mononuclear cell infiltration, severe vascular changes and disruption and separation of

muscle bundles and fibers due to oedema (Kumar et al., 1997). The most prominent

changes in the lung are congestion, oedema, and mononuclear cell infiltration with severe

vascular changes and haemorrhagic exudates in the bronchi and alveoli (Kumar et al.,

1997). In the kidneys, massive swelling of the tubular epithelium, necrosis, and extensive

haemorrhages are reported (Kumar et al., 1997). Changes in the bursa of Fabricius,

thymus, and spleen are lymphocytolysis and cyst formation that consequently lead to

Page 19: Sequence analysis, pathogenicity and cytokine gene

8

lymphocyte depletion in the medullae of the follicles in the bursa of Fabricius. Marked

congestion and mild lymphoid cell depletion are characteristic changes in the thymus

(Kumar et al., 1997; Abdul-Aziz and Hasan, 1995). The only differences between HHS

and IBH are mortality rate and the presence of hydropericardium, which is the most

characteristic gross change in HHS. Hydropericardium alone depresses cardiac function

and produces muffling of the heart sounds (Nakamura et al., 2002).

1.2.4. Virus excretion and transmission

The HHS virus is transmitted vertically and horizontally among birds by the oral-

fecal route (Toro et al., 2001; Cowen, 1992)

1.3. Other diseases caused by avian adenoviruses

There are several other important diseases affecting poultry which are caused by

viruses belonging to the genera Siadenovirus and Atadenovirus of the family

Adenoviridae. Viruses in the genus Siadenovirus typically produce disease without

immunosuppressive factors, virus from the genus Atadenovirus cause egg drop syndrome

(Adair and Fitzgerald, 2008).

Hemorrhagic enteritis virus

Turkey adenovirus 3 (TAdV-3), also known as hemorrhagic enteritis virus (HEV)

of turkeys, belongs to the genus Siadenovirus. Virus is transmitted by the oral-fecal route

and infects turkeys at 4 weeks and older. The clinical signs of infection include

Page 20: Sequence analysis, pathogenicity and cytokine gene

9

depression, bloody droppings, and death. Usually the disease persists in an affected flock

7 to 10 days (Itakura and Carlson, 1975).

Splenomegaly in chickens

Avian adenovirus splenomegaly virus (AASV) of chickens belongs to the genus

Siadenovirus and initially was reported in 1979 in the southeastern U.S.A. Splenomegaly

in chickens is a subclinical disease, although pulmonary oedema, congestion, depression,

weakness, and progressive dyspnea are reported by Fitzgerald and Reed (1989).

Egg drop syndrome

Egg drop syndrome virus (EDSV) infects chickens, quail, geese, and ducks, and

belongs to the genus Atadenovirus (Adair and Fitzgerald, 2008). This virus causes high

losses in egg production. One of the first signs of infection is loss of color in pigmented

eggs and this is quickly followed by production of thin-shelled, soft-shelled, or shell-less

eggs. EDSV is transmitted vertically and horizontally through the drinking water

contaminated by fecal material (Cook and Darbyshire, 1980).

Gizzard erosion

Natural outbreaks of gizzard erosions in broilers are characterized by mortality in

young broilers without clinical signs of infection (Manarolla et al., 2009). On gross

examination, distended gizzards with haemorrhagic fluid and multiple black patchy

erosions are found (Abe et al., 2001). Some researchers also reported pancreatitis,

hepatitis, cholecystitis, and cholangitis (Ono et al., 2004).

Page 21: Sequence analysis, pathogenicity and cytokine gene

10

1.4. Diseases in birds other than poultry

Over the past few decades, adenoviruses have been described and isolated from a

wide variety of other avian species, such as waterfowl, game birds, cage birds, and

raptors (Adair and Fitzgerald, 2008).

Pigeon adenoviruses are associated with two distinct diseases, both causing

serious harm to the racing pigeon population. The first disease, called adenovirosis type I

or ‘classical adenovirosis,’ occurs when young pigeons first participate in races at the age

of 6 months. Pigeons are infected orally by cross-contamination in common baskets.

Consequently, virus is shed in feces, thereby rapidly infecting all other susceptible

pigeons in the loft. Only birds less than 1 year of age develop clinical signs, while older

pigeons remain clinically healthy. The virus replicates in the liver and in the nucleus of

intestinal epithelial cells, causing severe intestinal damage with extensive loss of proteins

and ions. The enhanced permeability of the intestinal wall, with bacterial overgrowth in

the intestinal lumen, creates a serious risk of septicaemia. Adenovirosis type II or

‘necrotizing hepatitis’ is characterized by sudden death in pigeons of all ages throughout

the year (De Herdt et al., 1995). Clinical signs are minimal, since birds die within 24 to

48 hr. Occasional clinical signs are vomiting and production of yellow, liquid droppings.

Mortality is approximately 30%, but in some cases can reach 100%. Histological

examination consistently reveals the presence of nuclear inclusion bodies in hepatocytes

and extensive hepatic necrosis. All conditions related to pigeon adenovirus infection are

reviewed by Vereecken et al. (1998). Growth analysis of adenoviruses isolated from

pigeons in chicken cells is described by Hess et al. (1998).

Page 22: Sequence analysis, pathogenicity and cytokine gene

11

Adenovirus is also implicated in guineafowl disease. The virus was isolated from

the pancreas of 2-week-old guinea fowl affected with acute necrotizing pancreatitis with

prominent viral inclusions by Zellen et al. (1989). Gross examination showed emaciation

of all affected guinea fowl, with an enlarged, hard, and nodular pancreas, resulting in

distortion of the duodenal loop. Microscopic findings are generalized pancreatic necrosis

with large basophilic intranuclear inclusion bodies.

With the rapid expansion of the ostrich farming industry there is a consequent

increase in disease problems similar to those that are observed in the poultry industry.

Isolation of six adenoviruses from ostriches and their identification using restriction

endonuclease (RE) analysis is reported by Gough et al. (1997). RE analysis reveals that

five of the isolated viruses have a close genetic relationship to the FAdV-8 serotype (TR

59), while one isolate recovered from the pancreas of a 6-week-old ostrich with gross

lesions of pancreatitis has an RE pattern similar to the FAdV-2 serotype. In Italy, Capua

et al. (1994) described pancreatitis-like lesions in an ostrich chick from which an

adenovirus was isolated. Moreover, adenoviruses from ostriches can be transmitted to

guinea fowl, resulting in deaths and associated pancreatitis (Capua et al., 1994).

Falcon adenovirus (FaAdV-1) is a new species in the genus Aviadenovirus with

close similarity to FAdV-1 and FAdV-4 (for more detail see sections 2.1. and 2.2.2.). The

penton base of the FaAdV-1 lacks the Arg-Gly-Asp (RGD) and Leu-Asp-Val (LDV)

recognition motifs for integrin binding, suggesting that cell entry depends on a different

receptor. FaAdV-1 has broad cell tropism and is a primary pathogen that can infect

Page 23: Sequence analysis, pathogenicity and cytokine gene

12

several species of falcons. The virus is associated with outbreaks of disease in captive-

raised nestling Northern Aplomado falcons, peregrine falcons, a Vanuatu falcon, Taita

falcons, orange-breasted falcons, and a merlin (Schrenzel et al., 2005; Van Wettere et al.,

2005). The primary lesions are inclusion body hepatitis, splenomegaly, and moderate to

severe lymphoid atrophy and lympholysis in the bursa of Fabricius. Virus is spread

horizontally (Schrenzel et al., 2005). The peregrine falcon is the primary reservoir host

for the virus, and FaAdV-1 replicates in cell cultures of falcon origin (Oaks et al., 2005).

The only other avian species that is seropositive to FaAdV-1 is the single barred owl.

Interestingly, falcons and owls are phylogenetically disparate, but share susceptibility to

another typically host-specific virus, columbid herpesvirus 1 (Sileo et al., 1983; Aini et

al., 1993).

1.5. Diagnosis of aviadenovirus infection

Laboratory diagnosis is based on direct detection, including isolation and

identification of the causative agent, or indirect detection with antibodies (Abs). Usually

the diagnosis of adenovirus infection in poultry is based on histology and detection of

intranuclear inclusion bodies in hepatocytes, or on detection of the antigen or virus

particles using an immunofluorescence test or electron microscopy. During the last two

decades, several molecular biological tools, such as PCR, real-time PCR, and RE analysis

have been developed for detection of viral DNA (Erny et al., 1991; Raue and Hess, 1998;

Jiang et al., 1999; Hess et al., 1999; Hess, 2000; Raue et al., 2002; Luschow et al., 2007;

Steer et al., 2009; Romanova et al., 2009). Most avian adenovirus PCRs utilizes the

hexon gene for primer design. The hexon consists of conserved regions (pedestals P1 and

Page 24: Sequence analysis, pathogenicity and cytokine gene

13

P2) located inside the virion and variable loops (L1 to L4) that protrude from the surface

and contain type-specific neutralizing epitopes (Toogood et al., 1992). Consequently,

hexon pedestal and loop sequences permit the design of oligonucleotides that generate a

PCR product from all twelve serotypes; an RE analysis of the amplified PCR product

may be conducted to differentiate the serotypes. However, virus isolation from chicken

kidney (CK) or chicken embryo liver (CEL) cells is an essential procedure for further

typing of FAdVs (Sharpless et al., 1961). Embryonated eggs injected via the yolk sac

route are also a sensitive medium (Cowen, 1988). A group-specific ELISA is suitable to

detect less than 100 mean tissue-culture infective doses of adenovirus per gram of liver

tissue (Saiffudin and Wilks, 1991). The most common serologic test is immunodiffusion,

but its sensitivity is lower than the ELISA’s.

1.6. Control

Biosecurity practices are the most essential step to prevent infection. All factors

related to proper management including cleaning and disinfection of premises and

equipment, restricted entry of visitors and vaccination crews in the poultry houses play a

considerable role in prevention of the disease (Adair and Fitzgerald, 2008).

Immunosuppressive viruses such as IBDV and CIAV act as predisposing factors for

FAdV outbreaks (Adair and Fitzgerald, 2008). Therefore, control or elimination of

immunosuppressive pathogens is another important step in management procedures.

Despite the relatively high mortality and economic losses in the past several years in the

Canadian broiler industry due to IBH, no commercial vaccines are available in Canada,

but they have been successfully used in other countries. Comparison of two inactivated

Page 25: Sequence analysis, pathogenicity and cytokine gene

14

vaccines prepared from the livers of infected chickens, one without adjuvant and another

with a liquid paraffin adjuvant, showed that oil emulsion vaccine is highly effective (Roy

et al., 1999). Additionally, inactivated chicken liver cell culture and embryonated egg-

propagated vaccine for subcutaneous application also provided sufficient protection

against infection (Naeem et al., 1995). Evaluation of five inactivated vaccines used in

Mexico confirmed complete protection with the absence of histological changes in organs

of challenged chicks (Shane, 1996). In Australia, commercial live vaccine grown in chick

embryo liver cells containing the FAdV-8b strain is applied for vaccination of broiler

breeders (Grimes, 2007). Effective protection of the progeny of chickens against IBH-

HHS is achieved by dual vaccination of breeders for FAdV-4 and CIAV (Toro et al.,

2002).

2. Adenoviruses 2.1. Taxonomy

The family Adenoviridae comprises five genera. The members of the genus

Mastadenovirus infect only mammals. The genus Aviadenovirus infects only avian

species; the genus Atadenovirus infects a broad host range such as ruminants, marsupials,

reptiles, and birds, while the genus Siadenovirus infects both amphibian and avian

species. Recently added to the family is the genus Ichtadenovirus. Virus was isolated

from the white sturgeon (Kovács et al., 2003). The genome of Atadenovirus has very

high (>60%) AT content, while Siadenovirus contains a seemingly unique gene, a

sialidase gene homolog at the left-hand side of the genome (Benko et al., 2005). Each

genus is serologically distinct from other genera and also differs significantly in genome

Page 26: Sequence analysis, pathogenicity and cytokine gene

15

organization (Benko et al., 2005). The genus Aviadenovirus, also referred to as group I

avian adenoviruses, comprises five species of FAdVs with one or more serotypes in each.

Species in the genus Aviadenovirus are classified as Fowl adenovirus A (FAdV-1); Fowl

adenovirus B (FAdV-5); Fowl adenovirus C (FAdV-4, FAdV-10); Fowl adenovirus D

(FAdV-2, FAdV-3, FAdV-9, FAdV-11); Fowl adenovirus E (FAdV-6, FAdV-7, FAdV-

8a, FAdV-8b); Goose adenovirus GoAdV-1, GoAdV-2, GoAdV-3; and Falcon

adenovirus A FaAdV-1. Species designation depends on several criteria, such as

calculated phylogenetic distances, restriction fragment length polymorphism (RFLP),

host range, pathogenicity, cross neutralization, and ability to recombine. The first

molecular classification of fowl adenoviruses was based on RFLP of the DNA given by

BamHI and HindIII (Zsák and Kisary, 1984).

2.2. Characteristics of adenoviruses 2.2.1. Structure and physicochemical composition of the adenovirus

The members of the family Adenoviridae are non-enveloped viruses with regular

icosahedral capsids that are 70-90 nm in diameter and have a pseudo T number of 25

(Stewart et al., 1991). The key structural component of adenoviruses is a homotrimeric

hexon (Fig. 1.1.). There are 240 hexons on the faces and edges of the capsid. The pentons

lie at the 12 vertices of the icosahedral capsid and comprise penton bases and protruded

fibers (van Oostrum and Burnett, 1985). The hexon, a trimer of protein II, has the largest

mass and methionine content in the virus structure. The DNA sequence revealed 28

methionine residues among the 967 amino acids (aa) of protein II (Akusjarvi et al., 1984;

Jornvall et al., 1981). The penton consists of the penton base, a pentamer of polypeptide

Page 27: Sequence analysis, pathogenicity and cytokine gene

16

III, and the fiber, which is a trimer of polypeptide IV. The fiber is formed by interaction

of three polypeptide IVs, noncovalently bound into a multimeric structure (Hong and

Engler, 1996). The complete adenoviral structure was determined by cryoelectron

microscopy and three-dimensional image reconstruction methods in 1991 (Stewart et al.,

1991). Adenoviruses contain 13% DNA and 87% protein. The virus has no membranes or

lipids and hence is stable in solvents such as ether and ethanol (Hong and Engler, 1996).

Virions can be disrupted with 5 M urea, 10% pyridine, acetone, or multiple freeze-thaw

cycles (Maizel et al., 1968a). Cesium chloride densities from 1.32 to 1.37 g/ml have been

described, presumably due to differences in DNA contents (Adair and Fitzgerald, 2008).

2.2.2. Proteins

The viral capsid contains the structural proteins II, III, IIIa, IV, VI, VIII, and IX

(Fig. 1.1). Proteins V, VII, X, and terminal protein (TP), which directly interact with the

viral DNA, form the viral core (Hosokawa and Sung, 1976; van Oostrum and Burnett,

1985; Lehmberg et al., 1999). The polypeptides present in the mature virion are

designated by Roman numerals from II to XII in order of their decreasing molecular mass

in acrylamide gels (Maizel et al., 1968b). Protein II (hexon protein), 3 of which form a

hexon, is the most abundant protein in the virion and defines the type, group, and

subgroup antigenic determinants (Norrby, 1969). Protein III (penton protein) has diverse

functions, including pentamerization to form a penton base, and ensure stable fiber-

penton base interaction (Karayan et al., 1997). Protein IV (fiber protein) interacts with

the penton base at the vertices of the icosahedral capsid. The fiber is directly responsible

for viral attachment and internalization of the virus into the host cell (San Martin and

Page 28: Sequence analysis, pathogenicity and cytokine gene

17

Burnett, 2003). Commonly, the fiber structure is divided into three distinct regions

(Green et al., 1983). The first region consists of the N-terminus, also known as the tail,

which is embedded into the penton base. The second region, designated as the shaft,

contains numerous repeating motifs of approximately 15 aa. The third region located at

the C-terminus, the end of the shaft, is called the knob and is necessary and sufficient for

virion binding to the host cell. The mammalian adenoviruses, with a few exceptions

(HAdV-40 and HAdV-41), have one fiber. However, FAdVs have 2 fibers (Gelderblom

and Maichle-Lauppe, 1982) of similar length except in the chicken embryo lethal orphan

(CELO) virus, FAdV-1, in which the fibers are of different lengths. The long fiber of

FAdV-1 is four times longer than the short one and is placed at a 90º angle (Hess et al.,

1995). The differences in the FAdV-1 fibers might be responsible for some of the

properties of the virus, such as hemagglutination and abundant growth in embryonated

eggs that other fowl adenoviruses do not have.

Minor capsid components, including proteins IIIa, VI, VIII, and IX complete the

capsid structure and primarily act as cement proteins. The location of each of them and

their precise role in the virus life cycle has yet to be fully understood. The location of

polypeptide IIIa has been recently defined as a position below the penton base and this

protein is associated with others at the vertex, mainly with the hexon and protein V

(Saban et al., 2005). The location of polypeptide VI in the virion appears to be inside

cavity of every hexon trimer, connecting the bases of two adjacent peripentonal hexons

and tethering the capsid to the core region (Silvestry et al., 2009). Protein VI is liberated

from the virion when the acidic environment of the endosome triggers conformational

Page 29: Sequence analysis, pathogenicity and cytokine gene

18

changes, and this promotes membrane disruption to allow the transport of the virus to the

nucleus (Wiethoff et al., 2005). Moreover, protein VI harbors a PPxY motif involved in

reapid microtubule-dependent intracellular movement and infectivity. Inactivation of the

PPxY motif leads to post-entry delay with consequence on reduced infectivity and

prevent efficient accumulation of the virions at the microtubule organizing center

(Wodrich et al., 2010; Maier and Wiethoff, 2010).

Polypeptide VIII is located at the inner surface of the capsid and it is bound

between the peripentonal hexons and the rest of the capsid (Stewart et al., 1993). The

smallest minor capsid protein, protein IX, is present as a trimer and helps to stabilize the

virion, as mutant virions lacking this protein are less stable than the wild type virus

(Colby and Shenk, 1981). Proteins V, VII, µ, and TP directly interact with the linear

double-stranded DNA to form the viral core (van Oostrum and Burnett, 1985; Lehmberg

et al., 1999). Polypeptide VII is a major core protein with over 800 copies per virion and

together with DNA forms compact repeating structures called ‘adenosomes’ (Tate and

Philipson, 1979). Two copies of the TP are covalently linked to the 5’ end of the genome.

TP acts as a primer for DNA replication and may facilitate circularization of the virus

genome. The TP precursor, pTP, is cleaved by virus-encoded protease at two sites,

creating TP, that is essential for virus replication (Webster et al., 1997).

About 30 adenovirus non-structural proteins have been described (San Martin and

Burnett, 2003). They are generally produced in small quantities and the precise role for

most of them is still unknown. To date, the structures of two of the non-structural

Page 30: Sequence analysis, pathogenicity and cytokine gene

19

proteins have been described, DNA-binding protein (DBP) and viral protease (Tucker et

al., 1994; McGrath et al., 2003). DBP is a multifunctional nuclear phosphoprotein. It is a

major product of early region E2A and is essential for viral DNA replication. It is also

required for the control of early and late transcription (Ward et al., 1998; van der Vliet et

al., 1978). This protein has two domains: a highly phosphorylated N-terminal domain of

173 aa that contains a nuclear localization signal and the C-terminal domain of 356 aa

(Morin et al., 1989). The highly conserved and non-phosphorylated C-terminal domain

binds DNA and is active in DNA replication (Klein et al., 1979; Linne and Philipson,

1980). DBP binds single-stranded (ss) and dsDNA as well as RNA. DBP binds

cooperatively to ssDNA, protecting it against nuclease digestion and destabilizing the

double helix during the elongation phase of DNA replication. Binding to dsDNA occurs

in a non-cooperative fashion and is less stable (Tsernoglou et al., 1984; Zijderveld and

van der Vliet, 1994).

Proteinases play essential roles at various stages of viral replication, including

assembly and maturation of virions. Adenoviruses encode an endopeptidase that is crucial

to the assembly of the virus in the nucleus of infected cells (Weber, 1995). Between 10

and 30 molecules of the protease are packaged into each virion. Cotten and Weber (1995)

demonstrated that the adenovirus-encoded 23K protease is required for cleavage of virion

precursor proteins.

Page 31: Sequence analysis, pathogenicity and cytokine gene

20

The adenoviral enzyme is the only protease known to require DNA for maximum

activity. The atomic structure of the HAdV-2 protease revealed an ovoid-shaped particle

with α-helical regions at both poles and a central β-sheet (Ding et al., 1996).

2.2.3. Genome

The mastadenovirus genome is around 36 kilobase (kb) and consists of four early

regions (E1 to E4), and five late regions (L1 to L5), while the FAdV genome is around 45

kb and is the largest genome among members of the family Adenoviridae (Chiocca et al.,

1996; Ojkic and Nagy, 2000). So far, the complete nucleotide sequences of the genomes

of FAdV-1, FAdV-9, FAdV-8, FAdV-4, and turkey adenovirus type 1 (TAdV-1) are

determined (Chiocca et al., 1996; Ojkic and Nagy, 2000; Grgic et al., 2011; Griffin and

Nagy, 2011; Kajan et al., 2010). The nucleotide sequences at the left and right ends of the

genomes of FAdVs representing species C (FAdV-4 and FAdV-10), D (FAdV-2), and E

(FAdV-8) are also determined and compared to FAdV-1 (FAdV-A) and FAdV-9 (FAdV-

B) (Corredor et al., 2006; Corredor et al., 2008). Nucleotide sequence homology and aa

sequence identities are the highest among members of the same species group, while

different degrees of variations are present among all FAdVs. Moreover, all the studied

FAdVs have a similar gene arrangement, suggesting probably similar gene functions

(Corredor et al., 2008). Interestingly, at the right end of all analyzed FAdV genomes,

homologues to known ORFs such as lipase and Gam-1 are also found (Corredor et al.,

2008). GAM-1, encoded by ORF8, is highly conserved (82%-95.7%) among viruses

belonging to the same species group. In the following text, interesting features related to

FAdVs will be pointed out.

Page 32: Sequence analysis, pathogenicity and cytokine gene

21

FAdVs contain large genomes [FAdV-4 45,667 base pair (bp); FAdV-9 45,063

bp; FAdV-8 44,055 bp) (Griffin and Nagy, 2011; Ojkic and Nagy, 2000; Grgic et al.,

2011] although genes encoding a few structural proteins such as V and IX, and early

regions E1, E3, and E4 described in mastadenoviruses are missing from them. The

genomes of atadenoviruses and siadeoviruses are shorter. For example, the EDSV and

HEV are 33.2 and 26.3 kb, respectively (Hess et al., 1997; Pitcovski et al., 1998). The

highest degree of conservation in the adenovirus genome is in the middle part, which

encodes the structural proteins, whereas the two ends, containing mainly regulatory early

regions, exhibit a larger variation (Davison et al., 2003). In mastadenoviruses, early

regions E1A and E1B are present without exception, while in atadenoviruses, a p32K

gene is present in the position of the E1A of mastadenoviruses. The p32K gene is unique

to atadenoviruses and has been identified in every member of the genus Atadenovirus

studied so far. In addition, only the atadenoviruses seem to have some degree of

homology with the mastadenovirus E1B genes, namely the 55K protein. Moreover, at the

left-hand end of the siadenovirus genome, there is a putative sialidase gene in the place of

the E1 region of mastadenoviruses (Benko et al., 2005). Protein IX and V, found in

mastadenoviruses, are absent in the avi-, at- and siadenoviruses, suggesting that the

virions of these adenoviruses are different structurally. The E3 region seems to exist only

in the mastadenoviruses and no homologous genes were detected in any other genera.

Apparently, E3 represents one of the most commonly used insertion sites by the

adenovirus gene delivery system (Russell, 2000). However, the putative E3 region of

siadenoviruses has no homology with the E3 region of mastadenoviruses. Furthermore,

the right-hand end of the genome harbors an E4 region where the single gene for the 34K

Page 33: Sequence analysis, pathogenicity and cytokine gene

22

protein of mastadenoviruses has its homologue in atadenoviruses only (Benko et al.,

2009). The presence of tandem repeats (TR) at the right end of the genome is reported for

several FAdVs, namely FAdV-9, FAdV-8, FAdV-4, and FAdV-10 (Ojkic and Nagy,

2000; Corredor et al., 2008; Grgic et al., 2011; Griffin and Nagy, 2011). TR-2 of FAdV-

9, consisting of 13 repeats of 135 bp each, is dispensable for virus replication in vitro and

in vivo (Ojkic and Nagy, 2000; 2001). The role of these tandem repeats remains

unknown. The GAM-1 protein is unique to FAdVs only and has functional equivalence to

the human adenovirus E1A 243R, E1A 289R, and E1B 19 kDa proteins. GAM-1 of

FAdV-1 virus is involved in increasing cellular transcription by blocking pRb and

activating E2F-dependent transcription (Lehrmann and Cotten, 1999). Additionally,

lipase ORFs detected in the genome of FAdV-D and FAdV-E species have high identities

to lipase from pathogenic avian herpesviruses (Corredor et al., 2008). Lipase might be

important in virus replication because mutant Marek’s disease ( MDV-1) virus with

foreign gene insertion in the v-Lip LORF2 coding region showed reduced lytic replication

in vivo (Kamil et al., 2005).

Page 34: Sequence analysis, pathogenicity and cytokine gene

23

Figure 1.1. Schematic transection of the adenovirus particle. Taken from Nemerow et al. (2009).

Page 35: Sequence analysis, pathogenicity and cytokine gene

24

2.3. Adenovirus replication cycle 2.3.1. Attachment and penetration

Adenovirus entry into cells involves attachment of the fiber knob to the primary

receptor. Interaction of the penton base with a secondary receptor is responsible for virus

internalization. The virus enters the cell through receptor-mediated endocytosis in a

clathrin-coated vesicle and is transported to endosomes where acidification results in

partial disassembly of the capsid (Fig. 1.2). The altered virion escapes into the cytoplasm

and is transported to the nucleus, where replication occurs (Meier and Greber, 2004).

However, the current model of adenovirus entry is based mainly on studies with HAdVs

of species B and C.

The Coxsackievirus and adenovirus receptor (CAR) is the adenovirus receptor for

most but not all HAdVs (Bewley et al., 1999). CAR is a member of the immunoglobulin

superfamily and comprises two immunoglobulin-like extracellular domains. This receptor

is localized in specialized tight junctions near the tops of the cells and along the

basolateral membranes (Cohen et al., 2001). It is present in adherens junctions along the

lateral sides of airway cells (Walters et al., 2002). In humans, CAR is abundantly

expressed in a number of organs, including the heart, brain, pancreas, and intestines

(Tomko et al., 1997), as well as the lung, liver, and kidney (Fechner et al., 1999). The

presence of CAR on target cells permits virus attachment by HAdVs of species A, C, E,

and F, and also serves as a receptor for several animal adenoviruses: chimpanzee CV-68,

canine type 2, and FAdV-1 (Cohen et al., 2002; Soudais et al., 2000; Tan et al., 2001).

Some HAdVs of species B attach to the CD46 membrane cofactor protein (Cole et al.,

1985), whereas several species D viruses infect cells after attachment to α (2-3) linked

Page 36: Sequence analysis, pathogenicity and cytokine gene

25

sialic acid, a common carbohydrate component of glycoproteins and glycolipids

(Burmeister et al., 2004). Differences in primary receptors are reflected in virus tropism.

CAR-binding serotypes are mainly respiratory viruses, whereas HAdV-40 and HAdV-41

are enteric and their primary receptors are unknown. The main function of the fiber

receptor is to hold the virion in close proximity to the cell surface, permitting interaction

with an integrin molecule. The penton base protein is responsible for virus

internalization, through interaction of an RGD sequence with cell integrins. The RGD

peptide is localized in the flexible loop region of the penton base and serves as a

recognition site for several cellular integrins (Stewart et al., 1997). It has been shown that

the integrins that facilitate adenovirus entry include the vitronectin receptors αvβ3 and

αvβ5 (Wickham et al., 1993), as well as αvβ1 (Li et al., 2001), α3β1 (Salone et al., 2003),

α5β1 (Davison et al., 1997), and all of them recognize RGD ligands. The penton base-

integrin interaction induces activation of PI3 kinase (PI3K) (Li et al., 1998a), p130CAS (Li

et al., 2000), and Rho GTPases (Li et al., 1998b) that are important for virus

internalization and virus escape from the endosome. HAdV-40 and HAdV-41 lack the

RGD motif which explains why different receptors recognize these viruses; instead, the

HAdV-40 carries RGAD and HAdV-41carries IGDD (Albinsson and Kidd, 1999;

Davison et al., 1993). Interestingly, FAdVs also lack an RGD motif (Hess et al., 1995).

Page 37: Sequence analysis, pathogenicity and cytokine gene

26

Figure 1.2. Reproductive cycle of human adenovirus type 2. The virus attaches to a permissive cell and enters the cell via endocytosis 1), 2), partial disassembly takes place prior to entry of particles into the cytoplasm 3). Further uncoating occurs and viral genome is imported into the nucleus 4). Transcription of E1A gene by host cell RNA polymerase II 5). Export of E1A mRNAs to the cytoplasm 6). Synthesis of E1A proteins 7), and import into the nucleus 8), where regulate transcription of both cellular and viral genes 9a). Transcription of the VA genes by host cell RNA polymerase III 9b). The early pre-mRNA species are processed, exported to the cytoplasm 10) and translated 11). These early proteins are imported into the nucleus 12) and cooperate with cellular proteins in viral DNA synthesis 13). Replicated viral DNA molecules serve as template for replication 14) or for transcription of late genes 15). Processed late mRNA species are selectively exported from the nucleus 16). Their efficient translation in the cytoplasm 17) requires the major VA RNA, VA-RNA I and late L4 100kDa protein. The later protein serves as chaperone for assembly of hexons 18). Within nucleus capsids are assembled 19). Assembly requires a packaging signal located at the left of the genome and mature virions are formed when precursor protein are cleaved 20). Progeny virions are released 21). Taken from Flint et al., 2004.

Page 38: Sequence analysis, pathogenicity and cytokine gene

27

2.3.2. Activation of early viral genes

Replication cycle of adenoviruses is determined by the expression of early and

late genes, separated by the onset of viral DNA replication. The E1A is first

transcriptional unit expressed after infection. Initially, transcripts are processed into two

mRNAs that encode proteins 243R and 289R, while three additional alternatively spliced

mRNA congregate later in the infectious cycle (Yang et al., 2002). Furthermore, E1B

gene promoter requires readthrough transcription originating from the upstream E1a

gene, activates transcription factors necessary for transcription of the E1B genes (Shen

and Spector, 2003). Alternative splicing of the E1B pre-mRNA give a rise to E1B 55 kDa

and E1B 19 kDa proteins.

Two E1A proteins are expressed from differentially spliced mRNAs at 1-2 hours

post-infection (h.p.i.) (Yang et al., 2002). Role of these proteins is to promote cell cycle

progression to favor viral DNA replication, counteract the innate immune response

mediated by interferon and inhibit apoptosis. The large 289R protein activates

transcription of early genes from the E1B, E2 and E4 promoters in the infected cells

(Bandara and La Thangue, 1991). E1A proteins interact with cellular proteins and

modulate their function in order to activate the transcription of viral genes. Both 243R

and 289R E1A proteins interact with pRB/p130, TATA-binding protein (TBP),

p300/CBP (Arany et al., 1995), STAT (Look et al., 1998), p21 (Keblusek et al., 1999),

cyclins A and E/CDK complexes (Faha et al., 1993), etc. Binding of the E1A proteins to

Rb bypass the normal regulatory signals and release E2F proteins (Whyte et al., 1988;

Page 39: Sequence analysis, pathogenicity and cytokine gene

28

Chellappan et al., 1992) whose target genes are involved in cell cycle progression and

apoptosis.

2.3.3. DNA replication

Genome replication depends on three viral proteins encoded by the E2

transcription unit: pTP, DNA polymerase, and DBP. These three proteins with the core

origin can initiate DNA replication, but only inefficiently. Additionally, cellular nuclear

factor I (NFI) or CAAT transcription factor (CTF1) and nuclear factor III (NFIII) or

octamer-binding protein (Oct-1) are required for an efficient replication (de Jong and

Van der Vliet, 1999). Origins of DNA replication are present in both inverted terminal

repeats (ITR) of the genome. These origins contain a triplet repeat at genome termini

(GTAGTA) in HAdV-2 and HAdV-5, followed by a conserved region that serves as a

core origin where the pTP/polymerase complex binds. Upon initiation, DNA polymerase

and pTP form a heterodimer and DBP binds dsDNA to coat the entire genome. The

coating of DNA template increases the affinity of NFI for its recognition site in the origin

of replication. Distal to the NFI binding site is the DNA recognition site for NFIII (de

Jong et al., 2003). A pre-initiation complex is then formed and stabilized by direct

interaction of DNA polymerase with NFI and DBP bound to its recognition site (bp 23-

38) in the origin of replication (Armentero et al., 1994). The initiation of adenovirus

DNA replication occurs by a protein-priming process in which the first nucleotide

(dCMP) becomes covalently bound to a serine residue in the viral pTP. When the third

nucleotide is incorporated, the resulting trinucleotide complex (pTP-CAT) jumps back

three bases to the first GTA and disassociation of pre-initiation complex begins (de Jong

Page 40: Sequence analysis, pathogenicity and cytokine gene

29

et al., 2003). The pTP dissociates from DNA polymerase and the latter replicates the

genome by strand displacement. The displaced DNA strand forms a pan-handle structure

by complementary terminal sequences in the ITRs and the complex pTP-DNA

polymerase assembles to initiate the synthesis of the daughter strand (Evans and Hearing,

2002).

2.3.4. Late gene expression

Following the onset of viral DNA replication, the IVa2 and IX genes are

expressed at high levels to activate transcription through MLP (Lutz and Kedinger, 1996;

Lutz et al., 1997). Primary late transcript is about 28 kb in length and gives rise to 5

transcripts termed L1 to L5 by the use of 5 alternative polyadenylation signals. Following

that, each transcript undergoes alternative splicing to generate mature mRNAs for late

proteins (Chow et al., 1980). All alternatively spliced mRNAs contain a 5’ leader

sequence of 200 nts, which comes from the tripartite leader sequence. Leader sequence is

important for efficient translation of late mRNAs independent of host initiation factor

eIF4F (Dolph et al., 1988; Huang and Schneider, 1991).

2.3.5. Assembly and release

Onset of viral DNA replication coupled with the production of large quantities of

structural proteins provides an optimal setting for virus assembly. The assembly of

trimeric hexon capsomeres requires the L4 100K protein (Hong et al., 2005). The 100K

protein is required for the transport of newly synthesized hexon monomers to the nucleus

(Cepko and Sharp, 1982) and also acts as a scaffold to facilitate assembly of hexon

Page 41: Sequence analysis, pathogenicity and cytokine gene

30

trimers (Morin and Boulanger, 1986). Penton capsomeres, consisting of a pentameric unit

and trimeric fiber, assemble slowly in the cytoplasm. Following their production, hexon

and penton capsomeres are imported into the nucleus and virion assembly occurs. The

viral DNA is subsequently encapsidated into an empty capsid via a packaging signal

located at the left end of the viral genome (between nts 194 and 382 in HAdV-5) in a

polar, left-to-right fashion (Schmid and Hearing, 1995; Hammarskjold and Winberg,

1980). Proteins IVa2 and IX are important for packaging of full-length genomes into the

capsid (Ghosh-Choudhury et al., 1987; Zhang and Imperiale, 2003). The core proteins are

encapsidated along with the viral genome. During the maturation process, precursor

proteins pIIIa, pTP, pVI, pVII, pVIII, and pX are cleaved by AdV protease. The cleavage

of precursor proteins is required for the stabilization and infectivity of the viral particles

(Weber, 1995; Cotten and Weber, 1995). The adenovirus death protein (ADP), encoded

by the E3 region, is expressed at high levels late in infection and promotes the release of

virions from infected cells (Tollefson et al., 1996). The mature virions are released by

cell lysis.

3. Avian immune response

In order to understand the immunology of the lymphoid system, it is necessary to

know its basic structure. A number of morphologically and functionally different organs

and tissues have various functions in the development of immune responses. Based on

function, they can be differentiated as primary and secondary lymphoid organs. Thymus,

bursa, and bone marrow are considered primary lymphoid organs, while the secondary

lymphoid organs are lymph nodes, spleen, mucosal-associated lymphoid tissue (MALT),

Page 42: Sequence analysis, pathogenicity and cytokine gene

31

and gut-associated lymphoid tissue (GALT). The thymus-derived cells are a separate

lineage of antigen-specific lymphocytes that do not produce immunoglobulins. The

haematopoietic component that differentiates in the thymus into T lymphocytes,

macrophages, and medullary dendritic cells is derived from the mesoderm (Cooper et al.,

1966). On the other hand, the bursa of Fabricius is a lymphoepithelial organ in birds that

arises as a dorsal endomesodermal diverticulum of the cloaca. Lymphocytes which

differentiate in the bursa of Fabricius do not originate from the bursa rudiment itself, but

are derived from stem cells that invade the bursa between day 8 and 14 of embryonic

development (EID). Additionally, immunoglobulin diversity in chickens is generated by

somatic gene conversion (Ratcliffe and Jacobsen, 1994).

Several important characteristics differentiate innate from adaptive responses.

Firstly, pathogen recognition in the innate system is mediated by elements encoded in the

germline DNA, while B and T cell receptors are formed by rearrangement of receptor

gene elements, a process that provides large diversity of exquisitely different receptors.

Secondly, the different responses of innate and adaptive immunity are a direct

consequence of the diversity and frequency of receptors utilized to sense the pathogens. If

the frequency of the receptor is high, the response can be rapid, but specificity is low,

whereas with B and T cell receptors, which are clonally expressed, the rate of the

response is slow, but specificity is very high. Thirdly, immunological memory makes the

adaptive response more efficient at the second exposure, while the innate response is not

influenced by pre-exposure (Iwasaki and Medzhitov, 2004).

Page 43: Sequence analysis, pathogenicity and cytokine gene

32

3.1. Innate immune response

Innate immunity is based on physical and chemical barriers of infection, including

different cell types responsible for recognizing invading pathogens and activating

immune responses. The cellular components include antigen-presenting cells (APCs),

dendritic cells (DCs), phagocytic macrophages and granulocytes, cytotoxic natural killer

(NK) cells, and γδ T lymphocytes. The first response recognizes microorganisms via

germline-encoded pattern-recognition receptors (PRRs) (Akira et al., 2006). These

receptors are capable of recognizing invading pathogens by two mechanisms. Firstly, the

PRRs recognize microbial components via pathogen associated molecular patterns

(PAMPs), which are essential for the survival of the microorganism and are conserved

within a group of pathogens. For example, lipopolysaccharide and double-stranded RNA

(dsRNA) are PAMPs associated with Gram-negative bacterial infections and viral

infections, respectively. The second way for receptors to sense infection is indirect, via an

encounter between PRRs and host factors, when the latter are present in unusual locations

as a direct consequence of infection, inflammation, or other types of cellular stress.

Examples of such events are host-derived intracellular heat-shock proteins (HSPs) that

are recognized by some PRRs when exposed to the extracellular environment (Beg,

2002). The main function of PRRs comprise enhancement of phagocytosis, activation of

complement cascades, production of inflammatory cytokines, and maturation of DCs,

that all together orchestrate the early host response to infection and are an important link

to the adaptive immune response (Akira et al., 2006).

Page 44: Sequence analysis, pathogenicity and cytokine gene

33

Toll-like receptors (TLRs) as pattern-recognition receptors are most extensively

studied and were originally identified as a gene product essential for dorso-ventral

patterning during embryogenesis as well as for the antifungal response in Drosophila

(Lemaitre et al., 1996). Structurally, TLRs are type I integral membrane glycoproteins

characterized by an extracellular ligand-binding domain containing leucine-rich repeat

(LRR) motifs and a cytoplasmic signaling domain homologous to that of the interleukin 1

receptor (IL-1R), named the Toll/IL-1R homology (TIR) domain (Bowie and O’Neill,

2000). Pathogen binding to TLRs through PAMP-TLR interaction causes receptor

oligomerization which triggers intracellular signal transduction (O’Neill and Bowie,

2007). To date, a total of 13 TLRs (TLR1 to TLR13) have been identified in mammals

and mice, and each recognizes distinct PAMPs derived from different microbial

pathogens, including viruses, bacteria, fungi, and protozoa (Kaisho and Akira, 2006).

Temperley et al. (2008) found a total of ten TLRs in chicken, with only five human

orthologs, one fish ortholog, and three unique to chickens. Chromosome mapping of

chicken TLR genes shows that half of theTLR genes are located on chromosome 4

(Temperley et al., 2008). Commonly, TLRs separate into subgroups expressed at the cell

surface (TLRs 1, 2, 4, 5, 6, and 10) or at endosomes (TLRs 7, 8, and 9). Distribution can

be cell type dependent. TLR3 is expressed on the surface of human fibroblasts but

monocyte derived immature DCs express TLR3 intracellularly and not on the cell surface

(Matsumoto et al., 2003; Wagner, 2004). Usually, TLRs expressed on endosomal

membranes recognize RNA and DNA motifs, whereas cell membrane bound TLRs

recognize proteins, lipoproteins, and lipopolysaccharide.

Page 45: Sequence analysis, pathogenicity and cytokine gene

34

A second large class of PRRs is the family of intracellular sensors of bacterial

motifs known as nucleotide-binding oligomerization domain (NOD) proteins. The NOD

proteins contain an amino-terminal signaling domain and a carboxy-terminal ligand-

binding domain. The ligand binding domain is an LRR domain very similar to that of

TLRs (Inohara and Nunez, 2003).

A third class of PRRs, RNA helicases, were identified by Yoneyama et al. (2004).

The authors demonstrated that two cytoplasmic proteins, retinoic acid inducible gene

(RIG) I and melanoma differentiation associated gene 5, each with two amino-terminal

signaling domains and a carboxy-terminal helicase domain, become activated in response

to virus infection. Moreover, the authors concluded that helicases are probably kept in a

latent state in the absence of ligand engagement, and become activated through a

mechanism dependent on the ATPase activity of the helicase domain (Yoneyama et al.,

2004).

Lastly, the dsRNA-activated protein kinase R (PKR) has emerged as an important

signal transducer in the proinflammatory response to different agents. The activation of

PKR during infection by viral dsRNA down-modulates protein synthesis in virus infected

cells and inhibits viral replication (Nanduri et al., 2000).

In order to understand the cellular responses raised by different viruses, it is

necessary to know which PRRs become activated during viral infection and which viral

molecules serve as ligands for different PRRs. TLRs expressed on the cell surface, such

Page 46: Sequence analysis, pathogenicity and cytokine gene

35

as TLRs 2 and 4, are capable of recognizing viral proteins. Innate sensing of the members

of the herpesvirus, retrovirus, and paramyxovirus families by TLRs has been

demonstrated by Compton et al. (2003) and Rassa et al. (2002). They showed that

detection of measles virus and cytomegalovirus (CMV) particles by TLR2 results in the

activation of innate immune responses. On the other hand, respiratory syncytial virus and

mouse mammary tumor virus are potent triggers of proinflammatory cytokine expression

that occurs via ligation to TLR4.

Moreover, it has been shown by Beg (2002) that host proteins may also activate

the innate immune system during viral infections. Several endogenous proteins activate

the immune system via TLR2 and TLR4 when located extracellularly. Some of them are

HSPs found in all prokaryotes and eukaryotes which under normal physiological

conditions are expressed at low levels. In addition, stress stimuli such as environmental

(UV radiation, heavy metals), pathological (viral, bacterial infection, fever), or

physiological (cell differentiation, growth factors) stimuli can cause a stress response

characterized by increased intracellular HSP synthesis. Asea et al. (2002) showed that

HSP exits mammalian cells and interacts with cells of the immune system, exerting

immunoregulatory effects. The same authors showed that exogenously added HSP70 has

potent cytokine activity, activates NF-κB, up-regulates the expression of

proinflammatory cytokines in human monocytes, and utilizes both TLR2 and TLR4 to

transduce its proinflammatory signal in a CD14 dependent fashion (Asea et al., 2002). It

has been reported by Taddeo et al. (2002) that expression of HSP70 is robustly elevated

during lytic herpes simplex virus infection and TLR2 contributes to HSV-induced

Page 47: Sequence analysis, pathogenicity and cytokine gene

36

inflammation. This result shows a potential role of endogenous ligands in the

inflammatory response raised during lytic viral infection by HSV-1.

As described previously, TLR3, TLR7/8, and TLR9 are expressed primarily in

endosomes and are activated only if the ligands reach this cellular compartment. It has

been shown by Matsumoto et al. (2003) that TLR3 recognizes dsRNA and transduces

signals to activate NF-κB and interferon (IFN)-β promoters. Regulation and localization

of TLR3 are different in each cell type, reflecting participation of cell type-specific

multiple pathways in antiviral IFN induction via TLR3. It is known that dsRNA serves as

an alarm signal associated with viral infection and leads to stimulation of the innate

immune response. In contrast, the immunostimulatory potential of ssRNA is still poorly

understood. Heil et al. (2004) reported that guanosine (G)- and uridine (U)-rich ssRNA

oligonucleotides derived from the human immunodeficiency virus-1 (HIV-1) stimulate

dendritic cells and macrophages to secrete interferon α and proinflammatory as well as

regulatory cytokines. It has been shown that murine TLR7 and human TLR8 mediate

species-specific recognition of GU-rich ssRNA. These results suggest that ssRNA

represents a physiological ligand for TLR7 and TLR8 (Heil et al., 2004).

Recent studies have shown that viral DNA genomes serve as TLR9 ligands for

herpesviruses. Lund et al. (2003) provided the first evidence that plasmacytoid dendritic

cells (pDCs) recognize cytosine-phosphate-guanosine (CpG) motifs present in the

genomic DNA of viruses via TLR9, and that this recognition is required for IFN-α

secretion by these cells. Additionally, TLR9-mediated recognition of HSV suggests the

importance of this recognition pathway for other DNA viruses by the pDCs (Lund et al.,

Page 48: Sequence analysis, pathogenicity and cytokine gene

37

2003). McGuire et al. (2011) showed that HAdV-5 infection induces reactive oxygen

species (ROS) to activate immune cells. ROS induction depends on virally induced

endosomal rupture and release lysosomal cathepsin which leads to mitochondrial

membrane disruption and release of ROS from mitochondria. Moreover, disruption of

lysosomal membranes and the relase of cathepsin B in cytoplasm are essential for HAdV-

5 induced Nod-like receptor (NLRP3) activation (Barlan et al., 2011a, b).

3.1.1. Cytokines

Cytokines are soluble, low molecular weight polypeptides and glycopeptides,

often further subclassified as monokines and lymphokines produced by the

monocyte/macrophages and lymphocytes, respectively. According to Murphy et al.

(2000), a strong cellular immune response against intracellular pathogens is mediated by

IFN-γ-secreting Th1 cells, while Th2 cells, producing IL-4, IL-5, and IL-13, are involved

in responses to parasitic pathogens.

IFN is a glycoprotein produced by white blood cells and viral-infected cells in

response to viral infection. The IFNs are divided into three distinct classes, termed IFN-α,

IFN-β, and IFN-γ. Type I interferons IFN-α and IFN-β are the key cytokines produced

after viral infection and mediate induction of both innate and adaptive immunity to

viruses (Theofilopoulos et al., 2005). They also induce the maturation of DCs, mediate

antigen presentation via major histocompatibility complex (MHC) class I, and mediate

induction of antigen-specific CD8+ T cell responses. Cellular production of IFN-α/β after

virus infection is triggered by dsRNA and virtually any cell type can express IFN-α/β

Page 49: Sequence analysis, pathogenicity and cytokine gene

38

following virus infection. Type II IFN (IFN-γ or immune IFN) is derived from T

lymphocytes and NK cells. It has been shown that IFN-γ has some involvement in most

stages of the inflammatory and immune responses. To date, all three types of chicken IFN

have been cloned (Sekellick et al., 1994; Digby and Lowenthal, 1995). IFN-γ is an

essential cytokine in the host defense against various pathogens. It is directly responsible

for inhibition of MDV replication by inducing nitric oxide (NO) synthesis (Xing and

Schat, 2000), also reducing replication of coccidian parasites and inhibiting src

oncogene-induced tumor growth (Plachy et al., 1999; Gobel et al., 2003).

IL-18 is considered a proinflammatory cytokine expressed by a wide variety of

cells such as macrophages, DCs, and Kupffer cells (Kaiser et al., 2003). It is produced as

a 24-kD inactive precursor (pro-IL-18) that needs to be cleaved by endoprotease IL-1β-

converting enzyme (caspase-1) to generate biologically active IL-18. However, cleavage

of pro-IL-18 is not exclusive to caspase-1: proteinase 3 also can generate biologically

active IL-18 (Gracie et al., 2003). Although originally identified as an inducer of IFN-γ,

IL-18, in combination with IL-12, enhances T and NK cell maturation, cytokine

production, and cytotoxicity (Kaiser et al., 2003). Based on its potent immunomodulatory

properties, IL-18 is essential to host defenses against a variety of infections. During viral

infection, IL-18 effectively induces IFN-γ and activates CD8+ T cells which are important

for viral clearance. In an in vivo model of vaccinia virus infection, administration of IL-

18 significantly suppresses pock formation (Tanaka-Kataoka et al., 1999). Moreover,

clearance of neurovirulent influenza A virus is impaired in IL-18-deficient mice (Mori et

al., 2001). Cho et al. (2001) reported that down-modulation of IL-18-induced immune

Page 50: Sequence analysis, pathogenicity and cytokine gene

39

responses by human papilloma virus (HPV) oncoproteins might contribute to viral

pathogenesis or carcinogenesis. Kaiser et al. (2003) showed differential IL-18 expression

in the spleens of chickens of susceptible and resistant genotypes following MDV

infection.

IL-10 is a regulatory cytokine, originally named cytokine-synthesis inhibitory

factor (CSIF) because of its ability to modulate the immune response towards a type 2

response by inhibiting IFN-γ synthesis (Rothwell et al., 2004). Interestingly, IL-10 can

have both an immunosuppressive and an immunostimulating effect on cell mediators of

adaptive immunity. IL-10 is produced by multiple cell types, including DC, B cells,

macrophages, CD4 T cells, CD8 T cells, and NK cells, as well as innate and adaptive

regulatory T cells (O’Garra and Vieira, 2007). Several authors report that IL-10 acts

directly on antigen presenting cells in order to decrease stimulatory molecule expression

and alter cytokine production (Carbonneil et al., 2004; Moore et al., 2001). It has been

shown by Sloan and Jerome (2007) that herpes virus remodels the intracellular signaling

pathway of T cells to induce production of IL-10 in order to evade the host immune

response. Abdul-Careem et al. (2007) showed that the presence of IL-10 is closely

correlated to vaccine failure in MDV-infected chickens. IL-10 is also a contributing

factor in susceptibility of chickens to Eimeria maxima infection (Rothwell et al., 2004).

Viruses have evolved a number of diverse strategies to avoid the host

inflammatory response. Such strategies have been observed for poxviruses and

herpesviruses, which modulate cytokine action by encoding secreted forms of receptors

Page 51: Sequence analysis, pathogenicity and cytokine gene

40

for cytokines. On the other hand, adenoviruses modulate cytokine expression by

encoding intracellular proteins that counteract TNF-α (Lesokhin et al., 2002; Wold et al.,

1994). Moreover, E1A and E3 gene products down-regulate transcription of IL-8

cytokine (Lesokhin et al., 2002). Zhou et al. (2007) showed that the porcine adenovirus

type 3 E1Blarge protein inhibits the expression of the IL-8 cytokine. A similar observation

has also been reported in vitro in Marek’s disease lymphoblastoid cell line (MSB-1),

where treatment with M. gallisepticum suppresses IL-8 gene expression (Lam, 2004).

However, Booth and Metcalf (1999) reported type-specific induction of IL-8 in human

pulmonary epithelial cells upon infection with HAdV-7. It has been suggested that

Ras/Raf/MEK/Erk pathway is necessary for the HAdV-7 induction of IL-8 in lung

epithelial cells and lung tissue (Alcorn et al., 2001). Wu et al. (2006) showed that HAdV-

7 internalization is calcium dependant and is essential for IL-8 induction following

infection. Additionally, HAdV-7 infection initiate IFN-γ inducible protein 10 (IP-10)

secretions from macrophages and epithelial cells (Wu et al., 2010). IL-8 is a potent

chemotactic cytokine for neutrophils and T lymphocytes and is produced by macrophages

and other cell types such as epithelial cells.

3.2. Adaptive immunity

Immune mechanisms mediated by adaptive immunity are divided in antibody

(humoral) immunity and cell-mediated immunity. Humoral immunity is provided by

antibodies produced in response to antigen exposure. It is specific to an antigen and

comes in two forms: passive and active immunity. Passive immunity may be either

naturally or artificially acquired. Naturally acquired immunity refers to antibody-

Page 52: Sequence analysis, pathogenicity and cytokine gene

41

mediated immunity conveyed to an embryo by its mother during egg formation or via the

colostrum in mammals. Artificially acquired passive immunity is short-term

immunization achieved by the transfer of blood serum containing antibodies, either orally

or through injection. Passive immunity provides immediate protection, but the immune

system is not stimulated and the chick will not produce its own antibodies and does not

develop memory. On the other hand, active immunity is obtained from the chick’s own

immune system and is subdivided into naturally and artificially acquired immunity.

Naturally acquired active immunity occurs through contact with a disease-causing

pathogen, while artificially acquired active immunity is developed through vaccination.

Active immunity has the ability to remember specific pathogens, resulting in quicker and

more effective responses to repeated exposure. Immunoglobulins (Ig), the key proteins of

the specific immune response, are produced by plasma cells and lymphocytes. Their basic

unit structure contains four polypeptide chains, two heavy (H) and two light (L) chains,

that form the monomeric unit (H2L2) (Litman et al., 1993). Using immunocytochemical

and genetic techniques, five classes of mammalian Ig have been identified: IgA, IgD,

IgE, IgG, and IgM, while avian Igs have been classified into three classes: IgA, IgY

(IgG), and IgM. Avian IgA is the predominant form of antibody (Ab) activity in body

secretions. IgY is the major avian Ig with similarities to both mammalian IgG and IgE.

IgY is produced after IgM in the primary antibody response and is main isotype produced

in the secondary response (Warr et al., 1995). Chicken IgM is the first isotype to be

expressed after initial exposure to a new antigen and is structurally and functionally

homologous to its mammalian counterpart (Da Silva Martins et al., 1991).

Page 53: Sequence analysis, pathogenicity and cytokine gene

42

Infection with fowl adenoviruses elicits a strong Ab response against viral

proteins. Chickens infected with FAdVs seroconvert at 1 week p.i. and the highest titers

of anti-FAdV Abs are detected 3 weeks after inoculation (Ojkic and Nagy, 2003; Maiti

and Sarkar., 1997). However, the Ab response depends on the dosage and the route of

inoculation (Ojkic and Nagy, 2003). Moreover, fowl adenovirus-specific Abs are

commonly found in breeder and layer flocks and exposure to multiple serotypes is

reported by several authors (Adair et al., 1980; Adair and Fitzgerald, 2008). Serological

monitoring of Ontario broiler breeder flocks affected with IBH caused by an FAdV-2 at

an early age showed that seroconversion occurs within 4 weeks and birds are resistant to

reinfection with the same serotype 8 weeks after the primary infection (Philippe et al.,

2007).

On other hand, the antigen specific component of the cell-mediated immunity is

the T cell. Different types of T cells are differentiated based on their functional

capabilities and cell surface markers. T cells which play regulatory role in adaptive

immunity are usually referred to as T helper (Th) cells and express CD4 molecules on

their surface. Both avian and mammalian adaptive immunity depend on CD4+ Th cells

(Arstila et al., 1994). However, not much is known about this part of immune response

with respect to adenovirus infection.

Page 54: Sequence analysis, pathogenicity and cytokine gene

43

Hypothesis and objectives 1. Hypotheses

1.1. Hypothesis 1

Viruses representing the fowl adenovirus species group C (FAdV-C) (FAdV-4 field

isolate) and fowl adenovirus species group E (FAdV-E) (FAdV-8 field isolate) are not

primary pathogens.

1.2. Hypothesis 2

Specific sequences of the fiber gene contribute to virulence of FAdVs.

2. Objectives

In the last several decades, IBH has been reported in almost every country around

the world (Philippe et al., 2005; Gomis et al., 2006; Goodwin et al., 1993; Young et al.,

1972; Saiffudin and Wilks, 1991; Erny et al., 1991). Although fowl adenoviruses are

associated with numerous clinical conditions, their primary role in disease is still not

understood. Therefore, my first and second objective were focused on determining the

pathogenicity of two Ontario isolates, one isolated from a flock with clinical signs of

IBH, with IBH as the final diagnosis, while another virus was isolated from a flock

without clinical signs of the disease. Due to an increased interest in adenoviruses as

vaccine and gene delivery vectors, it is important to understand the molecular aspects of

infection and the host responses to them (Corredor and Nagy, 2010b). Therefore, the

following part of the research was focused on cytokine gene expression following FAdV

infection. This is the first study related to FAdV-associated cytokine gene expression and

Page 55: Sequence analysis, pathogenicity and cytokine gene

44

particular cytokines (IFN-γ, IL-18, IL-10, and IL-8) were selected because of their

important roles against intracellular pathogens. IFN-γ is involved in most stages of the

inflammatory and immune response and is the essential cytokine in the host defense

against various pathogens (Sekellick et al., 1994; Digby and Lowenthal, 1995). IL-18 is a

potent inducer of IFN-γ and is important for viral clearance (Kaiser et al., 2003). IL-10 is

a regulatory cytokine which has both an immunosuppressive and an immunostimulating

effect on cell mediators of adaptive immunity (O’Garra and Vieira, 2007). IL-8 is a

potent chemotactic cytokine critical to the process of inflammation (Oppenheim et al.,

1991).

The third objective focuses on fiber protein, a major constituent of the adenovirus

capsid involved in facilitating virus entry. It has been reported that FAdV-8 virulence is

determined by a difference in fiber protein alone (Pallister et al., 1996). Variations in

fiber gene sequences have also been implicated in the differences in tissue tropism and

virulence for the canine adenoviruses (Rasmussen et al., 1995). Yet the role of fiber

protein in the virulence of fowl adenoviruses is not well understood. Therefore, the last

objective examines the possible role of the fiber gene in FAdV virulence by analysis of

sequences from pathogenic and non-pathogenic FAdV-11 and FAdV-8 field isolates.

2.1. To determine the pathogenicity of a fowl adenovirus serotype 8 (FAdV-8) isolate in

specific pathogen free (SPF) layer chickens following oral and intramuscular

inoculations.

Page 56: Sequence analysis, pathogenicity and cytokine gene

45

2.1.1. To sequence and analyze the genome of FAdV-8 and study the relationship with

other fowl adenovirus genomes.

2.1.2. To analyze cytokine IFN-γ, IL-10, IL-18, and IL-8 gene expression in SPF

chickens after infection with FAdV-8.

2.2. To determine the pathogenicity of a fowl adenovirus serotype 4 (FAdV-4) isolate in

SPF layer chickens following oral and intramuscular inoculations.

2.2.1. To analyze cytokine IFN-γ, IL-10, IL-18, and IL-8 gene expression in SPF

chickens after infection with FAdV-4.

2.3. To analyze and compare the fiber gene sequences of FAdV isolates from field cases

in order to determine differences that may contribute to virulence.

Page 57: Sequence analysis, pathogenicity and cytokine gene

46

CHAPTER 2 Pathogenicity and complete genome sequence of a fowl adenovirus serotype 8 isolate

Helena Grgić1, Dan-Hui Yang2 and Éva Nagy1*

1Department of Pathobiology, Ontario Veterinary College, University of Guelph, Guelph, Ontario, N1G 2W1, Canada. 2Department of Molecular and Cellular Biology, University of Guelph, N1G 2W1 Guelph, Ontario, Canada

*Corresponding author: Éva Nagy E-mail: [email protected] Tel: +1 519 824 4120 ext 54783 FAX: +1 519 824 5930 The GenBank accession number of the sequences reported in this paper is GU734104.

Author’s contributions

HG designed and performed experiment, carried out analysis and wrote the paper. D-HY

provided guidance for qPCR. EN provided guidance during the research and critically

revised manuscript.

Published in Virus Research. 2011. 156:91-97.

Page 58: Sequence analysis, pathogenicity and cytokine gene

47

Abstract

In this study we determined and analyzed the complete nucleotide sequence of the

genome of a fowl adenovirus serotype 8 (FAdV-8) isolate and examined its pathogenicity

in chickens. The full genome of FAdV-8 was 44,055 nucleotides in length with a similar

organization to that of FAdV-1 and FAdV-9 genomes. No regions homologous to early

regions E1, E3 and E4 of mastadenoviruses were recognized. Along with FAdV-9,

FAdV-8 has only one fiber gene and with regard to sequence composition and genome

organization, FAdV-8 is closer to FAdV-9 than to FAdV-1. Moreover, our findings

suggest that FAdV-1 of species Fowl adenovirus A as the current type species despite its

historical priority is not representative of the genus Aviadenovirus, and that FAdV-8 or

FAdV-9 in species Fowl adenovirus E and Fowl adenovirus D, respectively, would be

more suitable for that designation. Additionally, pathogenicity of FAdV-8 was studied in

specific pathogen free chickens following oral and intramuscular inoculations. Despite

lack of clinical signs and pathological changes, virus was found in tissues and cloacal

swabs of all birds with the highest viral copy numbers present in the cecal tonsils. The

highest virus titers in the feces for orally and intramuscularly inoculated chickens were

recorded at day 10 and 3 post-infection, respectively.

Keywords

FAdV-8 complete nucleotide sequence; pathogenesis; virus shedding; highest viral copy

numbers in the cecal tonsils

Page 59: Sequence analysis, pathogenicity and cytokine gene

48

Abbreviations

FAdV-8, fowl adenovirus 8; d.p.i., days post-infection; ORF, open reading frame; pfu,

plaque forming unit; SPF, specific pathogen free

Page 60: Sequence analysis, pathogenicity and cytokine gene

49

1. Introduction

Adenoviruses are in the family Adenoviridae and grouped into five genera. While

mastadenoviruses and aviadenoviruses infect mammals and birds respectively, viruses in

the genera Atadenovirus and Siadenovirus have a broader host range. The genus

Aviadenovirus currently comprises five fowl (Fowl adenovirus A-E), one falcon (Falcon

adenovirus A) and one goose adenovirus (Goose adenovirus) species (ICTV, Virus Taxonomy:

2009 Release). The type species is Fowl adenovirus A and is represented by fowl

adenovirus 1 (FAdV-1) (Benkő et al., 2005).

The first molecular classification of fowl adenoviruses was based on restriction

fragment length polymorphisms (RFLP) of their DNA genome (Zsák and Kisary, 1984).

The genomes of fowl adenoviruses (FAdVs) are larger than that of the human

adenoviruses (HAdVs), in spite of the fact that open reading frames (ORFs) for a few

structural proteins such as V and IX, and early regions E1, E3 and E4 responsible mainly

for host cell interaction and immune modulation in mastadenoviruses, are absent in

FAdVs. The complete nucleotide sequences are available for the genomes of FAdV-1

(CELO virus) and FAdV-9 (A2-A strain) from the species Fowl adenovirus A and Fowl

adenovirus D, respectively. The FAdV-9 genome (45,063 bp) is 1,259 bp longer than

that of FAdV-1 (43,804 bp) (Ojkic and Nagy, 2000; Chiocca et al., 1996). Partial

sequences are also published for the left and right ends of the genomes of FAdV-2,

FAdV-4 and FAdV-10 (Corredor et al., 2006; 2008), and for the late region of FAdV-4

and FAdV-10 (Mase et al., 2010; McCoy and Sheppard, 1997; Sheppard et al., 1998).

Although it is known that the highest degree of conservation among all adenovirus

Page 61: Sequence analysis, pathogenicity and cytokine gene

50

genomes is in the middle region that encodes the structural proteins only a few complete

genome sequences are available for FAdVs.

FAdVs are common infectious agents of poultry although most of them are

subclinical or associated with mild clinical signs (McFerran and Smyth, 2000). However,

some FAdV strains are associated with a variety of diseases, such as inclusion body

hepatitis (IBH) which is diagnosed in almost every part of the world (Adair and

Fitzgerald, 2008). The disease affects mostly 3 to 7 weeks old meat-type chickens

(McFerran et al., 1976; Ojkic et al., 2008a,b), although Bickford (1974) and Hoffmann et

al. (1975) also reported cases in pullets and replacement layers, respectively. IBH

outbreaks are characterized by sudden onset of mortality which usually ranges from 2-

10%. (Adair and Fitzgerald, 2008). The varying degree of mortality correlates to the

pathogenicity of the virus, secondary infections and susceptibility of the chickens. In

Canada, the most commonly isolated FAdVs belong to serotypes 2, 7, 8, and 11 (Ojkic et

al., 2008a,b). Following initial multiplication the virus spreads to virtually every organ,

although the main sites of virus replication are the respiratory and alimentary tracts

(Cook, 1983; Saifuddin and Wilks, 1991; Hess, 2000).

The goals of this study were to (1) determine the complete nucleotide sequence of

the genome of a fowl adenovirus serotype 8 (FAdV-8) isolate from an IBH outbreak

representing species Fowl adenovirus E, (2) investigate its genomic organization and

relationship with other adenoviruses (AdVs) and (3) study its pathogenicity in specific

pathogen free (SPF) chickens.

Page 62: Sequence analysis, pathogenicity and cytokine gene

51

2. Materials and methods 2.1. Virus and cells

The serotype 8 FAdV was isolated by the Animal Health Laboratory (AHL),

University of Guelph from an IBH outbreak that occurred in an Ontario broiler farm. The

virus was plaque purified in our laboratory. Virus propagation and titration were done in

a chicken hepatoma cell line (CH-SAH) as described by Alexander et al. (1998).

2.2. Nucleic acid preparation

CH-SAH cells were infected at a multiplicity of infection (m.o.i.) of 0.5. Cells and

supernatants were collected when an extensive cytopathic effect was seen. Virus

purification and viral DNA preparation were performed as described (Ojkic and Nagy,

2001).

2.3. Cloning, sequencing and sequence analysis

The FAdV-8 DNA was digested with BamHI and the fragments were separated in

0.8% agarose gel. DNA from the desired fragments was extracted and cloned into BamHI

digested pBluescript (SK-). Sequencing was first carried out with universal forward and

reverse primers and then completed by primer walking. Multiple nucleotide sequence

alignments using Vector NTI Advance TM 11 were done to identify the conserved regions.

The sequence data were subjected to BLAST analysis with the GenBank database to

determine the nucleotide sequence homologies and amino acid identities and similarities.

ORFs encoding polypeptides over 50 amino acids were identified by the ORF finder from

Vector NTI Advance TM 11.

Page 63: Sequence analysis, pathogenicity and cytokine gene

52

2.4. Pathogenicity assessment of FAdV-8 in SPF chickens

A trial was conducted to assess the pathogenic potential of the virus. One hundred

white Leghorn SPF chickens were hatched at the Arkell Poultry Station from eggs

obtained from the Canadian Food Inspection Agency (CFIA), Ottawa. The birds were

housed at the University Isolation Unit according to the Animal Care Committee of the

University of Guelph in accordance with the Guide to the Care and Use of Experimental

Animals of the Canadian Council on Animal Care and received food (21% unmedicated

starter ration) and water ad libitum. At 9 days of age the chickens were tagged with wing

bands (Ketchum, ON) and sera were collected from all of them. Ten-day-old chicks were

randomly divided into four groups: Group I (FAdV-8)-oral, Group II (FAdV-8)-

intramuscular (im), Group III (control)-oral and Group IV (control)-im. Groups I and II

had 30 birds each and negative control groups III and IV had 20 birds each. Chicks were

each inoculated with 2 x 108 plaque forming units (pfu) of the virus. The control birds

received PBS. All chickens were re-inoculated at 14 days of age with the same dose of

virus.

The birds were observed daily (3x) for clinical signs. Four chicks from treatment

groups I and II and three chicks from the control groups were drawn randomly for

necropsy at each of 0, 3, 5, 7, 14, 21, and 28 days post-infection (d.p.i.) (except day 0

when only two chicks from the controls were euthanized). The chickens were euthanized,

necropsied, examined for the presence of gross lesions and tissues from thymus, lung,

heart, liver, kidney, bursa of Fabricius and cecal tonsils were collected. Each organ was

sectioned into three portions: one was placed into 10% formalin for histology, the second

Page 64: Sequence analysis, pathogenicity and cytokine gene

53

was stored at -70°C for virology and the third was stored at -70°C for qPCR. Cloacal

swabs were taken from all birds at 0, 3, 5, 10, 14, 21, 28 d.p.i. and their virus titers were

determined in CH-SAH cells. In addition, chickens were bled at weekly intervals, 0, 7,

14, 21, 28 d.p.i. and serum samples were tested for FAdV-specific antibodies (Abs) by

enzyme-linked immunosorbent assay ELISA as described (Ojkic and Nagy, 2003).

2.5. Quantitative real time PCR (qPCR) analysis

To determine the viral load in tissues, real time PCR using SYBR Green as

intercalating dye was developed. The FAdV-8 ORF-1 A/B gene (Corredor et al., 2006)

was used as an indicator for the presence of viral DNA with primers; forward primer:

5’-AAATGGTAAACGCGTGGGATC-3’ and reverse primer:

5’-TTCTCCGTCTCCGATCTGG-3’. The 20 l qPCR reaction containing 1 l (10 pmol)

of each primer, 10 l of QuantiTec SYBR Green I PCR Master Mix (Qiagen), 2 l DNA

and 6 l nuclease free water, was prepared using the computerized automated liquid

handling robotics CAS-1200 (Corbett Research). The qPCR program was first run at 15

min at 95C to activate the Taq polymerase, followed by 40 cycles of 95C for 20 sec,

57C for 15 sec, 72C for 20 sec, with an acquiring step at 75C for 15 sec, using the

Corbett Research Rotor-Gene 6000 System. Quantitative measurement of the PCR

product was carried out by incorporation of the SYBR Green I fluorescent dye. To

establish the standard curves, 10-fold serial dilutions (10-1 to 10-7) of FAdV-8 genomic

DNA starting at 12.2 ng/l (2.4 x 108 copies/l) were made and used as qPCR templates.

Melting curve analysis was performed after each run to monitor the specificity of the

Page 65: Sequence analysis, pathogenicity and cytokine gene

54

PCR products. Negative control using DNA extracted from uninfected chicken and no

template control (NTC) were routinely included in each experiment.

2.6. Statistical analysis

Kruskal-Wallis test was performed to evaluate whether viral copy numbers

differed among tissues. Following statistically significant P-value, pair-wise comparison

was performed between different tissues. P-values of 0.05 were considered to be

statistically significant.

3. Results 3.1. Genome size and organization of FAdV-8

The full genome of FAdV-8 was found to be 44,055 base pairs (bp) with a G+C

content of 58%. The position and size of 46 ORFs with potential to encode polypeptides

of at least 50 amino acid (aa) residues are listed in Supplementary Table 2.1. A schematic

of the genome map is presented in Fig. 2.1. The GenBank accession number of the

sequence reported in this paper is GU734104.

The sequences of the left and right ends of this FAdV-8 isolate were determined

earlier in our laboratory (Corredor et al., 2006; 2008). Briefly, the first 6 kb at the left

terminal region of the genome have no significant sequence homology to known

mastadenoviral sequences either at the DNA or protein level. Homologues to the first 10

ORFs (ORF0 to IVa2) in this region described previously for FAdV-1 and FAdV-9 were

present in the genome of FAdV-8 (Corredor et al., 2006). None of the leftward oriented

ORFs at the right end of FAdV-8 showed homology to identified E4 sequences from

Page 66: Sequence analysis, pathogenicity and cytokine gene

55

other AdVs. The FAdV-8 genome encodes three small ORFs (ORF 38652-39047, ORF 39052-

39486, and ORF 39291-39551) which are homologoues to ORF8 and are unique to FAdVs

only. ORF8 encodes the highly conserved (82-95%) GAM-1 protein found in all fowl

adenoviruses. The FAdV-8 genome has a short region of repeated sequences. The repeat

region consists of two consecutive 51 bp-long units, located within ORF19 at nts 35607-

35708, with no homology to the repeats of FAdV-9 TR-1 and TR-2 (Ojkic and Nagy,

2000; Corredor et al., 2008).

The central portion of the FAdV-8 genome, from IVa2 gene at position nt 6208 on the

left strand toward the fiber gene at position nt 30863 on the right strand showed a similar

organization to the central region of genomes of other adenoviral genera. The % identities

of 17 of those proteins are summarized in Table 2.1. Three proteins crucial for viral DNA

replication, DNA polymerase (DNA pol), terminal protein precursor (pTP) and DNA

binding protein (DBP) encoded by E2 were well conserved when compared with their

FAdV-1 and FAdV-9 counterparts. The percentages of aa identity between FAdV-8 and

FAdV-1 DNA pol, pTP and DBP were 66.2%, 59.8% and 58%, respectively. Protein

sequence identities between DNA pol, pTP and DBP of FAdV-8 and FAdV-9 were

79.5%, 74.3% and 73.7%, respectively. The intermediate gene pIVa2 of FAdV-8 showed

64.9% and 87.3% aa sequence identity to its FAdV-1 and FAdV-9 counterparts, whereas

identity to pIVa2 sequences from other AdVs varied from 73% to 94%. For each FAdV-8

protein the highest % identity was observed with the FAdV-9 homologue. Another

intermediate gene, the gene encoding the minor capsid protein IX that functions as capsid

cement could not be identified in FAdV-8. The FAdV-8 equivalent of early region 3 (E3),

Page 67: Sequence analysis, pathogenicity and cytokine gene

56

if it exists, does not seem to be in the same location as that in mastadenoviruses, between

the protein VIII and fiber protein genes. This intergenic region in FAdV-8 was only 253

nts in length and contained one ORF partially overlapping with the protein VIII gene.

The predicted protein sequence had no identity with any described E3 proteins. However,

this ORF had 62% and 70% identity with U-exon of FAdV-1 and FAdV-9, respectively.

In the FAdV-1 genome there are 476 nts between the pVIII and fiber genes while, in

FAdV-9 there are only 254 nts.

AdV late genes are expressed after the onset of viral DNA replication and encode

mostly structural proteins. A region located between nts 8101 and 8386 had 90% identity

to FAdV-9 major late promoter (MLP) and a TATA box was also identified between nts

8360 and 8365. The late genes of FAdV-8 (IIIa, penton base, hexon, pVI and pVIII) were

in the same order as seen for other aviadenoviruses and mastadenoviruses. The genes for

core proteins pVII and pX were also identified on the FAdV-8 genome whereas

sequences for pV, the third core protein found in mammalian adenoviruses could not be

identified. The hexon protein, the major component of the virion, carries the type, group

and subgroup specific antigenic determinants of FAdVs. Nevertheless, a very high

percent identity was noted for the hexon of FAdV-8, 81.2% and 88.3% with FAdV-1 and

FAdV-9, respectively. The FAdV-8 viral protease essential for virus maturation and

cleavage of the precursors of protein VI, VII, VIII, IIIa, X and pTP showed 39% and 37%

homology to its counterparts in FAdV-1 and FAdV-9 genomes, respectively. The 100k

protein required for assembly of hexon trimers and late viral translation was found to

share 50.9% and 53.7% homology with their counterparts in FAdV-1 and FAdV-9,

Page 68: Sequence analysis, pathogenicity and cytokine gene

57

respectively. FAdV-8 has only one fiber gene, similar to FAdV-9, and aa identity

between these two fibers was 53.4%. On the other hand, protein identity with FAdV-1

short fiber gene was 29% while, with the long one was only 17.4%. (Table 2.1). FAdV-8

fiber protein was comprised of 524 aa and the ORF started at nt 30863. Similar to other

adenovirus fiber proteins, it was organized into three regions: tail, shaft and head. The tail

region was 71 aa including the poly(G) stretch in the tail domain that may also form a

connection between the tail and the shaft without belonging to either. The nuclear

localization sequence (NLS) KRAR of the human adenovirus 2 (HAdV-2) fiber (Hong

and Engler, 1996) was not detected in FAdV-8, although the identified RKRP (residues

17-20) could have an NLS function. In many mammalian adenovirus fibers there is a

conserved VYPY sequence which is probably involved in the interaction with the penton

base. The equivalent sequence in FAdV-8 fiber tail was VYPF (residues 53-56). Because

of the absence of the conserved TLWT in FAdV-8 fiber, the actual start of the head

domain could not be determined with certainty.

The short rightward-oriented virus-associated (VA) RNAs transcribed by RNA

polymerase III and located between the pTP and 52K genes in the genome of most

mastadenovirus species could not be identified for the FAdV-8 genome.

3.2. Pathogenicity of FAdV-8 (clinical signs, gross lesions and histology)

Clinical signs or pathologic changes of inclusion body hepatitis were not seen in

groups of chicks after im and/or oral administration of virus. Histopathological

examination revealed mononuclear cell infiltration, degeneration and necrosis of

Page 69: Sequence analysis, pathogenicity and cytokine gene

58

hepatocytes in only three cases. The classical signs of IBH characterized by intranuclear

inclusion bodies in hepatocytes were not present in any of the inoculated chickens.

3.2.1. Viral genome copy number in tissues

Quantitative real-time PCR was employed to establish viral genome copy

numbers in liver, bursa of Fabricius and cecal tonsil of infected birds. No viral DNA was

detected in any tissues from any groups before inoculation and in mock-infected

chickens. The results over the entire study period are summarized in Table 2.2.

Viral DNA was detected in 95.1% of tissue samples collected at 3, 5, 7 14, 21 and 28

d.p.i. Of the 4.9% of tissue samples that were negative all were from the bursa of

Fabricius. Table 2.2 shows the arithmetic means of viral genome copy numbers of these

birds in each group of birds inoculated by different inoculation routes.

The results of the Kruskall-Wallis test indicated significant differences in orally

inoculated birds (P<0.001) and in im-inoculated birds (P=0.0012), and the im-inoculated

group of chickens had higher viral copy numbers. The cecal tonsil was the organ with the

highest number of viral copies through the study period irrespective to inoculation route

followed by liver and the bursa of Fabricius.

3.2.2 Virus titers in feces

No virus was present in cloacal swabs in any groups of chickens before

inoculation and in the mock infected group at all time. The virus titers in feces are given

in Table 2.3. The results showed that the difference in titers between chickens inoculated

orally and im was not statistically significant (P=0.88), when it was tested by 2-sample

Page 70: Sequence analysis, pathogenicity and cytokine gene

59

Wilcoxon test. The highest virus titer in the orally inoculated group was found at day 10

p.i. (1.6x104) while in the im inoculated group it was at day 3 p.i. (3.1x104). Virus titer

began to decline in the orally inoculated group after 10 d.p.i. while, in the im group

decline in virus titer was found after 3 d.p.i. Chickens shed the virus through the entire

study period in both oral and im group.

3.2.3. Antibody response

Abs against FAdV-8 were not detected in any groups before inoculation (0 d.p.i)

and in the mock-infected group at all times. Ab response to viral proteins appeared at 14

d.p.i. with significant differences between inoculated groups and negative control. Ab

titer was slightly lower at 21 d.p.i. but difference in titer between inoculated groups and

negative controls was significant. The differences in titers between birds inoculated im

and orally were significant (P<0.001). Chickens inoculated im had higher titers than

chickens inoculated orally (Fig. 2.2).

4. Discussion

We determined and analyzed the full genome sequence of a FAdV-8 isolate, the

first virus of species Fowl adenovirus E. So far the complete genome nucleotide

sequences of only FAdV-1 (CELO virus) and FAdV-9 (strain A-2A) have been reported

(Chiocca et al., 1996; Ojkic and Nagy, 2000). Nucleotide sequences of the right and left

ends of FAdV genomes representing viruses of four species groups had already been

determined and compared to those of the sequences of FAdV-1 and FAdV-9 (Corredor et

al., 2006; 2008).

Page 71: Sequence analysis, pathogenicity and cytokine gene

60

The genome of FAdV-8 was 44,055 nt in length, with a G+C content of 58% compared

to 54% for FAdV-1 and 54.3% FAdV-9. There were 46 ORFs with potential to encode

polypeptides that were at least 50 aa residues long. Based on this analysis the FAdV-8

genome is 251 nt longer than that of FAdV-1 and 1008 nt shorter than that of FAdV-9

genome. There are several reasons for such differences in size. First, two sets of tandemly

repeated sequences (51 bp long) identified within ORF19 in the FAdV-8 genome, were

much shorter than the repeat TRs found in FAdV-9. The FAdV-9 genome contains two

blocks of tandem repeats located at the right end of the viral genome and are incorporated

within a region rich in ORFs. The shorter region (TR-1) contains five 33 bp long repeats

while the longer region (TR-2) contains thirteen 135 bp long repeats (Cao et al., 1998;

Ojkic and Nagy, 2000). The function of these repeats is unknown, although it has been

reported that the longer subunit is dispensable for in vitro virus replication (Ojkic and

Nagy, 2001). Interestingly tandemly repeated sequences have also been documented in

vaccinia virus and Herpes simplex virus type 1 (HSV-1) (Baroudy and Moss, 1982;

Umene et al., 1984.) Secondly, the short, partially double-stranded VA RNAs transcribed

by RNA polymerase III involved in translational control and inhibition of the interferon

response (Mathews and Shenk, 1991) were absent in the FAdV-8 genome and a similar

finding was reported for the FAdV-9 genome (Ojkic and Nagy, 2000). Currently

identified VA RNAs of mammalian AdVs are transcribed rightward, and human AdVs

contain one or two VA RNA genes located between the pTP and 52k ORFs. FAdV-1 and

EDS viruses have only one VA RNA gene, which is much shorter and located near the

right end of the genome. Although they share a similar position on the genome, the

FAdV-1 VA RNA gene is transcribed leftward which is opposite to its EDSV counterpart

Page 72: Sequence analysis, pathogenicity and cytokine gene

61

(Hess et al., 1997; Chiocca et al., 1996). Thirdly, the complete DNA sequence of FAdV-

1 revealed that FAdV-1 is the only fowl adenovirus sequenced to date with two fiber

genes. The actual fiber ORF sizes differed greatly in length, 793 and 412 aa, respectively

(Hess et al., 1995; Chiocca et al., 1996). In contrast to FAdV-1, only one fiber gene was

identified in the corresponding region of the FAdV-8 genome. The fiber protein is the

major constituent of the outer capsid involved in virus entry into permissive cells.

Interestingly, an equivalent sequence motif in the FAdV-8 fiber tail VYPF was also

found in the fiber tail of OAV287 (Vrati et al., 1995) and EDSV (Hess et al., 1997).

Moreover, based on a conserved TLWT sequence it was possible to define where the

knob region of the fiber begins (Xia et al., 1994). Although, such a motif is conserved in

many adenoviruses, it is missing in the FAdV-8 fiber. Absence of the motif has been also

reported for OAV287 and EDSV (Vrati et al., 1995; Hess et al., 1997). Lastly, in most

adenoviruses including HAdV-5 (Chroboczek et al., 1992), canine (CAdV-1) (Morrison

et al., 1997), bovine (BAdV-3) (Mittal et al., 1992), porcine (PAdV-3) (Reddy et al.,

1995) and murine (MAdV-1) (Beard et al., 1990) adenoviruses, the E3 region is found

between the pVIII and fiber protein genes. E3 proteins are involved in the inhibition of

presentation of antigenic protein via the major class I histocompatibility complex (MHC-

I). The E3 genes of HAdVs are involved in the modulation of the host’s immune response

including cytotoxic T lymphocyte (CTL)-induced apoptosis (Wold and Gooding, 1991).

The E3 region represents one of the most commonly used insertion sites in the adenovirus

gene delivery system (Russell, 2000). However, deletion of this region results in an

increased inflammatory response when compared to the wild type virus counterpart

(Braithwaite and Russell, 2001). The size of the E3 region varies considerably among the

Page 73: Sequence analysis, pathogenicity and cytokine gene

62

mastadenoviruses and the proteins encoded by this region are also different. Lack of an

E3 region has been reported in ovine (OAV287; Vrati et al., 1995) and several avian

adenoviruses, such as EDSV (Hess et al., 1997), FAdV-1 and FAdV-9 (Chiocca et al.,

1996; Ojkic and Nagy, 2000).

Our data demonstrated that genes in the central region of the genome were in the same

order as in every adenovirus studied to date. These include the genes of structural

proteins and the viral protease gene coded on the r strand.

In an investigation of any disease outbreak, commonly asked questions are about the

pattern of infection and how transmission occurs. From the available evidence it seems

that most of the strains of FAdVs follow the same pattern of infection. Following initial

multiplication there is probably viraemia that results in virus spread to virtually all

organs. The main sites of virus replication appear to be respiratory and alimentary tracts

(Saifuddin and Wilks, 1991). In our study the organ with the highest viral load was cecal

tonsil in both orally and im inoculated chickens. Ojkic and Nagy (2003) demonstrated

that replication and distribution in chicken tissues can be markedly influenced by route of

virus inoculation. However, horizontal transmission of FAdVs occurs through contact

with nasal and tracheal excretions and feces, the latter of which contains the highest virus

titers (Grgic et al., 2006). Therefore, we compared im and the more natural oral routes of

inoculation and the results showed that the im inoculated group had higher concentrations

of viral copies in general. Furthermore, virus was isolated from cloacal swabs up to 28 d.

p.i. and that was probably the longest recorded shedding period for fowl adenoviruses to

Page 74: Sequence analysis, pathogenicity and cytokine gene

63

date (Corredor and Nagy, 2010a). The highest virus titers in the feces for the orally and

im-inoculated birds were recorded at days 10 and 3 p.i., respectively. Virus shedding

could be indirectly correlated to the Ab response of chickens. Indeed the development of

neutralizing Abs coincides with cessation of virus excretion (Clemmer, 1972) as was

possibly the case in this study. The lower Ab response in orally-inoculated chickens, the

lower viral genome copy number in tissues in the oral group and the lower virus titer in

the feces of orally-inoculated chickens could be attributed to the oral route of inoculation

and the amount of virus that indeed infected chickens. It is important to note, that

although not the natural route of viral infection, intramuscular inoculation ensures that all

birds receive the same amount of virus. That might be the reason why im-inoculated

chickens exhibited virus recovery and Ab results with higher values.

Our original hypothesis was that this virus was pathogenic since it was isolated from a

flock with clinical signs of inclusion body hepatitis. Thus a complete genome sequence

was determined and a pathogenicity study was performed. However, based on the

pathogenicity data we concluded that under the conditions of this study this isolate was

not pathogenic.

In summary our results for FAdV-8 are concordant with findings reported for the other

genomes of fowl adenoviruses. Along with FAdV-1 and FAdV-9, FAdV-8 lacked

mastadenoviral E1, E3 and E4 region, and sequences for protein V and IX. Similar to

FAdV-9, FAdV-8 has only one fiber gene. We also found that with regard to sequence

composition and genome organization, FAdV-8 is closer to FAdV-9 than to FAdV-1.

Page 75: Sequence analysis, pathogenicity and cytokine gene

64

Genetic distance between species E (FAdV-8) and D (FAdV-9) also indicates a closer

relationship between them and this observation is supported by the high identities among

their ORFs (Table 2.1). Our findings suggest that FAdV-1 of species Fowl adenovirus A

as the current type species is, despite its historical priority not representative of the genus

Aviadenovirus, and that FAdV-8 or FAdV-9 in species Fowl adenovirus E and Fowl

adenovirus D, respectively, would be more suitable for that designation. Despite lack of

clinical signs and pathological changes virus was found in tissues and cloacal swabs in

the chickens inoculated with FAdV-8. However, chickens inoculated im had higher

antibody titers than the orally inoculated ones. This information is important in devising

vaccination strategies against inclusion body hepatitis and for developing FAdV-8 as a

recombinant vector for gene delivery or for vaccine application, as we have already

described and discussed for FAdV-9 (Corredor and Nagy, 2010a,b).

Acknowledgements

This work was supported by the Natural Sciences and Engineering Research

Council of Canada (NSERC), the Poultry Industry Council (PIC) of Ontario and the

Ontario Ministry of Agriculture, Food and Rural Affairs (OMAFRA). The authors thank

Mrs. Sheila Watson for her technical help, and the Isolation Unit personnel for their

professional animal care and assistance.

Page 76: Sequence analysis, pathogenicity and cytokine gene

65

Table 2.1. Percent identities/similarities of selected FAdV-8 proteins from this study and their homologues.

Serotype/Species ACCESSION IVa2 DNA pol pTP 52/55kDa pIIIa pIII pVII pX pVI

FAdV-1/A NC001720 64.9/77.9 66.2/75.1 59.8/65.5 67.8/78.1 66.8/77 73.5/81.8 62.4/69.4 55.5/65.6 32.3/43.3

FAdV-9/D NC000899 87.3/91 79.5/84.9 74.3/78.3 82.5/88.1 90.8/95.2 86/89.8 68.8/72.9 77.5/85.8 38.7/44.3

FAdV-4/C ACR54094 _a _ _ _ _ _ _ _ _

FAdV-8/E AAC55302 _ _ _ _ _ _ _ _ _

FAdV- 8/E CFA40 AAF17339 _ _ _ _ _ _ _ _ _

FAdV-10/C AAB88670 _ _ _ _ _ _ _ _ _

HAdV-2/C J01917 30.8/44.8 34.8/48 29.2/42.9 23.6/37 23.6/36.8 40.5/51.4 17.5/22.5 _ 17.4/27.8

Serotype/Species ACCESSION Hexon Protease DBP 100K 33kDa pVIII Fiber/sb Fiber/lc

FAdV-1/A NC001720 81.2/88.5 76.1/86.6 58/70.1 50.9/61 5.6/13.1 68.5/76.2 29/42.7 17.4/25.1

FAdV-9/D NC000899 88.3/92.2 87.1/93.8 73.7/82.1 53.7/61.6 24.7/29.1 84.1/86.6 53.4/67.2 _

FAdV-4/C ACR54094 _ _ _ _ _ _ 39/54.4 _

FAdV-8/E AAC55302 _ _ _ _ _ _ 67.4/72.3 _

FAdV-8/E CFA40 AAF17339 _ _ _ _ _ _ 66.5/69.2 _

FAdV-10/C AAB88670 _ _ _ _ _ _ 36.7/52.5 _

HAdV-2/C J01917 46.8/59.7 39/62.9 26.5/42 26.4/37.6 14.6/23 24.9/37.5 19.2/30.2 _

a= data not available b= short fiber c= long fiber Percent identities and similarities were calculated using FASTA3 (EMBL-EBI).

Page 77: Sequence analysis, pathogenicity and cytokine gene

66

Table 2.2. Numbers of positive tissues and viral genome copy numbers in tissues of chickens inoculated with FAdV-8.

days p.i.

Inoculation route

tissue oral Im

Liver 6.7x102* 4/4§ 2.9x103 4/4

3 Bursa 4.8x100 3/4 9.6x103 4/4

Cecal tonsil 2.6x103 4/4 4.4x103 4/4

Liver 8.6x102 4/4 3.3x104 4/4

5 Bursa 1.3x102 4/4 1.6x105 4/4

Cecal tonsil 7.5x103 4/4 1.5x105 4/4

Liver 2x104 4/4 1.9x102 4/4

7 Bursa 1.3x101 4/4 3.8x102 4/4

Cecal tonsil 6.5x104 4/4 5.5x105 4/4

Liver 1.4x100 4/4 1.2x102 4/4

14 Bursa 6.1x101 3/4 5.8x100 3/4

Cecal tonsil 3.5x102 4/4 5.7x102 4/4

Liver 3.4x103 4/4 4.5x102 4/4

21 Bursa 2.2x100 2/4 7.9x102 3/4

Cecal tonsil 2x103 4/4 1.7x103 4/4

Liver 1.4x101 4/4 2.3x101 4/4

28 Bursa 5.3x100 4/4 1.1x101 3/4

Cecal tonsil 2.9x102 4/4 2.8x102 4/4

* = viral genome copy number/10 µg of total tissue DNA was determined by qPCR § = number of positive tissues/number analyzed tissues

Page 78: Sequence analysis, pathogenicity and cytokine gene

67

Table 2.3. Virus titers in the feces of chickens inoculated with FAdV-8.

Days Inoculation route post-infection Titers (pfu/ml) in swabs

Chickens shedding the virus (%)†

Oral Im 9.4x103±1.7x103 3.1x104±7x103

3 P=0.33 (100) (100) 1.3x104±4.2x103 1.9x104±4.4x103

5 P=0.32 (96) (100)

1.6x104±8x103 9.6x102±3x102

10 P<0.01 (96) (25)

7.9x102±3x102 1x102±3x101

14 P=0.18 (86) (31)

0.8x101±0.8x101 2.9x102±1x101

21 P<0.006 (8) (58)

2.2x101±1x101 3.2x102±1.7x101

28 P<0.003 (22) (85)

†30 chickens at day 0; 26 and 22 chickens at days 3 and 5, respectively; 18

and 14 chickens at days 10 and 14, respectively; 10 and 6 chickens at days 21 and 28, respectively. The virus titers were determined by plaque assay.

P values < 0.05 are statistically significant.

Page 79: Sequence analysis, pathogenicity and cytokine gene

68

Supplementary Table 2.1. Position of 46 predicted genes potentially encoded by the genome of FAdV-8. Gene Strand Location No. of aa G+C content (%) ORF0 r 515-799 95 61 ORF1 r 837-1298 154 61 ORF1A/B r 1493-1705 71 56 ORF1C r 1605-1871 89 55 ORF2 r 1922-2728 269 58 ORF24 l 3470-2901 190 56 ORF14 r 3947-3600 116 60 ORF13 l 4620-4315 102 58 Unknowna r 4891-5328 146 60 ORF12 l 6010-5255 252 55 IVa2 l 7350-6208 381 57 DNA pol l 11027-7347 1227 63 pTP l 13478-11259 740 63 52/55kDa r 13655-14884 410 64 pIIIa r 15017-16636 540 64 pIII r 16796-18454 553 61 pVII r 18492-18728 79 63 pX r 18952-19542 197 65 pVI r 19718-20416 233 66 Hexon r 20531-23374 948 57 Protease r 23405-24034 210 61 DBP l 25542-24142 467 63 100K r 26164-29262 1033 60 33kDa r 28910-29338 143 62 pVIII r 29885-30610 242 63 U-exon l 30864-30565 100 54 Fiber r 30863-32424 524 55 ORF32 l 31431-31204 76 57 ORF22a l 32950-32501 150 59 ORF22b r 33080-32916 55 59 ORF20A l 33638-33084 185 57 ORF20 r 34457-33648 270 53 ORF19 l 36807-34726 694 49 ORF28 r 37926-38243 106 42 ORF29 r 38470-38670 67 59 ORF8a r 38652-39047 132 58 ORF8b r 39052-39486 145 60 ORF8c r 39291-39551 87 61 ORF17 l 40152-39682 157 53 ORF33a r 40106-40300 65 58 ORF33b r 40502-40666 55 58 ORF11 r 40513-41202 230 55 ORF11A l 41553-41344 70 47 ORF23 l 42715-41852 288 55 ORF25a r 43147-43773 209 57 ORF25b r 43248-43706 153 59

The size of the deduced proteins and the G+C content of the individual genes are also shown. a Novel unidentified open reading frame r = rightward transcribed strand l = left transcribed strand.

Page 80: Sequence analysis, pathogenicity and cytokine gene

69

Figure 2.1. Schematic representation of FAdV-8, FAdV-9 and FAdV-1 genomes, members of the genus Aviadenovirus. Pink arrows depict genus common genes, blue arrows show genes conserved in all adenovirus genera, green arrows represent unique genes in FAdV-8 and FAdV-1 genome, red arrows depict genus specific genes and the yellow arrow shows unidentified ORF.

Page 81: Sequence analysis, pathogenicity and cytokine gene

70

Figure 2.2. Antibody response to viral proteins in chickens inoculated with FAdV-8 by the oral and intramuscular routes and mock infected as measured by S/P ratios. Serum samples were collected at 7, 14, 21 and 28 days post-infection. Dark gray and white columns represent im and orally inoculated chickens, respectively. Light gray and black columns represent mock-infected chickens. The error bars correspond to 95% confidence intervals. Significant differences among mock infected and infected birds are indicated by asterisks (*).

Page 82: Sequence analysis, pathogenicity and cytokine gene

71

CHAPTER 3 Key cytokines activated during infection of chickens with a serotype 8 fowl

adenovirus

Helena Grgića, Shayan Sharifa, Hamid R. Haghighia and Éva Nagya,*

aDepartment of Pathobiology, Ontario Veterinary College, University of Guelph, Guelph, Ontario, N1G 2W1, Canada

*Corresponding author: Éva Nagy E-mail: [email protected] Tel: +1 519 824 4120 ext 54783 FAX: +1 519 824 5930

Author’s contributions

HG designed and performed experiment, carried out analysis and wrote the paper. HRH

provided guidance for real-time PCR. SS and EN provided guidance during the research

and critically revised manuscript

Submitted to Virus Research

Page 83: Sequence analysis, pathogenicity and cytokine gene

72

Abstract

This study explored the ability of fowl adenovirus (FAdV) to subvert the host

cells expression of cytokines in response to infection as an important viral mechanism for

immune evasion during infection. The production of certain cytokine mRNA was

quantified by real-time quantitative reverse transcription-PCR (RT-PCR) in spleen, liver

and cecal tonsil during the course of infection of chickens with a serotype 8 FAdV

(FAdV-8). Compared to uninfected chickens, infected birds had higher mRNA

expression of interleukin (IL) 18 and IL-10 in spleen and liver, respectively. Interferon

gamma (IFN-γ) mRNA expressed in spleen and liver of infected chickens was

significantly up-regulated, while the expression of IL-8 mRNA in spleen and liver of

infected chickens was significantly down-regulated. There was no significant difference

between infected and uninfected groups in terms of cytokine gene expression in cecal

tonsil. These results indicate that these four cytokines might play a crucial role in driving

the immune responses following FAdV-8 infection.

Keywords:

Fowl adenovirus serotype 8; interleukins; IFN-γ; IL-18; IL-10; IL-8.

Abbreviations: FAdV-8, fowl adenovirus 8; d.p.i., days post-infection; pfu, plaque

forming unit; SPF, specific pathogen free; IFN-γ, interferon gamma; IL, interleukin;

h.p.i., hours post-infection; IBH, inclusion body hepatitis; MOI, multiplicity of infection;

IRF3, interferon regulatory factor; CAR, coxsackie and adenovirus receptor; PAMs,

pathogen-associated molecular patterns; TLR, toll like receptors; NK, natural killer cells;

DCs, dendritic cells.

Page 84: Sequence analysis, pathogenicity and cytokine gene

73

Highlights:

First study of cytokine gene expression following fowl adenovirus infection of chickens

FAdV-8 infection is associated with up-regulation of IL-18 and IFN-γ mRNA expression

Up-regulation of IL-10 is indicative of an immune evasive mechanism used by FAdV-8

FAdV-8 infection suppresses IL-8 gene expression in chickens

Page 85: Sequence analysis, pathogenicity and cytokine gene

74

1. Introduction

Fowl adenoviruses (FAdVs) are in the family Adenoviridae, genus Aviadenovirus

and are classified into five species, Fowl adenovirus A-E, based on their DNA genomes

and further divided into 12 serotypes based on virus neutralization (Benkő et al., 2005).

FAdVs have a worldwide distribution and some are associated with diseases such

as inclusion body hepatitis (IBH). IBH is mainly seen in meat type chickens 3-7 weeks of

age (Adair and Fitzgerald, 2008). However, some FAdVs cause no or mild clinical signs

and could be isolated from asymptomatic chickens, and are considered ubiquitous in

poultry populations. Clinical signs of IBH are characterized by sudden onset of mortality

that usually ranges from 2-10% in affected flocks. Mortality varies and is closely related

to the pathogenicity of the virus, secondary infection with other pathogens and

susceptibility of the chickens. At gross examination the main lesions are seen in the liver

which is swollen, light brown to yellow, occasionally with hemorrhages and necrotizing

hepatitis. Histologically the basophilic nuclear inclusion bodies present in the hepatocytes

are characteristic of infection (Philippe et al., 2005). Most of the FAdV strains follow the

same pattern of infection, after initial multiplication the virus spreads to virtually all

organs. Viraemia is detected at 24 hours post infection (h.p.i.) and has two peaks, the first

at 2 and the second at 7 day post infection (d.p.i.) (Saifuddin and Wilks, 1991). FAdVs

are transmitted both vertically and horizontally, and the virus is present in all excretions

such as those from the feces, trachea and nasal mucosa. The usual virus transmission

within a flock is via the oral-fecal route (Cook, 1974; Adair and Fitzgerald, 2008; Grgić

et al., 2006).

Page 86: Sequence analysis, pathogenicity and cytokine gene

75

Members of the genus Aviadenovirus have the largest genome (43-45 kb) among

adenoviruses, despite the fact that a few structural protein genes such as V and IX, and

early regions E1, E3 and E4 described in mastadenoviruses are not identified for

aviadenoviruses (Chiocca et al., 1996; Ojkic and Nagy, 2000; Kajan et al., 2010; Griffin

and Nagy, 2011; Grgić et al., 2011).

The early gene products of mammalian adenoviruses are responsible for host cell

interactions and modulation of the immune response including cytokine expression by

encoding proteins, which counteracts TNF-α (Lesokhin et al., 2002) or the virus-encoded

RNA (VA RNA), which prevents interferon-mediated shut-off of protein synthesis (Wold

et al., 1994) among others.

Earlier we determined the pathogenicity of a fowl adenovirus serotype 8 (FAdV-

8) isolate based on clinical signs, pathological and histological examination.

Additionally, viral genome copy numbers in the bursa of Fabricius, cecal tonsil and liver

following infection were evaluated (Grgic et al., 2011). Moreover, virus titer in feces and

antibody response were also determined. Despite a lack of clinical signs and pathological

changes virus was found in all selected tissues and cloacal swabs of all birds. The current

work is a continuation of our efforts towards better understanding the pathogenesis of

FAdV infection with the emphasis on the role of interleukins, which is an unknown area.

Our objectives were to investigate the effects of FAdV infection on the

transcription of a number of avian cytokines, both in vitro and in vivo. First, the dynamics

Page 87: Sequence analysis, pathogenicity and cytokine gene

76

of cytokine gene expression in FAdV-8 infected chicken hepatoma cells (CH-SAH) at

different time points was determined. Then in a chicken experiment interferon gamma

(IFN-γ), interleukin (IL) 10, IL-18 and IL-8 gene expression was followed in spleen, liver

and cecal tonsil at different days after infection with FAdV-8, and compared to that of

uninfected chicken samples by quantitative real-time RT-PCR.

2. Materials and methods 2.1. Virus and cells

FAdV-8 was isolated by the Animal Health Laboratory (AHL) at the University

of Guelph and characterized in our laboratory (Grgić et al., 2011). Virus propagation and

titration were performed in CH-SAH cells as described by Alexander et al. (1998).

CH-SAH cells were maintained in Dulbecco’s Modified Eagle’s Medium/Nutrient

Mixture F-12 Ham (DMEM-F12) with 10% non-heat inactivated fetal bovine serum

(FBS) as described (Alexander et al., 1998). To determine the effect of FAdV-8 on the

induction of cytokines, CH-SAH cells were plated onto 35-mm-diameter Petri dishes and

infected with a multiplicity of infection (MOI) of 10. At 1, 4, 8, 10, 12 and 24 hours post-

infection (h.p.i.), the cells were harvested and processed for RNA extraction. Total RNA

was prepared from triplicate cultures at each time point using TriZol (Invitrogen, Canada

Inc., Burlington, ON, Canada) as per the manufacturer’s protocol.

2.2. Experimental design

Specific pathogen free (SPF) eggs were obtained from the Canadian Food

Inspection Agency (CFIA) (Ottawa, Ontario, Canada), hatched at the Arkell Poultry

Page 88: Sequence analysis, pathogenicity and cytokine gene

77

Station, University of Guelph and the chicks were housed at the University Isolation

Unit. Ten-day old chicks were randomly divided into two groups, each group containing

50 birds. Group I was assigned for intramuscular (im) administration of the virus, each

bird received 2x108 plaque forming units (pfu). Group II, the uninfected chicks (negative

control) received PBS by im administration. Following inoculation, five chicks from each

group were randomly drawn for necropsy at 1, 2, 3, 4, 5, 6, 7, 10, and 14 d.p.i. Birds were

assigned for necropsy by simple random sampling of their identification numbers

stratified by cage, using the statistic software (MINITAB 14). Euthanasia was performed

according to the University of Guelph Animal Care Committee regulations (Isolation

Unit, University of Guelph, Ontario, Canada) and the spleen, cecal tonsil and liver were

collected, and immediately submerged in RNA stabilization reagent (RNAlater, Qiagen,

Mississauga, ON, Canada) and stored at -80°C until RNA extraction.

2.3. RNA extraction

Total RNA was extracted from the tissue samples by TriZol reagent (Invitrogen,

Canada Inc., Burlington, ON, Canada) as described by the manufacturer. Briefly, tissues

were homogenized in 1 ml TriZol followed by chloroform extraction, isopropanol

precipitation and a 75% ethanol wash of the RNA. The pellet was dissolved in 20 µl of

RNAse free water and the RNA concentration was measured by NanoVue

spectrophotometry. To remove possible contaminating DNA the samples were treated

with a DNase based inactivation reagent (Invitrogen, Canada Inc., Burlington, ON,

Canada). The cDNA was synthesized with Random Primer (Invitrogen, Canada Inc.,

Burlington, ON, Canada) and by using SuperScriptTM II Reverse Trancriptase

Page 89: Sequence analysis, pathogenicity and cytokine gene

78

(Invitrogen, Canada Inc., Burlington, ON, Canada) according to the manufacturer’s

instructions. The cDNAs were stored at -80°C until analyzed by real-time PCR.

2.4. Preparation of constructs for producing standard curves

Preparation of constructs and creation of standard curves for all cytokine genes,

which were used in this study, as well as the β-actin gene were as described (Abdul-

Careem et al., 2006; 2007).

2.5. Real-time PCR

Real-time PCR was used to quantify the expression of β-actin and cytokine genes

from the cDNA samples. The PCR reaction was prepared in a total volume of 20 µl of

Qiagen Quantitect SYBR Green PCR kit (Qiagen) containing hot start Taq polymerase

and cDNA. All the real-time PCR reactions were carried out in 96-well plates (Roche

Diagnostics GmbH, Mannheim, Germany). Each assay was run along with no-template

control and three dilution series from the standard. The sequences and other features of

specific primers for chicken IFN-γ, IL-18, IL-10, IL-8, and β-actin have been described

(Abdul-Careem et al., 2006; 2007). The primers were synthesized by Laboratory

Services, University of Guelph. The PCR conditions for different segments of each cycle

were optimized for all target genes because the thermal cycling parameters differ for each

gene (Abdul-Careem et al., 2006). All samples were amplified in a LightCycler (LC480)

instrument (Roche Diagnostics GmbH, Mannheim, Germany).

Page 90: Sequence analysis, pathogenicity and cytokine gene

79

2.6. Data analysis

Real-time PCR efficiency was calculated by 10(-1/slope of the standard curve) (Rasmussen,

2001). The mRNA expression was calculated relative to β-actin gene expression using

the equation (Etarget)ΔCPtarget(control –sample)/(Eref)

ΔCPref(control –sample) (Pfaffl, 2001). In this

equation, CP (crossing point) is the point at which the fluorescence rises significantly

above the background. The difference in cytokine expression between groups was

assessed by Mann-Whitney test and comparisons were considered significant at P≤0.05.

3. Results 3.1. Cytokine gene expression in CH-SAH cells

In our laboratory FAdVs are propagated in CH-SAH cells and since there are no

data available about the expression of cytokines by these cells we followed IFN-γ, IL-18,

IL-10 and IL-8 mRNA production upon infection with FAdV-8. The expression of IFN-γ,

IL-18, IL-10 and IL-8 of uninfected and infected CH-SAH cells at various time points is

illustrated in Fig. 3.1. The gene expression of all selected cytokines was down-regulated

in the first four hours after infection when compared to uninfected controls. The

expression of IL-10, IL-18 and IL-8 was up-regulated at 8 h.p.i. and the difference

between infected and uninfected samples was statistically significant (P≤0.05).

Significant up-regulation was detected up to 10 h.p.i. for IL-10 and IL-8 while for IL-18

up-regulation persisted up to 12 h.p.i. IFN-γ exhibited significant (P≤0.05) transient up-

regulation measured only at 10 h.p.i.

Page 91: Sequence analysis, pathogenicity and cytokine gene

80

3.2. Cytokine gene expression in experimentally infected chickens

The cytokine gene expression in the spleen, liver and cecal tonsil of chickens

infected with FAdV-8 was evaluated by a real-time quantitative reverse transcription-

PCR.

3.2.1. Gene expression in spleen

The expression of IFN-γ, IL-18, IL-10 and IL-8 mRNA in spleen of FAdV-8

infected and uninfected control chickens is shown in Fig. 3.2. The expression of selected

interleukins was detected at all time points. There was a statistically significant increase

(P≤0.05) in IFN-γ gene expression in spleen of infected chickens on day 3 p.i. Also, at

this time point, there was a significant (P≤0.05) decrease in IL-10 gene expression.

However, when IL-18 gene expression was compared to that of uninfected controls

statistically significant (P≤0.05) increase in expression was seen among infected chickens

at day 1 p.i. Expression of this cytokine changed on day 7 p.i., when it was significantly

down-regulated (P≤0.05) in the infected group. IL-8 gene expression showed similar

patterns to those of other cytokines in that it was expressed across all time points in the

spleen of chickens belonging to both infected and uninfected groups. Expression of this

cytokine was significantly down-regulated (P≤0.05) in the infected group at days 7 and

10 p.i.

3.2.2. Gene expression in cecal tonsil

There were no significant differences (P>0.05) in expression of mRNA for any of

the tested cytokine genes, compared with the values for the appropriate uninfected

Page 92: Sequence analysis, pathogenicity and cytokine gene

81

controls (Fig. 3.3). Increased levels of IFN-γ, IL-18 and IL-10 mRNA were evident in

infected birds at 1 d.p.i. although the results were not statistically significant. Patterns of

mRNA expression changed after 1 d.p.i. for IFN-γ, IL-18, IL-10 showing a downward

trend over the course of the experiment. IL-8 gene expression was down-regulated and

although the expression of IL-8 was detected at all time points in both groups, there was

no significant difference in the expression between groups (Fig. 3.3).

3.2.3. Gene expression in liver

The expression of IFN-γ, IL-18, IL-10 and IL-8 mRNA in liver of infected and

control chickens are illustrated in Fig. 3.4. Marked differences in expression of mRNA

between infected and uninfected chickens were evident for all studied cytokines. There

was a statistically significant increase (P≤0.05) in IFN-γ gene expression in liver of

infected chickens on days 3, 5, and 7 p.i. Infected chickens also had a significantly higher

IL-10 gene expression compared to chickens in the uninfected control group at days 1, 3,

and 5 p.i. Additionally, adenoviral infection altered the expression levels of IL-8 and IL-

18 and resulted in an increased expression (P≤0.05) of IL-18 at 3 d.p.i. On the other hand

the expression of IL-8 in the liver was down regulated and a decrease in expression

became more prominent and statistically significant (P≤0.05) at days 5, 7, and 10 p.i.

4. Discussion

Fowl adenoviruses are widely distributed in poultry operations and infect a wide

range of tissues, but their ability to cause disease depends on numerous factors such as

the type and age of the chickens, virulence of the virus, and secondary infection. The host

Page 93: Sequence analysis, pathogenicity and cytokine gene

82

defense mechanism in healthy birds usually functions effectively, however, some isolates

of certain serotypes can cause considerable morbidity and mortality especially among

nutritionally or immunologically compromised birds caused by a different viral infection,

for example chicken anemia or infectious bursal disease viruses (Toro et al., 2000). Due

to an increased interest in adenoviruses as vaccine and gene delivery vectors it is

important to understand the molecular aspects of infection and the host responses to them

(Corredor and Nagy, 2010b).

Innate early response occurs as a direct consequence of interaction of the virus

with the cell and does not always depend on the transcription of any virus genes.

Nevertheless, the nature of the innate response can differ significantly depending on

receptor usage and cell type, activating a complex signaling cascade with diverse

outcomes. Rendall and Goodbourn (2008) showed that the majority of these pathways

lead to the up-regulation of transcription factor NF-κB and interferon regulatory factor 3

(IRF3). Muruve (2004) confirmed that the interaction between adenoviral capsid and host

cell is sufficient to activate the pathways that lead to inflammatory responses. A study of

Tamanini et al. (2006) demonstrated that the fiber is sufficient for the immediate pro-

inflammatory response. The interaction of the fiber with the Coxsackie and adenovirus

receptor (CAR) is a critical initial event for such a response. Following adenovirus

infection, pathogen-associated molecular patterns (PAMs) are recognized by cellular

sensors, pattern-recognition receptors (PRRs) (Akira et al., 2006). The main function of

PRRs comprise enhancement of phagocytosis, activation of complement cascades,

production of inflammatory cytokines and maturation of dendritic cells (DCs), that all

Page 94: Sequence analysis, pathogenicity and cytokine gene

83

together orchestrate the early host response to infection and are important link to the

adaptive immune response (Akira et al., 2006). The most extensively studied PRRs are

toll-like receptors (TLR). To date 13 TLRs (TLR1 to TLR13) have been identified in

mammals and mice. In chickens Temperley et al. (2008) found nine TLRs (1, 2, 3, 4, 5, 7,

9, 15, and 21) with five human and one fish orthologs, and only three are unique to

chicken (Fukui et al., 2001; Kogut et al., 2005; Boyd et al., 2007; Jenkins et al., 2009).

Data on the interaction of FAdV-encoded molecular patterns and TLRs that recognize

these PAMPs are not available. However, there are differences in innate immune

response pathways depending on the infected cells and on the adenovirus species. Kim et

al. (2008) reported the recognition of unmethylated CpG sequences in human adenovirus

by TLR9. There are essential differences in the TLR repertoires of mammals and birds,

TLR15 does not exist in mammals and chickens lack TLR9. Keestra et al. (2010)

reported chicken TLR21 as a CpG DNA receptor that shares functional characteristics but

displays minimal sequence similarity with mammalian TLR9. Activation of innate

immune response without the participation of TLR9 has been reported by Takaoka et al.

(2007). Cerullo et al. (2007) showed that empty adenoviral particles devoid of genomic

DNA are poor inducers of innate responses suggesting that the DNA is a major

immunostimulatory molecule of adenoviruses. Engagement of some of the innate

receptors, such as TLRs with PAMPs, results in triggering of downstream pathways,

including IFN. Another type I or Th1 cytokine, IL-18 is critical for IFN- γ production by

natural killer (NK) cells and T cells (Rathinam and Fitzgerald, 2010). On the other hand,

the potent chemotactic cytokine IL-8 is known as an important mediator of inflammation

which recruits and activates neutrophils to sites of infection (Baggiolini, 1998). The

Page 95: Sequence analysis, pathogenicity and cytokine gene

84

overproduction of cellular IL-8 in viral infections is a reactive response of the host cells

aimed to reducing viral load (Parcells et al., 2001). On other hand, IL-10 has a key effect

on the suppression of Th1 cell response. This cytokine inhibits the production of the pro-

inflammatory cytokines by DCs and macrophages (Moore et al., 2001). Additionally, IL-

10 modulates the immune response towards a type 2 response by inhibiting IFN-γ

synthesis (Rothwell et al., 2004).

We examined cytokine gene expression patterns associated with FAdV infection

and to our knowledge this is the first report on this subject. We determined that IFN-γ,

IL-10, IL-18 and IL-8 were activated at early times after FAdV-8 infection. The studied

cytokines represent chicken Th1 and regulatory cytokines and all have previously been

shown to be involved in immunity to viral infections.

IL-18 is an important regulator of innate and acquired immune response and is

expressed at sites of inflammation (Gobel et al., 2003). Although originally identified as

an inducer of IFN-γ production, this cytokine also enhances T and NK cell maturation,

cytokine production and cytotoxicity (Okamura et al., 1995; Yoshimoto et al., 1998;

Micallef et al., 1996; Dao et al., 1998). During viral infection IL-18 activates CD8+ T

cells and is important for viral clearance. Abdul-Careem et al. (2007) showed that IL-18

expression is significantly higher in spleen of chickens in the unvaccinated/infected

group in which IL-10 expression was enhanced, whereas IFN-γ expression was minimal.

A possible explanation for such result was that chicken IL-18 could also be involved in

inducing type 2 cytokines, such as IL-10.

Page 96: Sequence analysis, pathogenicity and cytokine gene

85

Similar to mammals (Endharti et al., 2005; Minang et al., 2006) one of the roles of

chicken IL-10 is to repress the Th1 response. IL-10 aborts T cell responses and can

inhibit ongoing T cell activity to viral infections. It has been shown by several groups

(Carbonneil et al., 2004; Fiorentino et al., 1991; Moree et al., 2001; Steinbrink et al.,

1997) that IL-10 alters cytokine production and prevents maturation depressing T cell

activation.

In agreement with the findings of Abdul-Careem et al. (2007) we discovered that

expression of IL-18 and IL-10 genes in infected CH-SAH cells was significantly up-

regulated at 8, 10 and 12 h.p.i. in comparison with mock infected cells. Hence, IFN-γ

expression was significantly up-regulated at a single time point (10 h.p.i) and shortly

after exhibited a tendency to down-regulation. Our potential explanation of such a shift

in expression of IFN-γ is the inhibitory effect of IL-10 as observed by Endharti et al.

(2005) and Minang et al. (2006).

A slightly different trend was seen in chickens where expression of IL-18 gene

was significantly higher in spleen and liver of infected chickens at day 1 and 3 p.i.,

respectively. Accordingly, IL-10 expression was significantly up-regulated in liver of

infected chickens at 1, 3, and 5 d. p .i. In our study a significant induction of IFN-γ gene

was noted in the spleen of infected chickens at day 3 p.i., while in the liver the expression

of IFN-γ was significantly higher at day 3, 5 and 7 p.i. The reported variation might be

directly related to the pathogen used in the study. Moreover, although IL-10 is considered

a regulatory cytokine, which is inversely correlated with the expression of Th1 cytokines

Page 97: Sequence analysis, pathogenicity and cytokine gene

86

such as IFN-γ, under certain circumstances could be produced by Th1 cells (O’Garra and

Vieira, 2007). Further, IL-12 stimulates human CD4 and CD8 T cells for increased

production of both IFN-γ and IL-10 (Gerosa et al., 1996). Simultaneous expression of

IFN-γ and IL-10 genes has been also reported by Johensen et al. (2004) following

infection with porcine reproductive and respiratory syndrome virus.

Adenovirus encodes numerous products that counteract host defenses, ensure

prolonged survival in the infected host and consequently facilitate viral transmission.

Zhou et al. (2007) have demonstrated that porcine adenovirus 3 early region 1Blarge

(E1Blarge) protein down-regulates the induction of proinflammatory cytokine IL-8 by

inhibiting the NF-κB dependent gene transcription from human IL-8 promoter. The same

authors suggested that the E1Blarge protein acts as a IkappaB (IκB) homolog and retains

the ability to bind and inactivate NF-κB. Moreover, Lesokhin et al. (2002) reported that

the E3 genes can drastically reduce level of IL-8 mRNA. Inhibition of the amounts of

chemokine mRNA by E3 is directly related to decreased secretion of IL-8 proteins as

showed by ELISA. Four out of six proteins of the human adenovirus E3 region are

involved in immunoregulation (Wold et al., 1994). Bennett et al. (1999) showed that E3-

19K retains the major histocompatibility complex (MHC) I in the endoplasmic reticulum

and inhibits tapasin processing of peptides that bind to MHC I. The E3-10.4K and E3-

14.5K complex stimulates the internalization of proapoptotic receptors from the cell

surface (Tollefson et al., 1998; Elsing and Burgert 1998). The E3-14.7 inhibits tumor

necrosis factor alpha (TNF-α)-induced apoptosis (Gooding et al., 1990; Li et al., 1999).

Additionally, E3-14.7K inhibits TNF-α-induced production of arachidonic acid (Krajcsi

Page 98: Sequence analysis, pathogenicity and cytokine gene

87

et al., 1996). It is unclear which E3 protein interactions are important in the inhibition of

IL-8. The E3 region is placed between the pVIII and fiber protein genes and deletion of

this region results in an increased inflammatory response (Braithwaite and Russell 2001).

We recently showed that FAdV-8 along with FAdV-1, FAdV-4 and FAdV-9 lacks

mastadenoviral E1, E3 and E4 regions (Grgic et al., 2011).

In CH-SAH cells expression of IL-8 in infected cells was significantly down-

regulated in comparison with mock infected cell at 1, 4, and 24 h.p.i. A similar

observation has been reported in Marek’s disease virus (MDV)-transformed

lymphoblastoid cell line where infection with Mycoplasma gallisepticum suppressed IL-8

gene expression (Lam, 2004). On other hand, IL-8 is up-regulated in splenocytes after

Marek’s disease virus infection (Xing and Schat, 2000).

In this study, the expression of IL-8 mRNA in spleen of infected chickens was

significantly down-regulated at 7 and 10 d.p.i., similarly, significant down-regulation was

detected in the liver at 5, 7 and 10 p.i. These results are in accordance with the findings of

Zhou et al. (2007) and Lesokhin et al. (2002). Interestingly, the GAM-1 protein, encoded

by the Gam 1 gene (ORF8) at the right end of the FAdV-1 genome, has mastadenoviral

E1 functions such as inhibition of apoptosis (Chiocca et al., 1996; Lehrmann and Cotten,

1999). Homologue of ORF8 has been also identified for FAdV-8 genome (Grgic et al.,

2011). Therefore, we speculate that some parts of the FAdV-8 genome share functions

similar to mastadenoviral E1, E3 and E4 regions despite a lack of obvious and high

sequence homology.

Page 99: Sequence analysis, pathogenicity and cytokine gene

88

In conclusion the results presented here showed that FAdV-8 infection is

associated with enhanced expression of IFN-γ, IL-18 and IL-10 and reduced expression

of IL-8 in spleen and liver of infected chickens. There was no significant difference in

any of the studied cytokine gene expression between infected and uninfected groups in

cecal tonsil. Our finding that IL-10 was significantly up-regulated in liver might be

indicative of an immune evasive mechanism used by the virus to skew the Th-1-like

immune response against FAdV-8 and to aid the persistence of the virus. Another

possible explanation could be that IL-10 was induced by the chicken immune response

and inflammation caused by the virus. Further work is necessary to shed more light on

the role of cytokines in elicitation of the immune response to FAdV infections.

Acknowledgements

This work was supported by the Natural Sciences and Engineering Research Council of

Canada (NSERC), the Canadian Poultry Research Council (CPRC), and the Ontario

Ministry of Agriculture, Food and Rural Affairs (OMAFRA). The authors thank the

Isolation Unit personnel for their professional animal care and assistance.

Page 100: Sequence analysis, pathogenicity and cytokine gene

89

Figure 3.1. Cytokine mRNA expression in CH-SAH following FAdV-8 infection. The groups were as follows: Uninfected (negative control) CH-SAH, and FAdV-8 infected CH-SAH. Target and reference gene expression was quantified by real-time RT-PCR using SYBR Green. Target gene expression is presented relative to β-actin expression and normalized to a calibrator. The results are diagrammed as whisker boxes with medians. The difference in cytokine expression between groups was assessed by Mann-Whitney test and comparisons were considered significant at P≤0.05 (*). Numbers 1, 4, 8, 10, 12, and 24 represent hours post infection when samples were collected.

Page 101: Sequence analysis, pathogenicity and cytokine gene

90

Page 102: Sequence analysis, pathogenicity and cytokine gene

91

Figure 3.2. Cytokine mRNA expression in spleen of chickens infected with FAdV-8. The groups were as follows: Uninfected (negative control) chickens, and FAdV-8 infected chickens. Target and reference gene expression in spleen was quantified by real-time RT-PCR using SYBR Green. Target gene expression is presented relative to β-actin expression and normalized to a calibrator. The results are diagrammed as whisker boxes with medians. Boxes represent interquartile ranges and whiskers indicate extreme values. The difference in cytokine expression between groups was assessed by Mann-Whitney test and comparisons were considered significant at P≤0.05 (*). ○, ● symbols represent values which were identified as outliers. Numbers 1, 3, 5, 7, and 10 represent days post infection when samples were collected.

Page 103: Sequence analysis, pathogenicity and cytokine gene

92

Page 104: Sequence analysis, pathogenicity and cytokine gene

93

Figure 3.3. Cytokine mRNA expression in cecal tonsil of chickens infected with FAdV-8. The groups were as follows: Uninfected (negative control) chickens, and FAdV-8 infected chickens. Target and reference gene expression in spleen was quantified by real-time RT-PCR using SYBR Green. Target gene expression is presented relative to β-actin expression and normalized to a calibrator. The results are diagrammed as whisker boxes with medians. Boxes represent interquartile ranges and whiskers indicate extreme values. The difference in cytokine expression between groups was assessed by Mann-Whitney test and comparisons were considered significant at P≤0.05 (*). ○, ● symbols represent values which were identified as outliers. Numbers 1, 3, 5, 7, and 10 represent days post infection when samples were collected.

Page 105: Sequence analysis, pathogenicity and cytokine gene

94

Page 106: Sequence analysis, pathogenicity and cytokine gene

95

Figure 3.4. Cytokine mRNA expression in liver of chickens infected with FAdV-8. The groups were as follows: Uninfected (negative control) chickens, and FAdV-8 infected chickens. Target and reference gene expression in spleen was quantified by real-time RT-PCR using SYBR Green. Target gene expression is presented relative to β-actin expression and normalized to a calibrator. The results are diagrammed as whisker boxes with medians. Boxes represent interquartile ranges and whiskers indicate extreme values. The difference in cytokine expression between groups was assessed by Mann-Whitney test and comparisons were considered significant at P≤0.05 (*). ○, ● symbols represent values which were identified as outliers. Numbers 1, 3, 5, 7, and 10 represent days post infection when samples were collected.

Page 107: Sequence analysis, pathogenicity and cytokine gene

96

Page 108: Sequence analysis, pathogenicity and cytokine gene

97

CHAPTER 4

Pathogenicity and cytokine gene expression patterns of a serotype 4 fowl adenovirus isolate, a vaccine candidate

Helena Grgića, Zvonimir Poljakb, Shayan Sharifa and Éva Nagya,*

aDepartment of Pathobiology, Ontario Veterinary College, University of Guelph, Guelph, Ontario, N1G 2W1, Canada bDepartment of Population Medicine, Ontario Veterinary College, University of Guelph, Guelph, Ontario, N1G 2W1, Canada

*Corresponding author: Éva Nagy E-mail: [email protected] Tel: +1 519 824 4120 ext 54783 FAX: +1 519 824 5930

Author’s contributions

HG designed and performed experiment, carried out analysis and wrote the paper. ZP

advised on statistical analysis. SS and EN provided guidance during the research and

critically revised manuscript

To be submitted to Vaccine.

Page 109: Sequence analysis, pathogenicity and cytokine gene

98

Abstract

Infectious hydropericardium-hepatitis syndrome (HHS) is a recently emerged

disease of poultry associated with some strains of fowl adenovirus serotype 4 (FAdV-4).

In this study, a Canadian FAdV-4 isolate was evaluated for pathogenicity after oral and

intramuscular (im) infection of specific pathogen free (SPF) chickens. Pathogenicity was

evaluated by observation of clinical signs and gross and histological lesions. The highest

viral copy numbers, irrespective of the inoculation route, were detected in the cecal

tonsils. Virus titers in cloacal swabs collected over the entire study period were compared

between the orally and im inoculated chickens, and the difference in titers between the

two groups was significant (P<0.001). The antibody response of infected chickens tested

by an adenovirus-specific ELISA showed a statistically significant (P<0.001) difference

between the orally and im inoculated chickens. In addition, the effects of FAdV-4

infection on transcription of a number of avian cytokines were studied in vivo. Interferon

(IFN)-γ, interleukin (IL)-10, IL-18, and IL-8 were activated at early times after FAdV-4

infection. Our study demonstrated that this FAdV-4 isolate under experimental conditions

was non-pathogenic and could be considered as a live vaccine against HHS.

Keywords:

Fowl adenovirus serotype 4 (FAdV-4); Pathogenesis; IFN-γ; IL-18; IL-10; IL-8

Abbreviations: FAdV-4, fowl adenovirus 4; d.p.i., days post-infection; pfu, plaque

forming unit; SPF, specific pathogen free; IFN-γ, interferon gamma; IL, interleukin;

HHS, hydropericardium-hepatitis syndrome; MOI, multiplicity of infection; NK, natural

killer cells; Ab, antibody

Page 110: Sequence analysis, pathogenicity and cytokine gene

99

Highlights:

Pathogenicity of an FAdV-4 isolate in SPF chickens

Expression of cytokine genes after fowl adenovirus serotype 4 infection of chickens

FAdV-4 infection is associated with up-regulation of IL-18 and IFN-γ mRNA expression

Up-regulation of IL-10 may be indicative of an immune evasive mechanism used by

FAdV-4

Page 111: Sequence analysis, pathogenicity and cytokine gene

100

1. Introduction

Fowl adenovirus serotype 4 (FAdV-4) plays a major role in the aetiology of a

disease in broiler chickens called infectious hydropericardium-hepatitis syndrome (HHS).

This is an emerging disease initially reported from Pakistan (Khawaja et al., 1988). In

addition to India and Pakistan, outbreaks have also been reported in the Middle East,

Japan, Mexico, Peru, Ecuador, and Chile (Abe et al., 1998; Hess et al., 1999; Toro et al.,

1999). Outbreaks of HHS have not been reported in Canada. The disease is prevalent in

3- to 7-week-old broiler chickens. The affected birds may not exhibit signs other than

high mortality, up to 75% (Jaffery, 1988). The predominant gross lesion is

hydropericardium, characterized by accumulation of a jelly-like fluid in the pericardial

sac. The heart appears misshapen and flabby, with its apex floating in the pericardial sac

(Adair and Fitzgerald, 2008). At histopathological examination, the most consistent

findings in the liver are small multifocal areas of necrosis and mononuclear cell

infiltration, including basophilic intranuclear inclusion bodies in hepatocytes surrounded

by a clear halo or filling the entire nucleus (Nakamura et al., 1999). Initially, it was

hypothesized that HHS is caused by a toxin or by a nutritional disorder (Jaffery, 1988).

However, all attempts to reproduce the disease with feed samples from infected farms

failed. On the basis of successful isolation of an agent from the liver, heart, and kidneys,

and on transmission of disease after subcutaneous inoculation of infected liver

homogenate, the aetiological agent was tentatively identified as an adenovirus (Khawaja

et al., 1988). Subsequent studies demonstrated typical icosahedral adenovirus particles in

the purified liver extracts by negative-staining electron microscopy. The virus is usually

transmitted horizontally in flocks by the oral-fecal route (Cowen, 1992). Existing

Page 112: Sequence analysis, pathogenicity and cytokine gene

101

vaccines against HHS are composed of liver homogenate extract or virus propagated in

chicken liver cells or embryonated eggs. These preparations are inactivated with

formalin, β-propiolactone or heat treatment at 56˚C (Anjum, 1990) and they provide

protection against natural outbreaks of the disease (Anjum, 1990).

The FAdV-4 isolate investigated in this study was from a Canadian broiler

breeder flock with no clinical signs of HHS. The complete nucleotide sequence of the

virus was determined by Griffin and Nagy (2011).

The goals of this study were to (1) study the pathogenicity of this FAdV-4 isolate

in specific pathogen free (SPF) chickens, (2) investigate virus dissemination in various

tissues, virus titers in cloacal swabs, and antibody response, and (3) examine the

dynamics of interferon (IFN)-γ, interleukin (IL)-10, IL-18, and IL-8 gene expression

following FAdV-4 infection.

2. Material and methods 2.1. Virus

The FAdV-4 was isolated from tissues collected from chickens of a Canadian

broiler breeder flock showing no clinical signs of HHS. Virus isolation was performed by

the Animal Health Laboratory (AHL) at the University of Guelph. The virus was plaque

purified in our laboratory, and virus propagation and titration were performed in CH-

SAH cells as described by Alexander et al. (1998).

Page 113: Sequence analysis, pathogenicity and cytokine gene

102

2.2. Chickens and housing

All birds were white Leghorn SPF chickens. The eggs were obtained from the

Canadian Food Inspection Agency (CFIA), Ottawa, and they were hatched at the Arkell

Poultry Station. The birds were housed at the University Isolation Unit according to the

Animal Care Committee of the University of the Guelph in accordance with the Guide to

the Care and Use of Experimental Animals of the Canadian Council on Animal Care and

received food (21% unmedicated starter ration) and water ad libitum.

2.3. Pathogenicity assessment of FAdV-4 in SPF chicks

A chicken trial was conducted to evaluate the pathogenic potential of the FAdV-4

isolate. One hundred chicks were randomly divided into four groups: Group I (FAdV-4) -

oral, Group II (FAdV-4) -intramuscular (im), Group III (control) -oral, and Group IV

(control) -im. Groups I and II comprised 30 chickens each, and the negative control

groups (III and IV) comprised 20 chicks each. At 9 days of age, the chickens were tagged

with wing bands (Ketchum, ON) and blood samples were collected from all birds. Ten-

day-old chickens in Groups I and II were inoculated with 2x108 plaque forming units

(pfu) of virus per chick. The control chickens (mock-infected) received PBS. Groups I

and II were re-inoculated at 14 days of age with the same dose of virus. The chickens

were observed daily (3x) for the development of clinical signs. Three chicks from the

control groups and four chicks from the treatment groups were randomly selected for

necropsy by simple random sampling of their identification numbers stratified by cage,

using statistical software (MINITAB 14) at 0, 3, 5, 7, 14, 21, and 28 days post-infection

(d.p.i.). The chickens were euthanized according to the University of Guelph Animal

Page 114: Sequence analysis, pathogenicity and cytokine gene

103

Care Committee regulations (Isolation Unit, University of Guelph, Ontario, Canada),

necropsied, and examined for the presence of gross lesions, and tissues from thymus,

lung, heart, kidney, liver, bursa of Fabricius, and cecal tonsils were collected. Each organ

was sectioned into three portions: one was placed into 10% formalin for histology, the

second was stored at -80˚C for quantitative polymerase chain reaction (qPCR), and the

third was stored at -80˚C for virology. Cloacal swabs were collected at 0, 3, 5, 10, 14, 21,

28 d.p.i. from all chicks and their virus titers were determined in CH-SAH cells. Blood

samples were collected weekly, at 0, 7, 14, 21, and 28 d.p.i. and tested for FAdV-specific

antibodies (Abs) by enzyme-linked immunosorbent assay (ELISA) as described (Ojkic

and Nagy, 2003). The results were expressed by calculating the sample-to-positive (S/P)

ratio for each sample from mean optical density (OD405) readings.

2.4. Experimental design for determination of cytokine mRNA expression

Eighty ten-day-old chicks were randomly divided into two groups. Group I

contained 50 birds and Group II (negative control) contained 30 birds. Each bird in

Group I received 2x108 pfu/ml of the virus im. The negative control chicks received PBS

by im administration. Five chicks from the infected group and three chicks from the

uninfected group were selected randomly for necropsy at 0, 1, 2, 3, 4, 5, 6, 7, 10, and 14

d.p.i. The rationale for selecting these time points was guided by the assumption that

rarely is one immune measure at one time point sufficient for any conclusion. Therefore,

we selected several time points in order to cover the turning point of the infection. The

chickens were euthanized and the spleen, cecal tonsil, and liver were collected and

Page 115: Sequence analysis, pathogenicity and cytokine gene

104

immediately submerged in RNA stabilization reagent (RNAlater, Qiagen, Mississauga,

ON, Canada) and stored at -80°C until RNA extraction.

2.5. Quantification of viral DNA in tissues

In order to determine the viral load in the tissues, real-time PCR was performed using

SYBR Green as an intercalating dye. The FAdV-4 ORF-14 gene was used as an indicator

for the presence of viral DNA. The forward primer was 5’-

AGTGTGTATGTGCGTTGGGTAG-3’ and reverse primer was 5’-CATTGTCATAAC

GATGGTGTAG-3’. The 20 l qPCR reaction mix, containing 1 l (10 pmol) of each

primer, 10 l of QuantiTec SYBR Green I PCR Master Mix (Qiagen), 2 l DNA, and 6

l nuclease-free water, was prepared using computerized automated liquid handling

robotics CAS-1200 (Corbett Research). The qPCR program was run as a first step of 15

min at 95C to activate Taq polymerase, followed by 40 cycles of 95C for 20 sec, 55C

for 15 sec, and 72C for 20 sec, with an acquiring step at 76C for 15 sec, using the

Rotor-Gene 6000 System (Corbett Research). Quantitative measurement of the PCR

product was carried out by incorporation of the SYBR Green I fluorescent dye into

double-stranded DNA. To establish standard curves, 10-fold serial dilutions (10-1 to 10-7)

of FAdV-4 genomic DNA starting at 12.2 ng/l (2.4 x 108 copies/l) were made and used

as qPCR templates. Melting curve analysis was performed after each run to control the

specificity of the PCR products. Negative control DNA extracted from uninfected

chickens and no template control were routinely included in each experiment.

Page 116: Sequence analysis, pathogenicity and cytokine gene

105

2.6. RNA extraction and real-time PCR for determination of cytokine gene expression

Total RNA was extracted from the spleen, cecal tonsil, and liver samples by

Trizol reagent (Invitrogen, Canada Inc., Burlington, ON, Canada) as described by the

manufacturer. The tissues were homogenized in 1 ml Trizol followed by chloroform

extraction, isopropanol precipitation, and a 75% ethanol wash. The RNA pellet was

dissolved in 20 µl of RNAse-free water. The concentration of the RNA was measured by

NanoVue spectrophotometer. In order to remove possible contaminating DNA, a DNase

inactivation reagent (Invitrogen) was applied. For cDNA synthesis from total RNA,

SuperScriptTM II Reverse Trancriptase (Invitrogen) was used according to the

manufacturer’s instructions and the random primers were from Invitrogen. The cDNAs

were stored at -80°C until used for real-time PCR.

A real-time PCR assay was employed to quantify the expression of β-actin and

cytokine genes in cDNA samples. The PCR reaction was prepared in a total volume of 20

µl of Qiagen Quantitect SYBR Green PCR kit (Qiagen) containing hot start Taq

polymerase and cDNA. Each assay was performed without a template control and with

three dilution series from the standard. The sequences and other features of specific

primers for chicken IFN-γ, IL-18, IL-10, IL-8, and β-actin have been described (Abdul-

Careem et al., 2006; 2007). The primers were synthesized by the Laboratory Services,

University of Guelph. The PCR conditions for different segments of each cycle were

optimized for all target genes.

Page 117: Sequence analysis, pathogenicity and cytokine gene

106

Previous research indicated that after FAdV infection the most variability in

cytokine gene expression has been in the liver (Grgic et al., 2012). Therefore, in the liver,

mRNA expression was determined for IFN-γ, IL-18, IL-10, and IL-8, while for cecal

tonsil and spleen, mRNA expression was determined only for IFN-γ and IL-18.

2.7. Statistical analysis

In order to evaluate whether viral copy number differed among tissues, a Kruskal-

Wallis test was performed. When statistically significant P-values were identified, pair-

wise comparisons were performed between different tissues. P-values of <0.05 were

considered to be statistically significant.

Real-time PCR efficiency was calculated by 10(-1/slope of the standard curve) (Rasmussen,

2001). The mRNA expression was calculated relative to β-actin gene expression using

the equation (Etarget)ΔCPtarget(control –sample)/(Eref)

ΔCPref(control –sample) (Pfaffl, 2001). In this

equation, CP (crossing point) is the point at which the fluorescence rises significantly

above the background. The difference in cytokine expression between groups was

assessed by a Mann-Whitney test and comparisons were considered significant at P<0.05.

3. Results 3.1. Pathogenicity

Clinical signs characterized by depression, chickens huddling in corners with

ruffled feathers and exhibiting characteristic posture, with the chest and beak resting on

the ground was not seen in any groups of chicks. Distinctive gross lesions,

hydropericardium, pale, swollen and friable liver were absent. The characteristic lesions

Page 118: Sequence analysis, pathogenicity and cytokine gene

107

of HHS on histological examination described as intranuclear inclusion bodies in

hepatocytes were not detected in any of the infected chickens.

3.2. Viral genome copy number in tissues

Viral genome copy numbers in liver, bursa of Fabricius, and cecal tonsil were

determined by quantitative real-time PCR and the results over the entire study are

summarized in Table 4.1. No viral DNA was detected in any tissues from any chickens

before inoculation or in the mock-infected chickens. Each virus and each inoculation

group (oral and im) were subjected separately to the Kruskal-Wallis test to evaluate the

differences in number of viral copies among tissues. The cecal tonsil was the organ with

the highest number of viral copies, followed by liver and then bursa, irrespective of

inoculation route. At group level for FAdV-4 orally inoculated birds, there were

significant differences (P<0.001) in number of viral copies, whereas for the im group, the

differences tended to be significant (P=0.06). For the orally inoculated birds, viral copy

number was higher in cecal tonsil than in liver (P=0.0207), and bursa (P=0.0001),

whereas viral copy numbers in liver ranked only marginally higher than viral copy

numbers in bursa (P=0.0759). For the im inoculated birds, viral copy number was higher

in cecal tonsil than in the bursa (P=0.0154), whereas differences among other organs

were not significant (P>0.25).

3.3. Virus titers in feces

No virus was detected in the cloacal swabs of any chickens before inoculation nor

in the mock-infected groups at any time. The virus titers are given in Table 4.2. The

difference in titers between the groups of orally and im inoculated chickens was

Page 119: Sequence analysis, pathogenicity and cytokine gene

108

statistically significant (P<0.001) when it was tested using a 2-sample Wilcoxon test. The

oral group had higher ranks. The highest titers were found in the orally inoculated group

at days 3 p.i. and 5 p.i. Chickens shed the virus through the entire study period in both the

oral and im groups.

3.4. Antibody response

Ab against fowl adenovirus was not detected in any birds before inoculation (0

d.p.i.) nor in the mock-infected birds at any time. An Ab response against FAdV-4

appeared at 7 d.p.i. and the difference between inoculated groups and negative controls

was statistically significant (P<0.001). The inoculated chickens had higher antibody

levels than the mock-infected birds (Fig. 4.1A). Chickens inoculated im had higher

values than birds inoculated orally (P<0.001). When the same sera were tested against

FAdV-9, a heterologous virus, the results were similar, but the levels were slightly lower

(Fig. 1B).

3.5. Expression of cytokine genes

Expression of selected cytokine genes after FAdV-4 inoculation was measured in

cecal tonsil, liver and spleen.

In the cecal tonsils there were no significant differences (P>0.05) in the

expression of IFN-γ and IL-18 mRNA when compared with the values for the appropriate

uninfected controls (Fig. 4. 2). Up-regulation of IFN-γ and IL-18 mRNA was evident in

Page 120: Sequence analysis, pathogenicity and cytokine gene

109

infected birds at 1, 3 and 7 d.p.i., although the results were not statistically significant.

The mRNA expression changed after 7 d.p.i., showing a downward trend at 10 d.p.i.

The expression of IFN-γ, IL-18, IL-10, and IL-8 mRNA in the liver of FAdV-4

infected and uninfected control chickens is illustrated in Fig. 4.3. Expression of all

selected cytokines appeared to differ between infected and uninfected chickens. In the

liver, an upward trend for all the studied cytokines was recorded for the infected chickens

throughout the study period up to 10 d.p.i. There was a statistically significant increase

(P<0.05) in IFN-γ gene expression of infected chickens at 3 d.p.i. when compared to

chickens in the uninfected group. The expression pattern of IL-10 was similar to the

expression of IFN-γ, being significantly higher at 3 d.p.i. in the infected group of

chickens when compared to the uninfected group (P<0.05). The expression of IL-8 was

not significantly different (P>0.05) between infected and uninfected groups, although an

upward trend in the infected group was observed throughout the study period.

The expression of IFN-γ and IL-18 mRNA in the spleen of FAdV-4 infected and

uninfected control chickens is shown in Fig. 4. 4. Both IFN-γ and IL-18 were detected at

all time points and were down-regulated, in contrast to expression levels in cecal tonsil

and liver. When gene expression of both IFN-γ and IL-18 in the spleen was compared

between infected groups and uninfected controls, statistically significant (P<0.05) down-

regulation was observed in the infected group at 10 d.p.i.

4. Discussion

Fowl adenoviruses are associated with numerous clinical conditions, but their

primary role in disease is still not understood. Most strains of FAdVs follow the same

Page 121: Sequence analysis, pathogenicity and cytokine gene

110

pattern of infection. Following initial multiplication, viraemia occurs and results in virus

spread to virtually all organs. The main sites of virus replication are the respiratory and

alimentary systems (Saiffudin and Wilks, 1991).

In this study, despite the lack of clinical signs and gross and histological changes,

virus was found in selected tissues and cloacal swabs of all infected birds. In both the

orally and im inoculated groups, the highest numbers of viral genome copies were found

in cecal tonsil. This finding is in agreement with previous work on FAdV-8 (Grgic et al.,

2011). Ojkic and Nagy (2003) reported that replication and distribution of another FAdV

serotype in chicken tissues can be distinctly influenced by the route of virus inoculation.

Since FAdV-4 is transmitted horizontally among birds by the oral-fecal route (Cowen,

1992), the im route and the more natural oral route of infection were compared. The

results showed that the orally inoculated group had a higher number of viral copies in

general. This finding is in contrast to the previously published work related to FAdV-8

(Grgic et al., 2011), where im-inoculated chicks had a higher number of viral copies. The

reason for this difference might be explained by the fact that different serotypes were

applied in those two studies. Moreover, in the present study, virus was isolated from

cloacal swabs through the entire study period (28 d.p.i.) in the im-inoculated group, while

for the orally inoculated chickens, shedding was recorded up to 21 d.p.i. Orally

inoculated chickens exhibited a significantly higher virus titer in feces than the im-

inoculated group throughout the entire study period. This result is in agreement with the

FAdV-8 pathogenicity study, where shedding was also reported up to 28 d.p.i.

Ab response was significantly higher in im-inoculated chicks than in orally inoculated

chicks. In agreement with this finding, FAdV-8 also induced higher antibodies when the

Page 122: Sequence analysis, pathogenicity and cytokine gene

111

birds were infected via the im route (Grgic et al., 2012). Additionally, all sera were tested

with heterologous FAdV-9 virus, and while results were similar, S/P ratios were lower,

supporting the hypothesis that heterologous virus can be used to detect adenovirus

antibodies against all 12 serotypes. More importantly, we determined that FAdV-4 is

immunogenic, i.e, can induce an adequate antibody response. Currently available

vaccines against HHS are composed of infected liver homogenates or virus propagated in

chicken liver cells and inactivated with 0.5% formalin or 0.1% β-propiolactone (Anjum,

1990). However, the inactivated vaccines currently available are based on virulent

viruses, with the potential risk of a vaccine break if inactivation of the virus is

incomplete.

Furthermore, cytokine responses in liver, spleen, and cecal tonsils were

investigated subsequent to infection with an FAdV-4 isolate, showing that IFN-γ, IL-18,

IL-10 and IL-8 are activated at early times after infection. The cytokines studied here are

associated with T helper (Th)1, pro-inflammatory, and regulatory activities. In this study,

the expression of these cytokines was altered during the course of FAdV-4 infection. The

expression of cytokines has been demonstrated in a chicken hepatoma cell line (CH-

SAH) and in chicken liver, spleen, and cecal tonsil following FAdV-8 infection (Grgic et

al., 2012). However, this is the first report on the cytokine repertoire associated

specifically with FAdV-4 infection. Enhanced expression of IFN-γ in liver and cecal

tonsil was noted, with a significant increase in expression at 3 d.p.i. in the liver. IFN-γ is

considered an essential cytokine in the host defense against a variety of pathogens. It has

been shown that IFN-γ inhibits Marek’s disease virus (MDV) replication by inducing

Page 123: Sequence analysis, pathogenicity and cytokine gene

112

nitric oxide synthesis (Xing and Schat, 2000), reduces replication of coccidian parasites

(Lillehoj and Choi, 1998), and inhibits src oncogene-induced tumor growth (Plachy et al.,

1999). IFN-γ is also a key factor in inhibiting reactivation of chronic infection with

murine gammaherpesvirus (Steed et al., 2007).

IL-18 was originally identified as a factor promoting IFN-γ production by T and

NK cells when costimulated by IL-12 (Okamura et al., 1995). IL-18 is a pro-

inflammatory cytokine (Göbel et al., 2003) involved in activation of neutrophils and

natural killer (NK) cells in mammals (Okamura et al., 1995). Differential expression of

IL-18 in chickens either genetically susceptible or resistant to Marek’s disease revealed

that susceptible birds have a higher expression of IL-18 in the spleen than resistant birds

(Kaiser et al., 2003). Mice lacking the IL-18 gene demonstrate impaired clearance of

neurovirulent influenza A virus-infected neurons from the brain (Mori et al., 2001).

Efficacy of feline leukemia virus (FLV) DNA vaccine is improved by coadministration

with IL-18 expression vectors (Hanlon et al., 2001). In the present study, enhanced

expression of IL-18 was noted in both liver and cecal tonsils. Enhanced expression of IL-

18 is probably an indicator of induction of inflammation and an immune response against

the virus. This finding is in agreement with a previous report where an increase in IFN-γ

and IL-18 expression was observed following FAdV-8 infection (Grgic et al., 2012).

Enhanced expression of IL-10 in the liver was recorded throughout the entire

study period, with significant up-regulation at 3 d.p.i. in the infected group. Our finding

that IL-10 was up-regulated following FAdV-4 infection is in agreement with a previous

Page 124: Sequence analysis, pathogenicity and cytokine gene

113

report (Grgic et al., 2012), and may indicate an immune evasive mechanism used by the

virus to skew the immune response, supporting persistence of the virus (Grgic et al.,

2006). IL-10 has both immunosuppressive and immunostimulating effects on cells of the

adaptive immune system (O’Garra and Vieira, 2007). Simultaneous expression of IFN-γ

and IL-10 also has been reported after infection with porcine reproductive and respiratory

syndrome virus (Johensen et al., 2002). More recently Parvizi et al. (2009) reported

persistent up-regulation of IFN-γ and IL-10 in a MDV-susceptible line of chickens. In

humans, up-regulation of both IFN-γ and IL-10 is directly related to IL-12 stimulation of

CD4+ and CD8+ T cells (Gerosa et al., 1996). IL-10 is a potent stimulator of NK cells, a

function that probably helps clear pathogens. Enhanced expression of IL-10 may be

induced by the chicken immune system in order to regulate the immune response and

inflammation caused by the virus.

Cytokines are considered important mediators of inflammation and regulators of

the immune response. The first defense against viral infection is an inflammatory

response, including release of inflammatory cytokines. IL-8 is a potent neutrophil and T

lymphocyte chemotactic factor released by a variety of mammalian cells (Oppenheim et

al., 1991), and according to Bruder and Kovesdi (1997) adenovirus infection triggers

Raf/mitogen-activated protein kinase pathways that induce synthesis of IL-8 and thus

contribute to the early inflammatory response. Increased expression of IL-8 in the present

study may be a reactive response of the host to reduce the viral load.

Page 125: Sequence analysis, pathogenicity and cytokine gene

114

In summary, in this study, the FAdV-4 virus was found in tissues and cloacal swabs from

the inoculated chickens and a strong antibody response was induced. However, no

clinical signs or gross or histological changes were detected. FAdV-4 infection was

associated with increased expression of IL-18, IFN-γ, IL-10 and IL-8 in liver and

increased expression of IL-18 and IFN-γ in cecal tonsil, providing useful information

about important mediators involved in immunity against HHS virus infection. Although

we have gained further insight into FAdV-associated cytokine gene expression, the

mechanisms and molecular events governing the production of cytokines during infection

still remain to be described. Finally, this study demonstrated that this FAdV-4 isolate

under experimental conditions was non-pathogenic and could be used as live vaccine

against HHS.

Acknowledgements

This work was supported by the Natural Sciences and Engineering Research

Council of Canada (NSERC), the Poultry Industry Council (PIC) of Ontario and the

Ontario Ministry of Agriculture, Food and Rural Affairs (OMAFRA). The authors thank

the Isolation Unit personnel for their professional animal care and assistance.

Page 126: Sequence analysis, pathogenicity and cytokine gene

115

Table 4.1. Numbers of positive tissues and viral genome copy numbers in tissues of chickens inoculated with FAdV-4.

* = viral genome copy number/10 µg of total tissue DNA was determined by qPCR § = number of positive tissues/number analyzed tissues

days p.i.

Inoculation route

tissue oral im

Liver 3.4x100* 3/4§ 7.5x100 2/4

3 Bursa 5.6x100 2/4 5x100 1/4

Cecal tonsil 2.6x101 4/4 5.3x102 4/4

Liver 1.4x102 3/4 8.3x100 4/4

5 Bursa 6.1x100 4/4 1.2x101 2/4

Cecal tonsil 5.7x102 4/4 1.3x102 4/4

Liver 2.2x100 2/4 2.3x100 2/4

7 Bursa 0x100 0/4 6x100 2/4

Cecal tonsil 2.4x101 1/4 3.5x100 3/4

Liver 2.8x100 4/4 5.9x100 2/4 14 Bursa 4.9x100 3/4 4.6x100 2/4

Cecal tonsil 1.2x101 4/4 8.6x100 4/4

Liver 1.1x101 4/4 2.4x101 4/4

21 Bursa 2.3x101 3/4 3.1x102 4/4

Cecal tonsil 6.4x103 4/4 3.6x102 3/4

Liver 6.5x102 4/4 3.5x102 4/4 28 Bursa 1.4x100 2/4 7.4x101 1/4

Cecal tonsil 1.7x102 4/4 6.5x102 4/4

Page 127: Sequence analysis, pathogenicity and cytokine gene

116

Table 4.2. Virus titers in the feces of chickens inoculated with FAdV-4.

Days Inoculation route post-infection Mean titers (pfu/ml) in swabs (log10 mean ± log10SEmean)

Chickens shedding the virus (%)†

Oral Im 9.9x103 (3.30±0.22) 1.4x103 (0.69±0.24)

3 P≤0.001 (93) (25) 1.2x104 (3.15±0.32) 1.2x103 (1.01±0.33)

5 P≤0.001 (83) (32)

1.3x102 (0.24±0.17) 6.8x102 (0.54±0.31)

10 P=0.49 (13) (20)

3.9x101 (0.49±0.23) 2.5x101 (0.15±0.10)

14 P=0.26 (27) (13)

3.3x101 (0.20±0.14) 2.5x101 (0.34±0.23)

21 P=0.25 (17) (17)

0 1.2x101 (0.12±0.12)

28 P=0.34 (0) (12)

The virus titers were determined by plaque assay.

P values < 0.05 are statistically significant. †30 chickens at day 0; 26 and 22 chickens at days 3 and 5, respectively; 18 and 14 chickens at days 10 and 14, respectively; 10 and 6 chickens at days 21 and 28, respectively.

Page 128: Sequence analysis, pathogenicity and cytokine gene

117

Figure 4.1. Antibody response to viral proteins in chickens inoculated with FAdV-4 by the oral and intramuscular routes and mock infected as measured by S/P ratios. Serum samples were collected at 7, 14, 21 and 28 days post-infection. Dark gray and white columns represent im and orally inoculated chickens, respectively. Ligth gray and black columns represent mock-infected chickens. The error bars correspond to 95% confidence intervals. Significant differences among mock infected and infected birds are indicated by asterisks (*). Panel A. Serum samples were tested against homologous virus, FAdV-4.

Page 129: Sequence analysis, pathogenicity and cytokine gene

118

Figure 4.1. Panel B. Serum samples were tested against heterologous virus, FAdV-9.

Page 130: Sequence analysis, pathogenicity and cytokine gene

119

Figure 4.2. Cytokine mRNA expression in cecal tonsil of chickens infected with FAdV-4. The groups were as follows: Uninfected (negative control) chickens, and FAdV-4 infected chickens. Target and reference gene expression in cecal tonsil was quantified by real-time RT-PCR using SYBR Green. Target gene expression is presented relative to β-actin expression and normalized to a calibrator. The results are diagrammed as whisker boxes with medians. Boxes represent interquartile ranges and whiskers indicate extreme values. The difference in cytokine expression between groups was assessed by Mann-Whitney test and comparisons were considered significant at P≤0.05 (*). ○, ● symbols represent values which were identified as outliers. Numbers 1, 3, 5, 7, and 10 represent days post infection when samples were collected.

Page 131: Sequence analysis, pathogenicity and cytokine gene

120

Figure 4.3. Cytokine mRNA expression in liver of chickens infected with FAdV-4. The groups were as follows: Uninfected (negative control) chickens, and FAdV-4 infected chickens. Target and reference gene expression in liver was quantified by real-time RT-PCR using SYBR Green. Target gene expression is presented relative to β-actin expression and normalized to a calibrator. The results are diagrammed as whisker boxes with medians. Boxes represent interquartile ranges and whiskers indicate extreme values. The difference in cytokine expression between groups was assessed by Mann-Whitney test and comparisons were considered significant at P≤0.05 (*). ○, ● symbols represent values which were identified as outliers. Numbers 1, 3, 5, 7, and 10 represent days post infection when samples were collected.

Page 132: Sequence analysis, pathogenicity and cytokine gene

121

Page 133: Sequence analysis, pathogenicity and cytokine gene

122

Figure 4.4. Cytokine mRNA expression in spleen of chickens infected with FAdV-4. The groups were as follows: Uninfected (negative control) chickens, and FAdV-4 infected chickens. Target and reference gene expression in spleen was quantified by real-time RT-PCR using SYBR Green. Target gene expression is presented relative to β-actin expression and normalized to a calibrator. The results are diagrammed as whisker boxes with medians. Boxes represent interquartile ranges and whiskers indicate extreme values. The difference in cytokine expression between groups was assessed by Mann-Whitney test and comparisons were considered significant at P≤0.05 (*). ○, ● symbols represent values which were identified as outliers. Numbers 1, 3, 5, 7, and 10 represent days post infection when samples were collected.

*

*

Page 134: Sequence analysis, pathogenicity and cytokine gene

123

CHAPTER 5 Molecular characterization of the fiber gene of fowl adenovirus serotypes 8 and 11

Helena Grgić1, Davor Ojkić2, Éva Nagy1*

1Department of Pathobiology, Ontario Veterinary College, University of Guelph, Guelph, Ontario, N1G 2W1, Canada. 2Animal Health Laboratory, University of Guelph, Guelph, Ontario, N1G2W1, Canada

*Corresponding author: Éva Nagy E-mail: [email protected] Tel: +1 519 824 4120 ext 54783 FAX: +1 519 824 5930

The GenBank accession numbers of the sequences reported in this paper are:

BankIt1493149 Fiber JQ034214 BankIt1493157 Fiber JQ034215 BankIt1493162 Fiber JQ034216 BankIt1493466 Fiber JQ034217 BankIt1493468 Fiber JQ034218 BankIt1493469 Fiber JQ034219 BankIt1493470 Fiber JQ034220 BankIt1494294 Fiber JQ034221 Author’s contributions

HG designed and performed experiment, carried out analysis and wrote the paper. DO

provided viruses for the research and revised manuscript. EN provided guidance during

the research and critically revised manuscript

Short communication to be submitted to Virus Genes

Page 135: Sequence analysis, pathogenicity and cytokine gene

124

Abstract

The fiber protein as one of the major constituents of the adenoviral capsid is

involved in virus entry and it has been implicated in the variation of virulence of fowl

adenoviruses (FAdVs). The most commonly isolated FAdVs in Canada belong to

serotype 8 and 11, according to current surveillance data. Therefore, four FAdV-8 and

four FAdV-11 isolates were selected to examine the possible role of the fiber gene

sequence in FAdV virulence. On the basis of the established fiber structural model, the

FAdV fiber sequence can be divided into tail, shaft and head domains comprising some

specific features. The conserved ‘RKRP’ (aa 17 to aa 20) motif fits the consensus

sequence predicted for the NLS, while the ‘VYPF’ (aa 53 to aa 56) sequence motif is

suggested to be involved in the penton base interaction. Similarly to mammalian

adenoviruses, 17 pseudorepeats have been detected in the FAdV-8 fiber shaft region, with

each repetition comprising 16 aa on average. The shaft regions of FAdV-11 fibers consist

of 20 pseudorepeats and each repetition comprising 18 aa on average. There was a 144 nt

difference between the fiber genes of the two FAdV serotypes. In the shaft region the

characteristic well-conserved TLWT motif that marks the beginning of the fiber head

domain was not evident among examined FAdVs. The FAdV-11 isolates had 99.3% aa

sequence identity and 99.1% similarity to each other and there was no conserved aa

substitution within the fibers The FAdV-8 fiber proteins showed an overall 89% aa

sequence identity and 93.4% similarity to each other and 21 nonsynonymous mutations

were detected. Virulence markers were not detected in the analyzed fiber genes of the

two FAdV serotypes.

Page 136: Sequence analysis, pathogenicity and cytokine gene

125

Keywords:

FAdV-8; FAdV-11; fiber gene sequences; conserved TLWT motif; virulence markers.

Abbreviations: FAdV-8, fowl adenovirus serotype 8; FAdV-11, fowl adenovirus

serotype 11; NLS, nuclear localization signal.

Page 137: Sequence analysis, pathogenicity and cytokine gene

126

Members of the family Adenoviridae are non-enveloped double-stranded DNA

viruses. The 12 serotypes of fowl adenoviruses (FAdV) are grouped into five species

(FAdV-A toFAdV-E) on the basis of restriction fragment length polymorphism (RFLP)

profiles and cross-neutralization (Benkő et al., 2005). Infection with several FAdV

serotypes is associated with diseases such as inclusion body hepatitis (IBH) and

infectious hydropericardium syndrome. The outcome of FAdV infection is sometimes

influenced by concurrent infection with immunosuppressive agents such as infectious

bursal disease or chicken anemia viruses (Fadly et al., 1975). However, significant

economic losses have also been encountered in the absence of co-infection with

immunosuppressive agents (Saifuddin and Wilks, 1990). The pathogenicity of FAdVs

differs even among strains of the same serotype. According to current surveillance data,

in Canada the most commonly isolated FAdVs belong to serotype 8 and 11 (Ojkic et al.,

2008a, b). It has been reported that the virulence of FAdV-8 is determined by the fiber

protein alone (Pallister et al., 1996). Variations in the fiber gene sequences have also

been implicated in the differences in tissue tropism and virulence of the canine

adenoviruses (Rasmussen et al., 1995). The objective of this study was to examine the

possible role of the fiber gene in FAdV virulence by analysis of sequences from

pathogenic and non-pathogenic FAdV-11 and FAdV-8 isolates.

Four FAdV-11 (species Fowl adenovirus D) and four FAdV-8 (species Fowl

adenovirus E) isolates were kindly provided by the Animal Health Laboratory (AHL),

University of Guelph. Two of the serotype 11 and two of the serotype 8 viruses were

isolated from flocks with clinical signs of IBH, and IBH was confirmed as the final

Page 138: Sequence analysis, pathogenicity and cytokine gene

127

diagnosis. The other four viruses were isolated from flocks with no clinical signs of IBH

(Table 5.1). The genotyping of the isolates was determined by sequence analysis of the

hexon gene as described before (Ojkic et al., 2008). The viruses were plaque purified and

propagated in chicken hepatoma (CH-SAH) cells as described (Alexander et al., 1998).

CH-SAH cells were infected at a multiplicity of infection of 0.5. Cells and supernatants

were collected when extensive cytopathic effect was observed. Virus purification and

viral DNA preparation were performed as described (Ojkic and Nagy, 2001).

FAdV-8 fiber genes were amplified by PCR with the primer set: FibFor8 (5’-CTG

GTG AAG TCC CTC TCC GC –3’) and FibRev8 (5’-GAC ATA GGG GAA CAT CAA

–3’) (Grgic et al., 2011). For FAdV-11 the primers were designed based on a comparison

of the sequences of the fiber genes of FAdV-9 and FAdV-2, since these viruses are in the

same species and they have a close genetic relationship (Corredor et al., 2008). The

following primer set was used to amplify FAdV-11 fiber genes: FibFor11 (5’-CAC TCG

CCC TAG CTT ATG-3’) and FibRev11 (5’-TGT TAA CAA AGG TCG ACT G –3’).

PCR was performed with 35 cycles of 94˚C for 30 sec, 58˚C (for FAdV-8) or 52˚C (for

FAdV-11), and 72˚C for 30 sec. The PCR products were sequenced by the Laboratory

Services, University of Guelph. Sequence assembly was performed with Vector NTI

AdvanceTM11. The coded proteins were identified using BLAST analysis with the

GenBank database and degree of identities and similarity was identified using FASTA3

(EMBL-EBI).

Page 139: Sequence analysis, pathogenicity and cytokine gene

128

The role of fiber protein in virulence of adenoviruses is not well understood,

therefore we compared the fiber sequences of serotypes 8 and 11 FAdVs isolated from

clinical submissions to the Animal Health Laboratory and originated from cases with or

without the clinical diagnosis of inclusion body hepatitis. Some FAdV serotypes have

two fiber genes, such as FAdV-1 and FAdV-4 (Chiocca et al., 1996; Griffin and Nagy,

2011), however there was only one fiber gene for all studied viruses of both serotypes.

The FAdV-8 and FAdV-11 fiber genes encoded from 524 to 525 amino acids (aa) and

572 aa, respectively.

On the basis of the established fiber structural model for mastadenoviruses (Green

et al., 1983) the FAdV-8 and FAdV-11 fiber sequences can also be divided into tail, shaft

and head domains and have some specific features.

For both serotypes the first 63 N-terminal aa with conserved clusters make up the

tail region (Fig.5.1). These conserved patterns were identical for both serotypes 8 and 11

viruses. Most adenovirus fiber sequences contain the nuclear localization signal (NLS).

Moreover, all of the known nuclear localization signals are rich in basic aa residues. A

potential NLS sequence with stretch of positively charged aa was identified within the

predicted FAdV-8 and FAdV-11 sequences. The KRAR (aa 2 to aa 5) sequence

conserved among fibers from several human adenovirus serotypes, such as HAdV-5,

HAdV-3, and HAdV-7 (Chroboczek and Jacrot, 1987; Hong et al., 1988; Signas et al.,

1985) was not detected for FAdV-8 and FAdV-11 isolates. However, a similar motif

‘RKRP’ (aa 17 to aa 20) was present in all analyzed FAdV fibers of both serotypes. The

Page 140: Sequence analysis, pathogenicity and cytokine gene

129

‘RKRP’ motif could have an NLS function that is essential for the targeting of the fiber to

the nucleus and fits the consensus sequence predicted for the NLS by PSORT II finder

for NLS. Another well conserved motif of the tail region, the ‘VYPY’ sequence found in

many mammalian adenovirus fiber proteins is most likely involved in the penton base

interaction (Caillet-Boudin, 1989; Hong and Boulanger, 1995). Equivalent sequence

motif in FAdV fiber tails of both serotypes FAdV-8 and 11 is ‘VYPF’ (aa 53 to aa 56)

(Fig.5.1). Similar motif has been reported in the fiber tail of ovine adenovirus (Vrati et

al., 1995), egg drop syndrome virus (Hess et al., 1997), and bovine adenovirus 7 (Li and

Tikoo, 2002). Interestingly, Hess et al. (1995) reported a ‘VYPY/F’ sequence in the tail

region of both FAdV-1 fibers, long and short. In Fig.5.1 we have also indicated the poly

(G) stretch found in the tail domain, though the poly (G) might be only the connection

between the tail and shaft domains without belonging to either. A poly (G) stretch has

been reported as part of the long fiber tail region of FAdV-1 (Hess et al., 1995), while the

same stretch was absent in the FAdV-10 fiber gene (Sheppard et al., 1998).

Similar to mammalian adenoviruses (Green et al., 1983; van Raaij et al., 1999),

17 repetitive elements (pseudorepeats) have been identified in the FAdV-8 fiber shaft

region, with each repetition comprising 16 aa on average (Fig. 5.2). The shaft regions of

FAdV-11 fibers consist of 20 pseudorepeats with an average of 18 aa each (Fig. 5.2). It is

important to note that in most mastadenovirus fiber shafts the last pseudorepeat before

the head domain contains a conserved amino acid sequence (KLGXGLXFD/N). This

conserved sequence was absent in both FAdV-8 and -11 fibers and has not been reported

for FAdV-1 and FAdV-10 (Hess et al., 1995; Sheppard et al., 1998).

Page 141: Sequence analysis, pathogenicity and cytokine gene

130

In human adenoviruses there is a characteristic well-conserved TLWT sequence

in the first β-strand of the head (Xia et al., 1994) that marks the beginning of the fiber

head domain. This typical knob sequence was absent in the examined FAdV-8 and

FAdV-11 sequences and was also absent in the reported FAdV-1, FAdV-4 and FAdV-10

sequences (Hess et al., 1995; Mase et al., 2010; Sheppard et al., 1998) making it

impossible to define the actual start of the head domain. The absence of TLWT motif has

also been reported for the fiber proteins of the OAV287 (Vrati et al., 1995), BAV7 (Li

and Tikoo, 2002), and EDSV (Hess et al., 1997). Further analysis of the C-terminal knob

domains has shown no homology with the mammalian fiber heads. Nevertheless, the C-

terminal 154 aa of the FAdV fiber sequences probably constitutes the head region.

The FAdV-11 isolates had 99.3% aa sequence identity and 99.1% similarity to

each other. There was no conserved amino acid substitution within the fibers of the

serotype 11 viruses. The complete aa sequence alignment of the four FAdV-11 fibers and

FAdV-1 short and long fibers using FASTA3 (EMBL-EBI) method revealed 27.3%

identity and 42.9% similarity for the short fiber while for the long fiber the percent

identity and similarity was only 18.6 and 27.2%, respectively (Table 5.2). The FAdV-11

fiber protein and FAdV-9 fiber protein, which belongs to the same species group showed

an overall 84.1% aa sequence identity and 91.6% similarity. Very low 19.1% identity

and 29.1% similarity, has been observed when the fiber protein sequences of FAdV-11

isolates were compared to the FAdV-4 fiber protein that belongs to species group C

(Table5. 2). When the aa sequences of the fiber proteins of species D and E were

compared there were 47 to 50 aa differences between the fiber proteins, showing an

Page 142: Sequence analysis, pathogenicity and cytokine gene

131

overall 51.4% identity and 64.5% of similarity (Table 5.2). The fiber proteins of the

analyzed four FAdV-8 isolated showed an overall 89% aa sequence identity and 93.4%

similarity to each other. In total, 21 nonsynonymous mutations were detected within the

fiber of FAdV-8 isolates from birds with signs of IBH compared to the isolates from non-

IBH cases (Table 5.3). Nonsynonymous mutations are highlighted in Table 5.3. Because

of such a large number of nonsynonymous mutations it is difficult to speculate about the

effect of these mutations on the pathogenicity of these isolates. Further, comparison of

the FAdV-8 fibers and the FAdV-1 short and long fibers revealed 28.2% identity and

41.8% similarity for the short, while for the long fiber the percent identity and similarity

was only 19.2% and 27.4%, respectively (Table 5. 2). When the Ontario FAdV-8 fiber

proteins were compared to the mildly virulent CFA3 and hypervirulent CFA40 (Pallister

et al., 1996) isolates, the percent identity and similarity on average was 70.8/74.2 and

65.8/69.9, respectively (Table 5.2). The significantly decreased identity and similarity

was due to difference in size, the Ontario isolates were on average 96 aa longer.

Moreover, the last 65 aa representing 15% of the fiber protein of CFA40 and CFA3

showed the highest divergence compared to the Ontario fiber proteins. Interestingly, the

highest divergence was in the knob region. These results are different from the data

published by Mase et al. (2010) who showed that nt sequences of Japanese FAdV-4

isolates were identical to each other and had 98.5% homology with isolates from India

and Pakistan. Similarly, FAdV-1 isolates from 4 outbreaks of gizzard erosions in broilers

shared 100% identity at the protein level. Moreover, the short fiber of the pathogenic

isolates was 100% identical to the apathogenic Ote strain (Marek et al., 2010).

Page 143: Sequence analysis, pathogenicity and cytokine gene

132

We have shown that FAdV-8 and FAdV-11, representing species E and D have

only one fiber gene. There was a 144 nt difference between the fiber genes of the two

FAdV serotypes, mapped to the shaft region. According to our data virulence might not

be associated only with sequences of the fiber gene. In order to identify the critical sites

responsible for the pathogenicity of FAdVs further molecular analyses would be required.

Moreover, to estimate the influence of adenoviral fiber genetics on pathogenesis,

experimental infections would be necessary.

Acknowledgements

This work was supported by the Natural Sciences and Engineering Research

Council of Canada (NSERC), the Canadian Poultry Research Council (CPRC), and the

Ontario Ministry of Agriculture, Food and Rural Affairs (OMAFRA).

Page 144: Sequence analysis, pathogenicity and cytokine gene

133

Table 5.1. Description of the isolates.

* = FAdV-8 isolate from the Chapter 2

Designation AHL code Pathogenic/Non-pathogenic GenBank #

FAdV-8-ON-P1 04-53357-125 IBH case BankIt1493157 Fiber JQ034215

FAdV-8-ON-P2 06-41265-07 IBH case BankIt1493162 Fiber JQ034216

FAdV-8-ON-NP1 05-50052-3180 Non-IBH case BankIt1494294 Fiber JQ034221

FAdV-8-ON-NP2 06-16340-336-OH Non-IBH case BankIt1493149 Fiber JQ034214

FAdV-8* N/A IBH case - non-pathogenic HG-GU734104

FAdV-11-ON-P1 06-25854-1 IBH case BankIt1493469 Fiber JQ034219

FAdV-11-ON-P2 05-17766 IBH case BankIt1493470 Fiber JQ034220

FAdV-11-ON-NP1 05-50052-2924-1 Non-IBH case BankIt1493466 Fiber JQ034217

FAdV-11-ON-NP2 05-50052-3181 Non-IBH case BankIt1493468 Fiber JQ034218

Page 145: Sequence analysis, pathogenicity and cytokine gene

134

Table 5.2. Percent identities/similarities between FAdV-8 and FAdV-11 fiber proteins and their homologues.

S= short fiber, L= long fiber

FAdV-8/E (U40587) = CFA 40 hypervirulent, FAdV-8/E (U40588) = CFA 3 mildly virulent

Percent identities and similarities were calculated using FASTA3 (EMBL-EBI).

IsolatesType/Accession FAdV-8-ON-

P1 FAdV-8-ON-

P2 FAdV-8-ON-

NP1 FAdV-8-ON-

NP2 FAdV-11-

ON-P1 FAdV-11-

ON-P2 FAdV-11-ON-NP1

FAdV-11-ON-NP2

FAdV-8-ON-P1 - 86.3/92 94.3/96.8 92.2/94.9 - - - -

FAdV-8-ON-P2 - - 82.3/90.1 81.4/88.8 - - - -

FAdV-8-ON-NP1 - - - 97.7/97.7 - - - -

FAdV-8-ON-NP2 - - - - - - - -

FAdV-8 (U40587) 67/71.9 66.5/69.2 66.5/71.5 64/68.5 38.6/49.6 40/50.8 40/50.8 40/50.8

FAdV-8 (U40588) 75.3/77.2 67.4/72.3 67.4/72.3 75.3/76.5 42.2/52.6 43.6/53.8 43.6/53.8 43.6/53.8

FAdV-8 (GU734104) 86.5/92.2 99.8/99.6 83.1/90.3 81.6/89 49.2/63.7 50.6/64.9 50.6/64.9 50.6/64.9

FAdV-11-ON-P1 51.6/64.2 52.9/65.4 52.9/65.4 52.9/65.4 - 97.2/97 97.2/97 100/99.8

FAdV-11-ON-P2 50.9/63.3 52.2/64.5 52.2/64.5 52.2/64.5 - - 100/99.8 100/99.8

FAdV-11-ON-NP1 51.3/65.1 52.7/66.3 52.7/66.3 52.7/66.3 - - - 100/99.8

FAdV-11-ON-NP2 49/63.6 50.3/64.8 50.3/64.8 50.3/64.8 - - - -

FAdV-1/S (AA46933) 29.2/41.9 29/41.4 28.7/42.1 28.3/41.8 27/42.6 27.4/43 27.4/43 27.4/43

FAdV-1/L (AA46933) 19.5/27.9 17.9/25 19.7/28.5 19.8/28.3 18.6/27.2 18.6/27.2 18.6/27.2 18.6/27.2

FAdV-4/1 (GU188428) 24.3/36.2 24.7/34.8 23.7/34.9 23.4/34.5 18.8/28.9 19.2/29.2 19.2/29.2 19.2/29.2

FAdV-4/2 (GU188428) 39.3/54.1 38/54.1 39.4/53.4 38.8/52.8 35.9/51.6 35.4/49.3 35.4/49.3 35.4/49.3

FAdV-9 (AF083975) 52.4/67.3 52.5/66.5 52.8/67 52.8/66.1 82/89.7 84.8/92.3 84.8/92.3 84.8/92.3

Page 146: Sequence analysis, pathogenicity and cytokine gene

135

Table 5.3. Differences in amino acids between FAdV-8 isolates derived from flocks with or without the clinical signs or IBH.

Amino acids at position

Origin of viruses: 374 375 390 408 443 444 452 458 468 471 472 473 474 477 484 494 503 506 513 522 523

No signs of IBH P S A E V D G S V N A T I Q S S S L Y D V

Signs of IBH A A/N E/T Q L E D - C T V R V R L A/T A M F N A

Page 147: Sequence analysis, pathogenicity and cytokine gene

136

FAdV-8 FAdV-8-ON-NP1 MATSTPHAFSFGQIGSRKRPAGGDGERDASKVPKMQTPAPSATANGNDELDLVYPFWLQN 60 FAdV-8-ON-NP2 MATSTPHAFSFGQIGSRKRPAGGDGERDASKVPKMQTPAPSATANGNDELDLVYPFWLQN 60 FAdV-8-ON-P2 MATSTPHAFSFGQIGSRKRPAGGDGERDASKVPKMQTPAPSATANGNDELDLVYPFWLQN 60 FAdV-8-ON-P1 MATSTPHAFSFGQIGSRKRPAGGDGERDASKVPKMQTPAPSATANGNDELDLVYPFWLQN 60 ************************************************************ FAdV-8-ON-NP1 GSTGGGGGGG 70 FAdV-8-ON-NP2 GSTGGGGGGG 70 FAdV-8-ON-P2 GSTGGGGGGG 70 FAdV-8-ON-P1 GSTGGGGGGG 70 ********** FAdV-11 FAdV-11-ON-NP1 MAKSTPFAFSMGQHSSRKRPADSENTQNASKVAKTQTSATRAGVDGNDDLNLVYPFWLQN 60 FAdV-11-ON-NP2 MAKSTPFAFSMGQHSSRKRPADSENTQNASKVAKTQTSATRAGVDGNDDLNLVYPFWLQN 60 FAdV-11-ON-P2 MAKSTPFAFSMGQHSSRKRPADSENTQNASKVAKTQTSATRAGVDGNDDLNLVYPFWLQN 60 FAdV-11-ON-P1 MAKSTPFAFSMGQHSSRKRPADSENTQNASKVRQNADFCHARRCRRKYDLNLVYPFWLQN 60 ******************************** :. . : ************ FAdV-11-ON-NP1 STSGGGGGG 69 FAdV-11-ON-NP2 STSGGGGGG 69 FAdV-11-ON-P2 STSGGGGGG 69 FAdV-11-ON-P1 STSGGGGGG 69 ********* Figure 5.1. Amino acid sequence alignment of the tail sequence of FAdV-8 and FAdV-11 isolates made by CLUSTALW. The yellow arrow indicates motif which has NLS function. The blue arrow shows conserved sequence motif found in mammalian adenovirus tail region involved in penton base interaction. The green arrow shows poly (G) stretch which is connection between tail and shaft region.

Page 148: Sequence analysis, pathogenicity and cytokine gene

137

MGTSTPHAFSFGQIGSRKRPAGGDGERDASEVPKMETPAPSAT ANGNDELDLVYPFWLQNGSTGGGGGGGSGGN Tail

β βββ βββ βββ β 1 PS LNP PFL DP NGP LTVQ 2 NN LLK VNT TA PIA VTN 3 KA LTL SYD PE SLE LTNQ 4 QQ LAV KID PE GPL KATT 5 EG IQL SVD PT TLE V 6 DD VDW ELT VK LDP NGPLDSSA 7 TG ITV RVD DT LLV EDDGSGQG 8 KE LGV HLN PD GPI TADQ Shaft 9 NG LDL EID NQ TLK ITPGSA

10 GG VLS VQL KP QGG LNSQS 11 DG IQV VTQ NS IEV DN 12 GA LDV KVV AN GPL STTP 13 NG LTL NYD TG DFT VNA 14 GT LSI LRN PS LVA NAYLTSGA 15 RT LQQ FTA KG ENS SQFSFP 16 CA YYL QQW LS DGL IF17 SS LYL KLD RT RFT

GMSSDPSYQNARYFTFWVGGGAAMNLSQLSTP TITPSTTEWTAFAPALNYSGAPAFVYDANPVF TIYFQPNTGRVDTYLPVLTGDWKTSSTYNPGT Head ITLVVRNATIQLQSQSTFTTSVCYNFRCQNSGIF NNNATSGTLTLGPIFYSCPALSTADVS

Figure 5. 2. Regions of FAdV-8 and FAdV-11 fibers. Bold indicates hydrophobic amino acids. Residues forming β strands are indicated above pseudorepeats. Shaft region pseudorepeats are numbered on the left side.

Page 149: Sequence analysis, pathogenicity and cytokine gene

138

MAKSTPFAFSMGQHSSRKRPADSENTQNASKVAKTQTSATR AGVDGNDDLNLVYPFWLQNSTSGGGGGGSGGN Tail

β βββ βββ βββ β 1 PS LNP PFI DP NGP LYVQ 2 NS LLY VKT TA PIE VEN 3 KS LAL AYD SS LDV DAQ 4 NQ LQV KVD AE GPI RISP 5 DG LDI AVD PS TLE VD 6 DE WEL TVK LD PAG PISSSS 7 AG INI RVD DT LLI EDDDT 8 AQ VKE LGV HL NPN GPITADQ 9 DG LDL EVD PQ TLT VTTSGAT 10 GG VLG VLL KP SGG LQTSI Shaft 11 QG IGV AVA DT LTI SSNT 12 GT VEV KTD PN GSI GSSS 13 NG IAV VTD PA GPL TTSS 14 NG LSL KLT PN GSI QSSS 15 TG LSV QTD PA GPI TSGA 16 NG LSL SYD TS DFT VSQ 17 GM LSI IRN PS AYP DAYLESG 18 TN LLN NYT AY AEN S 19 SN YKF NCA YF LQS WYS 20 NG LVT SSL YL KIN RDN

LTSLPSGQLSENAKYFTFWVPTYESMNLSNVAT PTITPSSVPWGAFLPAQNCTSNPAFKYYLTQP PSIYFEPESGSVQTFQPVLTGDWDTNTYNPGTV Head QVCILPQTVVGGQSTFVNMTCYNFRCQNPGIF KVAASSGTFTIGPIFYSCPTNKLTQP

Page 150: Sequence analysis, pathogenicity and cytokine gene

139

Figure 5.3. Supplementary data. Amino acid alignment of FAdV-8 fibers derived from flocks with and without signs of IBH. Highlighted amino acids represent conserved substitutions. Symbol “:” means that conserved substitutions have been observed while, “.” means that semi-conserved substitutions are observed. FAdV-8-ON-NP1 MATSTPHAFSFGQIGSRKRPAGGDGERDASKVPKMQTPAPSATANGNDELDLVYPFWLQN 60 FAdV-8-ON-NP2 MATSTPHAFSFGQIGSRKRPAGGDGERDASKVPKMQTPAPSATANGNDELDLVYPFWLQN 60 FAdV-8-ON-P1 MATSTPHAFSFGQIGSRKRPAGGDGERDASKVPKMQTPAPSATANGNDELDLVYPFWLQN 60 FAdV-8-ON-P2 MATSTPHAFSFGQIGSRKRPAGGDGERDASKVPKMQTPAPSATANGNDELDLVYPFWLQN 60 ************************************************************ FAdV-8-ON-NP1 GSTGGGGGGG-SGGNPSLNPPFLDPNGPLAVQNNLLKVNTAAPITVTNKALTLAYEPDSL 119 FAdV-8-ON-NP2 GSTGGGGGGG-SCGNPSLIPPFLDPNGPLTVKNNLLKVNTAAPITVTNKALTLAYEPDSL 119 FAdV-8-ON-P1 GSTGGGGGGG-SGGNPSLNPPFLDPNGPLAVQNNLLKVNTAAPITVTNKALTLAYEPDSL 119 FAdV-8-ON-P2 GSTGGGGGGGGSGGNPSLNPPFLDPNGPLAVQNNLLKVNTAAPITVANKALTLAYEPDSL 120 ********** * ***** **********:*:**************:************* FAdV-8-ON-NP1 ELTNQQQLAVKIDPEGPLKATTEGIQLSVDPTTLEVDDVDWELTVKLDPDGPLDSSATGI 179 FAdV-8-ON-NP2 ELTNQQQLAVKIDPEGPLKATTEGIQLSVDPTTLEVDDVDWELTVKLDPDGPLDSSATGI 179 FAdV-8-ON-P1 ELTNQQQLAVKIDPEGPLKATTEGIQLSVDPTTLEVDDVDWELTVKLDPDGPLDSSATGI 179 FAdV-8-ON-P2 ELTNQQQLAVKIDPEGPLKATTEGIQLSVDPTTLEVDDVDWELTVKLDPDGPLDSSATGI 180 ************************************************************ FAdV-8-ON-NP1 TVRVDETLLVEDDGSGQGKELGVHLNPDGPITADQNGLDLEIDNQTLKITPGSAGGVLSV 239 FAdV-8-ON-NP2 TVRVDETLLVEDDGSGQGKELGVHLNPDGPITADQNGLDLEIDNQTLKITPGSAGGVLSV 239 FAdV-8-ON-P1 TVRVDETLLVEDDGSGQGKELGVHLNPDGPITADQNGLDLEIDNQTLKITPGSAGGVLSV 239 FAdV-8-ON-P2 TVRVDETLLIEDVGSGQGKELGVNLNPTGPITADDQGLDLEIDNQTLKVNSVTGGGVLAV 240 *********:** **********:*** ******::************:.. :.****:* FAdV-8-ON-NP1 QLKPQGGLNSQSDGIQVVTQNSIEVDNGALDVKVVANGPLSTTPNGLTLNYDTGDFTVNA 299 FAdV-8-ON-NP2 QLKPQGGLNSQSDGIQVVTQNSIEVDNGALDVKVVANGPLSTTPNGLTLNYDTGDFTVNA 299 FAdV-8-ON-P1 QLKPQGGLNSQSDGIQVVTQNSIEVDNGALDVKVVANGPLSTTPDGLTLNYDTGDFTVNA 299 FAdV-8-ON-P2 QLKSQGGLTAQTDGIQVNTQNSITVTNGALDVKVAANGPLESTDTGLTLNYDPGDFTVNA 300 ***.****.:*:***** ***** * ********.*****.:* *******.******* FAdV-8-ON-NP1 GTLSILRNPSLVANAYLTSGARTLQQFTAKGENSSQFSFPCAYYLQQWLSDGLIFSSLYL 359 FAdV-8-ON-NP2 GTLSILRNPSLVANAYLTSGARTLQQFTAKGENSSQFSFPCAYYLQQWLSDGLIFSSLYL 359 FAdV-8-ON-P1 GTLSILRNPSLVANAYLTSGASTLQQFTAKGENSSQFSFPCAYYLQQWLSDGLIFSSLYL 359 FAdV-8-ON-P2 GTLSIIRDPALVANAYLTSGASTLQQFTAKSENSSQFSFPCAYYLQQWLSDGLVLSSLYL 360 *****:*:*:*********** ********.**********************::***** FAdV-8-ON-NP1 KLDRTRFTGMSSDPSYQNARYFTFWVGGGAAMNLSQLSTPTITPSTTEWTAFAPALNYSG 419 FAdV-8-ON-NP2 KLDRTRFTGMSSDPSYQNARYFTFWVGGGAAMNLSQLSTRTITPSPPEWMAFAPALNYSG 419 FAdV-8-ON-P1 KLDRTRFTGMSSDAAYQNAKYFTFWVGGGEAMNLSQISTPTITPSTTQWNAFAPALNYSG 419 FAdV-8-ON-P2 KLDRAQFTNMPTGANYQNARYFTFWVGAGTSFNLSTLTEPTITPNTTQWNAFAPALDYSG 420 ****::**.*.:.. ****:*******.* ::*** :: ****...:* ******:*** FAdV-8-ON-NP1 APAFVYDANPVFTIYFQPNTGRVDTYLPVLTGDWKTSSTYNPGTITLVVRNATIQLQSQS 479 FAdV-8-ON-NP2 APAFVYDGNAVCTIYFQPNPGRVDTYLPVLTGDWKTSSTYNPGTITLVVRNATIQLQSQS 479 FAdV-8-ON-P1 APAFVYDANPVFTIYFQPNTGRLETYLPVLTDDWKTS-TYNPGTITLCVRTVRVQLRSQG 478 FAdV-8-ON-P2 APPFIYDASSVVTIYFEPTSGRLESYLPVLTDNWSQT--YNPGTVTLCVKTVRVQLRSQG 478 **.*:**...* ****:*..**:::******.:*. : *****:** *:.. :**:**. FAdV-8-ON-NP1 TFTTSVCYNFRCQNSGIFNNNATSGTLTLGPIFYSCPALSTADVS 524 FAdV-8-ON-NP2 TFTTSVCYNFRCQNSGIFNNNATSGTLTLGPIFYSCPALSTADVS 524 FAdV-8-ON-P1 TFNMLVCYNFRCQNAGIFNNNATAGTMTLGPIFFSCPALSTANAP 523 FAdV-8-ON-P2 TFSTLVCYNFRCQNTGIFNSNATAGTMTLGLIFFSCPALSTANAP 523 **. *********:****.***:**:*** **:********:..

Page 151: Sequence analysis, pathogenicity and cytokine gene

140

Figure 5.4. Supplementary data. Amino acid alignment of FAdV-11 fibers derived from flocks with and without signs of IBH. FAdV-11-ON-NP1 MAKSTPFAFSMGQHSSRKRPADSENTQNASKVAKTQTSATRAGVDGNDDLNLVYPFWLQN 60 FAdV-11-ON-NP2 MAKSTPFAFSMGQHSSRKRPADSENTQNASKVAKTQTSATRAGVDGNDDLNLVYPFWLQN 60 FAdV-11-ON-P2 MAKSTPFAFSMGQHSSRKRPADSENTQNASKVAKTQTSATRAGVDGNDDLNLVYPFWLQN 60 FAdV-11-ON-P1 MAKSTPFAFSMGQHSSRKRPADSENTQNASKVRQNADFCHARRCRRKYDLNLVYPFWLQN 60 ******************************** :. . : ************ FAdV-11-ON-NP1 STSGGGGGGSGGNPSLNPPFIDPNGPLYVQNSLLYVKTTAPIEVENKSLALAYDSSLDVD 120 FAdV-11-ON-NP2 STSGGGGGGSGGNPSLNPPFIDPNGPLYVQNSLLYVKTTAPIEVENKSLALAYDSSLDVD 120 FAdV-11-ON-P2 STSGGGGGGSGGNPSLNPPFIDPNGPLYVQNSLLYVKTTAPIEVENKSLALAYDSSLDVD 120 FAdV-11-ON-P1 STSGGGGGGSGGNPSLNPPFIDPNGPLYVQNSLLYVKTTAPIEVENKSLALAYDSSLDVD 120 ************************************************************ FAdV-11-ON-NP1 AQNQLQVKVDAEGPIRISPDGLDIAVDPSTLEVDDEWELTVKLDPAGPISSSSAGINIRV 180 FAdV-11-ON-NP2 AQNQLQVKVDAEGPIRISPDGLDIAVDPSTLEVDDEWELTVKLDPAGPISSSSAGINIRV 180 FAdV-11-ON-P2 AQNQLQVKVDAEGPIRISPDGLDIAVDPSTLEVDDEWELTVKLDPAGPISSSSAGINIRV 180 FAdV-11-ON-P1 AQNQLQVKVDAEGPIRISPDGLDIAVDPSTLEVDDEWELTVKLDPAGPISSSSAGINIRV 180 ************************************************************ FAdV-11-ON-NP1 DDTLLIEDDDTAQVKELGVHLNPNGPITADQDGLDLEVDPQTLTVTTSGATGGVLGVLLK 240 FAdV-11-ON-NP2 DDTLLIEDDDTAQVKELGVHLNPNGPITADQDGLDLEVDPQTLTVTTSGATGGVLGVLLK 240 FAdV-11-ON-P2 DDTLLIEDDDTAQVKELGVHLNPNGPITADQDGLDLEVDPQTLTVTTSGATGGVLGVLLK 240 FAdV-11-ON-P1 DDTLLIEDDDTAQVKELGVHLNPNGPITADQDGLDLEVDPQTLTVTTSGATGGVLGVLLK 240 ************************************************************ FAdV-11-ON-NP1 PSGGLQTSIQGIGVAVADTLTISSNTGTVEVKTDPNGSIGSSSNGIAVVTDPAGPLTTSS 300 FAdV-11-ON-NP2 PSGGLQTSIQGIGVAVADTLTISSNTGTVEVKTDPNGSIGSSSNGIAVVTDPAGPLTTSS 300 FAdV-11-ON-P2 PSGGLQTSIQGIGVAVADTLTISSNTGTVEVKTDPNGSIGSSSNGIAVVTDPAGPLTTSS 300 FAdV-11-ON-P1 PSGGLQTSIQGIGVAVADTLTISSNTGTVEVKTDPNGSIGSSSNGIAVVTDPAGPLTTSS 300 ************************************************************ FAdV-11-ON-NP1 NGLSLKLTPNGSIQSSSTGLSVQTDPAGPITSGANGLSLSYDTSDFTVSQGMLSIIRNPS 360 FAdV-11-ON-NP2 NGLSLKLTPNGSIQSSSTGLSVQTDPAGPITSGANGLSLSYDTSDFTVSQGMLSIIRNPS 360 FAdV-11-ON-P2 NGLSLKLTPNGSIQSSSTGLSVQTDPAGPITSGANGLSLSYDTSDFTVSQGMLSIIRNPS 360 FAdV-11-ON-P1 NGLSLKLTPNGSIQSSSTGLSVQTDPAGPITSGANGLSLSYDTSDFTVSQGMLSIIRNPS 360 ************************************************************ FAdV-11-ON-NP1 AYPDAYLESGTNLLNNYTAYAENSSNYKFNCAYFLQSWYSNGLVTSSLYLKINRDNLTSL 420 FAdV-11-ON-NP2 AYPDAYLESGTNLLNNYTAYAENSSNYKFNCAYFLQSWYSNGLVTSSLYLKINRDNLTSL 420 FAdV-11-ON-P2 AYPDAYLESGTNLLNNYTAYAENSSNYKFNCAYFLQSWYSNGLVTSSLYLKINRDNLTSL 420 FAdV-11-ON-P1 AYPDAYLESGTNLLNNYTAYAENSSNYKFNCAYFLQSWYSNGLVTSSLYLKINRDNLTSL 420 ************************************************************ FAdV-11-ON-NP1 PSGQLSENAKYFTFWVPTYESMNLSNVATPTITPSSVPWGAFLPAQNCTSNPAFKYYLTQ 480 FAdV-11-ON-NP2 PSGQLSENAKYFTFWVPTYESMNLSNVATPTITPSSVPWGAFLPAQNCTSNPAFKYYLTQ 480 FAdV-11-ON-P2 PSGQLSENAKYFTFWVPTYESMNLSNVATPTITPSSVPWGAFLPAQNCTSNPAFKYYLTQ 480 FAdV-11-ON-P1 PSGQLSENAKYFTFWVPTYESMNLSNVATPTITPSSVPWGAFLPAQNCTSNPAFKYYLTQ 480 ************************************************************ FAdV-11-ON-NP1 PPSIYFEPESGSVQTFQPVLTGDWDTNTYNPGTVQVCILPQTVVGGQSTFVNMTCYNFRC 540 FAdV-11-ON-NP2 PPSIYFEPESGSVQTFQPVLTGDWDTNTYNPGTVQVCILPQTVVGGQSTFVNMTCYNFRC 540 FAdV-11-ON-P2 PPSIYFEPESGSVQTFQPVLTGDWDTNTYNPGTVQVCILPQTVVGGQSTFVNMTCYNFRC 540 FAdV-11-ON-P1 PPSIYFEPESGSVQTFQPVLTGDWDTNTYNPGTVQVCILPQTVVGGQSTFVNMTCYNFRC 540 ************************************************************ FAdV-11-ON-NP1 QNPGIFKVAASSGTFTIGPIFYSCPTNKLTQP 572 FAdV-11-ON-NP2 QNPGIFKVAASSGTFTIGPIFYSCPTNKLTQP 572 FAdV-11-ON-P2 QNPGIFKVAASSGTFTIGPIFYSCPTNKLTQP 572 FAdV-11-ON-P1 QNPGIFKVAASSGTFTIGPIFYSCPTNKLTQP 572 ********************************

Page 152: Sequence analysis, pathogenicity and cytokine gene

141

GENERAL DISCUSSION

Fowl adenoviruses are common infectious agents of poultry and often of low

virulence. The widespread occurrence of antibodies and the isolation of FAdVs from

clinically healthy chickens argue against a role of FAdVs in primary poultry disease

(Yates et al., 1977; Okuda et al., 2006; Ojkic et al., 2008a). Common to all virulent

FAdVs is infection of hepatocytes which results in a clinical manifestation known as IBH

occurring mainly in broilers. FAdVs are shed in feces and usually spread horizontally

through contaminated food, water, and litter (McFerran and Smyth, 2000). Evidence of

vertical transmission has also been reported (Grgic et al., 2006). From the available

evidence, it seems that most strains of FAdVs follow the same pattern of infection.

Following initial multiplication, viraemia occurs and results in virus spread to virtually

all organs. The main sites of virus replication are the respiratory and alimentary systems

(Saiffudin and Wilks, 1991). In our study, two Canadian FAdV isolates were evaluated:

FAdV-8 was isolated from an IBH outbreak that occurred in an Ontario broiler farm, and

FAdV-4 was isolated from a broiler breeder flock showing no clinical signs of IBH or

IBH/HHS. Pathogenicity was determined for both isolates on the basis of clinical signs

and pathological and histological examination. Additionally, viral genome copy numbers

in the bursa of Fabricius, cecal tonsil, and liver were evaluated following infection, and

virus titer in feces and antibody response were also determined.

According to Ojkic and Nagy (2003), replication and distribution of FAdVs in

chicken tissues is markedly influenced by the route of virus inoculation. Therefore, we

compared im and the more natural oral routes of inoculation. Despite a lack of clinical

Page 153: Sequence analysis, pathogenicity and cytokine gene

142

signs and pathological changes, both FAdV-8 and FAdV-4 were found in tissues and

cloacal swabs of all birds. For both FAdV-8 and FAdV-4 and in both the orally and

intramuscularly inoculated groups, the highest numbers of viral genome copies were

found in the cecal tonsil. Virus was isolated from cloacal swabs up to 28 days post

infection for both viruses, which is believed to be the longest recorded shedding period

for fowl adenoviruses. Both FAdV-8 and FAdV-4 were sufficiently immunogenic to

induce a strong antibody response.

The pathogenicity data indicated that under the conditions of this study, neither

isolate was pathogenic. Thus, the effects of FAdV infection on the transcription of a

number of avian cytokines were investigated. Cytokines are important mediators of

inflammation and regulators of the immune response. The first defense against viral

infection is an inflammatory response, with simultaneous release of inflammatory

cytokines. However, the role of interleukins in the pathogenicity of and immune response

to FAdVs is unknown. This study explored the ability of fowl adenoviruses to subvert the

host cells’ secretion of cytokines in response to infection as an important viral

mechanism for immune evasion during infection. In chicken experiments, IFN-γ, IL-10,

IL-18, and IL-8 gene expression were evaluated following FAdV-8 and FAdV-4

infection. Cytokine gene expression was examined in the liver, spleen, and cecal tonsils.

Results showed that IFN-γ, IL-18, IL-10 and IL-8 were activated early in FAdV

infection. The studied cytokines represent chicken Th1 and regulatory cytokines and all

have previously been shown to be involved in immunity to viral infections. IFN-γ,

considered an essential cytokine in host defense against a variety of pathogens, inhibits

Page 154: Sequence analysis, pathogenicity and cytokine gene

143

MDV replication (Xing and Schat, 2000), reduces replication of coccidian parasites

(Lillehoj and Choi, 1998), and inhibits src oncogene-induced tumor growth (Plachy et al.,

1999). In this study, the IFN-γ gene was induced to a significant degree following both

FAdV-8 and FAdV-4 infection. Another Th1 cytokine, IL-18, was originally identified as

a factor promoting IFN-γ production by T and NK cells when costimulated by IL-12

(Okamura et al., 1995). This proinflammatory cytokine (Gobel et al., 2003) is involved in

activation of mammalian neutrophils and NK cells (Okamura et al., 1995). Following

FAdV-8 infection, expression of the IL-18 gene was significantly higher in spleen and

liver in infected chickens than in mock-infected chickens. Enhanced expression of IL-18

in liver and cecal tonsil was detected in FAdV-4 infection. However, there was no

significant difference between infected and mock-infected chickens in terms of IL-18

gene expression. The increase in expression of IL-18 is probably an indicator of induction

of inflammation and an immune response against the virus. Furthermore, IL-10 was up-

regulated following FAdV-8 and FAdV-4 infections, indicating an immune evasive

mechanism used by the virus to skew the immune response against FAdVs, supporting

persistence of the virus. Simultaneous expression of IFN-γ and the IL-10 gene also has

been reported by Johensen et al, (2002) and Parvizi et al (2009). It is possible that

enhanced expression of IL-10 is induced by the chicken immune system to modulate the

immune response and inflammation caused by the virus. The increased expression of

mRNA of the IL-8 gene in the liver recorded after FAdV-4 infection could be a reactive

response of the host to reduce viral load. In contrast, IL-8 mRNA was significantly down-

regulated in the liver and spleen following FAdV-8 infection. Inhibition of IL-8 mRNA

expression by mammalian adenovirus E3 genes has been reported by Lesokhin et al,

Page 155: Sequence analysis, pathogenicity and cytokine gene

144

(2002). However, FAdV-8 lacks the mastadenoviral E3 region (Grgic et al., 2011). Thus,

some parts of the FAdV-8 genome probably share a function similar to the

mastadenoviral E3 region, despite the lack of obvious sequence homology. This study

provides useful information about important immunological mediators involved in

immunity against FAdVs and virus evasion that could lead to an effective vaccine against

IBH or IBH/HHS.

In the family Adenoviridae, viruses belonging to the genus Mastadenovirus have

been extensively investigated, and their genomic organizations have been well

characterized. Although aviadenoviruses have been less studied, recently Grgic et al,

(2011) and Griffin and Nagy (2011) reported the complete sequences of two FAdV

serotypes (FAdV-8 and FAdV-4). In addition, the sequence analysis of the left and right

end regions of FAdVs representing each species group has been also determined

(Corredor et al., 2006; Corredor et al., 2008).

Here, for the first time, a comparative analysis of the full genome sequence of an

FAdV-8 isolate of the species Fowl adenovirus E has been described. The FAdV-8

genome is 251 nt longer than that of the FAdV-1 genome and 1008 nt shorter than that of

the FAdV-9 genome (Chiocca et al., 1996; Ojkic and Nagy, 2000). Several important

features of FAdV-8 contribute to this size difference. The nucleotide and amino acid

sequence analyses revealed a lack of mastadenoviral early regions E1, E3, and E4 and the

sequences of proteins V and IX. Lack of early regions has been also reported for FAdV-1

and FAdV-9 genomes (Chiocca et al., 1996; Ojkic and Nagy, 2000). Tandem repeats,

Page 156: Sequence analysis, pathogenicity and cytokine gene

145

short repeated regions located at the right end of the FAdV-9 genome with unknown

function and non-essential for genome replication, have been identified within ORF19 in

the FAdV-8 genome. However, tandem repeats of FAdV-8 are much shorter than those of

FAdV-9. Virus-associated RNA found in mammalian adenoviruses is 150 to 170 nt long

and transcribed rightward (Ma and Mathews, 1993). Human AdVs contain one or two

VA RNA genes located between the pTP and 52k ORFs. In contrast, the VA RNA of

FAdV-1 is shorter, comprising only 100 nt and transcribed leftward (Larsson et al.,

1986). Although the VA RNA in the EDS virus is similar in size to that in FAdV-1,

comprising 91 nt, it is transcribed rightward as in mammalian adenoviruses (Hess et al.,

1997). Sequence analysis showed that the VA RNA gene was not identified in the FAdV-

8 genome. The lack of a VA RNA gene was also reported for FAdV-9, FAdV-4, and

OAdV287 (Ojkic and Nagy, 2000, Griffin and Nagy, 2011, Venktesh et al., 1998). In

contrast to FAdV-1, which has two fiber genes (Chiocca et al., 1996), only one fiber gene

was identified in the corresponding region of the FAdV-8 and FAdV-9 genomes (Grgic et

al., 2011; Ojkic and Nagy, 2000). The fiber gene has been implicated in pathogenicity of

HAVs (Mei and Wadell, 1995) and variation in virulence of FAdVs (Pallister et al.,

1996). Although extensive research with a focus on structure, sequence analysis, and

pathogenicity (El Bakkouri et al., 2008; Grgic et al., 2011; Griffin and Nagy, 2011) was

performed, determinants of virulence for FAdVs are still unknown. I hypothesized that

sequence variations of the fiber gene contribute to the virulence of FAdVs. Therefore, the

nucleotide and amino acid structure of the fiber gene of eight fowl adenoviruses

(pathogenic and non-pathogenic) representing species groups D (FAdV-11) and E

(FAdV-8) were analyzed. The FAdV-11 and FAdV-8 fiber genes encoded 572 aa and 522

Page 157: Sequence analysis, pathogenicity and cytokine gene

146

to 525 aa, respectively, and like all other adenovirus fibers, they are organized into three

regions, tail, shaft, and head. The tail region of both serotypes 8 and 11 contained a

poly(G) stretch identified previously in the tail of the long fiber of FAdV-1 (Hess et al.,

1995), while the same stretch is absent in the FAdV-10 fiber gene (Sheppard et al.,

1998). It was shown for mammalian AdVs (Green et al., 1983; van Raaij et al., 1999)

that the shaft region contains repetitive elements. Correspondingly, the shaft region of

FAdV-11 contains 20 pseudorepeates, while FAdV-8 contains 17 pseudorepeates, each

repetition comprising 18 and 16 aa on average, respectively. Moreover, mastadenovirus

conserved amino acid sequence (KLGXGLXFD/N) is absent from both FAdV-8 and -11

fibers and has not been reported for FAdV-1 and FAdV-10 (Hess et al., 1995; Sheppard

et al., 1998). Well-conserved TLWT sequence in the first β-strand of the head (Xia et al.,

1994), also was not evident in either FAdV-8 or FAdV-11 fibers. From these results it

may be concluded, that virulence might not be associated with the sequence of the fiber

gene alone. Experimental infection would be necessary to estimate the influence of the

adenoviral fiber gene on pathogenesis. In conclusion, this is the first study aimed at

determining cytokine gene expression patterns associated with FAdV infection in vitro

and in vivo.

Additionally, on the basis of a complete genome sequence analysis of FAdV-8, it

was concluded that FAdV-1 of the species Fowl adenovirus A as the current type species

is not the best representative of the genus Aviadenovirus, and that FAdV-8 in the species

Fowl adenovirus E would be more suitable for that designation.

Page 158: Sequence analysis, pathogenicity and cytokine gene

147

Data generated from this work provide the basis for future studies related to

cytokine gene expression patterns associated with FAdV infection and to virulence

factors in pathogenic FAdVs.

Page 159: Sequence analysis, pathogenicity and cytokine gene

148

References Abdul-Aziz, T.A. and Al-Attar, M.A., 1991. New syndrome in Iraqi chicks. Vet. Rec. 129, 272-275. Abdul-Aziz, T.A. and Hasan, S.Y., 1995. Hydropericardium syndrome in broiler chickens: its contagious nature and pathology. Res. Vet. Sci. 59, 219-221. Abdul-Careem, M.F., Hunter, B.D., Parvizi, P., Haghighi, H, R., Thanthrige-Don, N., Sharif, S., 2007. Cytokine gene expression patterns associated with immunization against Marek’s disease in chickens. Vaccine 25, 424-432. Abdul-Careem, M.F., Hunter, B.D., Sarson, A.J., Mayameei, A., Zhou, H., Sharif, S., 2006. Marek’s disease virus-induced transient paralysis is associated with cytokine gene expression in the nervous system. Viral Immunol. 19, 167-176. Abe, T., Nakamura, K., Tojo, T., Yuasa, N., 2001. Gizzard erosion in broiler chicks by group I avian adenovirus. Avian Dis. 45, 234-239. Abe, T., Nakamura, K., Tojo, H., Mase, H., Shibahara, T., Yamaguchi, S., Yuasa, N., 1998. Histology, immunohistochemistry, and ultrastructure of hydropericardium syndrome in adult broiler breeders and broiler chicks. Avian Dis. 42, 606-612. Adair, B.M. and Fitzgerald, S.D., 2008. Adenovirus infections. In: Diseases of poultry. Eds.:Saif, Y.M., Fadly, A.M., Glisson, J.R., McDougald, L.R., Nolan, L.K., Swayne. 12th

ed. Blackwell Publishing Professional, Ames, Iowa, USA. Pp. 251-291. Adair, B.M., McFerran, J.B., Calvert, V.M., 1980. Development of a microtitre fluorescent antibody test for serological detection of adenovirus infection in birds. Avian Pathol. 9, 291-300. Aini, I., Shih, I.M., Castro, A.E., Zee, Y.C., 1993. Comparison of herpesivirus isolates from falcons, pigeons and psittacines by restriction endonuclease analysis. J. Wildl. Dis. 29, 196-202. Akira, S., Uematsu, S., Takeuchi, O., 2006. Pathogen recognition and innate immunity. Cell 124, 783-801. Akira, S. and Takeda, K., 2004. Toll-like receptor signaling. Nat. Rev. Immunol. 4, 499-511. Akusjarvi, G., Alestrom, P., Pettersson, M., Lager, M., Jornvall, H., Pettersson, U., 1984. The gene for the adenovirus 2 hexon polypeptide. J. Biol. Chem. 259, 13976-13979. Alexander, H.S., Huber, P., Cao, J.X., Krell, P.J., Nagy, É. 1998. Growth characteristics of fowl adenovirus type 8 in a chicken hepatoma cell line. J. Virol. Methods 74:9-14.

Page 160: Sequence analysis, pathogenicity and cytokine gene

149

Albinsson, B. and Kidd, A., 1999. Adenovirus type 41 lacks an RGD α v-integrin binding motif on the penton base and undergoes delayed uptake in A549 cells. Virus Res. 64, 125-136. Alcorn, M.J., Booth, J.L., Coggeshall, K.M., Metcalf, J.P., 2001. Adenovirus type 7 induces interleukin-8 production via activation of extracellular regulated kinase 1/2. J.Virol. 75, 6450-6459. Anjum, A.D., 1990. Experimental transmission of hydropericardium syndrome and protection against it in commercial broiler chickens. Avian Pathol. 19, 655-660. Arany, Z., Newsome, D., Oldread, E., Livingston, D.M., Eckner, R., 1995. A family of transcriptional adaptor proteins targeted by the E1A oncoprotein. Nature 374, 81-84. Armentero, M.T., Horwitz, M., Mermod, N., 1994. Targeting of DNA polymerase to the adenovirus origin of DNA replication by interaction with nuclear factor 1. Proc. Natl. Acad. Sci. USA 91, 11537-11541. Arstila, P.T., Vainio, O., Lassila, O., 1994. Central role of CD4+T cells in avian immune response. Poult. Sci. 73, 1019-1016. Asea, A., Rehli, M., Kabingu, E., Boch, A.J., Bare, O., Auron, P.E., Stevenson, A.M., Calderwood, S., 2002. Novel signal transduction pathway utilized by extracellular HSP70: role of toll-like receptor (TLR) 2 and TLR4. J. Biol. Chem. 277, 15028-15034. Baggiolini, M., 1998. Chemokines and leukocyte traffic. Nature 392, 565-568. Bandara, L.R., La Thangue, N.B., 1991. Adenovirus E1a prevents the retinoblastoma gene product from complexing with a cellular transcription factor. Nature 351, 494-497. Barlan, A.U., Griffin, T.M., Mcguire, K.A., Weithoff, C.M., 2011a. Adenovirus membrane penetration activates the NLRP3 inflammasome. J. Virol. 85, 146-155. Barlan, A.U., Danthi, P., Wiethoff, C.M., 2011b. Lysosomal localization and mechanism of membrane penetration influence nonenveloped virus activation of the NLRP3 inflammasome. Virology 412, 306-314.

Baroudy, B.M., Moss, B., 1982. Sequence homologies of diverse length tandem repetitions near ends of vaccinia virus genome suggest unequal crossing over. Nucleic Acid Res. 10, 5673-5679. Beard, C.W., Ball, A.O., Wooley, E.H., Spindler, K.R., 1990. Transcription mapping of mouse adenovirus type 1 early region 3. Virology 175, 81-90. Beg, A.A., 2002. Endogenous ligands of Toll-like receptors: implications for regulating inflammatory and immune responses. Trends Immunol. 23, 509-512.

Page 161: Sequence analysis, pathogenicity and cytokine gene

150

Benko, M., Harrach, B., Both, G.W., Russell, W.C., Adair, B.M., Adam, E., deJong, J.C., Hess, M., Johnson, M., Kajon, A., Kidd, A.H., Lehmkuhl, H.D., Li, Q., Mautner, V., Pring-Akerblom, P., Wadell, G., 2005. Family Adenoviridae. In: Fauquet, C.M., Mayo, M.A., Maniloff, J. et al (Eds.), Virus Taxonomy: VIIIth Report of the International Committee on Taxonomy of Viruses. Elsevier Academic Press, New York, pp. 213-228. Bennett, E.M., Bennink, J.R., Yewdell, J.W., Brodsky, F.M., 1999. Cutting edge: adenovirus E19 has two mechanisms for affecting class I MHC expression. J. Immunol. 162, 5049-5052. Bewley, M., Springer, K., Zhang, Y-B., Freimuth, P., Flanagan, J., 1999. Structural analysis of the mechanism of adenovirus binding to its human cellular receptor, CAR. Science 286, 1578-1583. Bickford, A.A., 1974. The clinical and pathological features of inclusion body hepatitis of chickens. Proceedings of 23rd Western Poultry Disease Conference, University of California, Davis, Califonia. pp. 83-85. Booth, J.L. and Metcalf, J.P., 1999. Type specific induction of interleukin-8 by adenovirus. Am. J. Respir. Cell Mol. Biol. 21, 521-527. Bowie, A. and O’Neill, L.A., 2000. The interleukin-1 receptor/Toll-like receptor superfamily: signal generators for pro-inflammatory interleukins and microbial products. J.Leukoc. Biol. 67, 508-514. Braithwaite, A.W., Russell, I.A., 2001. Induction of cell death by adenoviruses. Apoptosis 6, 359-370. Bruder, J.T., Kovesdi, I., 1997. Adenovirus infection stimulates the Raf/MAPK signaling pathway and induces interleukin-8 expression. J. Virol. 71, 398-04. Boyd, A., Philbin, V.J., Smith, A.L., 2007. Conserved and distinct aspects of the avian Toll-like receptor (TLR) system: implications for transmission and control of bird-born zoonoses. Biochem. Soc. Trans. 35, 1504-1507. Burmeister, W.P., Guilligay, D., Cusack, S., Wadell, G., Arnberg, N., 2004. Crystal structure of species D adenovirus fiber knobs and their sialic acid binding sites. J. Virol. 78, 7727-7736. Caillet-Boudin, M.L., 1989. Complementary peptides sequences in partner proteins of the adenovirus capsid. J. Mol. Biol. 208, 195-198. Cao, J.X., Krell, P.J., Nagy, É. 1998. Sequence and transcriptional analysis of terminal regions of the fowl adenovirus type 8 genome. J. Gen. Virol. 79, 2507-2516.

Page 162: Sequence analysis, pathogenicity and cytokine gene

151

Capua, I., Gough, R.E., Scaramozzino, P., Lelli, R., Gatti, A., 1994. Isolation of an adenovirus from an ostrich (Struthio camelus) causing pancreatitis in experimentally infected guinea fowl (Numida meleagris). Avian Dis. 38, 642-646. Carbonneil, C., Donkova-Petrini, V., Aouba, A., Weiss, L., 2004. Defective dendritic cell function in HIV-infected patients receiving effective highly active antiretroviral therapy: neutralization of IL-10 production and depletion of CD4+CD25+ T cells restore high levels of HIV-specific CD4+ T cell responses induced by dendritic cells generated in the presence of IFN-alpha. J. Immunol. 172, 7832-7840. Cepko, C.L. and Sharp, P.A., 1982. Assembly of adenovirus major capsid protein is mediated by a nonvirion protein. Cell 31, 407-415. Cerullo, V., Seiler, M.P., Mane, V., Brunetti-Pierri, N., Clarke, C., Bertini, T.K., Rodgers, J.R., Lee, B., 2007. Toll-like receptor 9 triggers an innate immune response to helper-dependent adenoviral vectors. Mol. Ther. 15, 378-385. Chellappan, S., Kraus, V.B., Kroger, B., Munger, K., Howley, P.M., Phelps, W.C., Nevins, J.R., 1992. Adenovirus E1A , simian virus 40 tumor antigen, and human papillomavirus E7 protein share the capacity to disrupt the interaction between transcription factor E2F and the retinoblastoma gene product. Proc. Natl. Acad. Sci. U. S. A. 89, 4549-4553. Chiocca, S., Kurzbauer, R., Schaffner, G, Baker, A., Mautner, V., Cotten, M., 1996. The complete DNA sequence and genomic organization of the avian adenovirus CELO. J. Virol. 70, 2939-2949. Cho, Y.S., Kang, J.W., Cho, M., Cho, C.W., Lee, S., Choe, Y.K., 2001. Down modulation of IL-18 expression by human papillomavirus type 16 E6 oncogene via binding to IL-18. FEBS Lett. 501, 139-45. Chow, L.T., Lewis, J.B., Broker, T.R., 1980. RNA transcription and splicing at early and intermediate times after adenovirus-2 infection. Cold Spring Harb. Symp. Quant. Biol. 44 Pt 1, 401-414. Chroboczek, J., Bieber, F., Jacrot, B., 1992. The sequence of the genome of adenovirus type 5 and its comparison with genome of adenovirus type 2. Virology 186, 280-285. Chroboczek, J., Jacrot, B., 1987. The sequence of adenovirus fiber: Similarities and differences between serotypes 2 and 5. Virology 161, 549-554. Clemmer, D.I., 1972. Age-associated changes in fecal excretion patterns of strain 93 chick embryo lethal orphan virus in chicks. Infect. Immun. 5, 60-64. Clemmer, D.I. and Ichinose, H., 1968. The cellular site of virus replication in the intestine of chickens infected with an avian adenovirus. Arch. Virol. 25, 277-287.

Page 163: Sequence analysis, pathogenicity and cytokine gene

152

Cohen, C.J., Xiang, Z.Q., Gao, G-P., Ertl, H.C., Wilson, J.M., Bergelson, J.M., 2002. Chimpanzee adenovirus CV-68 adapted as a gene delivery vector interacts with coxsackievirus and adenovirus receptor. J. Gen. Virol. 83, 151-155. Cohen, C.J., Shieh, J.T., Pickles, R.J., 2001. The coxsackievirus and adenovirus receptor is a transmembrane component of the tight junction. Proc. Natl. Acad. Sci. USA 98, 15191-15196. Colby, W.W. and Shenk, T., 1981. Adenovirus type 5 virions can be assembled in vivo in the absence of detectable polypeptide IX. J. Virol. 39, 977-980. Cole, J.L., Housley, G.A., Dykman, T.R., McDermott, R.P., Atkinson, J.P., 1985. Identification of an additional class of C3-binding membrane proteins of human peripheral blood leukocytes and cell lines. Proc. Natl. Acad. Sci. USA 82, 859-863. Compton, T., Kurt-Jones, E.A., Boehme, K.W., Belko, J., Latz, E., Golenbock, D.T., Finberg, R.W., 2003. Human cytomegalovirus activates inflammatory cytokine responses via CD14 and toll-like receptor 2. J. Virol. 77, 4588-4596. Cook, J.K.A., 1983. Fowl adenoviruses: studies on aspects of the pathogenicity of six strains for 1-day-old chicks. Avian Pathol. 12, 35-43. Cook, J.K.A. and Darbyshire., 1980. Epidemiological studies with egg drop syndrome 1976 (EDS-76) virus. Avian Pathol. 9, 437-443. Cook, J.K.A., 1974. Spread of an avian adenovirus (CELO virus) to uninoculated fowls. Res. Vet. Sci. 16, 156-161. Cooper, M.D., Peterson, R.D.A. South, M.A., Good, R.A., 1966. The functions of the thymus system and the bursa system in the chicken. J. Exp. Med. 123, 75-102. Corredor, J.C., Nagy, É., 2010a. A region at the left end of the fowl adenovirus 9 genome that is non-essential in vitro has consequences in vivo. J. Gen. Virol. 91, 51-58. Corredor, J.C., Nagy, É., 2010b. The non-essential left end region of the fowl adenovirus 9 genome is suitable of foreign gene insertion/replacement. Virus Res. 149, 167-174. Corredor, J.C., Garceac, A., Krell, P.J., Nagy, E., 2008. Sequence comparison of the right end of fowl adenovirus genomes. Virus Genes 36, 331-334. Corredor, J.C., Krell, P.J., Nagy, E., 2006. Sequence analysis of the left end of fowl adenovirus genomes. Virus Genes 33, 95-106. Cotten, M. and Weber, J.M., 1995. The adenovirus protease is required for virus entry into host cells. Virology 213, 494-502.

Page 164: Sequence analysis, pathogenicity and cytokine gene

153

Cowen, B., 1992. Inclusion body hepatitis-anaemia and hydropericardium syndrome: aetiology and control. Worlds Poult. Sci. J. 48, 247-253. Cowen, B., 1988. Chicken embryo propagation of type I avian adenovirus. Avian Dis. 32, 347-352. Cowen, B., Calnek, B.W., Menendez, N.A., Ball, R.F., 1978. Avian adenoviruses: effect on egg production, shell quality and feed consumption. Avian Dis. 22, 459-470. Cowen, B., Mitchell, G.B., Calnek, B.W., 1977. An adenovirus survey of poultry flocks during the growing and laying periods. Avian Dis. 22, 115-121. Dao, T., Mehal, W.Z., Crispe, I.N., 1998. IL-18 augments perforin-dependent cytotoxicity of liver NK-T cells. J. Immunol. 161, 2217-2222. Da Silva Martins, N.R., Mockett, A.P.A., Barrett, A.D.T., Cook, J.K.A., 1991. IgM responses in chicken serum to live and inactivated infectious bronchitis virus vaccines. Avian Dis. 35, 470-475. Davison, A.J., Benko, M., Harrach, B., 2003. Genetic content and evolution of adenoviruses. J. Gen. Virol. 84, 2895-2908. Davison, E., Diaz, R.M., Hart, I.R., Santis, G., Marshall, J.F., 1997. Integrin α5β1-mediated adenovirus infection is enhanced by the integrin-activating antibody TS2/16. J. Virol. 71, 6204–6207. Davison, A.J., Telford, E.A.R., Watson, M.S., McBride, K., Mautner, V., 1993. The DNA sequence of adenovirus type 40. J. Mol. Biol. 234, 1308-1316. De Herdt, P., Ducatelle, R., Lepoudre, C., Chalier, G., Nauwynck, H., 1995. An epidemic of fatal hepatic necrosis of viral origin in racing pigeons (Columba livia). Avian Pathol. 24, 475-483. de Jong, R.N., van der Vliet, P.C., Brenkman, A.B., 2003. Adenovirus DNA replication: protein priming, jumping back and the role of the DNA binding protein DBP. Curr. Top. Microbiol. Immunol. 272, 187-212. de Jong, R.N. and van der Vliet, P.C., 1999. Mechanism of DNA replication in eukaryotic cells: cellular host factors stimulating adenovirus DNA replication. Gene 236, 1-12. Digby, M.R. and Lowenthal, J.W., 1995. Cloning and expression of the chicken interferon-gamma gene. J. J. Interferon Cytokine Res. 15, 939-945. Ding, J., McGrath, W.J., Sweet, R.M., Mangel, W.F., 1996. Crystal structure of the human adenovirus proteinase with its 11 amino acid cofactor. EMBO J. 15, 1778-1783.

Page 165: Sequence analysis, pathogenicity and cytokine gene

154

Dolph, P.J., Racaniello, V., Villamarin, A., Palladino, F., Schneider, R.J., 1988. The adenovirus tripartite leader may eliminate the requirement for cap-binding protein complex during translation initiation. J. Virol. 62, 2059-2066. Edvardsson, B., Everitt, E., Jornvall, H., Prage, L., Philipson, L., 1976. Intermediates in adenovirus assembly. J. Virol. 19, 533-547. El Bakkouri, M., Seiradake, E., Cusak, S., Ruigrok, R.W., Schoehn, G., 2008. Structure of the C-terminal head domain of the fowl adenovirus type 1 short fiber. J. Virol. 378, 169-176. Elsing, A.T., Burgert, H.G., 1998. The adenovirus E3/10.4K-14.5K proteins down-modulate the apoptosis receptor Fas/Apo-1 by inducing its internalization. Proc. Natl. Acad. Sci. USA 95, 10072-10077. Erny, K.M., Barr, D.A., Fahey, K.J., 1991. Molecular characterization of highly virulent fowl adenovirus associated with outbreaks on inclusion body hepatitis, Avian Pathol. 20, 597-606. Endharti, A.T., Rifa, M.I., Shi, Z., Fukuoka, Y., Nakahara, Y., 2005. Cutting edge: CD8 CD122 regulatory T cells produce IL-10 to suppress IFN-γ production and proliferation of CD8 T cells. J. Immunol. 175, 7093-7097. Evans, J.D. and Hearing, P., 2002. Adenovirus replication. In: Curiel, D.T., Douglas, J.T. (Eds), Academic Press, San Diego, CA, pp. 39-70. Fadly, A.M., Winterfield, R.W., Olander, H.J., 1975. Role of the bursa of Fabricius in the pathogenicity of inclusion body hepatitis and infectious bursal disease viruses. Avian Dis. 20, 467-472. Faha, B., Harlow, E., Lees, E., 1993. The adenovirus E1A-associated kinase consists of cyclin E-p33 cdk2 and cyclin A-p33 cdk2. J. Virol. 67, 2456-2465. Fechner, H., Haack, A., Wang, H., Eizema, K., Pauschinger, M., Schoemaker, R.G., van Veghel, R., Houtsmuller, A.B., Schultheiss, H-P., Lamers, J.M.J., Poller, W., 1999. Expression of coxsackie adenovirus receptor and alphav-integrin does not correlate with adenovector targeting in vivo indicating anatomical vector barriers. Gene Ther. 6, 1520-1535. Fiorentino, D.F., Zlotnik, A., Vieira, P., Mosmann, T.R., Howard, M., Moore, K.W., O’Garra, A., 1991. IL-10 acts on the antigen-presenting cell to inhibit cytokine production by Th1 cells. J. Immunol. 146, 3444-3451. Fitzgerald, S.D. and Reed, W.M., 1989. A review of marble spleen disease of ring-necked pheasants. J. Wild Dis. 25, 455-461.

Page 166: Sequence analysis, pathogenicity and cytokine gene

155

Flint, S.J., Enquist, L.W., Racaniello, V.R., Skalka, A.M., 2004. Principles of virology molecular biology, pathogenesis, and control of animal viruses. 2nd ed, ASM Press American Society for Microbiology, Washington, DC. Pp. 804-806. Fukui, A., Inoue, N., Matsumoto, M., Nomura, M., Yamada, K., Matsuda, Y., Toyoshima, K., Seya, Y., 2001. Molecular cloning and functional characterization of chicken toll-like receptors. A single chicken toll covers multiple molecular patterns. J. Biol. Chem. 276, 47143-47149. Gelderblom, H. and Maichle-Lauppe, I., 1982. The fibers of fowl adenoviruses. Arch. Virol. 72, 289-298. Gerosa, F., Paganin, C., Peritt, D., Paiola, F., Scupoli, M.T., Aste-Amezaga, M., Frank, I., Trinchieri, G., 1996. Interleukine-12 primes human CD4 and CD8 T cell clones for high production of both interferon-gamma and interleukine-10. J. Exp. Med. 183, 2559-2569. Ghosh-Choudhury, G., Haj-Ahmad, Y., Graham, F.L., 1987. Protein IX, a minor component of the human adenovirus capsid, is essential for the packaging of full length genomes. EMBO J. 6, 1733-1739. Gobel, T.W., Schneider, K., Schaerer, B., Mejri, J., Puehler, F., Weigend, S., Staeheli, P., Kaspers, B., 2003. IL-18 stimulates the proliferation and IFN-γ release of CD4+ T cells in the chicken: conservation of a Th1-like system in nonmammalian species. J. Immunol. 171, 1809-1815. Gomis, S., Goodhope, R., Ojkic, D., Willson, P., 2006. Inclusion body hepatitis as a primary disease in broilers in Saskatchewan, Canada. Avian Dis. 50, 550-555. Gooding, L.R., Sofola, I.O., Tollefson, A.E., Duerksen-Hughes, P.J., Wold, W.S.M., 1990. The adenovirus E3-14.7K protein is a general inhibitor of tumor necrosis factor-mediated cytolysis. J. Immunol. 145, 3080-3086. Goodwin, M.A., Hill, D.I., Dekich, M.A., Putnam, M.R., 1993. Multisystem adenovirus infection in broiler chicks with hypoglycemia and spiking mortality. Avian Dis. 37, 625-627. Gough, R.E., Drury, S.E., Capua, I., Courtenay, A.E., Sharp, M.W., Dick, A.C.K., 1997. Isolation and identification of adenoviruses from ostriches (Struthio camelus). Vet. Rec. 140, 402-403. Gracie, J.A., Robertson, S.E., McInnes, I.B., 2003. Interleukine-18. J. Leuk. Bio. 73, 213-224.

Page 167: Sequence analysis, pathogenicity and cytokine gene

156

Green, N.M., Wrigley, N.G., Russell, W.C., Martin, S.R., McLachlan, A.D., 1983. Evidence for a repeating cross-beta sheet structure in the adenovirus fibre. EMBO J. 2, 1357-1365. Grgić, H., Sharif, S., Haghighi, H.R., Nagy, É., 2012. Key cytokines activated during infection of chickens with a serotype 8 fowl adenovirus. Virus Res. Grgic, H., Yang, D.H., Nagy, E., 2011. Pathogenicity and complete genome sequence of a fowl adenovirus serotype 8 isolate. Virus Res. 156, 91-97. Grgic, H., Phillippe, C., Ojkic, D., Nagy, E., 2006. Study of vertical transmission of fowl adenoviruses. Can. J. Vet. Res. 70, 230-233. Griffin, B.D. and Nagy, E., 2011. Coding potential and transcript analysis of fowl adenovirus 4: insight into upstream ORFs as common sequence features in adenoviral transcripts. J. Gen. Virol. 92, 1260-1272. Grimes, T. M., 2007. Inclusion body hepatitis of chickens-occurence and control. Proc. 56th West. Poult. Dis. Conf., March 27-29, Las Vegas, Nev., 42-46. Grimes, T.M., Culver, D.H., King, D.J., 1977a. Virus-neutralizing antibody titers against 8 avian adenovirus serotypes in breeder hens in Georgia by a microneutralization procedure. Avian Dis. 21, 220-229. Grimes, T.M., King, D.J., Kleven, S.H., Fletcher, O.J., 1977b. involvement of a type 8 avian adenovirus in the etiology of inclusion body hepatitis. Avian Dis. 21, 26-38. Hammarskjold, M.L. and Winberg, G., 1980. Encapsidation of adenovirus 16 DNA is directed by a small DNA sequence at the left end of the genome. Cell 20, 787-795. Hanlon, L., Argyle, D., Bain, D., Nicolson, L., Dunham, S., Golder, M.C., 2001. Feline leukemia virus DNA vaccine efficacy is enhanced by coadministration with interleukine-12 (IL-12) and IL-18 expression vectors. J. Virol. 75, 8424-33. Heil, F., Hemmi, H., Hochrein, H., Ampenberger, F., Kirschning, C, Akira, S., Lipford, G., Wagner, H., Bauer, S., 2004. Species-specific recognition of single-stranded RNA via toll-like receptor 7 and 8. Science 303, 1526-1529. Helmboldt, C.F. and Frazier, M.N., 1963. Avian hepatic inclusion bodies of unknown significance. Avian Dis. 7, 446-450. Hess, M., 2000. Detection and differentiation of avian adenoviruses: review. Avian Pathol. 29, 195-206. Hess, M., Raue, R., Hafez, H.M., 1999. PCR for specific detection of haemorrhagic enteritis virus of turkeys an avian adenovirus. J. Virol. Methods. 81, 199-203.

Page 168: Sequence analysis, pathogenicity and cytokine gene

157

Hess, M., Raue, R., Prusas, C., 1999. Epidemiological studies on fowl adenoviruses isolated from cases of infectious hydropericardium. Avian Pathol. 28, 433-439. Hess, M., Prusas, C., Monreal, G., 1998. Growth analysis of adenoviruses isolated from pigeons in chicken cells and serological characterization of the isolates. Avian Pathol. 27, 196-199. Hess, M., Blocker, H., Brandt, P., 1997. The complete nucleotide sequence of the egg drop syndrome virus: and intermediate between mastadenoviruses and aviadenoviruses. Virology 238, 145-156. Hess, M., Cuzange, A., Ruigrok, R.W.H., Chroboczek, J., Jacrot, B., 1995. The avian adenovirus penton: two fibers and one base. J. Mol. Biol. 252, 379-385. Hoffmann, R., Wessling, E., Dorn, P., Dangschat, H., 1975. Lesions in chickens with spontaneous or experimental infectious hepato-myelopoietic disease (inclusion body hepatitis) in Germany. Avian Dis. 19, 224-236. Hong, S.S., Szolajska, E., Schoehn, G., Franqueville, L., Myhre, S., Lindholm, L., Ruigrok, R.W., Boulanger, P., Chroboczek, J., 2005. The 100K-Chaperone protein from adenovirus serotype 2 (subgroup C) assists in trimerization and nuclear localization of hexons from subgroups C and B adenoviruses. J. Mol. Biol. 352, 125-138. Hong, S.S., Boulanger, P., 1995. Protein ligands of the human adenovirus type 2 outer capsid identified by bipanning of a phage displayed peptide library on separate domains of wild-type and mutant penton capsomers. EMBO J. 14, 4714-4727. Hong, J.S. and Engler, J.A., 1996. Domains required for assembly of adenovirus type 2 fiber trimers. J. Virol. 70, 7071-7078. Hong, J.S., Mullis, K.G., Engler, J.A., 1988. Characterization of the early region 3 and fiber genes of Ad7. Virology 1967, 545-553. Hosokawa, K. and Sung, M.T., 1976. Isolation and characterization of an extremely basic protein from adenovirus type 5. J. Virol. 17, 924-934. Howell, J., MacDonald, D.W., Christian, R.G., 1970. Inclusion body hepatitis in chickens. Can. Vet. Jour. 11, 99-101. Huang, J.T. and Schneider, R.J., 1991. Adenovirus inhibition of cellular protein synthesis involves inactivation of cap-binding protein. Cell 65, 271-280. ICTV Virus Taxonomy: 2009 Release http://www.ictvonline.org/virusTaxonomy.asp?version=2009

Page 169: Sequence analysis, pathogenicity and cytokine gene

158

Inohara, N. and Nunez, G., 2003. NODs: intracellular proteins involved in inflammation and apoptosis. Nat. Rev. Immunol. 3, 371-382. Itakura, C. and Carlson, H.C., 1975. Electron microscopic findings of cells with inclusion bodies in experimental hemorrhagic enteritis of turkeys. Can. J. Com. Med. 39, 299-304. Itakura, C., Yasuba, M., Goto, M., 1974. Histopathological studies on inclusion body hepatits in broiler chickens. Jpn. J. Vet. Sci. 36, 329-340. Iwasaki, A. and Medzhitov, R., 2004. Toll-like receptor control of the adaptive immune responses. Nat. Immunol. 5, 987-995. Jaffery, M.A., 1988. A treatise on Angara disease (hydropericardium-pulmonary oedema-hepatonephritis syndrome) J. Pak. Med. Assoc. 34, 1-33. Jenkins, K.A., Lowenthal, J.W., Kimpton, W., Bean, A.G., 2009. The in vitro and in ovo responses to chickens to TLR9 subfamily ligands. Dev. Comp. Immunol. 33, 660-667. Jiang, P., Ojkic, D., Tuboly, T., Huber, P., Nagy, Ѐ., 1999. Application of the polymerase chain reaction to detect fowl adenoviruses. Can. J. Vet. Res. 63, 124-128. Johensen, C.K., Botner, A., Kamstrup, S., Lind, P., Nielsen, J., 2002. Cytokine mRNA profiles in bronchoalveolar cells of piglets experimentally infected in utero with porcine reproductive and respiratory syndrome virus: association of sustained expression of IFN-gamma and IL-10 after viral clearance. Viral Immunol. 15, 549-556. Jornvall, H., Akusjarvi, G., Alestrom, P., von Bahr-Lindstrom, H., Pettersson, U., Fowler, A.V., Philipson, L., 1981. The adenovirus hexon protein: the primary structure of the polypeptide and its correlation with the hexon gene. J. Biol. Chem. 256, 6181-6186. Kaiser, P., Underwood, G., Davison, F., 2003. Differential cytokine responses following Marek’s disease virus infection of chickens differing in resistance to Marek’s disease. J.Virol. 77, 762-768. Kaisho, T. and Akira, S., 2006. Toll-like receptor function and signalling. J. Allergy. Clin. Immunol. 117, 979-233. Kajan G.L., Stefancsik, R., Ursu, K., Palva, V., Benko, M., 2010. The first complete genome sequence of a non-chicken aviadenovirus proposed to be turkey adenovirus 1. Virus Res. 153, 226-233. Kamil, J.P., Tischer, B.K., Trapp, S., Nair, V.K., Osterrieder, N., Kung, H.J., 2005. VLIP, a viral lipase homologue, is a virulence factor of Marek’s disease virus. J. Virol. 79, 6984-6996.

Page 170: Sequence analysis, pathogenicity and cytokine gene

159

Karayan, L., Hong, S.S., Gay, B., Tournier, J., d’Angeac, A.D., Boulanger, P., 1997. Structural and functional determinants in adenovirus type 2 penton base recombinant protein. J. Virol. 71, 8678-8689. Keblusek, P., Dorsman, J.C., Teunisse, A.F., Teunissen, H., van der Eb, A.J., Zantema. A., 1999. The adenoviral E1A oncoproteins interfere with the growth-inhibiting effect of the Cdk-inhibitor p21 (CIP1/WAF1). J. Gen. Virol. 80, 381-390. Keestra, A.M., de Zoete, M.R., Bouwman, L.I., van Putten, J.P.M., 2010. Chicken TLR21 is an innate CpG DNA receptor distinct from mammalian TLR9. J. Immunol. 185, 460-467. Khawaja, D.A., Ahmad, S., Rauf, M.A., Zulfiqar, M.Z., Mahmood, S.M.I., Hasan, M., 1988. Isolation of an adenovirus from hydropericardium syndrome in broiler chicks. Pak. J. Vet. Res. 1, 2-17. Khan, S.A., Zaidi, F.H., Chaudhry, R.A., Ashraf, S.K., Mian, M.S., Quershi, I.R., 1995. Comparative study on pathology of haemic system of three broiler chicken strains suffering from experimental hydropericardium syndrome. Asian-australas. J. Anim. Sci. 8, 325-328. Kim, Y.M., Brinkmann, M.M., Paquet, M.E., Ploegh, H.L., 2008. UNC93B1 delivers nucleotide-sensing toll-like receptors to endolysosomes. Nature 452, 234-238. Klein, H., Maltzman, W., Levine. A.J., 1979. Structure-function relationships of the adenovirus DNA-binding protein. J. Biol. Chem. 254, 11051-11060. Krajcsi, P., Dimitrov, T., Hermiston, T.W., Tollefson, A.E., Ranheim, T.S., Vande Pol, S.B., Stephenson, A.H., Wold, W.S., 1996. The adenovirus E3-14.7K protein and the E3-10.4K/14.5K complex of proteins, which independently inhibit tumor necrosis factor (TNF)-induced apoptosis, also independently inhibit TNF-induced release of arachidonic acid. J. Virol. 70, 4904-4913. Kogut, M.H., Iqbal, M., He, H., Philbin, V., Kaiser, P., Smith, A. 2005. Expression and function of Toll-like receptors in chicken heterophils. Dev. Comp. Immunol. 29, 791-807. Kovács, G.M., LaPatra, S.E., D’Halluin, J.C., Benko, M., 2003. Phylogenetic analysis of the hexon and protease genes of a fish adenovirus isolated from white sturgeon (Acipenser transmountanus) supports the proposal for a new adenovirus genus. Virus Res. 98, 27-34. Kumar, R., Chandra, R., Shukla, S.K., Agrawal, D.K., Kumar, M., 1997. Hydropericardium syndrome (HPS) in India: a preliminary study on the causative agent and control of the disease by inactivated autogenous vaccine. Trop. Anim. Health. Prod. 29, 158-164.

Page 171: Sequence analysis, pathogenicity and cytokine gene

160

Lam, K.M., 2004. Mycoplasma gallisepticum-induced alterations in cytokine genes in chicken cells and embryos. Avian Dis. 48, 215-219. Larsson, S., Bellett, A., Akusjarvi, G., 1986. VA RNA from avian and human adenoviruses: dramatic difference in length, sequence and gene location. J. Virol. 58, 600-609. Lehmberg, E., Traina, J.A., Chakel, J.A., Chang, R.J., Parkman, M., McCaman, M.T., Murakami, P.K., Lahidji, V., Nelson, J.W., Hancock, W.S., Nestaas, E., Pungor, E.Jr., 1999. Reversed-phase high-performance liquid chromatographic assay for the adenovirus type 5 proteome. J. Chromatogr. B Biomed. Sci. Appl. 732, 411-423. Lehrmann, H. and Cotten, M., 1999. Characterization of CELO virus proteins that modulate the pRb/E2F pathway. J. Virol. 73, 6517-6525. Lemaitre, B., Nicolas, E., Michaut, L., Reichhart, J.M., Hoffmann, J.A., 1996. The dorsoventral regulatory gne cassette spatzle/Toll/cactus controls the potent antifungal response in Drosophila adults. Cell 86, 973-983. Lesokhin, A.M., Delgado-Lopez, F., Horwitz, M., 2002. Inhibition of chemokine expression by adenovirus early region three (E3) genes. J. Virol. 76, 8236-8243. Li, X., Tikoo, S.K., 2002. Genetic organization and sequence analysis of pVIII, fiber and early region 4 of bovine adenovirus type 7. Virus Genes 25, 59-65. Li, E., Brown, S.L., Stupack, D., Puente, X.S., Cheresh, D.A., Nemerow, G.R., 2001. Integrin αvβ1 is an adenovirus coreceptor. J. Virol. 75, 5405–5409. Li, E., Stupack, D., Brown, S.L., Klemke, R., Schlaepfer, D.D., Nemerow, G.R., 2000. Association of p130CAS with phosphatidylinositol-3-OH kinase mediates adenovirus cell entry. J. Biol. Chem. 275, 14729-14735. Li, Y., Kang, J., Friedman, J., Tarassishin, L., Ye, J., Kovalenko, A., Wallach, D., Horwitz, M.S., 1999. Identification of a cell protein (FIP-3) as a modulator of NF-kappaB activity and as a target of an adenovirus inhibitor of tumor necrosis factor alpha-induced apoptosis. Proc. Natl. Acad. Sci. USA 96, 1042-1047. Li, E., Stupack, D., Klemke, R., Cheresh, D.A., Nemerow, G.R., 1998a. Adenovirus endocytosis ia αv integrins requires phosphoinositide-3-OH kinase. J. Virol. 72, 2055-2061. Li, E., Stupack, D., Bokoch, G.M., Nemerow, G.R., 1998b. Adenovirus endocytosis requires actin cytoskeleton reorganization mediated by Rho family GTPases. J. Virol. 72, 8806-8812.

Page 172: Sequence analysis, pathogenicity and cytokine gene

161

Linne, T. and Philipson, L., 1980. Further characterization of the phosphate moiety of the adenovirus type 2 DNA-binding protein. Eur. J. Biochem. 103, 259-270. Litman, G.W., Rast, J.P., Shamblott, M.J., Haire, R.N., Hulst, M., Roess, W., Litman, R.T., Hinds-Frey, K.R., Zilch, A., Amemiva, C.T., 1993. Phylogenetic diversification of immunoglobulin genes and the antibody repertoire. Mol. Biol. Evol. 10, 60-72. Lillehoj, H.S., Choi, K,D., 1998. Recombinant chicken interferon-gamma-mediated inhibition of Eimeria tenella development in vitro and reduction of oocyst production and body weight loss following Eimeria acervulina challenge infection. Avian Dis. 42, 307-14. Look, D.C., Roswit, W.T., Frick, A.G., Gris-Alevy, Y., Dickhaus, D.M., Walter, M.J., Holtzman, M.J., 1998. Direct suppression of stat 1 function during adenoviral infection. Immunity 9, 871-880. Lund, J., Sato, A., Akira, S., Medzhitov, R., Iwasaki, A., 2003. Toll-like receptor 9-mediated recognition of Herpes simplex virus-2 by plasmacytoid dendritic cells. J. Exp. Med. 198, 513-520. Luschow, D., Prusas, C., Lierz, M., Gerlach, H., Soike, D., Hafez, H.M., 2007. Adenovirus of psittacine bires: investigation on isolation and development of a real-time polymerase chain reaction for specific detection. Avian Path. 36, 487-494. Lutz, P. And Kedinger, C., 1996. Properties of the adenovirus Iva2 gene product an effector of late-phase-dependent activation of the major late promoter. J. Virol. 70, 1396-1405. Lutz, P., Rosa-Calatrava, M., Kedinger, C., 1997. The product of the adenovirus intermediate gene IX is a transcriptional activator. J. Virol. 71, 5102-5109. Ma, Y. and Mathews, M.B., 1993. Comparative analysis of the structure and function of adenovirus virus-associated RNAs. J. Virol. 67, 6605-6617. Maiti, N.K. and Sarkar, P., 1997. Humoral immune response of chickens to different clinical isolates of avian adenovirus type-1. Comp. Immunol. Microbiol. Infect. Dis. 20, 59-62. Maier, O. and Wiethoff, C.M., 2010. N-terminal α-helix-independent membrane interactions facilitate adenovirus protein VI induction of membrane tubule formation. Virology 408, 31-38. Maizel, J.V., White, D.O., Scharff, M.D., 1968a. The polypeptide of adenovirus. II. Soluble proteins, cores, top components and structure of the virion. Virology 36, 126-136.

Page 173: Sequence analysis, pathogenicity and cytokine gene

162

Maizel, J.V., White, D.O., Scharff, M.D., 1968b. The polypeptides of adenovirus. I. Evidence for multiple protein components in the virion and a comparison of types 2, 7A, and 12. Virology 36, 115-125. Manarolla, G., Pisoni, G., Moroni, P., Gallazzi, D., Sironi, G., Rampin, T., 2009. Adenoviral gizzard erosions in Italian chicken flocks. Vet. Rec. 164, 754-756. Marek, A., Schulz, E., Hess, C., Hess, M., 2010. Comparison of the fibers of fowl adenovirus A serotype 1 isolates from chickens with gizzard erosions in Europe and apathogenic reference strains. J. Vet. Diagn. Invest. 6, 937-941. Martinez-Palomo, A., Le Buis, J., Bernhard, W., 1967. Electron microscopy of adenovirus 12 replication. J. Virol. 1, 817-829. Mase, M., Nakamura, K., Imada, T. 2010. Characterization of fowl adenovirus serotype 4 isolated from chickens with hydropericardium syndrome based on analysis of the short fiber protein gene. J. Vet. Diagn. Invest. 22, 218-223. Mathews, M.B., Shenk, T., 1991. Adenovirus virus-associated RNA and translation control. J. Virol. 65, 5657-5662. Matsumoto, M., Funami, K., Tanabe, M., Oshiumi, H., Shingai, M., Seto, Y., Yamamoto, A., Seya, T., 2003. Subcellular localization of toll-like receptor 3 in human dendritic cells. J. Immunol. 171, 3154-3162. Mazaheri, A., Prusas, C., Vo, M., Hess, M., 1998. Some strains of serotype 4 fowl adenovirus cause inclusion body hepatitis and hydropericardium syndrome in chickens. Avian Path. 27, 269-276. McCoy, R.J., Sheppard, M. 1997. DNA sequence analysis of an avian adenovirus terminal protein precursor. Virus Genes 14, 245-249. McCracken, R.M., McFerran, J.B., Evans, R.T., Connor, T.J., 1976. Experimental studies on the aetiology of inclusion body hepatitis. Avian Path. 5, 325-339. McFerran, J.B. and Smyth.J.A., 2000. Avian adenoviruses. Rev. Sci. Tech. 19, 589-601. McFerran, J.B., McCracken, R.M., Connor, T.J., Evans, R.T., 1976. Isolation of viruses from clinical outbreaks of inclusion body hepatitis. Avian Pathol. 5, 315-324. McGrath, W., Ding, J., Didwania, A., Sweet, R.M., Mabgek, W.F., 2003. Crystallographic structure at 1.6-Ǻ resolution of the human adenovirus proteinase in a covalent complex with its 11-amino-acid peptide cofactor: insights on a new fold. BBA 1648, 1-11.

Page 174: Sequence analysis, pathogenicity and cytokine gene

163

McGuire, K.A., Barlan, A.U., Griffin, T.M., Wiethoff, C.M., 2011. Adenovirus type 5 rupture of lysosomes leads to cathepsin B-dependent mitochondrial stress and production of reactive oxygen species. J. Virol. 85, 10806-10813. Meier, O. and Greber, U.F., 2004. Adenovirus endocytosis. J. Gene Med. 6, S152-S163. Micallef, M.J., Ohtsuki, T., Kohno, K., Tanabe, F., Ushio, S., Namba, M., Tanimoto, T., Torigoe, K., Fujii, M., Ikeda, M., Fukuda, S., Kurimoto, M., 1996. Interferon-gamma-inducing factor enhances T helper 1 cytokine production by stimulated human T cells: synergism with interleukin-12 for interferon-gamma production. Eur. J. Immunol. 26, 1647-1651. Minang, J.T., Arestrom, I., Zuber, B., Jonsson, G., Troye-Blomberg, M., Ahlborg, N., 2006. Nickel-induced IL-10 down-regulates Th1- but nto Th2-type cytokine responses to the contact allergen nickel. Clin. Exp. Immunol. 143, 494-502. Mittal, S.K., Prevec, L., Babiuk, L.A., Graham, F.L., 1992. Sequence analysis of bovine adenovirus type 3 early region 3 and fibre protein genes. J. Gen. Virol. 73, 3295-3300. Moore, K.W., de Waal Malefyt, R., Coffman, R.L., O’Garra, A., 2001. Interleukine-10 and the interleukin-10 receptor. Annu. Rev. Immunol. 19, 683-765. Mori, I., Hossain, M.J., Takeda, K., Okamura, H., Imai, Y., Kohsaka, S., 2001. Impaired microglial activation in the brain of IL-18-gene disrupted mice after neurovirulent influenza A virus infection. Virology 287, 163-70. Morin, N., Delsert, C., Klessig, D.F., 1989. Nuclear localization of the adenovirus DNA-binding protein: requirement for two signals and complementation during viral infection. Mol. Cell Biol. 9, 4372-4380. Morin, N. and Boulanger, P., 1986. Hexon trimerization occurring in an assembly defective, 100K temperature-sensitive mutant of adenovirus 2. Virology 152, 11-31. Morrison, M.D., Onions, D.E., Nicolson, L., 1997. Complete DNA sequence of canine adenovirus type 1. J. Gen. Virol. 78, 873-878. Murphy, K.M., Ouyang, W., Farrar, J.D., Yang, J., Ranganath, S., Asnagli, H., Afkarian, M., Murphy, T.l., 2000. Signaling and transcription in T helper development. Annu. Rev. Immunol. 18, 451-494. Muruve, D.A., 2004. The innate immune response to adenovirus vectors. Hum. Gene. Ther. 15, 1157-1166. Naeem, K., Rabbani, M., Hussain, M., Cheema, A.H., 1995. Development of cell culture vaccine against HPS in poultry. Pak. Vet. J. 15, 150-151.

Page 175: Sequence analysis, pathogenicity and cytokine gene

164

Nakamura, K., Mase, M., Yamaguchi, S., Shiobahara, T., Yuasa, N., 1999. Pathologic study of specific pathogen free chicks and hens inoculated with adenovirus isolated from hydropericardium syndrome. Avian Dis. 43, 414-423. Nakamura, K., Tanaka, H., Mase, M., Imada, T., Yamada, M., 2002. Pancreatic necrosis and ventricular erosion in adenovirus-associated hydropericardium syndrome of broilers. Vet. Pathol. 39, 403-406. Nakamura, K., Mase, M., Yamaguchi, S., Shiobahara, T., Yuasa, N., 1999. Pathologic study of specific pathogen free chicks and hens inoculated with adenovirus isolated from hydropericardium syndrome. Avian Dis. 43, 414-423. Nanduri, S., Raham, F., Williams, B.R.G., Qin, J., 2000. A dynamically tuned double-stranded RNA binding mechanism for the activation of antiviral kinase PKR. EMBO J. 19, 5567-5474. Nemerow, G.R., Pache, L., Reddy, V., Stewart, P.L., 2009. Insight into adenovirus host cell interactions from structural studies. Virology 384, 380-388. Norrby, E., 1969. The relationship between the soluble antigens and the virion of adenovirus type 3. IV. Immunological complexity of soluble components. Virology 37, 565-576. Oaks, J.L., Schrenzel, M., Rideout, B., Sandfort, C., 2005. Isolation and epidemiology of falcon adenovirus. J. Clin. Microbiol. 43, 3414-3420. O’Garra, A. and Vieira, P., 2007. T(H)1 cells control themselves by producing interleukin-10. Nat. Rev. Immunol. 7, 425-428. Ojkic, D., Krell, P.J., Tuboly, T., Nagy, É., 2008a. Characterization of fowl adenoviruses isolated in Ontario and Quebec, Canada. Can. J. Vet. Res. 72, 236-241. Ojkic, D., Martin, E., Swinton, S., Vaillancourt, P.J., Boulianne, M., Gomis, S., 2008b. Genotyping of Canadian isolates of fowl adenoviruses. Avian Pathol. 37, 95-100. Ojkic, D. and Nagy, E., 2003. Antibody response and virus tissue distribution in chickens inoculated with wild-type and recombinant fowl adenoviruses. Vaccine 22, 42-48. Ojkic, D. and Nagy, É., 2001. The long repeat region is dispensable for fowl adenovirus replication in vitro. Virology 283, 197-206. Ojkic, D. and Nagy, E., 2000. The complete nucleotide sequence of fowl adenovirus type 8. J. Gen. Virol. 81, 1833-1837. Okamura, H., Tsutsi, H., Komatsu, T., Yutsudo, M., Hakura, A., Tanimoto, T., Torigoe, K., Okura, T., Nukada, Y., Hattori, K., Akita, K., Namba, M., Tanabe, F., Konishi, K.,

Page 176: Sequence analysis, pathogenicity and cytokine gene

165

Fukuda, S., Kurimoto, M., 1995. Cloning of a new cytokine that induces IFN-gamma production by T cells. Nature 378, 88-91. Okuda, Y., Ono, M., Shibata, I., Sati, S., Akashi, H. 2006. Comparison of the polymerase chain reaction-restriction fragment length polymorphism pattern of the fiber gene and pathogenicity of serotype-1 fowl adenovirus isolates from gizzard erosions and from feces of clinically healthy chicks in Japan. J. Vet. Diagn. Invest. 18, 162-167. O’Neill, L.A. and Bowie, A.G., 2007. The family of five: TIR-domain-containing adaptors in Toll-like receptors signaling. Nat. Rev. Immunol. 7, 353-364. Ono, M., Okuda, Y., Shibata, I., Sato, S., Okada, K., 2004. Pathogenicity by parenteral injection of fowl adenovirus isolated from gizzard erosion and resistance to reinfection in adenoviral gizzard erosions in chickens. Vet. Patol. 41, 483-489. Oppenheim, J.J., Zachariae, N., Mukaida, N., Matsushima, K., 1991. Properties of the novel proinflammatory supergene intercrine cytokine family. Annu. Rev. Immunol. 9, 617-48. Pallister, J., Wright, P.J., Sheppard, M., 1996. A single gene encoding the fiber is responsible for variations in virulence in the fowl adenoviruses. J. Virol. 70, 5115-5122. Parcells, M.S., Lin, S-F., Dienglewicz, R.L., Majerciak, V., Robinson, D.R., Chen, C-H., Wu, Z., Dubyak, G.R., Brunovskis, P., Hunt, H.D., Lee, L.F., Kung, H-J., 2001. Marek’s disease virus (MDV) encodes an interleukin-8 homolog (vIL-8): characterization of the vIL-8 protein and a vIL-8 deletion mutant MDV. J. Virol. 5159-5173. Parvizi, P., Read, L.R., Abdul-Careem, M.F., Searson, A.J., Lusty, C., Lambourne, M., Thanthrige-Don, N., Burgess, S.C., Sharif, S., 2009. Cytokine gene expression in splenic CD4+ and CD8+ T cell subsets of genetically resistant and susceptible chickens infected with Marek’s disease virus. Vet. Immunol. Immunopathol. 132, 209-17. Pfaffl, M.W., 2001. A new mathematical model for relative quantification in real-time RT-PCR. Nucleic Acids Research. 29, 2003-2007. Philippe, C., Grgic, H., Ojkic, D., Nagy, E., 2007. Serologic monitoring of a broiler breeder flock previously affected by inclusion body hepatitis and testing of the progeny for vertical transmission of fowl adenoviruses. Can. J. Vet. Res. 71, 98-102. Philippe, C., Grgic, H., Nagy, E., 2005. Inclusion body hepatitis in young broiler breeders associated with a serotype 2 adenovirus in Ontario, Canada. J. Appl. Poult. Res. 14, 588-593. Pitcovski, J., Mualem, M., Rei-Koren, Z., Krispel, S., Shmueli, E., Peretz, Y., Gutter, B., Gallili, G.E., Michael, A., Goldberg, D., 1998. The complete DNA sequence and genome organization of the avian adenovirus, hemorrhagic enteritis virus. Virology 249, 307-315.

Page 177: Sequence analysis, pathogenicity and cytokine gene

166

Plachy, J., Weining, K.C., Kremmer, E., Puehler, F., Hala, K., Kaspers, B., Staeheli, P., 1999. Protective effects of type I and type II interferons toward Rous sarcoma virus-induced tumors in chickens. Virology 256, 85-91. Rasmussen, R., 2001. Quantification on the LightCycler. In Meuer, S., Wittwer, C. and Nakagawara , K (eds), Rapid Cycle Real-time PCR, Methods and Application. Springer Press, Heidelberg, pp. 21-34. Rasmussen, U.B., Schlesinger, Y., Pavirani, A., Mehtali, M., 1995. Sequence analysis of the canine adenovirus 2 fiber-encoding gene. Gene 159, 279-280. Rassa, J.C., Meyers, J.L., Zhang, Y., Kudaravalli, R., Ross, S.R., 2002. Murine retroviruses activate B cells via interaction with toll-like receptor 4. Proc. Natl. Acad. Sci. U.S.A. 99, 2281-2286. Ratcliffe, M.J.H. and Jacobsen, K.A., 1994. Rearrangement of immunoglobulin genes in chicken B cell development. Semin. Immunol. 6, 175-184. Rathinam, V., Fitzgerald, K., 2010. Inflammasomes and anti-viral immunity. J. Clin. Immunol. 632-637. Raue, R. and Hess, M., 1998. Hexon based PCRs combined with restriction enzyme analysis for rapid detection and differentiation of fowl adenoviruses and egg drop syndrome virus. J. Virol. Methods. 73, 211-217. Raue, R., Hafez, H.M., Hess, M., 2002. A fibre gene-based pigeon adenovirus. Avian Path. 31, 95-99. Rathinam, V.A.K., Fitzgerald, K.A., 2011. Innate immune sensing of DNA viruses. Virology 2, 153-162. Reece, R.L., Barr, D.A., Grix, D.C., 1985. An investigation of vertical transmission of a fowl adenovirus serotype 8. Aust. Vet. J. 62, 136-137. Reddy, P.S., Nagy, É., Derbyshire, J.B., 1995. Sequence analysis of putative pVIII, E3 and fibre regions of porcine adenovirus type 3. Virus Res. 36, 97-106. Rendall, R.E., Goodbourn, S., 2008. Interferons and viruses: an interplay between induction, signaling, antiviral responses and virus countermeasures. J. Gen. Virol. 89, 1-47. Romanova, N., Corredor, J.C., Nagy, É., 2009. Detection and quantitation of fowl adenovirus genome by a real-time PCR assay. J. Virol. Methods. 159, 58-63. Rothwell, L., Young, J.R., Zoorob, R., Whittaker, C.A., Hesketh, P., Archer, A., Smith, A.L., Kaiser, P., 2004. Cloning and characterization of chicken IL-10 and its role in the immune response to Eimeria maxima. J. Immunol. 173, 2675-2682.

Page 178: Sequence analysis, pathogenicity and cytokine gene

167

Roy, P., Koteeswaran, M., Manickam, R., 1999. Efficacy of an inactivated oil emulsion vaccine against hydropericardium syndrome in broilers. Vet. Rec. 145, 458-459. Saban, S.D., Nepomuceno, R.R., Gritton, L.D., Nemerow, G.R., Stewart, P.L., 2005. CryoEM structure at 9 Å resolution of an adenovirus vector targeted to hematopoietic cells. J. Mol. Biol. 349, 526-537. Saifuddin, M. and Wilks, C., 1991. Pathogenesis of an acute viral hepatitis: inclusion body hepatitis in the chicken. Arch. Virol. 116, 33-43. Salone, B., Martina, Y., Piersanti, S., Cundari, E., Cherubini, G., Franqueville, L., Failla, C.M., Boulanger, P., Saggio, I., 2003. Integrin α3β1 is an alternate cellular receptor for adenovirus serotype 5. J. Virol. 77, 13448–13454. San Martin, C. and Burnett, R.M., 2003. Structural studies on adenoviruses. Curr. Top. Microbiol. Immunol. 272, 57-94. Schmid, S.I. and Hearing, P., 1995. Selective encapsidation of adenovirus DNA. Curr. Top. Microbiol. Immunol. 199 (Pt 1), 67-80. Schrenzel, M., Oaks, L.J., Rotstein, D., Maalouf, G., Snook, E., Sandfort, C., Rideout, B., 2005. Characterization of a new species of adenovirus in falcons. J. Clin. Microbiol. 43, 3402-3413. Sekellick, M.J., Ferrandino, A.F., Hopkins, D.A., Marcus, P.I., 1994. Chicken interferon gene: cloning, expression, and analysis. J. Interferon Res. 14, 71-79. Shafique, M. and Shakoori, A.R., 1994 Transmission of hydropericardium syndrome in poultry. Pak. J. Zool. 26, 145-148. Shane, S.M., 1996. Hydropericardium-hepatitis syndrome, the current world situation. Zootecnica Int. 18, 20-27. Sharpless, G.R. and Jungherr, E.L., 1961. Characterization of 2 viruses obtained from lymphomatous liver. Am. J. Vet. Res. 22, 937-943. Shen, L. and Spector, D.J., 2003. Local character of readthrough activation in adenovirus type 5 early region 1 transcription control. J. Virol. 77, 9266-9277. Sheppard, M., Tsatas E., Johnson, M., 1998. DNA sequence analysis of the genes for the fowl adenovirus serotype 10 putative 33K and pVIII. DNA Seq 9, 37-43. Signas, C., Akusjarvi, G., Petterson, U., 1985. Adenovirus 3 fiber polypeptide gene: Implications for the structure of the fiber protein. J. Virol. 53, 672-678.

Page 179: Sequence analysis, pathogenicity and cytokine gene

168

Sileo, I., Franson, J.C., Graham, D.I., Domermuth, C.H., Rattner, B.A., Pattee, O.H., 1983. Hemorrhagic enteritis in captive American kestrels (Falco sparverius). J. Wild. Dis. 19, 244-247. Silvestry, M., Lindert, S., Smith, J.G., Maier, O., Wiethoff, C.M., Nemerow, G.R., Stewart, L.P., 2009. Cryo-electron microscopy structure of acenovirus type 2 temperature-sensitive mutant 1 reveals insight into the cell entry defect. J. Virol. 83, 7375-83. Sloan, D.D. and Jerome, K.R., 2007. Herpes simplex virus remodels T-cell receptors signaling resulting in p38-dependent selective synthesis of interleukine 10. J. Virol. 81, 12504-12514. Soudais, C., Boutin, S., Hong, S.S., Chillon, M.I., Danos, O., Bergelson, J.M., Boulanger, P., Kremer, E.J., 2000. Canine adenovirus type 2 attachment and internalization: coxsackievirus-adenovirus receptor, alternative receptors, and an RGD-independent pathway. J. Virol. 74, 10639-10649. Steed, A., Buch, T., Waisman, A., Virgin, H.W.4th., 2007. Gamma interferon blocks gammaherpesvirus reactivation from latency in a cell type-specific manner. J. Virol. 81, 6134-40. Steer, P.A., Kirkpatrick, N.C., O’Rourke, D., Noormohammad, A.H., 2009. Classification of fowl adenovirus serotypes by use of High-Resolution Melting-Curve Analysis of the hexon gene region. J. Clin. Microbiol. 47, 311-321. Steinbrink, K., Wolfl, M., Jonuleit, H., Knop, J., Enk, A.H., 1997. Induction of tolerance by IL-10-treated dendritic cells. J. Immunol. 159, 4772-4780. Stewart, P.L., Chiu, C.Y., Huang, S., Muir, T., Zhao, Y., Chait, B., Mathias, P., Nemerow, G.R., 1997. Cryo-EM visualization of an exposed RGD epitope on adenovirus that escapes antibodyneutralization. EMBO J. 16, 1189-1198. Stewart, P.L., Fuller, S.D., Burnett, R.M., 1993. Difference imaging of adenovirus: bridging the resolution gap between X-Ray crystallography and electron microscopy. EMBO J. 12, 2589-2599. Stewart, P.L., Burnett, R.M., Cyrklaff, M., Fuller, S.D., 1991. Image reconstruction reveals the complex molecular organization of adenovirus. Cell 67, 145-154. Taddeo, B., Esclatine, A., Roizman, B., 2002. The patterns of accumulation of cellular RNAs in cells infected with a wild-type and a mutant herpes simplex virus 1 lacking the virion host shutoff gene. Proc. Natl. Acad. Sci. USA 99, 17031-17036.

Page 180: Sequence analysis, pathogenicity and cytokine gene

169

Takaoka, A., Wang, Z., Choi, M.K., Yanai, H., Negishi, H., Ban, T., Lu, Y., Miyagishi, M., Kodama, T., Honda, K., Ohba, Y., Taniguchi, T., 2007. DAI (DLM-1/ZBP1) is cytosolic DNA sensor and an activator of innate immune response. Nature 448, 501-505. Tamanini, A., Nicolis, E., Bonizzato, E., Bezzerri, V., Melotti, P., Assael, B.M., Cabrini, G., 2006. Interaction of adenovirus type 5 fiber with the coxsackievirus and adenovirus receptor activates inflammatory response in human respiratory cells. J. Virol. 80, 11241-11254. Tan, P.K., Michou, A-I., Bergelson, J.M., Cotten, M., 2001. Defining CAR as a cellular receptor for the avian adenovirus CELO using a genetic analysis of the two viral fibre proteins. J. Gen. Virol. 82, 1465-1472. Tanaka-Kataoka, M., Kunikata, T., Takayama, S., Iwaki, K., Ohashi, K., Ikeda, M., 1999. In vivo antiviral effect of interleukin 18 in a mouse model of vaccinia virus infection. Cytokine 11, 593-99. Tate, V.E. and Philipson, L., 1979. Parental adenovirus DNA accumulates in nucleosome-like structures in infected cells. Nucleic Acids Res. 6, 2769-2785. Temperley, N.D., Berlin, S., Paton, I.R., Griffin, K.D., Burt, D.W., 2008. Evolution of the chicken Toll-like receptor gene family: A story of gene gain and gene loss. BMC Genomics 9, 62-74. Thaci, B., Ulasov, I.V., Wainwright, D.A., Lesniak, M.S., 2011. The challenge for gene therapy: innate immune response to adenoviruses. Oncotarget 2, 113-121. Tomaszewski, E.K. and Phalen, D.N., 2007. Falcon adenovirus in an American Kestrel (Falco sparverius). J. Avian Med. Surg. 21, 135-139. Theofilopoulos, A.N., Baccala, R., Beutler, B., Kono, D.H., 2005. Type I interferons (α/β) in immunity and autoimmunity. Annu. Rev. Immunol. 23, 307-336. Tollefson, A.E., Hermiston, T.W., Lichteinstein, D.L., Colle, C.F., Tripp, R.A., Dimitrov, T., Toth, K., Wells, C.E., Doherty, P.C., Wold, W.S., 1998. Forced degradation of Fas inhibits apoptosis in adenovirus-infected cells. Nature 392, 726-730. Tollefson, A.E., Ryerse, J.S., Scaria, A., Hermiston, T.W., Wold, W.S., 1996. The E3-11.6-kDa adenovirus death protein (ADP) is required for efficient cell death: characterization of cells infected with Adp mutants. Virology 220, 152-162. Tomko, R.P., Xu, R., Philipson, L., 1997. HCAR and MCAR: the human and mouse cellular receptors for subgroup C adenoviruses and group B coxsackieviruses. Proc. Natl. Acad. Sci. USA 94, 3352-3356.

Page 181: Sequence analysis, pathogenicity and cytokine gene

170

Toogood, C.I., Crompton, J., Hay, R.T., 1992. Antipeptide antisera define neutralizing epitopes on the adenovirus hexon. J. Gen. Virol. 73, 1429-1435. Toro, H., Gonzalez, C., Cerda, L., Morales, M.A., Dooner, P., Salamero, M., 2002. Prevention of inclusion body hepatitis/hydropericardium syndrome in progeny chickens by vaccination of breeders with fowl adenovirus and chicken anemia virus. Avian Dis. 46, 547-554. Toro, H., Gonzalez, O., Escobar, C., Cerda, L., Morales, M.A., Gonzalez, C., 2001. Vertical induction of inclusion body hepatitis/hydropericardium syndrome with fowl adenovirus and chicken anemia virus. Avian Dis. 45, 215-222. Toro, H., Gonzales, C., Cerda, L., Hess, M., Reyes, E., Geissea, C., 2000. Chicken anemia virus and fowl adenoviruses: Association to induce the inclusion body hepatitis/hydropericardium syndrome. Avian Dis. 44, 51-58. Toro, H., Prusas, R., Raue, R., Cerda, L., Geisse, C., Gonzalez, C., Hess, M., 1999. Characterization of fowl adenoviruses from outbreaks of inclusion body hepatitis hydropericardium syndrome in Chile. Avian Dis. 43, 262-270. Tsernoglou, D., Tucker, A.D., van der Vliet, P.C., 1984. Crystallization of a fragment of the adenovirus DNA binding protein. J. Mol. Biol. 172, 237-239. Tucker, P.A., Tsernouglou, D., Tucker, A.D., Coenjaerts, F.E., Leenders, H., van der Vliet, P.C., 1994. Crystal structure of the adenovirus DNA binding protein reveals a hook-on model for cooperative DNA binding. EMBO J. 13, 2994-3002. Umene, K., Watson, R.J., Enquist, L.W., 1984. Tandem repeated DNA in an intergenic region of Herpes simplex virus type 1 (Patton). Gene 30, 33-39. van der Vliet, P.C., Keegstra, W., Jansz, H.S., 1978. Complex formation between the adenovirus type 5 DNA-binding protein and single-stranded DNA. Eur. J. biochem. 83, 389-398. van Oostrum, J. and Burnett, R.M., 1985. Molecular composition of the adenovirus type 2 virion. J. Virol. 56, 434-448. van Raaij, M.J., Mitraki, A., Lavigne, G., Cusack, S., 1999. A triple beta-spiral in the adenovirus fibre shaft reveals a new structural motif for a fibrous protein. Nature 401, 935-938. Van Wettere, A.J., Wiinschmann, A., Latimer, K.S., Redig, P.T., 2005. Adenovirus infection in taita falcons (Falco fasciinucha) and hybrid falcons (Falco rusticolus x Falco peregrinus). J Avian Med Surg. 19, 280-285.

Page 182: Sequence analysis, pathogenicity and cytokine gene

171

Venktesh, A., Watt, F., Xu, Z.Z. & Both, G.W., 1998. Ovine adenovirus (OAV287) lacks a virus-associated RNA gene. J.Gen. Virol 79, 509-516. Vereecken, M., de Herdt, P., Ducatelle, R., 1998. Adenovirus infection in pigeons: a review. Avian Pathol. 27, 333-338. Vrati S., Boyle, D., Kocherhans, R., Both, G.W., 1995. Sequence of ovine adenovirus homologues for 100K hexon assembly, 33K, pVIII and fiber genes: early region E3 is not in the expected location. Virology 209, 400-408. Wagner, H., 2004. The immunobiology of the TLR9 subfamily. Trends Immunol. 25, 381-386. Walters, R.W., Freimuth, P., Moninger, T.O., 2002. Adenovirus fiber disrupts CAR-mediated intercellular adhesion allowing virus escape. Cell 110, 789-799. Ward, P., Dean, F.B., O’Donnell, M.E., Berns, K.I., 1998. Role of the adenovirus DNA-binding protein in in vitro adeno-associated virus DNA replication. J. Virol. 72, 420-427. Warr, G.W., Magor, K.E., Higgins, D.A., 1995. IgY: clues to the origins of modern antibodies. Immunol. Today 16, 392-398. Weber, J.M., 1995. Adenovirus endopeptidase and its role in virus infection. Curr. Top. Microbiol. Immunol. 199, 227-235. Webster, A., Leith, I.R., Nicholson, J., Hounsell, J., Hay, R.T., 1997. Role of preterminal protein processing in adenovirus replication. J. Virol. 71, 6381-6389. Whyte, P., Buchkovich, K.J., Horowitz, J.M., Friend, S.H., Raybuck, M., Weinberg, R.A., Harlow, E., 1988. Association between an oncogene and an anti-oncogene: the adenovirus E1A proteins bind to the retinoblastoma gene product. Nature 334, 124-129. Wickham, T.J., Mathias, P., Cheresh, D.A., Nemerow, G.R., 1993. Integrins alpha v beta 3 and alpha v beta 5 promote adenovirus internalization but not virus attachment. Cell 73, 309-319. Wiethoff, C.M., Wodrich, H., Gerace, L., Nemerow, G., 2005. Adenovirus protein VI mediates membrane disruption following capsid disassembly. J. Virol. 79, 1992-2000. Wold, W.S., Hermiston, T.W., Tollefson, A.E., 1994. Adenovirus proteins that subvert host defenses. Trends Microbiol. 2, 437-443. Wold, W.S., Gooding, L.R., 1991. Region E3 of adenovirus: a cassette of genes involved in host immunosurveillance and virus-cell interactions. Virology 184, 1-8.

Page 183: Sequence analysis, pathogenicity and cytokine gene

172

Wodrich, H., Henaff, D., Jammart, B., Segura-Morales, C., Seelmeir, S., Coux, O., Ruzsics, Z., Weithoff, C.M., Kremer, E.J., 2010. A capsid-encoded PPxY-motif facilitates adenovirus entry. PLoS Pathog. 3, 1-18. Wu, W., Booth, J.L., Duggan, E.S., Patel, K.B., Coggeshall, K.M., Metcalf, P.J., 2010. Human lung innate immune cytokine response to adenovirus type 7. J. Gen. Virol. 91, 1155-1163. Wu, W., Booth, J.L., Coggeshall, K.M., Metcalf, J.P., 2006. Calcium-dependant viral internalizationis required for adenovirus type 7 induction of IL-8 protein. Virology 355, 18-29. Xia, D., Henry, L.J., Gerard, R.D., Deisenhofer, J., 1994. Chrystal structure of the receptor-binding domain of adenovirus type 5 fiber protein at 1.7 Å resolution. Structure 2, 1259-1270. Xing, Z. and Schat, K.A., 2000. Inhibitory effects on nitric oxide and γ interferon on in vitro and in vivo replication of Marek’s disease virus. J. Virol. 74, 3605-12. Yang, Y., McKerlie, C., Borenstein, S.H., Lu, Z., Schito, M., Chamberlain, J.W., Buchwald, M., 2002. Transgenic expression in mouse lung reveals distinct biological roles for the adenovirus type 5 E1A243-and 289-amino-acid proteins. J. Virol. 76, 8910-8919. Yates, V.J., Rhee, Y.O., Fry, D.E., Al Mishad, A.M., McCormick, K.J. 1977. The presence of avian adenoviruses and adenovirus associated viruses in healthy chickens. Avian Dis. 20, 146-152. Yoneyama, M., Kikuchi, M., Natsukawa, T., Shinobu, N., Imaizumi, T., Miyagishi, M., Taira, K., Akira, S., Fujita, T., 2004. The RNA helicase RIG-I has an essential function in double-stranded RNA-induced innate antiviral responses. Nat Immunol. 5, 730-737. Yoshimoto, T., Takeda, K., Tanaka, T., Ohkusu, K., Kashiwamura, S., Okamura, H., Akira, S., Nakanishi, K., 1998. IL-12 up-regulates IL-18 receptor expression on T cells, Th1 cells, and B cells: synergism with IL-18 for IFN-gamma production. J. Immunol. 161, 3400-3407. Young, J.A., Purcell, D.A., Kavanagh, P.J., 1972. Inclusion body hepatitis outbreak in broiler flocks. Vet. Rec. 90, 72-76. Zellen, G.K., Key, D.W., Jack, S.W., 1989. Adenoviral pancreatitis in guinea fowl (Numida meleagris). Avian Dis. 33, 586-589. Zhang, W. and Imperiale, M.J., 2003. Requirement of the adenovirus IVa2 protein for virus assembly. J. Virol. 77, 3586-3594.

Page 184: Sequence analysis, pathogenicity and cytokine gene

173

Zhou, Y., Ficzycz, A., Tikoo, S.K., 2007. Porcine adenovirus type 3 E1Blarge protein downregulates the induction of IL-8. Virol. J. 4, 60-68. Zijderveld, D.C. and van der Vliet, P.C., 1994. Helix-destabilizing properties of the adenovirus DNA-binding protein. J. Virol. 68, 1158-1164. Zsak, L. and Kisary, J., 1984. Grouping of fowl adenoviruses based upon the restriction patterns of DNA generated by BamHI and HindIII. Intervirology 22, 110-114.