Download Oral inoculation of chickens with a candidate fowl

Document related concepts

Monoclonal antibody wikipedia , lookup

Cancer immunotherapy wikipedia , lookup

Psychoneuroimmunology wikipedia , lookup

Molecular mimicry wikipedia , lookup

Innate immune system wikipedia , lookup

DNA vaccination wikipedia , lookup

Henipavirus wikipedia , lookup

Immunomics wikipedia , lookup

Hepatitis B wikipedia , lookup

Transcript
Oral inoculation of chickens with a candidate fowl adenovirus 9 vector
and functional studies of fowl adenovirus 9 ORF1
by
Li Deng
A Thesis
presented to
The University of Guelph
In partial fulfilment of requirements
for the degree of
Doctor of Philosophy
in
Pathobiology
Guelph, Ontario, Canada
© Li Deng, May, 2015
ABSTRACT
ORAL INOCULATION OF CHICKENS WITH A CANDIDATE
FOWL ADENOVIRUS 9 VECTOR AND FUNCTIONAL STUDIES OF
FOWL ADENOVIRUS 9 ORF1
Li Deng
University of Guelph, 2015
Advisor:
Dr. Éva Nagy
FAdV-9Δ4, lacking six open reading frames (ORFs) 0, 1, 1A, 1B, 1C and 2 at the
left end of FAdV-9 genome, shows potential as a vaccine vector. In vivo studies
demonstrated that FAdV-9Δ4 does not replicate as efficiently as wild type (wt) virus in
intramuscularly inoculated chickens, suggesting the important roles of the left-end ORFs.
The fecal-oral route is the natural route of FAdV infection, and the oral administration
confers several advantages compared to administration through other routes when
developing a vaccine. Therefore, the objectives of this study were to investigate the
effects of oral inoculation with both FAdV-9Δ4 and wtFAdV-9 in chickens and to
explore the functions of one of the left-end ORFs, ORF1.
Chickens were orally inoculated with FAdV-9Δ4 or wtFAdV-9. Compared to
wtFAdV-9 group, reduced virus shedding in feces, lower viral loads in tissues, and lower
antibody response were found in FAdV-9Δ4 group. There were also significant
differences between FAdV-9Δ4 and wtFAdV-9 in terms of the induction of cytokine
mRNA expression, including interferon α (IFN-α), IFN-γ, and interleukin-12 (IL-12).
This indicated the important roles of the six ORFs in modulation of the host immune
response.
ORF1, a homolog of deoxyuridine 5′-triphosphate pyrophosphatase (dUTPase),
was hypothesized to be important in virus replication and regulation of host immune
response. In this project, FAdV-9 ORF1 was verified as a functional dUTPase and its
molecular features were characterized, including transcription and protein expression
patterns and cellular localization. A dUTPase knockout virus (ORF1stop) and its rescued
revertant (resORF1) were generated. Functional studies showed that FAdV-9 dUTPase
was not required for virus replication in vitro, but played a role in virus replication in vivo.
FAdV-9 dUTPase also contributed to the regulation of the expression of IFN-α, IFN-β,
and IFN-γ both in vitro and in vivo, and the host antibody response as well.
This is the first study to functionally identify an early gene of FAdV-9. The data
presented are helpful for better understanding of the molecular biology of FAdVs and for
exploring the mechanism of the host immune response against FAdV infections.
ACKNOWLEDGEMENTS
I wish to express my sincere gratitude and thanks to my dear advisor Dr. Éva Nagy, for
her continued guidance and support in all aspects of my research in the past years, and for
her kindness and patience throughout my PhD program. I am also very grateful for her
financial support.
I would also like to sincerely thank Dr. Peter Krell, Dr. Shayan Sharif, and Dr. Ray Lu,
who served on my advisory committee and offered valuable advice, suggestions, ideas
and comments throughout this study. I am also thankful for Dr. Davor Ojkić and Dr.
Byram Bridle, who served on my qualifying examination committee and motivated me to
make achievements in science.
I would like to thank Bryan Griffin and all other members of the Nagy lab, both past and
present, including James Ackford, David Leishman, Betty-Anne McBey, Dr. Yanlong Pei,
Dr. Robert Kozak, Dr. Helena Grgić, Dr. Andres Diaz, Dexter Endozo, Javier Hernandez,
Nathan Orr, and Dr. Xiaobing Qin. You are all good colleagues and made the Nagy lab a
wonderful environment.
I am also thankful for Dr. Sarah Wootton and her post-docs and students, including Dr.
Jondavid de Jong, Dr. Nicolle Petrik, Dr. Darrick Yu, Dr. Scott Walsh, and Lisa Santry,
for all the assistance in the past years. I would also like to thank Leah Read and Tony
Cengija for their help during my experimental periods.
I would also like to acknowledge China Scholarship Council, University of Guelph and
Northwest Agriculture and Forestry University (Yangling, China) for providing me the
financial support to conduct my graduate studies.
iv
Last but never be the least, I am very grateful to my family in China, and my wife Ying
(Yvonne) He, for all the love, support, assistance, encouragement and all the other things
they have offered to me.
v
TABLE OF CONTENTS
ABSTRACT........................................................................................................................ ii
ACKNOWLEDGEMENTS ............................................................................................... iv
TABLE OF CONTENTS ................................................................................................... vi
LIST OF TABLES ............................................................................................................. ix
LIST OF FIGURES ............................................................................................................ x
LIST OF ABBREVIATIONS ............................................................................................ xi
Chapter 1 ............................................................................................................................. 1
GENERAL INTRODUCTION ................................................................................................ 1
LITERATURE REVIEW ....................................................................................................... 2
1. Adenovirus taxonomy ............................................................................................. 2
2. Virion structure ....................................................................................................... 4
3. Genome organization of mastadenoviruses and early gene functions .................... 7
3.1 The E1A region ................................................................................................. 8
3.2 The E1B region ................................................................................................. 9
3.3 E2 proteins ...................................................................................................... 14
3.4 The E3 region .................................................................................................. 15
3.4 E4 proteins ...................................................................................................... 16
3.5 Intermediate and late genes ............................................................................. 16
4. Molecular biology of aviadenoviruses .................................................................. 17
4.1 Genome organization of aviadenoviruses ....................................................... 17
4.2 Functionally identified early genes in FAdV genome .................................... 19
4.3 Study of FAdV-9 genome ............................................................................... 23
5. Disease associated with aviadenoviruses .............................................................. 24
6. Diagnosis of fowl adenovirus infection ................................................................ 25
7. Immune response to human adenovirus ................................................................ 28
7.1 Innate immune responses ................................................................................ 28
7.2 Adaptive immune responses ........................................................................... 31
8. Immune responses to fowl adenoviruses .............................................................. 31
HYPOTHESES AND OBJECTIVES ...................................................................................... 34
Chapter 2. Oral inoculation of chickens with a candidate fowl adenovirus 9 vector ....... 35
vi
ABSTRACT ..................................................................................................................... 36
INTRODUCTION .............................................................................................................. 38
MATERIALS AND METHODS ........................................................................................... 40
Viruses and cells ....................................................................................................... 40
Animal experiment.................................................................................................... 40
Statistical analysis ..................................................................................................... 41
RESULTS ........................................................................................................................ 41
Virus shedding .......................................................................................................... 42
Viral genome copy number in tissues ....................................................................... 42
Antibody response .................................................................................................... 43
Cytokine gene expression in tissues ......................................................................... 43
DISCUSSION ................................................................................................................... 45
ACKNOWLEDGMENTS .................................................................................................... 50
Chapter 3. Characterization and functional studies of fowl adenovirus 9 dUTPase ........ 58
ABSTRACT ..................................................................................................................... 59
INTRODUCTION .............................................................................................................. 60
MATERIALS AND METHODS ........................................................................................... 62
Viruses and cells ....................................................................................................... 62
dUTPase enzyme activity assay ................................................................................ 63
Generation of mutant viruses .................................................................................... 64
RT-PCR and qRT-PCR ............................................................................................. 65
Western blot .............................................................................................................. 66
Immunofluorescence assay ....................................................................................... 66
Statistical analysis ..................................................................................................... 66
RESULTS ........................................................................................................................ 67
Bioinformatics analysis of FAdV-9 ORF1 as a potential dUTPase ......................... 67
FAdV-9 ORF1 has dUTPase enzymatic activity ...................................................... 68
In vitro characteristics of ORF1 mutant viruses ....................................................... 69
Transcription and protein expression profiles of ORF1 ........................................... 69
Cellular localization of ORF1 protein....................................................................... 70
Cytokine gene mRNA expression induced by wild type and mutant viruses ........... 71
DISCUSSION ................................................................................................................... 73
ACKNOWLEDGEMENTS................................................................................................... 77
vii
Chapter 4. Fowl adenovirus 9 dUTPase plays a role in virus replication in vivo and in the
regulation of the host immune response ........................................................................... 89
ABSTRACT ..................................................................................................................... 90
INTRODUCTION .............................................................................................................. 91
MATERIALS AND METHODS ........................................................................................... 93
Viruses and cells ....................................................................................................... 93
Experimental animals................................................................................................ 93
Animal experimental design ..................................................................................... 93
DNA and RNA extraction ......................................................................................... 94
qPCR and qRT-PCR ................................................................................................. 94
Enzyme-linked immunosorbent assay (ELISA) ....................................................... 95
Statistical analysis ..................................................................................................... 95
RESULTS ........................................................................................................................ 95
Virus shedding .......................................................................................................... 95
Viral loads in tissues ................................................................................................. 96
Antibody response .................................................................................................... 97
Cytokine gene expression in tissues ......................................................................... 97
DISCUSSION ................................................................................................................. 100
ACKNOWLEDGEMENT .................................................................................................. 103
Chapter 5. General discussion......................................................................................... 111
References ....................................................................................................................... 119
viii
LIST OF TABLES
TABLE 1.1 Summary of 13 adenoviral proteins and their functions ................................. 6
TABLE 1.2 Adenovirus gene products and their functions .............................................. 11
TABLE 2.1 Virus titers (pfu/ml) in the feces of chickens orally inoculated with FAdV9Δ4 and wtFAdV-9. .................................................................................................. 51
TABLE 2.2 Viral genome copy number in tissues ........................................................... 52
TABLE 3.1 Primers for generating mutant viruses ......................................................... 78
TABLE 3.2 Primers used for cytokines for qRT-PCR .................................................... 79
SUPPLEMENTARY TABLE S1 Pairwise identities of dUTPase amino acid sequences of
adenoviruses. ............................................................................................................. 80
TABLE 4.1 Virus titers in the feces of infected chickens .............................................. 105
TABLE 4.2 Viral genome copy number in tissues ......................................................... 106
ix
LIST OF FIGURES
FIGURE 1.1 Phylogenetic tree of adenovirus genomes. .................................................... 3
FIGURE 1.2 Structure of adenovirus.................................................................................. 5
FIGURE 1.3 Transcription map of the mastadenovirus genome. ..................................... 10
FIGURE 1.4 ORFs in the left end of FAdV-9 genome. ................................................... 27
FIGURE 2.1 Antibody response in orally inoculated chickens ........................................ 53
FIGURE 2.2 Cytokine mRNA expression in spleen samples........................................... 54
FIGURE 2.3 Cytokine mRNA expression in liver samples ............................................. 55
FIGURE 2.4 Cytokine mRNA expression in bursa of Fabricius samples ........................ 56
FIGURE 2.5 Cytokine mRNA expression in cecal tonsil samples ................................... 57
SUPPLEMENTARY FIGURE S1 Generation of the mutant viruses. ............................. 81
FIGURE 3.1 Multiple alignment of amino acid sequences of dUTPase homologs. ........ 82
FIGURE 3.2 PCR-based dUTPase enzyme activity assay. .............................................. 83
FIGURE 3.3 Viral DNA accumulation curve and one-step growth curve. ...................... 84
FIGURE 3.4 Transcription and protein expression profiles of ORF1. ............................. 85
FIGURE 3.5 Cellular localization of ORF1 protein ......................................................... 86
FIGURE 3.6 Cytokine mRNA expressions in CH-SAH cells .......................................... 87
FIGURE 3.7 Cytokine mRNA expressions in Celi cells .................................................. 88
FIGURE 4.1 FAdV-specific IgG antibody response in chickens. .................................. 107
FIGURE 4.2 Cytokine mRNA expression in spleen samples of chickens. .................... 108
FIGURE 4.3 Cytokine mRNA expression in liver samples of chickens. ....................... 109
FIGURE 4.4 Cytokine mRNA expression in cecal tonsil samples of chickens. ............ 110
x
LIST OF ABBREVIATIONS
aa
Ab
Ad
ADP
AdV
AGPT
ANOVA
APCs
BAdV
bp
CAV
CBP
Celi
CFIA
CH-SAH
CPE
CTLs
DBP
d.p.i.
DCs
DMEM-F12
dsDNA
dUTPase
E. coli
E1A
EDSV
EDTA
EGFP
ELISA
ER
FAdV
FBS
GAM-1
GoAdV
gp19k
h.p.i.
HAdV
HAPI
HAT
HDAC1
HHV
HI
amino acid
antibody
adenoviral
adenovirus death protein
adenovirus
agar gel precipitation test
analysis of variance
antigen presenting cells
bovine adenovirus
base pair
chicken anemia virus
CREB-binding protein
chicken embryo liver
Canadian Food Inspection Agency
chicken hepatoma
cytopathic effect
cytotoxic T lymphocytes
DNA binding protein
days post-inoculation
dendritic cells
Dulbecco's modified Eagle's medium
and nutrient mixture Ham's F-12 medium
double-strand DNA
deoxyuridine 5′-triphosphate pyrophosphatase
Escherichia coli
early region 1A
egg drop syndrome virus
ethylenediamine tetraacetic acid
enhanced green fluorescence protein
enzyme-linked immunosorbent assay
endoplasmic reticulum
fowl adenovirus
fetal bovine serum
Gallus anti morte protein 1
goose adenovirus
glycoprotein 19k
hours post-infection
human adenovirus
highly pathogenic avian influenza
histone-acetyltransferase
histone deacetylase 1
human herpesvirus
hemagglutination inhibition
xi
hMDM
HPS
hsp40
HSV-1
i.m.
IBDV
IBH
IBV
IFN
Ig
IL
IRF
ITR
kDa
mDCs
MHC-I
MHV-68
ml
MLP
MLTU
MOI
MTOC
NBs
NF
NF-κB
NK
nt
ORF
PAdV
PAMP
PBMCs
PBS
PCR
pDCs
pfu
PHA-P
PKR
PML
pol
PPi
pRb
PRRs
pTP
RFLP
RNA
RT
human monocyte-derived macrophage
hydropericardium syndrome
heat-shock proteins 40
herpes simplex virus type 1
intramuscularly
infectious bursal disease virus
inclusion body hepatitis
infectious bronchitis virus
interferon
immunoglobulin
interleukin
interferon regulatory factor
inverted terminal repeat
kilodalton
myeloid DCs
major histocompatability complex class I
murine gammaherpesvirus 68
mililiter
major late promoter
major late transcription unit
multiplicity of infection
microtubule organizing center
nuclear bodies
nuclear factor
nuclear factor kappa B
natural killer
nucleotide
open reading frame
porcine adenovirus
pathogen-associated molecular pattern
peripheral blood mononuclear cells
phosphate buffered saline
polymerase chain reaction
plasmacytoid dendritic cells
plaque forming unit
phytohemagglutinin-P
protein kinase R
promyelocytic leukemia
polymerase
pyrophosphate
retinoblastoma protein
pattern recognition receptors
precursor terminal protein
restriction fragment length polymorphisms
ribonucleic acid
reverse transcription
xii
SAdV
SEM
SPF
SRBC
ssDNA
TAdV
Th1
TLRs
TNF
TP
TPL
TR
TSA
TSAdV
VA-RNAs
VZV
wt
μl
simian adenovirus
standard error of the mean
specific-pathogen-free
sheep red blood cells
single-strand DNA
turkey adenovirus
T helper 1
toll-like receptors
tumor necrosis factor
terminal protein
tripartite leader sequence
tandem repeat region
trichostatin A
tree shrew adenovirus
virus-associated RNAs
varicella zoster virus
wild type
microliter
xiii
Chapter 1
General introduction
Adenoviruses (AdVs) have been extensively developed as vaccine vectors. Numerous
studies have proven that adenoviral (Ad) vectors meet the most important criteria of an
ideal vaccine vector in terms of efficacy, safety and stability (Tatsis and Ertl, 2004).
Compared to other viral vectors, Ad vectors offer a number of advantages: a broad range
of natural host, infection of a variety of both dividing and quiescent cell types, robust
transgene expression, easily grown to high titers, easy manipulation, lack of serious threat
of horizontal transmission, induction of potent innate and adaptive immune responses,
and transduction of antigen presenting cells (Ahi et al., 2011; Bangari and Mittal, 2006).
Fowl adenoviruses (FAdVs) can be isolated from both sick and healthy birds in poultry
farms and have a worldwide distribution (Hess, 2013). In recent years, a number of
researchers have focussed on developing FAdVs, such as FAdV-1, FAdV-8, FAdV-9 and
FAdV-10, as vaccine vectors and have shown promising results (Francois et al., 2004;
Sheppard et al., 1998; Johnson et al., 2003; Corredor and Nagy, 2010b; 2011; Ojkić and
Nagy, 2001). However, compared to the extensive studies in human adenoviruses
(HAdVs), the study of the molecular biology of FAdVs is far behind.
The genome of FAdV-9 has been fully sequenced (Ojkić and Nagy, 2000). Based on this,
a deleted virus FAdV-9Δ4, lacking six open reading frames (ORFs), was generated in our
laboratory and has shown potential as a vaccine vector (Corredor and Nagy, 2010b; 2011;
Ojkić and Nagy, 2001). However, to better understand the molecular biology of FAdV-9,
research on exploring the functions of viral genes is required.
1
Literature review
1. Adenovirus taxonomy
Adenoviruses, first isolated as respiratory pathogens and characterized in 1953 (Rowe et
al., 1953), belong to the family Adenoviridae and are classified into five genera (Harrach
et al., 2011): Mastadenovirus, isolated from mammals; Aviadenovirus, isolated from
birds; Atadenovirus, isolated from mammals, birds and reptiles and named due to their
high adenine and thymine contents in the genomes; Siadenovirus, isolated from a frog
and birds and named because they encode a putative sialidase (Davison and Harrach,
2002); and Ichtadenovirus, isolated from a sturgeon (Berk, 2013; Harrach et al., 2011).
Specifically, each genus has one or more species (Fig. 1.1). To date, 56 human
adenovirus (HAdVs) serotypes have been identified and classified into 7 species, Human
adenovirus A to Human adenovirus G, based on serology, genome sequencing,
phylogenetic distance, hemagglutination, and oncogenicity in rodents (Berk, 2013). Other
species in the genus Mastadenovirus include Murine adenovirus A to Murine adenovirus
C, Ovine adenovirus A, Ovine adenovirus B, Porcine adenovirus A to Porcine adenovirus
C, Simian adenovirus A, and Tree shrew adenovirus A (Harrach et al., 2011).
There are 12 fowl adenovirus (FAdV) serotypes. They are grouped into 5 species, Fowl
adenovirus A to Fowl adenovirus E, on the basis of phylogeny, genome organization,
restriction fragment length polymorphism (RFLP) profiles and the lack of significant
cross-neutralization (Zsak and Kisary, 1984; Harrach et al., 2011). Other species in the
genus Aviadenovirus are Falcon adenovirus A, Goose adenovirus A, and Turkey
adenovirus B. Recently, several potential species have been proposed to belong to the
2
Figure 1. 1 Phylogenetic tree of adenovirus genomes.
Adenoviruses, belonging to family Adenoviridae, are classified into five genera,
Mastadenovirus, Aviadenovirus, Atadenovirus, Siadenovirus, and Ichtadenovirus. Each
genus has one or more species. Taken from Harrach et al., 2011.
3
genus Aviadenovirus but have not been approved yet, including Pigeon aviadenovirus A,
Duck aviadenovirus B, Turkey adenovirus C and Turkey adenovirus D (Marek et al.,
2014a, 2014b).
Atadenovirus includes Ovine adenovirus D, Duck adenovirus A (egg drop syndrome virus,
EDSV), Bovine adenovirus D, Possum adenovirus and Snake adenovirus A. Viruses in
the genus Siadenovirus have a putative sialidase gene present at the left end of the
genome and include three species: Frog adenovirus (Davison et al., 2000), Raptor
adenovirus A and Turkey adenovirus A. Ichtadenovirus has been newly established and
contains one single speices Sturgeon adenovirus A (Doszpoly et al., 2009), which has
been found in white sturgeon (Benkö et al., 2002).
2. Virion structure
Adenoviruses are non-enveloped, icosahedral particles with a diameter from 70 to 100
nm. Each virion consists of a protein shell called a capsid that surrounds a viral core
containing the DNA genome. The capsid has two roles: it protects the core genome and
mediates the entry of the virus into cells. There are at least 13 proteins in the adenovirus
virion (Fig. 1.2, Table 1.1), including II-IX, IIIa, Iva2, terminal protein (TP), μ, and p23,
the viral protease (Russell, 2009). They are all structural proteins encoded by late regions
L1 to L5 of the viral DNA genome. Historically, these proteins were numbered (II–IX) in
order due to their molecular weight (smallest to greatest) when they were first identified
(Vellinga et al., 2005).
The capsid is composed of 252 subunits called capsomeres, which are collectively made
up of 240 hexons and 12 pentons. Each capsid consists of at least 7 proteins, three of
4
Figure 1. 2 Structure of Adenovirus
Taken from Russell, 2009
5
Table 1.1 Summary of 13 adenoviral proteins and their functions
Name
Location
Known Functions
II
Hexon monomer
Structural
III
Penton base
Penetration
IIIa
Associated with penton base
Penetration
IV
Fiber
Receptor binding; haemagglutination
V
Core: associated with DNA and
penton base
Histone-like; packaging?
VI
Hexon minor polypeptide
Stabilization/assembly of particle?
VII
Core
Histone-like
VIII
Hexon minor polypeptide
Stabilization/assembly of particle?
IX
Hexon minor polypeptide
Stabilization/assembly of particle?
TP
Genome - Terminal Protein
Genome replication
μ
Nucleoprotein
Genome replication?
IVa2
Nucleoprotein
Genome packaging
Protease
Associated with pentons?
Maturation
Taken from Russell, 2009
6
which are termed major proteins, hexon (protein II), penton base (pIII) and fiber (pIV);
and the other four are termed minor capsid proteins, IIIa, VI, VIII and IX (Vellinga et al.,
2005). The hexon protein comprises the majority of the outer shell of the capsid, which is
made up of 240 homotrimers that encapsidate most parts of the virus, including the viral
genome and associated proteins. The trimeric fiber projects from each of the 12 vertices
of the icosahedron, while the penton base lies at the base of each fiber. These three capsid
proteins contribute to the majority of activities required for the early stages of Ad
infection (Russell, 2009; Medina-Kauwe, 2013; Berk, 2013).
The remaining six structural components are situated in the virus core, five of which are
associated with the double-stranded DNA genome (V, VII, μ, IVa2 and TP), while the 23
K virion protease plays a vital role in the assembly of the virion. The TP is covalently
attached to the 5’ ends of the adenoviral DNA, circularising it. Newly synthesized TP
acts as a primer for DNA replication in the cell nucleus (Berk, 2013; Russell, 2009).
3. Genome organization of mastadenoviruses and early gene functions
Adenoviruses have a linear, double-stranded DNA (dsDNA) genome with a size range of
26-48 kb (Harrach et al., 2011). At the 5’ end of each strand, a terminal protein is
covalently linked which functions as a protein primer at the onset of viral DNA
replication. The genomes of all adenoviruses have inverted terminal repeat (ITR)
sequences, ranging in size from 36 to over 200 base pairs (bp), which function as origins
of DNA replication at each end of the viral genome. DNA sequences, which encode most
virion structural proteins and proteins involved in viral DNA replication and virion
assembly, are well conserved in all adenovirus genomes. In contrast, DNA sequences
encoding regulatory proteins, usually located at both ends of the genome, are not
7
conserved among the genera (Berk, 2013).
According to the time of expression during the viral life cycle, human adenovirus
genomes contain 5 early transcription units (E1A, E1B, E2, E3 and E4), 4 intermediate
transcription units at the onset of viral DNA replication (IX, IVa2, L4 intermediate, and
E2 late), and one late transcription unit (major late) that is processed to generate 5
families of late mRNAs (L1-L5) (Fig. 1.3, Table 1.2). All of these transcripts are
transcribed by RNA polymerase II. In addition, the human adenovirus genome also
carries one or two small virus-associated RNA (VA-RNA), which varies in the serotypes.
The VA-RNA is transcribed by RNA polymerase III. The early genes are involved in
regulation and virus replication (E1); viral DNA replication (E2); escape from the
immune system (E3) and cell cycle control (E4). The late genes are transcribed later than
the early genes and encode the structural proteins described above.
3.1 The E1A region
During adenovirus infection, E1A is the first viral transcription unit to be transcribed
once the viral DNA reaches the nucleus. It is essential for adenovirus mediated cell
transformation (Mymryk et al., 1994; Gallimore and Turnell, 2001; Zhang et al., 2004).
E1A has been extensively studied in adenovirus serotype 5, and most of the literature
relate to this serotype. The E1A gene encodes, using alternative splicing, two mRNAs
during the early phase of the infection and three other mRNAs (11S, 10S and 9S) during
the late phase. The two early E1A proteins are called 13S and 12S, which are composed
of 298 and 243 amino acids (aa), respectively. The 13S has a 46 aa internal region that is
not found in the 12S form. Based on sequence comparison of different serotypes, these
two proteins both contain two conserved regions, CR1 and CR2, while 13S contains the
8
CR3 internal 46 aa region. The various functions of E1A are mediated mainly through
interactions with many key cellular proteins via these conserved regions.
The CR1 conserved region induces host cell cycling through interaction with many
chromatin-remodelling proteins for gene activation and repression. E1A interacts with
p300 and its paralog CREB-binding protein (CBP), which are co-activators for
transcription. It has been reported that E1A can inhibit the histone-acetyltransferase
(HAT) activity of p300 (Chakravarti et al., 1999; Hamamori et al., 1999) and stimulate
transcription through p300/CBP. Other E1A-CR1 interacting HAT complexes include
PCAF (Lang and Hearing, 2003), p400, a member of the SWI/SNF family (Fuchs et al.,
2001), and TRAPP, a component of HAT complexes (Deleu et al., 2001; Nikiforov et al.,
2002). However, the exact mechanism for how chromosome remodelling induces cell
cycling by E1A is still unclear. E1A CR3 region is involved in the activation of other
early viral transcription units. Its transcription activation function is dependent on the
interaction of E1A-CR3 with MED23 (formally called SUR-2), a component of the
Mediator complex (Ablack et al., 2010).
Cell transformation induced by expression of E1A alone is unstable. Creation of stable
transformed cells requires both E1A and E1B products. E1A induces the activation of
p53 and thus E1A-transformed cells die rapidly from apoptosis (Debbas and White, 1993;
Grand et al., 1994). It has been demonstrated that the function of both E1B products in
transformation is to block p53-activated apoptosis in order to establish stably transformed
cells (Grand et al., 1994).
3.2 The E1B region
The E1B transcription unit encodes five products by alternative splicing. The two major
9
polypeptides E1B19K and E1B55K are transcribed using two different ORFs. In addition,
Figure 1.3 Transcription map of the mastadenovirus genome.
The mastadenovirus genome is organized in early (red), intermediate (blue) and late
(green) transcriptional units.
Taken from Weitzman, 2005.
10
Table 1.2 Adenovirus gene products and their functions
Phase
Gene
E1A
E1B
E2A
Early
E2B
E3
E4
IVa2
IX
Intermediate
VAI
VAII
Late
L1-L5
Products and their function
Inactivates pRB to release E2F - cell cycle deregulation
Transactivates viral promoters
55K targets p53 and participate in transport of late viral
mRNA
19K is a Bcl-2 homologue - anti-apoptotic
Preterminal protein (pTP) and DNA polymerase (pol) - DNA
replication
Single-strand DNA binding protein (DBP) - DNA replication
gp19K inhibits MHC I expression
10.4K/14.5K (RID complex) inhibits tumor necrosis factor
(TNF) apoptosis, internalizes TNF receptor and degrades Fas
ligand
14.7K inhibits TNF apoptosis, stabilized NFκB
11.6K (ADP) induces cell lysis
Products: orf1, orf2, orf3, orf4, orf6 and orf6/7 modulate viral
mRNA metabolism, promote virus DNA replication and block
host protein synthesis
Initiate the major late promoter (MLP), which regulates late
genes
Non-coding RNA that stimulates translation of viral genes
and blocks double stranded RNA activated protein kinase R
(PKR) during interferon response
Non-coding RNA that blocks PKR during interferon response
Structural proteins: L1 (IIIa); L2 (penton base, V, VII); L3
(hexon, VI, virus protease); L4 (VIII); L5 (fiber)
Taken from Weitzman, 2005.
11
three other minor products E1B-156R, E1B-93R and E1B-84R are transcribed from the
ORF for E1B55K, (Virtanen and Pettersson, 1985). It is well known that the major
function of the E1B proteins, at least when functioning on their own, is to counteract
E1A-induced apoptosis in order to prevent premature lysis of the host cell, thus ensuring
an efficient viral replication (Berk, 2013).
E1B19K protein prevents p53 independent apoptosis. Upon activation of p53, the
proapoptotic BCL-2 family member BAK forms oligomers with another proapoptotic
protein BAX to form pore-like structures in the mitochondrial membrane. This process
leads to the release of cytochrome C and Smac/DIABOLO which induce the caspasedependent pathway, resulting in apoptosis. This effect can be suppressed by high levels
of expression of BCL-2 in uninfected cells. E1B19K protein is a viral mimic of BCL-2,
which sequesters BAK, thus preventing its association with BAX to prevent apoptosis.
E1B19K alone is sufficient to support E1A-mediated cell transformation; however,
E1B55K is also required for viral replication.
E1B55K is involved in many functions during both early and late phases of
mastadenovirus infection, most of which are dependent on the formation and the activity
of the E4orf6/E1B55K E3 ligase complex. The most studied and key early function of
E1B55K is to counteract p53-induced cell cycle arrest and early apoptosis which is
caused by the interaction of E1A protein interaction with either retinoblastoma protein
(pRB) family proteins or p300/CBP/p400, resulting in activation and accumulation of p53
(Chiou and White, 1997). The p53 is a well-known tumour suppressor that functions
mainly as a sequence specific transcription factor, activating or repressing transcription
and leading to ultimate cell cycle arrest or apoptosis (Woods and Vousden, 2001).
12
E1B55K antagonises the effect of p53 by at least two separate mechanisms: proteasomal
degradation of p53 by the E4orf6/E1B55K E3 ligase and E1B55K-mediated inhibition of
p53 transcriptional activity.
The localization of E1B55K is largely dependent on two other adenoviral proteins,
E4orf3 and E4orf6. In the context of expression of E1B55K alone, HAdV-5 E1B55K
adopts a predominantly cytoplasmic localization, exhibiting in immunofluorescence
studies bright staining of perinuclear structures generally believed to be aggresomes
(Goodrum et al., 1996). Aggresomes are perinuclear bodies formed at the microtubule
organizing center (MTOC) in response to accumulation of misfolded and aggregated
proteins. E4orf6 is localized in the nucleus when expressed alone. When E1B55K is coexpressed with E4orf6, E1B55K co-localizes with E4orf6 in the nucleus and the
aggresome structures remain unaffected (Goodrum et al., 1996). E4orf3 can also bind to
E1B55K. E4orf3 has been found to modify promyelocytic leukemia (PML) nuclear
bodies (NBs) into track-like structures in the nucleus. Co-expression of these two
proteins results in the translocation of E1B55K into the PML NBs (Leppard and Everett,
1999). The interactions of E1B55K with E4orf6 and E4orf3 are mutually exclusive but
when these three proteins are co-expressed, E1B55K co-localizes with E4orf6 (Konig et
al., 1999). In the context of infection, E1B55K is re-localized by E4orf6 mainly to
nuclear viral replication centers where DNA replication and subsequent transcription take
place. E1B55K is also found in the nuclear filamentous structures of PML and the
aggresomes (Ornelles and Shenk, 1991). E4orf6 and E1B55K nuclear co-localization
requires a primate-specific factor, RUNX1 (Marshall et al., 2008).
Apart from the direct effects described above on p53, E1B55K is believed to inactivate
13
p53 by sequestration to the perinuclear aggresome structures (Zantema et al., 1985).
During infection, as E4orf3 is believed to be expressed earlier than E4orf6, E1B55K is
first observed to associate with PML alone with p53 (Konig et al., 1999).
3.3 E2 proteins
The E2 transcription unit encodes three proteins, all of which are directly involved in
DNA replication. From the same promoter, two different transcripts, E2A and E2B, are
produced by alternative splicing. E2A transcripts encode a DNA binding protein (DBP)
and E2B encodes both the adenovirus DNA polymerase (pol) and the precursor terminal
protein (pTP). E2A expression is mediated by the E1A region. Viral DNA replication
requires all three E2 proteins, plus two cellular proteins, nuclear factor I (NF-I) and
nuclear factor III (NF-III) (Challberg and Kelly, 1981).
DBP is the first non-structural adenoviral protein identified due to its abundance in
infected cells (van der Vliet and Levine, 1973). It is a 72 kDa protein that binds to singlestrand DNA (ssDNA) with high affinity and to dsDNA and RNA with lower affinity. It
has many roles in both the initiation and elongation steps of viral DNA replication,
especially to unwind DNA templates. Ad pol is a 140 kDa viral DNA polymerase, a
major component of the initiation complex of viral DNA replication. Unlike cellular
polymerase, it possesses 5’ to 3’ polymerase activity as well as the 3’ to 5’ exonuclease
activity. The pTP is initially produced as an 80 kDa protein, which is responsible for
protein priming for the initiation of DNA replication. At the late phase of infection, pTP
is processed by proteolysis into a smaller fragment of 55 kDa during the assembly of
virions involving the adenovirus protease. TP acts as a primer for viral DNA replication
and hence ends up being covalently attached to the 5’ termini of each end of the viral
14
DNA (Challberg and Kelly, 1981).
3.4 The E3 region
The E3 region, which is dispensable for the replication of adenovirus, encodes a range of
proteins that subvert the host immune response. The E3 promoter contains nuclear factor
kappa B (NF-κB) -binding sites that can be induced by cytokines such as tumour necrosis
factors (TNFs) (Deryckere et al., 1995).
The adenovirus E3-19K glycoprotein (gp19k) is localized in the membrane of the
endoplasmic reticulum (ER), where it binds to major histocompatability complex (MHC)
class I antigens, preventing their export from the ER to the cell surface. It is the presence
of a lysine based motif (KKXX) that is responsible for retention in the ER.
The E3-10.4K/14.5K complex (RID α/β) also inhibits TNF-α and Fas ligand-induced cell
death by internalising their receptors (TNFR1, Fas and TNF-related apoptosis-inducing
ligand) and promoting their degradation in lysosomes (Elsing and Burgert, 1998;
Tollefson et al., 1998). RID α/β localize to the plasma membrane, Golgi, ER and vesicles
within the cell (Tollefson et al., 1998).
The E3-14.7K protein inhibits apoptosis through the TNFR, Fas and TRAIL pathways,
with the effect on Fas being less pronounced. The E3-14.7 kDa protein is localized in the
cytosol and nucleus (Li et al., 1999) and functions by binding to cellular proteins that
mediate apoptosis including NEMO/IKK-γ. Apoptosis induced by transfecting cells with
NEMO/IKK-γ, is reversed by 70% in the presence of E3-14.7K protein (Li et al., 1999).
The E3-11.6K, adenovirus death protein (ADP), is the only E3 protein not involved in
subverting the host cell immune response. It is produced late in infection and induces cell
15
death, resulting in the release of progeny virus from the cell.
3.4 E4 proteins
Transcripts from the E4 region are subject to alternate splicing events, leading to the
production of approximately 18 distinct mRNAs, which are predicted to encode 7
different proteins named Orf1, Orf2, Orf3, Orf4, Orf3/4, Orf6, Orf6/7 (Virtanen et al.,
1985). However, one of these polypeptides (E4orf3/4) has never been detected in infected
cell. The E4 region has a similar genomic organisation throughout all serotypes, except
that E4orf1 is not conserved in HAd40 (Davison et al., 2003).
The E4 region is required for efficient viral DNA replication, late gene expression and
shutoff of the synthesis of host cell proteins. This is confirmed by Halbert et al. (1985)
through generating mutant HAdV-5 viruses carrying defined lesions in the E4 region. A
mutant virus dl366, lacking the majority of the E4 region, is severely defective and can
only be propagated in stable Vero cells expressing the E4 region (Halbert et al., 1985).
Another mutant virus dl355, lacking 14 bp within the segment encoding the E4Orf6
protein, shows a delayed onset of viral DNA synthesis. Expression of late viral proteins is
reduced in both dl355 and dl366 viruses, although more severely in dl366. Shutoff of the
host protein synthesis is also less efficient with the mutant viruses (Halbert et al., 1985).
3.5 Intermediate and late genes
After the onset of virus DNA replication, the intermediate transcription units IVa2 and IX
genes are expressed at high levels, activating transcription of the MLP and resulting in
expression of the late genes (Lutz et al., 1997; Tribouley et al., 1994).
The major late transcription unit (MLTU) encodes multiple proteins from five regions,
16
L1 to L5, by differential splicing and polyadenylation. It is the primary transcript that
encodes most of the structural proteins and some non-structural proteins needed for the
assembly of progeny virions. All late mRNAs contain a 5’ tripartite leader sequence
(TPL), which is important for their nuclear export, translation and stability. The major
structural proteins produced by MLTU include L2-III (penton), L3-II (hexon) and L5-IV
(fiber), and non-structural proteins include L1-52/55K, L4-22K, L4-33K and L4-100K.
Expression from the MLTU is temporally regulated. During the early phase of infection,
the MLP is active at low levels and transcription does not proceed to the end of the
MLTU. This results in the production of one major product, the L1-52/55K protein. After
viral DNA replication, the MLP becomes fully activated and its activation depends on
many known activators including the E1A protein, the product of the intermediate
transcription unit IVa2 and transcription factors MAZ and Sp1. Viral DNA replication is
also required prior to MLP activation, suggesting that a cis-acting type of regulation is
involved (Berk, 2013).
4. Molecular biology of aviadenoviruses
4.1 Genome organization of aviadenoviruses
Compared to mastadenoviruses, the molecular biology of aviadenoviruses is less studied.
The genome size of aviadenoviruses is considerably larger compared to that of
mastadenoviruses, representing the largest adenovirus DNA after that of white sturgeon
adenovirus (Harrach et al., 2011). To date, the complete genome sequences have been
determined for members of all FAdV species: Fowl adenovirus A (FAdV-1, Chiocca et
al., 1996), Fowl adenovirus B (FAdV-5, Marek et al., 2013), Fowl adenovirus C (FAdV4, Griffin and Nagy, 2011; Marek et al., 2012), Fowl adenovirus D (FAdV-9, Ojkić and
17
Nagy, 2000) and Fowl adenovirus E (FAdV-8, Grgić et al., 2011). In addition to FAdVs,
the genomes of two other aviadenovirus species have been fully sequenced: Turkey
adenovirus B (TAdV-1, Kaján et al., 2010) and Goose adenovirus A (GoAdV-4, Kaján et
al., 2012). Recently, more complete genomes of several other avian adenoviruses have
become available, including pigeon adenovirus 1 (PiAdV-1), duck adenovirus 2 (DAdV2), TAdV-4 and TAdV-5 (Marek et al., 2014a, 2014b). They represent the potential
species Pigeon aviadenovirus A, Duck aviadenovirus B, Turkey adenovirus C and Turkey
adenovirus D, respectively. All of these are suggested to belong to the genus
Aviadenovirus but have not been approved yet by the international committee on
taxonomy of viruses (Marek et al., 2014a, 2014b).
The organization of the central part of the aviadenovirus genomes (from IVa2 to fiber
gene) is similar to that of mastadenoviruses, including E2 proteins, IVa2, fiber, hexon
and other late proteins. However, no regions homologous to the E1, E3 and E4 of
mastadenoviruses are recognized in aviadenovirus genomes. Despite the lack of genetic
similarity to other genera, the ends of the aviadenovirus genomes are termed E1 and E4
(Davison et al., 2003; Marek et al., 2013). In the left-end region (E1 region) of
aviadenovirus genomes, there are few variations in ORF constitution. The E4 region,
larger in aviadenoviruses than in members of other genera, contains several transcription
units that are unique for aviadenoviruses. ORFs 22, 20A, 20, 19, 8 (GAM-1, for Gallus
anti morte), and 17 of this region seem to be well conserved in all currently sequenced
aviadenoviruses except that ORF17 is absent in GoAdV-4 genome (Kaján et al., 2012).
On the other hand, the order and orientation of shared ORFs are also well conserved
among different FAdVs (Marek et al., 2013).
18
The fiber of FAdVs has been implicated to play an important role in the infectivity and
pathogenicity of FAdVs (Pallister et al., 1996; Grgić et al., 2014; Schachner et al., 2014).
Contrary to most mastadenoviruses, all FAdVs that have been examined have two fibers
protruding from each penton base (Gelderblom and Maichle-Lauppe, 1982). Two fiber
genes are identified in members of Fowl adenovirus A and Fowl adenovirus C (Chiocca
et al., 1996; Griffin and Nagy, 2011; Marek et al., 2012), while only a single fiber gene is
found in sequenced members of Fowl adenovirus B, Fowl adenovirus D and Fowl
adenovirus E (Grgić et al., 2011; Marek et al., 2013; Ojkić and Nagy, 2000). Although it
is still unknown how many fibers per penton base they possess, two fiber genes are
identified in the genomes of TAdV-1, TAdV-5, GoAdV-4 and PiAdV-1 (Kaján et al.,
2010, 2012; Marek et al., 2014a, 2014b), whereas one fiber gene was found in the
genome of TAdV-4 and DAdV-2 (Marek et al., 2014a, 2014b). Nevertheless, it is
predicted that, based on the phylogenetic analysis of aviadenoviruses, the ancestor strain
of all current aviadenoviruses probably had two fiber genes, of which fiber-1 was lost in
the branch leading to FAdV-B, FAdV-D, FAdV-E and TAdV-C (Marek et al., 2014a).
4.2 Functionally identified early genes in FAdV genome
At least one complete genome from members of all FAdV species has been sequenced,
and the FAdV genomes encode numerous genes. However, unlike the early genes of
human adenoviruses that are extensively defined, the majority of early genes of FAdVs
have not yet been characterized in detail. To date, there are only 3 early genes that have
been characterized beyond their sequences for FAdVs: ORF8 (GAM-1), ORF22, and
ORF1 (dUTPase). In addition, yet not characterized, some genes are found to show
homology with proteins of other viruses, for example, ORF2, a homolog of nonstructural
19
protein NS1 (Rep protein) of parvoviruses, and ORF19, a homolog of lipase of Marek's
disease virus (Corredor et al., 2006, 2008; Washietl and Eisenhaber, 2003).
ORF8
ORF8, also named GAM-1, is a conserved early gene located at the right-end of the
genomes of all FAdVs. Encoding a 31 kDa nuclear protein, GAM-1 was originally
discovered upon screening for viral proteins that regulate cellular apoptosis (Chiocca et
al., 1997). Though sharing no sequence homology, it mimics the functions of Bcl-2 and
E1B-19K of human adenoviruses and blocks apoptosis (Chiocca et al., 1997).
Glotzer et al. (2000) have demonstrated that GAM-1 of FAdV-1 is required for virus
replication. They showed that the GAM-1 protein increases the cellular levels of heatshock proteins 40 (hsp40) and hsp70 and relocates these proteins to the nucleus. GAM-1
activates host heat-shock responses with hsp40. Heat shock or overexpression of hsp40
can partially replace the roles of GAM-1 in viral replication.
GAM-1 influences the expression of cellular genes. Chiocca and co-workers (2002) have
shown that expression of GAM-1 increases the level of transcription from a variety of
eukaryotic promoters, probably by effectively inactivating histone deacetylase 1 (HDAC1)
both in vitro and in vivo. They also demonstrated that a FAdV-1 lacking GAM-1 is
replication defective, but the defect can be overcome by either expressing an interfering
HDAC1 mutant or by treating infected cells with HDAC inhibitor trichostatin A (TSA).
Further evidence showed that GAM-1 interferes with SUMOylation of HDAC1, through
destroying promyelocytic leukemia nuclear bodies and delocalizing SUMO-1 into the
cytoplasm (Colombo et al., 2002). The SUMOylation inhibitory function of the purified
GAM-1 protein has been recently confirmed in an in vitro assay (Avila et al., 2015).
20
Hacker et al. (2005) also demonstrated that the FAdV-1 GAM-1 protein enhances
transient and stable recombinant protein expression in Chinese hamster ovary cells.
Everett et al. (2014) showed that GAM-1 increases the infection efficiency of a
regulatory protein ICP0-deficient mutant of herpes simplex virus 1 (HSV-1).
ORF22
ORF22 of FAdV-1 is identified as an early transcription product that is detectable from 2
hours post-infection (h.p.i.) (Payet et al., 1998). The ORF22 protein is expressed as early
as 6 h p.i. and is still accumulating at 30 h.p.i. (Lehrmann and Cotten, 1999). In terms of
its function, ORF22 is found to interact with the pRb to activate the E2F pathway, in
cooperation with GAM-1 (Lehrmann and Cotten, 1999). ORF22 binds to the pocket
domain of pRb, similar to other DNA tumor virus proteins, while GAM-1 interacts with
pRb regions outside the pocket domain.
ORF1
Localized in the left end of FAdV genomes, ORF1 is a homolog of deoxyuridine 5′triphosphate pyrophosphatase (dUTPase). Corredor et al. (2006) showed that ORF1 is
well conserved in all genomes of FAdVs, with an amino acid identity ranging from 56 to
100% among FAdV genomes. Cao et al. (1998) showed that FAdV-9 ORF1 (formerly
named ORF LTR1) is transcribed at 2 h.p.i., while no data are available regarding its
protein expression. ORF1 of FAdV-1 has been demonstrated to possess dUTPase enzyma
activity (Weiss et al., 1997); however, its function in the virus life cycle is unknown.
dUTPase is a ubiquitous enzyme which exists widely in eukaryotic and prokaryotic cells,
viruses and some other biological organisms. This enzyme catalyzes the hydrolysis of
21
dUTP to dUMP and pyrophosphate (PPi), thereby preventing incorporation of uracil into
DNA and reducing dUTP/dTTP ratio in cells. To date, the dUTPase gene has been
characterized among a number of viruses, including herpesvirus (Fisher and Preston,
1986; Zhao et al., 2008), retrovirus (Payne and Elder, 2001), lentiviruses (Threadgill et
al., 1993), African swine fever virus (Oliveros et al., 1999), parapoxvirus (Cottone et al.,
2002), white spot syndrome virus (Liu and Yang, 2005), chlorella virus (Zhang et al.,
2005), Rana grylio virus (Zhao et al., 2007). According to a previous report (McGeoch,
1990), dUTPases from different species are classified into two types: class 1 and class 2,
based on the arrangements of five conserved motifs. The class 1 dUTPases have a chain
length of around 150 aa residues, and the arrangement of the motifs is 1-2-3-4-5, which is
known from bacteria, fungi, plants, metazoans, and a range of viruses. The class 2
dUTPases are about twice as long as those in class 1 with the 3-1-2-4-5 motifs, which are
found only in the alpha- and gamma- herpesviruses (McGeehan et al., 2001).
Based on the molecular assembly, dUTPases are grouped into three families: the
monomeric form of the enzyme encoded by mammalian and avian herpesviruses; the
dimeric form of dUTPase encoded by protozoan parasites and the bacterium; and the
trimeric form, which is the most studied family, discovered in eukaryotes, prokaryotes
and many viruses. The monomeric form is believed to have evolved via gene duplication
from a standard dUTPase-coding sequence of the trimeric form followed by a subsequent
loss of one copy of each motif from the double-length chain (McGeehan et al., 2001).
These two forms have similar enzymatic properties. In contrast, the dimeric enzymes
possess no similarity to members of other classes in sequence, structure or enzymatic
characteristics.
22
4.3 Study of FAdV-9 genome
FAdV-9 genome has been fully sequenced, with the size of 45,063 bp (Ojkić and Nagy,
2000). Transcriptional organization of FAdV-9 has been further analyzed as well (Ojkić
et al., 2002). The genome does not contain identifiable mammalian E1, E3, E4 regions, or
protein IX sequences, while sequences for E2 region, delayed-early protein IVa2, and late
proteins are relatively well conserved. There are two regions of tandemly repeated
sequences, TR-1 (nt 37,648–37,812) and TR-2 (nt 38,807–40,561), although their
functions are unknown. In previous studies in our laboratory, FAdV-9 has been used as a
recombinant virus vector to express the enhanced-green fluorescent protein (EGFP) with
the insertion of an EGFP gene in the TR-2, located at the right end of the genome, and its
growth in vitro has been characterized (Ojkić and Nagy, 2001). In addition, it has been
demonstrated that a 2.4 kb region at the left end of the FAdV-9 genome containing two
putative motifs of the packaging signal domain and six ORFs is dispensable for virus
replication in vitro. However, a mutant virus FAdV-9Δ4, which is devoid of 6 ORFs (0, 1,
1A, 1B, 1C and 2) (Fig. 1.4) but contains all packaging motifs, cannot replicate at wildtype level in vivo (Corredor and Nagy, 2010a). In another study of this group (Corredor
and Nagy, 2010b), an EGFP gene is inserted into the left end of genome of FAdV-9 at the
tenth codon of ORF1B and the genome of FAdV-9Δ4 to replace the deleted ORFs,
respectively. They showed that the left end region of the FAdV-9 genome is suitable as
an insertion/replacement site for foreign gene expression. All these data suggest the
importance of the left and right end genes of the FAdV-9 genome during their replication
cycle, vector design, and pathogenesis. However, further research about the function of
each ORF of FAdV-9 genome still needs to be conducted.
23
5. Disease associated with aviadenoviruses
Despite mild or no apparent clinical signs in birds, most members of Aviadenovirus have
been associated with a number of diseases including inclusion body hepatitis (IBH),
hydropericardium syndrome (HPS), gizzard erosions, proventriculitis and tenosynovitis
(Hess, 2013). Of these diseases, IBH and HPS are the most important ones that have been
reported from different parts of the world (Hess, 2013; Mittal et al., 2014).
IBH
IBH was first reported by Helmboldt and Frazier in 1963 in the United States (Helmboldt
and Frazier, 1963). IBH is prevalent in broiler chickens at 3-7 weeks of age, but case
studies have reported that it could be found in birds as young as 7 days old and as old as
20 weeks (Hess, 2013). IBH is characterized by sudden onset of mortality with a peak
after 3-4 days. It usually stops on day 5 but occasionally continues for 2-3 weeks.
Morbidity is low, and mortality may reach 10%. Outbreaks of IBH in chickens less than
3-weeks of age could cause the mortality to rise to 30%. Dar et al. (2012) reported that
the experimental infection of 2-day-old and 2-week-old chickens with FAdV-8b results in
83% and 43% mortalities, respectively. In the past several decades, IBH has been
reported in broiler flocks in numerous countries, including Canada (Nakamura et al.,
2011; Kim et al., 2008; Ojkić et al., 2008b; Philippe et al., 2005; Gomis et al., 2006;
Zadravec et al., 2013).
Almost all serotypes of FAdV have been reported to cause IBH in broiler chickens (Hess,
2013; Mittal et al., 2014). In Canada, the most commonly isolated serotypes are FAdVs-2,
-7, -8, and -11 (Ojkić et al., 2008b), of which FAdV-8 had become the dominant serotype
in Ontario since 2001 (Ojkić et al., 2008a).
24
Immunosuppressive virus agents such as infectious bursal disease virus (IBDV) and
chicken anemia virus (CAV) are thought to be involved in the outbreak of IBH (Fadly et
al., 1976; Hess, 2013). However, it is reported that IBH occurs in chickens before IBD is
present in New Zealand (Christensen and Saifuddin, 1989).
HPS
HPS is a emerged disease, which has caused huge economic losses to the poultry industry
in Pakistan since 1987 when it was first recognized. Compared to IBH, the mortality rate
and incidence of hydropericardium is higher for HPS, resulting in between 20% and 80%
mortality, though with low morbidity (Cowen, 1992; Hess, 2013). In addition to Pakistan,
the disease has been recognized in India, Middle East, Japan, Mexico, and South America
(Abe et al., 1998; Hess et al., 1999; Toro et al., 1999).
Unlike IBH that could be caused by any of 12 serotypes, HPS has been solely associated
with FAdV-4 from Fowl adenovirus C (Kim et al., 2008; Mittal et al., 2014). HPS
primarily occurs in broilers of 3-6 weeks of age, and it also occurs in breeding and layer
flocks as well, with lower mortality rates (Hess, 2013).
6. Diagnosis of fowl adenovirus infection
Diagnosis of FAdV infections can be carried out by the observation of gross and
histopathological changes in the liver, electron microscopic detection of viral particles,
various serological tests, such as enzyme-linked immunosorbent assay (ELISA), agar gel
precipitation test (AGPT), indirect hemagglutination, immunofluorescence, PCR, nested
PCR and real-time PCR (Hess, 2000, 2013; Ganesh et al., 2002; Philippe et al., 2007;
Romanova et al., 2009).
25
Several methods have been described for isolation and identification of FAdV. Cell lines
of chicken liver-origin such as primary chicken embryo liver (Celi) and hepatoma cell
lines (CH-SAH) can be used for the recovery of FAdV from specimens (Alexander et al.,
1998; Ojkić et al., 2008a). They produce typical CPE characterized by rounding,
syncytium formation (which resembles a bunch of grapes), and cell detachment. Negative
stain electron microscopy and thin-section electron microscopy have proven to be useful
for rapid detection of FAdV in suspected materials and tissues based on their
characteristic morphology (Hess, 2013). Immunocytochemistry can also be used to detect
the virus in infected cells by immunofluorescence staining with avian adenovirus
antiserum (Hess, 2013).
Several researchers have employed PCR alone (Jiang et al., 1999), or in combination with
RFLP of different regions of the hexon gene as a very sensitive and specific method for
detection, differentiation and phylogenetic analysis of FAdV (Meulemans et al., 2004;
Raue et al., 2005; Ojkić et al., 2008b). For these molecular methods, PCR primers are
frequently chosen in the region of the hexon gene and fiber gene because they have
hypervariable regions among serotypes (Ojkić et al., 2008b). Recently, a few research
groups developed a genotyping technique for FAdVs based on real-time PCR and highresolution melting-curve analysis (Marek et al., 2010; Steer et al., 2009, 2011), which is
demonstrated to be accurate, rapid, and robust for the identification of FAdV serotypes.
Serologic methods are also employed to identify FAdV infection. The AGPT is a widely
used serologic test for the detection of FAdV antibodies, since it is fast and economical.
However, its use is limited when detecting FAdV infections in SPF flocks, due to its lack
of sensitivity (Hess, 2013). ELISA for the detection of group-specific or type-specific
26
Figure 1.4 ORFs in the left end of FAdV-9 genome.
Taken from Corredor and Nagy, 2010a
27
antibodies to FAdVs has been described (Corredor and Nagy, 2011; Ojkić and Nagy,
2003; Junnu et al., 2014). Philippe et al. (2007) compared AGPT offered by diagnostic
laboratories that uses FAdV-1 as the antigen with a FAdV group-specific ELISA. They
demonstrated that ELISA is considerably more sensitive than the AGPT in early stages of
infection and shows good potential for practical application to monitor for the presence of
adenovirus antibodies in commercial flocks. Recently, Xie and co-workers (2013)
developed an ELISA for the diagnosis of FAdVs, based on two non-structural proteins,
100K and 33K of FAdV-1, and demonstrated that the 100K-33K-ELISA method is
sensitive, specific and can distinguish an acute FAdV infection from an inactivated virusbased vaccination response. In addition, it is reported that monoclonal antibody against
FAdV has been also used to improve diagnostic assays, study pathogenesis and identify
strains (Ahmad and Burgess, 2001).
7. Immune response to human adenovirus
7.1 Innate immune responses
Adenoviruses are highly immunogenic and elicit potent innate and adaptive immune
responses. Since adenoviral infections are common, the majority of adults have acquired
immunity, often against multiple Ad serotypes (Nayak and Herzog, 2009).
The innate response is comprised of various cells, including macrophages, dendritic cells
(DCs), neutrophils and natural killer (NK) cells as well as serum proteins such as
chemokines, cytokines and complement. Structural components of AdV, both DNA and
capsid proteins, play roles in triggering innate responses and the outcomes can vary
depending on Ad species and infected cell type (Liu and Muruve, 2003; Muruve, 2004).
28
Many of these pathways lead to activation of NF-κB and interferon regulatory factor 3
(IRF3), followed by production of inflammatory mediators and interferons (IFNs)
(Nociari et al., 2009; Randall and Goodbourn, 2008).
Adenoviral components are detected by pattern recognition receptors (PRRs) such as tolllike receptors (TLRs) (Nociari et al., 2009). In mammals, there are at least 13 TLRs with
unique specificities. Both TLR2, TLR9 have been implicated in AdV recognition
(Appledorn et al., 2008). TLR2 is a cell surface receptor that is known to bind
peptidoglycan and zymosan moieties present in bacteria. Studies have shown that TLR2
is a key mediator in responses to some dsDNA viruses as well; however, the exact ligand
has not been characterized. TLR9 is an endosomal receptor recognizing DNA with
unmethylated CpGs motifs (Appledorn et al., 2008; Compton et al., 2003).
The complement system is an important defense mechanism, which consists of plasma
proteins that are important in the defense against pathogens and induces rapid destruction
and phagocytosis of pathogens (Cichon et al., 2001). The main functions of the
complement components contain lysis of pathogens, opsonization, activation of
inflammatory response, and clearance of immune complexes (Kiang, et al., 2006).
In addition to complement activation, another early defense against pathogens is the
recognition of conserved microbial structures known as pathogen-associated molecular
patterns (PAMPs). Members of the TLR family are transmembrane proteins that
recognize PAMPs (Takeda and Akira, 2005). Adenoviral DNA is sensed by TLR9.
Through the adaptor protein myeloid differentiation factor 88 (MyD88), a signalling
cascade is initiated, whereupon production of type I interferons in plasmacytoid dendritic
cells (pDCs) is induced (Yamaguchi et al., 2007). Type I IFNs (IFN-α and IFN-β) as well
29
as the pro-inflammatory cytokines interleukin-6 (IL-6), IL-12 and tumor necrosis factor α
(TNF-α) are also secreted by other antigen presenting cells (APCs) such as myeloid DCs
(mDCs), macrophages and Kupffer cells. These cells recognize adenoviral DNA through
a TLR-independent pathway that is not fully characterized (Zhu et al., 2007a). One
proposed pathway is the induction of interferon regulatory factor 3 (IRF3) by double
stranded viral DNA that leads to transcription of type I IFNs (Nociari et al., 2007). The
IFNs activate a positive feedback loop, which results in maturation of the APCs (Nociari
et al., 2009). It is believed that DCs are activated by binding of the RGD motif on the
adenoviral penton base to αVβ integrins on DCs, which leads to TNF-α secretion and
maturation by autocrine TNF-α stimulation (Philpott et al., 2004).
Type I IFN represents one of the most important antiviral defense mechanisms (O'Neill
and Bowie, 2010). First of all, they promote innate immune responses by the activation of
NF-κB that induces production of pro-inflammatory cytokines and chemokines. Effector
cells like neutrophils, natural killer (NK) cells and monocytes are hence recruited to the
site of infection. Type I IFNs also induce adaptive immune response by promoting T and
B cell responses. How the adaptive immunity is regulated by type I IFNs is not
completely known but they do induce production of IL-15, which stimulates NK and T
cell proliferation (Bonjardim et al., 2009). In addition, type I IFNs upregulate IFN-γ
production in NK cells, which induces T helper 1 (Th1) response that will activate
cytotoxic T lymphocytes (CTLs) against virus-infected cells (Zhu et al., 2007b). B-cell
activation is characterized by IL-10, which peaks 72-96 hours after intravenous infusion
of a replication-selective adenovirus in cancer patients (Nemunaitis et al., 2001). Reid et
al. (2002) reported that patients who receive HAdV-5 systemically are positive for
30
anti-HAdV-5 antibodies after the first viral dose, which complicates re-administration.
7.2 Adaptive immune responses
Cell-mediated immune responses to HAdVs have been observed both for CD8+ cytotoxic
T cells (CTLs) and for CD4+ T helper cells. Multiple MHC class I and class II-restricted
epitopes have been mapped within the conserved region of the hexon (Leen et al., 2004,
2008). Only minimal cell-mediated immune responses to other capsid proteins, fiber or
penton base, are detected. CTLs kill infected cells by multiple mechanisms including
perforin, Fas-L and TNF-α, and thereby the life cycle of HAdV is disrupted before
progeny viruses are released (Leen et al., 2008). Specific CD4+ T cells play a critical role
in driving B-cell activation and differentiation. Activated B-cells undergo formation into
plasma cells, which produce antibodies directed against adenoviral epitopes located on
the major capsid proteins, i.e., hexon, penton and fiber (MacLennan et al., 1997). Binding
of antibodies induces effective clearance of viruses from the circulation and enhances the
interaction of AdVs with leukocytes through the Fcγ- and complement receptors (Spear et
al., 2001). Adenoviral infection generates serotype-specific neutralizing antibodies
(NAbs). Anti-hexon NAbs were initially proposed as the most important antibodies in the
neutralization process (Roberts et al., 2006). However, recent data indicate that NAbs
against fiber and penton proteins may also have great relevance (Myhre et al., 2007;
Sarkioja et al., 2008). It seems that antibodies directed against various Ad capsid
components synergize in the neutralization process (Gahery-Segard et al., 1998).
8. Immune responses to fowl adenoviruses
Despite an increasing interest from researchers, the immune response against fowl
adenovirus infection is less well studied, compared to that of human adenoviruses.
31
Recently, it has been demonstrated that FAdVs regulate a variety of cytokine gene
expression in infected birds. Grgić et al. (2013b) demonstrated that chickens
intramuscularly inoculated with FAdV-8 have significantly higher IFN-γ mRNA
expression and significantly lower IL-8 mRNA expression in spleen and liver, compared
to mock-infected chickens. FAdV-8 also upregulates, although not significantly, the
mRNA expression of IL-18 and IL-10 in spleen and liver. In another study by this group
(Grgić et al., 2013a), they showed that FAdV-4 induces significantly higher mRNA
expression of IFN-γ and IL-10, compared to the uninfected group, while no significant
difference was found in terms of the mRNA expression of IL-8 and IL-18.
Several studies have shown that FAdVs infection is involved in inducing
immunosuppression. For example, Shivachandra et al (2003) demonstrated that chickens
infected with serotype FAdV-4 have a significant decrease in the percentage of IgMproducing B cells in bursa of Fabricius. Schonewille et al. (2008) showed that the virulent
FAdV-4 infection results in a severe reduction in CD3+, CD4+ (helper T cells) and CD8+
cells (cytotoxic T cells) in spleen, and a decrease of CD4+ and CD8+ T-lymphocytes in
the thymus. More recently, Hussain et al. (2012) showed that broilers inoculated with a
FAdV-4 field strain have immune system dysfunction in terms of lymphoid organ
integrity, antibodies, and cell-mediated immune responses, as demonstrated by atrophy of
bursa of Fabricius, thymus, spleen, lower antibody titers against sheep red blood cells
(SRBC) and significantly reduced phytohemagglutinin-P (PHA-P), compared to the
uninfected group. Similarly, chickens orally-inoculated with a FAdV-1 field strain,
isolated from field outbreaks of IBH, have a significant decrease in the antibody response
to Brucella abortus (T-cell-independent antigen) and a significant decrease in
32
blastogenesis response of peripheral blood lymphocytes to PHA-P (Singh, et al., 2006).
These data indicate that suppression of the antibody and cell-mediated immune responses
might be a universal phenomenon of virulent fowl adenoviruses.
Fowl adenoviruses elicit a strong humoral response in the hosts. FAdV-specific Abs are
commonly detected in breeder and layer chickens (Hess, 2013). Following infection,
birds rapidly developed detectable NAbs (type-specific) at 1 week p.i., and the peak of
the Ab level is around 3 week p.i. (Ojkić and Nagy, 2003; Hess, 2013). Kim et al. (2014)
demonstrated that chickens vaccinated with an inactivated FAdV-4 still remain high Ab
level at 7 week p.i.. However, it is reported that the Ab response to FAdV-8 appears at 2
week p.i., followed by a slight decrease at 3 week p.i., and increases again at 4 week p.i.
(Grgić et al., 2011). Corredor and Nagy (2010a) showed that the Ab response to FAdV-9
(A-2A) appears at 1 week p.i. and continues to increase until the end of the experiment at
4 week p.i.. Schachner et al. (2014) reported that 1-day-old chickens intramuscularly
injected with fiber 2 of FAdV-4 induced anti-fiber 2 Abs which peak at the age of 4
weeks and protect chicken from the virulent FAdV-4 challenge.
Several studies have demonstrated that the host Ab response depends on the dosage of the
inoculum and the route of inoculation. For example, Ojkić and Nagy (2003) showed that
the chickens intramuscularly inoculated with a higher dose of FAdV-9 have significantly
higher Ab level than those inoculated with a lower dose from 1 to 4 weeks p.i.. They also
demonstrated that intramuscularly inoculated chickens have significantly higher Ab level
than chickens that receive virus through water or feed. Similarly, Grgić et al. (2011)
showed that the level of Ab response to FAdV-8 is significantly higher in intramuscularly
inoculated chickens than in orally inoculated chickens.
33
Hypotheses and Objectives
Hypotheses
1. FAdV-9 ORF1 is a genuine dUTPase, which plays important roles in virus replication
in vitro and in vivo.
2. FAdV-9 ORF1 modulates the host immune response against the virus infection.
Objectives
1. To explore the effects of the left-end genes on virus replication and modulation of
immune response on orally inoculated chickens.
2. To generate ORF1 knockout virus and HA-tagged ORF1 recombinant FAdV-9 virus.
3. To characterize the molecular features of FAdV-9 ORF1.
4. To explore the roles of FAdV-9 ORF1 on virus replication and modulation of the host
response in vitro and in vivo, through comparing the ORF1 knockout virus and wild
type virus.
34
Chapter 2. Oral inoculation of chickens with a candidate fowl adenovirus 9 vector
Li Deng, Shayan Sharif, Éva Nagy*
Department of Pathobiology, Ontario Veterinary College, University of Guelph, Guelph,
Ontario, Canada
*Corresponding author: Éva Nagy
Tel.: +1-(519) 824-4120 Ext. 54783
Fax: +1-(519) 824-5930
E-mail address: [email protected]
Mail address: Department of Pathobiology, Ontario Veterinary College, University of
Guelph, 50 Stone Road East, Guelph, ON, Canada, N1G 2W1
Author’s contributions
LD designed and performed all experiments, conducted the data analysis and wrote the
first draft of the manuscript. SS and ÉN provided guidance during the experiments and
critical review for the manuscript.
Published in Clinical and Vaccine Immunology. 2013. 20(8):1189-1196.
35
Abstract
Fowl adenoviruses (FAdVs) are a potential alternative to human adenovirus-based
vaccine vectors. Our previous studies demonstrated that a 2.4-kb region at the left end of
the FAdV-9 genome is nonessential for virus replication and is suitable for the insertion
or replacement of transgenes. Our in vivo study showed that the virus FAdV-9Δ4, lacking
six open reading frames (ORFs) at the left end of its genome, replicates less efficiently
than wild-type FAdV-9 (wtFAdV-9) in chickens that were infected intramuscularly.
However, the fecal-oral route is the natural route of FAdV infection, and the oral
administration of a vaccine confers some advantages compared to administration through
other routes, especially when developing an adenovirus as a vaccine vector. Therefore,
we sought to investigate the effects of FAdV-9 in orally inoculated chickens. In the
present study, we orally inoculated specific-pathogen-free (SPF) chickens with FAdV-9
and FAdV-9Δ4 and assessed virus shedding, antibody response, and viral genome copy
number and cytokine gene expression in tissues. Our data showed that FAdV-9Δ4
replicated less efficiently than did wtFAdV-9, as evidenced by reduced virus shedding in
feces, lower viral genome copy number in tissues, and lower antibody response, which
are consistent with the results of the intramuscular route of immunization. Furthermore,
we found that both wtFAdV-9 and FAdV-9Δ4 upregulated the mRNA expression of
alpha interferon (IFN-α), IFN-γ, and interleukin-12 (IL-12). In addition, there was a trend
toward downregulation of IL-10 gene expression caused by both viruses. These findings
indicate that one or more of the six deleted ORFs contribute to modulating the host
response against virus infection as well as virus replication in vivo.
36
Key words
fowl adenovirus, oral inoculation, virus replication, host response
37
Introduction
Fowl adenoviruses (FAdVs), of the genus Aviadenovirus and the family Adenoviridae
(Harrach et al., 2011), have a worldwide distribution and can be isolated from both sick
and healthy birds (Adair and Fitzgerald, 2008). Infection with pathogenic FAdVs can
lead to inclusion body hepatitis (IBH) in broiler chickens, causing very significant losses
to the poultry industry worldwide, including in Canada (Dar et al., 2012; Ojkić et al.,
2008). FAdVs are transmitted horizontally and vertically, can cause persistent infections,
and are excreted through feces and the respiratory tract (Adair and Fitzgerald, 2008;
Grgić et al., 2006).
To date, the genomes of four fowl adenoviruses (those of FAdV-1, FAdV-9, FAdV-8,
and FAdV-4) have been fully sequenced (Chiocca et al., 1996; Grgić et al., 2011; Griffin
and Nagy, 2011; Ojkić and Nagy, 2000), and they are about 10 kb larger than those of
mastadenoviruses.
Human adenoviruses (HAdVs) and other mammalian adenoviruses are used both as
oncolytic viruses (Cody and Douglas, 2009; Gallo et al., 2005; Shashkova et al., 2005)
and vaccine vectors (Lasaro and Ertl, 2009; Sharma et al., 2010). FAdVs are also suitable
vectors; for example, FAdV-1- and FAdV-8-based recombinant viruses have induced
protective immune responses against infectious bursal disease virus and infectious
bronchitis virus, respectively (Francois et al., 2004; Johnson et al., 2003).
The nonpathogenic FAdV-9 has also been developed as a virus vector. We demonstrated
that the tandem repeat region 2 (TR-2) at the right end of the genome is dispensable and
is suitable for foreign gene insertion (Ojkić and Nagy, 2001). More recently, a 2.4-kb
38
region at the left end of the FAdV-9 genome, containing two putative motifs of the
packaging signal domain and six open reading frames (ORFs), was shown to be
nonessential for virus replication in vitro. However, a deletion virus (FAdV-9Δ4) that
lacks the six ORFs (0, 1, 1A, 1B, 1C, and 2) replicated less efficiently than the wild-type
(unmodified) FAdV-9 (wtFAdV-9) in chickens inoculated intramuscularly, and the
antibody (Ab) level was lower in the FAdV-9Δ4-inoculated birds (Corredor and Nagy,
2010a). We have also demonstrated that the left end of the FAdV-9 genome is a suitable
site for the insertion and replacement of foreign genes (Corredor and Nagy, 2010b).
Moreover, in chickens immunized with a recombinant virus containing the enhanced
green fluorescence protein (EGFP) gene, antibodies were detected against the foreign
protein (Corredor and Nagy, 2011). All these studies suggest the importance of the leftend genes of the FAdV-9 genome in virus replication, immune response modulation, and
vector design. Moreover, the optimization of delivery routes and regimens is important
for overcoming the potential limitations of AdV-based vaccines for both human and
animal applications (Thacker et al., 2009).
FAdVs are normally transmitted by the fecal-oral route, so we wanted to learn more
about the replication of our vector virus and its effect on the chicken immune system
after oral administration. Therefore, the aims of the present work were to study virus
replication and host response in chickens that were inoculated orally with an FAdV
vector virus (FAdV-9Δ4) and a wild-type virus (wtFAdV-9). Specifically, virus shedding
in feces, viral genome copy number in tissues, antibody response, and expression levels
in tissues of selected cytokine genes, alpha interferon (IFN-α), IFN-γ, interleukin-10 (IL10), and IL-12 were determined.
39
Materials and Methods
Viruses and cells
FAdV-9 (strain A-2A) and FAdV-9Δ4 were propagated and titrated in chicken hepatoma
cells (CH-SAH) as described previously (Alexander et al., 1998). The cells were
maintained in Dulbecco's modified Eagle's medium and nutrient mixture Ham's F-12
medium (DMEM-F12) supplemented with 10% non-heat-inactivated fetal bovine serum
(FBS), 2 mM l-glutamine, 100 U/ml penicillin, and 100 μg/ml streptomycin.
Animal experiment
The experiment was reviewed and approved by 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. One hundred thirty-five 1-day-old
specific-pathogen-free (SPF) White Leghorn chickens were obtained from the Canadian
Food Inspection Agency (CFIA) (Ottawa, Canada) and were housed in the isolation unit
of the University of Guelph. At 7 days of age, the chickens were wing tagged and
randomly divided into three groups (groups I, II, and III). On day 10, the chickens were
inoculated orally with 1.5 × 107 PFU/chick with wtFAdV-9 (group I), FAdV-9Δ4 (group
II), or PBS (group III). The chickens were observed daily for clinical signs of infection.
To detect virus in the feces, cloacal swabs were collected in 1 ml PBS with antibiotics at
0, 1, 3, 5, 7, 10, 14, 21, and 28 days postinoculation (d.p.i.) and were stored at −80°C
until processing. Sample preparation and virus titration were performed as described
previously (Corredor and Nagy, 2010a). A sample was regarded as negative if it tested
negative at least twice and at two different times. Blood samples were collected from all
chickens on 0, 7, 14, 21, and 28 d.p.i., and sera were tested for antibodies by
40
enzyme-linked immunosorbent assay (ELISA) using purified FAdV-9 as an antigen and
following the method described previously (Ojkić and Nagy, 2003). Five chickens from
each group were randomly drawn for euthanasia and necropsy on 1, 2, 3, 4, 5, 7, 14, 21,
and 28 d.p.i.. Liver, cecal tonsil, spleen, and bursa of Fabricius samples were collected
and sectioned into two portions: one was placed in a sample bag and stored at -80°C for
viral genome copy number determination, as described previously (Romanova et al.,
2009), and the other was collected in a 1.5-ml sterile Eppendorf tube containing
RNAlater (Invitrogen Canada, Inc., Burlington, Ontario, Canada) and stored at -80°C.
The expression levels of the IFN-α, IFN-γ, IL-10, and IL-12 p40 cytokine genes were
evaluated by real-time quantitative PCR (RT-qPCR), with β-actin as a reference gene, as
described previously (Abdul-Careem et al., 2006, 2007; Grgić et al., 2013).
Statistical analysis
Statistical analyses were performed using GraphPad Prism 5.0 software (San Diego, CA).
A one-way analysis of variance (ANOVA) was used to determine significant differences
between the groups. The critical level for significance was set at a P value of <0.05. The
data were expressed as mean ± standard error of the mean (SEM), determined from five
individual birds at the designated days.
Results
Throughout the experiment, no clinical signs of infection were seen in any groups of
chickens, and there were no pathological lesions at necropsy.
41
Virus shedding
Virus titers in cloacal swabs were determined by the plaque assay. No virus was detected
in any groups of chickens before inoculation and in the mock-infected group throughout
the study. The virus titers in groups inoculated with wtFAdV-9 and FAdV-9Δ4 are shown
in Table 2.1. For FAdV-9Δ4, virus was detected only at days 1 and 7 p.i., and the titers
were significantly lower than those of wtFAdV-9-infected chickens. In the wtFAdV-9infected group, virus was detected with high titers at 1 to 14 d.p.i., but virus was not
detected at the later days (21 and 28 d.p.i.). The highest titer appeared at 5 d.p.i. with 4.0
×103 PFU/ml.
Viral genome copy number in tissues
Viral genome copy numbers in liver, cecal tonsil, bursa of Fabricius, and spleen samples
were determined by quantitative PCR (qPCR), and the results are summarized in Table
2.2. No viral DNA was detected in the mock-infected chickens. Throughout the study,
viral DNA was detected in cecal tonsil and spleen samples from both virus-infected
groups from 1 d.p.i. until 21 d.p.i. and also in liver samples collected at 1, 3, 5, and 7 d.p.i.
At day 14 p.i., 40% and 60% of the liver samples had detectable virus levels for the
FAdV-9Δ4 and wtFAdV-9 groups, respectively. At 21 d.p.i., only 20% of the samples,
and only from the wtFAdV-9 group, were positive for virus. Viral DNA was also
detected in some samples of the bursa of Fabricius until day 14 p.i.; however, the genome
copy numbers were low compared to those in other tissue samples. The viral genome
copy number was highest in cecal tonsil samples, and it was higher in the wtFAdV-9
group than in FAdV-9Δ4-infected chickens.
42
Antibody response
The presence of FAdV-9-specific Ab was determined by ELISA as described previously
(Ojkić and Nagy, 2003) and is shown in Fig. 2.1. No antibodies were detected in any
groups before inoculation or in the mock-infected group at any time. Over the study,
antibody levels increased in both the wtFAdV-9- and FAdV-9Δ4-infected groups from
week 1 p.i. until the end of the experiment at week 4 p.i.. The antibody response to
wtFAdV-9 was significantly higher (P<0.001) than that to FAdV-9Δ4 throughout the
experiment.
Cytokine gene expression in tissues
The expression of mRNA of cytokines in the spleen, liver, bursa of Fabricius, and cecal
tonsil samples was measured by RT-qPCR.
The expression of IFN-α, IFN-γ, IL-10, and IL-12 genes in spleen samples is shown in
Fig. 2.2. There was a statistically significant upregulation (P<0.05) in the expression of
IFN-α in the spleen samples from wtFAdV-9-infected chickens at 7 d.p.i. compared to
that in the mock-infected group. In addition, the expression of IFN-γ was significantly
upregulated (P<0.05) at both 5 and 7 d.p.i. upon wtFAdV-9 infection compared to that in
both the FAdV-9Δ4-infected and mock-infected groups. IL-12, similar to the pattern of
IFN-γ, was also significantly upregulated (P<0.05) in the spleen samples from wtFAdV9-infected chickens at 3, 5, and 7 d.p.i. compared to that in the other two groups.
Moreover, there was also a significant upregulation (P<0.05) of the expression of IL-12
in the spleen samples from FAdV-9Δ4-infected chickens. The expression of IL-10
showed some variations, including both upregulation and downregulation, upon
wtFAdV-9 or FAdV-9Δ4 infection, although they were not significant (P>0.05). It should
43
be noted that IL-10 was downregulated, although not significantly, by wtFAdV-9 at 5 and
7 d.p.i., while IFN-γ was significantly upregulated.
The expression of IFN-α, IFN-γ, IL-10, and IL-12 genes in liver samples is presented in
Fig. 2.3. Similar to the cytokine patterns in spleen samples, upregulation in the
expression of IFN-α, IFN-γ, and IL-12 was found in the wtFAdV-9-infected group at
certain d.p.i. not seen in the mock-infected group. For example, IFN-α was significantly
upregulated (P<0.05) at 3, 5 and 7 d.p.i., as was the case for IFN-γ at 3, 5, and 14 d.p.i.
and for IL-12 at all designated time points except 1 d.p.i. Additionally, compared to
FAdV-9Δ4 infection, the wtFAdV-9 caused a greater level (P<0.05) of induction of IFNα, IFN-γ, and IL-12 at 7, 3, and 7 d.p.i., respectively. In FAdV-9Δ4 infection,
upregulation was noted for only IFN-γ at 3 and 5 d.p.i. (P<0.05). The expression of IL-10
was downregulated (P<0.05) in wtFAdV-9-infected chickens at 5 d.p.i, while IFN-γ was
significantly upregulated (0.001<P<0.05) at that time.
The expression of IFN-α, IFN-γ, IL-10, and IL-12 genes in bursa of Fabricius samples is
illustrated in Fig. 2.4. There was a statistically significant upregulation (P<0.05) of the
expression of IFN-α in bursa samples from wtFAdV-9-infected chickens at 5, 7, and 14
d.p.i. and in bursa samples from FAdV-9Δ4-infected chickens at 14 d.p.i. compared to
that in the mock-infected group. For the expression of IFN-γ, significant upregulation
(P<0.05) was found at 5 and 14 d.p.i. in only the wtFAdV-9-infected chickens. IL-12 was
upregulated significantly by both wtFAdV-9 at 7 and 14 d.p.i. and by FAdV-9Δ4 at 14
d.p.i. On the other hand, the expression of IL-10 was noted for downregulation at 14
d.p.i., while IFN-γ was significantly upregulated (0.001<P<0.05) in the wtFAdV-9 group
at that time.
44
The expression of the IFN-α, IFN-γ, IL-10, and IL-12 genes in cecal tonsil samples is
presented in Fig. 2.5. There was no significant difference (P>0.05) in the expression of
IFN-α, IFN-γ, and IL-10 between any two groups. However, there might be some
downregulation of IL-10 expression at 5 and 7 d.p.i. Unlike other cytokines, IL-12 was
significantly upregulated (P<0.05) by wtFAdV-9 infection at 7 and 14 d.p.i. compared to
the mock-infected group.
Discussion
In the present study, we investigated virus replication and host responses in chickens that
were orally inoculated with our adenovirus vector candidate, FAdV-9Δ4, which lacks six
ORFs at the left end of the viral genome. FAdV-9Δ4, although it replicated less
efficiently in vivo than did wtFAdV-9, induced an antibody response, albeit at a lower
level than in wtFAdV-9-inoculated birds. The cytokine gene expression profiles upon
virus infection showed that wtFAdV-9 significantly upregulated the mRNA expression of
IFN-α, IFN-γ, and IL-12 in all tested tissues except cecal tonsils at least at one tested time
point throughout the experiment, while FAdV-9Δ4 did not.
Human adenoviruses, such as HAdV-5, have been extensively investigated for vectored
vaccine and gene therapy due to their aptitude for inducing potent innate and adaptive
immune responses (Cody and Douglas, 2009; Lasaro and Ertl, 2009; Yamamoto and
Curiel, 2010). However, the use of HAdV-based vectors is hampered by the widespread
preexisting immunity in humans (Lasaro and Ertl, 2009). This initiated interest in the
development of nonhuman AdVs, including FAdVs, which are an attractive choice both
as vaccine vectors for poultry (Francois et al., 2004; Johnson et al., 2003) and as gene
therapy vectors. The optimization of delivery routes and application regimens of AdV
45
vectors are also needed to counteract the limitations of HAdV-based vaccines (Thacker et
al., 2009). Moreover, oral administration of AdV vectors is better able to avoid systemic
neutralizing antibodies than are other routes of administration (Tucker et al., 2008; Xiang
et al., 2003).
Nonpathogenic FAdV-9 is being studied and developed as a vector in our laboratory.
Earlier, we employed both oral and intramuscular administration routes to evaluate the
Ab response to FAdV-9 (Ojkić and Nagy, 2003). However, in that study, virus was given
through water and feed, which means the amount of virus dose taken up by the chickens
was unknown. In more recent studies (Corredor and Nagy, 2010a, 2011), we evaluated
the FAdV-9Δ4 vector virus administered intramuscularly (i.m.), and in the present work,
the chickens were inoculated orally. Similar to with i.m. administration, virus was rarely
detected (only at days 1 and 7 p.i.) in the feces of the orally inoculated FAdV-9Δ4 group
and with titers very significantly lower than those of the wtFAdV-9 group. The route of
inoculation did not alter the period of virus shedding for wtFAdV-9-infected chickens.
After oral inoculation, virus was detected by both plaque assay and quantitative PCR at 1
d.p.i., which showed the highest viral genome copy number throughout the study. One
explanation is that the detected viruses were from the initial inoculum, i.e., the parental
viruses. At 3 d.p.i., the viral genome copy number dropped markedly in all tissue samples.
A second peak of viral genome copy numbers in the liver and bursa of Fabricius samples
from the wtFAdV-9 group occurred at 5 d.p.i., which was well in accord with the highest
titer detected at that time point. Similar trends have been seen for FAdV-8 (Grgić et al.,
2011). Viral genome copy numbers, indicating the virus load in different tissues, were
significantly higher in the wtFAdV-9 group than in the FAdV-9Δ4 group tissue samples
46
except in spleen and were shown to be the highest in the cecal tonsil samples. Similar
results have been obtained for both FAdV-8 (Grgić et al., 2011) and FAdV-4
(unpublished data).
The antibody response after oral inoculation in the wtFAdV-9 group was significantly
higher (P<0.001) than in the FAdV-9Δ4 group throughout the study (Fig. 2.1), which is
similar to the i.m. inoculation results of Corredor and Nagy (Corredor and Nagy, 2010a).
The fact that the mutant virus elicits a less-robust antibody response than the wild-type
virus might be advantageous when the same vector virus is considered in a secondary
treatment or vaccination. The i.m. inoculation induced a higher Ab level not only for
wFAdV-9 but for FAdV-8 (Grgić et al., 2011) and FAdV-4 as well (our unpublished
data).
In addition to the Ab response, we investigated the expression of IFN-α, IFN-γ, IL-12,
and IL-10 genes at different days after oral inoculation. Type I IFNs are essential for the
mediation of potent antiviral responses, and they also upregulate IFN-γ production in
natural killer (NK) cells, which induces a T helper 1 (Th1) response that will activate
cytotoxic T lymphocytes (CTLs) against virus-infected cells. One of the major roles of
IL-10 is to counteract the effects of Th1 responses by inhibiting IFN-γ synthesis
(Endharti et al., 2005). Chicken IL-10 also possesses a similar function (Rothwell et al.,
2004). In the present study, a trend of downregulation of IL-10 gene expression in both
wtFAdV-9 and FAdV-9Δ4 groups was found, which was not surprising considering that
IFN-γ expression was upregulated by both viruses. Likewise, it was also apparent that IL10 gene expression, similar to expression of other cytokine genes, was downregulated to
a larger extent in wtFAdV-9-infected birds than in FAdV-9Δ4 birds. Previous studies
47
(Liu et al., 2003; Muruve et al., 2004) showed that the inflammatory response against an
AdV vector in mice was transient and did not extend beyond 24 h, followed by a
somewhat resting period of inflammatory gene expression that occurred in the liver
samples lasting until 72 h.p.i. At days 4 to 5 p.i. a second dominant peak of inflammatory
gene expression appeared in the liver samples, which is consistent with the adaptive
immune response (Lieber et al., 1997). The cytokine gene expression in our study was
investigated from only day 1 p.i., and thus, the first peak of inflammatory gene
expression might have been missed, although this is unlikely. Nevertheless, the second
peak beginning about 5 d.p.i. was confirmed.
IFN-α and IFN-γ mRNA expression was upregulated in all tissues, except in the cecal
tonsils, of the wtFAdV-9-infected group. The upregulation of these two cytokines was
not remarkable soon after infection (day 1 p.i.), but it became statistically significant at
days 3, 5, and 7 p.i. These data were similar to the results of our study on FAdV-8 (Grgić
et al., 2013). FAdV-9Δ4 infection also upregulated the expression of IFN-α and IFN-γ
mostly in the liver and bursa of Fabricius samples. However, the upregulation by FAdV9Δ4 was less than that by wtFAdV-9 and was statistically significant in the liver samples
only at days 3 and 5 p.i. A significant difference was also noted for the expression of
IFN-α and IFN-γ between wtFAdV-9 and FAdV-9Δ4 groups, which might be due to the
less-efficient replication of FAdV-9Δ4 in inoculated chickens. However, it might also be
due to the deleted ORFs (0, 1, 1A, 1B, 1C, and 2) that potentially have roles in
modulating the host immune response against FAdV infection, as wtFAdV-9 induced a
significantly higher IFN-γ expression than FAdV-9Δ4 at 3 and 5 d.p.i. in spleen samples,
where no significant difference was found in terms of the viral genome copy number.
48
IL-12 is a pleiotropic heterodimeric cytokine comprising two subunits (p35 and p40) and
is secreted by monocytes, macrophages, and dendritic cells (Kato et al., 1996). In
mammals, the key role of IL-12 is the initiation and progression of the Th1-type immune
response that is typically associated with IFN-γ induction by resting and activated T and
NK cells, through inducing the proliferation of the activated T and NK cells (Cho et al.,
1996; Trinchieri, 2003). Both the p40 and p35 genes of chicken interleukin-12 (chIL-12)
are cloned and characterized (Degen et al., 2004), and similar to the mammalian IL-12,
chIL-12 also induces IFN-γ synthesis and proliferative activity of freshly exposed
chicken spleen cells. We found that the mRNA expression of IL-12 (p40) was
upregulated throughout the study in all selected tissues from both the wtFAdV-9 and
FAdV-9Δ4 groups. We found that the mRNA expression of IL-12 (p40) was upregulated
in all selected tissues of wtFAdV-9-infected chickens. Moreover, we also noted that on
the days when IL-12 was significantly upregulated, IFN-γ was also upregulated, which is
in agreement with the results of Degen and coworkers (Degen et al., 2004).
In conclusion, we investigated virus replication and host responses of orally inoculated
chickens with a candidate FAdV vector virus, FAdV-9Δ4. Based on virus shedding and
the number of viral genome copies in selected tissues, virus replication in FAdV-9Δ4 was
less efficient than that in wtFAdV-9, which was similar to that of intramuscular
inoculation. We also demonstrated that both wtFAdV-9 and FAdV-9Δ4 generally
upregulated the mRNA expression of IFN-α, IFN-γ, and IL-12 and had a trend of
downregulation of IL-10 gene expression in vivo. wtFAdV-9 normally caused a larger
extent of regulation than FAdV-9Δ4. Our data suggest that the six deleted ORFs of
FAdV-9Δ4 play an important role not only in virus replication in vivo but also in
49
modulating the host response against FAdV infection, the areas we are currently studying.
Acknowledgments
Li Deng is a recipient of a China Scholarship Council Ph.D. fellowship. This work was
supported by the Natural Sciences and Engineering Research Council of Canada, the
Canadian Poultry Research Council, and the Ontario Ministry of Agriculture and Food.
We thank Sara Languay and Betty-Anne McBey for their technical assistance and the
personnel in the Isolation Unit for their animal care.
50
Table 2.1 Virus titers (pfu/ml) in the feces of chickens orally inoculated with FAdV9Δ4 and wtFAdV-9.
Days
p.i.
FAdV-9Δ4
Titer
wtFAdV-9
%a
Titer
%
0
NDb
0
ND
0
1
1.04 × 102 ± 1.50 × 102
88.9
8.26 × 102 ± 5.81 × 102
100
3
ND
0
2.14 × 103 ± 1.46 × 103
100
5
ND
0
3.97 × 103 ± 4.07 × 103
100
7
1.60 × 101 ± 3.51 × 101
25
4.16 × 102 ± 3.89 × 102
100
10
ND
0
1.94 × 102 ± 2.65 × 102
80
14
ND
0
3.21 × 101 ± 3.12 × 101
53.3
21
ND
0
ND
0
28
ND
0
ND
0
a
b
The percentage of chickens shedding the virus
ND, not detected.
51
Table 2.2 Viral genome copy number in tissues of chickens orally inoculated with
FAdV-9Δ4 and wtFAdV-9.
Days
p.i.
1
3
5
Tissuesa
FAdV-9Δ4
copy number
%
wtFAdV-9
copy number
%
L
4.27×102±1.84×101
100
1.72×105±1.76×104
100
CT
1.68×104±1.58×103
100
1.88×106±1.32×105
100
B
1.15×102±4.84×101
100
2.17×103±1.71×102
100
S
1
8.77×10 ±6.43×10
L
2.08×101±6.42×100
CT
14
21
2
1
0
0
2.98×10 ±1.71×10
100
2
2.29×10 ±4.62×10
100
100
4.12×102±1.18×101
100
100
3
6
3
100
0
0
1.21×10 ±3.08×10
B
6.52×10 ±3.76×10
40
9.65×10 ±2.71×10
60
S
3.75×102±2.05×101
100
6.03×102±2.36×101
100
L
3.39×101±3.88×100
100
8.92×102±5.64×101
100
CT
2.65×103±7.46×101
100
9.80×105±9.35×102
100
B
7
2
b
1
0
1.86×10 ±2.68×10
1
60
0
2
1
60
2
1
3.42×10 ±7.14×10
S
9.54×10 ±7.1×10
100
1.75×10 ±2.01×10
100
L
4.55×101±4.56×100
100
2.12×102±2.46×101
100
CT
2
1
0
0
6.54×10 ±2.79×10
100
3
1
100
1
0
7.53×10 ±5.86×10
B
8.90×10 ±2.76×10
40
8.28×10 ±2.93×10
80
S
8.15×101±3.71×100
100
1.46×102±1.71×101
100
L
1.62×101±3.24×100
40
7.64×101±2.72×100
60
CT
7.30×101±5.45×100
100
1.77×103±1.63×101
100
B
6.73×100±2.26×100
40
3.22×101±5.50×100
40
S
2.65×101±2.35×100
100
1.27×102±5.21×101
100
L
NDc
0
1.56×101±4.35×100
20
CT
4.35×101±1.15×100
100
1.44×103±2.93×101
100
B
ND
0
ND
0
S
1
0
1.52×10 ±5.76×10
100
a
52
1
1.15×10 ±2.71×10
L, liver; CT, cecal tonsil; B, bursa of Fabricius; S, spleen.
The percentage of tissues in which the viral genome was detected
c
ND, not detected.
b
2
100
Figure 2.1 Antibody response in orally inoculated chickens
Antibody (IgG) response to FAdV-9 in chickens orally inoculated with FAdV-9Δ4
(checkered bars) or wtFAdV-9 (striped bars) and in mock-inoculated chickens (white
bars), as measured by ELISA, shown as S/P (sample-to-positive) ratios. *, statistical
significance (P<0.05) compared to the mock-infected group. Brackets above the bars
indicate comparison between wtFAdV-9- and FAdV-9Δ4-infected chickens.
53
Figure 2.2 Cytokine mRNA expression in spleen samples
Cytokine mRNA expression in spleen samples from wtFAdV-9 (striped bars), FAdV-9Δ4
(checkered bars), and mock-infected (white bars) chickens. Target and reference gene
expression levels were quantified by RT-qPCR, and levels are presented relative to βactin expression and normalized to a calibrator. Error bars represent standard error of the
means. The significance of the regulation level between any two groups was analyzed. *,
significant (P<0.05) upregulation compared to the mock-infected group; **, very
significant (0.001<P<0.01) upregulation compared to the mock-infected group. Brackets
above the bars indicate comparison between wtFAdV-9- and FAdV-9Δ4-infected
chickens.
54
Figure 2.3 Cytokine mRNA expression in liver samples
Cytokine mRNA expression in liver samples from wtFAdV-9 (striped bars), FAdV-9Δ4
(checkered bars), and mock-infected (white bars) chickens. Target and reference gene
expression levels were quantified by RT-qPCR and are presented relative to β-actin
expression and normalized to a calibrator. Error bars represent the standard error of the
means. The significance of the regulation level between any two groups was analyzed. *,
significant (P<0.05) up- or downregulation compared to the mock-infected group; **,
very significant (0.001<P<0.01) up- or downregulation compared to the mock-infected
group. Brackets above the bars indicate comparison between wtFAdV-9- and FAdV-9Δ4infected chickens.
55
Figure 2.4 Cytokine mRNA expression in bursa of Fabricius samples
Cytokine mRNA expression in bursa of Fabricius samples from wtFAdV-9 (striped bars),
FAdV-9Δ4 (checkered bars), and mock-infected (white bars) chickens. Target and
reference gene expression levels were quantified by RT-qPCR and are presented relative
to β-actin expression and normalized to a calibrator. Error bars represent the standard
error of the means. The significance of the regulation level between any two groups was
analyzed. *, significant (P<0.05) up- or downregulation compared to the mock-infected
group; **, very significant (0.001<P<0.01) up- or downregulation compared to the
mock-infected group. Brackets above the bars indicate comparison between wtFAdV-9and FAdV-9Δ4-infected chickens.
56
Figure 2.5 Cytokine mRNA expression in cecal tonsil samples
Cytokine mRNA expression in cecal tonsil samples from wtFAdV-9 (striped bars),
FAdV-9Δ4 (checkered bars), and mock-infected (white bars) chickens. Target and
reference gene expression levels were quantified by RT-qPCR and are presented relative
to β-actin expression and normalized to a calibrator. Error bars represent the standard
error of the means. The significance of the regulation level between any two groups was
analyzed. *, significant (P<0.05) up- or downregulation compared to the mock-infected
group; **, very significant (0.001<P<0.01) up- or downregulation compared to the
mock-infected group. Brackets above the bars indicate comparison between wtFAdV-9and FAdV-9Δ4-infected chickens.
57
Chapter 3. Characterization and functional studies of fowl adenovirus 9 dUTPase
Li Deng1, Xiaobing Qin2, Peter Krell3, Ray Lu3, Shayan Sharif1, Éva Nagy1*
1. Department of Pathobiology, University of Guelph, Guelph, Canada
2. College of Animal Science and Veterinary Medicine, Qingdao Agricultural
University, Qingdao, China
3. Department of Molecular and Cellular Biology, University of Guelph, Guelph,
Canada
*
Corresponding author: Éva Nagy
Tel.: +1-(519) 824-4120 Ext. 54783
Fax: +1-(519) 824-5930
E-mail address: [email protected]
Mail address: Department of Pathobiology, Ontario Veterinary College, University of
Guelph, 50 Stone Road East, Guelph, ON, Canada, N1G 2W1
Author’s contributions
LD designed all experiments and performed the majority of them, conducted the data
analysis and wrote the first draft of the manuscript. XQ made an intermediate construct
for the FAdV-9HA-ORF1 virus. RL, SS and ÉN provided guidance during the
experiments. PK and ÉN provided critical review for the manuscript.
To be submitted to Virology.
58
Abstract
Fowl adenoviruses (FAdVs) are being developed as recombinant vaccine vectors and
cancer therapy tools due to their large genomes that allow for a large transgene capacity.
dUTPase, a ubiquitous enzyme that catalyzes the hydrolysis of dUTP to dUMP, found in
many viruses including adenoviruses, has yet to be identified in FAdV-9. A multiple
alignment of dUTPase amino acid sequences suggested that dUTPase exists in most
AdVs including FAdVs. FAdV-9 ORF1 was very similar to other viral dUTPases and
contained the five conserved motifs that define the protein family. Indeed, ORF1 was
verified as a functional dUTPase through a PCR-based dUTPase enzymatic assay.
Moreover, the transcription and protein expression patterns were characterized by RTPCR and Western blot, which showed that FAdV-9 dUTPase was transcribed from 2
hours post-infection (h.p.i.) and the protein was translated from 6 h.p.i., and both
continued to the late phase of virus infection. By immunofluorescence microscopy with a
HA-tagged dUTPase the recombinant virus FAdV-9 dUTPase was localized in both the
cytoplasm and nucleus. Using a dUTPase knockout virus generated through site-directed
mutagenesis and homologous recombination, we found that FAdV-9 dUTPase was not
required for virus replication in vitro, but upregulated the expression of type I interferons.
This is the first study that functionally characterized an early gene in FAdV-9. Our
findings on the FAdV-9 dUTPase have shed new light on the mechanism of host immune
response against FAdV infection.
Key words
fowl adenovirus 9, dUTPase, virus replication, type I interferon
59
Introduction
Fowl adenoviruses (FAdVs) are distributed worldwide in poultry farms. Some FAdVs are
important causative agents of inclusion body hepatitis (IBH) and hydropericardium
syndrome (HPS) that result in significant economic losses to the poultry industry (Hess,
2013). FAdVs belong to the genus Aviadenovirus in the family Adenoviridae. The 12
serotypes are classified into five species: Fowl adenovirus A to Fowl adenovirus E. Due
to the large DNA genome that allows for a large transgene capacity and no or low
pathogenicity, FAdVs are considered as attractive alternatives to human adenoviruses
(HAdVs) as recombinant vaccine vectors and gene therapy tools (Thacker et al., 2009;
Corredor and Nagy, 2010b).
To date, five FAdV genomes have been completely sequenced, including FAdV-1, -4, -5,
-8 and -9 (Chiocca et al., 1996; Griffin and Nagy, 2011; Grgić et al., 2011; Marek et al.,
2012; Marek et al., 2013; Ojkić and Nagy, 2000). Unlike HAdVs, however, very limited
data are available in terms of the molecular biology of FAdVs, as most of the putative
FAdV ORFs share no homology to any known genes, especially for the genus-specific
open reading frames (ORFs) at both ends of the genome. To date, only a few genes have
been functionally characterized and functional studies of them lag behind those of
HAdVs. GAM-1 and ORF22 of FAdV-1 (CELO virus) show the ability of activating
E2F-dependent transcription through interacting with the retinoblastoma protein (pRb),
although neither exhibits any significant sequence homology to E1A proteins of HAdVs
(Lehrmann and Cotten, 1999). In addition, FAdV-1 GAM-1 is verified as an antiapoptotic protein that functions similarly to Bcl-2 and human adenovirus E1B-19K
(Chiocca et al., 1997). Genomic sequence analysis (Cao et al., 1998; Davison et al., 2003)
60
showed that FAdVs encode a deoxyuridine 5′-triphosphate pyrophosphatase (dUTPase)
homolog, which has already been functionally verified in FAdV-1 (Weiss et al., 1997)
but not yet in other FAdVs. However, its function has not been demonstrated.
dUTPase is a ubiquitous enzyme that catalyzes the hydrolysis of dUTP to dUMP and
pyrophosphate (PPi), thereby reducing the dUTP/dTTP ratio in cells and preventing the
incorporation of uracil into DNA. The hydrolysis of dUTP generates dUMP that serves as
a precursor for the biosynthesis of thymidine nucleotides (Harris et al., 1999). All freeliving organisms, as well as many viruses encode dUTPase, including herpesvirus (Glaser
et al., 2006), retrovirus (Payne and Elder, 2001), African swine fever virus (Oliveros et
al., 1999), poxvirus (Cottone et al., 2002), and adenovirus (Weiss et al., 1997). Based on
the oligomerization state, there are three forms of dUTPase representing three distinct
families. The first family contains the monomeric dUTPase encoded exclusively by
mammalian and avian herpesviruses. The second family contains the homodimeric
dUTPases, which are present in protozoa and bacteria. The largest family forms
homotrimers, which are found in prokaryotes, most eukaryotes, some DNA viruses such
as poxvirus, and a number of retroviruses (Tarbouriech et al., 2005).
It has been demonstrated that viral dUTPase is important for virus replication (Threadgill
et al., 1993; Voronin et al., 2014). Other studies showed that the dUTPase of EpsteinBarr virus (EBV) is able to upregulate pro-inflammatory cytokines in unstimulated
peripheral blood mononuclear cells (PBMCs) (Glaser et al., 2006) and the dUTPase of
the murine gammaherpesvirus 68 (MHV-68) can inhibit type I interferon signaling
(Leang et al., 2011).
Interferons (IFNs) are characterized by their potent antiviral properties (Takaoka and
61
Yanai, 2006). They are grouped into two classes: type I IFNs (majorly IFN-α and IFN-β),
which are produced in virally infected cells to confer an antiviral state on uninfected cells,
and type II IFN, which comprises a sole member, IFN-γ, and is strongly produced by
activated T cells or NK cells and implicated in macrophage activation and antiviral
response (Vilcek, 2003; Takaoka and Yanai, 2006).
The objectives of this study were to characterize FAdV-9 ORF1 as a dUTPase enzyme
and to investigate its molecular features including transcription and protein expression
patterns; subcellular localization and its potential role in virus replication in vitro and
modulating cytokine gene expression, including IFN-α, IFN-β, IFN-γ, and IL-10, a
pleiotropic cytokine that inhibits the synthesis of proinflammatory cytokines (Rothwell et
al., 2004).
Materials and Methods
Viruses and cells
The wild type FAdV-9 (wtFAdV-9; strain A-2A) and all mutant viruses including
ORF1stop, resORF1, and FAdV-9HA-ORF1 were propagated and titrated in chicken
hepatoma cells (CH-SAH) as described previously (Alexander et al., 1998). The one-step
growth curves in CH-SAH cells were also done as described (Alexander et al., 1998).
Cytokine gene expression, ORF1 transcription and protein translation, were determined
in Celi cells that were collected from the livers of 15-day-old chicken embryos as
described by Schat and Purchase (1989) and in CH-SAH cells. All cells were grown in
Dulbecco’s modified Eagle’s medium/nutrient mixture F-12 Ham (DMEM-F12)
supplemented with 10% fetal bovine serum (FBS), 2 mM L-glutamine, 100 U/ml
penicillin, and 100 µg/ml streptomycin.
62
dUTPase enzyme activity assay
The assay for testing the dUTPase enzyme activity was adapted from a protocol (Leang et
at., 2011). This is based on the ability of dUTP to be incorporated into a PCR product
during amplification in lieu of dTTP. If presented with active dUTPase, the dUTP would
be hydrolyzed to dUMP which cannot be incorporated and will prevent forming a PCR
product. Briefly, to produce ORF1 protein, ORF1 was cloned into pET-28a first, and the
His-tagged ORF1 protein was expressed in BL21(DE3) E. coli following the induction
with isopropyl-beta-D-thiogalactopyranoside, followed by purification on a Ni-column.
To test the dUTPase enzyme activity, five microliters of purified His-tagged ORF1
protein was incubated with 5 μl of 5 mM dUTP (Roche U1191) for 24 hours in 10 μl of
dUTPase reaction buffer (100 mM Tris pH 7.5, 20 mM MgCl2, 20 mM DTT, and 0.2
mg/ml BSA) at 37 °C. Reactions were terminated by freezing. Instead of a dNTP mix,
dUTP, or His-ORF1-treated dUTP, together with a mixture of individual dATP, dCTP,
and dGTP, were used for PCR amplification of ORF1 as a template (ORF1-For: 5’ACTGAAACCTCGCAGAGGTCC-3’;
ORF1-Rev:
5’-
ACTGAAACCTCGCAGAGGTCC-3’). PCR was conducted using Taq polymerase with
a high concentration of MgCl2 (5 mM). In combination with individual dATP, dCTP, and
dGTP, dTTP served as a positive control, while DNase and RNase-free H2O served as the
negative control. Heat-inactivated His-tagged ORF1 protein (95 °C, 5 mins) was also
used as a control. Cycle conditions were 95°C 10 min; 95 °C 30 seconds, 57 °C 30
seconds, 72 °C 30 seconds for 35 cycles; 72°C 10 minutes; hold at 4°C. PCR products
were run in a 1.0 % agarose gel.
63
Generation of mutant viruses
The FAdmid clone pPacFAdV-9 containing the whole FAdV-9 genome was constructed
previously in our laboratory to manipulate the FAdV-9 genome (Ojkić and Nagy, 2001).
First, the left-end 3.1 kb (nt 210–3316) of FAdV-9 genome was cloned to the pGEM-T
vector by primers pleft-For and pleft-Rev (Table 3.1) to generate the intermediate
construct pGEM-T-pleft (Step 2 in Supplementary Fig. S1). Primers (Table 3.1) were
designed to introduce three stop codons (TAA at nt 850, TGA at nt 865, and TAG at nt
871) and one SwaI site (ATTTAAAT at nt 852; no other SwaI cutting site would be
present in the final construct) within ORF1 (nt 847-1338) for knocking out ORF1,
through two rounds of PCR-based site-directed mutagenesis (SDM) (Step 3 in Fig. S1).
The PCR product was digested with DpnI to get rid of the original methylated pGEM-Tpleft template plasmid, followed by transformation into DH5α. The ORF1-mutated
pGEM-T-pleft was screened by SwaI digestion and sequencing, and the positive plasmid
clone was named ORF1-SDM-pleft. Secondly, the ORF1-mutated left-end FAdV-9
genome was amplified from ORF1-SDM-pleft with pleft-F and pleft-R (Table 3.1),
followed by gel purification (Step 4 in Fig. S1). The FAdmid pPacFAdV-9 was then
digested with SgfI and purified through ethanol precipitation (Step 1 in Fig. S1).
Homologous recombination was performed by co-transforming 500-800 ng purified
mutated PCR product and 50-80 ng SgfI-digested pPacFAdV-9 into E. coli BJ5183 strain
(Step 5 in Fig. S1), as described (Corredor and Nagy, 2010a). The positive mutant
FAdmid was screened by SwaI digestion and sequencing; followed by PacI digestion and
transfection into CH-SAH cells with 2-4 µg DNA. Upon the appearance of cytopathic
effect (CPE) at around 5-6 days post-transfection, cell cultures were frozen and thawed
64
three times, followed by virus propagation in CH-SAH cells in a 150 mm dish. The virus
was obtained through ultra-centrifugation and viral DNA extraction was conducted as
described (Ojkić and Nagy, 2001). The positive mutant virus was confirmed by
sequencing the DNA and named ORF1stop. Similarly, the ORF1-rescued virus resORF1
and HA-tagged virus FAdV-9HA-ORF1 were obtained with primers listed in Table 3.1.
RT-PCR and qRT-PCR
CH-SAH cells were infected with FAdV-9HA-ORF1 at a multiplicity of infection (MOI)
of 5. At 0, 1, 2, 4, 6, 8, 10, 12, 16, 20, and 24 h.p.i., cells were collected for RNA
extraction with Tri Reagent (Molecular Research Center), followed by DNase I treatment
(Fermentas), cDNA synthesis with SuperScript® II Reverse Transcriptase (Invitrogen)
and random primer. The cDNA was amplified by primers HA-tag-For and ORF1-Rev
(Table 3.1). The pcDNA-HA-ORF1 plasmid and RNA of mock-infected cells served as
positive and negative controls, respectively. PCR products were run in a 1.0 % agarose
gel.
RNAs were obtained from CH-SAH cells at 4, 8, 12, 16, 24 h.p.i. and primary chicken
embryo liver cells at 8, 12, and 16 h.p.i. mock-infected or infected with wtFAdV-9,
ORF1stop, resORF1FAdV-9 at a MOI of 5, followed by cDNA synthesis as described
above. The 1/10 diluted cDNA was used for real-time PCR with the gene specific primers
for different cytokines (Table 3.2) and LightCycler® 480 SYBR Green I Master as
described (Grgić et at., 2011). β-actin was used as a housekeeping gene. All data were
normalized to a positive plasmid, which contains the same sequence as the amplicon of
the gene as a calibrator.
65
Western blot
CH-SAH cells infected with FAdV-9HA-ORF1 at an MOI of 5 were washed with PBS
and harvested at 2, 4, 6, 8, 12, 16, 20 and 24 h.p.i. by removing the monolayer with a cell
scraper in PBS, followed by centrifugation at 6000 rpm at 4°C for 10 mins. The cell
pellets were then lysed with RIPA buffer containing a protease inhibitor cocktail on ice
for 20 mins. The supernatant was harvested by centrifugation at 6000 rpm at 4°C for 20
mins. HA-ORF1 fusion protein was probed with mouse anti-HA monoclonal antibody
(Sigma-Aldrich, 1:750) and horseradish peroxidase-conjugated goat anti-mouse IgG
antibody (Invitrogen, 1:10, 000).
Immunofluorescence assay
CH-SAH cells were infected with FAdV-9HA-ORF1 at an MOI of 5. At 6, 12, 18, and 24
h.p.i., cells were fixed with 3.7% paraformaldehyde for 30 mins and permeabilized with
0.1% NP-40 for 20 mins at room temperature. Afterwards, cells were incubated with
mouse anti-HA monoclonal antibody (Sigma-Aldrich, 1:750) for 1 hour, and DyLight
549-conjugated goat anti-mouse IgG antibody (Jackson ImmunoResearch, 1:10, 000) for
1 hour. The ProLong® Gold antifade reagents with DPAI (Life Technologies) was added
and incubated for 10 minutes prior to analysis by confocal microscopy.
Statistical analysis
Statistical analyses were performed with GraphPad Prism 6.0 software (San Diego, CA).
Significant differences of the cytokine gene expressions among the groups were
determined through the two-way ANOVA analysis. The critical level for significance was
set at P<0.05. The data were expressed as mean ± S.E.M., determined from three
66
independent experiments.
Results
Bioinformatics analysis of FAdV-9 ORF1 as a potential dUTPase
The 489 bp nucleotides of FAdV-9 ORF1, located at the left end of the genome (nt 8471335), encodes a polypeptide of 163 amino acids (aa) with a calculated molecular mass
of 17.4 kDa. Previous studies suggested that ORF1 of FAdVs is a dUTPase homolog
(Davison et al., 2003; Corredor and Nagy, 2006). The amino acid sequences of 20
identified or putative viral dUTPases, including 11 mastadenoviruses and 9
aviadenoviruses, were obtained from GeneBank (dUTPase homolog in HAdV-F genome
is not available). The pairwise identities of dUTPase sequences between any two species
were calculated and summarized in the Supplementary Table S1. The data showed that all
HAdV dUTPases share high amino acid sequence identities (40.8% - 68.8%) with each
other, but low (less than 29.1%) with other members in the genus Mastadenovirus, with
the exception of Simian adenovirus A (SAdV-A) (39% - 51.3%). However, dUTPases
from all aviadenoviruses show high amino acid sequence identities (44.3% - 85.0%) of
dUTPase with each other. Low identities were found between the two genera
Mastadenovirus and Aviadenovirus (less than 33.3%) except for that between
aviadenoviruses and Porcine adenovirus C (PAdV-C) (46.5 to 59.9%), Bovine
adenovirus A (BAdV-A) (45.8 to 54.9%), and Tree shrew adenovirus A (TSAdV-A)
(43.0 to 53.0%). These results indicat that the dUTPase genes, while conserved, are
highly variable within the family Adenoviridae, though more conserved within the genus
Aviadenovirus. In addition, 30 dUTPase amino acid sequences from several viruses and
different organisms were aligned to identify the conserved regions. As illustrated in Fig.
67
3.1, the 5 motifs forming the active site are well conserved in all sequences except in
human adenoviruses, SAdV-A and PAdV-A, in which motif 5 is missing. Due to the high
identities of amino acid sequences between FAdV-9 ORF1 and other dUTPases and the
conserved motifs found among them, we concluded that FAdV-9 ORF1 very likely
encodes a dUTPase.
FAdV-9 ORF1 has dUTPase enzymatic activity
To determine if FAdV-9 ORF1 indeed encodes a functional dUTPase, a PCR-based
enzymatic assay was employed. The rationale of the assay is explained in Materials and
Methods. Initially, to express the dUTPase, ORF1 was cloned into pET-28a. The Histagged ORF1 protein was expressed in Escherichia coli (E. coli) BL21 (DE3) following
induction with isopropyl-beta-D-thiogalactopyranoside at 37 ºC. The fusion protein HisORF1, calculated to have a molecular mass of approximately 22.0 kDa, was detected by
SDS-PAGE. The His-ORF1 protein was further confirmed by Western blot, using mouse
anti-His monoclonal antibody (Sigma-Aldrich) and horseradish peroxidase-conjugated
goat anti-mouse IgG antibody (Invitrogen). The His-ORF1 protein was purified through a
Ni-column for the dUTPase enzymatic assay. As shown in Fig. 3.2, PCR products were
detected when using a mixture of dCTP, dGTP, and dATP with either dTTP (lane 1) or
dUTP (lane 2), which served as positive controls. In lanes 3 and 4, PCR products were
detected in the presence of dUTPase reaction buffer itself, indicating that the buffer did
not interfere with the PCR reaction. In lanes 5 and 7, in which dTTP or dUTP was preincubated with His-tagged ORF1 protein before use, PCR product was present in lane 5
(dTTP) but not in lane 7 (dUTP), suggesting that the dUTP was hydrolyzed in the
presence of His-tagged ORF1 protein and the hydrolysis was specific to dUTP (lane 7)
68
but not dTTP (lane 5). To demonstrate that it was the enzymatic activity and not just the
ORF1 protein itself that was responsible for the lack of PCR product in lane 7, the HisORF1 protein was heat-inactivated prior to the incubation with dUTP in dUTPase
reaction buffer in lane 6, in which the PCR product was present. Together, it is concluded
that the ORF1 protein possesses dUTPase enzymatic activity.
In vitro characteristics of ORF1 mutant viruses
Many studies have shown that the viral dUTPase is important for virus replication in nondividing cells (Oliveros et al., 1999; Threadgill et al., 1993). However, the role of FAdV9 ORF1 during the virus life cycle is still unknown. To determine if ORF1 affects virus
replication, we generated an ORF1 knockout virus, ORF1stop, and its rescued revertant
resORF1, as depicted in Supplementary Fig. 3.1. Viral DNA accumulation and one-step
growth curves were analyzed among ORF1stop, resORF1, and wild type FAdV-9 viruses.
As shown in Fig. 3.3, both viral DNA accumulation and virus growth kinetics of
ORF1stop were similar to the wild type and rescued virus, indicating that ORF1 deletion
did not affect viral DNA accumulation or virus replication in vitro.
Transcription and protein expression profiles of ORF1
To study protein expression of FAdV-9 ORF1, an HA-tagged recombinant FAdV-9
(FAdV-9HA-ORF1) was constructed by inserting an HA tag into the N-terminus of
ORF1 in the genome through site-directed mutagenesis and homologous recombination.
CH-SAH cells were infected with FAdV-9HA-ORF1, and the transcription profile of
ORF1 was determined by RT-PCR, which was expected to generate a 516 bp PCR
fragment with primers ORF1-For and ORF1-Rev (Table 3.1). As shown in Fig. 3.4,
ORF1 was transcribed as early as 2 h.p.i. and it continued through the entire time-course.
69
This is consistent with previous results in our laboratory (Cao et al., 1998).
CH-SAH cells were infected with FAdV-9HA-ORF1, and protein expression was
evaluated by Western blot using mouse anti-HA monoclonal antibody (Sigma-Aldrich)
and horseradish peroxidase-conjugated goat anti-mouse IgG antibody (Invitrogen). As
shown in Fig. 3.4, no band was detected at 2 and 4 h.p.i.. However, two bands were
observed from 6 h.p.i. to 24 h.p.i.. Based on the predicted size of HA-ORF1
(approximately 18.5 kDa), the faster migrating band of the two presumably represents
HA-ORF1 protein. The slower migrating band (at ~21 kDa) was present with the faster
migrating one in all lanes from 6 to 24 h.p.i. but not detected at earlier time (2 and 4
h.p.i.). This suggests that the higher molecular weight protein detected by the anti-HA
antibody was at least virus infection dependent and could represent an alternative form of
HA-ORF1. To determine if expression of the larger protein is cell-dependent, DF-1 cells,
primary chicken embryo liver cells (Celi), and primary duck fibroblast cells were also
infected and analyzed by Western blot. The slower migrating band (~21 kDa) and the
band of HA-ORF1 protein appeared for all cells (data not shown).
Cellular localization of ORF1 protein
Studies have shown that other viral dUTPases localize in both the nucleus and cytoplasm;
however, no data is available for any adenoviral dUTPase. In order to determine the
cellular localization of FAdV-9 ORF1, immunofluorescence assay was conducted in CHSAH cells infected with FAdV-9HA-ORF1. As shown in Fig. 3.5, the red fluorescence
representing HA-ORF1 protein first appeared at 6 h.p.i. in the cytoplasm, though
fluorescence in some cells was weak. Starting from 12 h.p.i., some of the red
fluorescence was also seen in the nucleus, and continued to accumulate up to 18 h.p.i.. By
70
24 h.p.i., the red fluorescence was close to be evenly distributed in both the cytoplasm
and nucleus and was quite strong.
Cytokine gene mRNA expression induced by wild type and mutant viruses
The dUTPases of some gammaherpesviruses are demonstrated to modulate the host
immune response against virus infection (Glaser et al., 2006; Leang et al., 2011).To
investigate if FAdV-9 ORF1 has similar functions, CH-SAH cells were infected with
ORF1stop, resORF1 and wtFAdV-9, respectively. Mock-infected cells were used as a
negative control. The mRNA expression of IFN-α, IFN-β, IFN-γ, and IL-10 was
determined by qRT-PCR in each group, relative to β-actin. As illustrated in Fig. 3.6, the
resORF1 group showed very similar results with wtFAdV-9 for all cytokines and all time
points. This demonstrated that the only change made in the genome of ORF1stop was the
knockout of ORF1, to which any difference found between ORF1stop and wtFAdV-9
could be ascribed. Hence the comparisons are reported only for wtFAdV-9. As shown in
Fig. 3.6, despite a slight increase, transcript levels of IFN-α in ORF1stop group were not
statistically different from the mock group from 8 to 24 h.p.i.. However, IFN-α
transcripts were upregulated by wtFAdV-9 at 12 (2.7 fold, 0.001<P<0.01) and 16 h.p.i.
(1.9 fold, 0.001<P<0.01), compared to mock control. In addition, there was a statistically
significant difference of IFN-α transcripts between wtFAdV-9 group and ORF1stop
group at 12 (1.8 fold, 0.001<P<0.01) and 16 h.p.i. (1.6 fold, 0.001<P<0.01). IFN-β
transcripts in both ORF1stop and wtFAdV-9 groups were significantly upregulated at 12
(2.0 fold, P<0.05; 3.5 fold, 0.001<P<0.01) and 16 h.p.i. (1.5 fold, P<0.05; 2.3 fold,
0.001<P<0.01), compared to mock control. However, wtFAdV-9 upregulated the IFN-β
transcripts to a larger extent than ORF1stop did at 12 (1.7 fold, 0.001<P<0.01) and 16
71
h.p.i. (1.5 fold, 0.001<P<0.01). This may suggest that, in addition to ORF1, there is
another FAdV gene that is responsible for the up-regulation of IFN-β.
wtFAdV-9 generally seemed to have induced higher IFN-γ mRNA expression than
ORF1stop at all time points (except at 4 h.p.i.), although the differences were not
statistically significant. Transcript levels of IL-10 were significantly downregulated by
ORF1stop (2.0 fold, P<0.05), resORF1 (2.0 fold, P<0.05), and wtFAdV-9 (1.8 fold,
P<0.05) at 8 h.p.i.. At 12 and 16 h.p.i., despite a downregulation by all viruses compared
to mock control, no significant difference (P>0.05) was found between any two groups.
To investigate if any of the differences were cell line-dependent in modulating cytokine
transcript levels, primary chicken embryo liver cells (Celi) were also tested. Only 8, 12,
16 h.p.i. were selected, since any major differences were found at these time points for
CH-SAH cells. As shown in Fig. 3.7, all viruses upregulated transcript levels of IFN-α,
IFN-β, and IFN-γ, compared to the mock control. The results were very similar to those
of CH-SAH cells. Specifically, despite an increase, IFN-α transcripts in ORF1stop group
were not statistically different with the mock control at all time points, while it was
significantly up-regulated by wtFAdV-9 at 12 h.p.i. (3.6 fold, 0.001<P<0.05) and 16 h.p.i.
(5.6 fold, 0.001<P<0.05). There was a significant difference of the IFN-α transcripts
between ORF1stop and wtFAdV-9 groups at 12 (2.7 fold, 0.001<P<0.05) and 16 h.p.i.
(3.4 fold, 0.001<P<0.05), which was similar to that in CH-SAH cells as well. While there
was some minor increase of transcript level of IFN-β and IFN-γ compared to mock
control, no statistically significant differences were found for any two groups. Downregulation of IL-10 transcripts was found by all viruses at 16 h.p.i., though the differences
between any two groups were not statistically significant.
72
Discussion
In contrast to human adenoviruses, the molecular biology of fowl adenoviruses remains
poorly understood. In the present study, we showed that FAdV-9 ORF1 encodes a
genuine and active dUTPase enzyme that could catalyze the hydrolysis of dUTP. This
dUTPase would be the first early gene functionally characterized in FAdV-9. FAdV-9
dUTPase was first transcribed at 2 h.p.i. and its protein was expressed at 6 h.p.i. in the
cytoplasm, which was localized in both the nucleus and cytoplasm later. We also found
that the presence of FAdV-9 dUTPase did not affect virus replication in vitro but
contributed to the up-regulation of type I interferons.
To date, a number of viral dUTPases have been characterized (Cottone et al., 2002;
Glaser et al., 2006; Oliveros et al., 1999; Payne and Elder, 2001; Weiss et al., 1997). In
this study, we aligned 20 dUTPase amino acid sequences from different adenoviral
species and showed that the dUTPase, though conserved among aviadenoviruses, is
highly variable within the family Adenoviridae. The multiple alignment of 30 dUTPase
amino acid sequences from different viruses and organisms demonstrated that FAdV-9
ORF1 shares high dUTPase amino acid identity with many other viruses and organisms,
with the 5 conserved motifs forming the active site well conserved. On the other hand,
HAdVs dUTPases less well conserved with others, in which of the 5 conserved motifs,
motif 5 is not present and motifs 1, 2 and 3 are not well conserved. This is probably the
reason why HAdV dUTPases lack dUTPase enzymatic activity (Weiss et al., 1997), as
the motifs 3 and 5 were demonstrated to be critical for the catalytic activity in the
dUTPases of herpesvirus and equine infectious anemia virus (Shao et al., 1997; Harris et
al., 1999; Freeman et al., 2009). It is speculated that the HAdVs-encoded dUTPases have
73
evolved from an ancestral dUTPase enzyme to an oncogene leading to cell transformation
(Weiss et al., 1997).
Viral dUTPase is dispensable in dividing cells but is required for efficient virus
replication in non-dividing macrophages, for example, for African swine fever virus
(Oliveros et al., 1999) and equine infectious anemia virus (Threadgill et al., 1993).
However, Voronin and co-workers (2014) have recently showed that even in dividing
cells (Cf2Th cells), the dUTPase of bovine immunodeficiency virus is critical for the
production of progeny virus, although both viral cDNA synthesis and integration into the
host cell DNA were unaltered. In the present study, viral DNA accumulation and onestep growth curves were identical between ORF1stop and wtFAdV-9 in both CH-SAH
cells and primary chicken embryo liver cells, although, it could not be ruled out that
FAdV-9 dUTPase may be important for virus replication in vivo. The dUTPase of herpes
simplex virus does not affect the replication of the virus in vitro, but severely reduces the
replication in the central nervous system of mice (Pyles et al., 1992). Similarly, MHV-68encoded dUTPase is necessary for efficient virus replication in the lungs of infected mice
(Leang et al., 2011).
Previous studies showed that FAdV-9 ORF1 (formerly named ORF LTR1) was
transcribed at 2 h.p.i. (Cao et al., 1998). In this study, we generated a recombinant virus
FAdV-9HA-ORF1 so that the HA-ORF1 protein could be detectable by Western blot
with an anti-HA antibody. An 18.5 kDa HA-ORF1 protein band and a slower migrating
band (~21 kDa) were detected as early as 6 h.p.i., and continued to the late phase of the
virus life cycle and only in virus-infected cells. Other viral dUTPases are also found to be
transcribed and translated at the early stage of virus replication. Zhao et al. (2008)
74
reported that duck enteritis virus dUTPase is transcribed as early as 30 min p.i. and its
protein expression first occurred at 8 h.p.i., whereas the Rana grylio virus-encoded
dUTPase is both transcribed and expressed at 4 h.p.i. (Zhao et al., 2007). It should be
noted that the level of transcription and expression of other viral dUTPase generally
peaked at a certain time point, while that of FAdV-9 dUTPase continuously maintained a
very high level from the beginning to the late phase. In addition, from 12 h.p.i., ORF1
protein appeared and accumulated in the nucleus, and was distributed evenly in both
cytoplasm and nucleus at 24 h.p.i., which is consistent with the results of other viral
dUTPase proteins (Zhao et al., 2007, 2008). These data suggest that FAdV-9 dUTPase
might play an important role in the nucleus during the virus life cycle, perhaps related to
its role in maintaining a low dUTP/dTTP ratio to minimize the incorporation of uracil
into the viral DNA.
The dUTPases encoded by the gammaherpesviruses have novel functions, which are
independent of their enzymatic activity. For instance, Glaser and co-workers (2006)
showed that the EBV-encoded dUTPase is able to induce immune dysregulation in vitro
as demonstrated by the up-regulation of TNF-α, IL-1β, IL-8, IL-6, and IL-10 in
unstimulated PBMCs. In the present study, we compared the transcript level of INFs -α, β, -γ, and one anti-inflammatory cytokine IL-10 in cells infected with ORF1stop,
resORF1, and wtFAdV-9. Our results showed that wtFAdV-9 induced 1.8 and 1.7 fold
higher levels of IFN-α and IFN-β transcripts than ORF1stop, respectively, indicating a
role of FAdV-9 dUTPase in the upregulation of type I IFNs. This is similar to the data on
EBV-encoded dUTPase which upregulated IFN-β by 11-fold (Ariza et al., 2013). Other
studies have also demonstrated that the CD4 and cytokine receptor interacting region of
75
gp120 evolved from an ancestral dUTPase gene (Abergel et al., 1999) and that gp120 of
human immunodeficiency virus (HIV) is involved in higher levels of TNF-α, IL-6, IL-10,
INF-α, and IFN-γ (Rychert et al., 2010).
Type I IFNs represent the first line of defense for the innate immune system (Kawai and
Akira, 2006), which plays an essential role in the host response against viral infections
(Taniguchi and Takaoka, 2001). Moreover, many viruses have developed a myriad of
strategies to evade the host immune response through targeting the type I IFN system.
Viruses encode certain proteins that can inhibit IFN-α and IFN-β synthesis, inactivate
secreted IFN molecules, interfere with IFNAR signaling, and/or block the activation of
antiviral effector proteins upregulated by IFNs (Weber et al., 2004). The examples
include the NS1 protein of influenza A virus (Hale et al., 2008), the NS3/4A of hepatitis
C virus (Li et al., 2005), and the V protein of paramyxovirus (Andrejeva et al., 2004).
Previously, we found that FAdV-9 grows as well in CH-SAH cells pre-treated with
recombinant chicken IFN-α (AbD Serotec®) as in untreated cells (data not shown),
suggesting that FAdV-9 can evade the antiviral response of IFN-α. This indicated that
there should be some viral protein that can counteract the type I IFN response against
FAdV-9 infection. It was hypothesized that dUTPase can perform this function, as is the
case for the dUTPase of the MHV-68 (Leang et al., 2011). However, our data showed
that FAdV-9 dUTPase was able to induce type I interferons, which was surprising but
very interesting as this may indicate that dUTPase might function like a pathogenassociated molecular pattern (PAMP) instead. Ariza and co-workers (2014) have verified
the identity of dUTPases of human herpesvirus 6A (HHV-6A), HHV-8, and varicella
zoster virus (VZV) as a PAMP, as evidenced by activating NF-κB through ligation of
76
TLR2/TLR1 heterodimers and inducing the secretion of the inflammatory cytokines and
IFN-γ. It should be noted, however, that the dUTPase of herpesviruses is monomeric and
is normally twice the size of the sequence of the trimeric dUTPase (typically around 150
aa). Moreover, in the herpesvirus dUTPases, the 5 conserved motifs are reshuffled to an
order of 3-1-2-4-5 instead of the order 1-2-3-4-5 in the trimeric dUTPase (McGeehan et
al., 2001).
It is well known that AdVs induce robust innate and adaptive immune responses,
although the detailed mechanism remains to be clarified. Yamaguchi and co-workers
(2010) have shown that the virus-associated RNAs (VA-RNAs) induce the production of
IFN-α and IFN-β. FAdVs do not have VA-RNAs; therefore, if FAdV-9 dUTPase indeed
functions like a PAMP, it would help our understanding of the mechanism of the FAdVinduced immune response. Further studies are required to test this hypothesis,.
In conclusion, we verified that FAdV-9 ORF1 encodes a functional dUTPase enzyme and
characterized its molecular features including transcription and translation patterns and
cellular localization. We also demonstrated that, although FAdV-9 dUTPase did not
affect virus replication in vitro, it contributed to the up-regulation of type I interferons in
vitro. Our data lays the foundation for further exploration on the mechanism of the host
immune response against FAdV infection.
Acknowledgements
Li Deng was a recipient of China Scholarship Council PhD fellowship. This work was
supported by the Natural Sciences and Engineering Research Council of Canada, the
Canadian Poultry Research Council, and the Ontario Ministry of Agriculture, Food and
Rural Affairs.
77
Table 3.1 Primers for generating mutant viruses
Letters in bold indicate nucleotides that were changed; the letters in italic represent
enzyme sites; the underlined letters indicate stop codons.
Length
Primer Name Sequence (5’-3’)
Functions
(bp)
pleft-F
GCACAGTCCCAATGGCTT
18
Amplify the leftend FAdV-9
pleft-R
AAATTCTGGTCCGTTACCGA
20
genome
SDMORF1-1st- TCCCGTTTAGGTGAGGAGATGTAATTT
Introduce one stop
50
For
AAATCCGGTTGCCCCCCGACTCC
codon and one
SwaI site right
SDMORF1-1st- GGAGTCGGGGGGCAACCGGATTTAAA
after the start
50
Rev
TTACATCTCCTCACCTAAACGGGA
codon of ORF1
SDMORF1AGATGTAATTTAAATCCGGTTGACCC
Introduce another
50
2nd-For
TAGACTCCACCGGTAAAGCTGCTG
two stop codons
further away from
SDMORF1CAGCAGCTTTACCGGTGGAGTCTAGG
the start codon of
50
2nd-Rev
GTCAACCGGATTTAAATTACATCT
ORF1
TCCCGTTTAGGTGAGGAGATGTATCC
N-HA-ORF1ATATGATGTTCCAGATTATGCTTCTT 68
For
Fuse HA tag at the
TCGATTCCGGTTGCCC
N-terminus of
GGGCAACCGGAATCGAAAGAAGCAT
ORF1
N-HA-ORF1AATCTGGAACATCATATGGATACAT 68
Rev
CTCCTCACCTAAACGGGA
Sequence the leftSeq pleftGACCCTGGTCGGAAACGAT
19
end of FAdV-9
FAdV-9
genome
Determine the
transcription
HA-tag-For
TATCCATATGATGTTCCAGATTATGCT 27
profile of HAORF1
CAGGAATTCTCTTTCGATTCCGGTTGC
Clone ORF1 into
ORF1-For
30
CCC
pET-28a plasmid
Clone ORF1 into
pET-28a plasmid;
GCTGCGGCCGCCTAAGAAAAGGAGG
determine
ORF1-Rev
30
AGGGA
transcription
profile of HAORF1
78
Table 3.2 Primers used for cytokines for qRT-PCR
Name
Sequence
Length (bp)
β-actin For
β-actin Rev
IFN-α For
IFN-α Rev
IFN-β For
IFN-β Rev
IFN-γ For
IFN-γ Rev
IL-10 For
IL-10 Rev
CAACACAGTGCTGTCTGGTGGTA
ATCGTACTCCTGCTTGCTGATCC
ATCCTGCTGCTCACGCTCCTTCT
GGTGTTGCTGGTGTCCAGGATG
GCCTCCAGCTCCTTCAGAATACG
CTGGATCTGGTTGAGGAGGCTGT
ACACTGACAAGTCAAAGCCGCACA
AGTCGTTCATCGGGAGCTTGGC
AGCAGATCAAGGAGACGTTC
ATCAGCAGGTACTCCTCGAT
23
23
23
22
23
23
24
22
20
20
79
Mastadenovirus
Supplementary Table S1. Pairwise identities of dUTPase amino acid sequences of adenoviruses.
Twenty adenovirus dUTPase amino acid sequences were obtained from NCBI and aligned with Genious 8.0 program. The horizontal
and vertical lines divides mastadenoviruses and aviadenoviruses. The highest and lowest identity in each group is in bold.
HAdV-A
HAdV-B
HAdV-C
HAdV-D
HAdV-E
HAdV-G
SAdV-A
PAdV-A
PAdV-C
BAdV-A
Aviadenovirusa
TSAdV-A
FAdV-A
FAdV-B
FAdV-C
FAdV-D
FAdV-E
GoAdV-A
TAdV-B
TAdV-C
TAdV-D
HA
dVA
HA
dVB
HA
dVC
HA
dVD
HA
dVE
HA
dVG
50.0
52.0
48.0
45.2
52.4
51.3
14.3
25.8
22.8
17.5
24.2
22.7
25.8
25.0
25.8
22.7
25.8
25.8
21.9
46.0
44.8
68.8
50.0
45.8
16.7
26.0
24.4
19.8
22.8
21.3
22.8
22.8
22.0
19.7
22.8
21.3
21.3
44.8
46.8
46.0
44.2
18.3
23.4
19.7
18.3
22.5
18.6
20.9
20.9
20.2
20.9
21.7
20.2
20.9
43.2
40.8
40.2
22.4
20.6
18.3
18.3
21.4
22.2
26.2
22.2
22.2
21.4
23.8
22.2
19.8
50.0
39.0
17.5
23.6
22.0
22.2
23.6
22.8
24.4
24.4
23.6
18.9
21.3
23.6
21.3
41.5
17.5
29.1
26.0
21.4
27.6
23.6
25.2
26.0
26.0
24.4
26.8
24.4
25.2
SAd
V-A
PAd
V-A
PAd
V-C
13.4
23.3
23.5
17.8
25.6
23.1
24.8
24.0
26.4
23.1
26.4
25.6
24.0
32.5
36.5
34.4
28.6
29.4
27.0
32.5
30.2
31.0
26.2
33.3
28.6
70.9
49.7
59.9
55.6
58.5
56.3
52.8
46.5
47.9
57.0
55.6
a
BA
dVA
TS
Ad
V-A
48.6
53.5
51.4
51.4
53.5
50.7
45.8
46.5
54.9
52.8
52.3
47.0
47.0
49.0
47.0
50.0
43.0
45.6
53.0
FAd
V-A
FAd
V-B
FAd
V-C
FAd
V-D
FAd
V-E
56.7
51.2
58.3
60.8
53.5
45.2
59.6
72.7
48.4
68.2
67.5
46.8
51.9
67.5
53.8
55.7
57.3
45.4
56.8
52.2
49.7
85.0
50.7
53.0
73.9
55.7
50.7
59.1
73.9
56.1
Go
Ad
V-A
TA
dVB
TA
dVC
46.8
51.0
53.8
55.4
44.3
55.6
TA
dVD
Genomes of TAdV-C and TAdV-D have been recently fully sequenced (Marek et al., 2014a). Though not officially accepted yet,
TAdV-C and TAdV-D were suggested to belong to the genus Aviadenovirus.
Abbreviations: HAdV: human adenovirus; SAdV: Simian adenovirus; PAdV: Porcine adenovirus; BAdV: Bovine adenovirus; TSAdV:
Tree shrew adenovirus; FAdV: fowl adenovirus; GoAdV: Goose adenovirus; TAdV: Turkey adenovirus
80
Supplementary Figure S1 Generation of the mutant viruses.
(1) FAdmid pPacFAdV-9 was linearized through digestion with SgfI for later use; (2) the
intermediate construct pGEM-T-pleft was generated by cloning the left-end of the FAdV9 genome onto the pGEM-T easy vector by primers pleft-For and pleft-Rev; (3) two
rounds of site-directed mutagenesis were conducted to introduce the unique SwaI site and
3 stop codons in ORF1 to generate the mutated intermediate construct ORF1-SDM-pleft;
(4) linear mutated left-end of FAdV-9 genome was obtained through amplification of
ORF1-SDM-pleft with primers pleft-For and pleft-Rev for later use; (5) homologous
recombination was carried out with products of (1) and (4) in E.coli BJ5183. Positive
recombinant construct was transformed and propagated in E.coli DH5ɑ, followed by PacI
digestion and transfection into CH-SAH cells. The mutant virus was harvested following
observation of cytopathic effects.
81
Figure 3.1 Multiple alignment of amino acid sequences of dUTPase homologs.
Thirty identified or putative dUTPase amino acid sequences were aligned by ClustalW. The alignment was conducted by ClastW
through Geneious 8.0 program. Letters in different shadow indicates different similarities among all the sequences as below: black
(100%), dark grey (80 to 100%), light grey (60 to 80%), and no shadow (less than 60%). The five conserved motifs 1-5 are indicated
on top of the sequences.
82
Figure 3.2 PCR-based dUTPase enzyme activity assay.
PCR amplification of ORF1 was conducted with a mixture of individual dATP, dGTP,
dCTP, and either dTTP (lanes 1, 3, 5) or dUTP (lanes 2, 4, 6, 7). In the presence of
dUTPase reaction buffer, dTTP or dUTP used for the PCR was incubated with water
(V/V=1:1, lanes 3 and 4), or His-tagged protein (V/V=1:1, lanes 5 and 7), respectively
for 24 hours at 37 °C. Heat-inactivated His-tagged ORF1 protein (95 °C, 5 mins) was
also used as a negative control (lane 6). All PCR products were run in a 1.0% agarose gel.
83
Figure 3.3 Viral DNA accumulation curve and one-step growth curve.
CH-SAH cells were infected at an MOI of 5 with ORF1stop, resORF1 or wtFAdV-9. At
the indicated time points, (A) cells were collected for DNA extraction, followed by qPCR
of viral DNA genome; (B) cells with supernatant were collected for virus titration.
84
Figure 3.4 Transcription and protein expression profiles of ORF1.
CH-SAH cells were infected with FAdV-9HA-ORF1 at an MOI of 5. At the indicated
time points, RNA and cell lysates were collected for RT-PCR (A) and Western blot (B),
respectively. The HA-tag-For and ORF1-Rev for amplification of ORF1 template are
described in Table 1. PCR products were run in an 1.0% agarose gel. Protein was run in a
12% SDS-PAGE gel and transferred to PVDF membrane (GE Healthcare). Mouse antiHA monoclonal antibody and horseradish peroxidase-conjugated goat anti-mouse IgG
antibody were used at 1:750 and 1:10, 000 dilutions, respectively. Anti-β-actin antibody
was used to indicate equivalent loading.
85
Figure 3.5 Cellular localization of ORF1 protein
Immunofluorescence of HA-ORF1 in CH-SAH cells infected with FAdV-9HA-ORF1
virus at an MOI of 5. At the indicated time points, cells were fixed and permeabilized ,
followed by incubation with mouse anti-HA monoclonal antibody (1:750, Sigma-Aldrich)
and DyLight 549-conjugated goat anti-mouse IgG antibody (1:10 000, Jackson
ImmunoResearch). The ProLong® Gold antifade reagents with DAPI (Life Technologies)
was added and incubated for 10 minutes prior to analysis with confocal microscopy.
86
Figure 3.6 Cytokine mRNA expressions in CH-SAH cells
CH-SAH cells were infected with ORF1stop virus and cytokine mRNA expressions were
analyzed at different time points relative to β-actin. Mock-infected cells or cells infected
with resORF1 and wtFAdV-9 were used as controls. RNA were collected for qRT-PCT
with gene-specific primers as described above. Data was analyzed with Graphpad prism
6.0. Two-way ANOVA was used for the statistical analysis between any of the two
groups. * denotes P<0.05 and ** denotes 0.001<P<0.01 when compared to mock control.
Comparison between wtFAdV-9 and ORF1stop-infected group is indicated with brackets
above the bars.
87
Figure 3.7 Cytokine mRNA expressions in Celi cells
Celi cells were infected with ORF1stop virus and cytokine mRNA expressions were
analyzed at different time points relative to β-actin. Mock-infected cells or cells infected
with resORF1 and wtFAdV-9 were used as controls. RNA were collected for qRT-PCT
with gene-specific primers as described for Fig. 6. Data was analyzed with Graphpad
prism 6.0. Two-way ANOVA was used for the statistical analysis between any of the two
groups. * denotes P<0.05 and ** denotes 0.001<P<0.01 when compared to mock control.
Comparison between wtFAdV-9 and ORF1stop-infected group is indicated with brackets
above the bars.
88
Chapter 4. Fowl adenovirus 9 dUTPase plays a role in virus replication in vivo and
in the regulation of the host immune response
Li Deng, Bryan Griffin, Yanlong Pei, David Leishman, Betty-Anne McBey, Shayan
Sharif, Éva Nagy*
Department of Pathobiology, University of Guelph, Guelph, Canada
*
Corresponding author: Éva Nagy
Tel.: +1-(519) 824-4120 Ext. 54783
Fax: +1-(519) 824-5930
E-mail address: [email protected]
Mail address: Department of Pathobiology, Ontario Veterinary College, University of
Guelph, 50 Stone Road East, Guelph, ON, Canada, N1G 2W1
Author’s contributions
LD designed all experiments and performed the majority of them, conducted the data
analysis and wrote the first draft of the manuscript. BG, YP, DL and B-AM helped with
the chicken experiment. DL and B-AM provided technical assistance. SS and ÉN
provided guidance during the experiments. ÉN provided critical review for the
manuscript.
To be submitted to Viral Immunology
89
Abstract
Fowl adenoviruses (FAdVs) are distributed worldwide in poultry farms. Some of FAdVs
are important causative agents of inclusion body hepatitis and hydropericardium
syndrome that cause significant economic losses to the poultry industry. Previously, we
identified FAdV-9 ORF1 as a deoxyuridine 5′-triphosphate pyrophosphatase (dUTPase)
enzyme, which is not required for virus replication in vitro but contributes to the upregulation of type I interferons. In the present study, we compared virus replication in
vivo and the host immune response in chickens orally inoculated with ORF1stop
(dUTPase knockout virus), resORF1 (the rescued revertant of ORF1stop), and wtFAdV-9,
respectively. Our data showed that replication of ORF1stop was delayed on days 1 and 3
post inoculation (p.i.), as evidenced by significantly less virus shedding in feces and
lower viral loads in tissues, compared to those of wtFAdV-9 group. However, there was
no significant difference between the two groups from 5 d.p.i. to the end of the study.
Moreover, we found that there were significant differences between ORF1stop and
wtFAdV-9 in terms of the induction of cytokine gene mRNA expression in tissues and
IgG antibody responses, suggesting the important roles of dUTPase in modulating the
host immune response. Our study clarified the roles of FAdV-9 dUTPase in vivo that are
helpful for better understanding the virus-host interaction.
Key words
fowl adenovirus 9, dUTPase, in vivo, host immune response, cytokine gene expression
90
Introduction
Fowl adenoviruses (FAdVs), belonging to the genus Aviadenovirus in the family
Adenoviridae, are ubiquitous infectious agents in poultry farms and are distributed
worldwide. Although most of them are non-pathogenic or associated with only mild
clinical signs, some FAdV strains are associated with a number of diseases including
inclusion body hepatitis (IBH) and hydropericardium syndrome (HPS) (Hess, 2013).
There are 12 FAdV serotypes, which are subgrouped into five species (Fowl adenovirus A
to Fowl adenovirus E), and each consists of one or more serotypes. The genomes of the
representatives of all five speices have been fully sequenced, such as FAdV-1, -4, -5, -8
and -9 (Chiocca et al., 1996; Griffin and Nagy, 2011; Grgić et al., 2011; Marek et al.,
2013; Ojkić and Nagy, 2000), which are considerably larger than those of
mastadenoviruses. FAdVs are promising alternatives to human adenovirus as vaccine
vectors (Francois et al., 2004; Johnson et al., 2003; Corredor and Nagy, 2010b).
Previously we identified ORF1 of FAdV-9 as a functional dUTPase enzyme (unpublished
data), which was the first early gene functionally identified in FAdV-9. dUTPase is a
ubiquitous enzyme that catalyzes the cleavage of dUTP to dUMP and pyrophosphate
(PPi), maintaining a low dUTP/dTTP ratio in cells and preventing the misincorporation of
deoxyuridine into DNA. To date, the dUTPase gene has been characterized in many
viruses, such as adenovirus (Weiss et al., 1997), herpesvirus (Glaser et al., 2006),
retrovirus (Payne and Elder, 2001), African swine fever virus (Oliveros et al., 1999), and
poxvirus (Cottone et al., 2002). In addition, a dUTPase knockout virus ORF1stop and its
rescued revertant resORF1 were previously generated in our laboratory for functional
studies. Our data showed that FAdV-9 dUTPase was not essential for virus replication in
91
vitro as the knockout of dUTPase did not affect either the viral DNA replication or the
production of progeny virus in CH-SAH cells and primary chicken embryo liver (Celi)
cells. However, it could be not ruled out that FAdV-9 dUTPase might be important for
replication in vivo. For herpes simplex virus (HSV), the deletion of dUTPase did not
affect virus replication in vitro but resulted in reduced replication in the central nervous
system of mice (Pyles et al., 1992). Leang et al. (2011) also demonstrated that dUTPase
of murid herpesvirus 68 (MHV-68) is necessary for efficient virus replication in the lungs
of infected mice. Therefore, it is worth investigating the roles of FAdV-9 dUTPase in vivo.
It has been demonstrated that the dUTPase of Epstein-Barr virus (EBV) up-regulates
several pro-inflammatory cytokines including TNF-α, IL-1β, IL-8, IL-6, and IL-10 in
unstimulated peripheral blood mononuclear cells (PBMCs) (Glaser et al., 2006).
Another group showed that the dUTPase of the murine gammaherpesvirus 68 (MHV-68)
can inhibit type I interferon signaling (Leang et al., 2011). We demonstrated earlier that
although FAdV-9 dUTPase is not required for virus replication in vitro, it contributes to
the up-regulation of the mRNA expressions of type I interferons (unpublished data).
However, how FAdV-9 dUTPase functions in vivo remains unknown.
The aims of this study were to explore the functions of FAdV-9 dUTPase in virus
replication in orally inoculated chickens and its roles in modulating the host immune
response against virus infection. Clocal swabs were collected to test virus shedding
among virus-infected groups. Due to the relatively high tropism of the virus in chickens
(Ojkić and Nagy, 2003), spleen, liver and cecal tonsil were collected for the
determination of viral loads and cytokine gene expression in tissues. Serum samples were
collected to compare host antibody responses induced by all viruses.
92
Materials and Methods
Viruses and cells
ORF1stop, the dUTPase knockout virus, and resORF1, the rescued revertant, were both
generated through site-directed mutagenesis and homologous recombination, based on a
FAdmid clone pPacFAdV-9 that contains the whole FAdV-9 genome (Ojkić and Nagy,
2001). These two viruses, together with wild type virus, wtFAdV-9, were propagated and
titrated in chicken hepatoma cells (CH-SAH) as described previously (Alexander et al.,
1998). CH-SAH cells were maintained in Dulbecco’s modified Eagle’s medium/nutrient
mixture F-12 Ham (DMEM-F12), supplemented with 10% non-heat-inactivated fetal
bovine serum (FBS), 2 mM L-glutamine, 100 U/ml penicillin and 100 μg/ml
streptomycin.
Experimental animals
Specific pathogen free (SPF) white Leghorn chickens were obtained from the Canadian
Food Inspection Agency (Ottawa, ON, Canada) and housed in the Isolation Unit of
University of Guelph throughout the study. All animal experiments were reviewed and
approved by the Animal Care Committee of the University of Guelph according to the
Guide to the Care and Use of Experimental Animals of the Canadian Council on Animal
Care.
Animal experimental design
To explore the functions of FAdV-9 dUTPase in virus replication in vivo and modulating
the host immune responses, a chicken experiment was carried out. One hundred and sixty
four day-old chickens were divided into four groups (n≥40). At the age of 10 days,
93
chickens were orally inoculated with (1) PBS, served as a mock control, (2) 2×107
p.f.u./chick ORF1stop, (3) 2×107 p.f.u./chick wtFAdV-9, and (4) 2×107 p.f.u./chick
resORF1, respectively. At 0, 1, 3, 5, 7, 10, 14, 21, 28 and 35 days post inoculation (d.p.i.),
cloacal swabs were collected from all birds in 1 ml PBS supplemented with 100 U/ml
penicillin and 100 μg/ml streptomycin to determine the virus titer in feces. At 10 hours p.i.
(h.p.i.), 1, 3, 5, 7, 14 and 35 d.p.i., liver and cecal tonsil were collected from 5 birds in
each group to determine the viral load in tissues; spleen, another portion of liver and
cecal tonsil were collected from 5 birds in each group to investigate the cytokine gene
expression in tissues, including IFN-α, IFN-β, IFN-γ, IL-10, and IL-8. In addition, serum
samples were collected at 0, 7, 14, 21, 28, and 35 d.p.i. to test FAdV-specific IgG
antibodies.
DNA and RNA extraction
Viral DNA was extracted from tissues stored at -80ºC with QIAamp DNA Mini Kit
(QIAGEN Inc) following the manufacturer’s instructions. The DNA concentration was
measured using a NanoDrop 2000 spectrophotometer. RNA was extracted with Trizol
(Invitrogen) from tissues stored in RNAlater as described previously (Deng et al., 2013),
prior to digestion with DNase I (Fermentas) and quantification with NanoDrop 2000
spectrophotometer.
qPCR and qRT-PCR
To determine the viral load in tissues, viral DNA was quantified by qPCR with
LightCycler® 480 SYBR Green I Master Kit (Roche Diagnostics) and FAdV-9-specific
primers as described previously (Romanova et al., 2009). To determine the cytokine gene
expression in tissues, cDNA was first synthesized from total RNA with the
94
SuperScriptTM II Reverse Trancriptase (Invitrogen) and random primer. The 1:10 diluted
cDNA was used for qPCR with LightCycler® 480 SYBR Green I Master Kit (Roche
Diagnostics) and gene-specific primers as described previously (Abdul-Careem et al.,
2006, 2007).
Enzyme-linked immunosorbent assay (ELISA)
Heat-inactivated serum samples were diluted 1:100 in wash buffer (0.05% Tween 20 in
PBS) and used for ELISA as previously described (Ojkić and Nagy, 2003). The sample to
positive (S/P) ratio was used to indicate the antibody level (Ojkić and Nagy, 2003). A
serum sample from a previous chicken trial, collected from wtFAdV-9-infected chicken at
28 d.p.i, was used as a positive control.
Statistical analysis
Statistical analyses were performed using GraphPad Prism 6.0 software (San Diego, CA).
A two-way analysis of variance (ANOVA) was used to determine significant differences
between any two groups. The critical level for significance was set at a P value of <0.05.
The data were expressed as mean ± standard error of the mean (SEM).
Results
Virus shedding
To compare virus shedding in chickens among ORF1stop, resORF1 and wtFAdV-9
groups, cloacal swabs were collected for the determination of virus titers in feces by
plaque assay. No virus was detected in any samples from chickens prior to inoculation
and mock-infected group (data not shown), and the virus titers for other groups are
summarized in Table 4.1. Virus was detected in chickens of all three virus-infected
95
groups from day 1 to day 14 p.i., but no virus was detected at days 21, 28 and 35 p.i.. The
peak of virus shedding in all three groups occurred at day 5 p.i., with the titer around
5.0×104 p.f.u./ml. At days 1 and 3 p.i., the virus titers in the samples of ORF1stop group
were significantly lower (P<0.05) than that of wtFAdV-9 group (approximately 4 and 14
fold at 1 and 3 d.p.i., respectively), while no significant difference was found at later days
(5 to 14 d.p.i.). The virus titers of samples from resORF1 group were quite similar to that
of wtFAdV-9 throughout the experiment.
Viral loads in tissues
To determine the viral loads in tissues, the viral genome copy numbers in liver and cecal
tonsil were quantified by real-time PCR with specific FAdV-9 primers. No virus was
detected in samples from chickens before inoculation and mock-infected groups (data not
shown), and the results of other groups are summarized in Table 4.2. As shown in Table
4.2, virus was detected in cecal tonsils of all groups at 10 h.p.i., with the viral genome
copy numbers around 3.0×104/μg total tissue DNA. However, no virus was detected in
livers of any group at this time point. At days 1 and 3 p.i., the viral genome copy numbers
were significantly lower (P<0.05) in both liver and cecal tonsil of ORF1stop group than
those in wtFAdV-9 group (4.3 fold in liver and 3.3 fold in cecal tonsil at 1 d.p.i.; 3.8 fold
in liver and 2.5 fold in cecal tonsil at 3 d.p.i.). At days 5 and 7 p.i., although wtFAdV-9
group had generally higher viral genome copy numbers in both liver and cecal tonsil than
ORF1stop group, there was no significant difference between the two groups. The
resORF1 group showed very similar results with wtFAdV-9, and there was no significant
difference between these two groups at all time points.
96
Antibody response
Antibody levels against FAdV-9 were determined to investigate the role of ORF1 in
modulating the immune response. No antibodies were detected in samples from any
group prior to inoculation (0 d.p.i.). The results of other day points, based on the sample
to positive (S/P) ratio, are shown in Fig. 4.1. Throughout the experiment, no antibodies
were detected in the mock group, while antibody levels increased in all virus-infected
groups from 1 week p.i. until the end of the experiment at 35 d.p.i.. Except at 7 d.p.i., the
antibody level (S/P ratio) in ORF1stop group was significantly lower than that of
wtFAdV-9, while the resORF1 group had very similar antibody response with wtFAdV-9
throughout the study.
Cytokine gene expression in tissues
In order to determine whether FAdV-9 dUTPase could modulate the innate immune
response as it does in vitro, the mRNA expression of IFN-α, IFN-β, IFN-γ, and IL-10 was
investigated in the spleen, liver, and cecal tonsil by real-time PCR. The mRNA
expression of IL-8 was also followed.
In spleens, as is shown in Fig. 4.2, compared to mock control, all viruses resulted in an
increase of the mRNA expression of IFN-α at all time points except for 10 h.p.i.. At 1 and
3 d.p.i. the increase was significantly different with mock control for both resORF1 and
wtFAdV-9 but not for ORF1stop. At these time points, there were significant differences
between ORF1stop and wtFAdV-9. No significant difference was found between any two
groups at 5 and 7 d.p.i.. For the mRNA expression of IFN-β, in spite of some increase in
all virus-infected groups, no significant difference was found between any two groups at
all time points. Similar to IFN-α, the mRNA expression of IFN-γ was upregulated by all
97
three viruses at all time points, compared to mock control. Significant increase was found
in both resORF1 and wtFAdV-9 groups throughout the study. However, no significant
difference was found in the ORF1stop group. There were significant differences between
wtFAdV-9 and ORF1stop at 1, 5 and 7 d.p.i. as well. The mRNA expression of IL-10 was
relatively stable, although increased expression was found at 3 d.p.i. and decreased
expression was found at 5 and 7 d.p.i. However, differences between any two groups
were not significant. In terms of the mRNA expression of IL-8, in spite of some upregulation by all three viruses, there was no significant difference between any two
groups at all time points.
The mRNA expression of IFN-α, IFN-β, IFN-γ, IL-10, and IL-8 in liver samples is
presented in Fig. 4.3. All viruses induced higher IFN-α mRNA expression compared to
mock control at 1, 3, 5 and 7 d.p.i.. At 5 d.p.i., the increase compared to mock found in
both resORF1 and wtFAdV-9 groups but not in ORF1stop was significant. There was no
significant difference between any two virus-infected groups at all time points. Similarly,
the mRNA expression of IFN-β was upregulated by all viruses throughout the study. The
increase in both resORF1 and wtFAdV-9 groups but not in ORF1stop was significant at 5
d.p.i.. Also at this time point, a significant difference was found between ORF1stop and
wtFAdV-9. For the mRNA expression of IFN-γ, all viruses upregulated its expression
compared to mock control at all time points. Both resORF1 and wtFAdV-9 had
significantly higher level of IFN-γ mRNA expression than mock control at 3 and 5 d.p.i..
In addition, there was a significant difference between ORF1stop and wtFAdV-9 at 3
d.p.i.. The production of IL-10 mRNA expression by ORF1stop was very close to that by
mock control at all time points. However, compared to mock control, there were
98
significant decreases of IL-10 mRNA expression in both resORF1 and wtFAdV-9 groups
at 10 h.p.i., 1, 3, and 5 d.p.i.. At 7 d.p.i., decreased IL-10 mRNA expression was also
found in both resORF1 and wtFAdV-9 groups but not in ORF1stop, although there was
no significant difference between any two groups. The mRNA expression of IL-8 was
upregulated by all viruses at all time points, compared to mock control. Only the increase
in both resORF1 and wtFAdV-9 groups but not in ORF1stop group at 1 d.p.i. was found
to be significant.
The mRNA expression of IFN-α, IFN-β, IFN-γ, IL-10, and IL-8 in cecal tonsil samples is
presented in Fig. 4.4. Compared to mock control, there was increased mRNA expression
of all cytokines including IFN-α, IFN-β, IFN-γ, IL-10, and IL-8 in all virus-infected
groups throughout the study (except for a slight decrease for IFN-γ by ORF1stop at 3
d.p.i.), Although not significant, resORF1 and wtFAdV-9 had higher level of IFN-α
mRNA expression at 3, 5, and 7 d.p.i. than ORF1stop. Increased mRNA expression of
IFN-β was found in all virus-infected groups. In addition, resORF1 and wtFAdV-9 but
not ORF1stop showed significant increase of the mRNA expression of IFN-β at 1 d.p.i.,
compared to mock control. resORF1 and wtFAdV-9 had generally higher level of IFN-β
mRNA expression than ORF1stop at 1, 3, 5, and 7 d.p.i.. Increased IFN-γ mRNA
expression was found in all virus-infected groups at all time points (except for ORF1stop
at 1 d.p.i.), while significant increase was only found at 7 d.p.i., compared to mock
control. There was no significant difference in the mRNA expression of IL-8 between any
two groups at all time points, in spite of some up-regulation by all viruses.
99
Discussion
The early genes of human adenoviruses have been extensively studied (reviewed in
Chakraborty and Tansey, 2009; Horwitz, 2004; Weitzman, 2005), while only three early
genes are functionally identified in fowl adenoviruses (Chiocca et al., 1997; Lehrmann
and Cotten, 1999; Weiss et al., 1997). We previously characterized FAdV-9 ORF1 as a
dUTPase enzyme, which is not required for virus replication but is able to upregulate the
expression of type I interferons in CH-SAH cells and Celi cells (unpublished data). In the
present study, we demonstrated that ORF1stop, the dUTPase knockout virus, did not
replicate as efficiently as the wild type virus at the early phase of infection (up to day 3
p.i.). However, no significant difference was observed between the two groups from 5
d.p.i. to the end of the study. Moreover, ORF1stop induced significantly less expression
of IFN-α, IFN-β, IFN-γ and significantly more expression of IL-10 in tissues, and elicited
significantly lower antibody response, compared to wtFAdV-9, suggesting that FAdV-9
dUTPase plays important roles in virus replication in vivo and in modulating the host
immune response.
A number of researchers have reported the roles of viral dUTPase in virus replication
both in vitro and in vivo. For the in vitro studies, the most recognized conclusion is that
the viral dUTPase is not essential in dividing cells but is required for efficient virus
replication in non-dividing host macrophages. This has been demonstrated for the
dUTPase of African swine fever virus (Oliveros et al., 1999) and equine infectious
anemia virus (Threadgill et al., 1993). Consistent with these results, our data also
demonstrated that FAdV-9 dUTPase did not affect virus replication in CH-SAH cells and
Celi cells (unpublished data). However, Voronin and co-workers (2014) showed that the
100
dUTPase of bovine immunodeficiency virus is critical for the virus replication even in
dividing cells, although synthesis of the viral cDNA was not altered. On the other hand,
different results were reported for the roles of viral dUTPase in vivo. Pyles et al. (1992)
showed that the deletion of dUTPase of HSV resulted in reduced replication of virus in
the central nervous system of mice. In contrast, the dUTPase of vaccinia virus is not
required for virus replication in vivo (Prichard et al., 2008). In the present study, we
showed that the replication of ORF1stop, the dUTPase knockout virus, was retarded in
chickens at the early phase of infection (1 and 3 d.p.i.) but the overall virus level was the
same as wild type from 5 d.p.i. to the end of the experiment. This suggests that FAdV-9
dUTPase has an important role in virus replication in vivo, at least at the early phase of
infection. Interestingly, Francois et al. (2001) demonstrated that the transfection of the
dUTPase-deficient FAdV-1 cosmid did not produce cytopathic effects (CPE) in chicken
hepatocarcinoma cells (LMH), and the infection with the transfected cell lysates
produced a CPE later than the wild type FAdV-1 cosmid-transfected cell lysates. Along
with our data, this suggests that the lack of dUTPase might be able to retard virus
replication among all FAdVs.
The antibody level in wtFAdV-9 group increased throughout the experiment, which is
consistent with the result of our previous study (Deng et al., 2013). The antibody level in
ORF1stop group is significantly lower than that in wtFAdV-9 group at all time points
except for 7 d.p.i., suggestting that FAdV-9 dUTPase plays significant roles in the
regulation of the host antibody response. This is possibly related to the ability of FAdV-9
dUTPase in up-regulating the expression of type I interferons. Theofilopoulos et al. (2005)
reported that type I IFNs can promote antibody-mediated immune responses through
101
promoting adaptive T and B cell responses. Zhu et al. (2007b) showed that type I IFNs
induced upon adenoviral infection are critical for the formation virus-specific IgG2a and
IgG2b isotypes, as they mediate the activation of early B cells, formation of the germinal
center, Ig isotype switching and plasma cell differentiation. They also found that Ig
isotype switching and the formation of neutralizing Ab rely on type I IFN signaling in
both B and CD4 T cells. Other studies have also shown that antibody response to
adenoviral vectors is CD4 Th cell dependent (Yang et al., 1995, 1996). Similarly,
Swanson and co-workers (2010) demonstrated that type I IFNs promote production of
antigen-specific IgG2c by follicular B cells, which therefore enhances the magnitude and
quality of the T cell-independent type 2 antibody responses. Previously, we demonstrated
that despite inducing a significantly higher antibody response, compared to the mock
control, a multiORF-deleted virus FAdV-9Δ4, lacking ORFs 0, 1, 1A, 1B, 1C and 2,
induced significantly lower level of antibody than the wild type virus (Corredor and Nagy,
2010a; Deng et al., 2013). Together with our data, this suggests that of the six deleted
genes, dUTPase, if not the only, plays significant roles in the mediation of antibody
response. Indeed, ORF1C shares some amino acid identities to the E5 oncoprotein of
bovine papillomavirus type 1 (Ojkić and Nagy, 2000) and is hypothesized to
downregulate major histocompatibility complex class I molecules (Corredor and Nagy,
2010a). However, to determine if there are more genes other than dUTPase (ORF1)
involved in the mediation of antibody response, more research has to be done.
Among numerous innate immune cytokines, type I interferons are pleiotropic cytokines
that are capable of exerting direct antiviral immune response against virus infection.
Induction of type I IFNs is part of the innate immune response to adenovirus infection.
102
We previously found that FAdV-9 dUTPase is important for the up-regulation of type I
IFNs in CH-SAH cells and Celi cells. In this study, the mRNA expression of IFN-α and
IFN-β was investigated in spleen, liver and cecal tonsil. According to our data, the
dUTPase knockout virus, ORF1stop, induced significantly less mRNA expression of
IFN-α in spleen at 1 and 3 d.p.i., and significantly less mRNA expression of IFN-β in
liver at 5 d.p.i., compared to the wild type virus. This indicates that the role of FAdV-9
dUTPase in up-regulating type I IFNs is not limited to in vitro, but also applies to in vivo.
IFN-γ is the only type II interferon that plays key roles in modulating immunity to
infectious diseases. In our in vitro study, in which both CH-SAH cells and Celi cells were
infected with ORF1stop or wtFAdV-9, no significant difference was found in terms of the
mRNA expression of IFN-γ between these two groups. However, the in vivo study
showed that there was significantly less expression of IFN-γ in livers of ORF1stop group
than that of wtFAdV-9 group at 3 d.p.i.; so was the case in spleens at 1, 5, and 7 d.p.i..
This could be due to that organs like spleen and liver have more types of cells capable of
producing INF-γ, compared to a single type of cell lines.
Overall, our data demonstrated that FAdV-9 dUTPase retarded the virus replication at the
early stage of the infection and played important roles in modulating the host immune
response against virus infection as well. Our study provides useful data that will shed
more lights on the mechanism of the host immune response against the fowl adenovirus
infection.
Acknowledgement
Li Deng was a recipient of China Scholarship Council PhD fellowship. This work was
supported by the Natural Sciences and Engineering Research Council of Canada, the
103
Canadian Poultry Research Council, and the Ontario Ministry of Agriculture, Food and
Rural Affairs. The authors wish to thank the personnel of the Isolation Unit for their
professional animal care and assistance.
104
Table 4.1 Virus titers in the feces of infected chickens
Days
ORF1stop
resORF1
wtFAdV-9
Titer (pfu/ml)
%a
Titer (pfu/ml)
%
Titer (pfu/ml)
%
1
1.6×103±3.5×102
100
5.7×103±1.0×103
100
6.4×103±1.2×103
100
3
3.6×103±5.5×102
100
4.8×104±2.1×104
100
4.9×104±1.1×104
100
5
5.0×104±1.1×104
100
5.2×104±1.4×104
100
5.5×104±1.1×104
100
7
5.4×103±1.9×103
100
4.9×103±1.4×103
100
5.3×103±1.3×103
100
10
7.3×102±4.4×102
66.7
7.9×102±1.9×102
85.7
7.1×102±1.9×102
80
14
3.4×101±5.8×100
33.3
3.8×101±4.4×100
71.4
3.8×101±7.1×100
66.7
21
NDb
0
ND
0
ND
0
28
ND
0
ND
0
ND
0
35
ND
0
ND
0
ND
0
p.i.
a
percentage of chickens shedding viruses
b
ND = not detected
105
Table 4.2 Viral genome copy number in tissues
Viral genome copy number in liver (L) and cecal tonsil (CT) of chickens orally
inoculated with ORF1stop, resORF1 and wtFAdV-9, expressed as copies/μg total tissue
DNA.
Days p.i.
Tissue
ORF1stop
10 hpi
L
NDa
1
3
5
7
a
resORF1
wtFAdV-9
ND
3
3
ND
3
CT
2.74×10 ±1.6×10
2.88×10 ±7.8×10
3.12×103±7.4×102
L
4.91×101±9.8×100
1.90×102±3.7×101
2.10×102±6.2×101
CT
2.18×105±4.1×104
6.37×105±1.3×105
7.16×105±1.9×105
L
1.31×102±4.2×101
5.35×102±2.2×102
4.92×102±1.1×102
CT
3.27×105±1.4×105
8.32×105±3.5×105
8.11×105±2.8×105
L
1.15×103±3.6×102
1.56×103±2.9×102
1.46×103±4.8×102
CT
2.61×105±8.2×104
3.26×105±9.7×104
3.20×105±1.0×105
L
3.89×102±1.1×102
5.75×102±2.3×102
5.29×102±1.8×102
CT
2.0×104±9.6×103
2.76×104±7.1×103
3.19×104±1.7×104
ND = not detected
106
2
Figure 4.1 FAdV-specific IgG antibody response in chickens.
Chickens were orally inoculated with ORF1stop, or resORF1, or wtFAdV-9, or mockinfected. Heat-inactivated serum samples were diluted 1:100 in wash buffer (0.05%
Tween 20 in PBS) and used for ELISA. The optical density (OD) value was read in a
Bio-Tek ELISA microplate reader at 405 nm. The sample to positive (S/P) ratio was used
to indicate the antibody level. A serum sample that is collected from wtFAdV-9-infected
chicken at 28 d.p.i in a previous chicken trial was used as a positive control. Statistical
significance was indicated by * (P<0.05), ** (0.001<P<0.01), *** (P<0.001). Brackets
above the bars indicate comparison between ORF1stop and wtFAdV-9.
107
Figure 4.2 Cytokine mRNA expression in spleen samples of chickens.
Chickens were orally inoculated with ORF1stop, resORF1, or wtFAdV-9, or mockinfected. The mRNA expression of target and reference genes was quantified by qRTPCR, and is presented relative to β-actin expression and normalized to a positive plasmid
calibrator. Error bars represent standard error of the means. Statistical significance was
indicated by * (P<0.05) or ** (0.001<P<0.01), compared to mock control. Brackets
above the bars indicate comparison between ORF1stop group and wtFAdV-9 group.
108
Figure 4.3 Cytokine mRNA expression in liver samples of chickens.
Chickens were orally inoculated with ORF1stop, resORF1, or wtFAdV-9, or mockinfected. The mRNA expression of target and reference genes was quantified by qRTPCR, and is presented relative to β-actin expression and normalized to a positive plasmid
calibrator. Error bars represent standard error of the means. Statistical significance was
indicated by * (P<0.05) or ** (0.001<P<0.01), compared to mock control. Brackets
above the bars indicate comparison between ORF1stop group and wtFAdV-9 group.
109
Figure 4.4 Cytokine mRNA expression in cecal tonsil samples of chickens.
Chickens were orally inoculated with ORF1stop, resORF1, or wtFAdV-9, or mockinfected. The mRNA expression of target and reference genes was quantified by qRTPCR, and is presented relative to β-actin expression and normalized to a positive plasmid
calibrator. Error bars represent standard error of the means. Statistical significance was
indicated by * (P<0.05) or ** (0.001<P<0.01), compared to mock control.
110
Chapter 5. General discussion
It is well known that the major limitation of HAdV-based vectors for clinical use is the
universal pre-existing immunity in humans (Lasaro and Ertl, 2009). To overcome this,
alternative serotypes and non-human adenoviruses have been investigated for the ability
to avoid acute toxicity and evade anti-AdV immune clearance. Tordo et al. (2008) and
Xiang et al. (2014) used an E1-deleted canine adenoviruses type-2 (CAdV-2) and an
experimental chimpanzee adenovirus vector Simian adenovirus serotype 24 (SAdV-24),
respectively, to express the rabies virus glycoprotein, both of which induced high levels
of anti-rabies virus neutralizing antibody and conferred protection against the rabies virus
lethal challenge. Fowl adenoviruses are also investigated as vaccine vectors for avian
diseases. For example, FAdV-1- and FAdV-10-based vectors have been demonstrated to
induce protective immunity against IBDV (Francois et al., 2004; Sheppard et al., 1998);
and a FAdV-8-based vector was studied against infectious bronchitis virus (IBV)
(Johnson et al., 2003). FAdV-9, which is being studied in our laboratory, also shows
promising results, as is described in more detail below.
In addition to searching alternatives of HAdV vectors, the optimization of delivery routes
and regimens is considered as another potential solution (Thacker et al., 2009). For
example, Steitz et al. (2010) compared intratracheal, conjunctival, subcutaneous, and in
ovo routes to evaluate the optimal vaccine administration for the HAdV-5-based H5N1
highly pathogenic avian influenza (HAPI) vaccine in chickens and revealed that the
subcutaneous injection induces the highest humoral immune responses measurable by
hemagglutination inhibition test (HI). In particular, it is demonstrated that oral
administration of AdV vectors is better able to avoid systemic neutralizing antibodies
111
compared with other routes of administration (Xiang et al., 2003) and that the efficacy of
oral vaccination with HAdV-5-based vectors is unaffected by pre-existing immunity in
BALB/c mice (Tucker, 2008). Previous studies in our laboratory employed intramuscular
inoculation as the administration route to develop FAdV-9 as a vaccine vector. However,
the fecal-oral route is a more natural transmission route of FAdVs infection. In the study
of Chapter 2, chickens were orally inoculated with FAdV-9Δ4 or wtFAdV-9. Results
showed that wtFAdV-9 resulted in virus shedding in feces with titers similar to that of
intramuscular inoculation of chickens. However, it should be noted that a higher
inoculum was used in this study than that in the previous intramuscular inoculation study
(1.5×107 pfu/chick versus 2.0×106 pfu/chick), with the consideration of the strong innate
mucosal immunity in the gastrointestinal tract. The results showed that the highest titer in
feces appeared earlier, compared to that of the intramuscular route study (4.0 ×103 pfu/ml
at 5 d.p.i versus 9.6 ×103 pfu/ml at 7 d.p.i), which might indicate that the oral inoculation
triggers the host immune response to clear the virus at an earlier stage than the
intramuscular inoculation. In addition, the antibody response induced by both wtFAdV-9
and FAdV-9Δ4 through oral administration are similar to those by intramuscular
administration, suggesting that the oral administration could be, if not better, at least
competitive with the intramuscular route and easier to administer.
FAdV-9 has been developed as a vaccine vector in our laboratory. It has been
demonstrated that the tandem repeat region 2 (TR-2) at the right end of FAdV-9 genome
is dispensable for virus propagation in vitro and suitable for insertion of foreign genes
(Ojkić and Nagy, 2001). FAdV-9Δ4, a deleted virus lacking ORFs 0, 1, 1A, 1B, 1C and 2,
replicates less efficiently and induces lower FAdV-specific IgG antibody level than wild
112
type virus in intramuscularly inoculated chickens (Corredor and Nagy, 2010a). These data
suggest the importance of the left end genes of FAdV-9 genome in terms of virus
replication and modulation of immune response.
Among the six deleted ORFs, ORF1, a dUTPase homolog, is the most conserved one in
all FAdV genomes, and its amino acid identities ranges from 56% to 100% among FAdV
dUTPases (Corredor and Nagy, 2006). This implies that ORF1 is very likely to play
important roles in the FAdV life cycle. In this study, FAdV-9 ORF1, containing 5
conserved motifs that form the active sites of the enzyme, was verified as a functional
dUTPase, as demonstrated by its ability of catalyzing the hydrolysis of dUTP.
dUTPase exists in all free-living organisms, as well as in many viruses. In recent years,
functional studies of viral dUTPases have demonstrated that this gene is important for
virus replication (Pyles et al., 1992; Glaser et al., 2006; Leang et al., 2011; Ariza et al.,
2013). In the study of Chapter 3, it was demonstrated that FAdV-9 dUTPase was not
required for virus replication in CH-SAH cells and Celi cells, which is consistent with the
results of studies of other viral dUTPases (Oliveros et al., 1999; Threadgill et al., 1993).
However, in the in vivo study of Chapter 4, ORF1stop, the dUTPase knockout virus, did
not replicate as well as wild type virus in orally inoculated chickens at days 1 and 3 p.i.,
indicating that FAdV-9 dUTPase plays a role in virus replication in vivo.
In the study of Chapter 3, the mRNA expression of INFs -α, -β, -γ, and IL-10 was
investigated in CH-SAH cells and Celi cells infected with ORF1stop or wtFAdV-9.
Results showed that wtFAdV-9 induced significantly higher mRNA expression of IFN-α
and IFN-β than ORF1stop, indicating the role of FAdV-9 dUTPase in the upregulation of
type I IFNs. This was contrary with the results of Leang et al. (2011), which showed that
113
the MHV-68-encoded dUTPase counteracts the antiviral reponse of type I IFNs.
Nevertheless, it is demonstrated that EBV-encoded dUTPase is able to upregulate the
expression of IFN-β (~11 fold) (Ariza et al., 2013). Another group also showed that
gp120 of HIV is involved in the higher levels of TNF-α, IL-6, IL-10, IFN-α, and IFN-γ
(Rychert et al., 2010). It should be noted that cytokine receptor interacting region of
gp120 is evolved from an ancestral dUTPase gene (Abergel et al. 1999). In the study of
Chapter 4, similar results were found in terms of the mRNA expression of INFs -α, -β, -γ,
IL-10 in tissues of chickens inoculated with ORF1stop or wtFAdV-9. Together, these
data indicate that FAdV-9 dUTPase plays important roles in the upregulation of type I
IFNs both in vitro and in vivo.
Recent studies demonstrated that functions of several dUTPases in affecting virus
replication and modulating the immune response are not related to their enzymatic
activity. For instance, Voronin et al. (2014) showed that despite the lack of dUTPase
activity, the dUTPase-related gene of bovine immunodeficiency virus is critical for viral
replication. Leang et al. (2011) showed that the inhibition of the type I interferon
response by MHV-68-encoded dUTPase is independent of its enzymatic activity.
Therefore, it would be interesting to investigate whether the capacity of FAdV-9 dUTPase
to affecting virus replication and upregulating type I IFNs is independent of its enzymatic
activity. To achieve this, it is necessary to determine the key residues and peptide for its
catalysis through site-directed mutagenesis.
Several studies have shown that the motifs 3 and 5 of dUTPase are critical for the
catalysis. For example, Freeman et al. (2009) showed that deletion of the flexible Cterminal tail carrying motif 5 of EBV-encoded dUTPase resulted in a protein completely
114
devoid of enzymatic activity. Results of site-directed mutagenesis within motif 5 further
demonstrated that replacement of arginine at nt 268 with alanine largely affects the
enzyme activity, and replacements of phenylalanine at nt 273 with alanine leads to a nondetectable enzymatic activity (Freeman et al., 2009). It is very likely that motif 5 is also
critical for the dUTPase of HAdVs, as dUTPase of HAdV-9 lacking the motif 5 does not
show detectable enzymatic activity (Weiss et al., 1997). Whether dUTPases of other
HAdVs possess the enzymatic activity is unknow, however, it was demonstrated in
Chapter 3 that the absence of the motif 5 is universal for dUTPases of all human
adenoviruses (Fig 3.1). To test if the motif 5 is also critical for FAdV-9 dUTPase, future
research needs to be done. Also, it would be interesting to explore if there is other critical
motif and core sequence responsible for the enzymatic activity of FAdV-9 dUTPase.
Truncated mutants of FAdV-9 dUTPase could be generated through site-directed
mutagenesis to achieve this aim.
In the present work, FAdV-9 dUTPase was demonstrated to contribute to the upregulation of type I interferons in vitro and in vivo. However, the detailed mechanism is
still obscure. Glaser et al. (2006) showed that purified EBV-encoded dUTPase is able to
induce immune dysregulation in vitro by up-regulating the expression of proinflammatory cytokines including TNF-α, IL-1β, IL-8, IL-6, and IL-10 in PBMCs. It was
further demonstrated that this increased expression in response to the EBV-encoded
dUTPase treatment was dependent on the activation of NF-ĸB (Waldman et al., 2008).
Ariza et al. (2009) showed that EBV-encoded dUTPase activates NF-ĸB in a dosedependent manner through TLR2, but not TLR3, TLR4, or TLR4/MD2, and that the
activation of NF-ĸB requires the recruitment of the adaptor molecule MyD88, indicating
115
that EBV-encoded dUTPase could potentially modulate the innate immune response in
EBV-permissive cells through the TLR2 and MyD88 signal transduction pathway. To
determine how FAdV-9 dUTPase up-regulates the cytokine gene expression, future
research needs to be done.
For human adenoviruses, several TLRs are involved in sensing the virus. TLR-9 was
found to sense HAdV-B in peripheral blood mononuclear cells and plasmacytoid
dendritic cells (pDCs) (Sirena et al., 2004; Hendrickx et al., 2014). In addition, TLR2
knockout mice showed reduced NF-κB activation and humoral responses to HAdV
vectors (Appledorn et al., 2008). In mammals, high TLR9 expression in pDCs serves as a
sensor for DNA virus infection (such as HSV-1 and HSV-2) by exerting an effective
antiviral immune response by producing type I IFN (Lund et al., 2003; Krug et al., 2004).
Chicken TLR2 has been identified, while there is no orthologous gene of mammalian
TLR9 idetified in chickens (Juul-Madsen et al., 2011; Chen et al., 2013). It is reported
that TLR21, identified in chickens but not in mammals, has a similar function to
mammalian TLR9 in recognition of CpG DNA (Keestra et al., 2010). Therefore, it would
be worth exploring whether FAdV-9 dUTPase up-regulates the type I interferons through
TLR21 or other TLRs, and investigating what up-stream cytokines of type I IFNs
pathway are involved and how their gene expressions are regulated.
Generally, viruses encode certain proteins that can counteract the antiviral response of
type I IFNs, for example, the NS1 protein of influenza A virus (Hale et al., 2008), the
NS3/4A of hepatitis C virus (Li et al., 2005), and the V protein of paramyxovirus
(Andrejeva et al., 2004). In human adenoviruses, E1A is known to block this type I IFNmediated response, although the exact mechanisms have not been elucidated. This
116
brought up the question that what viral genes of FAdV-9 would function as an antiviral
response inhibitor. Initially, ORF1 of FAdV-9 was hypothesized to be one potential target,
as is the case for the MHV-68-encoded dUTPase (Leang et al., 2011). However, based on
the study of Chapter 3, it seems that dUTPase of FAdV-9 does not possess this function.
Previously, it was found that both wtFAdV-9 and FAdV-9Δ4 grow in CH-SAH cells pretreated with recombinant chicken IFN-α as well as in untreated ones (data not shown),
suggesting that both viruses can evade the antiviral response of IFN-α. This indicates that
there should be some viral proteins in FAdV-9 genome, rather than the six ORFs (0, 1,
1A, 1B, 1C and 2) deleted in FAdV-9Δ4, which are responsible for this inhibition of IFNα antiviral response.
The E4 region might be translocated from the right end of human adenovirus genomes to
the left end in FAdV genomes. For example, E4orf1, the dUTPase homolog, is located in
the right end of HAdV genomes, while dUTPases of FAdVs are located in the left end of
the genome (Weiss et al., 1997; Harrach et al., 2011). Therefore, it would be interesting
to investigate whether the genes that function like E1A, at the left end of mastadenovirus
genome, would be translocated to the right end of FAdV genomes.
In conclusion, in this work, FAdV-9 ORF1 was characterized as a functional dUTPase
enzyme and its molecular features including transcription and translation patterns,
cellular localization were determined. Functional studies of FAdV-9 dUTPase
demonstrated that although FAdV-9 dUTPase did not affect virus replication it
contributed to the up-regulation of type I interferons in vitro. Moreover, in vivo study
showed that FAdV-9 dUTPase plays significant roles in virus replication at the early
117
stage of the infection (up to 3 d.p.i.) and in modulating the host immune response against
virus infection.
The data presented in this work contribute to the better understanding of molecular
biology of FAdVs, and are also helpful for exploring the mechanism of the host immune
response against the fowl adenovirus infection.
118
References
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, 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.
Abergel, C., Robertson, D.L., Claverie, J., 1999. “ Hidden ” dUTPase sequence in human
immunodeficiency virus type 1 gp120. J. Virol. 73, 751–753.
Ablack, J.N., Pelka, P., Yousef, A.F., Turnell, A.S., Grand, R.J., Mymryk, J.S., 2010.
Comparison of E1A CR3-dependent transcriptional activation across six different
human adenovirus subgroups. J. Virol. 84, 12771-12781.
Adair, B.M., Fitzgerald, S.D., 2008. Group 1 adenovirus infections. In: Saif, Y.M., Fadly,
A.M., Glisson, J.R., McDougald, L.R., Nolan, L.K., Swayne, D.E., Diseases of
poultry (12th Edition). Wiley-Blackwell, Ames, IA. 260–286.
Ahi, Y.S., Bangari, D.S., Mittal, S.K., 2011. Adenoviral vector immunity: its implications
and circumvention strategies. Curr. Gene Ther. 11, 307–320.
doi:10.2174/156652311796150372
Ahmad, M.U., Burgess, G.W., 2001. Production and characterization of monoclonal
antibodies to fowl adenoviruses. Avian Pathol. 30, 457–463.
Alexander, H.S., Huber, P., Cao, J., 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.
Andrejeva, J., Childs, K.S., Young, D.F., Carlos, T.S., Stock, N., Goodbourn, S., Randall,
R.E., 2004. The V proteins of paramyxoviruses bind the IFN-inducible RNA
helicase, mda-5, and inhibit its activation of the IFN-beta promoter. Proc. Natl. Acad.
Sci. U. S. A. 101, 17264–17269. doi:10.1073/pnas.0407639101
Appledorn, D.M., Patial, S., McBride, A., Godbehere, S., Van Rooijen, N.,
Parameswaran, N., Amalfitano, A., 2008. Adenovirus vector-induced innate
inflammatory mediators, MAPK signaling, as well as adaptive immune responses
are dependent upon both TLR2 and TLR9 in vivo. J. Immunol. 181, 2134–2144.
119
Ariza, M.E., Glaser, R., Kaumaya, P.T.P., Jones, C., Williams, M.V., 2009. The EBVencoded dUTPase activates NF-kappa B through the TLR2 and MyD88-dependent
signaling pathway. J. Immunol. 182, 851–859.
Ariza, M.E., Glaser, R., Williams, M. V, 2014. Human herpesviruses-encoded dUTPases:
a family of proteins that modulate dendritic cell function and innate immunity. Front.
Microbiol. 5, 504. doi:10.3389/fmicb.2014.00504
Ariza, M.E., Rivailler, P., Glaser, R., Chen, M., Williams, M.V., 2013. Epstein-Barr virus
encoded dUTPase containing exosomes modulate innate and adaptive immune
responses in human dendritic cells and peripheral blood mononuclear cells. PLoS
One. 8, e69827. doi:10.1371/journal.pone.0069827
Avila, G. A., Ramirez, D.H., Hildenbrand, Z.L., Jacquez, P., Chiocca, S., Sun, J., RosasAcosta, G., Xiao, C., 2015. Expression and in vitro functional analyses of
recombinant Gam1 protein. Protein Expr. Purif. 105, 47–53.
doi:10.1016/j.pep.2014.10.005
Bangari, D.S., Mittal, S.K., 2006. Development of nonhuman adenoviruses as vaccine
vectors. Vaccine. 24, 849–862. doi:10.1016/j.vaccine.2005.08.101
Benkö, M., Elo, P., Ursu, K., Ahne, W., LaPatra, S.E., Thomson, D., Harrach, B., 2002.
First molecular evidence for the existence of distinct fish and snake adenoviruses. J.
Virol. 76, 10056–10059. doi:10.1128/JVI.76.19.10056-10059.2002
Berk, A. J. 2013. Adenoviridae. In: Knipe, D. M., Howley, P., Fields Virology (6th
Edition). Lippincott Williams & Wilkins, Philadelphia, PA. 1704-1731.
Bonjardim, C.A., Ferreira, P.C., Kroon, E.G., 2009. Interferons: Signaling, antiviral and
viral evasion. Immunol. Lett. 122, 1-11.
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.
Chakraborty, A.A., Tansey, W.P., 2009. Adenoviral E1A function through Myc. Cancer
Res. 69, 6-9. doi:10.1158/0008-5472.CAN-08-3026
Chakravarti, D., Ogryzko, V., Kao, H.Y., Nash, A., Chen, H., Nakatani, Y., Evans, R.M.,
1999. A viral mechanism for inhibition of p300 and PCAF acetyltransferase activity.
Cell. 96, 393-403.
Challberg, M.D., Kelly, T.J., 1981. Processing of the adenovirus terminal protein. J. Virol.
38, 272-277.
Chen, S., Cheng, A., Wang, M., 2013. Innate sensing of viruses by pattern recognition
receptors in birds. Vet. Res. 44, 1. doi:10.1186/1297-9716-44-82
120
Chiocca, S., Baker, A., Cotten, M., 1997. Identification of a novel antiapoptotic protein,
GAM-1, encoded by the CELO adenovirus. J. Virol. 71, 3168–3177.
Chiocca, S., Kurtev, V., Colombo, R., Boggio, R., Teresa Sciurpi, M., Brosch, G., Seiser,
C., Draetta, G.F., Cotten, M., 2002. Histone deacetylase 1 inactivation by an
adenovirus early gene product. Curr. Biol. 12, 594–598. doi:10.1016/S09609822(02)00720-0
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.
Chiou S.K., White, E., 1997. p300 binding by E1A cosegregates with p53 induction but
is dispensable for apoptosis. J. Virol. 71, 3515-3525.
Cho, D., Lee, W.J., Halloran, P.J., Trinchieri, G., Kim, Y.B., 1996. Enhancement of
porcine natural killer cell activity by recombinant human and murine IL-12. Cell.
Immunol. 172, 29–34.
Christensen, N.H., Saifuddin, M., 1989. A primary epidemic of inclusion body hepatitis
in broilers. Avian Dis. 33, 622-630.
Cichon, G., Boeckh-Herwig, S., Schmidt, H.H., Wehnes, E., Muller, T., Pring-Akerblom,
P., Burger, R., 2001. Complement activation by recombinant adenoviruses. Gene.
Ther. 8, 1794-1800.
Cody, J.J., Douglas, J.T., 2009. Armed replicating adenoviruses for cancer virotherapy.
Cancer Gene Ther. 16, 473–488.
Colombo, R., Boggio, R., Seiser, C., Draetta, G.F., Chiocca, S., 2002. The adenovirus
protein Gam1 interferes with sumoylation of histone deacetylase 1. EMBO Rep. 3,
1062–1068. doi:10.1093/embo-reports/kvf213
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.
Corredor, J.C., Garceac, A., Krell, P.J., Nagy, É., 2008. Sequence comparison of the right
end of fowl adenovirus genomes. Virus Genes. 36, 331–344. doi:10.1007/s11262007-0194-9
Corredor, J.C., Krell, P.J., Nagy, É., 2006. Sequence analysis of the left end of fowl
adenovirus genomes. Virus Genes. 33, 95–106. doi:10.1007/s11262-005-0031-y
121
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.
doi:10.1099/vir.0.013839-0
Corredor, J.C., Nagy, É., 2010b. The non-essential left end region of the fowl adenovirus
9 genome is suitable for foreign gene insertion/replacement. Virus Res. 149, 167–
174. doi:10.1016/j.virusres.2010.01.014
Corredor, J.C., Nagy, É., 2011. Antibody response and virus shedding of chickens
inoculated with left end deleted fowl adenovirus 9-based recombinant viruses. Avian
Dis. 55, 443-446. doi:10.1637/9819-971011-DIGEST.1
Cottone, R., Büttner, M., McInnes, C.J., Wood, A.R., Rziha, H.J., 2002. Orf virus
encodes a functional dUTPase gene. J. Gen. Virol. 83, 1043–1048.
Cowen, B., 1992. Inclusion body hepatitis-anaemia and hydropericardium syndrome:
aetiology and control. Worlds Poult. Sci. J. 48, 247-253.
Dar, A., Gomis, S., Shirley, I., Mutwiri, G., Brownlie, R., Potter, A., Gerdts, V., Tikoo,
S.K., 2012. Pathotypic and molecular characterization of a fowl adenovirus
associated with inclusion body hepatitis in Saskatchewan chickens. Avian Dis. 56,
73–81. doi:10.1637/9764-041911-Reg.1
Davison, A.J., Benkö, M., Harrach, B., 2003. Genetic content and evolution of
adenoviruses. J. Gen. Virol. 84, 2895-2908. doi:10.1099/vir.0.19497-0
Davison, A.J., Harrach, B., 2002. Siadenovirus. In: Tidona, C.A., Darai,G.,The Springer
index of viruses (2nd Edition). Springer –Verlag, New York, NY. 29–33.
Davison, A.J., Wright, K.M., Harrach, B., 2000. DNA sequence of frog adenovirus. J.
Gen. Virol. 81, 2431–2439.
Debbas, M., White, E., 1993. Wild-type p53 mediates apoptosis by E1A, which is
inhibited by E1B. Genes Dev. 7, 546-554.
Degen, W.G., van Daal, N., van Zuilekom, H.I., Burnside, J., Schijns, V.E., 2004.
Identification and molecular cloning of functional chicken IL-12. J. Immunol. 172,
4371–4380.
Deleu, L., Shellard, S., Alevizopoulos, K., Amati, B., Land, H., 2001. Recruitment of
TRRAP required for oncogenic transformation by E1A. Oncogene. 20, 8270-8275.
Deng, L., Sharif, S., Nagy, É., 2013. Oral inoculation of chickens with a candidate fowl
adenovirus 9 vector. Clin. Vaccine Immunol. 20, 1189–1196.
doi:10.1128/CVI.00187-13
122
Deryckere, F., Ebenau-Jehle, C., Wold, W.S., Burgert, H.G., 1995. Tumor necrosis factor
alpha increases expression of adenovirus E3 proteins. Immunobiology. 193, 186-192.
Doszpoly, A., Harrach, B., Benkö, M., 2009. Genome analysis of a fish adenovirus
confirms the proposal for a fifth adenovirus genus. Abstracts of the 9th international
adenovirus meeting, Dobogókő, Hungary, 26-30 April, p127.
Elsing, A., Burgert, H.G., 1998. The adenovirus E3/10.4K-14.5K proteins downmodulate the apoptosis receptor Fas/Apo-1 by inducing its internalization. Proc. Natl.
Acad. Sci. U.S.A. 95, 10072-10077.
Endharti, A.T., Rifa'l, M., Shi, Z., Fukuoka, Y., Nakahara, Y., Kawamoto, Y., Takeda, K.,
Isobe, K., Suzuki, H., 2005. Cutting edge: CD8+ CD122+ regulatory T cells produce
IL-10 to suppress IFN-gamma production and proliferation of CD8+ T cells. J.
Immunol. 175, 7093–7097.
Everett, R.D., Chiocca, S., Orr, A., 2014. The chicken adenovirus Gam1 protein, an
inhibitor of the sumoylation pathway, partially complements ICP0-null mutant
herpes simplex virus 1. J. Virol. 88, 5873–5876. doi:10.1128/JVI.00080-14
Fadly, A.M., Winterfield, R.W., Olander, H.J., 1976. Role of the bursa of Fabricius in the
pathogenicity of inclusion body hepatitis and infectious bursal disease viruses.
Avian Dis. 20, 467-472.
Fisher, F.B., Preston, V.G., 1986. Isolation and characterisation of herpes simplex virus
type 1 mutants which fail to induce dUTPase activity. Virology. 148, 190–197.
doi:10.1016/0042-6822(86)90414-9
Francois, A, Eterradossi, N., Delmas, B., Payet, V., Langlois, P., 2001. Construction of
avian adenovirus CELO recombinants in cosmids. J. Virol. 75, 5288–5301.
doi:10.1128/JVI.75.11.
Francois, A., Chevalier, C., Delmas, B., Eterradossi, N., Toquin, D., Rivallan, G.,
Langlois, P., 2004. Avian adenovirus CELO recombinants expressing VP2 of
infectious bursal disease virus induce protection against bursal disease in chickens.
Vaccine. 22, 2351–2360. doi:10.1016/j.vaccine.2003.10.039
Freeman, L., Buisson, M., Tarbouriech, N., Heyden, V. Der, Labbe, P., 2009. The
flexible motif v of Epstein-Barr virus deoxyuridine 5 -triphosphate pyrophosphatase
is essential for catalysis. J. Biol. Chem. 284, 25280–25289.
doi:10.1074/jbc.M109.019315
Fuchs, M., Gerber, J., Drapkin, R., Sif, S., Ikura, T., Ogryzko, V., Lane, W.S., Nakatani,
Y., Livingston, D.M., 2001. The p400 complex is an essential E1A transformation
target. Cell. 106, 297-307.
123
Gahery-Segard, H., Farace, F., Godfrin, D., Gaston, J., Lengagne, R., Tursz, T.,
Boulanger, P., Guillet, J.G., 1998. Immune response to recombinant capsid proteins
of adenovirus in humans: Antifiber and anti-penton base antibodies have a
synergistic effect on neutralizing activity. J. Virol. 72, 2388-2397.
Gallimore, P.H., Turnell, A.S., 2001. Adenovirus E1A: Remodelling the host cell, a life
or death experience. Oncogene. 20, 7824-7835.
Gallo, P., Dharmapuri, S., Cipriani, B., Monaci, P., 2005. Adenovirus as vehicle for
anticancer genetic immunotherapy. Gene Ther. 12, S84–S91.
Ganesh, K., Suryanarayana, V.V., Raghavan, R., 2002. Detection of fowl adenovirus
associated with hydropericardium hepatitis syndrome by a polymerase chain
reaction. Vet. Res. Commun. 26, 73-80.
Gelderblom, H., Maichle-Lauppe, I., 1982. The fibers of fowl adenoviruses. Arch. Virol.
72, 289–298. doi:10.1007/BF01315225
Glaser, R., Litsky, M.L., Padgett, D.A., Baiocchi, R.A., Yang, E.V., Chen, M., Yeh, P.E.,
Green-Church, K.B., Caligiuri, M.A., Williams, M.V., 2006. EBV-encoded
dUTPase induces immune dysregulation: Implications for the pathophysiology of
EBV-associated disease. Virology. 346, 205–218. doi:10.1016/j.virol.2005.10.034
Glotzer, J.B., Saltik, M., Chiocca, S., Michou, A.I., Moseley, P., Cotten, M., 2000.
Activation of heat-shock response by an adenovirus is essential for virus replication.
Nature. 407, 207–211. doi:10.1038/35025102
Gomis, S., Goodhope, A.R., Ojkić, A.D., Willson, P., 2006. Inclusion body hepatitis as a
primary disease in broilers in Saskatchewan, Canada. Avian Dis. 50, 550–555.
doi:10.1637/7767.1
Goodrum, F.D., Shenk, T., Ornelles, D.A., 1996. Adenovirus early region 4 34-kilodalton
protein directs the nuclear localization of the early region 1B 55-kilodalton protein
in primate cells. J. Virol. 70, 6323-6335.
Grand, R.J., Grant, M.L., Gallimore, P.H., 1994. Enhanced expression of p53 in human
cells infected with mutant adenoviruses. Virology. 203, 229-240.
Grgić, H., Krell, P.J., Nagy, É., 2014. Comparison of fiber gene sequences of inclusion
body hepatitis (IBH) and non-IBH strains of serotype 8 and 11 fowl adenoviruses.
Virus Genes. 48, 74–80. doi:10.1007/s11262-013-0995-y
Grgić, H., Philippe, C., Ojkić, D., Nagy, É., 2006. Study of vertical transmission of fowl
adenoviruses. Can. J. Vet. Res. 70, 230–233.
124
Grgić, H., Poljak, Z., Sharif, S., Nagy, É., 2013a. Pathogenicity and cytokine gene
expression pattern of a serotype 4 fowl adenovirus isolate. PLoS One. 8.
doi:10.1371/journal.pone.0077601
Grgić, H., Sharif, S., Haghighi, H.R., Nagy, É., 2013b. Cytokine patterns associated with
a serotype 8 fowl adenovirus infection. Viral Immunol. 26, 143–149.
doi:10.1089/vim.2012.0078
Grgić, H., Yang, D.H., Nagy, É., 2011. Pathogenicity and complete genome sequence of
a fowl adenovirus serotype 8 isolate. Virus Res. 156, 91–97.
doi:10.1016/j.virusres.2011.01.002
Griffin, B.D., Nagy, É., 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. doi:10.1099/vir.0.030064-0
Hacker, D.L., Derow, E., Wurm, F.M., 2005. The CELO adenovirus Gam1 protein
enhances transient and stable recombinant protein expression in Chinese hamster
ovary cells. J. Biotechnol. 117, 21–29. doi:10.1016/j.jbiotec.2005.01.006
Halbert, D.N., Cutt, J.R., Shenk, T., 1985. Adenovirus early region 4 encodes functions
required for efficient DNA replication, late gene expression, and host cell shutoff. J.
Virol. 56, 250-257.
Hale, B.G., Randall, R.E., Ortin, J., Jackson, D., 2008. The multifunctional NS1 protein
of influenza A viruses. J. Gen. Virol. 89, 2359-2376.
doi:10.1099/vir.0.2008/004606-0
Hamamori, Y., Sartorelli, V., Ogryzko, V., Puri, P. L., Wu, H.Y., Wang, J.Y., Nakatani,
Y., Kedes, L. 1999. Regulation of histone acetyltransferases p300 and PCAF by the
bHLH protein twist and adenoviral oncoprotein E1A. Cell. 96, 405-413.
Harrach, B., Benkö, M., Both, G.W., Brown, M., Davison, A.J., Echavarria, M., Hess, M.,
Jones, M.S., Kajon, A., Lehmkuhl, H.D., Mautner, V., Mittal, S.K., Wadell, G. 2011.
Family Adenoviridae. In: King, A.M.Q., Adams, M.J., Carstens, E.B., Lefkowitz,
E.J., Virus taxonomy: classification and nomenclature of viruses: ninth report of the
International Committee on Taxonomy of Viruses. Elsevier Academic Press. San
Diego, CA. 125–141.
Harris, J.M., McIntosh, E.M., Muscat, G.E., 1999. Structure/function analysis of a
dUTPase: catalytic mechanism of a potential chemotherapeutic target. J. Mol. Biol.
288, 275–287. doi:10.1006/jmbi.1999.2680
Helmboldt, C.F., Frazier, M.N., 1963. Avian hepatic inclusion bodies of unknown
significance. Avian Dis. 7, 446–450.
125
Hess, M. 2013. Aviadenovirus infections. In: Swayne, D.E., Glisson, J.R., McDougald,
L.R., Nolan, L.K., Suarez, D.L., Nair, V., Diseases of Poultry (13th Edition). John
Wiley & Sons, Inc. Ames, IA. 289-300.
Hendrickx, R., Stichling, N., Koelen, J., Kuryk, L., Lipiec, A., Greber, U.F., 2014. Innate
immunity to adenovirus. Hum. Gene Ther. 25, 265–84. doi:10.1089/hum.2014.001
Hess, M., 2000. Detection and differentiation of avian adenoviruses: a review. Avian
Pathol. 29, 195–206. doi:10.1080/03079450050045440
Hess, M., Raue, R., Prusas, C., 1999. Epidemiological studies on fowl adenoviruses
isolated from cases of infectious hydropericardium. Avian Pathol. 28, 433-439.
Horwitz, M.S., 2004. Function of adenovirus E3 proteins and their interactions with
immunoregulatory cell proteins. J. Gene Med. Suppl 1, S172-183.
doi:10.1002/jgm.495
Hussain, I., Mahmood, M.S., Arshad, M.I., 2012. Immune system dysfunction in broiler
chickens experimentally inoculated with fowl adenovirus serotype-4 associated with
inclusion body hepatitis hydropericardium syndrome. Turk. J. Vet. Anim. Sci. 36,
223–230. doi:10.3906/vet-0807-21
Jiang, P., Ojkić, D., Tuboly, T., Huber, P., Nagy, É., 1999. Application of the polymerase
chain reaction to detect fowl adenoviruses. Can. J. Vet. Res. 63, 124–128.
Johnson, M.A., Pooley, C., Ignjatovic, J., Tyack, S.G., 2003. A recombinant fowl
adenovirus expressing the S1 gene of infectious bronchitis virus protects against
challenge with infectious bronchitis virus. Vaccine. 21, 2730–2736.
doi:10.1016/S0264-410X(03)00227-5
Junnu, S., Lertwatcharasarakul, P., Jala, S., Phattanakunanan, S., Moonjit, P., Songserm,
T., 2014. Developing an indirect ELISA based on recombinant hexon protein for
serological detection of inclusion body hepatitis in chickens. J. Vet. Med. Sci. 76,
289-293.
Juul-Madsen, H.R., Viertlböeck, B., Härtle, S., Smith, A.L., Göbel, T.W., 2014. Innate
immune responses. In: Schat, K.A., Kaspers, B., Kaiser, P., Avian immunology (2nd
Edition). Elsevier Academic Press. San Diego, CA. 121-147.
Kaján, G.L., Davison, A.J., Palya, V., Harrach, B., Benkö, M., 2012. Genome sequence
of a waterfowl aviadenovirus, goose adenovirus 4. J. Gen. Virol. 93, 2457–2465.
doi:10.1099/vir.0.042028-0
Kaján, G.L., Stefancsik, R., Ursu, K., Palya, V., Benkö, M., 2010. The first complete
genome sequence of a non-chicken aviadenovirus, proposed to be turkey adenovirus
1. Virus Res. 153, 226–233. doi:10.1016/j.virusres.2010.08.006
126
Kato, T., Hakamada, R., Yamane, H., Nariuchi, H., 1996. Induction of IL-12 p40
messenger RNA expression and IL-12 production of macrophages via CD40-CD40
ligand interaction. J. Immunol. 156, 3932–3938.
Kawai, T., Akira, S., 2006. Innate immune recognition of viral infection. Nat. Immunol. 7,
131–137. doi:10.1038/ni1303
Keestra, A.M., de Zoete, M.R., Bouwman, L.I., van Putten, J.P., 2010. Chicken TLR21 is
an innate CpG DNA receptor distinct from mammalian TLR9. J. Immunol. 185,
460–467.
Kiang, A., Hartman, Z.C., Everett, R.S., Serra, D., Jiang, H., Frank, M.M., Amalfitano,
A., 2006. Multiple innate inflammatory responses induced after systemic adenovirus
vector delivery depend on a functional complement system. Mol. Ther. 14, 588-598.
Kim, J.N., Byun, S.H., Kim, M.J., Kim, J.J., Sung, H.W., Mo, I.P., 2008. Outbreaks of
hydropericardium syndrome and molecular characterization of Korean fowl
adenoviral isolates. Avian Dis. 52, 526–530. doi:10.1637/8178-112207-Case
Kim, M.S., Lim, T.H., Lee, D.H., Youn, H.N., Yuk, S.S., Kim, B.Y., Choi, S.W., Jung,
C.H., Han, J.H., Song, C.S., 2014. An inactivated oil-emulsion fowl Adenovirus
serotype 4 vaccine provides broad cross-protection against various serotypes of fowl
Adenovirus. Vaccine. 32, 3564–3568. doi:10.1016/j.vaccine.2014.03.015
Konig, C., Roth, J., Dobbelstein, M., 1999. Adenovirus type 5 E4orf3 protein relieves
p53 inhibition by E1B-55-kilodalton protein. J. Virol. 73, 2253-2262.
Krug, A., Luker, G.D., Barchet, W., Leib, D.A., Akira, S., Colonna, M., 2004. Herpes
simplex virus type 1 activates murine natural interferon-producing cells through tolllike receptor 9. Blood. 103, 1433–1437.
Lang, S.E., Hearing, P., 2003. The adenovirus E1A oncoprotein recruits the cellular
TRRAP/GCN5 histone acetyltransferase complex. Oncogene. 22, 2836-2841.
Lasaro, M.O., Ertl, H.C.J., 2009. New insights on adenovirus as vaccine vectors. Mol.
Ther. 17, 1333–1339. doi:10.1038/mt.2009.130
Leang, R.S., Wu, T.T., Hwang, S., Liang, L.T., Tong, L., Truong, J.T., Sun, R., 2011.
The anti-interferon activity of conserved viral dUTPase ORF54 is essential for an
effective MHV-68 infection. PLoS Pathog. 7, e1002292.
doi:10.1371/journal.ppat.1002292
Leen, A.M., Christin, A., Khalil, M., Weiss, H., Gee, A.P., Brenner, M.K., Heslop, H.E.,
Rooney, C.M., Bollard, C.M. 2008. Identification of hexon-specific CD4 and CD8
T-cell epitopes for vaccine and immunotherapy. J. Virol. 82, 546-554.
127
Leen, A.M., Sili, U., Vanin, E.F., Jewell, A.M., Xie, W., Vignali, D., Piedra, P.A.,
Brenner, M.K., Rooney, C.M. 2004. Conserved CTL epitopes on the adenovirus
hexon protein expand subgroup cross-reactive and subgroup-specific CD8+ T cells.
Blood. 104, 2432-2440.
Lehrmann, H., Cotten, M., 1999. Characterization of CELO virus proteins that modulate
the pRb/E2F pathway. J. Virol. 73, 6517–6525.
Leppard, K.N., Everett, R.D., 1999. The adenovirus type 5 E1b 55K and E4 Orf3 proteins
associate in infected cells and affect ND10 components. J. Gen. Virol. 80, 997-1008.
Li, X.D., Sun, L., Seth, R.B., Pineda, G., Chen, Z.J., 2005. Hepatitis C virus protease
NS3/4A cleaves mitochondrial antiviral signaling protein off the mitochondria to
evade innate immunity. Proc. Natl. Acad. Sci. U. S. A. 102, 17717-17722.
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 NFkappaB activity and as a target of an adenovirus inhibitor of tumor necrosis factor
alpha-induced apoptosis. Proc. Natl. Acad. Sci. U. S. A. 96, 1042–1047. doi:DOI
10.1073/pnas.96.3.1042
Lieber, A., He, C.Y., Meuse, L., Schowalter, D., Kirillova, I., Winther, B., Kay, M.A.,
1997. The role of Kupffer cell activation and viral gene expression in early liver
toxicity after infusion of recombinant adenovirus vectors. J Virol. 71, 8798–8807.
Liu, Q., Muruve, D.A., 2003. Molecular basis of the inflammatory response to adenovirus
vectors. Gene Ther. 10, 935-940.
Liu, X., Yang, F., 2005. Identification and function of a shrimp white spot syndrome
virus (WSSV) gene that encodes a dUTPase. Virus Res. 110, 21–30.
doi:10.1016/j.virusres.2005.01.003
Lund, J., Sato, A., Akira, S., Medzhitov, R., Iwasaki, A., 2003. Toll-like receptor 9mediated recognition of Herpes simplex virus-2 by plasma cytoid dendritic cells. J.
Exp. Med. 198, 513–520.
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.
MacLennan, I.C., Gulbranson-Judge, A., Toellner, K.M., Casamayor-Palleja, M., Chan,
E., Sze, D.M., Luther, S.A., Orbea, H.A. 1997. The changing preference of T and B
cells for partners as T-dependent antibody responses develop. Immunol. Rev. 156,
53-66.
Marek, A., Ballmann, M.Z., Kosiol, C., Harrach, B., Schlötterer, C., Hess, M., 2014a.
Whole-genome sequences of two turkey adenovirus types reveal the existence of
128
two unknown lineages that merit the establishment of novel species within the genus
Aviadenovirus. J. Gen. Virol. 95, 156–170. doi:10.1099/vir.0.057711-0
Marek, A., Kaján, G.L., Kosiol, C., Harrach, B., Schlötterer, C., Hess, M., 2014b.
Complete genome sequences of pigeon adenovirus 1 and duck adenovirus 2 extend
the number of species within the genus Aviadenovirus. Virology. 462-463, 107–114.
doi:10.1016/j.virol.2014.04.033
Marek, A., Kosiol, C., Harrach, B., Kaján, G.L., Schlötterer, C., Hess, M., 2013. The first
whole genome sequence of a Fowl adenovirus B strain enables interspecies
comparisons within the genus Aviadenovirus. Vet. Microbiol. 166, 250–256.
doi:10.1016/j.vetmic.2013.05.017
Marek, A., Nolte, V., Schachner, A., Berger, E., Schlötterer, C., Hess, M., 2012. Two
fiber genes of nearly equal lengths are a common and distinctive feature of Fowl
adenovirus C members. Vet. Microbiol. 156, 411–417.
doi:10.1016/j.vetmic.2011.11.003
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. 22, 937–941.
doi:10.1177/104063871002200613
Marshall, L.J., Moore, A.C., Ohki, M., Kitabayashi, I., Patterson, D., Ornelles, D.A.,
2008. RUNX1 permits E4orf6-directed nuclear localization of the adenovirus E1B55K protein and associates with centers of viral DNA and RNA synthesis. J. Virol.
82, 6395-6408.
Mcgeehan, J.E., Depledge, N.W., Mcgeoch, D.J., 2001. Evolution of the dUTPase gene
of mammalian and avian herpesviruses. Curr. Protein Pept. Sci. 44, 325–333.
McGeoch, D.J., 1990. Protein sequence comparisons show that the “pseudoproteases”
encoded by poxviruses and certain retroviruses belong to the deoxyuridine
triphosphatase family. Nucleic Acids Res. 18, 4105–4110.
Medina-Kauwe, L.K., 2013. Development of adenovirus capsid proteins for targeted
therapeutic delivery. Ther Deliv. 4, 267–277. doi:10.4155/tde.12.155
Meulemans, G., Couvreur, B., Decaesstecker, M., Boschmans, M., van den Berg, T.P.,
2004. Phylogenetic analysis of fowl adenoviruses. Avian Pathol. 33, 164–170.
doi:10.1080/03079450310001652086
Mittal, D., Jindal, N., Tiwari, A.K., Khokhar, R.S., 2014. Characterization of fowl
adenoviruses associated with hydropericardium syndrome and inclusion body
hepatitis in broiler chickens. Indian J. Virol. 25, 114–119. doi:10.1007/s13337-0130183-7
129
Muruve, D.A., 2004. The innate immune response to adenovirus vectors. Hum. Gene
Ther. 15, 1157-1166.
Myhre, S., Henning, P., Granio, O., Tylo, A.S., Nygren, P.A., Olofsson, S., Boulanger, P.,
Lindholm, L., Hong, S.S. 2007. Decreased immune reactivity towards a knobless,
affibody-targeted adenovirus type 5 vector. Gene Ther. 14, 376-381.
Mymryk, J.S., Shire, K., Bayley, S.T., 1994. Induction of apoptosis by adenovirus type 5
E1A in rat cells requires a proliferation block. Oncogene. 9, 1187–1193.
Nakamura, K., Mase, M., Yamamoto, Y., Takizawa, K., Kabeya, M., Wakuda, T.,
Matsuda, M., Chikuba, T., Yamamoto, Y., Ohyama, T., Sato, N., Akiyama, N.,
Honma, H., Imai, K., Mase, A.B.M., Yamamoto, A.Y., Takizawa, A.K., Kabeya,
C.M., Wakuda, D.T., Chikuba, F.T., Yamamoto, G.Y., Ohyama, H.T., Takahashi,
I.K., Sato, J.N., 2011. Inclusion body hepatitis caused by fowl adenovirus in broiler
chickens in case report - inclusion body hepatitis caused by fowl adenovirus in
broiler chickens in Japan 2009–2010. Avian Dis. 55, 719-723.
Nayak, S., Herzog, R.W., 2009. Progress and prospects: immune responses to viral
vectors. Gene Ther. 17, 295–304. doi:10.1038/gt.2009.148
Nemunaitis, J., Cunningham, C., Buchanan, A., Blackburn, A., Edelman, G., Maples, P.,
Netto, G., Tong, A., Randlev, B., Olson, S., Kirn, D., 2001. Intravenous infusion of
a replication-selective adenovirus (ONYX-015) in cancer patients: Safety, feasibility
and biological activity. Gene Ther. 8, 746-759.
Nikiforov, M.A., Chandriani, S., Park, J., Kotenko, I., Matheos, D., Johnsson, A.,
McMahon, S.B., Cole, M.D., 2002. TRRAP-dependent and TRRAP-independent
transcriptional activation by myc family oncoproteins. Mol. Cell. Biol. 22, 50545063.
Nociari, M., Ocheretina, O., Murphy, M., Falck-Pedersen, E., 2009. Adenovirus
induction of IRF3 occurs through a binary trigger targeting jun N-terminal kinase
and TBK1 kinase cascades and type I interferon autocrine signaling. J. Virol. 83,
4081-4091.
Nociari, M., Ocheretina, O., Schoggins, J.W., Falck-Pedersen, E., 2007. Sensing infection
by adenovirus: Toll-like receptor-independent viral DNA recognition signals
activation of the interferon regulatory factor 3 master regulator. J. Virol. 81, 41454157.
Ojkić, D., Krell, P.J., Nagy, É., Nagy, E., 2002. Unique Features of Fowl Adenovirus 9
Gene Transcription. Virology. 302, 274–285. doi:10.1006/viro.2002.1583
Ojkić, 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.
130
Ojkić, D., Martin, E., Swinton, J., Vaillancourt, J.P., Boulianne, M., Gomis, S., 2008b.
Genotyping of Canadian isolates of fowl adenoviruses. Avian Pathol. 37, 95–100.
doi:10.1080/03079450701805324
Ojkić, D., Nagy, É., 2000. The complete nucleotide sequence of fowl adenovirus type 8. J.
Gen. Virol. 81, 1833–1837.
Ojkić, D., Nagy, É., 2001. The long repeat region is dispensable for fowl adenovirus
replication in vitro. Virology. 283, 197–206. doi:10.1006/viro.2000.0890
Ojkić, D., Nagy, É., 2003. Antibody response and virus tissue distribution in chickens
inoculated with wild-type and recombinant fowl adenoviruses. Vaccine. 22, 42–48.
doi:10.1016/S0264-410X(03)00544-9
Oliveros, M., García-Escudero, R., Alejo, A., Viñuela, E., Salas, M.L., Salas, J., 1999.
African swine fever virus dUTPase is a highly specific enzyme required for efficient
replication in swine macrophages. J. Virol. 73, 8934–8943.
O'Neill, L.A. , Bowie, A.G., 2010. Sensing and signaling in antiviral innate immunity.
Curr. Biol. 20, R328-333.
Ornelles, D.A., Shenk, T., 1991. Localization of the adenovirus early region 1B 55kilodalton protein during lytic infection: Association with nuclear viral inclusions
requires the early region 4 34-kilodalton protein. J. Virol. 65, 424-429.
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.
Payet, V., Arnauld, C., Picault, J.P., Jestin, A., Langlois, P., 1998. Transcriptional
organization of the avian adenovirus CELO. J. Virol. 72, 9278–9285.
Payne, S., Elder, J., 2001. The role of retroviral dUTPases in replication and virulence.
Curr. Protein Pept. Sci. 2, 381–388. doi:10.2174/1389203013381008
Philippe, C., Grgić, H., Nagy, É., 2005. Inclusion body hepatitis in young broiler breeders
associated with a serotype 2 adenovirus in Ontario, Canada. J. Appl. Poult. Res. 14,
588–593.
Philippe, C., Grgić, H., Ojkić, D., Nagy, É., 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.
Philpott N.J., Nociari M., Elkon K.B., Falck-Pedersen, E., 2004. Adenovirus-induced
maturation of dendritic cells through a PI3 kinase-mediated TNF-alpha induction
pathway. Proc. Natl. Acad. Sci. U. S. A. 101, 6200-6205.
131
Prichard, M.N., Kern, E.R., Quenelle, D.C., Keith, K.A., Moyer, R.W., Turner, P.C.,
2008. Vaccinia virus lacking the deoxyuridine triphosphatase gene (F2L) replicates
well in vitro and in vivo, but is hypersensitive to the antiviral drug (N)methanocarbathymidine. Virol. J. 5, 39. doi:10.1186/1743-422X-5-39
Pyles, R.B., Sawtell, N.M., Thompson, R.L., 1992. Herpes simplex virus type 1 dUTPase
mutants are attenuated for neurovirulence, neuroinvasiveness, and reactivation from
latency. J. Virol. 66, 6706–6713.
Randall, R.E., Goodbourn, S., 2008. Interferons and viruses: An interplay between
induction, signalling, antiviral responses and virus countermeasures. J. Gen. Virol.
89, 1-47.
Raue, R., Gerlach, H., Müller, H., 2005. Phylogenetic analysis of the hexon loop 1 region
of an adenovirus from psittacine birds supports the existence of a new psittacine
adenovirus (PsAdV). Arch. Virol. 150, 1933–1943. doi:10.1007/s00705-005-0578-x
Reid, T., Galanis, E., Abbruzzese, J., Sze, D., Wein, L.M., Andrews, J., Randlev, B.,
Heise, C., Uprichard, M., 2002. Hepatic arterial infusion of a replication-selective
oncolytic adenovirus (dl1520): Phase II viral, immunologic, and clinical endpoints.
Cancer Res. 62, 6070-6079.
Roberts, D.M., Nanda, A., Havenga, M.J., Abbink, P., Lynch, D.M., Ewald, B.A., Liu, J.,
Thorner, A.R., Swanson, P.E., 2006. Hexon-chimaeric adenovirus serotype 5 vectors
circumvent pre-existing anti-vector immunity. Nature. 441, 239-243.
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.
doi:10.1016/j.jviromet.2009.02.026
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.
doi:10.1016/j.it.2004.06.005
Rowe, W.P., Huebner, R.J., Gilmore, L.K., Parrott, R.H., Ward, T.G., 1953. Isolation of a
cytopathogenic agent from human adenoids undergoing spontaneous degeneration in
tissue culture. Proc. Soc. Exp. Biol. Med. 84, 570–573. doi:10.3181/00379727-8420714
Russell, W.C., 2009. Adenoviruses: Update on structure and function. J. Gen. Virol. 90,
1-20. doi:10.1099/vir.0.003087-0
Rychert, J., Strick, D., Bazner, S., Robinson, J., Rosenberg, E., 2010. Detection of HIV
gp120 in plasma during early HIV infection is associated with increased
132
proinflammatory and immunoregulatory cytokines. AIDS Res. Hum. Retroviruses.
26, 1139–1145. doi:10.1089/aid.2009.0290
Sarkioja M., Pesonen S., Raki M., Hakkarainen T., Salo J., Ahonen M. T., Kanerva A.,
Hemminki A. 2008. Changing the adenovirus fiber for retaining gene delivery
efficacy in the presence of neutralizing antibodies. Gene Ther. 15, 921-929.
Schachner, A., Marek, A., Jaskulska, B., Bilic, I., Hess, M., 2014. Recombinant FAdV-4
fiber-2 protein protects chickens against hepatitis-hydropericardium syndrome
(HHS). Vaccine. 32, 1086–1092. doi:10.1016/j.vaccine.2013.12.056
Schat, K.A., Purchase, H.G., 1989. Cell culture methods. In: Purchase, H.G., Arp, L.H.,
Domermuth, C.H., Pearson, J.E., A laboratory manual for the isolation and
identification of avian pathogens (3rd Edition). Kendall/Hunt Publishing Company.
Dubuque, IA. 167-175.
Schonewille, E., Singh, A., Göbel, T.W., Gerner, W., Saalmüller, A., Hess, M., 2008.
Fowl adenovirus (FAdV) serotype 4 causes depletion of B and T cells in lymphoid
organs in specific pathogen-free chickens following experimental infection. Vet.
Immunol. Immunopathol. 121, 130–139. doi:10.1016/j.vetimm.2007.09.017
Shao, H., Robek, M.D., Threadgill, D.S., Mankowski, L.S., Cameron, C.E., Fuller, F.J.,
Payne, S.L., 1997. Characterization and mutational studies of equine infectious
anemia virus dUTPase. Biochim. Biophys. Acta - Protein Struct. Mol. Enzymol.
1339, 181–191. doi:10.1016/S0167-4838(96)00229-4
Sharma, A., Tandon, M., Ahi, Y.S., Bangari, D.S., Vemulapalli, R., Mittal, S.K., 2010.
Evaluation of cross-reactive cell-mediated immune responses among human, bovine
and porcine adenoviruses. Gene Ther. 17, 634–642. doi:10.1038/gt.2010.1
Shashkova, E. V, Cherenova, L. V, Kazansky, D.B., Doronin, K., 2005. Avian
adenovirus vector CELO-TK displays anticancer activity in human cancer cells and
suppresses established murine melanoma tumors. Cancer Gene. Ther. 12, 617–626.
doi:10.1038/sj.cgt.7700822
Sheppard, M., Werner, W., Tsatas, E., McCoy, R., Prowse, S., Johnson, M., 1998. Fowl
adenovirus recombinant expressing VP2 of infectious bursal disease virus induces
protective immunity against bursal disease. Arch. Virol. 143, 915–930.
doi:10.1007/s007050050342
Shivachandra, S.B., Sah, R.L., Singh, S.D., Kataria, J.M., Manimaran, K., 2003.
Pathogenesis of FAV serotype–4 induced hydropericardium syndrome in broilers.
Ind. J. Vet. Pathol. 27, 1–4.
133
Singh, A., Grewal, G.S., Maiti, N.K., Oberoi, M.S., 2006. Effect of fowl adenovirus-1
(IBH isolate) on humoral and cellular immune competency of broiler chicks. Comp.
Immunol. Microbiol. Infect. Dis. 29, 315–321. doi:10.1016/j.cimid.2006.08.001
Sirena, D., Lilienfeld, B., Eisenhut, M., Kälin, S., Boucke, K., Beerli, R.R., Vogt, L.,
Ruedl, C., Bachmann, M.F., Greber, U.F., Hemmi, S., 2004. The human membrane
cofactor CD46 is a receptor for species B adenovirus serotype 3. J. Virol. 78, 4454–
4462. doi:10.1128/JVI.78.9.4454-4462.2004
Spear, G.T., Hart, M., Olinger, G.G., Hashemi, F.B., Saifuddin, M., 2001. The role of the
complement system in virus infections. Curr. Top. Microbiol. Immunol. 260, 229245.
Steer, P.A., Kirkpatrick, N.C., O’Rourke, D., Noormohammadi, 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.
doi:10.1128/JCM.01567-08
Steer, P.A., O’Rourke, D., Ghorashi, S.A., Noormohammadi, A.H., 2011. Application of
high-resolution melting curve analysis for typing of fowl adenoviruses in field cases
of inclusion body hepatitis. Aust. Vet. J. 89, 184–192. doi:10.1111/j.17510813.2011.00695.x
Steitz, J., Wagner, R.A., Bristol, T., Gao, W., Donis, R.O., Gambotto, A., 2010.
Assessment of route of administration and dose escalation for an adenovirus-based
influenza A virus (H5N1) vaccine in chickens. Clin. Vaccine Immunol. 17, 1467–
1472. doi:10.1128/CVI.00180-10
Swanson, C.L., Wilson, T.J., Strauch, P., Colonna, M., Pelanda, R., Torres, R.M., 2010.
Type I IFN enhances follicular B cell contribution to the T cell-independent
antibody response. J. Exp. Med. 207, 1485–1500. doi:10.1084/jem.20092695
Takaoka, A., Yanai, H., 2006. Interferon signalling network in innate defence. Cell.
Microbiol. 8, 907–922. doi:10.1111/j.1462-5822.2006.00716.x
Takeda, K., Akira, S., 2005. Toll-like receptors in innate immunity. Int. Immunol. 17, 114.
Taniguchi, T., Takaoka, A., 2001. A weak signal for strong responses: interferon-α/β. Nat.
Rev. Cell Biol. 2, 378–386.
Tarbouriech, N., Buisson, M., Seigneurin, J.M., Cusack, S., Burmeister, W.P., 2005. The
monomeric dUTPase from Epstein-Barr virus mimics trimeric dUTPases. Structure.
13, 1299–1310. doi:10.1016/j.str.2005.06.009
134
Tatsis, N., Ertl, H.C., 2004. Adenoviruses as vaccine vectors. Mol. Ther. 10, 616–629.
doi:10.1016/j.ymthe.2004.07.013
Thacker, E.E., Timares, L., Matthews, Q.L., 2009. Strategies to overcome host immunity
to adenovirus vectors in vaccine development. Expert Rev. Vaccines. 8, 761–777.
doi:10.1586/erv.09.29
Theofilopoulos, A.N., Baccala, R., Beutler, B., Kono, D.H., 2005. Type I interferons
(alpha/beta) in immunity and autoimmunity. Annu. Rev. Immunol. 23, 307–336.
doi:10.1146/annurev.immunol.23.021704.115843
Threadgill, D.S., Steagall, W.K., Flaherty, M.T., Fuller, F.J., Perry, S.T., Rushlow, K.E.,
Grice, S.F.J.L.E., Payne, S.L., 1993. Characterization of equine infectious anemia
virus dUTPase : growth properties of a dUTPase-deficient mutant. J. Virol. 67,
2592–2600.
Tollefson, A.E., Hermiston, T.W., Lichtenstein, 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.
Tordo, N., Foumier, A., Jallet, C., Szelechowski, M., Klonjkowski, B., Eloit, M., 2008.
Canine adenovirus based rabies vaccines. Dev. Biol. (Basel). 131, 467–476.
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.
Tribouley, C., Lutz, P., Staub, A., Kedinger, C., 1994. The product of the adenovirus
intermediate gene IVa2 is a transcriptional activator of the major late promoter. J.
Virol. 68, 4450–4457.
Trinchieri, G., 2003. Interleukin-12 and the regulation of innate resistance and adaptive
immunity. Nat. Rev. Immunol. 3, 133–146. doi:10.1038/nri1001
Tucker, S.N., Tingley, D.W., Scallan, C.D., 2008. Oral adenoviral-based vaccines:
historical perspective and future opportunity. Expert Rev. Vaccines. 7, 25–31.
doi:10.1586/14760584.7.1.25
van der Vliet, P.C., Levine, A.J., 1973. DNA-binding proteins specific for cells infected
by adenovirus. Nat. New. Biol. 246, 170-174.
Vellinga, J., Van der Heijdt, S., Hoeben, R.C., 2005. The adenovirus capsid: Major
progress in minor proteins. J. Gen. Virol. 86, 1581-1588. doi:10.1099/vir.0.80877-0
Vilcek, J., 2003. Novel interferons. Nat. Immunol. 4, 8-9. doi:10.1038/ni0103-8
135
Virtanen, A., Pettersson, U., 1985. Organization of early region 1B of human adenovirus
type 2: Identification of four differentially spliced mRNAs. J. Virol. 54, 383-391.
Voronin, N., Herzig, E., Hizi, A., 2014. The dUTPase-related gene of bovine
immunodeficiency virus is critical for viral replication, despite the lack of dUTPase
activity of the encoded protein. Retrovirology. 11, 60. doi:10.1186/1742-4690-11-60
Waldman, W.J., Williams, M. V., Lemeshow, S., Binkley, P., Guttridge, D., KiecoltGlaser, J.K., Knight, D.A., Ladner, K.J., Glaser, R., 2008. Epstein-Barr virusencoded dUTPase enhances proinflammatory cytokine production by macrophages
in contact with endothelial cells: Evidence for depression-induced atherosclerotic
risk. Brain. Behav. Immun. 22, 215–223. doi:10.1016/j.bbi.2007.07.007
Washietl, S., Eisenhaber, F., 2003. Reannotation of the CELO genome characterizes a set
of previously unassigned open reading frames and points to novel modes of host
interaction in avian adenoviruses. BMC Bioinformatics. 4, 55. doi:10.1186/14712105-4-55
Weber, F., Kochs, G., Haller, O., 2004. Inverse interference: how viruses fight the
interferon system. Viral Immunol. 17, 498–515. doi:10.1089/vim.2004.17.498
Weiss, R.S., Lee, S.S., Prasad, B. V, Javier, R.T., 1997. Human adenovirus early region 4
open reading frame 1 genes encode growth-transforming proteins that may be
distantly related to dUTP pyrophosphatase enzymes. J. Virol. 71, 1857–1870.
Weitzman, M.D., 2005. Functions of the adenovirus E4 proteins and their impact on viral
vectors. Front. Biosci. 10, 1106–1117. doi:10.2741/1604
Woods, D.B., Vousden, K.H., 2001. Regulation of p53 function. Exp. Cell. Res. 264, 5666.
Xiang, Z.Q., Gao, G.P., Reyes-Sandoval, A., Li, Y., Wilson, J.M., Ertl, H.C.J., 2003.
Oral vaccination of mice with adenoviral vectors is not impaired by preexisting
immunity to the vaccine carrier. J. Virol. 77, 10780–10789.
doi:10.1128/JVI.77.20.10780-10789.2003
Xiang, Z.Q., Greenberg, L., Ertl, H.C., Rupprecht, C.E., 2014. Protection of non-human
primates against rabies with an adenovirus recombinant vaccine. Virology. 450,
243–249. doi:10.1016/j.virol.2013.12.029
Xie, Z., Luo, S., Fan, Q., Xie, L., Liu, J., Xie, Z., Pang, Y., Deng, X., Wang, X., 2013.
Detection of antibodies specific to the non-structural proteins of fowl adenoviruses
in infected chickens but not in vaccinated chickens. Avian Pathol. 42, 491–496.
doi:10.1080/03079457.2013.829553
136
Yamaguchi, T., Kawabata, K., Koizumi, N., Sakurai, F., Nakashima, K., Sakurai, H.,
Sasaki, T., Okada, N., Yamanishi, K., Mizuguchi, H., 2007. Role of MyD88 and
TLR9 in the innate immune response elicited by serotype 5 adenoviral vectors. Hum.
Gene. Ther. 18, 753-762.
Yamaguchi, T., Kawabata, K., Kouyama, E., Ishii, K.J., Katayama, K., Suzuki, T.,
Kurachi, S., Sakurai, F., Akira, S., Mizuguchi, H., 2010. Induction of type I
interferon by adenovirus-encoded small RNAs. Proc. Natl. Acad. Sci. U. S. A. 107,
17286–17291. doi:10.1073/pnas.1009823107
Yamamoto, M., Curiel, D.T., 2010. Current issues and future directions of oncolytic
adenoviruses. Mol. Ther. 18, 243–250. doi:10.1038/mt.2009.266
Yang, Y., Greenough, K., Wilson, J.M., 1996. Transient immune blockade prevents
formation of neutralizing antibody to recombinant adenovirus and allows repeated
gene transfer to mouse liver. Gene Ther. 3, 412–420.
Yang, Y., Li, Q., Ertl, H.C., Wilson, J.M., 1995. Cellular and humoral immune responses
to viral antigens create barriers to lung-directed gene therapy with recombinant
adenoviruses. J. Virol. 69, 2004–2015.
Zadravec, M., Brigita, S., Krapež, U., Kaján, G.L., Račnik, J., Juntes, P., Juršič-Cizerl,
R., Benkö, M., Zorman-Rojs, O., 2013. Inclusion body hepatitis (IBH) outbreak
associated with fowl adenovirus type 8b in broilers. Acta veterinaria. 63, 101-110.
doi:10.2298/AVB1301101Z
Zantema, A., Schrier, P.I., Davis-Olivier, A., van Laar, T., Vaessen, R.T., van der EB,
A.J., 1985. Adenovirus serotype determines association and localization of the large
E1B tumor antigen with cellular tumor antigen p53 in transformed cells. Mol. Cell.
Biol. 5, 3084-3091.
Zhang, X., Turnell, A.S., Gorbea, C., Mymryk, J.S., Gallimore, P.H., Grand, R.J., 2004.
The targeting of the proteasomal regulatory subunit S2 by adenovirus E1A causes
inhibition of proteasomal activity and increased p53 expression. J. Biol. Chem. 279,
25122-25133.
Zhang, Y., Moriyama, H., Homma, K., Van Etten, J.L., 2005. Chlorella virus-encoded
deoxyuridine triphosphatases exhibit different temperature optima. J. Virol. 79,
9945–9953. doi:10.1128/JVI.79.15.9945-9953.2005
Zhao, L., Cheng, A., Wang, M., Yuan, G., Jia, R., Zhou, D., Qi, X., Ge, H., Sun, T., 2008.
Identification and characterization of duck enteritis virus dUTPase gene. Avian Dis.
52, 324–331. doi:10.1637/8169-110607-ResNote.1
137
Zhao, Z., Ke, F., Gui, J., Zhang, Q., 2007. Characterization of an early gene encoding for
dUTPase in Rana grylio virus. Virus Res. 123, 128–137.
doi:10.1016/j.virusres.2006.08.007
Zhu, J., Huang, X., Yang, Y. 2007a. Innate immune response to adenoviral vectors is
mediated by both toll-like receptor-dependent and -independent pathways. J. Virol.
81, 3170-3180.
Zhu, J., Huang, X., Yang, Y., 2007b. Type I IFN signaling on both B and CD4 T cells is
required for protective antibody response to adenovirus. J. Immunol. 178, 3505–
3510. doi:10.4049/jimmunol.178.6.3505
Zsák, L., Kisary, J., 1984. Grouping of fowl adenoviruses based upon the restriction
patterns of DNA generated by BamHI and HindIII. Intervirology. 22, 110–114.
doi:10.1159/000149541
138