Download University of Veterinary Medicine Hannover - Ti

Document related concepts

Meningococcal disease wikipedia , lookup

African trypanosomiasis wikipedia , lookup

Whooping cough wikipedia , lookup

Norovirus wikipedia , lookup

Human cytomegalovirus wikipedia , lookup

Middle East respiratory syndrome wikipedia , lookup

Ebola virus disease wikipedia , lookup

HIV wikipedia , lookup

West Nile fever wikipedia , lookup

Eradication of infectious diseases wikipedia , lookup

Chickenpox wikipedia , lookup

Influenza A virus wikipedia , lookup

Orthohantavirus wikipedia , lookup

Marburg virus disease wikipedia , lookup

Henipavirus wikipedia , lookup

Neisseria meningitidis wikipedia , lookup

Herpes simplex virus wikipedia , lookup

Pandemic wikipedia , lookup

Hepatitis B wikipedia , lookup

Herpes simplex research wikipedia , lookup

Transcript
University of Veterinary Medicine Hannover
Clinic for Poultry
Molecular epidemiology of infectious bursal disease
viruses and development of a microparticle based vaccine
THESIS
Submitted in partial fulfillment of the requirements for the degree
DOCTOR OF PHILOSOPHY (PhD)
at the University of Veterinary Medicine Hannover, Germany
By
Tamiru Negash Alkie
(Arsi, Ethiopia)
Hannover, Germany 2013
Supervisor:
Prof. Dr. Silke Rautenschlein (Clinic for
Poultry, University of Veterinary Medicine
Hannover, Germany)
Advisory Committee:
Prof. Dr. Silke Rautenschlein
Prof. Dr. Beatrice Grummer (Institute of Virology,
University
of
Veterinary
Medicine
Hannover,
Germany)
PD Dr. Reimar Johne (Federal Institute for Risk
Assessment, Berlin, Germany)
1st Evaluation:
Prof. Dr. Silke Rautenschlein
Prof. Dr. Beatrice Grummer
PD Dr. Reimar Johne
2nd Evaluation:
Prof. Dr. H. M. Hafez (Institute of Poultry Diseases,
Faculty of Veterinary Medicine, Free University
Berlin, Germany)
Date of oral exam:
25 April, 2013
Dedicated to
My wife, Eskedar Hailegebral
&
My children, Heldana and Yafet
V
Table of Contents
Table of Contents ....................................................................................................... V Publications .............................................................................................................. VII List of abbreviations ................................................................................................... IX List of Figures .......................................................................................................... XIII List of Tables ........................................................................................................... XIV 1. Introduction ............................................................................................................. 1 2. Literature review ..................................................................................................... 4 2.1. IBDV genomic organization and functions ....................................................... 4 2.2. Pathobiology of IBDV ....................................................................................... 8 2.2.1. Pathogenetic mechanisms ........................................................................ 8 2.2.1.1. Polyprotein (PP) cleavage sites .......................................................... 8 2.2.1.2. Capsid and polymerase proteins......................................................... 8 2.2.1.3. Nonstructural protein ......................................................................... 10 2.2.2. IBDV host cell receptor and virus entry ................................................... 11 2.2.3. Pathogenesis of infectious bursal disease ............................................... 11 2.2.4. Clinical disease and pathology ................................................................ 14 2.3. Immune responses to IBDV ........................................................................... 16 2.3.1. Innate immunity ....................................................................................... 16 2.3.2. Humoral immunity.................................................................................... 17 2.3.3. Cellular immunity ..................................................................................... 18 2.4. Field evolution and molecular epidemiology of IBDV ..................................... 19 2.4.1. IBDV reassortment .................................................................................. 20 2.4.2. IBDV recombination................................................................................. 22 2.4.3. IBDV quasispecies and reversion to virulence ........................................ 24 2.5. Diagnostic methods........................................................................................ 25 2.5.1. Embryo inoculation .................................................................................. 25 2.5.2. In vitro virus propagation ......................................................................... 25 2.5.3. Immunological methods ........................................................................... 25 2.5.4. Molecular characterization ....................................................................... 26 2.6. Vaccines and vaccination against IBDV ......................................................... 28 VI
2.7. DNA vaccines ................................................................................................ 32 2.7.1. Stimulation of immune cells by DNA vaccines ......................................... 32 2.7.2. IBDV DNA vaccines................................................................................. 33 2.8. Adjuvants ....................................................................................................... 34 2.8.1. Cytokine adjuvants for avian viral vaccines ............................................. 35 2.8.1.1. Chicken IL-2 and IFN-γ as adjuvants for IBDV-DNA vaccines .......... 40 2.8.2. Toll like receptor (TLR) ligands as adjuvants ........................................... 40 2.9. Microparticulate vaccine and adjuvant carrier systems .................................. 42 2.9.1. Enhancing specific immunity by PLGA MPs ............................................ 45 3. Goals and objectives ............................................................................................ 47 4. Molecular evidence of very virulent infectious bursal disease virus in chickens in
Ethiopia .................................................................................................................... 48 5. Mucosal application of cationic poly(D, L-lactide-co-glycolide) microparticles as
carriers of DNA vaccine and adjuvants to protect chickens against infectious bursal
disease ..................................................................................................................... 51 6. Discussion and conclusions.................................................................................. 81 6.1. Molecular epidemiology of IBDV field isolates ............................................... 81 6.2. Immune responses induced by candidate IBDV DNA vaccines and correlation
to protection .......................................................................................................... 85 6.3. Do molecular adjuvants enhance protectivity of an IBDV DNA vaccine? ....... 86 6.4. Cationic PLGA MPs as particulate carriers for DNA vaccine and molecular
adjuvants............................................................................................................... 88 6.5. Promise of cationic PLGA MPs in improving an IBDV DNA vaccine .............. 89 6.6. Concluding remarks and future perspectives ................................................. 92 7. Summary .............................................................................................................. 94 8. Zusammenfassung ............................................................................................... 96 9. References ........................................................................................................... 99 10. Acknowledgements .......................................................................................... 143 VII
Publications
Research articles:
NEGASH, T., E. GELAYE, H. PETERSEN, B. GRUMMER u. S. RAUTENSCHLEIN
(2012):
Molecular evidence of very virulent infectious bursal disease virus in chickens in
Ethiopia.
Avian Dis 56 :605-610.
NEGASH, T., M. LIMAN u. S. RAUTENSCHLEIN (Vaccine revised):
Mucosal application of cationic poly(D, L-lactide-co-glycolide) microparticles as
carriers of DNA vaccine and adjuvants to protect chickens against infectious bursal
disease
Oral presentations at scientific meetings:
NEGASH, T., B. GRUMMER u. S. RAUTENSCHLEIN (2012):
Molecular identification and differentiation of infectious bursal disease viruses
(IBDVs) in field outbreaks.
82. Fachgespräch über Geflügelkrankheiten, Hannover, Germany; November 2012.
VIII
NEGASH, T.u. S. RAUTENSCHLEIN (2013):
Does mucosal application of cationic poly(D, L-lactide-co-glycolide) (PLGA)
microparticles as carriers of DNA vaccine and adjuvants enhance immunity against
infectious bursal disease virus?
23rd Annual meeting of the society for virology, Kiel, Germany; March 2013.
Poster presentations at scientific meetings:
NEGASH, T., B. GRUMMER u. S. RAUTENSCHLEIN (2011):
Molecular evidence of very virulent infectious bursal disease virus in chickens in
Ethiopia.
21st Annual meeting of the society for virology, Freiburg, Germany; March 2011.
NEGASH, T u. S. RAUTENSCHLEIN (2011):
Development of new generation vaccination strategies against infectious bursal
disease virus of chickens.
4th Graduate school day, Bad Salzdetfurth, Germany; November 2011.
NEGASH, T. u. S. RAUTENSCHLEIN (2012):
Mucosal application of cationic PLGA microparticles as carriers of DNA vaccines and
adjuvants enhances viral clearance in SPF chickens after IBDV challenge.
XIIth Avian immunology research group conference. Roslin institute, Edinburgh, UK
August 2012.
IX
List of abbreviations
aa
Amino acids
Abs
Antibodies
AC-ELISA
Antigen capture ELISA
Ag
Antigen
AIV
Avian influenza virus
aMPV
Avian metapneumovirus
APCs
Antigen presenting cells
BF
bursa of Fabricius
CAM
Chorioallantoic membrane
CD
Cluster of differentiation
CEFs
Chicken embryo fibroblasts
chIFN
Chicken interferon
cHsp90
Chicken heat shock protein 90
cIBDV
Classical infectious bursal disease
CMI
Cell mediated immunity
CMV
Cytomegalovirus
CpG-ODN
CpG-deoxynucleoside
CTAB
Cetyltrimethylammonium bromide
DCs
Dendritic cells
DNA
Deoxyribonucleic acid
dpi
Days postinfection
dsRNA
Double stranded RNA
ELISA
Enzyme-linked immunosorbent assay
FDCs
Follicular dendritic cells
GC
Germinal center
GILZ
Glucocorticoid-induced leucine zipper
GM-CSF
Granulocyte-macrophage cell stimulating factor
Gzm A
Granzyme A
HA
Hemagglutinin
X
HI
Hemagglutination inhibition
HIV
Human immunodeficiency virus
HN
Hemagglutinin-neuraminidase
hVP2
Hypervariable region of VP2
HVT
Herpesvirus of Turkeys
IBD
Infectious bursal disease
IBDV
Infectious bursal disease virus
IBD-ICX
IBD immune complex
IBV
Infectious bronchitis virus
IFN-γ
Interferon-γ
Ig
Immunoglobulin
iIELs
Intraepithelial lymphocytes
IL-2
Interleukin-2
IL-2R
IL-2 receptor
ILTV
Infectious laryngotracheitis virus
IM
Intramuscular
iNOS
Inducible NO synthase
JAK/STAT
Januskinase/signal
transducers
and
activators
of
transcription
MAB
Maternal Ab
mAbs
Monoclonal Abs
MDV
Marek’s disease virus
MHC
Major Histocompatibility complex
MPs
Microparticles
mRNA
Messanger RNA
ND
Newcastle disease
NDV
Newcastle disease virus
NF-κB
Nuclear factor kappa-light-chain-enhancer of activated B
cells
NK cells
Natural killer cells
NO
Nitric oxide
XI
nt
Nucleotide
OAS
20-50-oligoadenylate synthetase
P
Projection
PAMPs
Pathogen-associated molecular patterns
PBL
Peripheral blood lymphocytes
PCR
Polymerase chain reaction
PEI
Polyethylenimine
PFN
Perforin
pi
Postinfection
PI3K)/Akt
Phosphatidylinositol 3-kinase
PKR
Protein kinase R
PLGA
poly(D, L-lactide-co-glycolide)
PP
Polyprotein
PRRs
Pathogen recognition receptor
pVP2
Precapsid VP2
qRT-PCR
Quantitative RT-PCR
rAdV
Recombinant Adenovirus
RFLP
Restriction fragment length polymorphism
rFPV
Recombinant fowlpox virus
RNA
Ribonucleic acid
RT-PCR
Reverse transcriptase PCR
SPF
Specific pathogen free
SRs
Scavenger receptors
SVPs
Subviral particles
Th1
T helper-1
TLRs
Toll-like receptors
tMRCA
Time of the most recent common ancestor
TNF-α
Tumor necrosis factor-α
Tregs
Regulatory T-cells
VDAC2
Voltage-dependent anion channel 2
VLPs
Virus-like particles
XII
VNT
Virus neutralization test
VP2
Virus protein 2
vvIBDV
Very virulent IBDV
vvVP2
Very virulent VP2
XIII
List of Figures
Chapter 2
Fig. 1.
Method of vaccine (DNA/protein) microencapsulation in PLGA
microspheres using the w/o/w emulsion solvent evaporation
method ........................................................................................... 44
Chapter 5
Fig. 1.
Quantification of postchallenge chIFN-α mRNA expression levels
in the bursa at 3 (A), and 7 dpi (B); and chIL-4 mRNA expression
at 3 (C) and 7 dpi (D). .................................................................... 78
Fig. 2.
Quantification of postchallenge intrabursal CD4+ T-cells at 3 (A),
and 7 dpi (B); and CD8+ T-cells at 3 (C) and 7 dpi (D). ................. 80
XIV
List of Tables
Chapter 2
Table 1.
Functions of IBDV proteins .............................................................. 7
Table 2.
Examples of natural reassortant IBDV isolates ............................. 21
Table 3.
Examples for natural recombinant IBDV isolates ........................... 23
Table 4.
Experimental IBDV vaccines .......................................................... 31
Table 5.
Improvement of the protective efficacy of avian viral vaccines
after co-administration of plasmid encoded avian cytokines .......... 38
Table 6.
Examples for avian cytokines delivered by live viral vectors as
adjuvants for avian viral vaccines .................................................. 39
Chapter 5
Table 1.
Experiment 1: Evaluation of protection conferred by IBDV DNA
vaccinations in comparison to a live IBDV vaccine ........................ 70
Table 2.
Experiment 2: Evaluation of protection induced by IBDV DNA
vaccination in combination with molecular adjuvants ..................... 72
Table 3.
Experiment 3: Study design showing experimental groups and
the timing of vaccine and adjuvant application ............................... 74
Table 4.
Experiment 3: Determination of the protective efficacy of MP
based DNA vaccine and adjuvants after mucosal application to
SPF chickens ................................................................................ 75
1
1. Introduction
Immunosuppressive pathogens are major constraints of poultry production. They can
affect the efficacy of the immune system, which leads to vaccination failure and
increased susceptibility to many pathogens. Infectious bursal disease virus (IBDV) is
a relevant immunosuppressive virus of chickens. It is a dsRNA virus targeting
primarily the immature IgM+ B-cells residing in the bursa of Fabricius (BF), which is a
primary lymphoid organ in avian species. Worldwide, the poultry industry has
encountered heavy economic losses associated with very virulent (vv) IBDV strains
during the last several years. These strains may cause high mortality in affected
chicken flocks and severe immunosuppression that involves both innate and adaptive
immune responses.
During recent years, significant progress has been made to understand the molecular
epidemiology of vvIBDV. After the emergence of the virus in Europe in the late
1980’s, it was reported to have spread to several poultry producing regions
worldwide. By the mid-1990’s, almost 80% of the World Organization for Animal
Health (OIE) member countries have reported the occurrence of vvIBDV based on
pathological and molecular investigations from field outbreaks. A recent study
addressing the global molecular epidemiology of IBDV from four continents including
Africa showed that 60-76% of IBDV isolates were vvIBDVs. However, limited
information is available regarding the molecular epidemiology of IBDV, particularly
that of vvIBDV in Africa. Although an acute infectious bursal disease (IBD) infection
was suspected in major field outbreaks involving commercial farms and breeding
centers in Ethiopia in recent years, the nature and epidemiology of the virus remain
unknown.
Effective vaccination programs in combination with biosecurity are vital to control
IBD. Inactivated vaccines are applied mainly to breeder flocks and may require
multiple boosts to induce strong humoral immunity. Conversely, live attenuated
vaccines are applied to layers and broilers and stimulate not only humoral immunity
but also cell mediated immunity (CMI). Major challenges of live IBDV vaccines
2
include the possible risk of reversion to more virulent IBDVs and generation of
chimeric viruses by exchange of viral segments between vaccinal and wild IBDV
strains.
New generation IBDV-subunit vaccines may circumvent some of these problems and
provide protective levels of immunity. DNA vaccines may be good candidates due to
their ability to elicit both humoral and cellular immunity. Nevertheless, they did not
provide protective levels of immunity in many studies and may require well
characterized adjuvants to enhance their immunogenicity and protection. A number
of studies have demonstrated that co-administration of chicken interleukin-2 (chIL-2)
and synthetic unmethylated oligodeoxynucleotides (ODN) containing cytosineguanosine in succession (CpG-ODN) improved IBDV DNA vaccine efficacy. CpGODN stimulates the host defense system by mimicking host invasion by pathogens to
keep antigen presenting cells (APCs) in a state of alert. Generally, both molecular
adjuvants have multifunctional immunoregulatory properties and are known to
activate dendritic cells, macrophages, B- and T-cells and they can enhance the
expression of costimulatory molecules by these cells. This may enhance Ag
presentation and subsequent immunological responses to the co-administered IBDV
DNA vaccines.
Despite measurable progress that has been made in developing DNA vaccines,
further improvements are required for possible large scale applications. Vaccines and
adjuvants can be made more effective by employing appropriate particulate delivery
systems to enhance the uptake and processing by APCs. Microparticles (MPs)
prepared from biodegradable poly(D, L-lactide-co-glycolide) (PLGA) polymers are
one of the most extensively characterized particulate carriers for delivering vaccines
and adjuvants either parenterally or mucosally to enhance systemic and mucosal
immunity.
The goals of this study were to determine the molecular epidemiology of recent IBDV
isolates in Ethiopia and furthermore to develop and improve IBDV DNA vaccines by
using adjuvants and MPs as mucosal vaccine and adjuvant carriers.
3
The objectives of the first part of this study were to evaluate the pathogenicity of
Ethiopian IBDV field isolates experimentally and to characterize their molecular
nature by investigating the hypervariable region of the virus protein (VP) 2 (hVP2)
and the 5` two thirds of VP1.
The objectives of the second part of the thesis were to assess the immunogenicity
and protective efficacy of candidate recombinant IBDV DNA vaccines that encode
the vvIBDV-VP2 genes of a selected Ethiopian and reference IBDV strain; we
included the molecular adjuvants, CpG-ODN and plasmid encoded chIL-2 to
enhance protection. Furthermore, we evaluated cationic PLGA MPs as a carrier for
mucosal delivery of the DNA vaccine and adjuvants to improve the efficaciousness of
the IBDV DNA vaccine.
4
2. Literature review
IBDV is one of the economically most important immunosuppressive viruses of
chickens. IBDV is an Avibirnavirus and belongs to the family of Birnaviridae (BROWN
1989). Two serotypes of the virus have been described. Serotype 1 IBDV strains are
pathogenic to chickens (MÜLLER et al. 2003; VAN DEN BERG et al. 2004), whereas
serotype 2 strains are non-pathogenic (MCFERRAN et al. 1980). Serotype 1 IBDV
isolates comprise the variant, calssical virulent and vvIBDV strains, which greately
differ in their pathogenicity to chickens. Variant IBDVs do not cause mortality,
whereas the classical strains cause up to 20% mortality (MÜLLER et al. 2003).
vvIBDV causes mortality exceeding 50% in susceptible chickens (CHETTLE et al.
1989; BERG et al. 1991; MÜLLER et al. 2003).
2.1. IBDV genomic organization and functions
IBDV is a non-enveloped virus with a bipartite dsRNA genome (DOBOS et al. 1979;
MULLER et al. 1979). The main open reading frame of genome segment A encodes
a polyprotein (PP) (NH2-pVP2-VP4-VP3-COOH) that is cleaved by the virus encoded
protease into pre-capsid virus protein (VP)2, VP4 and VP3 within infected cells
(BIRGHAN et al. 2000; LEJAL et al. 2000). The pre-capsid VP2 undergoes defined
sequential C-terminal cleavage by VP4 (viral protease) (SANCHEZ u. RODRIGUEZ
1999), host protease (puromycin-sensitive aminopeptidase) (IRIGOYEN et al. 2012),
and by the endopeptidase activity of VP2 (LUQUE et al. 2007; IRIGOYEN et al.
2009) to release the mature VP2 protein. Other small peptides released during the
PP processing such as pep46 (46 aa) remain associated with the outer capsid
(CHEVALIER et al. 2005), but can not be visualized in IBDV particles by X-ray
crystallography (COULIBALY et al. 2005). The N-terminus moiety of pep46 bears a
positively charged hydrophobic domain and may be responsible for virus penetration
into the cytoplasm of infected cells (GALLOUX et al. 2007). The C-terminal moiety of
pep46 may assist in the formation of larger pore sizes to enhance viral entry into
infected cells (GALLOUX et al. 2010).
5
The VP2 crystal structure indicates three domains: the base (B), shell (S), and
projection (P) domains (COULIBALY et al. 2005; GARRIGA et al. 2006; LEE et al.
2006). The B and S domains are formed by the conserved N-and C-terminal
stretches of VP2. The P domain is the middle part containing the host cell receptor
binding motifs and the hypervariable region of VP2 (hVP2). The hVP2 harbours
antigenic major hydrophilic peak A (amino acid; aa 212-224) and B (aa 314-325)
(SCHNITZLER et al. 1993) that form loops PBC (aa 219-224) and PHI (aa 316-324)
(COULIBALY et al. 2005), respectively. The minor hydrophilic peak 1 (aa 248-254)
and peak 2 (aa 279-290) in the hVP2 form loops PDE (aa 249-254) and PFG (aa 279284) (COULIBALY et al. 2005).
The IBDV outer capsid is composed of a single shell of 260 trimeric spikes formed by
the P domain of VP2 radially projected from the capsid, and organized in a T=13
icosahedral lattice (BOTTCHER et al. 1997). This organization is determined by the
electrostatic interactions of precapsid VP2 with the VP3 C-terminal residues
(SAUGAR et al. 2010). The expression of the mature VP2 alone as a recombinant
protein forms dodecahedral T=1 subviral particles (SVPs) containing 20 VP2 trimers
(CASTON et al. 2001; GARRIGA et al. 2006), whereas precapsid VP2 expression
forms a tubular structure (CASTON et al. 2001).
The last five acidic residues at the C-terminus of VP3 interact with pVP2 during
particle morphogenesis for correct capsid assembly (CHEVALIER et al. 2004;
SAUGAR et al. 2005; LUQUE et al. 2009). The sixteen C-terminal residues of VP3
interact with VP1 (TACKEN et al. 2002; MARAVER et al. 2003; GARRIGA et al.
2007). VP3 binding to the genomic dsRNA and VP1 forms the ribonucleoprotein
complex (LUQUE et al. 2009).
The VP5 protein is encoded by another open reading frame on segment A that
partially overlaps with the 5’ end of the PP gene. It is a host membrane-associated
and highly basic protein with a cytoplasmic N-terminus and an extracellular Cterminal domain (LOMBARDO et al. 2000).
6
Segment B encodes the polymerase (SAUGAR et al. 2010), which is present in the
virion as a free protein or covalently linked to the 5′ ends of both genome segments
(VPg) (MÜLLER u. NITSCHKE 1987). PAN et al. (2007) has recently characterized
the VP1 crystal structure. VP1 has three domains: the N- terminus (aa 1-167), central
polymerase (aa 168-658) and C-terminal (aa 659-878) regions. The N-terminus of
VP1 is involved in protein priming as it possesses the putative guanylylation site (XU
et al. 2004; PAN et al. 2007). The central polymerase domain folds into a right-hand
shape (fingers-palm-thumb) structure. The five RNA polymerase motifs (C, A, B, D
and E) are located in the palm region of the polymerase. Each of the aa motifs
function during virus replication, for example, in nucleotide recognition and binding (
e.g. motifs A, B & F), phosphoryl group transfer (A & C), in a metal ion like Mn2+ and
Mg2+ binding (C), nucleotide guidance to active sites (D), and primer gripping (E)
(PAN et al. 2007). Motif C forms the polymerase active site (SHWED et al. 2002;
PAN et al. 2007). The finger sub-domains contain polymerase motifs F and G that
are involved in virus replication (PAN et al. 2007).
The 5’ non-coding regions of both segments contain promoter elements
(NAGARAJAN u. KIBENGE 1997) as well as 18S rRNA binding element and play
roles in the initiation of virus replication (MUNDT u. MÜLLER 1995).
No N-linked glycosylation of any of the virion proteins has been detected. The
biological functions of the different proteins of IBDV are summarized in table 1.
7
Table 1: Functions of IBDV proteins
Proteins
Functions
References
VP2
Host receptor binding
(OGAWA et al. 1998)
Contains neutralizing epitopes
(AZAD et al. 1987)
Virulence determinant
(BRANDT et al. 2001)
Cell culture adaptation
(MUNDT 1999)
Apoptosis
(FERNANDEZARIAS et al.
1997)
VP3
Endopeptidase activity
(IRIGOYEN et al. 2009)
Chaperone activity
(CHEVALIER et al. 2004)
Antiapoptosis by interacting with PKR
(BUSNADIEGO et al.
2012)
Suppresses hosts RNA silencing mechanism (VALLI et al. 2012)
VP4
Transcriptional activator
(TACKEN et al. 2002)
Forms ribonucleoprotein complex
(LUQUE et al. 2009)
Viral protein processing (viral protease)
(BIRGHAN et al. 2000)
Trans-activate VP1 synthesis
(BIRGHAN et al. 2000)
Suppresses type I IFN by interacting with
(LI et al. 2013b)
GILZ
VP5
VP1
Early antiapoptotic effects
(LIU u. VAKHARIA 2006)
Late apoptotic effects
(LI et al. 2012)
Viral polymerase
(SAUGAR et al. 2010)
Virulence determinant
(LE NOUEN et al. 2012)
8
2.2. Pathobiology of IBDV
2.2.1. Pathogenetic mechanisms
Early studies revealed segment A as the sole determinant of IBDV virulence.
Nevertheless, the molecular basis of IBDV pathogenicity likely depends on the
synergism between both segments of the virus (BOOT et al. 2000; BRANDT et al.
2001; BOOT et al. 2005; ESCAFFRE et al. 2012).
2.2.1.1. Polyprotein (PP) cleavage sites
To determine the molecular determinants of IBDV virulence, segment A and B of
several strains were sequenced and analysed (XIA et al. 2008). vvIBDV isolates with
aa substitutions adjacent to the pVP2 maturation site (aa 441–442), close to the
VP2–VP4 (aa 512–513) and VP4–VP3 (aa 755–756) cleavage sites (BROWN u.
SKINNER 1996; CHEVALIER et al. 2004; XIA et al. 2008) and at the C-terminus of
VP3 (CHEVALIER et al. 2004a) were reported. An aa substitution at the predicted
protease active-site of the VP4 gene of one of the European highly virulent strains
UK661 was suspected to contribute to its virulence. Generally, these mutations may
speed up the proteolytic activity of VP4, the processing and maturation of VP2, and
the capsid assembly efficiency. This may enhance the replication of vvIBDV and lead
to higher yields of virus particles in the host tissues (YAMAGUCHI et al. 1997b;
CHEVALIER et al. 2004; XIA et al. 2008).
2.2.1.2. Capsid and polymerase proteins
Exposed at the virion surface, VP2 contributes to IBDV virulence (YAMAGUCHI et al.
1996b; BOOT et al. 2000). Most vvIBDV isolates have key aa marker of virulence at
the hVP2 region (BROWN et al. 1994; BROWN u. SKINNER 1996; YAMAGUCHI et
al. 1997a; ISLAM et al. 2001). Specific aa residues responsible for tissue culture
9
adaptation, virulence and cell tropism have been mapped onto the VP2 gene using
reverse genetics (MUNDT u. VAKHARIA 1996). Amino acid mutations at positions
253 (Q →H), 279 (D →N) and 284 (A →T) in the VP2 were generated by reverse
genetics and completely attenuated the vvIBDV isolates. The modified viruses
propagated well in cell-culture and showed reduced pathogenicity in chickens (LIM et
al. 1999; MUNDT 1999; VAN LOON et al. 2002). The adaptation of vvIBDV strain
OKYM to cell culture introduced comparable aa mutations at positions 279 (D →N)
and 284 (A →T) (YAMAGUCHI et al. 1996b). These mutations were also detected in
other cell culture-adapted classical and vvIBDV strains (YAMAGUCHI et al. 1996a).
A single aa mutation at position 253 (H →Q/N) in VP2 markedly increased the
virulence of an attenuated IBDV strain (JACKWOOD et al. 2008).
Substantial molecular evidence showed the role of segment B in IBDV pathogenicity
as well (LIU u. VAKHARIA 2004; BOOT et al. 2005). Field reassortant IBDV isolates
comprising segment A of vvIBDV and segment B of attenuated strains show reduced
pathogenicity under field conditions and when evaluated experimentally in
susceptible chickens compared to typical vvIBDV isolates (LE NOUEN et al. 2006).
This field observation of reduced pathogenicity was proven with a classical virulent
IBDV that was genetically modified to contain the VP2 gene from a vvIBDV strain.
This virus failed to cause morbidity and mortality in SPF chickens unless the
genetically modified virus contained a typical segment B from vvIBDV (LIU u.
VAKHARIA 2004). Recombinant IBDVs generated by exchanging different regions of
VP1 of a vvIBDV with their counterparts of VP1 of an attenuated IBDV in a vvIBDV
segment A background showed reduced virulence in SPF chickens (LE NOUEN et
al. 2012). These viruses replicated to reduced virus titers in the bursa indicating the
important role of VP1 in vvIBDV virulence. Putative virulence marker aa residues
across the VP1 protein were predicted (YU et al. 2010). JACKWOOD et al. (2012)
recently described the presence of aa motif TDN at positions 145, 146 and 147 in the
VP1 gene of all vvIBDVs tested, and their absence in non-vvIBDV isolates. An IBDV
strain designated as 94432 that maintained all virulence markers aa in its VP2 gene
like other prototype vvIBDVs failed to cause mortality in chickens. The presence of
10
threonine (T) at position 276 in the exposed hydrophobic groove of the finger domain
of VP1 has been then shown to contribute to the reduced pathogenicity. Exchanging
this aa at position 276 from T →V restores the pathogenicity of the molecularly
cloned virus (ESCAFFRE et al. 2012).
2.2.1.3. Nonstructural protein
VP5 is believed to play an important role in IBDV pathogenicity; nonetheless it is not
essential for viral replication (MUNDT et al. 1997). In vitro, the VP5 protein showed
extensive accumulation within the plasma membrane of infected cells at later time
points during IBDV replication (LOMBARDO et al. 2000). VP5 may have antiapoptotic effects on infected cells during early time points (LIU u. VAKHARIA 2006),
and at the same time it may trigger apoptosis at later stages of IBDV replication in
infected cells (YAO u. VAKHARIA 2001). Furthermore, it was suggested that VP5
plays a significant role in virus release from infected cells by triggering cell death
(WU et al. 2009).
IBDV infection induces the activation of the phosphatidylinositol 3-kinase (PI3K)/Akt
signaling pathway during the early phase of IBDV replication in DF-1 cells. The
regulatory subunit of PI3K may suppress premature apoptosis of infected cells, which
may sustain IBDV replication and production of large quantities of infectious virus
progeny (WEI et al. 2011). When PI3K was inhibited most of the infected cells
showed early apoptotic signatures (WEI et al. 2011). Voltage-dependent anion
channel 2 (VDAC2) and VP5 (aa residues from 1-50) interaction later during IBDV
infection may be responsible for the release of infectious virus particles from infected
cells (LI et al. 2012). VDAC2 is a host molecule, which forms pores in the outer
mitochondrial membrane and is involved in apoptosis (SHOSHAN-BARMATZ et al.
2010). Blocking VDAC2 by small interfering RNA inhibited IBDV-induced apoptosis,
reduced virus release and virus titer.
11
2.2.2. IBDV host cell receptor and virus entry
IBDV may bind to host cell proteins such as N-glycosylated polypeptide(s) expressed
on the cell membrane of immature IgM+ B-cells during viral entry process (OGAWA
et al. 1998; LUO et al. 2010). A pore forming peptide of the virus (pep46), which is
associated with the outer capsid of the IBDV particle, may facilitate viral entry into the
cytoplasm of infected cells (GALLOUX et al. 2007; GALLOUX et al. 2010). A lipid raft
mediated endocytic mechanism was suggested based on the results of an in vitro
study to support entry of attenuated IBDV to the cells (YIP et al. 2012). An in vitro
binding assay showed the interaction of subviral particles (SVPs) derived from the
mature VP2 with the chicken heat shock protein 90 (cHsp90) expressed on the
surface of DF-1 cells (LIN et al. 2007). This binding of IBDV to cHsp90α was inhibited
by cHsp90α interfering microRNAs that resulted in reduced virus titer (YUAN et al.
2012). The binding of VP2-SVP with α4β1 integrin suggested an integrin-binding
domain in VP2 of IBDV, which was later confirmed that aa residues at positions 234236 are responsible (DELGUI et al. 2009). This integrin-binding motif is conserved in
all IBDV strains. A single point mutation within this motif might completely abrogate
the binding of SVPs to cells and virus infectivity (DELGUI et al. 2009). Immature Bcells have abundant α4β1 on their surfaces (ROSE et al. 2002).
2.2.3. Pathogenesis of infectious bursal disease
Natural IBDV infection occurs by the oral route. Other mucosal routes have been
demonstrated for experimental IBDV infection. The mononuclear phagocytic cells and
lymphoid cells of the gut mucosa may serve as targets for IBDV infection and
replication (MÜLLER et al. 1979). Infected macrophages transport the virus to the
bursa of Fabricius (BF), the prime target organ for extensive IBDV replication in the
cytoplasm of intrabursal IgM+ B-cells (KAUFER u. WEISS 1980; HIRAGA et al.
1994). Virus dissemination to other lymphoid organs such as to the thymus, bone
marrow, spleen, Peyer's patches, cecal tonsils, and Harderian glands may take place
12
mainly during vvIBDV infection of susceptible chickens (ETERRADOSSI u. SAIF
2008). The cecal tonsils and bone marrow may serve as non-bursal lymphoid tissues
supporting virus replication at later time points (ELANKUMARAN et al. 2002).
As early as 48 hr pi, IBDV infection induces prominent inflammation in the BF. By day
3 to 4 pi, all bursal IgM+ B-cells are infected and show cytolytic changes (CHEVILLE
1967). vvIBDV strains such as UK661 can infect Bu-1+ cells and IgY+ B-cells in
several lymphoid tissues indicating both immature and mature B-cells can be infected
(WILLIAMS u. DAVISON 2005), whereas classical virulent and variant IBDVs mostly
target immature B-cells. Between 7 and 21 dpi, IBDV infection results in significant
reduction in the number of the B subpopulation compared to the A subpopulation of
IgM+ B-cells as determined by flow cytometric analysis (PETKOV et al. 2009). These
two phenotypes differed on their cell size and granularity as well as showed
differential expression levels of Lewis(x), IgM, Bu-1b, and MUI78 surface antigens. It
becomes apparent that macrophages are infected with IBDV and in vitro studies
indicated rapid virus replication with altered in vitro phagocytic activity of such cells
(KHATRI u. SHARMA 2009b). Other cells like the bone marrow-derived
mesenchymal stem cells may be infected with IBDV (KHATRI u. SHARMA 2009b).
The haemopoietically derived reticular cells, which reside in the antigen-trapping
zone of the spleen (ellipsoid and periellipsoidal white pulp), were found to be more
susceptible to IBDV. The reticular cells of mesenchymal origin, which reside in the
bursal cortex, and periarteriolar lymphoid sheaths, germinal center (GC) and red pulp
of the spleen were relatively resistant to IBDV (BIRO et al. 2011). Bursal follicular
dendritic cells disappeared during IBDV infection probably due to lack of an intact Bcell microenvironment (JEURISSEN et al. 1998; KABELL et al. 2006).
Generally, the sequellae of IBDV infections such as severity of clinical signs, organ
lesions and immunosuppression correlate with the status of immunity, age and
genetic background of affected chickens and with the virulence of the infecting virus
strain (BERG 2000). SPF chickens infected with vvIBDV develop an earlier onset of
mortality and more severe bursal lesions compared to broiler chickens with MAB and
vaccinated chickens (HASSAN et al. 2002; ARICIBASI et al. 2010). Infections by
13
virulent IBDV resulted in an earlier onset and higher production of IFN-γ in the bursa
of young SPF chickens compared to their older counterpart. Massive interferon-γ
(IFN-γ) production from T-cells infiltrating the bursa during an acute IBDV infection
(ELDAGHAYES et al. 2006; RAUW et al. 2007) further activates macrophages to
release proinflammatory cytokines such as interleukin-6 (IL-6) and also nitric oxide
(NO), which may aggravate bursal lesions (KIM et al. 1998). IFN-γ is suggested to be
a potent apoptosis inducer in IBDV infected or adjacent healthy B-cells (LAM 1997;
TANIMURA u. SHARMA 1998). A massive mast cell influx detected in the bursa of
SPF chickens infected with vvIBDV may aggravate bursal lesions as typical
indicators of acute hypersensitivity responses were observed in the bursa of such
chickens (WANG et al. 2008; WANG et al. 2012a). These cytokine mediated bursal
lesions may result in an early onset of severe immunosuppression in younger
chickens (RAUTENSCHLEIN et al. 2007). Similar patterns of cytokine production and
bursal lesions were detected in SPF layer type chickens infected with virulent IBDV
compared to infection of age-matched 3 weeks old broiler type chickens (ARICIBASI
et al. 2010).
Bursal lesions with B-cell depletions lead to severe humoral immunosuppression
(SHARMA et al. 1994; SHARMA et al. 2000). When susceptible neonatal chickens
younger than two weeks of age are infected, they may lose the entire bursal B-cells,
which result in permanent immunologic damage (HUDSON et al. 1975; WITHERS et
al. 2005). Cytokine dysregulation may cause suppression of innate and cellular
immunity (RAUW et al. 2007). This was elucidated as recombinant chicken IFN-γ
(rchIFN-γ) inhibited an in vitro proliferation of naïve peripheral blood lymphocytes
(PBL) or splenocytes (RAUW et al. 2007). Lymphocytes harvested from the blood
and lymphoid organs of an IBDV infected or live IBDV vaccinated chickens showed
reduced lymphoproliferative responses when stimulated by conventional mitogens
(MAZARIEGOS et al. 1990; SHARMA et al. 2000). This inhibition has been detected
during early time points after IBDV infection coinciding with an increased IFN-γ
response. Recently, in vitro study described the interaction of IBDV VP4 with the
chicken glucocorticoid-induced leucine zipper (cGILZ) that inhibited the transcription
14
of NF-κB with a subsequent suppression of the innate immunity (LI et al. 2013b).
Others
suggested
T-cells
with
regulatory
property
may
mediate
cellular
immunosuppression (KIM et al. 1998; SHANMUGASUNDARAM u. SELVARAJ
2011). IBDV infection induced mucosal immunosuppression denoted by reduced
intraepithelial lymphocytes (iIELs) and their in vitro cytotoxicity activity (KUMAR et al.
1998). A reduced intestinal secretory IgA+ B-cells were also detected (WANG et al.
2009a).
The immunological functions of both B- and T-cells may be restored during recovery
of infected chickens (KIM et al. 1999; SHARMA et al. 2000). The mechanisms of Bcell functional restoration were described, whereas of T-cells remained unclear.
Bursal stem cells, which survive IBDV-induced depletion, proliferate to generate new
and larger bursal follicles. They became repopulated with IgM+ B-cells, and Bu-1+
cells expressing IgM or IgY (WILLIAMS u. DAVISON 2005) and dendritic-like cells
(WITHERS et al. 2005). The B-cells in these follicles may undergo immunoglobulin
(Ig) gene (hyper)conversion for Ig diversity and sustain specific immunological
functions (WITHERS et al. 2005; WITHERS et al. 2006). Previous study indicated the
strong expression of the Lex and chB1 genes in the recovering follicles as indicators
of Ig gene (hyper)conversion (IVAN et al. 2001). Medullary B-cells surviving IBDV
infection form small follicles, which lack Ig gene (hyper)conversion. Birds with only
small follicles do not produce Abs against IBDV or other Ags, and such chickens may
be in a state of permanent immunosuppression (WITHERS et al. 2006).
2.2.4. Clinical disease and pathology
Chickens infected between 3 and 6 weeks of age develop the most severe clinical
signs of IBD (ETERRADOSSI u. SAIF 2008). Susceptible chickens exposed to
vvIBDV and classical virulent strains show a sudden onset of clinical disease within
2-3 days of exposure, characterized by severe depression and ruffled feathers (VAN
DEN BERG et al. 2004). Chickens younger than 2 weeks of age and birds older than
6
weeks
rarely
develop
clinical
signs
(VERVELDE
u.
DAVISON
1997).
15
Experimentally, broilers especially with higher MAB levels may not show clinical
signs or mortality when infected with vvIBDV. Mortality peaks at 4 dpi with virulent
strains. A rapid recovery after 5-7 dpi is a prominent feature of acute IBD (VAN DEN
BERG et al. 2000a). In general viral shedding in the faeces of naturally infected or
live vaccine vaccinated chickens can last up to 2 weeks and viral RNA can be
detected by RT-PCR up to 4 weeks (KABELL et al. 2005). Age and immune status of
infected chickens (SKEELES et al. 1979; ABDEL-ALIM u. SAIF 2001; IVAN et al.
2005; SAPATS et al. 2005), route of infection and nature of infecting viruses
(WINTERFIELD et al. 1972; SKEELES et al. 1979; ELANKUMARAN et al. 2002)
influence the development of clinical IBD and virus shedding.
On the first few days after infection, an increased in bursa to body weight ratio is
usually observed due to edematous bursae. Occasionally, extensive hemorrhage
throughout the entire bursa has been observed in the case of vvIBDV. Compared
with a moderately pathogenic strain of the virus, the vvIBDV strains caused a greater
decrease in thymic weight index. Generally, hemorrhagic inflammation of the bursa is
the main pathological feature of infection by virulent strains, whereas variant strains
(e.g. GLS and E/Del) cause rapid bursal atrophy mostly without an inflammatory
response (LAM 1997; TANIMURA u. SHARMA 1998). IBDV-induced lymphoid cell
depletion is responsible for bursal atrophy as early as 7-8 dpi (CHEVILLE 1967).
Histological bursal lymphoid depletions are comparable between vvIBDV and virulent
IBDV infection in the early few days of infection (TANIMURA et al. 1995;
TSUKAMOTO et al. 1995; INOUE et al. 1999; STOUTE et al. 2009), soon followed
by heterophilic infiltration. A more virulent IBDV strains caused severe lymphoid
depletions in the cecal tonsils, thymus, spleen, and bone marrow. The pathogenicity
may correlate with lesion production in non-bursal lymphoid organs. As the
inflammatory reaction subsides, cystic cavities develop in the medulla of affected
bursal follicles followed by fibrosis in interfollicular areas (ETERRADOSSI u. SAIF
2008). Proliferation of the bursal epithelial layer produced a glandular structure of
columnar epithelial cell containing globules of mucin. During recovery stage,
scattered foci of lymphocytes appeared in the bursal follicles.
16
2.3. Immune responses to IBDV
Apart from its immunosuppressive effects, IBDV infection in chickens activates all
branches of the immune system. However, the level of activation varies depending
on the virulence of infecting strains, age, immune status and genetic background of
affected chickens.
2.3.1. Innate immunity
The influx of macrophages, heterophils and mast cells in the bursa of Fabricius
constitutes the early innate immune response to IBDV (KHATRI et al. 2005;
PALMQUIST et al. 2006; RAUTENSCHLEIN et al. 2007; WANG et al. 2008). The
influx of these cells may be mediated by chemokines (IL-8, iNOS) (KHATRI et al.
2005; ELDAGHAYES et al. 2006; PALMQUIST et al. 2006; RAUTENSCHLEIN et al.
2007; RAUW et al. 2007; RAUF et al. 2011a). Toll-like receptors (TLRs) such as
TLR3 and TLR7 expressed by these inflammatory cells detect IBDV nucleic acids.
Their mRNA expressions have been found upregulated during an acute IBDV
infection (RAUF et al. 2011a; GUO et al. 2012). These interactions between IBDV
and TLR3 or TLR7 have been shown to activate the interferon (IFN) system and also
induced proinflammatory cytokines (IL-6, IL-1β, and IL-18) (RAUF et al. 2011a; GUO
et al. 2012). The release of these cytokines was suggested to be tightly regulated by
NF-κB, whereby its expression was found to be elevated in the bursa during the early
phase of IBDV infection (KHATRI u. SHARMA 2006; GUO et al. 2012). The
upregulation of IFN-α/β mRNA expression was reported in lymphoid organs and PBL
of chickens experimentally infected with virulent IBDV (KIM et al. 1998; RAUF et al.
2011a; MAHGOUB et al. 2012). The expression levels of these cytokines in the
bursa differ in relation to the virulence of infecting strains, genetic background and
age of infected chickens (ELDAGHAYES et al. 2006; RAUTENSCHLEIN et al. 2007;
ARICIBASI et al. 2010; RAUF et al. 2011a). IFNs may protect chickens against IBDV
infection. This was verified experimentally whereby chickens pretreated with
17
recombinant chicken IFN-α/β and latter challenged with IBDV had reduced challenge
virus replication and pathological lesions in the bursa (CAI et al. 2012).
IBDV infection of chicken embryo fibroblasts (CEFs) resulted in upregulation of IFNinducible 20-50-oligoadenylate synthetase (OAS), IFN regulatory factors, IL-6 and IL8 mRNA expression (LI et al. 2007). The OAS and RNase L pathway interferes with
viral infection through the cleavage of viral ssRNA, which is one of the recognized
viral suppressor activities of OAS (MALATHI et al. 2007). Recombinant type I IFN
pretreated CEFs resisted IBDV replication and resulted in reduced viral titer after
infection (O'NEILL et al. 2010; CAI et al. 2012).
Nitric oxide released by macrophages may constitute an early host defence against
IBDV and promotes the killing of IBDV-infected and possibly virus-free cells (KHATRI
et al. 2005; KHATRI u. SHARMA 2006; PALMQUIST et al. 2006; KHATRI u.
SHARMA 2009a).
2.3.2. Humoral immunity
Significant titers of systemic IBDV specific-Abs have been detected in the
convalescent sera of chickens that are naturally or experimentally infected with IBDV
(ETERRADOSSI u. SAIF 2008). All classes of Igs can be produced, but the Ab
response may not protect chickens from antigenetically different IBDV strains.
Neutralizing Abs are directed against the conformation dependent neutralizing
epitopes of VP2 (FAHEY et al. 1991; SNYDER et al. 1992). Abs against VP3
(BECHT et al. 1988; FAHEY et al. 1991) and conformation-independent antigenic
domains of VP2 (AZAD et al. 1987) are non-neutralizing. Live and inactivated IBDV
vaccines may induce vigorous Ab responses in the first few weeks postvaccination
(MAAS et al. 2001; ARICIBASI et al. 2010). Compared to cell culture derived strains,
bursal and embryo derived strains induce higher neutralizing Ab titers (RODRIGUEZCHAVEZ et al. 2002).
18
Humoral immunity plays a significant role in protection against IBDV. Maternal
antibody (MAB) provides passive protection in the first few weeks after hatch (ALNATOUR et al. 2004). MAB positive chickens developed significantly less bursal
lesions than Ab negative chickens after IBDV challenge supporting the role of
passive immunity in protection (HASSAN et al. 2002; ARICIBASI et al. 2010). MAB
may interfere with the development of an active immune response after IBDV
vaccination (RAUTENSCHLEIN et al. 2005a).
Although Ab mediated immunity is crucial against IBDV, an important role of the cell
mediated immunity (CMI) is suggested by several groups (RAUTENSCHLEIN et al.
2002a; YEH et al. 2002).
2.3.3. Cellular immunity
During acute IBD, while bursal follicles are B-cell depleted, T-cells accumulate at the
site of virus replication (TANIMURA u. SHARMA 1997; KIM et al. 1998; KIM et al.
2000; SHARMA et al. 2000). A notable influx of CD4+- and CD8+ T-cells was
detected as early as 1 dpi and peaked at around 7 dpi (KIM et al. 2000). Although
viral Ag was cleared by week 3 pi, T-cell influx and activation continued to week 12
pi. No T-cell depletion was detected from the bursa during IBDV infection. However,
IBDV particles were detected in intrabursal T-cells (MAHGOUB et al. 2012).
Infiltrating T-cells in the bursa show markers of activation such as upregulated IL-2,
major histocompatibility complex (MHC) class II molecules, and IFN-γ mRNA
expression (KIM u. SHARMA 2000; RAUW et al. 2007; RAUF et al. 2011b). T-cells
are not only involved in bursal recovery by killing virus infected cells, but also
contribute to bursal lesions. T-cell compromised SPF-chickens had the highest viral
Ag load and milder inflammatory bursal lesions compared to T-cell intact birds after
IBDV infection (RAUTENSCHLEIN et al. 2002a). T-cells infiltrating the bursa after
IBDV infection expressed higher levels of the mRNA for cell membrane-disrupting
proteins such as perforin (PFN) and granzyme A (Gzm A), and other cytolytic
molecules such as the high mobility group proteins. PFN and Gzm mediated
19
cytotoxic activity may contribute to rapid viral clearance from the bursa (RAUF et al.
2011b). The role of T-cells in IBDV protection was supported in vaccination studies
with T-cell or B-cell compromised chickens. Chickens depleted of functional T-cells
either by neonatal thymectomy or Cyclosporin A treatment showed insufficient
protection against IBDV challenge after immunization with an inactivated IBDV
vaccine, whereas chickens with intact T-cells had significantly higher IBDV protection
rates (RAUTENSCHLEIN et al. 2002b). Chickens with severely compromised Abproducing ability following treatment with cyclophosphamide were sufficiently
protected against IBDV challenge despite the absence of detectable vaccine-induced
Abs. This implies that T-cells may have role in protection (YEH et al. 2002).
2.4. Field evolution and molecular epidemiology of IBDV
The first outbreak of infectious bursal disease (IBD) that had occurred in 1957 in a
broiler farm near Gumboro, the Delaware area in the USA, was caused by the
classical serotype 1 IBDV (COSGROVE 1962). The variant IBDV strains then
emerged in the 1980’s in IBDV-vaccinated farms in the Delmarva area and were
antigenetically different from the former isolates. In the late 1980’s, vvIBDV emerged
in Europe (CHETTLE et al. 1989) and rapidly spread across continental Europe and
Asia (LIN et al. 1993; SHCHERBAKOVA et al. 1998), Middle East (PITCOVSKI et al.
1998), South America (DI FABIO et al. 1999), and Africa (ZIERENBERG et al. 2000).
IBDV undergoes genetic variation during its evolution to adapt to new hosts and to
escape the host immune responses. Different biological mechanisms may play
important roles for the emergence of novel viruses, particularly in segmented RNA
viruses, such as IBDV.
Early IBDV isolates frequently showed mutations at the major hydrophilic domains
particularly in the loops PBC and PHI, which affected the antigenicity of the strains and
induced vaccination failure (BAYLISS et al. 1990; HEINE et al. 1991; LANA et al.
1992; DORMITORIO et al. 1997). In the past few years, several field IBDV strains
20
isolated from different geographic areas showed aa substitutions at the minor
hydrophilic domains mainly at position 254 (loop PDE) and 284 (loop PFG)
(JACKWOOD u. SOMMER-WAGNER 2005; MARTIN et al. 2007; DURAIRAJ et al.
2011; JACKWOOD u. SOMMER-WAGNER 2011). Most of these viruses have been
identified from areas where the viruses have been circulating for a long period of time
(MARTIN et al. 2007). In the USA, one-third of the investigated field IBDV isolates
(out of 300) failed to react with any of the described monoclonal abs (mAbs) that
have been used to identify IBDV strains for the last 2 decades, which may reveal the
circulation of new IBDV subtypes (DURAIRAJ et al. 2011). A new variant IBDV
differing from the Delaware (Del E) variant of the Delmarva Peninsula was identified,
which did not react with those mAbs (GELB et al. 2012). IBDV isolates, which contain
epitopes of both variant and classical IBDVs in their VP2 genes were demonstrated,
which can affect mAb reactivity (JACKWOOD 2012). This may provide an
explanation for the increased antigenic and virulence diversity of the recent IBDV
isolates. Atypical IBDVs, which harbour aa residues characteristics of variant,
classical, and vvIBDV in their VP2 were characterized and showed atypical
pathogenicity and reactivity patterns to most of the mAbs (MARTIN et al. 2007).
2.4.1. IBDV reassortment
Genetic reassortment might be accountable for the emergence of vvIBDV in the late
1980’s in Europe (HON et al. 2006). The time of the appearance of the most recent
common ancestor (tMRCA) of very virulent (vv) VP2 is approximated around 1960,
whereas of vvVP1 around 1980 (HON et al. 2006), in which the latter coincided with
the emergence of vvIBDV in the late 1980’s (CHETTLE et al. 1989). Thus a newly
appeared vvVP1 from an unidentified avian reservoir was suggested to recombine
with an already existing vvVP2 to evolve to the vvIBDV genotype, which then caused
massive mortality in Europe (HON et al. 2006). This indicates the independent
evolutionary history of the two segments of vvIBDV (ISLAM et al. 2001; LE NOUEN
et al. 2006). Recently, several natural reassortant IBDV isolates were characterized
21
during field outbreaks (Table 2). The most common reassortant IBDVs contain
segment A of vvIBDV and segment B from attenuated strains indicating the
drawbacks of extensive application of live IBDV vaccines. Attempted experimental
generation of reassortant viruses by co-infecting specific pathogen free (SPF)
chickens with vvIBDV and attenuated serotype 1 IBDV has failed. The process of
reassortment may be more complex in the field than expected and may involve the
interactions of several factors: time, environment and vaccine pressure (WEI et al.
2008).
Table 2: Examples of natural reassortant IBDV isolates
Isolates
Sources of segments
Country
Year
References
Segment A
Segment B
Unknown
vvIBDV
vvIBDV
Europe
1980
(HON et al. 2006)
SH95
vvIBDV
Variant E
China
Unknown
(SUN et al. 2003)
02015.1
vvIBDV
Attenuated
France
Unknown
(LE NOUEN et al. 2006)
ZJ2000
Attenuated
vvIBDV
China
2000
(WEI et al. (2006)
TL2004
Attenuated
vvIBDV
China
2004
(WEI et al. 2008)
CA-K785
vvIBDV
Serotype 2
USA
2009
(JACKWOOD et al. 2011)
KZC-104
vvIBDV
Attenuated
Zambia
2004
(KASANGA et al. 2012)
22
2.4.2. IBDV recombination
Natural homologous intragenic recombination is described for many animal viruses
(LEE et al. 2013). The risk of live vaccines recombining to generate virulent natural
recombinants have been well described, and disease outbreaks associated with
these viruses have recently been described for infectious laryngotracheitis virus
(ILTV) infections of chickens (LEE et al. 2012). Recombination may lead to
antigenetically and genetically diverse IBDV populations and the emergence of novel
vvIBDV groups (HON et al. 2008; HE et al. 2009a). It has the potential to alter the
interactions of IBDV proteins and the orientation of the capsid domains preventing
neutralization by pre-existing Abs, which lead to vaccine failure. Almost all IBDV
recombinant viruses identified from field outbreaks are VP2 recombinants (Table 3).
Intrasegment recombination was also detected in segment B of two vvIBDV strains
possibly due to the recombination between two vvIBDV donors (HON et al. 2008). A
recently isolated IBDV strain, GX-NN-L, has reassortant characteristics, whereby its
segment A derived from vvIBDV, and segment B from an attenuated strain. But
interestingly, segment B contains putative aa residues typical for vvIBDV isolates
(CHEN et al. 2012a). An attenuated vaccinal strain, ViBursaCE, is suggested to be a
potential recombinant whereby its segment A is a mosaic between variant (Variant E)
and an attenuated French vaccine strain (Rhone-Merieux, strain-CT) (HON et al.
2008).
23
Table 3: Examples for natural recombinant IBDV isolates
Isolates
VP2 recombinants
Country
Year
References
SH-h
hVP2 from vvIBDV (HLJ-5 strain)
China
Unknown
(HON
KSH/KK1
849VB
et
al.
within segment A of an attenuated
2008; HE et al.
(D78) strain
2009a)
hVP2 from vvIBDV (SH.92 strain)
Korea
1992/1997
(HON
et
al.
within segment A from an
2008; HE et al.
attenuated (D78) strain
2009a)
Part of segment A from attenuated
Belgium
1987
(D78), and the other part from
(HON
et
al.
2008)
vvIBDV (D6948)
Several
aa sequences at loop regions PBC and
Venezuela 2001-2005 (JACKWOOD
isolates
PHI from classical IBDV & aa at PDE
Colombia
2012)
and PFG from variant IBDV
Several
aa sequences at loop regions PBC
isolates
from variant & PDE and PFG from
Mexico
2004-2011 (JACKWOOD
2012)
classical IBDV
157776
aa at positions 294 to 299 from
Italy
2003
vvIBDV & residues from 222 to 279
(MARTIN et al.
2007)
from an attenuated strain
VP1 recombinants
OE/G2
Segment B recombinant between
Turkey
Unknown
OKYM & OA/G1 vvIBDV strains
Harbin-1
Segment B recombinant between
(SILVA
et
al.
et
al.
2012)
China
Unknown
(HON
HLJ-7 or HENAN & GZ/96 vvIBDV
2008; SILVA et
strains
al. 2012)
24
2.4.3. IBDV quasispecies and reversion to virulence
The existence of RNA virus quasispecies may have a paramount contribution to virus
evolution. An RNA virus population is made up of heterogeneous viruses, which
share the consensus sequence but differ from each other by one or many mutations
(DOMINGO et al. 1985). In IBDV vaccine and field strains, the quasispecies
phenomenon has been described by real time RT-PCR and melting curve analysis
(JACKWOOD u. SOMMER 2002; HERNANDEZ et al. 2006). Pre-existing selection
pressure, for example, altered host immune status may favor one clone of a virulent
virus to overwhelm the virus population to maintain its endemicity (MORIMOTO et al.
1998).
Attenuated live IBDV vaccines are most frequently used to vaccinate commercial
chickens. Reversion of these attenuated vaccinal strains to more virulent phenotypes
under field and experimental conditions has been frequently reported (YAMAGUCHI
et al. 2000; JACKWOOD et al. 2008) possibly due to a lack of IBDV polymerase
fidelity during vaccine viral genome replication in the host cells. A tissue cultureadapted IBDV generated by reverse genetics from a vvIBDV strain reverted
phenotypically and genotypically to the vvIBDV pathotype after inoculation into SPF
chickens and maintained this pathotype afterwards (RAUE et al. 2004). Genetic
reversion of vaccine strains is most likely to be one of the mechanisms that may
contribute to the dissemination and persistence of virulent IBDV in the chicken
population worldwide.
25
2.5. Diagnostic methods
2.5.1. Embryo inoculation
The inoculation of bursal homogenates from IBDV infected chickens per the
chorioallantoic membrane of 9-10 days old embryonated SPF chicken eggs is the
most sensitive diagnostic method for virus isolation. The embryos die mostly within 35 days in the case of classical and very virulent viruses (HITCHNER 1970;
ROSALES et al. 1989). Embryos may not die from infection by the variant viruses,
yet show hepatic necrosis.
2.5.2. In vitro virus propagation
Adaptation of IBDV field isolates to a cell culture system requires passages in
embryonated chicken eggs and subsequent passages in cell culture system further
attenuates the virus to the extent that these viruses do not induce bursal lesions
(YAMAGUCHI et al. 1996a). The cell culture adapted viruses replicate in primary
avian cells such as CEFs (SHI et al. 2009) and continuous cell lines of avian (QT35)
and mammalian origins (Vero cells) (LUKERT u. DAVIS 1974). Compared to classic
and variant strains, adaptation of the vvIBDV viruses to cell culture has been very
difficult.
2.5.3. Immunological methods
ELISA and virus neutralization test (VNT) can be used to determine IBDV Ab levels.
Different IBDV-ELISA procedures have been described for routine diagnostic
purposes (MARQUARDT et al. 1980; KECK et al. 1993). Polyclonal based Agcapture ELISA is more sensitive (HASSAN et al. 1996) compared to the mAb-based
ones to detect IBDV Abs (LEE u. LIN 1992; VAN DEN BERG et al. 2004).
26
Commercially available ELISA systems use either intact viral particles or a mixture of
recombinant VP2 based proteins coated plates. VNT is mostly used for research
purposes (SKEELES et al. 1979). The VN titers accurately correlate with protection
of chickens against IBDV (KNOBLICH et al. 2000). Only VNT allows to differentiate
serotype 1 from serotype 2 IBDVs (JACKWOOD et al. 1985).
Determination of the antigenic properties of IBDV field isolates is necessary when
recurrent outbreaks are observed in poultry farms that had been previously
vaccinated against IBDV. Although in vitro VN tests can be used for detection of
antigenic differences between virus strains, in vivo cross protection studies are
essential to determine the antigenicity of a virus and complete evaluation of host
immune responses (JACKWOOD u. SAIF 1987). In vitro antigenicity is determined by
the reaction patterns of IBDV isolates to panels of mAbs, which target conformation
dependent neutralizing epitopes at the hydrophilic regions of VP2 (SCHNITZLER et
al. 1993; BERG et al. 1996; ETERRADOSSI et al. 1997) using VNT or ELISA
(SNYDER et al. 1988; VAKHARIA et al. 1994). By sequence analysis and sitedirected mutagenesis, aa residues, which influence the reactivity patterns of IBDVs
with specific mAbs were identified in the past (LANA et al. 1992; LETZEL et al. 2007;
ICARD et al. 2008; DURAIRAJ et al. 2011). Recombinant Abs developed from a
chicken single chain variable Ab fragments (scFv) also differentiate IBDV isolates in
a sandwich ELISA platform (SAPATS et al. 2005; SAPATS et al. 2006).
2.5.4. Molecular characterization
The classical methods for molecular characterization and differentiation of IBDV field
isolates include RT-PCR and restriction fragment length polymorphism (RFLP),
nucleotide sequence analysis, and quantitative real time RT-PCR (qRT-PCR)
(JACKWOOD 2004; WU et al. 2007a). The hypervariable region of VP2 (hVP2) of
IBDV between aa residues 206 and 350 shows significant aa sequence variations.
Sequencing of this region is used to categorize isolates into different pathogenic
27
strains (YAMAGUCHI et al. 1997b; JACKWOOD u. SOMMER-WAGNER 2005;
JACKWOOD u. SOMMER-WAGNER 2007; SREEDEVI et al. 2007).
In early studies, RT-PCR-RFLP analysis was performed mainly on hVP2 and seldom
on VP1 to distinguish IBDV isolates (MEIR et al. 2001; JUNEJA et al. 2008).
According to their restriction profiles, the viruses form molecular groups or genotypes
(JACKWOOD et al. 2001). However, the limitation of this method is that viruses in a
molecular group or with matching RFLPs may differ in their antigenic or virulence
property and require sequence analysis.
Direct sequencing of RT-PCR products of the hVP2 and 5’two thirds of VP1 may
provide deeper insights into the nucleotide and predicted aa identity of strains and
may determine mutations, recombination and reassortment in the viral genomes (LE
NOUEN et al. 2005). The nucleotide sequences of VP2 and VP1 can be further used
to construct phylogenetic trees to demonstrate genetic relationships between IBDV
isolates (HON et al. 2006).
A TaqMan qRT-PCR and melting curve analysis can be used to trace mutations in
the hVP2 region (JACKWOOD et al. 2003). This method allows comparing
sequences between field and vaccinal strains (JACKWOOD u. SOMMER 2002; GAO
et al. 2007). It determines a single nucleotide polymorphism in VP2 (WU et al.
2007a). qRT-PCR quantifies viral load (MOODY et al. 2000).
RT-loop-mediated isothermal amplification (RT-LAMP) is a rapid field test used as a
screening method, particularly when direct nt sequencing facilities are unavailable
(XU et al. 2009). The LAMP assay is based on the principle of autocycling strand
displacement DNA synthesis performed by the Bst DNA polymerase and a set of two
inner and two outer primers that recognize 6–8 regions of target DNA. LAMP has
high specificity as the primers recognize six specific regions of the target amplicon
(TSAI et al. 2012).
28
2.6. Vaccines and vaccination against IBDV
Vaccination is widely used to prevent IBD outbreaks in the field. Most of the
commercially available vaccines against IBDV are live attenuated and inactivated
ones; recombinant and subunit vaccines have been licensed in some countries.
Live vaccines are produced from classical and variant IBDV strains by passaging
these viruses in tissue cultures or embryonated chicken eggs (YAMAGUCHI et al.
1996a; LASHER et al. 1997; JACKWOOD u. SOMMER-WAGNER 2011). They can
be classified as mild, intermediate or intermediate plus vaccines based on the level of
attenuation and residual virulence for SPF chickens (VAN DEN BERG et al. 2000a).
The intermediate plus vaccines are regularly applied to protect chickens against
vvIBDV challenges. The Deventer formula may help to determine the optimal time for
IBDV vaccination to circumvent the neutralizing activity of MAB (DE WIT 1998). Live
vaccines are favourable for mass application through drinking water and can induce
strong humoral and cellular immunity (MÜLLER et al. 2003; MÜLLER et al. 2012).
The proven reversion to virulence (YAMAGUCHI et al. 2000) and their residual
immunosuppressive effects (RAUTENSCHLEIN et al. 2005b; RAUTENSCHLEIN et
al. 2007) are major safety concern of their extensive field applications. Breeder
vaccination by priming with live vaccines and boosting with inactivated oil-emulsion
vaccines prior to laying ensures higher levels of MAB transfer to the progeny (MAAS
et al. 2001; MÜLLER et al. 2012) and is applied in some countries.
Commercially available IBD immune complex (IBD-ICX) vaccines are found to be
safe and efficacious for in ovo and posthatch vaccination of broilers (HADDAD et al.
1997; GIAMBRONE et al. 2001; IVAN et al. 2005). They are prepared by combining
an IBDV-hyperimmune serum with live intermediate plus IBDV (WHITFILL et al.
1995; JOHNSTON et al. 1997). The entrapment and retention of ICX on bursal
follicular dendritic cells (FDCs) and on splenic FDCs in the germinal center were
suggested as the immune enhancing mechanism of such vaccines (JEURISSEN et
al. 1998). The viruses are released from the ICX when the levels of MAB declined to
induce specific humoral immune responses that protect chickens against challenge
29
virus. A recombinant neutralizing Ab has been evaluated for formulation of an IBDICX vaccine (IGNJATOVIC et al. 2006).
The protective effects of many recombinant IBDV vaccines were evaluated under
experimental and field conditions. The polyprotein (PP), mature VP2 or immunogenic
domains of VP2 of pathogenic IBDV strains were targeted to produce candidate
vaccines: subunit, vectored, virus-like particles (VLPs) and chimeric virus particles.
Some of these experimental vaccines are presented in table 4.
An IBDV-VP2 subunit vaccine expressed in Pichia pastoris is licensed for commercial
uses (PITCOVSKI et al. 2003). An E. coli expressed subunit vaccine has been
evaluated under field conditions (RONG et al. 2007). The use of peptide epitope
mimics, i.e. mimotopes as candidate IBDV vaccines have become promising
strategy. Mimotopes are chemically synthesized and resembled the neutralizing
epitopes of VP2. Their expression in prokaryotic expression vector resulted in a
bioactive peptide that can induce significant neutralizing Abs and protection against
IBDV challenge (WANG et al. 2007). These types of vaccines induce strong humoral
immunity and always require adjuvants and multiple injections for inducing protective
levels of neutralizing Abs. Many live vectored IBDV vaccines, which mimic natural
infection have been developed and tested for efficacy. A live Newcastle disease virus
(NDV) vectored VP2 vaccine has been experimentally evaluated (HUANG et al.
2004) and recently HVT-IBD vaccine was licensed for in ovo and posthatch
vaccination of broilers and layers in various countries (BUBLOT et al. 2007; LE
GROS et al. 2009). These vectored vaccines induce strong systemic neutralizing Ab
levels and mucosal Abs, but pre-existing immunity for example against NDV-vector
may affect their efficacy.
Other IBDV-candidate vaccines include virus-like particles (VLPs). These vaccines
lack viral genomes and are non-infectious. They preserve the native conformation of
the capsid protein and present multiple copies of these immunogenic epitopes
(BRUN et al. 2011). However, the expression systems determine the nature of the
VLPs. The expression of IBDV PP by a recombinant vaccinia virus in mammalian
30
cells resulted in true VLPs (FERNANDEZ-ARIAS et al. 1998), whereas defective
VLPs were detected when the PP was expressed in insect cells by a baculovirus (HU
et al. 1999; KIBENGE et al. 1999; CHEVALIER et al. 2002). The main reason for the
lack of true VLP formation in the yeast and insect cells may be the absence of the
host protease, puromycin-sensitive aminopeptidase that is required for the
processing of the PP (IRIGOYEN et al. 2012). The VP2 icosahedral capsid had been
shown to induce the higher neutralizing Ab levels and better protection against IBDV
challenge than the PP-derived structures and the VPX tubules (MARTINEZTORRECUADRADA et al. 2003).
Candidate attenuated live IBDV vaccines generated by reverse genetics have been
shown to induce strong protective immunity (BOOT et al. 2002; MUNDT et al. 2003;
ZIERENBERG et al. 2004; QIN et al. 2010; GAO et al. 2011), but vaccinated
chickens developed milder bursal lesions after a challenge study. These tailored
chimeric IBDV vaccines were generated to contain VP2 regions of two different
strains of serotype 1 IBDV (MUNDT et al. 2003; GAO et al. 2011) or were chimeric
between segment A of serotype 1 and segment B of serotype 2 IBDVs
(ZIERENBERG et al. 2004). A VP5 mutant IBDV vaccine induced better protection
than its molecular cloned PP counterpart (QIN et al. 2010). BOOT et al. (2002)
produced a chimeric virus containing the C-terminal serotype 2 VP3 inserted into
genome segment A of serotype 1 IBDV. Nevertheless, the risk of reversion to
virulence of these genetically modified viruses hinders their field applications (RAUE
et al. 2004).
Recombinant IBDV-VP2 vaccines may possibly be used as ‘’marker vaccines’’
(MÜLLER et al. 2012) by allowing the differentiation of infected from vaccinated
animals (DIVA) by the detection of anti-VP3 Abs in naturally infected birds.
31
Table 4: Experimental IBDV vaccines
Candidate
Targeted immunogenic
Expression system or
Evaluation of immunity/protection
vaccines
domain
vectors
after challenge
Subunit
hVP2
Pichia pastoris
30% morbidity and mortality, IBDV Ag detected
References
(VILLEGAS et al. 2008)
in the bursa
N-terminal VP2 (aa 18–139)
E. coli
↑ELISA-Ab titer, 100% protection from mortality,
(PRADHAN et al. 2012)
IBDV Ag detected in the bursa
Mimotope
E. coli
↑ELISA- and VN-Ab titer, 100% survival rate
(WANG et al. 2007)
VP2
Plants
Seroconverted, 80% protection from mortality
(WU et al. 2004; WU et al. 2007b)
Chimeric virus
Neutralizing epitope from the
Bamboo mosaic virus
↑ ELISA-Ab titer, mild to moderate bursal lesions
(CHEN et al. 2012b)
particles
PBC loop
Mimotope polypeptide
Human hepatitis B virus
↑ELISA- and VN-Ab titer, 100% survival rate
(WANG et al. 2012c)
VP2
Fowlpox virus
14% and 33% of the chickens protected from
(TSUKAMOTO et al. 2000)
Live vectored
virus
gross and histological lesions, respectively
Marek’s disease virus
55% protection from bursal lesions, IBDV
(TSUKAMOTO et al. 1999)
infection was not prevented
Live bacterial-
VP2
Semliki forest virus
Some levels of neutralizing Abs detected
(PHENIX et al. 2001)
Vaccinia virus
VN-Ab titer demonstrated
(ZANETTI et al. 2012)
Avian adenovirus
↑VN-Ab titer, mortality up to 20%
(FRANCOIS et al. 2004)
T4 bacteriophage
Ab detected, partial protection
(CAO et al. 2005)
E. coli
Over 95% protection from mortality,
(MAHMOOD et al. 2007)
delivered
seroconversion detected
PP
S. Typhimurium
73% protection from mortality and
seroconversion detected
(LI et al. 2006)
32
2.7. DNA vaccines
Over the last two decades, recombinant DNA vaccines have been developed and
evaluated for protection against several infectious and non-infectious diseases in
humans and animals. These DNA vaccines have been shown to elicit cellular and
humoral immunity. Whilst several experimental studies confirmed their protection in
challenge model studies, in almost all cases the results are not transferable to a large
scale commercial application. A plasmid based DNA vaccine against West Nile virus
was licensed for horses (DAVIS et al. 2001) and against infectious hematopoietic
necrosis virus for fish (GARVER et al. 2005). Normally DNA vaccines are applied
intramuscularly. The intramuscular (IM) administration is, however, less efficient in
inducing strong immunity. Plasmids are distributed systemically from the site of
injection, and may then be degraded by the endonuclease activity of the host tissues
(FAUREZ et al. 2010). In mammals, several other DNA vaccine delivery methods
have been investigated to improve the immunogenicity of these candidate vaccines.
Examples include parenteral administration with electroporation and gene gun
delivery methods (YAGER et al. 2009; MURAKAMI u. SUNADA 2011). In the avian
species, the IM route is the most frequently used one, however, mucosal (PETERS
et al. 2004) and in ovo (PARK et al. 2009) routes have been used for DNA vaccine
delivery. Very recently, the gene gun technology has been evaluated to administer
DNA vaccine and the subsequent immunological responses (NIEDERSTADT et al.
2012). In ovo electroporation of exogenous plasmid DNA encoding green fluorescent
protein (GFP) had been conducted to study the biological development of chicken
embryos (FARLEY et al. 2011).
2.7.1. Stimulation of immune cells by DNA vaccines
The eukaryotic expression cassette of a typical plasmid DNA vaccine consists of a
transcriptional promoter (for example the cytomegalovirus intron A, CMV), the
translatable gene and transcriptional terminator and polyadenylation signal sequence
(for example the bovine growth hormone) (STEPENSON et al. 2008). Following
33
inoculation of the plasmid DNA by the IM route, the de novo production of the Ag
mimics infection by a pathogen. The Ag will be efficiently presented by MHC class I
and II molecules for the stimulation of humoral and cell mediated immunity. Two
scenarios are proposed regarding the mechanisms by which the immune responses
are mounted after IM application of DNA vaccines (PARDOLL u. BECKERLEG
1995). The APCs may be directly transfected and immunity is initiated in the context
of MHC class I molecules. The intramuscularly applied plasmid DNA transfects
mostly myocytes (ULMER et al. 1996). The Ag may be nibbled from the surface of
myocytes by APCs (the so-called cross presentation) via the scavenger receptors
(SRs) and presented in the context of MHC class II molecules to induce the desired
Ag specific immunity (HARSHYNE et al. 2003). SRs are expressed on the surface of
macrophages and dendritic cells and as secreted proteins. SRs have been described
in chickens (HE et al. 2009b). The transfected muscle cells show increased
expression of MHC class I and co-stimulatory molecules suggesting their roles in
endogenous Ag presentation directly to CD8+ T-cells (SHIROTA et al. 2007;
MARINO et al. 2011).
2.7.2. IBDV DNA vaccines
A number of studies were conducted to evaluate the immunogenicity and protective
efficacy of IBDV DNA vaccines following their IM administration to commercial
broilers and SPF chickens. The VP2 or the PP gene of pathogenic IBDV strains were
thus cloned into eukaryotic expression vectors as candidate DNA vaccines.
Experimental studies demonstrated that IBDV-DNA vaccinated chickens may or may
not develop neutralizing Abs (LI et al. 2003; SUN et al. 2005; HSIEH et al. 2010;
CHEN et al. 2011c). Previously, it has been speculated that CMI may be involved in
IBDV DNA vaccine-induced protection. This has been forwarded due to protection of
challenged chickens in the absence of higher levels of Ab responses (CHANG et al.
2001; CHANG et al. 2003; KIM et al. 2004). In vitro and in vivo lymphoproliferative
assays showed that IBDV DNA vaccines might induce significant T-cells responses
(KIM et al. 2004; LI et al. 2006; HSIEH et al. 2007; MAHMOOD et al. 2007; HSIEH et
34
al. 2010) including memory T-cells in vaccinated chickens (HSIEH et al. 2007, 2010).
Broiler chickens vaccinated with DNA vaccines showed increased T-cell responses
indicating no interference of T-cell immunity with existing MAB (HSIEH et al. 2010).
DNA vaccines applied intramuscularly induced only few prechallenge T-cells in the
bursa as demonstrated immunohistochemically (CHEN et al. 2011c). These
responses might provide some levels of protection presumably by clearing virus
infected cells during challenge.
However, the results of most challenge studies showed variable success of
protection. Chickens vaccinated with these candidate DNA vaccines at higher doses
(200 µg to 7 mg/bird) showed reduced clinical signs and bursal lesions after
challenge (HSIEH et al. 2010). IBDV DNA vaccines evaluated after in ovo application
and through mucosal routes conferred partial protection (PETERS et al. 2004; PARK
et al. 2009). Generally, IBDV DNA vaccines showed good immunogenicity when
administered at higher doses, however, this remains impractical for field applications.
Several adjuvants and adjuvant formulations have been investigated to improve the
immunogenicity and protection provided by these vaccines (section 2.8 and 2.9).
2.8. Adjuvants
Vaccine adjuvants are compounds that can modulate the intrinsic immunogenicity of
an Ag and elicit strong and long lasting immune responses without having any
specific antigenic effect by themselves (VOGEL 1998). They can be derived from
natural compounds such as minerals, microbial components or natural products of
animal and plant origins (BEREZIN et al. 2010).
Most inactivated and subunit avian vaccines are formulated as oil-in-water emulsion
or in alum (SWAYNE et al. 2012). In this case, montanide was used as oil-in-water
preparation for experimental and commercial vaccines. This vaccine emulsion has
been shown to induce strong humoral responses with reduced reactogenicity at the
injection sites (JANG et al. 2011; LIU et al. 2011). These adjuvants provide slow
35
release of Ags at the injection sites. Microbial components such as cholera and E.
coli heat-labile toxins have been investigated for example with inactivated avian
influenza vaccines as mucosal adjuvants in chickens and induced strong mucosal
and systemic immunity (HIKONO et al. 2012; MILLER et al. 2012). These
conventional adjuvants in general promote strong Ab responses, yet show limited
ability to induce CMI. New classes of molecular adjuvants are required, which
enhance immune responses by the new generation vaccines such as DNA vaccines.
2.8.1. Cytokine adjuvants for avian viral vaccines
Cytokines are small soluble proteins secreted mainly by immune cells. They play
pivotal roles in cell-to-cell signaling and thus control immune responses to
pathogenic infections and vaccination (BAROUCH et al. 2004). Recently, several
avian cytokines have been cloned and their biological activities have been
investigated in vitro and in vivo. Over 23 interleukins, eight type I IFNs, IFN-γ, IFN-λ,
and GM-CSF are characterized in chickens (KAISER et al. 2005; KARPALA et al.
2008; KAISER 2010). The functional domains of most of the avian cytokines (GU et
al. 2010a; GU et al. 2010b; KIM et al. 2012), which interact with the receptors on the
surface of the immune cells are sufficiently characterized (GU et al. 2010c; JEONG
et al. 2012; KIM et al. 2012). The intracellular signalling cascades (JAK/STAT) that
result in the modulation of the immune response have been well determined (KAMPA
u. BURNSIDE 2002; OSINALDE et al. 2011).
The multifunctional effects of these cytokines on the innate and adaptive immune
system of the host include the activation and differentiation of APCs, B- and T-cell
proliferation, and induction of the expression of costimulatory molecules (HILTON et
al. 2002). Chickens inoculated with recombinant chIL-2 (HILTON et al. 2002) or
plasmid encoded chIL-2/IL-15 (LILLEHOJ et al. 2001) developed an increase in the
numbers of peripheral blood CD4+- and CD8+ T-cells. These cells expressed higher
levels of the IL-2 receptor (IL-2R) (HILTON et al. 2002). Potent therapeutics and
vaccine adjuvant effects of several avian cytokines were described for poultry
36
(LILLEHOJ et al. 2001; KAISER 2010; KUMAR et al. 2010). Cytokine adjuvants can
be classified as Th1 promoting and include IL-2, IFN-γ, IL-12, IL-15, and IL-18. These
cytokines can augment CMI. Th2 promoting cytokines like granulocyte-macrophage
cell stimulating factor (GM-CSF) and IL-4 enhance humoral immunity (RAHMAN u.
EO 2012). Simultaneous application of Th1 and Th2 cytokine adjuvants may result in
balanced Th1/Th2 immune responses to an Ag.
The major advantages of cytokine adjuvants are their suitability to be administered
with DNA, killed, and live viral vaccines (Table 5 and 6). They may show synergistic
effects when two or more of them are administered simultaneously. They can be
delivered as recombinant cytokines (XIAOWEN et al. 2009), and have been shown to
induce significant humoral immune responses when co-administered for example
together with inactivated Newcastle disease (ND) and infectious bronchitis (IB)
vaccines (DEGEN et al. 2005; WANG et al. 2006; HUNG et al. 2010). However, the
major limitations of these recombinant cytokines as vaccine adjuvants are their
extraordinary short in vivo half-life time. To circumvent this problem, cytokineencoding plasmids are utilized and co-administered with Ags (BAROUCH et al.
2004). Large numbers of plasmid encoded cytokines have been evaluated as DNA or
conventional vaccine adjuvants for several viral Ags in chickens (Table 5). They may
be cloned into a bicistronic eukaryotic expression vector together with the
immunogenic target gene to enhance the protective levels of both humoral and
cellular immunity (KUMAR et al. 2009).
Alternatively, cytokine genes encoded by live viral vectors such as HVT, fowlpox
(FPV) and adenovirus (Ad) vectors may enhance neutralizing Ab production of the
co-administered vaccines (Table 6). Orally administered live attenuated S.
Typhimurium that expressed IL-18 modulated both humoral and Th1-biased CMI
against the co-delivered inactivated avian influenza vaccine (RAHMAN et al. 2012).
Expression of chIL-2 as fusion protein with influenza neuraminidase VLPs has been
shown to induce strong humoral immunity when applied together with inactivated
influenza A (H3N2) vaccine (YANG et al. 2009).
37
Although the adjuvant properties are well recognized, none of the cytokine-based
adjuvants are licensed so far. Appropriate methods of cytokine delivery still need to
be developed.
38
Table 5: Improvement of the protective efficacy of avian viral vaccines after co-administration of plasmid encoded avian cytokines
Cytokines
Vaccines
Evaluation of immunity/protection after challenge
References
IFN-γ
Live cell-free HVT
↓tumor incidence after vMDV challenge
(HAQ et al. 2011)
NDV-DNA
↑HI-Ab titer, ↑T-cell proliferation
(SAWANT et al. 2011)
Live ND
44% of the live chicks had IgM after vaccination
(DILAVERIS et al. 2007)
Inactivated ND
↑HI-Ab, ↑↑IFN-γ, ↑↑CD8+ T-cells, 100% protection
(HUNG et al. 2010)
IL-18
against NDV
ILTV-DNA
↑CD8+ T-cells, ↑IFN-γ, 80% protection against ILTV
(CHEN et al. 2011b)
AIV-DNA
↑HI-Ab, ↑CD4+ T-cells
(LIM et al. 2012)
IL -15
AIV-DNA
↑↑HI-Ab, ↑↑CD4+ T-cells
(LIM et al. 2012)
IL-4
NDV-DNA
↑↑↑HI-Ab, 40% survival after challenge with
(SAWANT et al. 2011)
vNDV
GM-CSF
IBV-DNA
↑ELISA-Ab, ↑T-cell proliferation, 86% protection
(TAN et al. 2009)
IL-18
IBDV-DNA
↑ELISA-Ab, ↑T-cell proliferation, ↑IL-4 & IFN-γ mRNA expression, 93%
(LI et al. 2013a)
protection from clinical signs & mortality, milder bursal lesions
IL-6
IBDV-DNA
↑ELISA-Ab, 60% protection, milder bursal lesions
(SUN et al. 2005)
IBDV-DNA*
↑ELISA-Ab, 90% protection, milder bursal lesions
(SUN et al. 2005)
NDV-DNA is a bicistronic plasmid encoding the HN & F gene of NDV; ILTV-DNA represents a plasmid encoding glycoprotein B (gB) of ILTV; AIV-DNA encodes the
HA of H5N1; IBV-DNA encodes S1 gene of IBV; IBDV-DNA represents a plasmid encoding VP2 of IBDV & IBDV-DNA* encodes the PP gene. IBDV= infectious bursal
disease virus; AIV= avian influenza virus; ILTV= infectious laryngotracheitis virus; HVT= herpesvirus of Turkeys; NDV= Newcastle disease virus. ↑ denotes increase
and ↓ denotes decrease.
39
Table 6: Examples for avian cytokines delivered by live viral vectors as adjuvants for avian viral vaccines
Cytokine vectors
Vaccines/target immunogen Evaluation of immunity/protection after challenge
References
rHVT-IL-2
HVT
↑VN-Ab, 43% protection against vMDV
(TARPEY et al. 2007a)
Live IBV
↑ELISA-Ab, 83% protection against vIBV, ↓ciliostasis
(TARPEY et al. 2007b)
rFPV-HN-IL-12
HN-NDV
↑↑HI-Ab, 100% protection against vNDV
(SU et al. 2011)
rFPV-IL-12
rFPV-HN-NDV
↑HI-Ab, ↑IFN-γ, 83% protection against vNDV
(SU et al. 2011)
rFPV-HA-IL-6
HA-H5N1
↑HI-Ab, ↑T-cell proliferation, 80% protection against H5N1, (QIAN et al. 2012)
↓virus shedding
rFPV-gB-IL-18
gB-ILTV
↑CD4: CD8+ T-cell ratio, 100% protection against vILTV
(CHEN et al. 2011a)
rFPV-HA-IL-18
rFPV-HA (H9N2)
↑HI-Ab,↑↑ T-cell proliferation, no viral shedding
(CHEN et al. 2011a)
rAd-GM-CSF-S1
S1-IBV
↑HI-Ab, ↑↑IFN-γ, 100% protection, mild renal lesions
(ZESHAN et al. 2011)
rAd-GM-CSF
rAd-S1-IBV
↑HI-Ab, ↑CMI, 100% protection against IBV, mild renal (ZESHAN et al. 2011)
lesions
rFPV-IL-12
rFPV-VP2
↑VN-Ab, 100% protection from challenge, some birds with (SU et al. 2011)
bursal lesions
rHVT-IL-2 represents recombinant HVT expressing IL-2; rFPV-HN-IL-12 coexpresses HN of NDV & IL-12; rFPV-HA-IL-6 coexpresses HA of H5N1 & IL-6; rFPV-gB-IL18 represents FPV co-expressing gB of ILTV & IL-18; rAd-GM-CSF-S1 co-expresses GM-CSF & S1of IBV; rFPV-IL-12 expresses IL-12; rFPV-HN-NDV expresses the
HN gene of NDV; rFPV-HA (H9N2) expresses HA of H9N2; rAd-S1-IBV expresses S1 gene of IBDV; rFPV-VP2 expresses VP2 of IBDV. rFPV= recombinant fowlpox
virus; ILTV= infectious laryngotracheitis virus; rHVT= recombinant herpesvirus of Turkeys; rAd= recombinant Adenovirus. ↑ denotes increase and ↓ denotes decrease.
40
2.8.1.1. Chicken IL-2 and IFN-γ as adjuvants for IBDV-DNA vaccines
IL-2 is mainly secreted by T helper cells. Apart from its proliferative effects on T-cells
and APCs (HILTON et al. 2002), IL-2 promotes growth of chickens (FORD et al.
2002). Co-administration of plasmid encoded chIL-2 together with plasmid encoded
IBDV-VP2 or PP genes promoted IBD-specific neutralizing Ab production and T-cell
responses. The vaccinated SPF chickens were partially protected against lethal virus
challenge (HULSE u. ROMERO 2004; LI et al. 2004; KUMAR et al. 2009). The
immune potentiating effect of IL-2 was found to be significant when cloned together
with the VP2 gene into a bicistronic eukaryotic expression vector (KUMAR et al.
2009). In contrast to these findings, lack of adjuvant effects of plasmid-encoded chIL2 was demonstrated after in ovo application with an IBDV DNA vaccine (PARK et al.
2009). The co-administration of plasmids encoding chIFN-γ and IBDV-PP did not
enhance the immune response and protection against IBDV (ROH et al. 2006).
Repeated IM co-administration of chIFN-γ and PP encoding plasmids at the same
injection sites was found to be immunosuppressive and challenged chickens
developed more severe bursal lesions compared to the administration of the two
plasmids at different sites (HSIEH et al. 2006).
2.8.2. Toll like receptor (TLR) ligands as adjuvants
An antigen presenting cell (APC) uses its pathogen recognition receptors (PRRs)
such as TLRs to recognize pathogen-associated molecular patterns (PAMPs), which
include viral nucleic acids, bacterial DNA, and others (LI et al. 2010). For chickens,
ten different types of TLRs expressed by APCs such as DCs (WU u. KAISER 2011),
monocytes (HE et al. 2006), B-cells (ST PAUL et al. 2012d), thrombocytes (ST PAUL
et al. 2012c), and heterophils (KOGUT et al. 2005), CD4+- and CD8+ T-cells
(SHANMUGASUNDARAM u. SELVARAJ 2011; ST PAUL et al. 2012a) and
erythrocytes (ST PAUL et al. 2013) have been described. TLR3, TLR7 (PHILBIN et
al. 2005) and TLR21 are localized in the endosomes of APCs and sense viral and
41
bacterial nucleic acids, respectively. Core TLR signalling cascades will be initiated
after the recognition of PAMPs, which in turn leads to activation and upregulation of
the host defense genes such as type I IFNs, chemokines and cytokines (CIRACI u.
LAMONT 2011). The cytokine responses can amplify the innate immune responses
and activate adaptive immunity (CHEVRIER et al. 2011).
In mice and humans, CpG-ODNs may reach to the endosomally located TLR9
through the lectin receptor referred to as DEC-205 (LAHOUD et al. 2012) or CD14
(BAUMANN et al. 2010). Granulin, which is a soluble protein, binds to CpG-ODN and
may serve as co-factor to recruit CpG-ODN to TLR9 (MORESCO u. BEUTLER 2011;
PARK et al. 2011). The chicken TLR21, which is functionally orthologous to the
mammalian TLR9, recognizes synthetic oligodeoxynucleotides (ODN) containing
unmethylated CpG motifs (CpG-ODN) or bacterial DNA that normally contains this
motif (BROWNLIE et al. 2009; KEESTRA et al. 2010). The chicken bone marrowderived dendritic cells expressed higher DEC-205 levels in response to stimulation by
lipopolysaccharide (WU et al. 2010), yet its role as receptor for CpG-ODN uptake in
avian species remains unclear.
Synthetic class B CpG-ODN is a single stranded short DNA sequence that contains
phosphorothioated internucleotide linkage, which confer resistance of the DNA to
degradation by nucleases (SESTER et al. 2000). They have been extensively
investigated for their in vitro immunostimulatory properties in avian cells. They can
activate avian APCs (HE et al. 2006; WU u. KAISER 2011) and B-cells (WATTRANG
2009) that responded to CpG-ODN stimulation through the production of NO, IL-1β,
and IFN-γ (BHAT et al. 2010). Avian monocytes showed a strong Th1-biased immune
response to stimulation by CpG-ODN and TLR3 agonist (Poly I: C) (HE et al. 2012a).
CD4+ T-cells and B-cells isolated from the bursa responded to CpG-ODN through
the production of IFN-γ (ST PAUL et al. 2012a). B-cells showed upregulation of
transcripts associated with Ag presentation such as CD80 and MHC class II
molecules (ST PAUL et al. 2012d). In vivo treatment of chickens with CpG-ODN
promoted the upregulation of MHC class II and IFN-γ mRNA expression in the
42
spleen, indicating the role of CpG-ODN in enhancing Ag presentation (ST PAUL et
al. 2011).
Prophylactic effects of class B CpG-ODNs after in ovo or parenteral application have
been reported against Eimeria (DALLOUL et al. 2004), E. coli (GOMIS et al. 2004)
and S. Typhimurium (TAGHAVI et al. 2008) infections in chickens. High mRNA
expression levels of proinflammatory cytokines, IFN-γ and chemokines in the spleen
of chicken embryos inoculated with CpG-ODN at embryonation day 18 prior to IBV
challenge has limited IBV replication (DAR et al. 2009). Chickens that had received
CpG-ODN and were challenged with a low pathogenic avian influenza virus showed
reduced virus shedding (ST PAUL et al. 2012b).
Several synthetic class B CpG-ODNs have been investigated as vaccine adjuvants in
mammals and avian species (MAHMOOD et al. 2006; DUTHIE et al. 2011; MALLICK
et al. 2012). They have been shown to enhance the immunogenicity and efficacy of
inactivated viral vaccines (WANG et al. 2009b; XIAOWEN et al. 2009) and
recombinant poultry vaccines (MALLICK et al. 2012). Additional work showed that
chickens intramuscularly inoculated with CpG-ODN and IBDV-VP2 DNA vaccine
exhibited higher levels of protective immunity and survival following lethal virus
challenge (WANG et al. 2003; MAHMOOD et al. 2006). IBDV DNA vaccine combined
with CpG-ODN induced CMI as indicated by increased delayed hypersensitivity
reactions (MAHMOOD et al. 2006; MAHMOOD et al. 2007). In contrast, ROH et al.
(2006) reported that CpG-ODN had no significant effects on the immunogenicity of
an IBDV PP DNA vaccine when co-administered intramuscularly to chickens.
2.9. Microparticulate vaccine and adjuvant carrier systems
The hurdle of poor immunogenicity of DNA and peptide vaccines may be resolved
through effective microparticulate vaccine carrier systems (SINGH et al. 2000;
O'HAGAN et al. 2004b; FOGED 2011; DE TEMMERMAN et al. 2012). These carrier
systems are mostly prepared from poly(D, L-lactide-co-glycolide) (PLGA), which is a
43
versatile biodegradable polymer used to deliver drugs, hormones, and vaccines in
human medicine (LANGER et al. 1990). PLGA MPs undergo hydrolysis in vivo to
non-harmful and non-toxic compounds that are readily excreted from the body
(BENDIX 1998).
The formulation of vaccines with PLGA MPs increased the persistence of Ag at the
injection sites. MPs thus serve as tailored slow control release system for the Ags
that are expressed in situ in myocytes. Ags are released either by diffusion or by
hydrolysis of MPs and this quality of MPs may avoid the application of booster
vaccinations (SINGH et al. 1997). They are suitable carriers to deliver multiple copies
of different Ags to the same APCs. The targeted delivery of Ags to APCs by these
MPs may have Ag dose sparing effects and is an advantage if large numbers of
animals need to be vaccinated (O'HAGAN et al. 2006).
At the same time the recruitment of large numbers of phagocytes to the sites of IM
injection of PLGA MPs may allow Ag presentation by these cells following
phagocytosis (DENIS-MIZE et al. 2003). PLGA MPs with a diameter of 0.5-10 µm
can be readily phagocytosed by APCs (JILEK et al. 2005; O'HAGAN et al. 2006;
JAIN et al. 2011) with the help of β2 integrins expressed on the surface of APCs
(ROGERS u. BABENSEE 2011). This may be followed by an immediate increase in
the levels of expression of costimulatory molecules (CD86) and NF-κB, indicating
immune system stimulation by PLGA MPs (NICOLETE et al. 2011). The maturation
process of APCs in mammals was further improved by delivery of DNA vaccine
adsorbed onto the surface of positively charged PLGA MPs (JILEK et al. 2004;
REDDY et al. 2012).
The processes of microencapsulation or adsorption of vaccines and adjuvants are
outlined in Fig.1.
44
1st homogenization
2nd homogenization
Solvent evaporation
DNA/protein solution
PLGA solution
(Aqueous phase)
(Organic/oily phase)
Surfactant solution
Water-in-oil emulsion
(Aqueous phase)
Water-in-oil-in-water emulsion
and MPs separation
(W/O/W)
PLGA MPs hardening
Lyophilization &
Lyophilized PLGA MPs
ready to use
Fig.1. Method of vaccine (DNA/protein) microencapsulation in PLGA microspheres
using the w/o/w emulsion solvent evaporation method (SINGH et al. 1997; TINSLEYBOWN
et
al.
2000;
FREITAS
et
al.
2005;
OSTER
u.
KISSEL
2005).
PLGA/polyethylenimine (PEI) is dissolved in dichloromethane. Surfactant contains
polyvinyl alcohol to coat the outer PLGA matrix. This prevents MPs aggregation and
facilitates their uptake by phagocytic cells by influencing the structure and
hydrophobicity of MPs. PEI may be added to the oily phase to condense DNA and
reduces DNA vaccine degradation during microencapsulation (ZHANG et al. 2008).
Preferentially, DNA can be adsorbed onto cationic PLGA MPs to prevent the
destruction
of
the
supercoiled
DNA
structure
of
the
plasmid
during
microencapsulation (OSTER et al. 2005). Cetyltrimethylammonium bromide (CTAB)
is added to the surfactant solution to provide a positive surface charge to PLGA MPs
to readily adsorb DNA (O'HAGAN et al. 2004a; SINGH et al. 2006). CTAB binding to
the surface of MPs could be facilitated by addition of PEI in the oily/organic phase,
45
which neutralizes the anionic charges of the PLGA-matrix (OSTER et al. 2005).
During surface adsorption, plasmid DNA is incubated with cationic PLGA MPs for
adsorption through electrostatic interactions (OSTER et al. 2005; LIMAN et al. 2007).
Vaccine adjuvants can be entrapped in or adsorbed onto PLGA MPs in similar ways.
2.9.1. Enhancing specific immunity by PLGA MPs
Several early studies have shown the induction of strong immunological responses
following parenteral or mucosal administration of microencapsulated protein or DNA
vaccines against several infectious diseases including HIV-AIDS (O'HAGAN 2001;
SINGH et al. 2001b; WILKHU et al. 2011). The IM administration of cationic PLGA
MPs with an adsorbed hepatitis B virus DNA vaccine to mice was shown to induce
significantly higher Ab levels, CMI, and protection compared to the naked DNA
vaccine. This may probably be due to improved transfection of inflammatory
phagocytic cells or due to persistent Ag expression from the transfected myocytes
(HE et al. 2005). An intramuscularly delivered anti-foot and mouth disease DNA
vaccine adsorbed onto cationic MPs induced a long-term immune response against
challenge (REDDY et al. 2012). The combination of using electroporation before or
after IM application of PLG-encapsulated plasmid DNA enhanced the expression of
the Ag, which resulted in a more robust T-cell response (BARBON et al. 2010).
Mucosal delivery of DNA vaccines entrapped in PLGA MPs or adsorbed onto the
surface of cationic MPs have been shown to induce systemic and mucosal immunity
in several animal models for different infectious diseases (SINGH et al. 2001b;
WANG et al. 2011; BHOWMIK et al. 2012). MPs may prevent DNA vaccine
degradation by mucosal enzymes (JONES et al. 1998). Recently, the usefulness of
PLGA MPs as vaccine delivery systems to avian species was introduced (PEISER
2006; LIMAN et al. 2007). An oculonasally administered avian Metapneumovirus
(aMPV) fusion protein-encoding plasmid DNA and the corresponding subunit vaccine
by PLGA MPs in a prime-boost approach to turkeys induced mucosal and cellular
immunity (LIMAN et al. 2007).
46
There is a need to further enhance the immunogenicity of MP-entrapped or adsorbed
vaccines by additional adjuvants (O'HAGAN et al. 2006; FISCHER et al. 2009). While
CpG-ODN shows extremely potent immunomodulating effects on innate and adaptive
immune cells in vitro, it has become apparent that rapid in vivo degradation and an
inefficient delivery to target cells hinder its application as a novel vaccine adjuvant.
The adjuvant effects of several immunostimulatory molecules including CpG-ODN
and cytokines in non-avian vaccines were found to be superior when they were
administered adsorbed or microencapsulated in PLGA MPs compared to their
applications in soluble forms (SINGH et al. 2001a; MALYALA et al. 2008; MALYALA
et al. 2009; WANG et al. 2012b). The delivery of chIL-2 and Newcastle disease DNA
vaccine in the form of nanoparticles improved immunity against NDV challenge
(ZHANG et al. 2010). This formulation resulted in a higher HI-Ab titers and serum
IFN-γ levels compared to the naked DNA vaccine (ZHANG et al. 2010).
47
3. Goals and objectives
The present study was designed to understand the molecular epidemiology of IBDV
in Ethiopia and to develop a new generation vaccine against IBDV.
The objectives of the first part of the study include
-
To collect IBDV isolates from field outbreaks
-
To investigate their pathogenicity in vivo in chickens
-
To conduct genotypic characterization of the isolates and construct
phylogenetic trees.
The specific objectives in the second part of the study include
-
To construct IBDV DNA vaccines
-
To identify suitable molecular adjuvants
-
To improve the immunogenicity of the DNA vaccine by utilizing particulate
vaccine delivery system
-
To enhance the protection conferred by MP-based DNA vaccine by delivering
adjuvant loaded MPs
48
4. Molecular evidence of very virulent infectious bursal disease virus in
chickens in Ethiopia
Tamiru Negash, Esayas Gelaye, Henning Petersen, Beatrice Grummer and
Silke Rautenschlein
Avian Dis 2012 56:605-610.
http://www.aaapjournals.info/doi/full/10.1637/10086-022012-ResNote.1
49
SUMMARY.
Infectious
bursal
disease
virus
(IBDV)
is
an
important
immunosuppressive pathogen of chickens worldwide. The introduction and evolution
of IBDV in most African countries, especially in Ethiopia remains unclear. We have
investigated IBDV isolates obtained from commercial broilers, indigenous chickens
and pullets. The hypervariable region of the virus protein (VP) 2 (hVP2) and the 5`
two third of VP1 of eleven IBDV isolates were characterized by RT-PCR and further
sequencing. All isolates were identified as very virulent (vv) IBDV based on the
predicted amino acid (aa) sequences of the VP2 protein. Interestingly, the sequence
analysis of the 5` two third of VP1 indicated that the Ethiopian IBDV strains have aa
residues typical for vvIBDV and also for attenuated IBDV strains. Amongst all IBDV
strains included in this study for phylogenetic comparison of VP2 nucleotide
sequences, Ethiopian strains form a cluster within the vvIBDV lineage. We have
shown that Ethiopian IBDV strains have mutations in the VP1 region. Their roles in
IBDV virulence may require further in vivo studies. As depicted in this study, the
nucleotide and aa sequence analysis of VP1 in addition to VP2 is necessary to obtain
a clear picture of the molecular evolution of IBDV.
Key words: chicken, Ethiopia, infectious bursal disease virus, molecular
epidemiology
50
The extent of contribution from Tamiru Negash to this article is evaluated according
to the following scale:
A. has contributed to collaboration
(0-33%).
B.
has contributed significantly
(34-66%).
C.
has essentially performed this study independently
(67-100%).
1. Design of the project including design of individual experiments
C
2. Performing of the experimental part of the study
C
3. Analysis of experiments
C
4. Presentation and discussion of the study in article form
C
51
5. Mucosal application of cationic poly(D, L-lactide-co-glycolide) microparticles
as carriers of DNA vaccine and adjuvants to protect chickens against
infectious bursal disease
Tamiru Negasha, Martin Liman1, Silke Rautenschlein a*
a
Clinic for Poultry, University of Veterinary Medicine Hannover, Bünteweg 17, 30559
Hannover, Germany
1
Present address: AniCon Labor GmbH, Mühlenstraße 13 49685, Höltinghausen,
Germany
Vaccine (submitted)
52
Abstract
Infectious bursal disease virus (IBDV) is an immunosuppressive virus of
chickens. The virus protein (VP) 2 induces neutralizing antibodies, which protect
chickens against the disease. The aim of this study was to develop a cationic poly(D,
L-lactide-co-glycolide) (PLGA) microparticle (MP) based IBDV-VP2 DNA vaccine
(MP-IBDV-DNA) for chickens to be delivered orally and by eye drop route. The
tested IBDV-VP2 DNA vaccines were immunogenic for specific-pathogen-free
chickens and induced an antibody response after intramuscular application. Coinoculation with a plasmid encoding chicken IL-2 (chIL-2) or CpG-ODN did not
significantly improve protection against IBDV challenge. However, the application of
a MP-IBDV-DNA vaccine alone or in combination with a delayed oral and eye drop
application of cationic MP loaded with CpG-ODN or chIL-2 improved protection
against challenge. The MP-IBDV-DNA-vaccinated chickens showed less pathological
and histopathological bursal lesions, a reduced IBDV antigen load as well as T-cell
influx into the bursa of Fabricius (BF) compared to the other groups (p<0.05). The
addition of chIL-2 loaded MP improved challenge virus clearance from the BF as
demonstrated by lower neutralizing antibody titers and reduced IL-4 and IFN-α
mRNA expression in the bursa at 7 days postchallenge compared to the other
challenged groups. Overall, the efficacy of the IBDV-DNA vaccine was improved by
adsorption of the DNA vaccine onto cationic PLGA-MP, which also allowed mucosal
application of the DNA vaccine.
Key words: Adjuvant, chicken, DNA vaccine, IBDV, poly(D, L-lactide-co-glycolide)
microparticles
53
Abbreviations
ChIL-2= chicken interleukin 2; CpG-ODN = oligodeoxynucleotides (ODN) containing
CpG; DNA= deoxyribonucleic acid; IBDV= infectious bursal disease virus; MP=
microparticles; PLGA= poly(D, L-lactide-co-glycolide); Tregs= regulatory T-cells;
VP2= virus protein 2
54
1. Introduction
Infectious bursal disease virus (IBDV) causes an acute and economically
important immunosuppressive disease in poultry worldwide. IBDV is an Avibirnavirus
[1] with a bisegmented dsRNA genome [2], of which segment A encodes the virus
protein 2 (VP2) responsible for the induction of neutralizing antibodies [3].
IBDV replicates in the gut-associated lymphoid tissues and macrophages. The
main target cells are immature intrabursal B-cells [4]. IBDV causes severe lymphoid
cell depletion in the bursa of Fabricius (BF) [4,5]. IBDV infection of young chickens is
marked by severe mortality and immunosuppression, which leads to vaccination
failure and increases susceptibility of chickens to other pathogens [6].
Live IBDV vaccines are widely applied to commercial chickens to control IBDV
[7,8]. Nevertheless, these vaccines may revert to virulent strains [9], induce
immunosuppression [8] and may undergo segment reassortment with field strains
[10]. IBDV VP2 subunit vaccines have been developed [11,12], but production costs
of these subunit vaccines are usually high. Recently a vector vaccine on the basis of
herpesvirus of turkeys-expressing IBDV VP2 was licensed [13].
DNA vaccines against IBDV have been developed encoding-VP2 or the
polyprotein gene of IBDV. These vaccines often conferred only partial protection [1416]. It was demonstrated that IBDV DNA vaccine efficacy can be improved by
coadministering plasmid encoded chicken interleukin-2 (chIL-2) or CpG-ODN [17,18].
In mice and humans it was shown that DNA vaccination together with a delayed
application of cytokine adjuvants may further improve the protective responses after
DNA vaccination [19,20]. It is speculated that this strategy may reduce the expansion
of immunosuppressive regulatory T-cells (Tregs) [20], which may otherwise suppress
antigen specific T-cells.
Microparticulate carriers for DNA vaccines may provide additional adjuvant
effects and protect the DNA from degradation after mucosal application [21-23].
Poly(D, L-lactide-co-glycolide) (PLGA) is the polymer of choice for cationic
microparticle (MP) preparation [24]. Targeted delivery and prolonged persistence of
the DNA when coated onto cationic MP improve the duration of DNA-vaccine
induced immunity [21,25]. The objective of our study was to develop and test the
55
efficacy of an improved cationic MP based IBDV-DNA vaccine to be delivered by
mucosal routes to chickens in combination with a delayed mucosal delivery of
cationic MP adsorbing chIL-2 or CpG-ODN.
2. Materials and Methods
2.1. Chickens
Eggs from specific pathogen free (SPF) layer type chickens (VALO®,
Lohmann LSL-LITE) were obtained from Lohmann Tierzucht (Cuxhaven, Germany).
The hatched chickens were randomly distributed and maintained in different groups
under isolated conditions in the animal facilities of the University of Veterinary
Medicine Hannover. All animal experimentations were approved and carried out
following the institutional guidelines for animal care. Feed and water was supplied ad
libitum
2.2. Cloning and expression of VP2 and chIL-2
cDNAs of VP2 of the two very virulent (vv) IBDV strains 82Eth [26] and
89163/7.3 (provided by N. Eterradossi, AFSSA, Ploufragan, France) were
synthesized from total RNA using Oligo(dT)20 SuperScriptâ„¢ III RT kit (Invitrogen).
The RNA had been isolated from bursae of IBDV infected birds. The vvIBDV strains
had been isolated from several chicken flocks that had experienced high mortality
rate [26]. The sequence coding for chIL-2 was amplified from splenocytes of
chickens, which were stimulated in vitro with Concanavalin (Con) A (5 µg/mL) for 6
hrs [27]. PCR was conducted with the TaKaRa Ex Taqâ„¢ Polymerase (Takara Bio
Inc.,
Shiga,
Japan)
using
the
primers;
GCCGGTACCGACGCAGCGATGACAAACCTGC-3’;
CGGGCGGCCGCTGATCACCTTATGGCCCGGATTA-3’,
GCCGGTACCGACGCAATGATGTGCAAAGTACTGATC-3’;
CGGGCGGCCGCTGATTATTTTTGCAGATATCTCACAAA-3’).
VP2f:
VP2r:
5’-
chIL-2f:
and
KpnI
5’5’-
chIL-2r:
and
5’NotI
restriction sites were included (underlined in the primers). The PCR products were
56
double digested by FastDigest® NotI and FastDigest® KpnI (Fermentas GmbH, St.
Leon-Rot, Germany) and purified with NucleoSpin® Extract II PCR clean up and Gel
extraction kit (Macherey-Nagel, Düren, Germany). Double digested PCR products
were cloned into the pCR3.1® eukaryotic expression vector (Invitrogen) and
transformed into chemically competent TOP10F' E. coli (Invitrogen) following the
guidelines of the manufacturer. The inserts were verified by sequence analysis.
The expression of VP2 and chIL-2 was verified by immunofluorescence after
transfection of chicken embryo fibroblasts (CEFs) with the respective plasmids
(TransFectinâ„¢ lipid reagent; Bio-Rad, Hercules, California, USA). We used the
following primary Abs: rabbit anti-IBDV polyclonal Ab [28], or VP2 mAb (provided by
Egbert Mundt, FLI, Germany) and the mouse anti-chicken IL-2 mAb (AbD Serotec
MorphoSys,
Düsseldorf,
Germany);
and
secondary
antibodies:
fluorescein
isothiocyanate (FITC)-conjugated goat anti-rabbit IgG or FITC-conjugated goat antimouse polyvalent Ab (Sigma-Aldrich, St. Louis, USA). VP2 and chIL-2 expressing
plasmids were designated DNA-VP2/Eth82, DNA-VP2/7.3 and DNA-chIL-2. DNAvec represented control vector (pCR3.1).
2.3. chIL-2 bioactivity assay
Splenocytes from SPF chickens were seeded (2x 105 cells/well) in complete
RPMI medium (Biochrom AG, Berlin, Germany). Triplicates were stimulated with
ConA (5 µg/mL), supernatants from DNA-chIL-2 or DNA-vec transfected CEFs at
41°C and 5% CO2 [18]. Lymphocyte proliferation was measured by MTT (3-(4,5dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide) (Sigma, St. Louis, MO)
colorimetric assay [29].
2.4. Cationic MP preparation and DNA adsorption
Cationic MP were prepared as described previously [25,30] with the following
modifications. Three-hundred mg of PLGA Resomer® RG 753S (Boehringer
Ingelheim, Ingelheim, Germany) and 30 mg of Polyethylenimine (PEI; Sigma-Aldrich)
were dissolved in 3 mL dichloromethane (DCM; Sigma-Aldrich) and in 2 mL DCM,
respectively. The two organic solutions were mixed and homogenized with 0.5 mL
57
PBS, and subsequently with 25 mL of 0.5% Cetyltrimethylammonium bromide
(Boehringer Ingelheim, Ingelheim, Germany). Finally, the organic solvents were
evaporated and the cationic PLGA-MP were washed with 5% sucrose and
lyophilized.
Plasmid DNAs were adsorbed onto the MP by incubating 100 mg of MP
resuspended in distilled water (pH 5.0) with 1 mg of DNA as previously described
[25,30]. MP were washed and lyophilized. Phosphorothioate backbone modified
CpG-ODN
(5'-
TCGTCGTTGTCGTTTTGTCGTT-3';
CpG
motifs
underlined;
Biomers.net, Ulm, Germany) was adsorbed onto the MP following the protocol of
plasmid DNA adsorption. This particular CpG-ODN sequence has been selected
based on its significant immunostimulatory properties on avian immune cells; while
its non-CpG-ODN sequence has been shown to induce only marginal upregulation of
transcripts associated with antigen presentation [31,32].
The MP adsorbed with the DNA-VP2/Eth82, DNA-chIL-2, CpG-ODN, DNA-vec
and non-adsorbed MP were designated as VP2-MP, chIL-2-MP, CpG-MP, Vec-MP
and Mock-MP.
2.5. Characterization of MP
The diameter of the lyophilized MP was determined by the Grimm Aerosol
Spectrometer (Grimm Technologies, Douglasville, USA). The MP adsorption
efficiency for DNA was determined spectrometrically by subtracting the unbound
DNA left in the supernatants at the end of the adsorption period from the amounts of
DNA used for adsorption [25]. The resistance of the MP-adsorbed DNA to
degradation by enzymes was determined by incubating the loaded MP with 15 U of
DNAse I (Sigma). The integrity of the plasmid DNA was verified by gel
electrophoresis [25]. A 24 hrs in vitro-release of the MP-adsorbed DNA was
determined by incubating 10 mg loaded MP in PBS at 37°C. The amount of DNA
released into the medium was determined spectrometrically.
58
2.6. In vitro phagocytosis of MP
FITC-labeled bovine serum albumin (FITC-BSA) was microencapsulated as
described previously [25]. Chicken mononuclear cells (MQ-NCSU) were grown in L15 Leibovitz-McCoy’s 5A modified medium and incubated with FITC-BSA-MP
(250µg/well/mL) for 4 hrs at 41°C in a humidified 5% CO2 incubator. After washing,
the MQ-NCSU cells were stained for MHC Class II surface antigen with an Rphycoerythrin (R-PE)-conjugated mouse anti-chicken MHC Class II Ab (B-L)
(Southern Biotech, Biozol, Eching, Germany), and FITC-BSA-MP uptake and MHC
class II staining was verified by immunofluorescence microscopy.
2.7. Antibody detection
IBDV-specific antibodies were detected by a commercial IBDV-ELISA
(Synbiotics IBD ProFLOK Plus®, San Diego, USA). Neutralizing Abs were also
measured [33]. Briefly, two-fold serially diluted sera were mixed with 100 TCID50 of
an intermediate IBDV strain. Chicken embryo fibroblasts (105/well) were added and
the development of cytopathic effects was evaluated. Geometric mean titers (log2)
per group are presented.
2.8. Histopathology and immunohistochemistry
IBDV induced bursal lesion scores were determined as described before [34].
The mean bursal lesion score for each bird was determined from five microscopic
fields
(400x)
and
group
mean
computed.
Bursal
sections
were
immunohistochemically (IHC) stained for IBDV antigen following a previously
published protocol [28]. The average number of IBDV antigen positive cells in the
bursa of each bird was determined from five microscopic fields (400×). A maximum of
100 brown cells/field was counted.
2.9. Detection of cytokine mRNA
A TaqMan real-time RT-PCR amplification of chIL-4 and IFN-α mRNA from
bursa samples was performed using the AgPath-IDâ„¢ One-Step RT-PCR Reagent kit
(Applied Biosystems®) and the detection and quantification was done by the
59
Mx3005PTM thermal cycler (STRATAGENE; Agilent Technologies Company). The
relative transcript levels of these cytokines were normalized to the chicken 28S rRNA
[35] and values are expressed as 35-CT. The primers, probes and amplification
conditions were described previously [28,35].
2.10. Intrabursal T-cells
T-cell influx in the bursa after IBDV challenge was evaluated by
immunohistochemistry. Cryosections from frozen bursa were stained for CD4+- and
CD8+ T-cells. Mouse anti-chicken CD4 and CD8 mAbs (Southern Biotech, Eching,
Germany) were used as primary Abs, and the anti-mouse IgG biotinylated Ab as a
secondary antibody (Vector Laboratories, Burlingame, USA). T-cell quantification
was conducted as described for IBDV Ag.
2.11. Experimental design
2.11.1. DNA vaccination (experiment 1)
Three weeks old SPF chickens were assigned into 5 groups (n=14-16/group).
Group 1, 2 and 3 were immunized into the thigh muscle (100 μg DNA/bird) with DNAVP2/Eth82, DNA-VP2/7.3 and DNA-vec, respectively (Table 1). Two boost DNA
immunizations were conducted every two weeks. Group 4 received PBS
intramuscularly. Group 5 was vaccinated orally with a commercially available chicken
embryo adapted live IBDV vaccine strain (103EID)50/bird). Blood samples were
collected weekly for Ab detection. Two weeks after the last vaccination each group
was subdivided. One subgroup was challenged with the vvIBDV strain 89163/7.3 (103
ELD50/bird) by eye drop [36] and the other subgroup remain unchallenged. Birds
were monitored for morbidity and mortality. At 3 and 10 days postinfection (dpi), 3-4
chickens/subgroup were sacrificed. Bursa to body weight (B/B) ratio was determined.
Bursae were collected for histology and IBDV Ag detection.
2.11.2. DNA vaccination and adjuvant application (experiment 2)
The vaccine groups (n=9-10/group; 3 weeks old SPF chickens) and
descriptions are presented in Table 2. The route of vaccination, dosage, frequency,
60
challenge and sampling in this experiment was similar to the first experiment.
Plasmid encoded chIL-2 (100 μg /bird) or CpG-ODN (25 µg/bird) were mixed with the
DNA vaccine and delivered in one shot. The dosage for CpG-ODN and chIL-2 were
based on results of previous studies [17,18]. Necropsy was done at 7 dpi.
2.11.3. Vaccination with cationic MP (experiment 3)
The protective efficacy of the MP based IBDV DNA vaccine was tested after
mucosal application to 3 days-old SPF chickens (n =10-12/group) intended to see if
even in immunologically immature birds this regime may lead to a detectable immune
responses. The vaccination protocol is summarized in Table 3. Briefly, chickens of
all groups received 4.5 mg MP (3.5 mg orally and 1 mg by eye drop), except the
CpG-MP group. One mg of VP2-MP or chIL-2-MP adsorbed 8.7 µg of DNAVP2/Eth82 or DNA-chIL-2. When CpG-ODN was administered, each chicken
received 3 mg CpG-MP (2 mg orally and 1 mg by eye drop). One mg of CpG-MP
adsorbed 5.6 µg CpG-ODN. Group 5 (G5) and 6 (G6) were administered respectively
with chIL-2-MP and CpG-MP as an adjuvant two days after inoculation with VP2-MP.
Two booster immunizations were administered every two weeks. Chickens were offfed for 2-3 hrs after MP administration. Two weeks after the last vaccination,
chickens were orally challenged with an intermediate plus IBDV vaccine strain (≥2.0
log10 EID50/bird), which induce significant macroscopical and microscopical bursal
lesions without high mortality rates [8]. Group 1 (G1) served as non-challenge
control. At 3 and 7 dpi, 5-6 chickens were sacrificed and B/B ratio was determined.
Bursae were collected for histology, Ag detection, cytokine mRNA analysis and T-cell
evaluation.
2.12. Statistics
ANOVA was used to analyze the data on lymphocyte proliferation and from
experiment 3. Group means were compared with Tukey HSD pot-hoc test. KruskalWallis one-way nonparametric ANOVA was used in other experiments. Student’s ttest was used to analyze B/B ratios between subgroups. p<0.05 indicates significant
differences.
61
3. Results
3.1. Cloning and expression of VP2 and chIL-2
Transfection studies revealed expression of VP2 and chIL-2 in CEFs (data not
shown). Chicken lymphocytes stimulated with supernatants from CEFs, which had
been transfected with DNA-chIL-2, showed a higher lymphoproliferative index
(2.0±0.5) (p<0.05) than treatment with supernatants from vector-transfected cells
(0.9±0.1).
3.2. Immunogenicity of DNA vaccines (experiment 1&2)
At five weeks post intramuscular immunization, 75% (12/16) of the DNAVP2/Eth82 vaccinated chickens had developed IBDV-ELISA antibody titers of
2070±1565, comparably lower than the live virus vaccinated group (9734±1824).
Only 40% of the DNA-VP2/7.3 (6/15) vaccine group had seroconverted at week 5
after immunization. Four out of eight DNA-VP2/Eth82 vaccinated bird showed
depression and ruffled feathers after challenge. The vector and PBS groups showed
severe clinical signs (Table 1). The DNA-VP2/Eth82 vaccinated group had a
significantly lower mean bursal lesion score (p<0.05) and fewer intrabursal Ag
positive cells compared to the other challenged groups, except the live vaccine group
(Table 1) indicating only partial protection. DNA-chIL-2 or CpG incorporation did not
improve the immunogenicity of the DNA vaccine (Table 2).
3.3. Characterization of cationic MP
The prepared MP had diameters of 3.5-8.5 µm, and the FITC-BSA-MP were
successfully phagocytized by MQ-NCSU cells (data not shown). The adsorption
efficiency was 83-87% for DNA-VP2/Eth82 and DNA-chIL-2, and 56% for CpG-ODN,
indicating the adsorption of 8.3-8.7 µg of plasmid DNA and 5.6 µg of CpG-ODN to 1
mg of cationic MP. Forty five percent of the adsorbed plasmid DNA and 60% of the
CpG-ODN were released into the incubation medium after 24 hrs in vitro incubation
at 37°C. The MP-adsorbed DNA was protected from degradation by DNAse I as the
62
supercoiled plasmid was preserved at the end of DNAse I degradation assay (data
not shown).
3.4. Evaluation of the protective efficacy of the cationic MP-delivery strategy for
IBDV-DNA vaccination (experiment 3)
At 3 dpi, no difference was observed in the B/B ratios between groups
(p>0.05). The chickens in G4 (VP2-MP), G5 (VP2-chIL-2-MP) and G6 (VP2-CpGMP) had a significantly higher B/B ratio at 7 dpi compared to the other challenged
groups (Table 4) (p<0.05). At 7 dpi bursal lesions were lower in G4, 5, and 6 (p<0.05)
compared to the G2 (Mock-MP+), 3 (Vec-MP), 7 (chIL-2-MP) and 8 (CpG-MP)
groups (p<0.05) (Table 4). All VP2-MP groups had a relatively lower intrabursal Ag
load compared to the other groups at 3 and 7 dpi (Table 4) (p<0.05). Chickens in G5
had the lowest intrabursal Ag load and three out of six chickens had cleared the virus
at 7 dpi.
Neutralizing antibodies were neither detected in prechallenge sera nor at 3 dpi
in any group. Chickens vaccinated with VP2-ChIL-2-MP had the lowest
postchallenge VN Ab titers at 7 dpi (Table 4).
At 3 dpi no significant differences were detected in bursal IFN-α and IL-4
mRNA expression between groups (Fig. 1A & C). At 7 dpi the expression levels of
chIFN-α mRNA were significantly lower in the bursae of G4, 5 and 6 (Fig. 1B)
compared to other groups (p<0.05). Significant upregulation of IL-4 mRNA was
observed in G2 and G3 at 7 dpi compared to G1, G4, and G5 (Fig. 1D).
Higher numbers of intrabursal CD4+ T-cells were detected in the Mock-MP+
group compared to the other groups at 3 dpi (Fig. 2A; p<0.05). The number of CD8+
T-cells increased in all challenged groups at 3 dpi and G2, G3, G7 and G8 (Fig. 2C)
had the highest number of CD8+ T-cells. At 7 dpi more CD4+ (Fig. 2B) and CD8+ Tcells (Fig. 2D) were recruited to the bursa of G2, 3, 7 and 8 (p<0.05) compared to the
VP2-MP and VP2-MP plus adjuvant groups.
63
4. Discussion
The aim of this study was to develop an improved MP based IBDV DNA
vaccine and to further enhance the protection efficacy of this vaccine by including MP
loaded with adjuvants.
The VP2-DNA vaccines were shown to be immunogenic and conferred partial
protection against challenge. Generally, DNA vaccinations are known to not induce
significant Ab responses, but may provide partial protection against virulent IBDV
[11,15,17,18]. Co-administration of CpG-ODN or plasmid encoded chIL-2 together
with the IBDV-VP2 or polyprotein gene induced variable levels of humoral and
cellular immunity. Yet, chickens mostly exhibited bursal atrophy after challenge
[17,37,38], which were even more pronounced than in groups without adjuvants [39],
coinciding with our observation.
The characterization of the MP revealed appropriate physical properties for in
vivo application [25, 30]. PEI incorporation into the MP may condense plasmid DNA
on the MP surfaces and likely prevents DNA degradation by mucosal enzymes
[23,30].
The mucosal delivery of the VP2-loaded MP (VP2-MP) in combination with a
delayed mucosal application of chIL-2 or CpG-ODN loaded MP did not induce virus
neutralizing Abs. We speculate that the DNA is slowly released from the MP, VP2 is
expressed, processed and presented by APCs to culminate in a CTL response [40].
After IBDV challenge, bursal pathology and intrabursal Ag load and T-cell
influx were reduced in all VP2-MP vaccinated groups. This provides the proof-ofconcept for the potential of MP to improve DNA vaccine efficacy. The lower antigen
load in VP2-MP plus chIL-2-MP-inoculated chickens may explain the lower
postchallenge neutralizing Ab titers, IL-4 and IFN-α mRNA expression possibly due
to reduced virus replication or faster clearance in these birds compared to other
groups.
CpG-ODN and chIL-2 were shown to exert potent adjuvant effects when
delivered by particulate carriers compared to their delivery in soluble forms [41,42].
Baden et al. [19] suggested that increased IL-2 receptor expression on antigen
primed T-cells may be necessary for an optimal activation and proliferation of these
64
cells by IL-2. However, during simultaneous administration of an Ag and cytokine, the
cytokine may simulate rather a broad cellular response with relatively fewer antigenspecific cells [19]. Tregs in mice and humans showed an increased level of toll like
receptor (TLR) expression [43] and may expand to an extent dominating antigen
specific CD4+ T-cells when CpG was administered simultaneously with a vaccine.
Thus, the timing of TLR agonist (CpG-ODN) application may influence the outcome
of vaccine-induced immune responses [44]. We hypothesize that in our vaccination
model the delayed application of chIL-2 after an MP-DNA delivery may enhance the
activation and proliferation of Ag specific T-cells in chickens before extensive
proliferation of immunosuppressive Tregs may take place.
Overall the results of this study indicate that DNA-vaccination against IBDV
using MP as a delivery system in combination with molecular adjuvants may be an
interesting alternative IBDV-vaccination approach, which may help to reduce the
risks associated with the use of live vaccine strains. In future studies, we not only
need to investigate additional routes of administration but also we need to elucidate
the mechanism of protection and to what extent the different types of T-cells are
involved in the control of IBDV challenge after MP-DNA-vaccination. This would allow
further improvement of this vaccination approach and possibly paves an option for
mass application in the field.
65
Acknowledgments
This project was supported partly by the Dorothea Erxleben Program (SR),
German research Foundation (RA767/2.1) and by the German Academic Exchange
Service (TN). The authors acknowledge Christine Haase for her technical support;
Sonja Bernhard, Sabrina Techel and Katja Stolpe for their excellent help in animal
care and necropsy.
66
References
[1]
Brown F. The Classification and Nomenclature of Viruses - Summary of Results of
Meetings of the International Committee on Taxonomy of Viruses in Edmonton, Canada
1987. Intervirology 1989;30:181-6.
[2]
Müller H, Scholtissek C, Becht H. The genome of infectious bursal disease virus
consists of two segments of double-stranded RNA. J Virol 1979;31:584-9.
[3]
Eterradossi N, Toquin D, Rivallan G, Guittet M. Modified activity of a VP2-located
neutralizing epitope on various vaccine, pathogenic and hypervirulent strains of infectious
bursal disease virus. Arch Virol 1997;142:255-70.
[4]
Eterradossi N, Saif YM. Infectious Bursal Disease. In: Saif YM, Fadly AM, Glisson JR,
McDougald LR, Nolan LK, Swayne DE, editors. Diseases of Poultry 12 ed. Iowa: Iowa State
University Press, 2008: 185-2008.
[5]
Cheville NF. Studies on the pathogenesis of Gumboro disease in the bursa of
Fabricius, spleen, and thymus of the chicken. Am J Pathol 1967;51:527-51.
[6]
Nakamura K, Yuasa N, Abe H, Narita M. Effect of infectious bursal disease virus on
infections produced by Escherichia coli of high and low virulence in chickens. Avian Pathol
1990;19:713-21.
[7]
Müller H, Mundt E, Eterradossi N, Islam MR. Current status of vaccines against
infectious bursal disease. Avian Pathol 2012;41:133-39.
[8]
Rautenschlein S, Kraemer C, Vanmarcke J, Montiel E. Protective efficacy of
intermediate and intermediate plus infectious bursal disease virus (IBDV) vaccines against
very virulent IBDV in commercial broilers. Avian Dis 2005;49:231-37.
[9]
Yamaguchi T, Setiyono A, Kobayashi M, Takigami S, Fukushi H, Hirai K. Infectious
bursal disease live vaccines: changes in the virus population during serial passage in
chickens and chicken embryo fibroblast cells. Avian Dis 2000;44:284-90.
[10]
Kong LL, Omar AR, Hair-Bejo M, Aini I, Seow HF. Sequence analysis of both genome
segments of two very virulent Infectious bursal disease virus field isolates with distinct
pathogenicity. Arch Virol 2004;149:425-34.
[11]
Rong J, Cheng T, Liu X, Jiang T, Gu H, Zou G. Development of recombinant VP2
vaccine for the prevention of infectious bursal disease of chickens. Vaccine 2005
23;23:4844-51.
[12]
Pitcovski J, Gutter B, Gallili G, Goldway M, Perelman B, Gross G, et al. Development
and large-scale use of recombinant VP2 vaccine for the prevention of infectious bursal
disease of chickens. Vaccine 2003;21:4736-43.
67
[13]
Le Gros FX, Dancer A, Giacomini C, Pizzoni L, Bublot M, Graziani M, et al. Field
efficacy trial of a novel HVT-IBD vector vaccine for 1-day-old broilers. Vaccine 2009;27:5926.
[14]
Chen YY, Hsieh MK, Tung CY, Wu CC, Lin TL. Infectious bursal disease DNA
vaccination conferring protection by delayed appearance and rapid clearance of invading
viruses. Arch Virol 2011;156:2241-50.
[15]
Li J, Huang Y, Liang X, Lu M, Li L, Yu L, et al. Plasmid DNA encoding antigens of
infectious bursal disease viruses induce protective immune responses in chickens: factors
influencing efficacy. Virus Res 2003;98:63-74.
[16]
Hsieh MK, Wu CC, Lin TL. DNA-mediated vaccination conferring protection against
infectious bursal disease in broiler chickens in the presence of maternal antibody. Vaccine
2010;28:3936-43.
[17]
Mahmood MS, Siddique M, Hussain I, Khan A, Mansoor MK. Protection capability of
recombinant plasmid DNA vaccine containing VP2 gene of very virulent infectious bursal
disease virus in chickens adjuvanted with CpG oligodeoxynucleotide. Vaccine 2006;24:483846.
[18]
Hulse DJ, Romero CH. Partial protection against infectious bursal disease virus
through DNA-mediated vaccination with the VP2 capsid protein and chicken IL-2 genes.
Vaccine 2004;22:1249-59.
[19]
Baden LR, Blattner WA, Morgan C, Huang Y, Defawe OD, Sobieszczyk ME, et al.
Timing of plasmid cytokine (IL-2/Ig) administration affects HIV-1 vaccine immunogenicity in
HIV-seronegative subjects. J Infect Dis 2011;204:1541-9.
[20]
Chen H, Gao N, Fan D, Wu J, Zhu J, Li J, et al. Suppressive effects on the immune
response and protective immunity to a JEV DNA vaccine by co-administration of a GM-CSFexpressing plasmid in mice. PLoS One 2012;7:e34602.
[21]
Reddy KS, Rashmi BR, Dechamma HJ, Gopalakrishna S, Banumathi N,
Suryanarayana VV, et al. Cationic microparticle [poly(D,L-lactide-co-glycolide)]-coated DNA
vaccination induces a long-term immune response against foot and mouth disease in guinea
pigs. J Gene Med 2012;14:348-52.
[22]
Wang G, Pan L, Zhang Y, Wang Y, Zhang Z, Lu J, et al. Intranasal delivery of cationic
PLGA nano/microparticles-loaded FMDV DNA vaccine encoding IL-6 elicited protective
immunity against FMDV challenge. PLoS One 2011;6:e27605.
68
[23]
Wegmann F, Gartlan KH, Harandi AM, Brinckmann SA, Coccia M, Hillson WR, et al.
Polyethyleneimine is a potent mucosal adjuvant for viral glycoprotein antigens. Nat
Biotechnol 2012;30:883-8.
[24]
O'Hagan DT, Singh M, Ulmer JB. Microparticles for the delivery of DNA vaccines.
Immunol Rev 2004;199:191-200.
[25]
Liman M, Peiser L, Zimmer G, Prosting M, Naim HY, Rautenschlein S. A genetically
engineered prime-boost vaccination strategy for oculonasal delivery with poly(D,L-lactic-coglycolic acid) microparticles against infection of turkeys with avian Metapneumovirus.
Vaccine 2007;25:7914-26.
[26]
Negash T, Gelaye E, Perersen H, Grummer B, Rautenschlein S. Molecular Evidence
of Very Virulent Infectious Bursal Disease Viruses in Chickens in Ethiopia. Avian Dis
2012;56:605-10.
[27]
Zhou JY, Wang JY, Chen JG, Wu JX, Gong H, Teng QY, et al. Cloning, in vitro
expression and bioactivity of duck interleukin-2. Mol Immunol 2005;42:589-98.
[28]
Rautenschlein S, von Samson-Himmelstjerna G, Haase C. A comparison of immune
responses to infection with virulent infectious bursal disease virus (IBDV) between specificpathogen-free chickens infected at 12 and 28 days of age. Vet Immunol Immunopathol
2007;115:251-60.
[29]
Gerlier D, Thomasset N. Use of MTT colorimetric assay to measure cell activation. J
Immunol Methods 1986;94:57-63.
[30]
Oster CG, Kim N, Grode L, Barbu-Tudoran L, Schaper AK, Kaufmann SHE, et al.
Cationic microparticles consisting of poly(lactide-co-glycolide) and polyethylenimine as
carriers systems for parental DNA vaccination. J Control Release 2005;104:359-77.
[31]
St Paul M, Paolucci S, Read LR, Sharif S. Characterization of responses elicited by
Toll-like receptor agonists in cells of the bursa of Fabricius in chickens. Vet Immunol
Immunopathol 2012;149:237-44.
[32]
St Paul M, Mallick AI, Haq K, Orouji S, Abdul-Careem MF, Sharif S. In vivo
administration of ligands for chicken toll-like receptors 4 and 21 induces the expression of
immune system genes in the spleen. Vet Immunol Immunopathol 2011;144:228-37.
[33]
Winterfield RW, Fadly AM, Bickford A. Infectivity and distribution of infectious bursal
disease virus in the chicken. Persistence of the virus and lesions. Avian Dis 1972;16:622-32.
[34]
Sharma JM, Dohms JE, Metz AL. Comparative pathogenesis of serotype 1 and
variant serotype 1 isolates of infectious bursal disease virus and their effect on humoral and
cellular immune competence of specific-pathogen-free chickens. Avian Dis 1989;33:112-24.
69
[35]
Powell FL, Rothwell L, Clarkson MJ, Kaiser P. The turkey, compared to the chicken,
fails to mount an effective early immune response to Histomonas meleagridis in the gut.
Parasite Immunol 2009;31:312-27.
[36]
Aricibasi M, Jung A, Heller ED, Rautenschlein S. Differences in genetic background
influence the induction of innate and acquired immune responses in chickens depending on
the virulence of the infecting infectious bursal disease virus (IBDV) strain. Vet Immunol
Immunopathol 2010;135:79-92.
[37]
Park JH, Sung HW, Yoon BI, Kwon HM. Protection of chicken against very virulent
IBDV provided by in ovo priming with DNA vaccine and boosting with killed vaccine and the
adjuvant effects of plasmid-encoded chicken interleukin-2 and interferon-gamma. J Vet Sci
2009;10:131-9.
[38]
Li J, Liang X, Huang Y, Meng S, Xie R, Deng R, et al. Enhancement of the
immunogenicity of DNA vaccine against infectious bursal disease virus by co-delivery with
plasmid encoding chicken interleukin 2. Virology 2004;329:89-100.
[39]
Roh HJ, Sung HW, Kwon HM. Effects of DDA, CpG-ODN, and plasmid-encoded
chicken IFN-gamma on protective immunity by a DNA vaccine against IBDV in chickens. J
Vet Sci 2006;7:361-8.
[40]
Helson R, Olszewska W, Singh M, Megede JZ, Melero JA, O'Hagan D, et al.
Polylactide-co-glycolide (PLG) microparticles modify the immune response to DNA
vaccination. Vaccine 2008;26:753-61.
[41]
Malyala P, Chesko J, Ugozzoli M, Goodsell A, Zhou F, Vajdy M, et al. The potency of
the adjuvant, CpG oligos, is enhanced by encapsulation in PLG microparticles. J Pharm Sci
2008;97:1155-64.
[42]
Zhang WL, Yin Z, Liu NH, Yang T, Wang JW, Bu Z, et al. DNA-chitosan nanoparticles
improve DNA vaccine-elicited immunity against Newcastle disease virus through shuttling
chicken interleukin-2 gene. J Microencapsul 2010;27:693-702.
[43]
Kulkarni R, Behboudi S, Sharif S. Insights into the role of Toll-like receptors in
modulation of T cell responses. Cell Tissue Res 2011;343:141-52.
[44]
Jensen KM, Melchjorsen J, Dagnaes-Hansen F, Sorensen UBS, Laursen RR,
Ostergaard L, et al. Timing of Toll-Like Receptor 9 Agonist Administration in Pneumococcal
Vaccination Impacts Both Humoral and Cellular Immune Responses as Well as
Nasopharyngeal Colonization in Mice. Infect Immun 2012;80:1744-52.
70
Table 1
Experiment 1: Evaluation of protection conferred by IBDV DNA vaccinations in comparison to a live IBDV vaccineg
Mean±SD
Groups
1
2
3
4
5
Vaccines
DNA-VP2/Eth82
DNA-VP2/7.3
DNA-vec
PBS
Live virus vaccine
Challenge
Morbidity
Mortality
B/B ratio
Bursal lesion score
Bursal antigen load
3 dpi
10 dpi
3 dpi
10 dpi
3 dpi
10 dpi
-
0/8
0/8
3.8±0.9
3.7±0.9b
0.0±0.0
0.0±0.0
0.0±0.0
0.0±0.0
+
4/8
1/8
4.9±1.0
2.3±1.0a
0.3±0.1a
2.5±0.1a
65.0±9.0a
51.0±17.3a
-
0/7
0/7
4.2±0.7
3.7±0.9b
0.0±0.0
0.0±0.0
0.0±0.0
0.0±0.0
a
1.6±0.2
0.0±0.0
+
5/8
1/8
6.2±1.7
2.1±1.6
-
0/7
0/7
3.9±1.5
4.4±0.3b
a
4.0±0.0
b
3.5±0.1
b
0.0±0.0
c
71.0±13
a
0.0±0.0
c
65.0±17.3a
0.0±0.0
b
100±0.0b
+
7/7
1/7
4.6±1.0
1.7±1.6
-
0/7
0/7
4.6±0.3
4.4±0.3b
0.0±0.0
0.0±0.0
0.0±0.0
0.0±0.0
+
7/7
2/7
4.9±1.0
1.7±1.7a
4.0±0.0c
4.0±0.0c
100±0.0b
100±0.0b
-
0/8
0/8
3.3±0.6
3.4±1.5b
0.0±0.0
0.0±0.0
0.0±0.0
0.0±0.0
+
0/8
0/8
4.0±1.4
3.7±0.4b
0.0±0.0
0.0±0.0
0.0±0.0
0.0±0.0
4.0±0.0
100±0.0
71
g
Three weeks old SPF chickens (n=14-16/group) were immunized by the IM route with DNA vaccines (DNA-VP2/Eth82, DNA-
VP2/7.3) followed by two DNA booster immunizations (group 1-3). Group 5 was vaccinated orally with a commercially available
chicken embryo adapted live IBDV vaccine strain (103EID)50/bird). Two weeks after the last vaccination groups were divided: one
subgroup was challenged (+) with vvIBDV strain 89163/7.3. The other subgroup remained unchallenged (-).The vvIBDV strains had
been isolated from several chicken flocks that had experienced high mortality rate [26]. Morbidity, mortality, B/B ratio and microscopic
bursal lesions were determined. IBDV Ag load in the bursa was quantified by counting antigen positive cells in five microscopic
fields/section. Histopathology of the bursae in affected groups included lymphoid depletion in the cortex and medulla. Different
superscripts in a column show significant difference between challenged groups (Kruskal-Wallis test, p<0.05). Comparison of B/B
ratios of challenged and non-challenged subgroups was done by Student’s t-test. p<0.05.
72
Table 2
Experiment 2: Evaluation of protection induced by IBDV DNA vaccination in combination with molecular adjuvantsh
Groups
Vaccine group
Challenge Morbidity
Mortality
Mean±SD
B/B ratio
Bursal lesion
score
1
DNA-vec+DNA-chIL-2
+
9/9
4/9
2.1±0.7
4.0±0.0b
2
DNA-vec+CpG
+
9/9
4/9
2.4±0.4
4.0±0.0b
3
DNA-VP2/7.3
+
8/10
2/10
2.9±1.8
3.1±0.2a
4
DNA-VP2/7.3+DNA-chIL-2
+
10/10
5/10
3.3±2.2
4.0±0.0b
5
DNA-VP2/7.3+CpG
+
10/10
4/10
2.7±0.2
3.5±0.5b
6
DNA-chIL-2
+
10/10
2/10
2.8±0.9
4.0±0.0b
7
CpG
+
10/10
3/10
3.2±0.6
4.0±0.0b
8
DNA-vec
+
9/9
4/9
2.8±0.9
4.0±0.0b
9
PBS
-
0/10
0/10
3.5±0.7
0.0±0.0
10
PBS
+
9/9
4/9
2.7±0.9
4.0±0.0b
73
h
Three weeks old SPF chickens were randomly assigned into 10 groups (n=9-10/group) and each group received different
treatments. To improve the immune responses to the IBDV DNA vaccine, DNA-chIL-2 and CpG-ODN as adjuvants were included.
After two booster vaccinations, the chickens were challenged with vvIBDV strain 89163/7.3. Morbidity and mortality were recorded.
B/B ratio and microscopic bursal lesions were determined at 7 dpi. Different superscripts in a column show significant difference
between challenged groups. Kruskal-Wallis test, p<0.05.
74
Table 3
Experiment 3: Study design showing experimental groups and the timing of vaccine and adjuvant applicationi
Groups
Challenge
Cationic MP administration at days postvaccination (dpv)
42 dpv
0
2
14
16
28
30
G1 (Mock-MP)
-
Mock-MP
Mock-MP
Mock-MP
G2 (Mock-MP)
+
Mock-MP
Mock-MP
Mock-MP
G3 (Vec-MP)
+
Vec-MP
Vec-MP
Vec-MP
G4 (VP2-MP)
+
VP2-MP
VP2-MP
VP2-MP
G5 (VP2-chIL-2-MP)
+
VP2-MP
chIL-2-MP VP2-MP
chIL-2-MP VP2-MP
chIL-2-MP
G6 (VP2-CpG-MP)
+
VP2-MP
CpG-MP
CpG-MP
CpG-MP
G7 (chIL-2-MP)
+
chIL-2-MP
chIL-2-MP
chIL-2-MP
G8 (CpG-MP)
+
CpG-MP
CpG-MP
CpG-MP
i
VP2-MP
VP2-MP
In this experiment DNA vaccine and adjuvants were delivered through mucosal routes to 3-5 days old SPF chickens. The DNA
vaccine and adjuvants were adsorbed separately onto cationic PLGA-MP. G5 and G6 were administered with chIL-2-MP and CpGMP, respectively two days after the chickens were inoculated with the DNA vaccine adsorbed onto the MP, both at the primary and
booster vaccinations. G means group.
75
Table 4
Experiment 3: Determination of the protective efficacy of MP based DNA vaccine and adjuvants after mucosal application to SPF
chickensj
Mean±SD
Groups
Challenge VN titer (log2)/group
B/B ratio
7 dpi
3 dpi
7 dpi
Bursal lesion score Bursal antigen load
3 dpi
7 dpi
3 dpi
7 dpi
0.0±0.0
0.0±0.0
0.0±0.0
0.0±0.0
G1 (Mock-MP)
-
0.0±0.0
3.7 ±0.9 3.59 ±0.2b
G2 (Mock-MP)
+
5.0±0.0a
4.9±0.9
1.65 ±0.2a
2.6±0.2a 4.0±0.0b 100 ±0.0d 100±0.0a
G3 (Vec-MP)
+
5.0±0.0a
4.0±0.6
1.65 ±0.3a
0.9±0.4b 4.0±0.0b 100±0.0d
100±0.0a
G4 (VP2-MP)
+
3.3±0.0b
3.5±1.2
2.15 ±0.5c
0.9±0.1b 2.1±0.3a 45±13c
22±8.5b
G5 (VP2-chIL-2-MP) +
1.7±0.0c
4.2±0.6
2.75 ±0.3c
0.7±0.2b 1.9±0.1a 28±18c
9±2.1c
G6 (VP2-CpG-MP)
+
2.5±0.0d
3.8±0.5
2.85 ±0.6c
1.7±0.4c 1.8±0.2a 42±16c
25.0±8.9b
G7 (chIL-2-MP)
+
5.0±0.0a
4.1±0.9
1.47 ±0.3a
2.3±0.5a 4.0±0.0b 98±2.0d
100±0.0a
G8 (CpG-MP)
+
5.0±0.0a
4.3±1.3
1.96 ±0.4a
1.9±0.3c 4.0±0.0b 100±0.0d
100±0.0a
76
j
Three to five days old SPF chickens (n=10-12/group) received 4.5 mg MP of DNA vaccine or chIL-2. When CpG group was
administered, each chicken received 3 mg MP. Each milligram MP adsorbed either 8.7 µg of DNA (VP2 or chIL-2) or 5.6 µg CpG.
G5 and G6 were administered with chIL-2-MP and CpG-MP two days after the chickens were inoculated with the DNA vaccine
adsorbed to the MP. Two weeks after the last vaccination, chickens were orally challenged with an intermediate plus IBDV vaccine
strain (≥2.0 log10 EID50/bird), which induce significant macroscopical and microscopical bursal lesions without high mortality rates
(RAUTENSCHLEIN et al. 2005a). Different superscripts in a column show significant differences between challenged groups, while
identical superscripts show no statistically significant differences between challenged and nonchallenged groups. ANOVA; p<0.05.
IFN-alpha corrected mRNA
expression (35-CT)
77
25.00
20.00
15.00
10.00
5.00
0.00
G1
(MockMP-)
G2
(MockMP+)
G3 (Vec- G4 (VP2- G5 (VP2- G6 (VP2MP)
MP)
chIL2- CpG-MP)
MP)
G7
(chIL2MP)
G8 (CpGMP)
G3 (Vec- G4 (VP2- G5 (VP2- G6 (VP2G7
MP)
MP)
chIL2- CpG-MP) (chIL2MP)
MP)
G8 (CpGMP)
Experimental groups
A
IFN-alpha corrected mRNA
expression (35-CT)
25
20
*
15
*
10
5
0
G1
(MockMP-)
B
*
G2
(MockMP+)
Experimental groups
78
IL-4 corrected mRNA
expression (35-CT)
5
4
3
2
1
#
#
0
G1
(MockMP-)
G2
(MockMP+)
G3 (Vec- G4 (VP2- G5 (VP2- G6 (VP2G7
G8 (CpGMP)
MP)
chIL2- CpG-MP) (chIL2MP)
MP)
MP)
Experimental groups
IL-4 corrected mRNA
expression (35-CT)
C
5
*
*
4
3
2
1
0
G1
(MockMP-)
G2
(MockMP+)
G3 (Vec- G4 (VP2- G5 (VP2- G6 (VP2G7
G8 (CpGMP)
MP)
chIL2- CpG-MP) (chIL2MP)
MP)
MP)
Experimental groups
D
Fig. 1. Quantification of postchallenge chIFN-α mRNA expression levels in the bursa at 3
(A), and 7 dpi (B); and chIL-4 mRNA expression at 3 (C) and 7 dpi (D). Chickens were
challenged with an intermediate plus IBDV vaccine strain, which induce significant
macroscopical and microscopical bursal lesions without high mortality rates [8] two
weeks after the last booster vaccination. *indicates statistically significant differences
between groups. # Ct value of 35 was used to calculate the relative gene expression in
all samples. ANOVA, p<0.05. Group descriptions are presented in Table 3.
A
Average no. of CD4+ T cells/field
79
120
100
80
60
40
20
0
*
G1
(MockMP-)
G2
G3 (VecG4
(MockMP)
(VP2MP+)
MP)
G5
(VP2chIL2MP)
G6
(VP2CpGMP)
G7
(chIL2MP)
G8
(CpGMP)
Experimental groups
A
Average no. of CD4+ T cells/field
B
120
100
80
60
40
20
0
*
*
*
*
G1
G2
(Mock- (MockMP-) MP+)
G3
(VecMP)
G4
(VP2MP)
G5
(VP2chIL2MP)
G6
(VP2CpGMP)
G7
G8
(chIL2- (CpGMP)
MP)
Experimental groups
Average no. of CD8+ T cells/field
80
120
100
80
60
40
*
20
0
G7
G3 (Vec- G4 (VP2- G5 (VP2- G6 (VP2G2
(chIL2CpGMP)
MP)
chIL2(MockMP)
MP)
MP)
MP+)
G1
(MockMP-)
Experimental groups
Average no. of CD8+ T cells/field
C
D
G8
(CpGMP)
120
*
*
*
100
80
60
40
20
D
0
G2
G1
(Mock- (MockMP+)
MP-)
G3
(VecMP)
G4
(VP2MP)
G5
(VP2chIL2MP)
G6
(VP2CpGMP)
G7
(chIL2MP)
G8
(CpGMP)
Experimental groups
Fig. 2. Quantification of postchallenge intrabursal CD4+ T-cells at 3 (A), and 7 dpi
(B); and CD8+ T-cells at 3 (C) and 7 dpi (D). Chickens were challenged with an
intermediate plus IBDV vaccine strain, which induce significant macroscopical and
microscopical bursal lesions without high mortality rates [8] two weeks after the
last booster vaccination.*indicates statistically significant differences between
groups (ANOVA, p<0.05). Group descriptions are presented in Table 3.
81
6. Discussion and conclusions
Poultry pathogens change their nature in response to intensified poultry production
and present complex challenges to the poultry health and productivity. Classical
IBDV strains that emerged in the 1960s still cause mortality rates of 1% to 30% in
susceptible chickens (MÜLLER et al. 2003). Very virulent IBDVs, which emerged in
the late 1980s can cause mortality of up to 70% and induce severe
immunosuppression (CHETTLE et al. 1989). Vaccination pressure may likely
contribute to the emergence of highly pathogenic virus strains. Thus, understanding
the molecular epidemiology of IBDVs in the field has been important for the design of
preventive measures and for the development of a new generation of vaccines to
control the spread of the virus.
6.1. Molecular epidemiology of IBDV field isolates
Molecular approaches allow the identification and differentiation of IBDV strains
circulating in chicken populations and associate recent and past isolates (LE NOUEN
et al. 2005). The molecular epidemiology of IBDV has been studied in many
geographical areas and IBDV evolution well documented. Particularly, serotype 1
IBDV strains have been circulating in many poultry operations in North and South
America, Europe, Asia and in African countries (YAMAGUCHI et al. 1997a;
ZIERENBERG et al. 2000; VAN DEN BERG et al. 2004; JACKWOOD u. SOMMER
2005; JACKWOOD u. SOMMER-WAGNER 2007; JUNEJA et al. 2008; HE et al.
2012b; KASANGA et al. 2012).
In this study, several IBDV isolates were collected during field outbreaks in different
types of poultry flocks in Ethiopia. Experimental inoculation of chickens with most of
these isolates allowed us to confirm the pathogenicity of the viruses. Inoculated
chickens showed significant bursal lesions and an extensive IBDV Ag load in the
bursa. The characterization of the nucleotide (nt) sequences of the hVP2 and the 5’
two thirds of VP1 of 11 isolates demonstrated that the Ethiopian isolates showed
82
close homology to vvIBDV isolates from other countries. Based on a recent study, it
is expected that, worldwide, about 60 to 76% of IBDV isolates are of vvIBDV
genotype (JACKWOOD u. SOMMER-WAGNER 2007; HE et al. 2012b). The rest of
the isolates are classical and variant strains based on their hVP2 characteristics
(JACKWOOD u. SOMMER-WAGNER 2007; HE et al. 2012b). Although phenotypic
characterization of IBDV in experimentally infected SPF chickens is the gold standard
to determine IBDV virulence, genotypic features can support IBDV grouping (VAN
DEN BERG et al. 2004). vvIBDV has shown a rapid global spread; even in the USA,
where the classical and variant strains dominated field outbreaks for nearly five
decades, vvIBDV emerged in 2008 (STOUTE et al. 2009).
Almost all poultry producing regions report the co-existence of two or more strains of
varied pathogenicity (JACKWOOD u. SOMMER-WAGNER 2007; HE et al. 2012b).
Interestingly, only vvIBDV was identified in Ethiopia during the period of our study. A
recent countrywide study reported IBDV seropositivity rates in backyard chickens to
be close to 92% (CHAKA et al. 2012; JENBREIE et al. 2012). Detailed molecular
studies of IBDV in this segment of the chicken population may provide further
information about IBDV epidemiology.
Phylogenetic approaches have been particularly useful for studying the origin and
subsequent evolution of a virus. Based on the VP2 nt sequence, phylogenetic
analysis revealed the formation of a cluster by Ethiopian IBDV strains; yet this cluster
reclines in one major monophyletic lineage with other vvIBDVs. It was expected that
vvIBDV strains of the same geographic origins may cluster together, yet some minor
non-significant clustering outside the main cluster was reported among isolates of the
same origins (CORTEY et al. 2012; SILVA et al. 2012). This may indicate a strong
association between geographical origin and viral diversity. In contrast to this
observation, vvIBDVs isolated from the major chicken producing regions of China
showed very divergent phylogenetic clusters based on their VP2 sequences (HE et
al. 2012b). This may highlight the circulation of vvIBDVs with diverse genetic
mutations. Several recent IBDV field strains from USA did not form clusters with most
of the IBDV VP2 sequences retrieved from databases, implying the virus is under
83
continuous evolution (DURAIRAJ et al. 2011). Despite vvIBDV isolates maintaining
the identical virulence marker aa in their VP2 regions, they show differences in their
pathogenicity to SPF chickens even under standardized experimental conditions.
This suggests the importance of the role of VP1 in IBDV virulence (SILVA et al.
2012), which should be besides VP2 regularly included in the molecular analysis of
IBDV strains. Normally, the VP1 gene of IBDV strains shows multiple phylogenetic
lineages, additionally providing evidence of its contribution in IBDV pathogenicity
(JACKWOOD et al. 2012).
How the vvIBDV strains evolved in Ethiopia remains unclear. Literature suggests that
international trade of live poultry and poultry products may facilitate the global spread
of IBDV (COBB 2011). IBDV may spread through contaminated equipment
(FLENSBURG et al. 2002; JACKWOOD u. SOMMER-WAGNER 2010). The high
tenacity of the virus and its resistance to several disinfections and virucidal
procedures may contribute to the rapid distribution of the virus (VAN DEN BERG et
al. 2000b; GARRIGA et al. 2006). The risk of transmission via infected free living wild
birds (JEON et al. 2008) can not be excluded as IBDV or IBDV-specific Abs were
detected in other avian species and in backyard chickens (KASANGA et al. 2008).
Almost all acute disease outbreaks in backyard chickens in developing countries
remain undiagnosed. vvIBDV isolates from wild birds and backyard chickens were
shown to be highly pathogenic for SPF chickens under experimental conditions and
maintain virulence marker aa residues across their VP2 and VP1 genes
(HERNANDEZ-DIVERS et al. 2008).
Ethiopian IBDV isolates appear clonal and are very virulent. In the case of RNA
viruses, biological events including genetic reassortment or recombination alter the
phenotypes and genotypes of circulating viruses and compromise their genetic
stability. Also the first vvIBDV that had caused severe disease outbreak in Europe
was linked to the emergence of segment reassortant IBDV between a newly
appeared IBDV, which combined its very virulent VP1 segment with an endemic
IBDV bearing a very virulent VP2 segment (HON et al. 2006). Natural homologous
intragenic recombination apart from reassortment may lead to new variants of IBDV
84
(HON et al. 2008; HE et al. 2009a). Likewise, Ethiopian strains in our study showed
aa residues within the region of VP1 that are characteristic of attenuated as well as
vvIBDV, so we can not rule out the possibility of recombination within this gene
segment.
While prototype vvIBDVs present glycine (G) at position 254 (loop PDE) of the minor
hydrophilic peak 1 domain of the hypervariable region of VP2, we found a serine (S)
residue in all Ethiopian vvIBDV isolates at this position. A serine (S) residue at this
position was detected in most recent vvIBDV field isolates from Africa, Europe and
Asia from chickens vaccinated with classical live IBDV vaccines (KASANGA et al.
2007; MARTIN et al. 2007). Recent study revealed that most other IBDV isolates
show frequent aa exchange at this position as well (ICARD et al. 2008; DURAIRAJ et
al. 2011). This indicates a significant role of this domain in determining IBDV
antigenicity as demonstrated by previous studies (BAYLISS et al. 1990; HEINE et al.
1991; LANA et al. 1992; DORMITORIO et al. 1997). Also the crystal structure of the
VP2 protein suggests selection pressure associated with the hydrophilic domains due
to their location at the outermost exposed part of the VP2 protein (COULIBALY et al.
2005). A neutralizing Ab escape mutant virus at position 254 (S) was generated from
Del-E IBDV by site-directed mutagenesis (JACKWOOD u. SOMMER-WAGNER
2011). Chickens vaccinated with a live Del-E vaccine and challenged with the mutant
developed severe bursal lesions, whereas those vaccinated with Del-E and
challenged with a homologous virus were fully protected (JACKWOOD u. SOMMERWAGNER 2011) showing the contribution of this particular aa residue in vaccination
failure.
Vaccination failure associated with the application of plaque purified and cloned live
IBDV vaccines during and in subsequent IBD outbreaks in Ethiopia (personal
communication with poultry producers and veterinarians) may be due to the lack of
cross protection due to aa exchanges at the hydrophilic domains of field isolates
compared to the vaccinal strains. Vaccines on the basis of IBDV quasispecies
(JACKWOOD u. SOMMER 2002) may in contrast to a cloned live vaccine provide a
85
better cross protection in the face of heterologous IBDV field challenges
counteracting at least in part the development of Ab escape mutants.
6.2. Immune responses induced by candidate IBDV DNA vaccines and correlation to
protection
Considering the significant economic losses associated with IBDV, the development
and evaluation of new generation IBDV vaccines are important. With this goal,
candidate DNA vaccines were produced from the newly characterized vvIBDV strains
circulated in the chicken population in Ethiopia. In the current study, IM immunization
of SPF chickens with DNA encoding IBDV-VP2 from two different vvIBDV strains
induced humoral responses, although the Ab titers were lower compared to a
commercial live IBDV vaccine. Multiple boosts could not significantly increase the Ab
levels as confirmed by previous studies (LI et al. 2003). Similar investigations,
however, reported that IM administration of large doses (10 mg/bird) of DNA
encoding the IBDV polyprotein (PP) gene induced significantly higher Ab responses
compared to the VP2 DNA vaccine (HSIEH et al. 2010). The oral delivery of live
attenuated S. Typhimurium or E. coli that expressed the PP or VP2 induced higher
Ab levels comparable to a conventional live IBDV vaccine and provided better
protection (LI et al. 2006; MAHMOOD et al. 2007; HSIEH et al. 2010). Normally, DNA
encoding the PP of vvIBDV induced higher neutralizing Ab responses and protection
compared to PP originated from an attenuated strain (KIM et al. 2004). The highest
Ab titers to IBDV DNA vaccines were detected when they were used for priming
followed by boosters with killed or subunit vaccines (HSIEH et al. 2007; PARK et al.
2009; GAO et al. 2013).
In our study, DNA-VP2 vaccinated chickens were protected to some extent against
IBDV challenge. Earlier studies showed that DNA-PP vaccines were more effective
to induce protection against challenge than the DNA-VP2 vaccines possibly due to
higher IBDV-specific neutralizing Abs (LI et al. 2003; KIM et al. 2004; LI et al. 2004).
The expression of PP in vivo by transfected muscle cells following IM application may
86
lead to the formation of virus-like particles (MARTINEZ-TORRECUADRADA et al.
2003). This may be possible because of correct processing of the PP by viral and
host proteases that maintain the natural conformation of the immunodominant
epitopes of VP2, thus their immunogenicity and protective efficacy.
In most previous studies no complete protection against challenge was achieved by
IBDV-DNA vaccination. Bursal atrophy, viral RNA or Ags were detected after
challenge in the bursal follicles of recovered birds similar to our study (ZIERENBERG
et al. 2000; CHANG et al. 2001; CHANG et al. 2003; KIM et al. 2004; LI et al. 2006;
MAHMOOD et al. 2007; HSIEH et al. 2010). IBDV DNA vaccine induced complete
protection possibly depends on the levels of both humoral and cellular immunity
(CHEN et al. 2011c).
6.3. Do molecular adjuvants enhance protectivity of an IBDV DNA vaccine?
Since ‘first-generations’ IBDV DNA vaccines are poorly immunogenic, and rarely
generated detectable Ab levels, we next evaluated the inclusion of plasmids
encoding chIL-2 or synthetic unmethylated CpG-ODN as adjuvants. Previous studies
indicated the in vitro and in vivo immunoregulatory properties of chIL-2 and CpGODN on various cell populations of the immune system of avian species (HILTON et
al. 2002; HE u. KOGUT 2003; ASIF et al. 2004; SHANMUGASUNDARAM u.
SELVARAJ 2011). We wanted to extend these observations by utilizing these
molecular adjuvants as IBDV DNA vaccine adjuvants in various formulations.
With our vaccination protocol, which was similar to previously described ones, chIL-2
did not enhance the potency and protection conferred by the IBDV DNA vaccine. A
lack of an adjuvant effect of a plasmid-encoded chIL-2 was previously demonstrated
when it was applied in ovo with the IBDV DNA vaccine (PARK et al. 2009). Others
had shown that the inclusion of IL-2 encoding DNA enhanced humoral and cellular
immune responses induced by the DNA-VP2 vaccine; although protection varied
greatly (HULSE u. ROMERO 2004; LI et al. 2004). The co-administration of DNA
87
encoding chIL-2 with the DNA-PP intramuscularly to SPF chickens induced an early
and higher Ab response and protection against lethal IBDV compared to the chIL-2
plus DNA-VP2 group, which had the lowest Ab response (LI et al. 2004). The Ab
responses in the chIL-2 plus DNA-VP2 group remained lower compared to the PP or
VP2 alone group, but protection was comparably higher in the former group
indicating stimulation of T-cell responses by IL-2 treatment (LI et al. 2004). IL-2 is
known to promote a Th1-type response, and induces the proliferation and activation
of T-cells and NK cells (HILTON et al. 2002).
Higher IBDV-specific ELISA titers and CD4+ T-cell responses, and greater survival
rates were demonstrated when broiler chickens were vaccinated with a bicistronic
plasmid expressing chIL-2 and VP2 even by using a very low dose (50 µg/bird) of
plasmid (KUMAR et al. 2009). One limitation of this bicistronic plasmid study may be
the quality of the plasmid DNA. Immune modulating effects of lipopolysaccharide
from lysed bacterial cell wall during plasmid DNA isolation additionally have
contributed to the enhanced immunogenicity.
The in vitro immunoregulatory properties of CpG-ODNs on avian APCs, B- and Tcells have been well investigated; yet their applications as avian vaccine adjuvants
are very limited. Class B-CpG-ODN that has been used in previous in vitro and in
vivo studies in avian species was shown to be well tolerated by chickens and they
were used in our study. Despite a relatively high dose of CpG-ODN, DNA-VP2
vaccine did neither induce seroconversion nor protection in the vaccinated chickens.
Co-administration of CpG-ODN at higher doses and frequency together with a DNAPP vaccine aggravated bursal lesions and mortality rates in the challenged chickens
compared to the PP vaccine alone (ROH et al. 2006). No detectable prechallenge
Abs and CMI were demonstrated, which was comparable to our findings (ROH et al.
2006). Others had shown increased IBDV-Ab and CMI responses, and protection
against mortality by virulent IBDV when CpG-ODN was co-administered with a DNAVP2 vaccine; nevertheless clinically recovered birds showed prominent bursal
lesions (WANG et al. 2003; MAHMOOD et al. 2006; MAHMOOD et al. 2007).
88
Previous mouse studies suggested that the co-administration of phosphorothioate
backbone modified CpG-ODN mixed with a plasmid DNA vaccine may reduce gene
expression of the plasmid (WEERATNA et al. 1998), which may have happened in
our study. CpG-ODN stimulated mammalian and avian APCs have been shown to
secrete higher levels of TNF-α and IFN-γ (CIRACI u. LAMONT 2011). It was
suggested that in mice higher expression levels of TNF-α and IFN-γ at the site of
inoculation together with the plasmid DNA preparation may affect the activity of the
cytomegalovirus (CMV) promoter, which results in down-regulation of protein
expression from genes encoded by the vector (QIN et al. 1997). Additionally, in
mammals, CpG-ODN induces proliferation of regulatory T-cells (Tregs) (MOSEMAN
et al. 2004). Tregs may dampen immunity induced by DNA vaccination, which has to
be evaluated further in chickens.
It seems that the in vivo adjuvant property of IL-2 and CpG-ODN depends on the
nature of the vaccine, expression systems of IL-2, and vaccine dosage and
formulations. Previous studies tested the adjuvant property of IL-2 in a co-inoculation
protocol with IBDV DNA vaccines and only partial immune enhancing effects were
demonstrated.
The adjuvant effects of CpG-ODN may be better exploited when immunogens are
administered prior to CpG-ODN application. This hypothesis had been evaluated in a
pneumococcal vaccination model. Delayed CpG-ODN application improved the
protection of mice against pneumococcus challenge compared to simultaneous
administration of the vaccine and CpG-ODN (JENSEN et al. 2012).
6.4. Cationic PLGA MPs as particulate carriers for DNA vaccine and molecular
adjuvants
To improve the immunogenicity and protective efficaciousness of our candidate DNA
vaccine further we have exploited particulate carrier systems for the vaccines and
adjuvants. For this purpose cationic MPs were prepared following previously
89
published protocols, and their physical properties were characterized. Plasmid
encoding the VP2 gene of IBDV or chIL-2, and CpG-ODN were adsorbed on
preformed cationic MPs. These cationic PLGA MPs prepared with PEI had a
diameter of 3-10 µm and are readily phagocytosed by APCs of mammalian and avian
origin (OSTER et al. 2005; LIMAN et al. 2007). The adsorption efficacy of cationic
PLGA MPs was found to be higher for plasmid DNAs than for CpG-ODN as shown in
this study. One hundred percent adsorption efficacy for plasmid DNA was previously
reported (OSTER et al. 2005). Cationic MPs prepared from PLGA Resomer® RG
503H could achieve 100% adsorption efficiency compared to PLGA Resomer® RG
753S as shown in our preliminary tests (data not shown). The former polymer
induced rather highly aggregated MPs and may not be suitable to deliver Ags to
APCs. Our in vitro DNA release studies confirmed release of 45% of adsorbed
plasmid DNA and 60% of CpG-ODN in 24 hrs incubation period, which was lower
compared to a previous report of 78% release of CpG-ODN (SINGH et al. 2001a). An
in vitro DNAse I degradation assay demonstrated considerable DNA protection
against enzymatic degradation when the DNA was bound to MPs (HE et al. 2005). It
was speculated that inclusion of PEI during MPs preparation may condense DNA
molecules on the surface of MPs, which protects the DNA from in vitro degradation
(OSTER et al. 2005). Using mucosal or systemic routes, cationic MPs have been
shown to induce strong immunity to various infectious diseases in animals including
avian species (DENIS-MIZE et al. 2003; OTTEN et al. 2005; LIMAN et al. 2007;
WANG et al. 2011). This provides circumstantial evidence that even under in vivo
conditions MPs may protect DNA from degradation by enzymes.
6.5. Promise of cationic PLGA MPs in improving an IBDV DNA vaccine
We evaluated the efficacy of IBDV DNA vaccine adsorbed onto MPs in SPF chickens
following mucosal application. The adjuvant effects of chIL-2 or CpG-ODN adsorbed
onto cationic MPs were assessed after mucosal priming with DNA vaccine to
augment immune responses provoked by the DNA vaccine.
90
We used bursal pathology, IBDV Ag load in the bursa and intrabursal CD4+- and
CD8+ T-cell counts as markers of protection against IBDV challenge. Chickens
vaccinated with our MPs with adsorbed DNA-VP2 developed moderate bursal
lesions, showed lower IBDV Ag load and less T-cell influx into the bursa compared to
the challenged nonvaccinated controls, adjuvant and vector groups. Experimental
infections of non-vaccinated susceptible SPF chickens with IBDV is expected to
induce extensive intrabursal CD4+- and CD8+ T-cell influx and viral replication in the
bursa (KIM et al. 2000). Our results indicate that even in the absence of detectable
prechallenge
systemic
neutralizing
Abs,
DNA-VP2-MP
vaccinated
chickens
controlled IBDV replication in the bursa. In this case, cell mediated immunity may be
involved in rapid viral clearance and resulted in fewer IBDV Ag positive cells in the
bursa. An increased endolysosomal pH that is suggested to occur after phagocytosis
of the DNA loaded MPs may disrupt the endosomal membrane to facilitate DNA-MPs
complex escape to the cytoplasm (XU u. SZOKA 1996) and support CD8+ T-cell
stimulation (SHEN et al. 2006; HELSON et al. 2008; SCHLIEHE et al. 2011). More
hydrophobic polymers showed delayed degradation within APCs in previous studies
and may even enhance endosomal escape and induce stronger cellular immunity
(XU u. SZOKA 1996). Evaluation of an HIV DNA vaccine delivered by cationic PLGA
MPs through the IM route induced both humoral and cellular immune responses
(OTTEN et al. 2005). This underlines the critical role of polymer nature, its
hydrophobicity, nature of targeted gene and route of immunization for the phenotype
and robustness of immune responses (OTTEN et al. 2005; ROMAN et al. 2008).
Mucosal priming of turkeys with DNA encoding the fusion protein of aMPV adsorbed
onto cationic PLGA followed by boosting with encapsulated fusion protein did not
induce significant aMPV specific ELISA Abs, but elicited F protein-specific VN Abs.
Partial protection against challenge was observed in this vaccination modality
(LIMAN et al. 2007). Many studies have been published concerning the induction of
neutralizing secretory Abs and cell mediated immunity at mucosal surfaces following
oral and intranasal delivery of DNA vaccines loaded onto PLGA MPs (WANG et al.
2011). IgA is important in mucosal immunity. A DNA vaccine expressing βgalactosidase was shown to stimulate an IgA response after eye-drop inoculation in
91
chickens (RUSSELL u. MACKIE 2001). This aspects of mucosal immune responses
in protection against challenge needs to be further validated particularly when DNA
vaccines are being delivered by PLGA MPs.
The delayed application of plasmid encoded chIL-2, which was delivered by cationic
PLGA MPs at least by 48 hrs relative to VP2-MPs, did provide an increased level of
protection compared to the VP2 loaded MPs alone. The lower level of postchallenge
VN Abs, which also coincided with a low level of IL-4 and IFN-α mRNA expression,
suggests rapid viral clearance before sufficient IBDV replication could take place to
induce significant Ab levels. Previously, it was suggested that memory T-cells may
play a significant role in rapid challenge virus clearance from the bursa of chickens
vaccinated with DNA vaccines (CHEN et al. 2011c).
Delayed IL-2 administration has been suggested to amplify DNA vaccine-provoked
Ab and cellular immune responses in humans (BADEN et al. 2011). Effective IL-2
binding to its high-affinity IL-2Rα receptor after DNA vaccination may result in
significant APC and T-cell proliferation. In a lymphocytic choriomeningitis virus
infection model in mice, delayed application of IL-2 induced the proliferation of
memory T-cells that resulted in rapid viral clearance (BLATTMAN et al. 2003). DNA
vaccination preceding IL-2 application may minimize the immunosuppressive effects
of Tregs as well (BRANDENBURG et al. 2008). Tregs suppress T-cell activation
mainly by inhibiting IL-2 gene transcription in CD4+ T-cells (THORNTON et al. 2004).
They do not produce their own IL-2, but may compete for the available IL-2 with
effector CD4+ T-cells leading to decrease in the number of antigen responsive CD4+
T-cells.
It has been of interest in our study to determine whether CpG-ODN may modulate a
MP-IBDV DNA vaccine induced immune response. Our results show that CpG-ODN
indeed improved the immune responses by enhancing the clearance of challenge
virus from the bursa. Previously, it was demonstrated that the biologic effects of
CpG-ODN could be significantly improved by using a delivery vehicle to target APCs
instead of administering it in a soluble form (MALYALA et al. 2009). Our vaccination
92
protocol may have induced expansion of IBDV effector T-cells due to additional
stimulation by CpG-ODN (LIU u. ZHAO 2007).
6.6. Concluding remarks and future perspectives
We have investigated the molecular epidemiology of vvIBDV in Ethiopia and showed
the clonality of vvIBDV strains circulating in the chicken population. While the
information obtained from the nt sequences of the hVP2 of IBDV has been extremely
valuable to understand the genetic and antigenic variability of IBDV, an additional
VP1 characterization is necessary to determine IBDV evolution.
Ethiopian IBDV isolates showed antigenic drift both at the VP2 and VP1 genes
possibly through mutations or recombination. The antigenicity of Ethiopian IBDV
isolates needs to be characterized for example by comparing their reaction patterns
to panels of mAbs, which target the hydrophilic domains of hVP2 or by crossprotection studies with available live IBDV vaccines. By considering the new IBDV
epidemiological situations new vaccination strategy against IBDV may be required.
In spite of several years of experimental studies to validate candidate IBDV DNA
vaccines, their commercial application remains elusive. Thus, a standard DNA
vaccination protocol and optimization in terms of age at vaccination, dosage,
intervals and frequency of DNA vaccination need to be empirically determined.
Usage of proper adjuvants and improved vaccination strategies may enhance the
protective efficacy of these vaccines.
The efficaciousness of IBDV DNA vaccines can be possibly improved further by
adsorbing the DNA vaccine onto cationic PLGA MPs that would allow mucosal
application and additionally provide some adjuvant effects. The efficacy of this MPIBDV DNA vaccine was further potentiated by delayed application of CpG-ODN and
chIL-2 DNA also adsorbed onto cationic MPs.
93
Further studies are required to elucidate the mechanism of protection in our
vaccination protocol using cationic PLGA as delivery system. In particular, the extent
to which T-cells and including Tregs are involved in the immune response and the
distribution of the MP-DNA vaccine after mucosal application require further
investigations. Alternatively, encapsulating CpG-ODN by W/O/W method and
adsorbing DNA-VP2 on the surface of the MPs may mimic our vaccination protocol.
The CpG-ODN can be released from the interior of the MPs after complete hydrolysis
of the PLGA MPs to stimulate immune responses provoked by surface adsorbed
DNA vaccine.
Generally, the induction of higher IBDV-specific Ab titers requires a large dose of
soluble DNA vaccine. Broilers with MAB require at least 25-50 times higher doses of
DNA vaccine compared to SPF chickens for the induction of protective levels of
immune responses due to MAB interference to developing humoral immunity (HSIEH
et al. 2010). Our improved vaccination protocol may allow the administration of lower
doses of IBDV DNA vaccine and adjuvants.
94
7. Summary
Tamiru Negash Alkie
Molecular epidemiology of infectious bursal disease viruses and
development of a microparticle based vaccine
Infectious bursal disease virus (IBDV) is an immunosuppressive virus of chickens
with distribution across all continents. IBDV results in B-cell depletion primarily in the
bursa of Fabricius, but other lymphoid tissues may be affected. Virus proteins (VP) 2
and VP1 contribute to the pathogenicity of IBDV, whereby VP2 as the main structural
protein elicits neutralizing antibodies (Abs).
The present study was designed to understand the molecular epidemiology of IBDV
in Ethiopia and to develop a new generation IBDV DNA vaccine in combination with a
slow release vaccine delivery system on the basis of biodegradable poly(D, L-lactideco-glycolide) (PLGA) microparticles (MPs).
We characterized IBDV field isolates from different Ethiopian chicken flocks showing
clinical IBD. In vivo experimental studies demonstrated that these isolates replicated
extensively
and
induced
significant
bursal
lesions
in
chickens.
Molecular
characterization by RT-PCR and sequencing of the hypervariable region of VP2
(hVP2) and the 5’ two-thirds of VP1 revealed close homology to very virulent (vv)
IBDVs. Phylogenetic analysis of the VP2 region showed that all IBDV isolates had a
vv genotype. Molecularly identified virulence marker amino acids (aa) at the hVP2
region have been well conserved in these strains. The VP1 region encoded aa
residues typical for vvIBDV but also for attenuated IBDV strains. The implication on
virulence mechanism needs further investigation. This molecular epidemiological
study revealed vvIBDV as the major cause of severe losses in the different segments
of poultry production in Ethiopia despite vigorous vaccination with classical live IBDV
vaccines. Questions regarding vaccination failure need to be addressed in the future
including appraisal of the vaccination strategy itself and the selection of vaccine
strains.
95
Reversion of live IBDV vaccines to more virulent strains in the field justifies the
development of new generation IBDV vaccines. In this study plasmid based IBDVDNA vaccines were developed encoding the VP2 gene of two different vvIBDV
strains. Their immunogenicity and the consequent protection were evaluated.
Significant numbers of specific pathogen free (SPF) chickens intramuscularly
vaccinated with IBDV-DNA vaccines did seroconvert, and were partially protected
against oral challenge with vvIBDV. DNA vaccinated chickens developed less severe
bursal lesions compared to non-vaccinated ones. Co-administration of plasmid
encoded chicken IL-2 (chIL-2) or CpG-ODN and the VP2-DNA vaccine neither
enhanced seroconversion nor protection. Subsequently the DNA vaccine application
was modified by adsorption of vaccine and molecular adjuvants onto PLGA based
MPs.
We investigated mucosal (oral and intranasal) application of the VP2-DNA loaded
MP (VP2-MP) in combination with a delayed mucosal delivery of MP loaded either
with chIL-2 or CpG-ODN. Chickens inoculated with the VP2-MP with or without
adjuvants were significantly better protected against oral challenge compared to
chickens inoculated with unloaded-MPs or with MPs loaded with adjuvants or vector
only. VP2-MP-vaccinated chickens showed higher bursal to body weight (B/B) ratio,
lower bursal lesion scores, reduced IBDV antigen load and less T-cell influx into the
bursae compared to the different control groups. Rapid virus clearance particularly in
the group inoculated additionally with chIL-2 loaded MPs lead to lower postchallenge
virus neutralizing Ab titers, and lower IL-4 and IFN-α mRNA expression levels in the
bursa. The results of this vaccination study provide proof-of-concept for further
development and optimization of this vaccination platform as a possible alternative
strategy to protect chickens against IBDV.
96
8. Zusammenfassung
Tamiru Negash Alkie
Molekulare Epidemiologie Infektiöser Bursitis Viren und Entwicklung einer
Mikropartikel-basierten Vakzine
Das Infektiöse Bursitis Virus (IBDV) ist ein immunsuppressives Virus, welches
Hühner infiziert und eine Verbreitung über alle Kontinente aufweist. IBDV löst
vorwiegend in der Bursa Fabricii aber auch in anderen lymphatischen Geweben BZell-Depletion aus. Das Virusprotein (VP) 2 und das VP1 tragen zur Pathogenität von
IBDV bei. Neutralisierende Antikörper (Abs) werden gegen das Hauptstrukturprotein
VP2 gebildet.
Die vorliegende Studie hatte zum Ziel, die molekulare Epidemiologie von IBDV in
Äthiopien zu untersuchen und eine neue Formulierung für eine IBDV-DNA-Vakzine,
zu entwickeln, in dem diese an biologisch abbaubare Poly- (D, L-lactide-co-glycolide)
(PLGA) Mikropartikel (MPs) adsorbiert und nach Applikation langsam freigesetzt
werden kann.
IBDV-Feldisolate aus verschiedenen erkrankten äthiopischen Hühnerbeständen
wurden charakterisiert. In vivo Studien zeigten, dass diese Isolate stark replizierten
und signifikante Läsionen in der Bursa infizierter Hühner hervorriefen. Molekulare
Charakterisierung
durch
RT-PCR
mit
nachfolgender
Sequenzierung
der
hypervariablen Region des VP2 (hVP2) und Zweidritteln des 5´Endes des VP1
zeigten große Homologien zu hochvirulenten (vv) IBDVs. Anhand phylogenetischer
Analysen der VP2-Regionen wurden alle IBDV-Isolate dem hochvirulenten Genotyp
zugeordnet.
Molekulargenetisch
identifizierte
Aminosäuren
(aa),
die
als
Virulenzmarker dienen, waren in der hVP2-Region aller Stämme hochkonserviert.
Die VP1-Region wies sowohl charakterliche aa-Sequenzen von vvIBDV als auch
attenuierten Stämmen auf. Weitere Studien über die Bedeutung dieser aa-Positionen
zur Klärung der Virulenzmechanismen sind nötig. Obwohl die Herden mit klassischen
97
Lebendvakzinen gegen IBDV geimpft waren, weisen die Ergebnisse dieser
molekularepidemiologischen Studie auf die Bedeutung von vvIBDV als Ursache der
großen wirtschaftlichen Verluste in den verschiedenen Geflügel-Produktionsebenen
Äthiopiens hin. Daraus ergeben sich Fragen, wie durch neue Impfstrategien bzw.
durch die Auswahl eines geeigneten Impfstammes ein Impfversagen zukünftig
vermieden werden kann.
Die Möglichkeit der Reversion von IBDV-Lebendvakzinen im Feld zu Stämmen
höherer Virulenz, erfordert die Entwicklung einer neuen Generation von IBDVImpfstoffen. Es wurden zwei Plasmid basierte IBDV-DNA-Vakzinen hergestellt, die
die VP2 Gene von zwei verschiedenen vvIBDVs kodierten. Die Immunogenität, sowie
der hervorgerufene Impf-Schutz wurden evaluiert. Eine signifikante Anzahl an
spezifisch-pathogenfreien Hühnern, die intramuskulär mit den IBDV-DNA-Vakzinen
geimpft worden waren, serokonvertierten und waren partiell gegen eine orale
Belastungsinfektion mit vvIBDV-Stämmen geschützt. Hühner, die mit der DNAVakzine geimpft wurden, entwickelten weniger Bursaläsionen im Vergleich zu nicht
geimpften Tieren. Die gleichzeitige Gabe der DNA-Vakzine mit plasmidkodiertem
Hühner IL-2 (chIL-2) oder CpG-ODN führte weder zu einer besseren Serokonversion
noch zu einem besseren Schutz. Nachfolgend wurde die Applikation der DNAVakzine durch die Adsorption des Impfstoffes und molekularer Adjuvantien an PLGA
basierte MPs modifiziert.
Die orale und intranasale Schleimhautapplikation der VP2-DNA beladenen MP (VP2MP) wurde kombiniert mit einer verzögerten Applikation von mit chIL-2 oder CpGODN beladenen MPs. Hühner, die mit dem VP2-MP geimpft worden waren, waren
unabhängig vom Einsatz von Adjuvantien signifikant besser gegen eine orale
Belastungsinfektion geschützt als Hühner, die mit unbeladenen MPs, MPs mit
Adjuvanz oder Vektor inokuliert worden waren. VP2-MP geimpfte Hühner zeigten ein
besseres Bursa-Körpergewichtsverhältnis (B/B), einen geringeren Bursaläsionsscore, weniger IBDV-Antigen und eine geringere Anzahl an T-Zellen in der Bursa im
Vergleich zu ungeschützten Tieren. Eine schnelle Virus Elimination führte nach der
Belastungsinfektion insbesondere bei den zusätzlich mit chIL-2-MP inokulierten
98
Tieren zu niedrigeren Titern virusneutralisierender Antikörper, sowie zu einer
geringeren Expression von IL-4 und IFN-α mRNA in der Bursa. Das vorgestellte
Impfkonzept stellt eine wichtige Grundlage für weitere Entwicklungen und
Optimierungen der IBDV-DNA Impfung dar und bietet eine mögliche Alternative zum
Schutz von Hühnern gegen IBDV.
99
9. References
ABDEL-ALIM, G. A. u. Y. M. SAIF (2001):
Detection and persistence of infectious bursal disease virus in specific-pathogen-free and
commercial broiler chickens.
Avian Dis 45, 646-654
AL-NATOUR, M. Q., L. A. WARD, Y. M. SAIF, B. STEWART-BROWN u. L. D. KECK (2004):
Effect of diffferent levels of maternally derived antibodies on protection against infectious
bursal disease virus.
Avian Dis 48, 177-182
ARICIBASI, M., A. JUNG, E. D. HELLER u. S. RAUTENSCHLEIN (2010):
Differences in genetic background influence the induction of innate and acquired immune
responses in chickens depending on the virulence of the infecting infectious bursal disease
virus (IBDV) strain.
Vet Immunol Immunopathol 135, 79-92
ASIF, M., K. A. JENKINS, L. S. HILTON, W. G. KIMPTON, A. G. D. BEAN u. J. W.
LOWENTHAL (2004):
Cytokines as adjuvants for avian vaccines.
Immunol Cell Biol 82, 638-643
AZAD, A. A., M. N. JAGADISH, M. A. BROWN u. P. J. HUDSON (1987):
Deletion mapping and expression in Escherichia coli of the large genomic segment of a
birnavirus.
Virology 161, 145-152
BADEN, L. R., W. A. BLATTNER, C. MORGAN, Y. D. HUANG, O. D. DEFAWE, M. E.
SOBIESZCZYK, N. KOCHAR, G. D. TOMARAS, M. J. MCELRATH, N. RUSSELL, K.
BRANDARIZ, M. CARDINALI, B. S. GRAHAM, D. H. BAROUCH, R. DOLIN u. N. H. V. T.
NETWORK (2011):
Timing of Plasmid Cytokine (IL-2/Ig) Administration Affects HIV-1 Vaccine Immunogenicity in
HIV-Seronegative Subjects.
J Infect Dis 204, 1541-1549
BARBON, C. M., L. BAKER, C. LAJOIE, U. RAMSTEDT, M. L. HEDLEY u. T. M. LUBY
(2010):
In vivo electroporation enhances the potency of poly-lactide co-glycolide (PLG) plasmid DNA
immunization.
Vaccine 28, 7852-7864
BAROUCH, D. H., N. L. LETVIN u. R. A. SEDER (2004):
The role of cytokine DNAs as vaccine adjuvants for optimizing cellular immune responses.
Immunol Rev 202, 266-274
BAUMANN, C. L., I. M. ASPALTER, O. SHARIF, A. PICHLMAIR, S. BLUML, F. GREBIEN,
M. BRUCKNER, P. PASIERBEK, K. AUMAYR, M. PLANYAVSKY, K. L. BENNETT, J.
COLINGE, S. KNAPP u. G. SUPERTI-FURGA (2010):
CD14 is a coreceptor of Toll-like receptors 7 and 9.
100
J Exp Med 207, 2689-2701
BAYLISS, C. D., U. SPIES, K. SHAW, R. W. PETERS, A. PAPAGEORGIOU, H. MULLER u.
M. E. BOURSNELL (1990):
A comparison of the sequences of segment A of four infectious bursal disease virus strains
and identification of a variable region in VP2.
J Gen Virol 71 1303-1312
BECHT, H., H. MULLER u. H. K. MULLER (1988):
Comparative studies on structural and antigenic properties of two serotypes of infectious
bursal disease virus.
J Gen Virol 69 ( Pt 3), 631-640
BENDIX, D. (1998):
Chemical synthesis of polylactide and its copolymers for medical applications.
Polym Degrad Stab 59, 129-135
BEREZIN, V. E., A. P. BOGOYAVLENSKYI, S. S. KHUDIAKOVA, P. G. ALEXUK, E. S.
OMIRTAEVA, I. A. ZAITCEVA, G. B. TUSTIKBAEVA, R. C. BARFIELD u. R. H. FETTERER
(2010):
Immunostimulatory complexes containing Eimeria tenella antigens and low toxicity plant
saponins induce antibody response and provide protection from challenge in broiler
chickens.
Vet Parasitol 167, 28-35
BERG, T. P. (2000):
Acute infectious bursal disease in poultry: a review.
Avian Pathol 29, 175-194
BERG, T. P., M. GONZE u. G. MEULEMANS (1991):
Acute infectious bursal disease in poultry: Isolation and characterisation of a highly virulent
strain.
Avian Pathol 20, 133-143
BERG, T. P., M. GONZE, D. MORALES u. G. MEULEMANS (1996):
Acute infectious bursal disease in poultry: Immunological and molecular basis of antigenicity
of a highly virulent strain.
Avian Pathol 25, 751-768
BHAT, A., S. GOMIS, A. POTTER u. S. K. TIKOO (2010):
Role of Hsp90 in CpG ODN mediated immunostimulation in avian macrophages.
Mol Immunol 47, 1337-1346
BHOWMIK, T., B. D'SOUZA, M. N. UDDIN u. M. J. D'SOUZA (2012):
Oral delivery of microparticles containing plasmid DNA encoding hepatitis-B surface antigen.
J Drug Target 20, 364-371
BIRGHAN, C., E. MUNDT u. A. E. GORBALENYA (2000):
A non-canonical lon proteinase lacking the ATPase domain employs the ser-Lys catalytic
dyad to exercise broad control over the life cycle of a double-stranded RNA virus.
Embo J 19, 114-123
101
BIRO, E., K. KOCSIS, N. NAGY, D. MOLNAR, S. KABELL, V. PALYA u. I. OLAH (2011):
Origin of the chicken splenic reticular cells influences the effect of the infectious bursal
disease virus on the extracellular matrix.
Avian Pathol 40, 199-206
BLATTMAN, J. N., J. M. GRAYSON, E. J. WHERRY, S. M. KAECH, K. A. SMITH u. R.
AHMED (2003):
Therapeutic use of IL-2 to enhance antiviral T-cell responses in vivo.
Nat Med 9, 540-547
BOOT, H. J., A. J. W. HOEKMAN u. A. L. J. GIELKENS (2005):
The enhanced virulence of very virulent infectious bursal disease virus is partly determined
by its B-segment.
Archives of virology 150, 137-144
BOOT, H. J., A. A. H. M. TER HUURNE, A. J. W. HOEKMAN, B. P. H. PEETERS u. A. L. J.
GIELKENS (2000):
Rescue of very virulent and mosaic infectious bursal disease virus from cloned cDNA: VP2 is
not the sole determinant of the very virulent phenotype.
J Virol 74, 6701-6711
BOOT, H. J., A. A. H. M. TER HUURNE, A. J. W. HOEKMAN, J. M. POL, A. L. J. GIELKENS
u. B. P. H. PEETERS (2002):
Exchange of the C-terminal part of VP3 from very virulent infectious bursal disease virus
results in an attenuated virus with a unique antigenic structure.
J Virol 76, 10346-10355
BOTTCHER, B., N. A. KISELEV, V. Y. STELMASHCHUK, N. A. PEREVOZCHIKOVA, A. V.
BORISOV u. R. A. CROWTHER (1997):
Three-dimensional structure of infectious bursal disease virus determined by electron
cryomicroscopy.
J Virol 71, 325-330
BRANDENBURG, S., T. TAKAHASHI, M. DE LA ROSA, M. JANKE, G. KARSTEN, T.
MUZZULINI, Z. ORINSKA, S. BULFONE-PAUS u. A. SCHEFFOLD (2008):
IL-2 induces in vivo suppression by CD4(+)CD25(+)Foxp3(+) regulatory T cells.
Eur J Immunol 38, 1643-1653
BRANDT, M., K. YAO, M. LIU, R. A. HECKERT u. V. N. VAKHARIA (2001):
Molecular determinants of virulence, cell tropism, and pathogenic phenotype of infectious
bursal disease virus.
J Virol 75, 11974-11982
BROWN, F. (1989):
The Classification and Nomenclature of Viruses - Summary of Results of Meetings of the
International Committee on Taxonomy of Viruses in Edmonton, Canada 1987.
Intervirology 30, 181-186
BROWN, M. D., P. GREEN u. M. A. SKINNER (1994):
102
VP2 sequences of recent European 'very virulent' isolates of infectious bursal disease virus
are closely related to each other but are distinct from those of 'classical' strains.
J Gen Virol 75 675-680
BROWN, M. D. u. M. A. SKINNER (1996):
Coding sequences of both genome segments of a European 'very virulent' infectious bursal
disease virus.
Virus Res 40, 1-15
BROWNLIE, R., J. ZHU, B. ALLAN, G. K. MUTWIRI, L. A. BABIUK, A. POTTER u. P.
GRIEBEL (2009):
Chicken TLR21 acts as a functional homologue to mammalian TLR9 in the recognition of
CpG oligodeoxynucleotides.
Mol Immunol 46, 3163-3170
BRUN, A., J. BARCENA, E. BLANCO, B. BORREGO, D. DORY, J. M. ESCRIBANO, G. LE
GALL-RECULE, J. ORTEGO u. L. K. DIXON (2011):
Current strategies for subunit and genetic viral veterinary vaccine development.
Virus Res 157, 1-12
BUBLOT, M., N. PRITCHARD, F. X. LE GROS u. S. GOUTEBROZE (2007):
Use of a vectored vaccine against infectious bursal disease of chickens in the face of hightitred maternally derived antibody.
J Comp Pathol 137 S81-84
BUSNADIEGO, I., A. M. MAESTRE, D. RODRIGUEZ u. J. F. RODRIGUEZ (2012):
The Infectious Bursal Disease Virus RNA-Binding VP3 Polypeptide Inhibits PKR-Mediated
Apoptosis.
PLoS One 7, e46768
CAI, M., F. ZHU u. P. SHEN (2012):
Expression and purification of chicken beta interferon and its antivirus immunological activity.
Protein Expr Purif 84, 123-129
CAO, Y. C., Q. C. SHI, J. Y. MA, Q. M. XIE u. Y. Z. BI (2005):
Vaccination against very virulent infectious bursal disease virus using recombinant T4
bacteriophage displaying viral protein VP2.
Acta Biochim Biophys Sin (Shanghai) 37, 657-664
CASTON, J. R., J. L. MARTINEZ-TORRECUADRADA, A. MARAVER, E. LOMBARDO, J. F.
RODRIGUEZ, J. I. CASAL u. J. L. CARRASCOSA (2001):
C terminus of infectious bursal disease virus major capsid protein VP2 is involved in
definition of the T number for capsid assembly.
J Virol 75, 10815-10828
CHAKA, H., F. GOUTARD, S. P. BISSCHOP u. P. N. THOMPSON (2012):
Seroprevalence of Newcastle disease and other infectious diseases in backyard chickens at
markets in Eastern Shewa zone, Ethiopia.
Poult Sci 91, 862-869
CHANG, H.-C., T.-L. LIN u. C.-C. WU (2003):
103
DNA vaccination with plasmids containing various fragments of large segment genome of
infectious bursal disease virus.
Vaccine 21, 507-513
CHANG, H. C., T. L. LIN u. C. C. WU (2001):
DNA-mediated vaccination against infectious bursal disease in chickens.
Vaccine 20, 328-335
CHEN, F., J. J. LIU, Z. Q. YAN, D. LIU, J. JI, J. P. QIN, H. Y. LI, J. Y. MA, Y. Z. BI u. Q. M.
XIE (2012a):
Complete Genome Sequence Analysis of a Natural Reassortant Infectious Bursal Disease
Virus in China.
J Virol 86, 11942-11943
CHEN, H. Y., Y. H. SHANG, H. X. YAO, B. A. CUI, H. Y. ZHANG, Z. X. WANG, Y. D. WANG,
A. J. CHAO u. T. Y. DUAN (2011a):
Immune responses of chickens inoculated with a recombinant fowlpox vaccine coexpressing
HA of H9N2 avain influenza virus and chicken IL-18.
Antiviral Res 91, 50-56
CHEN, H. Y., H. Y. ZHANG, X. S. LI, B. A. CUI, S. J. WANG, J. W. GENG u. K. LI (2011b):
Interleukin-18-mediated enhancement of the protective effect of an infectious
laryngotracheitis virus glycoprotein B plasmid DNA vaccine in chickens.
J Med Microbiol 60, 110-116
CHEN, T. H., T. H. CHEN, C. C. HU, J. T. LIAO, C. W. LEE, J. W. LIAO, M. Y. LIN, H. J. LIU,
M. Y. WANG, N. S. LIN u. Y. H. HSU (2012b):
Induction of protective immunity in chickens immunized with plant-made chimeric Bamboo
mosaic virus particles expressing very virulent Infectious bursal disease virus antigen.
Virus Research 166, 109-115
CHEN, Y. Y., M. K. HSIEH, C. Y. TUNG, C. C. WU u. T. L. LIN (2011c):
Infectious bursal disease DNA vaccination conferring protection by delayed appearance and
rapid clearance of invading viruses.
Arch Virol 156, 2241-2250
CHETTLE, N., J. C. STUART u. P. J. WYETH (1989):
Outbreak of virulent infectious bursal disease in East Anglia.
Vet Rec 125, 271-272
CHEVALIER, C., M. GALLOUX, J. POUS, C. HENRY, J. DENIS, B. DA COSTA, J. NAVAZA,
J. LEPAULT u. B. DELMAS (2005):
Structural peptides of a nonenveloped virus are involved in assembly and membrane
translocation.
J Virol 79, 12253-12263
CHEVALIER, C., J. LEPAULT, B. DA COSTA u. B. DELMAS (2004):
The last C-terminal residue of VP3, glutamic acid 257, controls capsid assembly of infectious
bursal disease virus.
J Virol 78, 3296-3303
104
CHEVALIER, C., J. LEPAULT, I. ERK, B. DA COSTA u. B. DELMAS (2002):
The maturation process of pVP2 requires assembly of infectious bursal disease virus
capsids.
J Virol 76, 2384-2392
CHEVILLE, N. F. (1967):
Studies on the pathogenesis of Gumboro disease in the bursa of Fabricius, spleen, and
thymus of the chicken.
Am J Pathol 51, 527-551
CHEVRIER, N., P. MERTINS, M. N. ARTYOMOV, A. K. SHALEK, M. IANNACONE, M. F.
CIACCIO, I. GAT-VIKS, E. TONTI, M. M. DEGRACE, K. R. CLAUSER, M. GARBER, T. M.
EISENHAURE, N. YOSEF, J. ROBINSON, A. SUTTON, M. S. ANDERSEN, D. E. ROOT, U.
VON ANDRIAN, R. B. JONES, H. PARK, S. A. CARR, A. REGEV, I. AMIT u. N. HACOHEN
(2011):
Systematic discovery of TLR signaling components delineates viral-sensing circuits.
Cell 147, 853-867
CIRACI, C. u. S. J. LAMONT (2011):
Avian-specific TLRs and downstream effector responses to CpG-induction in chicken
macrophages.
Dev Comp Immunol 35, 392-398
COBB, S. P. (2011):
The spread of pathogens through trade in poultry meat: overview and recent developments.
Rev Sci Tech 30, 149-164
CORTEY, M., K. BERTRAN, J. TOSKANO, N. MAJO u. R. DOLZ (2012):
Phylogeographic distribution of very virulent infectious bursal disease virus isolates in the
Iberian Peninsula.
Avian Pathol 41, 277-284
COSGROVE, A. S. (1962):
An apparently new disease of chickens-avian nephrosis.
Avian Dis 6, 385-389
COULIBALY, F., C. CHEVALIER, I. GUTSCHE, J. POUS, J. NAVAZA, S. BRESSANELLI, B.
DELMAS u. F. A. REY (2005):
The birnavirus crystal structure reveals structural relationships among icosahedral viruses.
Cell 120, 761-772
DALLOUL, R. A., H. S. LILLEHOJ, M. OKAMURA, H. XIE, W. MIN, X. DING u. R. A.
HECKERT (2004):
In vivo effects of CpG oligodeoxynucleotide on Eimeria infection in chickens.
Avian Dis 48, 783-790
DAR, A., A. POTTER, S. TIKOO, V. GERDTS, K. LAI, L. A. BABIUK u. G. MUTWIRI (2009):
CpG Oligodeoxynucleotides Activate Innate Immune Response that Suppresses Infectious
Bronchitis Virus Replication in Chicken Embryos.
Avian Dis 53, 261-267
105
DAVIS, B. S., G. J. J. CHANG, B. CROPP, J. T. ROEHRIG, D. A. MARTIN, C. J. MITCHELL,
R. BOWEN u. M. L. BUNNING (2001):
West Nile virus recombinant DNA vaccine protects mouse and horse from virus challenge
and expresses in vitro a noninfectious recombinant antigen that can be used in enzymelinked immunosorbent assays.
J Virol 75, 4040-4047
DE TEMMERMAN, M. L., J. REJMAN, R. E. VANDENBROUCKE, S. DE KOKER, C.
LIBERT, J. GROOTEN, J. DEMEESTER, B. GANDER u. S. C. DE SMEDT (2012):
Polyelectrolyte LbL microcapsules versus PLGA microparticles for immunization with a
protein antigen.
J Control Release 158, 233-239
DE WIT, J. J. (1998):
Gumboro disease: estimation ofoptimal time of vaccination by the Deventer formula (1998). l,
(nr 25),.
Pol Vet J 3, 19-22
DEGEN, W. G. J., H. I. VAN ZUILEKOM, N. C. SCHOLTES, N. VAN DAAL u. V. E. J. C.
SCHIJNS (2005):
Potentiation of humoral immune responses to vaccine antigens by recombinant chicken IL-18
(rChIL-18).
Vaccine 23, 4212-4218
DELGUI, L., A. ONA, S. GUTIERREZ, D. LUQUE, A. NAVARRO, J. R. CASTON u. J. F.
RODRIGUEZ (2009):
The capsid protein of infectious bursal disease virus contains a functional alpha 4 beta 1
integrin ligand motif.
Virology 386, 360-372
DENIS-MIZE, K. S., M. DUPUIS, M. SINGH, C. WOO, M. UGOZZOLI, D. T. O'HAGAN, J. J.
DONNELLY, 3RD, G. OTT u. D. M. MCDONALD (2003):
Mechanisms of increased immunogenicity for DNA-based vaccines adsorbed onto cationic
microparticles.
Cell Immunol 225, 12-20
DI FABIO, J., L. I. ROSSINI, N. ETERRADOSSI, M. D. TOQUIN u. Y. GARDIN (1999):
European-like pathogenic infectious bursal disease viruses in Brazil.
Vet Rec 145, 203-204
DILAVERIS, D., C. L. CHEN, P. KAISER u. P. H. RUSSELL (2007):
The safety and immunogenicity of an in ovo vaccine against Newcastle disease virus differ
between two lines of chicken.
Vaccine 25, 3792-3799
DOBOS, P., B. J. HILL, R. HALLETT, D. T. KELLS, H. BECHT u. D. TENINGES (1979):
Biophysical and biochemical characterization of five animal viruses with bisegmented doublestranded RNA genomes.
J Virol 32, 593-605
106
DOMINGO, E., E. MARTINEZ-SALAS, F. SOBRINO, J. C. DE LA TORRE, A. PORTELA, J.
ORTIN, C. LOPEZ-GALINDEZ, P. PEREZ-BRENA, N. VILLANUEVA, R. NAJERA u. ET AL.
(1985):
The quasispecies (extremely heterogeneous) nature of viral RNA genome populations:
biological relevance--a review.
Gene 40, 1-8
DORMITORIO, T. V., J. J. GIAMBRONE u. L. W. DUCK (1997):
Sequence comparisons of the variable VP2 region of eight infectious bursal disease virus
isolates.
Avian Dis 41, 36-44
DURAIRAJ, V., H. S. SELLERS, E. G. LINNEMANN, A. H. ICARD u. E. MUNDT (2011):
Investigation of the antigenic evolution of field isolates using the reverse genetics system of
infectious bursal disease virus (IBDV).
Arch Virol 156, 1717-1728
DUTHIE, M. S., H. P. WINDISH, C. B. FOX u. S. G. REED (2011):
Use of defined TLR ligands as adjuvants within human vaccines.
Immunol Rev 239, 178-196
ELANKUMARAN, S., R. A. HECKERT u. L. MOURA (2002):
Pathogenesis and tissue distribution of a variant strain of infectious bursal disease virus in
commercial broiler chickens.
Avian Dis 46, 169-176
ELDAGHAYES, I., L. ROTHWELL, A. WILLIAMS, D. WITHERS, S. BALU, F. DAVISON u. P.
KAISER (2006):
Infectious bursal disease virus: Strains that differ in virulence differentially modulate the
innate immune response to infection in the chicken bursa.
Viral Immunol 19, 83-91
ESCAFFRE, O., C. LE NOUËN, M. AMELOT, X. AMBROGGIO, K. M. OGDEN, O.
GUIONIE, D. TOQUIN, H. MÜLLER, M. R. ISLAM u. N. ETERRADOSSI (2012):
Both genome segments contribute to the pathogenicity of very virulent infectious bursal
disease virus.
J Virol in press,
ETERRADOSSI, N u. Y. M. SAIF (2008).
Infectious Bursal Disease. In: Saif YM, Fadly AM, Glisson JR, McDougald LR, Nolan LK,
Swayne DE, editors. Diseases of Poultry 12 ed. Iowa: Iowa State University Press: 1852008.
ETERRADOSSI, N., D. TOQUIN, G. RIVALLAN u. M. GUITTET (1997):
Modified activity of a VP2-located neutralizing epitope on various vaccine, pathogenic and
hypervirulent strains of infectious bursal disease virus.
Arch Virol 142, 255-270
FAHEY, K. J., P. MCWATERS, M. A. BROWN, K. ERNY, V. J. MURPHY u. D. R. HEWISH
(1991):
107
Virus-Neutralizing and Passively Protective Monoclonal-Antibodies to Infectious Bursal
Disease Virus of Chickens.
Avian Dis 35, 365-373
FARLEY, E. K., E. GALE, D. CHAMBERS u. M. LI (2011):
Effects of in ovo electroporation on endogenous gene expression: genome-wide analysis.
Neural Dev 6, 17
FAUREZ, F., D. DORY, V. LE MOIGNE, R. GRAVIER u. A. JESTIN (2010):
Biosafety of DNA vaccines: New generation of DNA vectors and current knowledge on the
fate of plasmids after injection.
Vaccine 28, 3888-3895
FERNANDEZ-ARIAS, A., C. RISCO, S. MARTINEZ, J. P. ALBAR u. J. F. RODRIGUEZ
(1998):
Expression of ORF A1 of infectious bursal disease virus results in the formation of virus-like
particles.
J Gen Virol 79 (1047-1054
FERNANDEZARIAS, A., S. MARTINEZ u. J. F. RODRIGUEZ (1997):
The major antigenic protein of infectious bursal disease virus, VP2, is an apoptotic inducer.
J Virol 71, 8014-8018
FISCHER, S., E. SCHLOSSER, M. MUELLER, N. CSABA, H. P. MERKLE, M. GROETTRUP
u. B. GANDER (2009):
Concomitant delivery of a CTL-restricted peptide antigen and CpG ODN by PLGA
microparticles induces cellular immune response.
J Drug Target 17, 652-661
FLENSBURG, M. F., A. K. ERSBOLL u. P. H. JORGENSEN (2002):
Risk factors associated with the introduction of acute clinical infectious bursal disease among
Danish broiler chickens in 1998.
Avian Pathol 31, 23-29
FOGED, C. (2011):
Subunit vaccines of the future: the need for safe, customized and optimized particulate
delivery systems.
Ther Deliv 2, 1057-1077
FORD, G. W., J. P. THAXTON, T. L. FREDRICKSEN, J. K. TYCZKOWSKI, T. N.
CHAMBLEE u. G. W. MORGAN (2002):
Growth promotion in chickens by interleukin-2.
Growth Dev Aging 65, 73-81
FRANCOIS, A., C. CHEVALIER, B. DELMAS, N. ETERRADOSSI, D. TOQUIN, G.
RIVALLAN u. P. LANGLOIS (2004):
Avian adenovirus CELO recombinants expressing VP2 of infectious bursal disease virus
induce protection against bursal disease in chickens.
Vaccine 22, 2351-2360
FREITAS, S., H. P. MERKLE u. B. GANDER (2005):
108
Microencapsulation by solvent extraction/evaporation: reviewing the state of the art of
microsphere preparation process technology.
J Control Release 102, 313-332
GALLOUX, M., S. LIBERSOU, I. D. ALVES, R. MARQUANT, G. F. SALGADO, H. REZAEI,
J. LEPAULT, B. DELMAS, S. BOUAZIZ u. N. MORELLET (2010):
NMR Structure of a Viral Peptide Inserted in Artificial Membranes A VIEW ON THE EARLY
STEPS OF THE BIRNAVIRUS ENTRY PROCESS.
J Biol Chem 285, 19409-19421
GALLOUX, M., S. LIBERSOU, N. MORELLET, S. BOUAZIZ, B. DA COSTA, M. OULDALI, J.
LEPAULT u. B. DELMAS (2007):
Infectious bursal disease virus, a non-enveloped virus, possesses a capsid-associated
peptide that deforms and perforates biological membranes.
J Biol Chem 282, 20774-20784
GAO, H., K. LI, L. GAO, X. QI, Y. GAO, L. QIN, Y. WANG u. X. WANG (2013):
DNA prime-protein boost vaccination enhances protective immunity against infectious bursal
disease virus in chickens.
Vet Microbiol In press,
GAO, H. L., X. M. WANG, Y. L. GAO u. C. Y. FU (2007):
Direct Evidence of Reassortment and Mutant Spectrum Analysis of a Very Virulent Infectious
Bursal Disease Virus.
Avian Dis 51, 893-899
GAO, L., X. L. QI, K. LI, H. L. GAO, Y. L. GAO, L. T. QIN, Y. Q. WANG u. X. M. WANG
(2011):
Development of a tailored vaccine against challenge with very virulent infectious bursal
disease virus of chickens using reverse genetics.
Vaccine 29, 5550-5557
GARRIGA, D., A. NAVARRO, J. QUEROL-AUDI, F. ABAITUA, J. F. RODRIGUEZ u. N.
VERDAGUER (2007):
Activation mechanism of a noncanonical RNA-dependent RNA polymerase.
Proc Natl Acad Sci U S A 104, 20540-20545
GARRIGA, D., J. QUEROL-AUDI, F. ABAITUA, I. SAUGAR, J. POUS, N. VERDAGUER, J.
R. CASTON u. J. F. RODRIGUEZ (2006):
The 2.6-Angstrom structure of infectious bursal disease virus-derived T=1 particles reveals
new stabilizing elements of the virus capsid.
J Virol 80, 6895-6905
GARVER, K. A., S. E. LAPATRA u. G. KURATH (2005):
Efficacy of an infectious hematopoietic necrosis (IHN) virus DNA vaccine in Chinook
Oncorhynchus tshawytscha and sockeye O-nerka salmon.
Dis Aquat Organ 64, 13-22
GELB, J., JR., D. J. JACKWOOD, E. MUNDT, C. R. POPE, R. HEIN, G. SLACUM, J. M.
HARRIS, B. S. LADMAN, P. LYNCH, D. A. BAUTISTA, J. M. RUANO u. M. M. TROEBER
(2012):
109
Characterization of infectious bursal disease viruses isolated in 2007 from Delmarva
commercial broiler chickens.
Avian Dis 56, 82-89
GIAMBRONE, J. J., T. DORMITORIO u. T. BROWN (2001):
Safety and efficacy of in ovo administration of infectious bursal disease viral vaccines.
Avian Dis 45, 144-148
GOMIS, S., L. BABIUK, B. ALLAN, P. WILLSON, E. WATERS, N. AMBROSE, R. HECKER
u. A. POTTER (2004):
Protection of neonatal chicks against a lethal challenge of Escherichia coli using DNA
containing cytosine-phosphodiester-guanine motifs.
Avian Dis 48, 813-822
GU, J., X. RUAN, Z. HUANG, J. CHEN u. J. ZHOU (2010a):
Identification of functional domains of chicken interleukin 2.
Vet Immunol Immunopathol 134, 230-238
GU, J., J. XU, Z. HUANG, M. LIAO u. J. ZHOU (2010b):
The residues (35)proline and (41)cysteine of chicken IL-2 are critical for binding to chicken
CD25.
Dev Comp Immunol 34, 805-811
GU, J. Y., Q. Y. TENG, Z. Y. HUANG, X. Z. RUAN u. J. Y. ZHOU (2010c):
Identification of the functional interleukin-2 binding domain of the chicken common cytokine
receptor gamma chain.
Dev Comp Immunol 34, 258-263
GUO, X., L. WANG, D. CUI, W. RUAN, F. LIU u. H. LI (2012):
Differential expression of the Toll-like receptor pathway and related genes of chicken bursa
after experimental infection with infectious bursa disease virus.
Arch Virol Epub ahead of print,
HADDAD, E. E., C. E. WHITFILL, A. P. AVAKIAN, C. A. RICKS, P. D. ANDREWS, J. A.
THOMA u. P. S. WAKENELL (1997):
Efficacy of a novel infectious bursal disease virus immune complex vaccine in broiler
chickens.
Avian Dis 41, 882-889
HAQ, K., I. ELAWADLI, P. PARVIZI, A. I. MALLICK, S. BEHBOUDI u. S. SHARIF (2011):
Interferon-gamma influences immunity elicited by vaccines against very virulent Marek's
disease virus.
Antiviral Res 90, 218-226
HARSHYNE, L. A., M. I. ZIMMER, S. C. WATKINS u. S. M. BARRATT-BOYES (2003):
A role for class A scavenger receptor in dendritic cell nibbling from live cells.
J Immunol 170, 2302-2309
HASSAN, M. K., M. AFIFY u. M. M. ALY (2002):
Susceptibility of vaccinated and unvaccinated Egyptian chickens to very virulent infectious
bursal disease virus.
Avian Pathol 31, 149-156
110
HASSAN, M. K., Y. M. SAIF u. S. SHAWKY (1996):
Comparison between antigen-capture ELISA and conventional methods used for titration of
infectious bursal disease virus.
Avian Dis 40, 562-566
HE, C. Q., L. Y. MA, D. WANG, G. R. LI u. N. Z. DING (2009a):
Homologous recombination is apparent in infectious bursal disease virus.
Virology 384, 51-58
HE, H., K. J. GENOVESE, C. L. SWAGGERTY, K. M. MACKINNON u. M. H. KOGUT
(2012a):
Co-stimulation with TLR3 and TLR21 ligands synergistically up-regulates Th1-cytokine IFNgamma and regulatory cytokine IL-10 expression in chicken monocytes.
Dev Comp Immunol 36, 756-760
HE, H. Q., K. J. GENOVESE, D. J. NISBET u. M. H. KOGUT (2006):
Profile of Toll-like receptor expressions and induction of nitric oxide synthesis by Toll-like
receptor agonists in chicken monocytes.
Mol Immunol 43, 783-789
HE, H. Q. u. M. H. KOGUT (2003):
CpG-ODN-induced nitric oxide production is mediated through clathrin-dependent
endocytosis, endosomal maturation, and activation of PKC, MEK1/2 and p38 MAPK, and NFkappa B pathways in avian macrophage cells (HD11).
Cell Signal 15, 911-917
HE, H. Q., K. M. MACKINNON, K. J. GENOVESE, J. R. NERREN, C. L. SWAGGERTY, D. J.
NISBET u. M. H. KOGUT (2009b):
Chicken scavenger receptors and their ligand-induced cellular immune responses.
Mol Immunol 46, 2218-2225
HE, X., L. JIANG, F. WANG, Z. XIAO, J. LI, L. S. LIU, D. LI, D. REN, X. JIN, K. LI, Y. HE, K.
SHI, Y. GUO, Y. ZHANG u. S. SUN (2005):
Augmented humoral and cellular immune responses to hepatitis B DNA vaccine adsorbed
onto cationic microparticles.
J Control Release 107, 357-372
HE, X., P. WEI, X. YANG, D. GUAN, G. WANG u. A. QIN (2012b):
Molecular epidemiology of infectious bursal disease viruses isolated from Southern China
during the years 2000-2010.
Virus Genes 45, 246-255
HEINE, H. G., M. HARITOU, P. FAILLA, K. FAHEY u. A. AZAD (1991):
Sequence-Analysis and Expression of the Host-Protective Immunogen-Vp2 of a Variant
Strain of Infectious Bursal Disease Virus Which Can Circumvent Vaccination with Standard
Type-I Strains.
J Gen Virol 72, 1835-1843
HELSON, R., W. OLSZEWSKA, M. SINGH, J. Z. MEGEDE, J. A. MELERO, D. O. HAGAN u.
P. J. M. OPENSHAW (2008):
111
Polylactide-co-glycolide (PLG) microparticles modify the immune response to DNA
vaccination.
Vaccine 26, 753-761
HERNANDEZ-DIVERS, S. M., P. VILLEGAS, C. JIMENEZ, S. J. HERNANDEZ-DIVERS, M.
GARCIA, S. M. RIBLET, C. R. CARROLL, B. M. O'CONNOR, J. L. WEBB, M. J. YABSLEY,
S. M. WILLIAMS u. S. SANCHEZ (2008):
Backyard chicken flocks pose a disease risk for neotropic birds in Costa Rica.
Avian Dis 52, 558-566
HERNANDEZ, M., A. BANDA, D. HERNANDEZ, F. PANZERA u. R. PEREZ (2006):
Detection of very virulent strains of infectious bursal disease virus (vvIBDV) in commercial
broilers from Uruguay.
Avian Dis 50, 624-631
HIKONO, H., M. MASE, A. MATSUU, M. NAKAYAMA u. T. SAITO (2012):
Intraocular vaccination with an inactivated highly pathogenic avian influenza virus induces
protective antibody responses in chickens.
Vet Immunol Immunopathol In press,
HILTON, L. S., A. G. BEAN, W. G. KIMPTON u. J. W. LOWENTHAL (2002):
Interleukin-2 directly induces activation and proliferation of chicken T cells in vivo.
J Interferon cytokine Res 22, 755-763
HIRAGA, M., T. NUNOYA, Y. OTAKI, M. TAJIMA, T. SAITO u. T. NAKAMURA (1994):
Pathogenesis of highly virulent infectious bursal disease virus infection in intact and
bursectomized chickens.
J Vet Med Sci 56, 1057-1063
HITCHNER, S. B. (1970):
Infectivity of Infectious Bursal Disease Virus for Embryonating Eggs.
Poult Sci 49, 511
HON, C.-C., T.-Y. LAM, A. DRUMMOND, A. RAMBAUT, Y.-F. LEE, C.-W. YIP, F. ZENG, P.Y. LAM, P. T. W. NG u. F. C. C. LEUNG (2006):
Phylogenetic analysis reveals a correlation between the expansion of very virulent infectious
bursal disease virus and reassortment of its genome segment B.
J Virol 80, 8503-8509
HON, C. C., T. T. Y. LAM, C. W. YIP, R. T. Y. WONG, M. SHI, J. W. JIANG, F. Y. ZENG u.
F. C. C. LEUNG (2008):
Phylogenetic evidence for homologous recombination within the family Birnaviridae.
Journal of General Virology 89, 3156-3164
HSIEH, M. K., C. C. WU u. T. L. LIN (2006):
The effect of co-administration of DNA carrying chicken interferon-gamma gene on protection
of chickens against infectious bursal disease by DNA-mediated vaccination.
Vaccine 24, 6955-6965
HSIEH, M. K., C. C. WU u. T. L. LIN (2007):
Priming with DNA vaccine and boosting with killed vaccine conferring protection of chickens
against infectious bursal disease.
112
Vaccine 25, 5417-5427
HSIEH, M. K., C. C. WU u. T. L. LIN (2010):
DNA-mediated vaccination conferring protection against infectious bursal disease in broiler
chickens in the presence of maternal antibody.
Vaccine 28, 3936-3943
HU, Y. C., W. E. BENTLEY, G. H. EDWARDS u. V. N. VAKHARIA (1999):
Chimeric infectious bursal disease virus-like particles expressed in insect cells and purified
by immobilized metal affinity chromatography.
Biotechnol Bioeng 63, 721-729
HUANG, Z. H., S. ELANKUMARAN, A. S. YUNUS u. S. K. SAMAL (2004):
A recombinant newcastle disease virus (NDV) expressing VP2 protein of infectious bursal
disease virus (IBDV) protects against NDV and IBDV.
J Virol 78, 10054-10063
HUDSON, L., M. PATTISON u. N. THANTREY (1975):
Specific B lymphocyte suppression by infectious bursal agent (Gumboro disease virus) in
chickens.
Eur J Immunol 5, 675-679
HULSE, D. J. u. C. H. ROMERO (2004):
Partial protection against infectious bursal disease virus through DNA-mediated vaccination
with the VP2 capsid protein and chicken IL-2 genes.
Vaccine 22, 1249-1259
HUNG, L. H., H. P. LI, Y. Y. LIEN, M. L. WU u. H. C. CHAUNG (2010):
Adjuvant effects of chicken interleukin-18 in avian Newcastle disease vaccine.
Vaccine 28, 1148-1155
ICARD, A. H., H. S. SELLERS u. E. MUNDT (2008):
Detection of infectious bursal disease virus isolates with unknown antigenic properties by
reverse genetics.
Avian Dis 52, 590-598
IGNJATOVIC, J., G. GOULD, L. TRINIDAD u. S. SAPATS (2006):
Chicken recombinant antibodies against infectious bursal disease virus are able to form
antibody-virus immune complex.
Avian Pathol 35, 293-301
INOUE, M., A. FUJITA u. K. MAEDA (1999):
Lysis of myelocytes in chickens infected with infectious bursal disease virus.
Vet Pathol 36, 146-151
IRIGOYEN, N., J. R. CASTON u. J. F. RODRIGUEZ (2012):
Host Proteolytic Activity Is Necessary for Infectious Bursal Disease Virus Capsid Protein
Assembly.
J Biol Chem 287, 24473-24482
113
IRIGOYEN, N., D. GARRIGA, A. NAVARRO, N. VERDAGUER, J. F. RODRIGUEZ u. J. R.
CASTON (2009):
Autoproteolytic Activity Derived from the Infectious Bursal Disease Virus Capsid Protein.
J Biol Chem 284, 8064-8072
ISLAM, M. R., K. ZIERENBERG u. H. MULLER (2001):
The genome segment B encoding the RNA-dependent RNA polymerase protein VP1 of very
virulent infectious bursal disease virus (IBDV) is phylogenetically distinct from that of all other
IBDV strains.
Arch Virol 146, 2481-2492
IVAN, J., N. NAGY, A. MAGYAR, I. KACSKOVICS u. J. MESZAROS (2001):
Functional restoration of the bursa of Fabricius following in ovo infectious bursal disease
vaccination.
Vet Immunol Immunopathol 79, 235-248
IVAN, J., M. VELHNER, K. URSU, P. GERMAN, T. MATO, C. N. DREN u. J. MESZAROS
(2005):
Delayed vaccine virus replication in chickens vaccinated subcutaneously with an immune
complex infectious bursal disease vaccine: quantification of vaccine virus by real-time
polymerase chain reaction.
Can J Vet Res 69, 135-142
JACKWOOD, D. H. u. Y. M. SAIF (1987):
Antigenic Diversity of Infectious Bursal Disease Viruses.
Avian Dis 31, 766-770
JACKWOOD, D. J. (2004):
Recent trends in the molecular diagnosis of infectious bursal disease viruses.
Anim Health Res Rev 5, 313-316
JACKWOOD, D. J. (2012):
Molecular epidemiologic evidence of homologous recombination in infectious bursal disease
viruses.
Avian Dis 56, 574-577
JACKWOOD, D. J., B. M. CROSSLEY, S. T. STOUTE, S. SOMMER-WAGNER, P. R.
WOOLCOCK u. B. R. CHARLTON (2012):
Diversity of genome segment B from infectious bursal disease viruses in the United States.
Avian Dis 56, 165-172
JACKWOOD, D. J., Y. M. SAIF u. P. D. MOORHEAD (1985):
Immunogenicity and antigenicity of infectious bursal disease virus serotypes I and II in
chickens.
Avian Dis 29, 1184-1194
JACKWOOD, D. J. u. S. SOMMER-WAGNER (2007):
Genetic characteristics of infectious bursal disease viruses from four continents.
Virology 365, 369-375
JACKWOOD, D. J. u. S. E. SOMMER-WAGNER (2005):
114
Molecular epidemiology of infectious bursal disease viruses: distribution and genetic analysis
of newly emerging viruses in the United States.
Avian Dis 49, 220-226
JACKWOOD, D. J. u. S. E. SOMMER-WAGNER (2010):
Detection and characterization of infectious bursal disease viruses in broilers at processing.
Preven Vet Med 97, 45-50
JACKWOOD, D. J. u. S. E. SOMMER-WAGNER (2011):
Amino acids contributing to antigenic drift in the infectious bursal disease Birnavirus (IBDV).
Virology 409, 33-37
JACKWOOD, D. J., S. E. SOMMER-WAGNER, B. M. CROSSLEY, S. T. STOUTE, P. R.
WOOLCOCK u. B. R. CHARLTON (2011):
Identification and pathogenicity of a natural reassortant between a very virulent serotype 1
infectious bursal disease virus (IBDV) and a serotype 2 IBDV.
Virology 420, 98-105
JACKWOOD, D. J. u. S. E. SOMMER (2002):
Identification of infectious bursal disease virus quasispecies in commercial vaccines and field
isolates of this double-stranded RNA virus.
Virology 304, 105-113
JACKWOOD, D. J. u. S. E. SOMMER (2005):
Molecular studies on suspect very virulent infectious bursal disease virus genomic RNA
samples.
Avian Dis 49, 246-251
JACKWOOD, D. J., S. E. SOMMER u. H. V. KNOBLICH (2001):
Amino acid comparison of infectious bursal disease viruses placed in the same or different
molecular groups by RT/PCR-RFLP.
Avian Dis 45, 330-339
JACKWOOD, D. J., B. D. SPALDING u. S. E. SOMMER (2003):
Real-time reverse transcriptase-polymerase chain reaction detection and analysis of
nucleotide sequences coding for a neutralizing epitope on infectious bursal disease viruses.
Avian Dis 47, 738-744
JACKWOOD, D. J., B. SREEDEVI, L. J. LEFEVER u. S. E. SOMMER-WAGNER (2008):
Studies on naturally occurring infectious bursal disease viruses suggest that a single amino
acid substitution at position 253 in VP2 increases pathogenicity.
Virology 377, 110-116
JAIN, S., D. T. O'HAGAN u. M. SINGH (2011):
The long-term potential of biodegradable poly(lactide-co-glycolide) microparticles as the
next-generation vaccine adjuvant.
Expert Rev Vaccines 10, 1731-1742
JANG, S. I., H. S. LILLEHOJ, S. H. LEE, K. W. LEE, E. P. LILLEHOJ, F. BERTRAND, L.
DUPUIS u. S. DEVILLE (2011):
115
Montanide IMS 1313 N VG PR nanoparticle adjuvant enhances antigen-specific immune
responses to profilin following mucosal vaccination against Eimeria acervulina.
Vet Parasitol 182, 163-170
JENBREIE, S., G. AYELET, E. GELAYE, F. KEBEDE, S. E. LYNCH u. H. NEGUSSIE
(2012):
Infectious bursal disease: seroprevalence and associated risk factors in major poultry rearing
areas of Ethiopia.
Trop Anim Health Prod In press,
JENSEN, K. M., J. MELCHJORSEN, F. DAGNAES-HANSEN, U. B. S. SORENSEN, R. R.
LAURSEN, L. OSTERGAARD, O. S. SOGAARD u. M. TOLSTRUP (2012):
Timing of Toll-Like Receptor 9 Agonist Administration in Pneumococcal Vaccination Impacts
Both Humoral and Cellular Immune Responses as Well as Nasopharyngeal Colonization in
Mice.
Infect Immun 80, 1744-1752
JEON, W.-J., E.-K. LEE, S.-J. JOH, J.-H. KWON, C.-B. YANG, Y.-S. YOON u. K.-S. CHOI
(2008):
Very virulent infectious bursal disease virus isolated from wild birds in Korea: epidemiological
implications.
Virus Res 137, 153-156
JEONG, J., W. H. KIM, J. YOO, C. LEE, S. KIM, J. H. CHO, H. K. JANG, D. W. KIM, H. S.
LILLEHOJ u. W. MIN (2012):
Identification and Comparative Expression Analysis of Interleukin 2/15 Receptor beta Chain
in Chickens Infected with E. tenella.
PLoS One 7, e37704
JEURISSEN, S. H. M., E. M. JANSE, P. R. LEHRBACH, E. E. HADDAD, A. AVAKIAN u. C.
E. WHITFILL (1998):
The working mechanism of an immune complex vaccine that protects chickens against
infectious bursal disease.
Immunology 95, 494-500
JILEK, S., H. P. MERKLE u. E. WALTER (2005):
DNA-loaded biodegradable microparticles as vaccine delivery systems and their interaction
with dendritic cells.
Adv Drug Deliv Rev 57, 377-390
JILEK, S., M. ULRICH, H. P. MERKLE u. E. WALTER (2004):
Composition and surface charge of DNA-loaded microparticles determine maturation and
cytokine secretion in human dendritic cells.
Pharm Res 21, 1240-1247
JOHNSTON, P. A., H. LIU, T. O'CONNELL, P. PHELPS, M. BLAND, J. TYCZKOWSKI, A.
KEMPER, T. HARDING, A. AVAKIAN, E. HADDAD, C. WHITFILL, R. GILDERSLEEVE u. C.
A. RICKS (1997):
Applications in in ovo technology.
Poult Sci 76, 165-178
JONES, D. H., J. C. CLEGG u. G. H. FARRAR (1998):
116
Oral delivery of micro-encapsulated DNA vaccines.
Dev Biol Stand 92, 149-155
JUNEJA, S. S., RAMNEEK, D. DEKA, M. S. OBEROI u. A. SINGH (2008):
Molecular characterization of field isolates and vaccine strains of infectious bursal disease
virus.
Comp Immunol Microbiol Infect Dis 31, 11-23
KABELL, S., K. J. HANDBERG, Y. LI, M. KUSK u. M. BISGAARD (2005):
Detection of vvIBDV in vaccinated SPF chickens.
Acta Vet Scand 46, 219-227
KABELL, S., B. Z. IGYARTO, A. MAGYAR, Z. HAJDU, E. BIRO, M. BISGAARD u. I. OLAH
(2006):
Impact of heterophil granulocyte depletion caused by 5-fluorouracil on infectious bursal
disease virus infection in specific pathogen free chickens.
Avian Pathol 35, 341-348
KAISER, P. (2010):
Advances in avian immunology-prospects for disease control: a review.
Avian Pathol 39, 309-324
KAISER, P., T. Y. POH, L. ROTHWELL, S. AVERY, S. BALU, U. S. PATHANIA, S.
HUGHES, M. GOODCHILD, S. MORRELL, M. WATSON, N. BUMSTEAD, J. KAUFMAN u.
J. R. YOUNG (2005):
A genomic analysis of chicken cytokines and chemokines.
J Interferon Cytokine Res 25, 467-484
KAMPA, D. u. J. BURNSIDE (2002):
Jak3-regulated genes: DNA array analysis of concanavalin A-interleukin-2-activated chicken
T cells treated with a specific Jak3 inhibitor.
J Interferon Cytokine Res 22, 975-980
KARPALA, A. J., K. R. MORRIS, M. M. BROADWAY, P. G. MCWATERS, T. E. O'NEIL, K. E.
GOOSSENS, J. W. LOWENTHAL u. A. G. BEAN (2008):
Molecular cloning, expression, and characterization of chicken IFN -lambda.
J Interferon Cytokine Res 28, 341-350
KASANGA, C. J., T. YAMAGUCHI, H. M. MUNANG'ANDU, K. OHYA u. H. FUKUSHI (2012):
Genomic sequence of an infectious bursal disease virus isolate from Zambia: classical
attenuated segment B reassortment in nature with existing very virulent segment A.
Arch Virol
KASANGA, C. J., T. YAMAGUCHI, P. N. WAMBURA, A. D. MAEDA-MACHANG'U, K. OHYA
u. H. FUKUSHI (2007):
Molecular characterization of infectious bursal disease virus (IBDV): Diversity of very virulent
IBDV in Tanzania.
Arch Virol 152, 783-790
KASANGA, C. J., T. YAMAGUCHI, P. N. WAMBURA, H. M. MUNANG'ANDU, K. OHYA u. H.
FUKUSHI (2008):
117
Detection of infectious bursal disease virus (IBDV) genome in free-living pigeon and guinea
fowl in Africa suggests involvement of wild birds in the epidemiology of IBDV.
Virus Genes 36, 521-529
KAUFER, I. u. E. WEISS (1980):
Significance of Bursa of Fabricius as Target Organ in Infectious Bursal Disease of Chickens.
Infect Immun 27, 364-367
KECK, L. D., J. K. SKEELES u. R. W. MCNEW (1993):
Antibody detection in matched chicken sera and egg-yolk samples by commercial enzymelinked immunosorbent assay kits for Newcastle disease virus, infectious bronchitis virus,
infectious bursal disease virus, and avian reovirus.
Avian Dis 37, 825-828
KEESTRA, A. M., M. R. DE ZOETE, L. I. BOUWMAN u. J. P. M. VAN PUTTEN (2010):
Chicken TLR21 is an innate CpG DNA receptor distinct from mammalian TLR9.
J Immunol 185, 460-467
KHATRI, M., J. A. PALMQUIST, R. M. CHA u. J. A. SHARMA (2005):
Infection and activation of bursal macrophages by virulent infectious bursal disease virus.
Virus Res 113, 44-50
KHATRI, M. u. J. M. SHARMA (2006):
Infectious bursal disease virus infection induces macrophage activation via p38 MAPK and
NF-kappaB pathways.
Virus Res 118, 70-77
KHATRI, M. u. J. M. SHARMA (2009a):
Response of embryonic chicken lymphoid cells to infectious bursal disease virus.
Vet Immunol Immunopathol 127, 316-324
KHATRI, M. u. J. M. SHARMA (2009b):
Susceptibility of chicken mesenchymal stem cells to infectious bursal disease virus.
J Virol Methods 160, 197-199
KIBENGE, F. S. B., B. QIAN, E. NAGY, J. R. CLEGHORN u. D. WADOWSKA (1999):
Formation of virus-like particles when the polyprotein gene (segment A) of infectious bursal
disease virus is expressed in insect cells.
Can J Vet Res 63, 49-55
KIM, I. J., M. GAGIC u. J. M. SHARMA (1999):
Recovery of antibody-producing ability and lymphocyte repopulation of bursal follicles in
chickens exposed to infectious bursal disease virus.
Avian Dis 43, 401-413
KIM, I. J., K. KARACA, T. L. PERTILE, S. A. ERICKSON u. J. M. SHARMA (1998):
Enhanced expression of cytokine genes in spleen macrophages during acute infection with
infectious bursal disease virus in chickens.
Vet Immunol Immunopathol 61, 331-341
KIM, I. J. u. J. M. SHARMA (2000):
118
IBDV-induced bursal T lymphocytes inhibit mitogenic response of normal splenocytes.
Vet Immunol Immunopathol 74, 47-57
KIM, I. J., S. K. YOU, H. KIM, H. Y. YEH u. J. M. SHARMA (2000):
Characteristics of bursal T lymphocytes induced by infectious bursal disease virus.
J Virol 74, 8884-8892
KIM, S., L. FARIS, C. M. COX, L. H. SUMNERS, M. C. JENKINS, R. H. FETTERER, K. B.
MISKA u. R. A. DALLOUL (2012):
Molecular characterization and immunological roles of avian IL-22 and its soluble receptor IL22 binding protein.
Cytokine 60, 815-827
KIM, S. J., H. W. SUNG, J. H. HAN, D. JACKWOOD u. H. M. KWON (2004):
Protection against very virulent infectious bursal disease virus in chickens immunized with
DNA vaccines.
Vet Microbiol 101, 39-51
KNOBLICH, H. V., S. E. SOMMER u. D. J. JACKWOOD (2000):
Antibody titers to infectious bursal disease virus in broiler chicks after vaccination at one day
of age with infectious bursal disease virus and Marek's disease virus.
Avian Dis 44, 874-884
KOGUT, M. H., M. IQBAL, H. HE, V. PHILBIN, P. KAISER u. A. SMITH (2005):
Expression and function of Toll-like receptors in chicken heterophils.
Dev Comp Immunol 29, 791-807
KUMAR, P. A., S. K. DAS u. J. R. RAO (1998):
Effect of immunostimulation on cytotoxic activity of intestinal intraepithelial lymphocytes of
chickens in infectious bursal disease and Eimeria tenella infections.
Acta vet Hung 46, 1-11
KUMAR, S., Y. S. AHI, S. S. SALUNKHE, M. KOUL, A. K. TIWARI, P. K. GUPTA u. A. RAI
(2009):
Effective protection by high efficiency bicistronic DNA vaccine against infectious bursal
disease virus expressing VP2 protein and chicken IL-2.
Vaccine 27, 864-869
KUMAR, S., M. KOUL u. A. RAI (2010):
Role of Immunostimulatory Molecules in Poultry Vaccines.
Recent Pat Biotechnol 4, 235-241
LAHOUD, M. H., F. AHMET, J. G. ZHANG, S. MEUTER, A. N. POLICHENI, S. KITSOULIS,
C. N. LEE, M. O'KEEFFE, L. C. SULLIVAN, A. G. BROOKS, R. BERRY, J. ROSSJOHN, J.
D. MINTERN, J. VEGA-RAMOS, J. A. VILLADANGOS, N. A. NICOLA, M. C.
NUSSENZWEIG, K. J. STACEY, K. SHORTMAN, W. R. HEATH u. I. CAMINSCHI (2012):
DEC-205 is a cell surface receptor for CpG oligonucleotides.
Proc Natl Acad Sci U S A 109, 16270-16275
LAM, K. M. (1997):
Morphological evidence of apoptosis in chickens infected with infectious bursal disease virus.
119
J Comp Pathol 116, 367-377
LANA, D. P., C. E. BEISEL u. R. F. SILVA (1992):
Genetic Mechanisms of Antigenic Variation in Infectious Bursal Disease Virus - Analysis of a
Naturally-Occurring Variant Virus.
Virus Genes 6, 247-259
LANGER, R., L. G. CIMA, J. A. TAMADA u. E. WINTERMANTEL (1990):
Future directions in biomaterials.
Biomaterials 11, 738-745
LASHER, H. N., V. S. DAVIS, D. LASHER u. A. S. COSGROVE (1997):
History of infectious bursal disease in the U.S.A.--the first two decades.
Avian Dis 41, 11-19
LE GROS, F. X., A. DANCER, C. GIACOMINI, L. PIZZONI, M. BUBLOT, M. GRAZIANI u. F.
PRANDINI (2009):
Field efficacy trial of a novel HVT-IBD vector vaccine for 1-day-old broilers.
Vaccine 27, 592-596
LE NOUEN, C., G. RIVALLAN, D. TOQUIN, P. DARLU, Y. MORIN, W. BEVEN, C. DE
BOISSESON, C. CAZABAN, S. COMTE, Y. GARDIN u. N. ETERRADOSSI (2006):
Very virulent infectious bursal disease virus: reduced pathogenicity in a rare natural
segment-B-reassorted isolate.
J Gen Virol 87, 209-216
LE NOUEN, C., G. RIVALLAN, D. TOQUIN u. N. ETERRADOSSI (2005):
Significance of the genetic relationships deduced from partial nucleotide sequencing of
infectious bursal disease virus genome segments A or B.
Arch Virol 150, 313-325
LE NOUEN, C. L., D. TOQUIN, H. MULLER, R. RAUE, K. M. KEAN, P. LANGLOIS, M.
CHERBONNEL u. N. ETERRADOSSI (2012):
Different domains of the RNA polymerase of infectious bursal disease virus contribute to
virulence.
PLoS One 7, e28064
LEE, C.-C., T.-P. KO, C.-C. CHOU, M. YOSHIMURA, S.-R. DOONG, M.-Y. WANG u. A. H. J.
WANG (2006):
Crystal structure of infectious bursal disease virus VP2 subviral particle at 2.6A resolution:
implications in virion assembly and immunogenicity.
J Struct Biol 155, 74-86
LEE, L. H. u. Y. P. LIN (1992):
A monoclonal antibody capture enzyme-linked immunosorbent assay for detecting antibodies
to infectious bursal disease virus.
J Virol Methods 36, 13-23
LEE, S. W., J. M. DEVLIN, J. F. MARKHAM, A. H. NOORMOHAMMADI, G. F. BROWNING,
N. P. FICORILLI, C. A. HARTLEY u. P. F. MARKHAM (2013):
120
Phylogenetic and Molecular Epidemiological Studies Reveal Evidence of Multiple Past
Recombination Events between Infectious Laryngotracheitis Viruses.
PLoS One 8, e55121
LEE, S. W., P. F. MARKHAM, M. J. COPPO, A. R. LEGIONE, J. F. MARKHAM, A. H.
NOORMOHAMMADI, G. F. BROWNING, N. FICORILLI, C. A. HARTLEY u. J. M. DEVLIN
(2012):
Attenuated vaccines can recombine to form virulent field viruses.
Science 337, 188
LEJAL, N., B. DA COSTA, J. C. HUET u. B. DELMAS (2000):
Role of Ser-652 and Lys-692 in the protease activity of infectious bursal disease virus VP4
and identification of its substrate cleavage sites.
J Gen Virol 81, 983-992
LETZEL, T., F. COULIBALY, F. A. REY, B. DELMAS, E. JAGT, A. A. VAN LOON u. E.
MUNDT (2007):
Molecular and structural bases for the antigenicity of VP2 of infectious bursal disease virus.
J Virol 81, 12827-12835
LI, J., Y. HUANG, X. LIANG, M. LU, L. LI, L. YU u. E. AL. (2003):
Plasmid DNA encoding antigens of infectious bursal disease viruses induce protective
immune responses in chickens: factors influencing efficacy.
Virus Res 98, 63-74
LI, J., X. LIANG, Y. HUANG, S. MENG, R. XIE, R. DENG u. L. YU (2004):
Enhancement of the immunogenicity of DNA vaccine against infectious bursal disease virus
by co-delivery with plasmid encoding chicken interleukin 2.
Virology 329, 89-100
LI, K., H. GAO, L. GAO, X. QI, Y. GAO, L. QIN, Y. WANG u. X. WANG (2013a):
Adjuvant effects of interleukin-18 in DNA vaccination against infectious bursal disease virus
in chickens.
Vaccine In press,
LI, L., W. FANG, J. LI, Y. HUANG u. L. YU (2006):
Oral DNA vaccination with the polyprotein gene of infectious bursal disease virus (IBDV)
delivered by the attenuated Salmonella elicits protective immune responses in chickens.
Vaccine 24, 5919-5927
LI, X. Y., S. JIANG u. R. I. TAPPING (2010):
Toll-like receptor signaling in cell proliferation and survival.
Cytokine 49, 1-9
LI, Y. P., K. J. HANDBERG, H. R. JUUL-MADSEN, M. F. ZHANG u. P. H. JORGENSEN
(2007):
Transcriptional profiles of chicken embryo cell cultures following infection with infectious
bursal disease virus.
Arch Virol 152, 463-478
LI, Z., Y. WANG, X. LI, H. CAO u. S. J. ZHENG (2013b):
121
Critical roles of glucocorticoid-induced leucine zipper in infectious bursal disease virus
(IBDV)-induced suppression of type I Interferon expression and enhancement of IBDV
growth in host cells via interaction with VP4.
J Virol 87, 1221-1231
LI, Z. H., Y. Q. WANG, Y. F. XUE, X. Q. LI, H. CAO u. S. J. J. ZHENG (2012):
Critical Role for Voltage-Dependent Anion Channel 2 in Infectious Bursal Disease VirusInduced Apoptosis in Host Cells via Interaction with VP5.
Journal of Virology 86, 1328-1338
LILLEHOJ, H. S., W. MIN, K. D. CHOI, U. S. BABU, J. BURNSIDE, T. MIYAMOTO, B. M.
ROSENTHAL u. E. P. LILLEHOJ (2001):
Molecular, cellular, and functional characterization of chicken cytokines homologous to
mammalian IL-15 and IL-2.
Vet Immunol Immunopathol 82, 229-244
LIM, B. L., Y. C. CAO, T. YU u. C. W. MO (1999):
Adaptation of very virulent infectious bursal disease virus to chicken embryonic fibroblasts by
site-directed mutagenesis of residues 279 and 284 of viral coat protein VP2.
J Virol 73, 2854-2862
LIM, K. L., S. D. JAZAYERI, S. K. YEAP, N. B. ALITHEEN, M. H. BEJO, A. IDERIS u. A. R.
OMAR (2012):
Co-administration of avian influenza virus H5 plasmid DNA with chicken IL-15 and IL-18
enhanced chickens immune responses.
BMC Vet Res 8, 132
LIMAN, M., L. PEISER, G. ZIMMER, M. PROSTING, H. Y. NAIM u. S. RAUTENSCHLEIN
(2007):
A genetically engineered prime-boost vaccination strategy for oculonasal delivery with
poly(D,L-lactic-co-glycolic acid) microparticles against infection of turkeys with avian
Metapneumovirus.
Vaccine 25, 7914-7926
LIN, T. W., C. W. LO, S. Y. LAI, R. J. FAN, C. J. LO, Y. M. CHOU, R. THIRUVENGADAM, A.
H. J. WANG u. M. Y. WANG (2007):
Chicken heat shock protein 90 is a component of the putative cellular receptor complex of
infectious bursal disease virus.
J Virol 81, 8730-8741
LIN, Z., A. KATO, Y. OTAKI, T. NAKAMURA, E. SASMAZ u. S. UEDA (1993):
Sequence Comparisons of a Highly Virulent Infectious Bursal Disease Virus Prevalent in
Japan.
Avian Dis 37, 315-323
LIU, C. G., M. LIU, F. LIU, D. F. LIU, Y. ZHANG, W. Q. PAN, H. CHEN, C. H. WAN, E. C.
SUN, H. T. LI u. W. H. XIANG (2011):
Evaluation of several adjuvants in avian influenza vaccine to chickens and ducks.
Virol J 8, 321
LIU, G. W. u. Y. ZHAO (2007):
122
Toll-like receptors and immune regulation: their direct and indirect modulation on regulatory
CD4(+) CD25(+) T cells.
Immunology 122, 149-156
LIU, M. u. V. N. VAKHARIA (2004):
VP1 protein of infectious bursal disease virus modulates the virulence in vivo.
Virology 330, 62-73
LIU, M. H. u. V. N. VAKHARIA (2006):
Nonstructural protein of infectious bursal disease virus inhibits apoptosis at the early stage of
virus infection.
J Virol 80, 3369-3377
LOMBARDO, E., A. MARAVER, I. ESPINOSA, A. FERNANDEZ-ARIAS u. J. F.
RODRIGUEZ (2000):
VP5, the nonstructural polypeptide of infectious bursal disease virus, accumulates within the
host plasma membrane and induces cell lysis.
Virology 277, 345-357
LUKERT, P. D. u. R. B. DAVIS (1974):
Infectious bursal disease virus: growth and characterization in cell cultures.
Avian Dis 18, 243-250
LUO, J., H. ZHANG, M. TENG, J. M. FAN, L. M. YOU, Z. J. XIAO, M. L. YI, Y. B. ZHI, X. W.
LI u. G. P. ZHANG (2010):
Surface IgM on DT40 cells may be a component of the putative receptor complex
responsible for the binding of infectious bursal disease virus.
Avian Pathol 39, 359-365
LUQUE, D., I. SAUGAR, M. T. REJAS, J. L. CARRASCOSA, J. F. RODRIGUEZ u. J. R.
CASTON (2009):
Infectious Bursal Disease Virus: Ribonucleoprotein Complexes of a Double-Stranded RNA
Virus.
J Mol Biol 386, 891-901
LUQUE, D., I. SAUGAR, J. F. RODRIGUEZ, N. VERDAGUER, D. GARRIGA, C. S. MARTIN,
J. A. VELAZQUEZ-MURIEL, B. L. TRUS, J. L. CARRASCOSA u. J. R. CASTON (2007):
Infectious bursal disease virus capsid assembly and maturation by structural rearrangements
of a transient molecular switch.
J Virol 81, 6869-6878
MAAS, R. A., S. VENEMA, H. L. OEI, J. M. POL, I. J. CLAASSEN u. A. A. HUURNE (2001):
Efficacy of inactivated infectious bursal disease (IBD) vaccines: comparison of serology with
protection of progeny chickens against IBD virus strains of varying virulence.
Avian Pathol 30, 345-354
MAHGOUB, H. A., M. BAILEY u. P. KAISER (2012):
An overview of infectious bursal disease.
Arch Virol Epub ahead of print,
MAHMOOD, M. S., I. HUSSAIN, M. SIDDIQUE, M. AKHTAR u. S. ALI (2007):
123
DNA vaccination with VP2 gene of very virulent infectious bursal disease virus (vvIBDV)
delivered by transgenic E. coli DH5alpha given orally confers protective immune responses
in chickens.
Vaccine 25, 7629-7635
MAHMOOD, M. S., M. SIDDIQUE, I. HUSSAIN, A. KHAN u. M. K. MANSOOR (2006):
Protection capability of recombinant plasmid DNA vaccine containing VP2 gene of very
virulent infectious bursal disease virus in chickens adjuvanted with CpG
oligodeoxynucleotide.
Vaccine 24, 4838-4846
MALATHI, K., B. DONG, M. GALE, JR. u. R. H. SILVERMAN (2007):
Small self-RNA generated by RNase L amplifies antiviral innate immunity.
Nature 448, 816-819
MALLICK, A. I., R. R. KULKARNI, M. ST PAUL, P. PARVIZI, E. NAGY, S. BEHBOUDI u. S.
SHARIF (2012):
Vaccination with CpG-adjuvanted avian influenza virosomes promotes antiviral immune
responses and reduces virus shedding in chickens.
Viral Immunol 25, 226-231
MALYALA, P., J. CHESKO, M. UGOZZOLI, A. GOODSELL, F. ZHOU, M. VAJDY, D. T.
O'HAGAN u. M. SINGH (2008):
The potency of the adjuvant, CpG oligos, is enhanced by encapsulation in PLG
microparticles.
J Pharm Sci 97, 1155-1164
MALYALA, P., D. T. O'HAGAN u. M. SINGH (2009):
Enhancing the therapeutic efficacy of CpG oligonucleotides using
microparticles.
Adv Drug Deliv Rev 61, 218-225
biodegradable
MARAVER, A., R. CLEMENTE, J. F. RODRIGUEZ u. E. LOMBARDO (2003):
Identification and molecular characterization of the RNA polymerase-binding motif of
infectious bursal disease virus inner capsid protein VP3.
J Virol 77, 2459-2468
MARINO, M., F. SCUDERI, C. PROVENZANO u. E. BARTOCCIONI (2011):
Skeletal muscle cells: from local inflammatory response to active immunity.
Gene Therapy 18, 109-116
MARQUARDT, W. W., R. B. JOHNSON, W. F. ODENWALD u. B. A. SCHLOTTHOBER
(1980):
An indirect enzyme-linked immunosorbent assay (ELISA) for measuring antibodies in
chickens infected with infectious bursal disease virus.
Avian Dis 24, 375-385
MARTIN, A. M., F. FALLACARA, I. BARBIERI, G. TOSI, G. RIVALLAN, N. ETERRADOSSI,
R. CERUTI u. P. CORDIOLI (2007):
Genetic and antigenic characterization of infectious bursal disease viruses isolated in Italy
during the period 2002-2005.
124
Avian Dis 51, 863-872
MARTINEZ-TORRECUADRADA, J. L., N. SAUBI, A. PAGES-MANTE, J. R. CASTON, E.
ESPUNA u. J. I. CASAL (2003):
Structure-dependent efficacy of infectious bursal disease virus (IBDV) recombinant vaccines.
Vaccine 21, 3342-3350
MAZARIEGOS, L. A., P. D. LUKERT u. J. BROWN (1990):
Pathogenicity and immunosuppressive properties of infectious bursal disease "intermediate"
strains.
Avian Dis 34, 203-208
MCFERRAN, J. B., M. S. MCNULTY, E. R. MCKILLOP, T. J. CONNOR, R. M.
MCCRACKEN, D. S. COLLINS u. G. M. ALLAN (1980):
Isolation and serological studies with infectious bursal disease viruses from fowl, turkeys and
ducks: demonstration of a second serotype.
Avian Pathol 9, 395-404
MEIR, R., D. J. JACKWOOD u. Y. WEISMAN (2001):
Molecular typing of infectious bursal disease virus of Israeli field and vaccine strains by the
reverse transcription/polymerase chain reaction/restriction fragment length polymorphism
assay.
Avian Dis 45, 223-228
MILLER, T., M. FANTON, S. NICKELSON, H. MASON u. S. WEBB (2012):
Safety and immunogenicity of bacterial and tobacco plant cell line derived recombinant
native and mutant Escherichia coli heat-labile toxin in chickens.
Avian Pathol 41, 441-449
MOODY, A., S. SELLERS u. N. BUMSTEAD (2000):
Measuring infectious bursal disease virus RNA in blood by multiplex real-time quantitative
RT-PCR.
J Virol Methods 85, 55-64
MORESCO, E. M. Y. u. B. BEUTLER (2011):
Special Delivery: Granulin Brings CpG DNA to Toll-like Receptor 9.
Immunity 34, 453-455
MORIMOTO, K., D. C. HOOPER, H. CARBAUGH, Z. F. FU, H. KOPROWSKI u. B.
DIETZSCHOLD (1998):
Rabies virus quasispecies: implications for pathogenesis.
Proc Natl Acad Sci U S A 95, 3152-3156
MOSEMAN, E. A., X. LIANG, A. J. DAWSON, A. PANOSKALTSIS-MORTARI, A. M. KRIEG,
Y. J. LIU, B. R. BLAZAR u. W. CHEN (2004):
Human plasmacytoid dendritic cells activated by CpG oligodeoxynucleotides induce the
generation of CD4+CD25+ regulatory T cells.
J Immunol 173, 4433-4442
MÜLLER, H., M. R. ISLAM u. R. RAUE (2003):
Research on infectious bursal disease--the past, the present and the future.
125
Vet Microbiol 97, 153-165
MÜLLER, H., E. MUNDT, N. ETERRADOSSI u. M. R. ISLAM (2012):
Current status of vaccines against infectious bursal disease.
Avian Pathol 41, 133-139
MÜLLER, H. u. R. NITSCHKE (1987):
The two segments of the infectious bursal disease virus genome are circularized by a
90,000-Da protein.
Virology 159, 174-177
MULLER, H., C. SCHOLTISSEK u. H. BECHT (1979):
The genome of infectious bursal disease virus consists of two segments of double-stranded
RNA.
J Virol 31, 584-589
MÜLLER, R., I. KAUFER, M. REINACHER u. E. WEISS (1979):
Immunofluorescent studies of early virus propagation after oral infection with infectious
bursal disease virus (IBDV).
Zentralbl Veterinarmed B 26, 345-352
MUNDT, E. (1999):
Tissue culture infectivity of different strains of infectious bursal disease virus is determined by
distinct amino acids in VP2.
J Gen Virol 80, 2067-2076
MUNDT, E., N. DE HAAS u. A. A. W. M. VAN LOON (2003):
Development of a vaccine for immunization against classical as well as variant strains of
infectious bursal disease virus using reverse genetics.
Vaccine 21, 4616-4624
MUNDT, E., B. KOLLNER u. D. KRETZSCHMAR (1997):
VP5 of infectious bursal disease virus is not essential for viral replication in cell culture.
J Virol 71, 5647-5651
MUNDT, E. u. H. MÜLLER (1995):
Complete nucleotide sequences of 5'- and 3'-noncoding regions of both genome segments of
different strains of infectious bursal disease virus.
Virology 209, 10-18
MUNDT, E. u. V. N. VAKHARIA (1996):
Synthetic transcripts of double-stranded Birnavirus genome are infectious.
Proc Natl Acad Sci U S A 93, 11131-11136
MURAKAMI, T. u. Y. SUNADA (2011):
Plasmid DNA gene therapy by electroporation: principles and recent advances.
Curr Gene Ther 11, 447-456
NAGARAJAN, M. M. u. F. S. KIBENGE (1997):
The 5'-terminal 32 basepairs conserved between genome segments A and B contain a major
promoter element of infectious bursal disease virus.
126
Arch Virol 142, 2499-2514
NICOLETE, R., D. F. DOS SANTOS u. L. H. FACCIOLI (2011):
The uptake of PLGA micro or nanoparticles by macrophages provokes distinct in vitro
inflammatory response.
Int immunopharmacol 11, 1557-1563
NIEDERSTADT, L., O. HOHN, B. G. DORNER, R. SCHADE u. N. BANNERT (2012):
Stimulation of IgY responses in gene gun immunized laying hens by combined administration
of vector DNA coding for the target antigen Botulinum toxin A1 and for avian cytokine
adjuvants.
J Immunol Methods 382, 58-67
O'HAGAN, D. T. (2001):
Recent developments in vaccine delivery systems.
Curr Drug Targets Infect Disord 1, 273-286
O'HAGAN, D. T., M. SINGH, C. DONG, M. UGOZZOLI, K. BERGER, E. GLAZER, M.
SELBY, M. WININGER, P. NG, K. CRAWFORD, X. PALIARD, S. COATES u. M.
HOUGHTON (2004a):
Cationic microparticles are a potent delivery system for a HCV DNA vaccine.
Vaccine 23, 672-680
O'HAGAN, D. T., M. SINGH u. J. B. ULMER (2004b):
Microparticles for the delivery of DNA vaccines.
Immunol Rev 199, 191-200
O'HAGAN, D. T., M. SINGH u. J. B. ULMER (2006):
Microparticle-based technologies for vaccines.
Methods 40, 10-19
O'NEILL, A. M., E. J. LIVANT u. S. J. EWALD (2010):
Interferon alpha-induced inhibition of infectious bursal disease virus in chicken embryo
fibroblast cultures differing in Mx genotype.
Avian Dis 54, 802-806
OGAWA, H., T. YAMAGUCHI, A. SETIYONO, T. HO, H. MATSUDA, S. FURUSAWA, H.
FUKUSHI u. K. HIRAI (1998):
Some characteristics of a cellular receptor for virulent infectious bursal disease virus by using
flow cytometry.
Arch Virol 143, 2327-2341
OSINALDE, N., H. MOSS, O. ARRIZABALAGA, M. J. OMAETXEBARRIA, B. BLAGOEV, A.
M. ZUBIAGA, A. FULLAONDO, J. M. ARIZMENDI u. I. KRATCHMAROVA (2011):
Interleukin-2 signaling pathway analysis by quantitative phosphoproteomics.
J Proteomics 75, 177-191
OSTER, C. G., N. KIM, L. GRODE, L. BARBU-TUDORAN, A. K. SCHAPER, S. H. E.
KAUFMANN u. T. KISSEL (2005):
Cationic microparticles consisting of poly(lactide-co-glycolide) and polyethylenimine as
carriers systems for parental DNA vaccination.
127
J Control Release 104, 359-377
OSTER, C. G. u. T. KISSEL (2005):
Comparative study of DNA encapsulation into PLGA microparticles using modified double
emulsion methods and spray drying techniques.
J Microencapsul 22, 235-244
OTTEN, G. R., M. SCHAEFER, B. DOE, H. LIU, I. SRIVASTAVA, J. ZUR MEGEDE, J.
KAZZAZ, Y. LIAN, M. SINGH, M. UGOZZOLI, D. MONTEFIORI, M. LEWIS, D. A. DRIVER,
T. DUBENSKY, J. M. POLO, J. DONNELLY, D. T. O'HAGAN, S. BARNETT u. J. B. ULMER
(2005):
Enhanced potency of plasmid DNA microparticle human immunodeficiency virus vaccines in
rhesus macaques by using a priming-boosting regimen with recombinant proteins.
J Virol 79, 8189-8200
PALMQUIST, J. M., M. KHATRI, R. M. CHA, B. M. GODDEERIS, B. WALCHECK u. J. M.
SHARMA (2006):
In vivo activation of chicken macrophages by infectious bursal disease virus.
Viral immunol 19, 305-315
PAN, J., V. N. VAKHARIA u. Y. J. TAO (2007):
The structure of a birnavirus polymerase reveals a distinct active site topology.
Proc Natl Acad Sci U S A 104, 7385-7390
PARDOLL, D. M. u. A. M. BECKERLEG (1995):
Exposing the immunology of naked DNA vaccines.
Immunity 3, 165-169
PARK, B., L. BUTI, S. LEE, T. MATSUWAKI, E. SPOONER, M. M. BRINKMANN, M.
NISHIHARA u. H. L. PLOEGH (2011):
Granulin Is a Soluble Cofactor for Toll-like Receptor 9 Signaling.
Immunity 34, 505-513
PARK, J. H., H. W. SUNG, B. I. YOON u. H. M. KWON (2009):
Protection of chicken against very virulent IBDV provided by in ovo priming with DNA vaccine
and boosting with killed vaccine and the adjuvant effects of plasmid-encoded chicken
interleukin-2 and interferon-gamma.
J Vet Sci 10, 131-139
PEISER, L. (2006):
Development and characterization of a controlled release system for in ovo-vaccination of
chickens. .
Doctoral Thesis, University of Veterinary Medicine, Hannover, Germany
PETERS, M. A., T. L. LIN u. C. C. WU (2004):
Infectious bursal disease virus polyprotein expression arrests growth and mitogenic
stimulation of B lymphocytes.
Arch Virol 149, 2413-2426
PETKOV, D. I., E. G. LINNEMANN, D. R. KAPCZYNSKI u. H. S. SELLERS (2009):
Identification and characterization of two distinct bursal B-cell subpopulations following
infectious bursal disease virus infection of White Leghorn chickens.
128
Avian Dis 53, 347-355
PHENIX, K. V., K. WARK, C. J. LUKE, M. A. SKINNER, J. A. SMYTH, K. A. MAWHINNEY u.
D. TODD (2001):
Recombinant Semliki Forest virus vector exhibits potential for avian virus vaccine
development.
Vaccine 19, 3116-3123
PHILBIN, V. J., M. IQBAL, Y. BOYD, M. J. GOODCHILD, R. K. BEAL, N. BUMSTEAD, J.
YOUNG u. A. L. SMITH (2005):
Identification and characterization of a functional, alternatively spliced Toll-like receptor 7
(TLR7) and genomic disruption of TLR8 in chickens.
Immunology 114, 507-521
PITCOVSKI, J., D. GOLDBERG, B. Z. LEVI, D. DI-CASTRO, A. AZRIEL, S. KRISPEL, T.
MARAY u. Y. SHAALTIEL (1998):
Coding region of segment a sequence of a very virulent isolate of IBDV - Comparison with
isolates from different countries and virulence.
Avian Dis 42, 497-506
PITCOVSKI, J., B. GUTTER, G. GALLILI, M. GOLDWAY, B. PERELMAN, G. GROSS, S.
KRISPEL, M. BARBAKOV u. A. MICHAEL (2003):
Development and large-scale use of recombinant VP2 vaccine for the prevention of
infectious bursal disease of chickens.
Vaccine 21, 4736-4743
PRADHAN, S. N., P. R. PRINCE, J. MADHUMATHI, P. ROY, R. B. NARAYANAN u. U.
ANTONY (2012):
Protective immune responses of recombinant VP2 subunit antigen of infectious bursal
disease virus in chickens.
Vet Immunol Immunopathol 148, 293-301
QIAN, C., S. CHEN, P. DING, M. CHAI, C. XU, J. GAN, D. PENG u. X. LIU (2012):
The immune response of a recombinant fowlpox virus coexpressing the HA gene of the
H5N1 highly pathogenic avian influenza virus and chicken interleukin 6 gene in ducks.
Vaccine 30, 6279-6286
QIN, L., Y. DING, D. R. PAHUD, E. CHANG, M. J. IMPERIALE u. J. S. BROMBERG (1997):
Promoter attenuation in gene therapy: interferon-gamma and tumor necrosis factor-alpha
inhibit transgene expression.
Hum Gene Ther 8, 2019-2029
QIN, L. T., X. L. QI, Y. L. GAO, H. L. GAO, X. H. LU, Y. Q. WANG, Z. G. BU u. X. M. WANG
(2010):
VP5-deficient mutant virus induced protection against challenge with very virulent infectious
bursal disease virus of chickens.
Vaccine 28, 3735-3740
RAHMAN, M. M. u. S. K. EO (2012):
Prospects and challenges of using chicken cytokines in disease prevention.
Vaccine 30, 7165-7173
129
RAHMAN, M. M., E. UYANGAA, Y. W. HAN, S. B. KIM, J. H. KIM, J. Y. CHOI u. S. K. EO
(2012):
Enhancement of Th1-biased protective immunity against avian influenza H9N2 virus via oral
co-administration of attenuated Salmonella enterica serovar Typhimurium expressing
chicken interferon-alpha and interleukin-18 along with an inactivated vaccine.
BMC Vet Res 8, 105
RAUE, R., M. R. ISLAM, M. N. ISLAM, K. M. ISLAM, S. C. BADHY, P. M. DAS u. H.
MULLER (2004):
Reversion of molecularly engineered, partially attenuated, very virulent infectious bursal
disease virus during infection of commercial chickens.
Avian Pathol 33, 181-189
RAUF, A., M. KHATRI, M. V. MURGIA, K. JUNG u. Y. M. SAIF (2011a):
Differential modulation of cytokine, chemokine and Toll like receptor expression in chickens
infected with classical and variant infectious bursal disease virus.
Vet Res 42, 85
RAUF, A., M. KHATRI, M. V. MURGIA u. Y. M. SAIF (2011b):
Expression of perforin-granzyme pathway genes in the bursa of infectious bursal disease
virus-infected chickens.
Dev Comp Immunol 35, 620-627
RAUTENSCHLEIN, S., C. KRAEMER, J. VANMARCKE u. E. MONTIEL (2005a):
Protective efficacy of intermediate and intermediate plus infectious bursal disease virus
(IBDV) vaccines against very virulent IBDV in commercial broilers.
Avian Dis 49, 231-237
RAUTENSCHLEIN, S., C. KRAEMER, J. VANMARCKE u. E. MONTIEL (2005b):
Protective efficacy of intermediate and intermediate plus infectious bursal disease virus
(IBDV) vaccines against very virulent IBDV in commercial broilers.
Avian Dis 49, 231-237
RAUTENSCHLEIN, S., G. VON SAMSON-HIMMELSTJERNA u. C. HAASE (2007):
A comparison of immune responses to infection with virulent infectious bursal disease virus
(IBDV) between specific-pathogen-free chickens infected at 12 and 28 days of age.
Vet Immunol Immunopathol 115, 251-260
RAUTENSCHLEIN, S., H. Y. YEH, M. K. NJENGA u. J. M. SHARMA (2002a):
Role of intrabursal T cells in infectious bursal disease virus (IBDV) infection: T cells promote
viral clearance but delay follicular recovery.
Arch Virol 147, 285-304
RAUTENSCHLEIN, S., H. Y. YEH u. J. M. SHAMA (2002b):
The role of T cells in protection by an inactivated infectious bursal disease virus vaccine.
Vet Immunol Immunopathol 89, 159-167
RAUW, F., B. LAMBRECHT u. T. VAN DEN BERG (2007):
Pivotal role of ChIFNgamma in the pathogenesis and immunosuppression of infectious
bursal disease.
130
Avian Pathol 36, 367-374
REDDY, K. S., B. R. RASHMI, H. J. DECHAMMA, S. GOPALAKRISHNA, N. BANUMATHI,
V. V. S. SURYANARAYANA u. G. R. REDDY (2012):
Cationic microparticle [poly(D,L-lactide-co-glycolide)]-coated DNA vaccination induces a
long-term immune response against foot and mouth disease in guinea pigs.
J Gen Med 14, 348-352
RODRIGUEZ-CHAVEZ, I. R., J. K. ROSENBERGER, S. S. CLOUD u. C. R. POPE (2002):
Characterization of the antigenic, immunogenic, and pathogenic variation of infectious bursal
disease virus due to propagation in different host systems (bursa, embryo, and cell culture).
III. Pathogenicity.
Avian Pathol 31, 485-492
ROGERS, T. H. u. J. E. BABENSEE (2011):
The role of integrins in the recognition and response of dendritic cells to biomaterials.
Biomaterials 32, 1270-1279
ROH, H. J., H. W. SUNG u. H. M. KWON (2006):
Effects of DDA, CpG-ODN, and plasmid-encoded chicken IFN-gamma on protective
immunity by a DNA vaccine against IBDV in chickens.
J Vet Sci 7, 361-368
ROMAN, B. S., J. M. IRACHE, S. GOMEZ, N. TSAPIS, C. GAMAZO u. M. S. ESPUELAS
(2008):
Co-encapsulation of an antigen and CpG oligonucleotides into PLGA microparticles by
TROMS technology.
Eur J Pharm Biopharm 70, 98-108
RONG, J., T. JIANG, T. CHENG, M. SHEN, Y. DU, S. LI, S. WANG, B. XU u. G. FAN (2007):
Large-scale manufacture and use of recombinant VP2 vaccine against infectious bursal
disease in chickens.
Vaccine 25, 7900-7908
ROSALES, A. G., P. VILLEGAS, P. D. LUKERT, O. J. FLETCHER, M. A. MOHAMED u. J.
BROWN (1989):
Isolation, identification, and pathogenicity of two field strains of infectious bursal disease
virus.
Avian Dis 33, 35-41
ROSE, D. M., J. W. HAN u. M. H. GINSBERG (2002):
alpha 4 integrins and the immune response.
Immunol Rev 186, 118-124
RUSSELL, P. H. u. A. MACKIE (2001):
Eye-drop DNA can induce IgA in the tears and bile of chickens.
Vet Immunol Immunopathol 80, 327-332
SANCHEZ, A. B. u. J. F. RODRIGUEZ (1999):
Proteolytic processing in infectious bursal disease virus: Identification of the polyprotein
cleavage sites by site-directed mutagenesis.
131
Virology 262, 190-199
SAPATS, S., G. GOULD, L. TRINIDAD, L. H. PAREDE, C. DAVID u. J. IGNJATOVIC (2005):
An ELISA for detection of infectious bursal disease virus and differentiation of very virulent
strains based on single chain recombinant chicken antibodies.
Avian Pathol 34, 449-455
SAPATS, S. I., L. TRINIDAD, G. GOULD, H. G. HEINE, T. P. VAN DEN BERG, N.
ETERRADOSSI, D. JACKWOOD, L. PAREDE, D. TOQUIN u. J. IGNJATOVIC (2006):
Chicken recombinant antibodies specific for very virulent infectious bursal disease virus.
Arch Virol 151, 1551-1566
SAUGAR, I., N. IRIGOYEN, D. LUQUE, J. L. CARRASCOSA, J. F. RODRIGUEZ u. J. R.
CASTON (2010):
Electrostatic Interactions between Capsid and Scaffolding Proteins Mediate the Structural
Polymorphism of a Double-stranded RNA Virus.
J Biol Chem 285, 3643-3650
SAUGAR, I., D. LUQUE, A. ONA, J. F. RODRIGUEZ, J. L. CARRASCOSA, B. L. TRUS u. J.
R. CASTON (2005):
Structural polymorphism of the major capsid protein of a double-stranded RNA virus: An
amphipathic alpha helix as a molecular switch.
Structure 13, 1007-1017
SAWANT, P. M., P. C. VERMA, P. K. SUBUDHI, U. CHATURVEDI, M. SINGH, R. KUMAR
u. A. K. TIWARI (2011):
Immunomodulation of bivalent Newcastle disease DNA vaccine induced immune response
by co-delivery of chicken IFN-gamma and IL-4 genes.
Vet Immunol Immunopathol 144, 36-44
SCHLIEHE, C., C. REDAELLI, S. ENGELHARDT, M. FEHLINGS, M. MUELLER, N. VAN
ROOIJEN, M. THIRY, K. HILDNER, H. WELLER u. M. GROETTRUP (2011):
CD8(-) Dendritic Cells and Macrophages Cross-Present Poly( D,L-lactate-co-glycolate) Acid
Microsphere-Encapsulated Antigen In Vivo.
J Immunol 187, 2112-2121
SCHNITZLER, D., F. BERNSTEIN, H. MULLER u. H. BECHT (1993):
The genetic basis for the antigenicity of the VP2 protein of the infectious bursal disease virus.
J Gen Virol 74 1563-1571
SESTER, D. P., S. NAIK, S. J. BEASLEY, D. A. HUME u. K. J. STACEY (2000):
Phosphorothioate backbone modification modulates macrophage activation by CpG DNA.
J Immunol 165, 4165-4173
SHANMUGASUNDARAM, R. u. R. K. SELVARAJ (2011):
Regulatory T cell properties of chicken CD4+CD25+ cells.
J Immunol 186, 1997-2002
SHARMA, J. M., K. KARACA u. T. PERTILE (1994):
Virus-induced immunosuppression in chickens.
Poult Sci 73, 1082-1086
132
SHARMA, J. M., I. J. KIM, S. RAUTENSCHLEIN u. H. Y. YEH (2000):
Infectious bursal disease virus of chickens: pathogenesis and immunosuppression.
Dev Comp Immunol 24, 223-235
SHCHERBAKOVA, L. O., A. I. LOMAKIN, A. V. BORISOV, V. V. DRYGIN u. A. A. GUSEV
(1998):
[Comparative analysis of the VP2 variable region of the gene from infectious bursal disease
virus isolates].
Mol Gen Mikrobiol Virusol 35-40
SHEN, H., A. L. ACKERMAN, V. CODY, A. GIODINI, E. R. HINSON, P. CRESSWELL, R. L.
EDELSON, W. M. SALTZMAN u. D. J. HANLON (2006):
Enhanced and prolonged cross-presentation following endosomal escape of exogenous
antigens encapsulated in biodegradable nanoparticles.
Immunology 117, 78-88
SHI, L. X., H. B. LI, G. P. MA, J. Y. ZHOU, L. L. HONG, X. J. ZHENG, Y. P. WU, Y. Z.
WANG u. Y. YAN (2009):
Competitive replication of different genotypes of infectious bursal disease virus on chicken
embryo fibroblasts.
Virus Genes 39, 46-52
SHIROTA, H., L. PETRENKO, C. HONG u. D. M. KLINMAN (2007):
Potential of transfected muscle cells to contribute to DNA vaccine immunogenicity.
J Immunol 179, 329-336
SHOSHAN-BARMATZ, V., V. DE PINTO, M. ZWECKSTETTER, Z. RAVIV, N. KEINAN u. N.
ARBEL (2010):
VDAC, a multi-functional mitochondrial protein regulating cell life and death.
Mol Aspects Med 31, 227-285
SHWED, P. S., P. DOBOS, L. A. CAMERON, V. N. VAKHARIA u. R. DUNCAN (2002):
Birnavirus VP1 proteins form a distinct subgroup of RNA-dependent RNA polymerases
lacking a GDD motif.
Virology 296, 241-250
SILVA, F. M., P. M. VIDIGAL, L. W. MYRRHA, J. L. FIETTO, A. J. SILVA u. M. R. ALMEIDA
(2012):
Tracking the molecular epidemiology of Brazilian Infectious bursal disease virus (IBDV)
isolates
Infect Genet Evol 13C, 18-26
SINGH, M., M. BRIONES, G. OTT u. D. O'HAGAN (2000):
Cationic microparticles: A potent delivery system for DNA vaccines.
Proc Natl Acad Sci U S A 97, 811-816
SINGH, M., J. H. FANG, J. KAZZAZ, M. UGOZZOLI, J. CHESKO, P. MALYALA, R.
DHALIWAL, R. WEI, M. HORA u. D. O'HAGAN (2006):
A modified process for preparing cationic polylactide-co-glycolide microparticles with
adsorbed DNA.
133
Int J Pharm 327, 1-5
SINGH, M., J. P. MCGEE, X. M. LI, W. KOFF, T. ZAMB, C. Y. WANG u. D. T. O'HAGAN
(1997):
Biodegradable microparticles with an entrapped branched octameric peptide as a controlledrelease HIV-1 vaccine.
J Pharm Sci 86, 1229-1233
SINGH, M., G. OTT, J. KAZZAZ, M. UGOZZOLI, M. BRIONES, J. DONNELLY u. D. T.
O'HAGAN (2001a):
Cationic microparticles are an effective delivery system for immune stimulatory cpG DNA.
Pharm Res 18, 1476-1479
SINGH, M., M. VAJDY, J. GARDNER, M. BRIONES u. D. O'HAGAN (2001b):
Mucosal immunization with HIV-1 gag DNA on cationic microparticles prolongs gene
expression and enhances local and systemic immunity.
Vaccine 20, 594-602
SKEELES, J. K., P. D. LUKERT, E. V. DE BUYSSCHER, O. J. FLETCHER u. J. BROWN
(1979):
Infectious bursal disease viral infections. II. The relationship of age, complement levels,
virus-neutralizing antibody, clotting, and lesions.
Avian Dis 23, 107-117
SNYDER, D. B., D. P. LANA, P. K. SAVAGE, F. S. YANCEY, S. A. MENGEL u. W. W.
MARQUARDT (1988):
Differentiation of infectious bursal disease viruses directly from infected tissues with
neutralizing monoclonal antibodies: evidence of a major antigenic shift in recent field isolates.
Avian Dis 32, 535-539
SNYDER, D. B., V. N. VAKHARIA u. P. K. SAVAGE (1992):
Naturally Occurring-Neutralizing Monoclonal-Antibody Escape Variants
Epidemiology of Infectious Bursal Disease Viruses in the United-States.
Arch Virol 127, 89-101
Define
the
SREEDEVI, B., L. J. LEFEVER, S. E. SOMMER-WAGNER u. D. J. JACKWOOD (2007):
Characterization of infectious bursal disease viruses from four layer flocks in the United
States.
Avian Dis 51, 845-850
ST PAUL, M., N. BARJESTEH, S. PAOLUCCI, Y. PEI u. S. SHARIF (2012a):
Toll-like receptor ligands induce the expression of interferon-gamma and interleukin-17 in
chicken CD4+ T cells.
BMC Res Notes 5, 616
ST PAUL, M., A. I. MALLICK, K. HAQ, S. OROUJI, M. F. ABDUL-CAREEM u. S. SHARIF
(2011):
In vivo administration of ligands for chicken toll-like receptors 4 and 21 induces the
expression of immune system genes in the spleen.
Vet Immunol Immunopathol 144, 228-237
134
ST PAUL, M., A. I. MALLICK, L. R. READ, A. I. VILLANUEVA, P. PARVIZI, M. F. ABDULCAREEM, E. NAGY u. S. SHARIF (2012b):
Prophylactic treatment with Toll-like receptor ligands enhances host immunity to avian
influenza virus in chickens.
Vaccine 30, 4524-4531
ST PAUL, M., S. PAOLUCCI, N. BARJESTEH, R. D. WOOD, K. A. SCHAT u. S. SHARIF
(2012c):
Characterization of chicken thrombocyte responses to Toll-like receptor ligands.
PLoS One 7, e43381
ST PAUL, M., S. PAOLUCCI, N. BARJESTEH, R. D. WOOD u. S. SHARIF (2013):
Chicken erythrocytes respond to Toll-like receptor ligands by up-regulating cytokine
transcripts.
Res Vet Sci In press,
ST PAUL, M., S. PAOLUCCI, L. R. READ u. S. SHARIF (2012d):
Characterization of responses elicited by Toll-like receptor agonists in cells of the bursa of
Fabricius in chickens.
Vet Immunol Immunopathol 149, 237-244
STOUTE, S. T., D. J. JACKWOOD, S. E. SOMMER-WAGNER, G. L. COOPER, M. L.
ANDERSON, P. R. WOOLCOCK, A. A. BICKFORD, C. G. SENTIES-CUE u. B. R.
CHARLTON (2009):
The Diagnosis of Very Virulent Infectious Bursal Disease in California Pullets.
Avian Dis 53, 321-326
SU, B. S., H. S. YIN, H. H. CHIU, L. H. HUNG, J. P. HUANG, J. H. SHIEN u. L. H. LEE
(2011):
Immunoadjuvant activities of a recombinant chicken IL-12 in chickens vaccinated with
Newcastle disease virus recombinant HN protein.
Vet Microbiol 151, 220-228
SUN, J. H., P. LU, Y. X. YAN, X. G. HUA, J. JIANG u. Y. ZHAO (2003):
Sequence and analysis of genomic segment A and B of very virulent infectious bursal
disease virus isolated from China.
J Vet Med B 50, 148-154
SUN, J. H., Y. X. YAN, J. JIANG u. P. LU (2005):
DNA immunization against very virulent infectious bursal disease virus with VP2-4-3 gene
and chicken IL-6 gene.
J Vet Med B Infect Dis Vet Public Health 52, 1-7
SWAYNE, D. E., D. EGGERT u. J. R. BECK (2012):
Reduction of high pathogenicity avian influenza virus in eggs from chickens once or twice
vaccinated with an oil-emulsified inactivated H5 avian influenza vaccine.
Vaccine 30, 4964-4970
TACKEN, M. G. J., B. P. H. PEETERS, A. A. M. THOMAS, P. J. M. ROTTIER u. H. J. BOOT
(2002):
Infectious bursal disease virus capsid protein VP3 interacts both with VP1, the RNAdependent RNA polymerase, and with viral double-stranded RNA.
135
J Virol 76, 11301-11311
TAGHAVI, A., B. ALLAN, G. MUTWIRI, A. VAN KESSEL, P. WILLSON, L. BABIUK, A.
POTTER u. S. GOMIS (2008):
Protection of Neonatal Broiler Chicks Against Salmonella Typhimurium Septicemia by DNA
Containing CpG Motifs.
Avian Dis 52, 398-406
TAN, B., H. WANG, L. SHANG u. T. YANG (2009):
Coadministration of chicken GM-CSF with a DNA vaccine expressing infectious bronchitis
virus (IBV) S1 glycoprotein enhances the specific immune response and protects against IBV
infection.
Arch Virol 154, 1117-1124
TANIMURA, N. u. J. M. SHARMA (1997):
Appearance of T cells in the bursa of Fabricius and cecal tonsils during the acute phase of
infectious bursal disease virus infection in chickens.
Avian Dis 41, 638-645
TANIMURA, N. u. J. M. SHARMA (1998):
In-situ apoptosis in chickens infected with infectious bursal disease virus.
Journal of Comparative Pathology 118, 15-27
TANIMURA, N., K. TSUKAMOTO, K. NAKAMURA, M. NARITA u. M. MAEDA (1995):
Association between Pathogenicity of Infectious Bursal Disease Virus and Viral-Antigen
Distribution Detected by Immunohistochemistry.
Avian Dis 39, 9-20
TARPEY, I., P. J. DAVIS, P. SONDERMEIJER, C. VAN GEFFEN, I. VERSTEGEN, V. E.
SCHIJNS, J. KOLODSICK u. R. SUNDICK (2007a):
Expression of chicken interleukin-2 by turkey herpesvirus increases the immune response
against Marek's disease virus but fails to increase protection against virulent challenge.
Avian Pathol 36, 69-74
TARPEY, I., A. A. VAN LOON, N. DE HAAS, P. J. DAVIS, S. ORBELL, D. CAVANAGH, P.
BRITTON, R. CASAIS, P. SONDERMEIJER u. R. SUNDICK (2007b):
A recombinant turkey herpesvirus expressing chicken interleukin-2 increases the protection
provided by in ovo vaccination with infectious bursal disease and infectious bronchitis virus.
Vaccine 25, 8529-8535
THORNTON, A. M., E. E. DONOVAN, C. A. PICCIRILLO u. E. M. SHEVACH (2004):
Cutting edge: IL-2 is critically required for the in vitro activation of CD4(+)CD25(+) T cell
suppressor function.
J Immunol 172, 6519-6523
TINSLEY-BOWN, A. M., R. FRETWELL, A. B. DOWSETT, S. L. DAVIS u. G. H. FARRAR
(2000):
Formulation of poly(D,L-lactic-co-glycolic acid) microparticles for rapid plasmid DNA delivery.
J Control Release 66, 229-241
TSAI, S. M., H. J. LIU, J. H. SHIEN, L. H. LEE, P. C. CHANG u. C. Y. WANG (2012):
136
Rapid and sensitive detection of infectious bursal disease virus by reverse transcription loopmediated isothermal amplification combined with a lateral flow dipstick.
J Virol Methods 181, 117-124
TSUKAMOTO, K., C. KOJIMA, Y. KOMORI, N. TANIMURA, M. MASE u. S. YAMAGUCHI
(1999):
Protection of chickens against very virulent infectious bursal disease virus (IBDV) and
Marek's disease virus (MDV) with a recombinant MDV expressing IBDV VP2.
Virology 257, 352-362
TSUKAMOTO, K., T. SATO, S. SAITO, N. TANIMURA, N. HAMAZAKI, M. MASE u. S.
YAMAGUCHI (2000):
Dual-viral vector approach induced strong and long-lasting protective immunity against very
virulent infectious bursal disease virus.
Virology 269, 257-267
TSUKAMOTO, K., N. TANIMURA, M. MASE u. K. IMAI (1995):
Comparison of virus replication efficiency in lymphoid tissues among three infectious bursal
disease virus strains.
Avian Dis 39, 844-852
ULMER, J. B., D. L. MONTGOMERY, J. J. DONNELLY u. M. A. LIU (1996):
DNA vaccines.
Methods Mol Med 4, 289-300
VAKHARIA, V. N., J. HE, B. AHAMED u. D. B. SNYDER (1994):
Molecular basis of antigenic variation in infectious bursal disease virus.
Virus Res 31, 265-273
VALLI, A., I. BUSNADIEGO, V. MALIOGKA, D. FERRERO, J. R. CASTON, J. F.
RODRIGUEZ u. J. A. GARCIA (2012):
The VP3 Factor from Viruses of Birnaviridae Family Suppresses RNA Silencing by Binding
Both Long and Small RNA Duplexes.
PLoS One 7, e45957
VAN DEN BERG, T. P., N. ETERRADOSSI, D. TOQUIN u. G. MEULEMANS (2000a):
Infectious bursal disease (Gumboro disease).
Rev Sci Tech 19, 527-543
VAN DEN BERG, T. P., N. ETERRADOSSI, D. TOQUIN u. G. MEULEMANS (2000b):
Infectious bursal disease (Gumboro disease).
Rev Scie Et Tech De L Off International Des Epiz 19, 527-543
VAN DEN BERG, T. P., D. MORALES, N. ETERRADOSSI, G. RIVALLAN, D. TOQUIN, R.
RAUE, K. ZIERENBERG, M. F. ZHANG, Y. P. ZHU, C. Q. WANG, H. J. ZHENG, X. WANG,
G. C. CHEN, B. L. LIM u. H. MÜLLER (2004):
Assessment of genetic, antigenic and pathotypic criteria for the characterization of IBDV
strains.
Avian Pathol 33, 470-476
VAN LOON, A. A. W. M., N. DE HAAS, I. ZEYDA u. E. MUNDT (2002):
137
Alteration of amino acids in VP2 of very virulent infectious bursal disease virus results in
tissue culture adaptation and attenuation in chickens.
J Gen Virol 83, 121-129
VERVELDE, L. u. T. F. DAVISON (1997):
Comparison of the in situ changes in lymphoid cells during infection with infectious bursal
disease virus in chickens of different ages.
Avian Pathol 26, 803-821
VILLEGAS, P., M. HAMOUD, L. B. PURVIS u. F. PEROZO (2008):
Infectious bursal disease subunit vaccination.
Avian Dis 52, 670-674
VOGEL, F. R. (1998):
Adjuvants in perspective.
Dev Biol Stand 92, 241-248
WANG, D., X. JIA, R. SHE u. Y. LIU (2012a):
Acute hypersensitive-like injury in specific-pathogen-free chickens after infection with very
virulent infectious bursal disease virus.
Poult Sci 91, 334-339
WANG, D., X. ZHOU, R. SHE, J. XIONG, Q. SUN, K. PENG, L. LIU u. Y. LIU (2009a):
Impaired intestinal mucosal immunity in specific-pathogen-free chickens after infection with
very virulent infectious bursal disease virus.
Poult Sci 88, 1623-1628
WANG, D. C., J. M. XIONG, R. P. SHE, L. Q. LIU, Y. M. ZHANG, D. M. LUO, W. G. LI, Y. X.
HU, Y. H. WANG, Q. ZHANG u. Q. SUN (2008):
Mast cell mediated inflammatory response in chickens after infection with very virulent
infectious bursal disease virus.
Vet Immunol Immunopathol 124, 19-28
WANG, D. Y., X. R. LI, L. X. XU, Y. L. HU, B. K. ZHANG u. J. G. LIU (2006):
Immunologic synergism with IL-2 and effects of cCHMIs on mRNA expression of IL-2 and
IFN-gamma in chicken peripheral T lymphocyte.
Vaccine 24, 7109-7114
WANG, G., L. PAN, Y. G. ZHANG, Y. L. WANG, Z. W. ZHANG, J. L. LU, P. ZHOU, Y. Z.
FANG u. S. T. JIANG (2011):
Intranasal Delivery of Cationic PLGA Nano/Microparticles- Loaded FMDV DNA Vaccine
Encoding IL-6 Elicited Protective Immunity against FMDV Challenge.
PLoS One 6, e27605
WANG, N. X., D. A. BAZDAR, S. F. SIEG u. H. A. VON RECUM (2012b):
Microparticle delivery of Interleukin-7 to boost T-cell proliferation and survival.
Biotechnol Bioeng 109, 1835-1843
WANG, X., P. JIANG, S. DEEN, J. WU, X. LIU u. J. XU (2003):
Efficacy of DNA vaccines against infectious bursal disease virus in chickens enhanced by
coadministration with CpG oligodeoxynucleotide.
138
Avian Dis 47, 1305-1312
WANG, Y. M., C. Q. SHAN, S. X. MING, Y. LIU, Y. C. DU u. G. T. JIANG (2009b):
Immunoadjuvant effects of bacterial genomic DNA and CpG oligodeoxynucleotides on avian
influenza virus subtype H5N1 inactivated oil emulsion vaccine in chicken.
Res Vet Sci 86, 399-405
WANG, Y. S., H. J. FAN, Y. LI, Z. L. SHI, Y. PAN u. C. P. LU (2007):
Development of a multi-mimotope peptide as a vaccine immunogen for infectious bursal
disease virus.
Vaccine 25, 4447-4455
WANG, Y. S., W. OUYANG, X. J. LIU, K. W. HE, S. Q. YU, H. B. ZHANG, H. J. FAN u. C. P.
LU (2012c):
Virus-like particles of hepatitis B virus core protein containing five mimotopes of infectious
bursal disease virus (IBDV) protect chickens against IBDV.
Vaccine 30, 2125-2130
WATTRANG, E. (2009):
Phosphorothioate oligodeoxyribonucleotides induce in vitro proliferation of chicken B-cells.
Vet Immunol Immunopathol 131, 218-228
WEERATNA, R., C. L. BRAZOLOT MILLAN, A. M. KRIEG u. H. L. DAVIS (1998):
Reduction of antigen expression from DNA vaccines by coadministered
oligodeoxynucleotides.
Antisense Nucleic Acid Drug Dev 8, 351-356
WEI, L., L. HOU, S. S. ZHU, J. WANG, J. ZHOU u. J. LIU (2011):
Infectious bursal disease virus activates the phosphatidylinositol 3-kinase (PI3K)/Akt
signaling pathway by interaction of VP5 protein with the p85 alpha subunit of PI3K.
Virology 417, 211-220
WEI, Y. W., J. R. LI, J. T. ZHENG, H. XU, L. LI u. L. YU (2006):
Genetic reassortment of infectious bursal disease virus in nature.
Biochem Biophys Res Commun 350, 277-287
WEI, Y. W., X. P. YU, J. T. ZHENG, W. Y. CHU, H. XU, X. M. YU u. L. YU (2008):
Reassortant infectious bursal disease virus isolated in China.
Virus Res 131, 279-282
WHITFILL, C. E., E. E. HADDAD, C. A. RICKS, J. K. SKEELES, L. A. NEWBERRY, J. N.
BEASLEY, P. D. ANDREWS, J. A. THOMA u. P. S. WAKENELL (1995):
Determination of optimum formulation of a novel infectious bursal disease virus (IBDV)
vaccine constructed by mixing bursal disease antibody with IBDV.
Avian Dis 39, 687-699
WILKHU, J., S. E. MCNEIL, D. J. KIRBY u. Y. PERRIE (2011):
Formulation design considerations for oral vaccines.
Ther Deliv 2, 1141-1164
WILLIAMS, A. E. u. T. F. DAVISON (2005):
139
Enhanced immunopathology induced by very virulent infectious bursal disease virus.
Avian Pathol 34, 4-14
WINTERFIELD, R. W., A. M. FADLY u. A. BICKFORD (1972):
Infectivity and distribution of infectious bursal disease virus in the chicken. Persistence of the
virus and lesions.
Avian Dis 16, 622-632
WITHERS, D. R., T. F. DAVISON u. J. R. YOUNG (2006):
Diversified bursal medullary B cells survive and expand independently after depletion
following neonatal infectious bursal disease virus infection.
Immunol 117, 558-565
WITHERS, D. R., J. R. YOUNG u. T. F. DAVISON (2005):
Infectious bursal disease virus-induced immunosuppression in the chick is associated with
the presence of undifferentiated follicles in the recovering bursa.
Viral Immunol 18, 127-137
WU, C. C., P. RUBINELLI u. T. L. LIN (2007a):
Molecular detection and differentiation of infectious bursal disease virus.
Avian Dis 51, 515-526
WU, H., N. K. SINGH, R. D. LOCY, K. SCISSUM-GUNN u. J. J. GIAMBRONE (2004):
Expression of immunogenic VP2 protein of infectious bursal disease virus in Arabidopsis
thaliana.
Biotechnol Lett 26, 787-792
WU, J., L. YU, L. LI, J. HU, J. ZHOU u. X. ZHOU (2007b):
Oral immunization with transgenic rice seeds expressing VP2 protein of infectious bursal
disease virus induces protective immune responses in chickens.
Plant Biotechnol J 5, 570-578
WU, Y. P., L. L. HONG, J. X. YE, Z. Y. HUANG u. J. Y. ZHOU (2009):
The VP5 protein of infectious bursal disease virus promotes virion release from infected cells
and is not involved in cell death.
Arch Virol 154, 1873-1882
WU, Z. u. P. KAISER (2011):
Antigen presenting cells in a non-mammalian model system, the chicken.
Immunobiology 216, 1177-1183
WU, Z. G., L. ROTHWELL, J. R. YOUNG, J. KAUFMAN, C. BUTTER u. P. KAISER (2010):
Generation and characterization of chicken bone marrow-derived dendritic cells.
Immunology 129, 133-145
XIA, R. X., H. Y. WANG, G. M. HUANG u. M. F. ZHANG (2008):
Sequence and phylogenetic analysis of a Chinese very virulent infectious bursal disease
virus.
Arch Virol 153, 1725-1729
XIAOWEN, Z., Y. QINGHUA, Z. XIAOFEI u. Y. QIAN (2009):
140
Co-administration of inactivated avian influenza virus with CpG or rIL-2 strongly enhances
the local immune response after intranasal immunization in chicken.
Vaccine 27, 5628-5632
XU, H. T., W. D. SI u. P. DOBOS (2004):
Mapping the site of guanylylation on VP1, the protein primer for infectious pancreatic
necrosis virus RNA synthesis.
Virology 322, 199-210
XU, J. T., Z. M. ZHANG, Y. B. YIN, S. J. CUI, S. Z. XU, Y. Y. GUO, J. D. LI, J. L. WANG, X.
C. LIU u. L. M. HAN (2009):
Development of reverse-transcription loop-mediated isothermal amplification for the detection
of infectious bursal disease virus.
J Virol Methods 162, 267-271
XU, Y. H. u. F. C. SZOKA (1996):
Mechanism of DNA release from cationic liposome/DNA complexes used in cell transfection.
Biochemistry 35, 5616-5623
YAGER, E. J., H. J. DEAN u. D. H. FULLER (2009):
Prospects for developing an effective particle-mediated DNA vaccine against influenza.
Expert Rev vaccines 8, 1205-1220
YAMAGUCHI, T., T. KONDO, Y. INOSHIMA, M. OGAWA, M. MIYOSHI, T. YANAI, T.
MASEGI, H. FUKUSHI u. K. HIRAI (1996a):
In vitro attenuation of highly virulent infectious bursal disease virus: Some characteristics of
attenuated strains.
Avian Dis 40, 501-509
YAMAGUCHI, T., M. OGAWA, Y. INOSHIMA, M. MIYOSHI, H. FUKUSHI u. K. HIRAI
(1996b):
Identification of sequence changes responsible for the attenuation of highly virulent infectious
bursal disease virus.
Virology 223, 219-223
YAMAGUCHI, T., M. OGAWA, M. MIYOSHI, Y. INOSHIMA, H. FUKUSHI u. K. HIRAI
(1997a):
Sequence and phylogenetic analyses of highly virulent infectious bursal disease virus.
Arch Virol 142, 1441-1458
YAMAGUCHI, T., M. OGAWA, M. MIYOSHI, Y. INOSHIMA, H. FUKUSHI u. K. HIRAI
(1997b):
Sequence and phylogenetic analyses of highly virulent infectious bursal disease virus.
Arch Virol 142, 1441-1458
YAMAGUCHI, T., A. SETIYONO, M. KOBAYASHI, S. TAKIGAMI, H. FUKUSHI u. K. HIRAI
(2000):
Infectious bursal disease live vaccines: Changes in the virus population during serial
passage in chickens and chicken embryo fibroblast cells.
Avian Diseases 44, 284-290
141
YANG, Y. F., D. LEGGAT, A. HERBERT, P. C. ROBERTS u. R. S. SUNDICK (2009):
A Novel Method to Incorporate Bioactive Cytokines as Adjuvants on the Surface of Virus
Particles.
J Interferon Cytokine Res 29, 9-22
YAO, K. u. V. N. VAKHARIA (2001):
Induction of apoptosis in vitro by the 17-kDa nonstructural protein of infectious bursal disease
virus: possible role in viral pathogenesis.
Virology 285, 50-58
YEH, H. Y., S. RAUTENSCHLEIN u. J. M. SHARMA (2002):
Protective immunity against infectious bursal disease virus in chickens in the absence of
virus-specific antibodies.
Vet Immunol Immunopathol 89, 149-158
YIP, C. W., C. C. HON, F. Y. ZENG u. F. C. C. LEUNG (2012):
Cell culture-adapted IBDV uses endocytosis for entry in DF-1 chicken embryonic fibroblasts.
Virus Res 165, 9-16
YU, F., X. L. QI, Y. Q. YUWEN, Y. Q. WANG, H. L. GAO, Y. L. GAO, L. T. QIN u. X. M.
WANG (2010):
Molecular characteristics of segment B of seven very virulent infectious bursal disease
viruses isolated in china.
Virus Genes 41, 246-249
YUAN, W. F., X. Y. ZHANG, X. L. XIA u. H. C. SUN (2012):
Inhibition of infectious bursal disease virus infection by artificial microRNAs targeting chicken
heat-shock protein 90.
J Gen Virol 93, 876-879
ZANETTI, F. A., M. P. D. ZAJAC, O. A. TABOGA u. G. CALAMANTE (2012):
Evaluation of modified vaccinia virus Ankara expressing VP2 protein of infectious bursal
disease virus as an immunogen in chickens.
J Vet Sci 13, 199-201
ZESHAN, B., M. H. MUSHTAQ, X. WANG, W. LI u. P. JIANG (2011):
Protective immune responses induced by in ovo immunization with recombinant
adenoviruses expressing spike (S1) glycoprotein of infectious bronchitis virus fused/coadministered with granulocyte-macrophage colony stimulating factor.
Vet Microbiol 148, 8-17
ZHANG, W. L., Z. YIN, N. H. LIU, T. YANG, J. W. WANG, Z. G. BU u. D. L. WU (2010):
DNA-chitosan nanoparticles improve DNA vaccine-elicited immunity against Newcastle
disease virus through shuttling chicken interleukin-2 gene.
J Microencapsul 27, 693-702
ZHANG, X. Q., J. INTRA u. A. K. SALEM (2008):
Comparative study of poly (lactic-co-glycolic acid)-poly ethyleneimine-plasmid DNA
microparticles prepared using double emulsion methods.
J Microencapsul 25, 1-12
142
ZIERENBERG, K., H. NIEPER, T. P. VAN DEN BERG, C. D. EZEOKOLI, M. VOSS u. H.
MULLER (2000):
The VP2 variable region of African and German isolates of infectious bursal disease virus:
comparison with very virulent, "classical" virulent, and attenuated tissue culture-adapted
strains.
Arch Virol 145, 113-125
ZIERENBERG, K., R. RAUE, H. NIEPER, M. R. ISLAM, N. ETERRADOSSI, D. TOQUIN u.
H. MULLER (2004):
Generation of serotype 1/serotype 2 reassortant viruses of the infectious bursal disease virus
and their investigation in vitro and in vivo.
Virus Res 105, 23-34
143
10. Acknowledgements
The dedication and expertise given by my supervisor, Prof. Dr. Silke Rautenschlein,
from the inception of the PhD research to the end of my PhD was amazing. I thank
you very much for your encouragements in all circumstances and for the livable
guidance to complete my work.
The PhD supervisory committee members, Prof. Dr. Beatrice Grummer and PD Dr.
Reimar Johne, without your continual advise and support, my PhD would not have
been materialized.
The
role
of
Hannover
Graduate
School
for
Veterinary
Pathobiology,
Neuroinfectiology, and Translational Medicine (HGNI), University of Veterinary
Medicine Hanover (TiHo), Germany, was also critical for my PhD work mainly
through coordination, financing all my international conference fees and for providing
an additional 3 months financial support.
My very heartfelt acknowledgment directly goes to the German Academic Exchange
Service (DAAD) for the provision of the scholarship, family allowances and for
unreservedly providing all the necessary information during my PhD study.
I express my appreciation to the PhD secretaries for providing valuable ideas during
the PhD course work. Mrs. Maritta Ledwoch, from the International Office, you are so
wonderful and I always feel comfortable when I visit your office and for the nice
extracurricular dialogue which we had in the past.
Likewise, the University of Gondar, Ethiopia is highly acknowledged for granting me a
PhD study leave and the Institute of Biodiversity Conservation and Research for
handling my sample export permit favorably. The National Veterinary Institute,
Ethiopia equally shares my thanks for providing me some of the samples for my PhD
research.
144
My very special thanks to Christine Haase for her excellent technical support; Sonja
Bernhard, Sabrina Techel and Katja Stolpe for their excellent help in animal care and
necropsy. The institute for parasitology, TiHo also deserves my thanks for allowing
me to use some of the laboratory facilities.
I am very pleased and would like to take this opportunity to pay tribute to every single
member of the staff at the Clinic for Poultry, who have played a part in my research. I
do not have words to express my thanks to Lydia Teske for your kindness and
support in translating the summary part of my thesis into the German version. Dr.
Henning Petersen, from the Clinic for Poultry, I thank you very much for structuring
and formatting my thesis.
Dr. Awoke, Mohammed Yusuf Adem and Bidir, thank you for handling my animal
experiment at Gondar while I was processing my visa.
No amounts of words express my love and respect to my wife, Eskedar Hailegebral,
for providing me a lesson of patience in my life, also for sharing your mature political
thoughts and for your untiring responsibility in managing our family. I know that it is
stressful to handle the two very little kids. Eskedar you suffered, but never get
disappointed we are decorated with the two blessings. In the front line of my PhD
work is my lovely daughter Heldana, she is the fortitude of my life.