Scientific Research and Essay Vol. 4 (2), pp. 042-058, February, 2009
Available online at http://www.academicjournals.org/SRE
ISSN 1992-2248 © 2009 Academic Journals
Review
Oxidative stress and coxsackievirus infections as
mediators of beta cell damage: A review
Vibha Sharma
1
, Shahina Kalim
2
, Manoj K. Srivastava
2
, Surabhi Nanda
1
and Sanjay Mishra
1,3 *
1
Department of Biotechnology College of Egineering and Technology, IFTM campus, Delhi Road, Moradabad 244001,
U.P., India.
2
Department of Biochemistry, Bundelkhand University, Jhansi, U.P., India.
3
Department of Biotechnology and Microbiology, College of Applied Sciences, IFTM Campus, Lodhipur-Rajput, Delhi
Road, Moradabad 244 001, U.P., India.
Accepted 11 December, 2008
Type I diabetes (T1D) is an autoimmune disease resulting in gradual cell-mediated destruction of insulin
producing beta cells in the pancreatic islets of Langerhans. The most plausible environmental triggers
to launch and/or accelerate this process in a genetically predisposed organism including enterovirus
infections and oxidative stress. Among other enteroviruses the group B of coxsackieviruses is
associated with potential beta cell toxicity. Beta cells are weak in antioxidative defense, which makes
them hypersensitive to oxidative stress. Acknowledging the inhibitory potential of in vivo conditions
scientists have developed two models resembling a slowly progressing coxsackievirus infection first,
by restricting the production of viral progeny with a selective inhibitor of viral RNA replication and
second, by means of lowering the multiplicity of infection. Hydrogen peroxide has been established as
the oxidative stressor. Recent studies reflect that a productive CVB-infection results in lytic beta cell
death. When pharmacologically restricted by guanidine-HCl, the viability increases dramatically through
decreased necrosis and associates with simultaneous stimulation of apoptotic death. In summary, the
review introduces potential mechanistic models for enterovirus infections in beta cells.
Key words: Oxidative stress, type 1 diabetes, coxsackievirus infection, beta cells, human leukocyte antigen.
INTRODUCTION
T1D is considered to represent a multifactorial disorder
based on genetic predisposition triggered by an environ-
mental factor e.g. virus infection. The effector stage,
insulitis, is characterized by gradual invasion of macro-
phages, T-cells and the produced cytokines and oxidative
radicals into the islets (Rasilainen et al., 2004). The pre-
*Corresponding author. E-mail: [email protected].
Abbreviations: APC, antigen presenting cell; CAR,
coxackie/adenoviral receptor;
CVA, coxackievirus A strain; GAD, glutamic acid decarboxylase;
GPx, glutathione peroxidase; HLA, human leukocyte antigen;
IFN, interferon; IL-1, interleukin 1 beta; MHC, major histocom-
patibility complex; NF-B, nuclear factor-kappa B transcription
factor; NO, nitric oxide; NOD mouse, spontaneously diabetic
non-obese; ROS, reactive oxygen species; SOD, super oxide
dismutase; TNF – , tumor necrosis factor alpha.
diabetic period is usually long and characterized by
formation of islet cell autoantibodies even years before
the development of diabetic symptoms. Enterovirus infec-
tions are epidemiologically linked to the pathogenesis of
T1D (Rasilainen et al., 2004). Beta cells exhibit poor
intracellular antioxidative capacity, which renders them
vulnerable to oxidative stress. Thus, both enterovirus
infections and oxdative stress are considered to repre-
sent particularly potential triggers and/or accelerators of
beta cell destruction.
It is generally known that the amount of viral particles
needed to establish a local infection is low and all viruses
invading a physiological environment face natural resis-
tance by factors of the hosting immune system (Rasilai-
nen et al., 2002). Considering that coxsackievirus infec-
tions most often remain subclinical in vivo, the prog-
ression of infection and amount of virus are possibly limi-
ted and submaximal (Roivainen et al., 2000). The pre-
sent review is a comprehension of information on experi-
mental models of slowly progressing enterovirus infection
and reports regarding the patterns of mechanisms of
virus induced beta cell damage. Further, this scientific re-
port includes: characterizing the mechanisms of oxidative
beta cell injury; counteracting it by means of antioxidative
agents and studying the effects of virus infection on
intracellular redox balance.
Background of type I diabetes (T1D)
Juvenile, insulin-dependent diabetes mellitus (Type I
diabetes, T1D) is a consequence of selective autoim-
mune destruction of the insulin-producing beta cells in the
pancreatic islets of Langerhans (Bach, 1994; Schranz
and Lernmark, 1998). The resulting insulin deficiency
leads to a chronic metabolic derangement associated
with significant secondary morbidity. According to the tra-
ditional opinion consistent within most autoimmune
diseases, both a genetically predisposed individual and a
suitable environmental trigger are needed for the des-
tructive process, in the case of T1D against beta cells, to
begin. Nowadays, the group of human leukocyte antigen
(HLA) genes is considered to account for the major ge-
netic risk and virus infections to represent prime environ-
mental regulator. Various lines of evidence suggest that
the autoimmune destruction is launched by a local insult
to islet(s) exciting a pool of immune cells, dominantly T-
cells and macrophages, to invade the islets resulting in
insulitis (Anastasi et al., 2005) Considering the pathogen-
nesis, most results are obtained from studies in sponta-
neously diabetic non-obese (NOD) mice, whose
pathogen-nesis mimics human type I diabetes (Anita et
al., 2006). Most probably both cytotoxic (CD8+) and
helper (CD4+) T-cells are required for both the primary
attack and the later overt destruction, respectively, to pro-
gress (Bendelac et al., 1987; Hanafusa et al., 1988;
Jarpe et al., 1990; Jarpe et al., 1991; Miller et al., 1988;
Sumida et al., 1994; Wicker et al., 1994; Wong et al.,
1996). Besides, a pool of antigen presenting cells (APC)
is indispensable in the vicinity of pancreatic islets to intro-
duce the islet antigen to the autoreactive T-cells. Accord-
ing to current opinion, programmed cell death (apoptosis)
represents the dominant mechanism of beta cell death
during immune mediated T1D. The immune reaction rela-
ted proinflammatory cytokines, especially interferon gam-
ma (IFN-), interleukine 1 beta (IL-1) and tumor necrosis
factor alpha (TNF-), are thought to play a primary role in
activating signal transduction pathways leading to beta
cell dysfunction and apoptosis (Delaney et al., 1997;
Mandrup, 1996; Rabinovitch, 1998). The direct killing
through the granzyme-perforin pathway by cytotoxic T
cells is also believed to be a key mechanism ultimately
leading to apoptotic beta cell death.
Finland has the highest incidence of T1D in under 15-
year-old children worldwide. It has gradually increased
during the last 43 years from 12/100000 (1953) to
45/100000 (1996); the average relative annual increase
Sharma et al. 043
thus being 3.4 % between 1965 and 1996 (Somersalo,
1955; Tuomilehto et al., 1999). The incidence is not only
geographically, but also seasonally distributed peaking in
winter (Levy-Marchal et al., 1995). Finland and some
other Nordic high-incidence countries have also shown
gender-impact with a slight male predominance (Padaiga
et al., 1997). Approximately 10% of newly diagnosed T1D
patients have an affected first degree relative (Dahlquist
et al., 1985; Tuomilehto et al., 1992). The genetic risk
and familial clustering of T1D have traditionally been con-
cidered to originate from the class II major histocompati-
bility complex (MHC) genes on chromosome 6p21,
including HLA DP, DQ and DR (Campbell and Trowsdale,
1993; Davies et al., 1994;). These genes encode hetero-
dimeric proteins expressed on APC-cells and have a
function in presenting antigenic peptides to CD4+ T cells.
HLA class II genes have a high degree of polymorphism
and marked linkage disequilibrium that potentiate the risk
to form certain high-risk allele-pairs. Furthermore, these
susceptibility alleles may then choose potential autoreac-
tive T-cells during thymic sorting, and enhance the dis-
ease probability. Many recent studies have demonstrated
the DQ-locus to primarily harbor the susceptibility for T1D
(Dorman and Bunker, 2000; Heimberg et al., 1992;
Thorsby and Ronningen, 1993). Specifically, worldwide
approximately 30% of T1D patients are heterozygous for
the high risk HLA-DQA1*0501-DQB1*0201 / DQA1*0301-
DQB1*0302 alleles (previously referred to as HLA-DR3/4
or HLADQ2/ DQ8). In addition lack of the protective HLA-
DQA1*0102-DQB1*0602 allele (HLADR2 or HLA-DQ6)
increases T1D susceptibility through potentiating the
impacts of predisposing alleles (Pociot and McDermott,
2002). Genes present in the HLA-region, including genes
worthy for structure and function of both the HLA-DQ and
DR, are considered capable of accounting for less than
50% of the inherited disease risk (Pociot and McDermott,
2002). Largely by means of genome 13 scan dependent
linkage studies of affected sib-pair families at least 17
non-HLA T1D susceptibility loci have also been identified
(Pociot and McDermott, 2002). Among these, the insulin
gene locus (IDDM2) is the strongest and most consistent
(Bennett et al., 1995).
Necrosis of cell death and their association to T1D
Various routes and mechanisms have been described for
a cell to die. The specific pathway chosen depends large-
ly on the trigger, but the ultimate result may vary accord-
ing to the cellular capacity of defense and the net status
of the host (age, inflammation etc.). A general divi- sion is
traditionally done by the resulting morphological charac-
teristics: apoptotic, pyknotic, necrotic (Figure 1).
Necrotic cell death is an accidental, traumatic event
always associated with specific pathology. It is an unre-
gulated process and follows no specific pattern. The sa-
lient features include cytoplasmic and mitochondrial
swelling and breakdown of cell to cell junctions and com-
044 Sci. Res. Essays
Figure 1. Illustration of nuclear morphology associated with different forms of cell
death.
munication that lead to early rupture of cellular mem-
branes and thereafter leakage of the contents to the
extracellular space provoking inflammation (Buja et al.,
1993; Lemasters, 1999). Considering beta cells, compre-
hensive studies on cytokine-induced stress in rat islets
have revealed nitric oxide (NO) to mediate functional de-
terioration and mostly necrotic beta cell death (Hoorens
et al., 2001). More recently, necrotic beta cell damage
was shown to dominate the pattern of spontaneous, NO-
induced and streptozotocin (STZ)-induced diabetes in
diabetes prone BB rats (Fehsel et al., 2003). Particularly
considering cell culture models, environmental factors
such as high cell density may also lead the outcome of
cell death towards necrosis.
The term apoptosis descends from the Greek words:
'apo' (away) + 'ptosis' (drop) and characterizes a cellular
suicide (Ueda and Shah, 1994). Apoptosis is a physiolo-
gical, energy-requiring, orderly phenomenon used by
multicellular organisms to control the size of cell popula-
tions and to eliminate damaged, infected or mutated
cells. It is thus important for proper embryonic morpho-
genesis, development, and adult cellular homeostasis.
The apoptotic machinery is constitutively expressed in all
nucleated animal cells and is strictly regulated by a set of
evolutionarily conserved genes (Horvitz, 1999; Vaux et
al., 1992). The apoptotic process can be divided in three
stages: a) initiation: a cell receives an apoptotic stimulus;
b) execution: enzymatic events, and 3) degeneration:
disintergration and elimination of the cell. In generally
apoptotic stimuli may activate two different pathways: an
extrinsic death-receptor pathway or an intrinsic pathway
based on mitochondrial dysfunction (Hengartner, 2000).
The extrinsic trail is launched by triggering of a death
receptor on cell surface, the most common of which
belong to the family of tumor necrosis factor receptors
(TNF-R) (Armitage, 1994; Baker and Reddy, 1998).
These receptors include a death domain (Tartaglia et al.,
1993), which is responsible for the activation of the down-
stream cascade: the employment of intracellular adapter
proteins (FADD, TRADD, RAIDD) and the following acti-
vation of caspase (cysteine aspartic acidspecific pro-
teases) 8 and possibly caspase 2 (Alnemri et al., 1996;
Ashkenazi and Dixit, 1998). T-cells use alternative extrin-
sic routes by activating either Fas (discussed later) or
granzyme B leading to stimulation of either caspase 8 or
the intrinsic pathway (Pinkoski et al, 2001). Commonly
the intrinsic route is activated by stress factors including
DNA damage, cell cycle perturbation or growth factor
deprivation. They engage pro-apoptotic members of the
Bcl-2-family (Bak, Bax) to translocate to mitochondria and
form pores to permit cytochrome c to release from the
mitochondrial intermembrane space and to compose the
apoptosome with Apaf-1 and procaspase 9 (Rodriguez
and Lazebnik, 1999; Zou et al., 1999;). Recently, a new
mitochondrial flavoprotein (apoptosis inducing factor; AIF)
was observed to translocate to nucleus in response to
e.g. oxidative apoptotic stimuli and launch caspase- (in)
dependent chromatin condensation and DNA cleavage
(Susin et al., 1999). At the last, both routes activate the
effector caspases (3, 6 and 7) that proteolytically activate
other downstream caspases (Nicholson and Thornberry,
1997; Slee et al., 2001). Caspase 3 activates DFF/CAD,
which results in DNA cleavage to n x 180 bp fragments
(Enari et al., 1998; Liu et al., 1997). This step is called
'point of no return', as the proteolytic cascade irreversibly
results in to cellular collapse featured by the character-
ristic morphology: cell shrinkage, chromatin conden-
sation, blebbing of the plasma membrane and formation
of apoptotic bodies that become eliminated by neigh-
boring phagocytes (Kerr et al., 1972; Wyllie, 1981).
Apoptotic death mediated by immune effector cells is
considered to dominate beta cell death. Several mole-
cular pathways may be activated for this outcome to
occur. Macrophages and activated T-cells produce and
secrete proinflammatory cytokines, IL-6, IL-2, IL-10, IL-
1, TNF- and IFN-, the last three of which are known
for their potential to induce functional and structural beta
cell damage and to provoke apoptosis (Liu et al., 2000a;
Marselli et al., 2000; Saldeen, 2000; Zumsteg et al.,
2000). A recent microarray study on cytokine-stressed rat
islets revealed the transcription factor NF-B to be a core
regulator of cytokine induced signaling pathways (Cardozo
et al., 2001). By inhibiting its activity the cytokine induced
beta cell apoptosis could be prevented (Heimberg et al.,
2001). Attracted T-cells express Fas-ligand (FasL) on cell
surface, which by binding to the target cell Fas receptor
results in apoptosis (Kagi et al., 1994). In spontaneously
diabetic NOD mice the proinflammatory cytokines may
induce Fas expression in beta cells and thus contribute to
beta cell death through interaction with FasL on the sur-
Sharma et al. 045
Table 1. Species and subgroup division of enteroviruses.
Species Subgroup and species
Poliovirus Polioviruses 1-3
Human enterovirus A coxsackievirus A 2-8, 10, 12, 14, 16
Human enterovirus B coxsackievirus A9
coxsackievirus B 1-6
echovirus 1-7, 9, 11-21, 24-27, 29-33
echovirus 69
Human enterovirus C coxsackievirus A 1, 11, 13, 15, 17-22, 24
Human enterovirus D echovirus 68, 70
face of the neighboring CD4+ T-cells (Augstein et al.,
2003; Nakayama et al., 2002; Petrovsky et al., 2002;
Suarez-Pinzon et al., 1999). Also human beta cells in
pancreatic biopsies from newly diagnosed diabetics have
been reported to express Fas and to contain apoptotic
beta cells (Moriwaki et al., 1999). As mentioned earlier
CD8+ T-cells, suggested to have a critical role in the
early stages of beta cell destruction, may mediate apop-
tosis also through the perforin-granzyme pathway, de-
pendent on perforin-built channels on target cell mem-
brane allowing granzyme entry and caspase activation
(Garcia-Sanz et al., 1987; Kagi et al., 1994; Yoon and
Jun, 1999).
Beside the well-known apoptotic and necrotic path-
ways, a pyknotic form of cell death has also been differ-
rentiated. It morphologically resembles apoptosis with an
intact plasma membrane, but is characterized by con-
densed but intact chromatin, unlike the apoptotic frag-
mented DNA, demarcated at the margins of the nucleus
(Agol et al., 1998; Tolskaya et al., 1995). Pyknotic mor-
phology is typically seen in the early phase of a cytocidal
virus infection when the first virus-induced alterations in
host cell membrane organization and cytoskeleton have
on set (Koch and Koch, 1985). This change is often ire-
versible and results in lytic cell death. A comparable
phenomenon has also been observed during an apoptotic
process, which due to a sudden ATP depletion or some
other disability was interrupted and turned into secondary
necrosis (Hirsch et al., 1997; McCarthy et al., 1997).
Enteroviruses
Picornaviruses are a diverse group of small, non-enve-
loped animal viruses with single-stranded RNA genome
of positive polarity ('pico' Greek: very small – RNA vi-
ruses) (Racaniello, 2001). The first infections, later pin-
pointed to have been caused by a specific picorna-virus,
were described over 3000 years ago (poliomyelitis in a
temple record from Egypt ca. 1400 BC). Nowadays they
comprise the most common infections of humans in the
developed world (Rotbart, 2002). Picornaviruses were
previously classified according to their physicochemical
properties (particle density, pH-sensitivity) and serolo-
gical relatedness. More recently, the classification has
been based on nucleotide sequence comparisons. The
picornavirus genera infecting man include enteroviruses,
hepatoviruses, kobuviruses, parechoviruses and rhinovi-
ruses. The enterovirus genus is further divided into five
species, which are listed in Table 1 together with the pre-
vious subgroups (based on pathogenesis in experi-
mental animals) and serotypes (Hyypiä et al., 1997; King
et al., 1999).
Enteroviral genome and viral replication cycle
In order to multiply, the virus must first bind to its specific
receptor on the surface of a target cell. Several receptors
have been demonstrated for enteroviruses: coxsackiea-
denovirus receptor (CAR) is recognized by CVB 1-6 and
decay-accelerating factor (DAF, CD55) by CVB 1, 3, 5
(Bergelson et al., 1995, 1997, 1998). In addition, CVA 9
has been shown to use alpha v beta 3 integrin, the vitro-
nectin receptor, to penetrate host cells (Roivainen et al.,
1994). In studies using neutralizing antibodies, both alpha
and beta cells in human islets were observed to express
CAR (Chehadeh et al., 2000). After contacting the recep-
tor, the virus should successfully enter the cell, which
commonly requires some conformational changes of the
viral capsid. During entry or immediately afterwards, the
viral genomic material is released into the host’s cyto-
plasm in a process called uncoating (Figure 2). Once
released in the cytoplasm, the viral genomic RNA, about
7500 nucleotides in length, functions directly as a tem-
plate for the production of viral proteins. The single open
reading frame encodes a large precursor poly-protein.
This precursor includes two proteolytic sequences (2A
and 3C), which cleave the precursor into intermediate
products P1, P2 and P3 (Kitamura et al., 1981). P1 is fur-
ther digested into capsid (structural) proteins VP0, VP3
and VP1 (Korant, 1973). P2 and P3 are in the sequential
steps cleaved into seven non-structural proteins func-
tioning in the replication and encapsidation of the viral
RNA and in the processing of the proteins. One specific
product is the virus RNA-dependent RNA polymerase
(3D). It duplicate the genomic RNA into a negative sense
strand, which then acts as a template for positive -strand
046 Sci. Res. Essays
Figure 2. The replication cycle of enteroviruses.
RNA synthesis. Finally, the newly pro-duced genomic
material is packaged into the procapsids formed by
twelve pentameric structures, each of which contains five
protomers (VP0-VP3-VP1 –heterotrimers) (Korant, 1973).
During this RNA-encapsidation, the VP0-protein is further
cleaved into VP2 and VP4 (maturation cleavage; relevant
for enteroviruses, but not for all picornaviruses (Hyypiä et
al., 1992; Yamashita et al., 1998)) stabilizing the pro-
capsid.
Host Vs virus infection
As suggested earlier, the outcome of an enterovirus
infection may vary between aggressive fulminant and
acute subclinical (Ramsingh, 1997b; Yoon et al., 1979).
In suitable conditions persistence may develop (Adachi et
al., 1996; Chehadeh et al., 2000; Conaldi et al., 1997a;
1997b Klingel et al., 1992; Tam and Messner, 1999). In
general, a viral invasion into an organism supplied with a
properly functioning immune system always attracts it to
respond. The innate arm of immunity composed of natu-
ral killer cells, natural (nonspecific) antibodies, the com-
plement system, phagocytes, antimicrobial peptides and
interferons (IFNs) together constitute the first line of
defense. IFNs may mediate the antiviral response
through many routes. Firstly, through binding to type I
IFN receptor (Platanias et al., 1996) they stimulate pro-
tein tyrosine kinases (JAK 1, TYK2) which further act to
stimulate STAT-, Crk- and IRS- pathways (Ahmad et al.,
1997; Darnell et al., 1994; Uddin et al., 2000). These
routes finally lead to the induction of anti-viral interferon
stimulated genes (ISG) (Bose and Banerjee, 2003).
Interferons are also able to induce the production of
MHC-molecules on virus-infected cells and on antigen
presenting cells required for CD4+ T-cell action. Further-
more, they may initiate or contribute to the production of
antigen specific antibodies and thus activate the second
line of antiviral host defense, the adaptive or cell-media-
ted arm of immunity (Kadowaki et al., 2000). Recent
observation in human islets concluded CVB-infection to
stimulate INF- production selectively in beta cells, which
by inhibiting the efficiency of viral replication prevents the
infection from disseminating and maintains its persis-
tence (Chehadeh et al., 2000). Furthermore, infection–
induced IFN- has been speculated to participate in the
initiation of beta cell autoimmunity (Chakrabarti
et al., 1996; Chehadeh et al., 2000; Stewart et al., 1993).
It was recently found that efforts of the host to inhibit viral
dissemination and persistence often involve induction of
apoptosis through either perforin-granzyme or death
receptor mediated pathways (Ashkenazi and Dixit, 1998;
Froelich et al., 1998). IFNs are reported to sensitize infec-
ted cells to apoptosis (Balachandran et al., 2000; Tanaka
et al., 1998). As and when beta cells are stimulated with
IFNs in combination with double-stranded viral RNA or
the cytokine IL-1ß, a synergistic apoptotic effect is
observed (Liu et al., 2002). To defend them-selves, many
viruses have evolved to counteract the host responses.
To secure the replicative potential and production of pro-
geny virions, it is favorable for the virus to maintain the
host cell viable and for this purpose many routes of the
host signaling machinery may be manipulated. For
example, various steps of the IFN cascade may be block-
ed by several different viruses (Katze, 1995; Komatsu et
al., 2000; Munoz-Jordan et al., 2003; Ronco et al., 1998).
Also cleavage of p21(ras) GTPase-activating protein
(RasGAP) and activation of the MAPK family members
ERK 1/2, registered essential for CVB3 replication, could
be affected (Huber et al., 1999a; Luo et al., 2002; Opav-
sky et al., 2002). After stimulation of the innate arm of
immunity a cell-mediated immune response is launched.
The dendritic cells, stimulated during the innate res-
ponse by cytokines and the pathogen itself, present anti-
gen to naive T-cells, which start to develop differential
markers. Specifically the Th1 type CD4+ T-cells secrete
proinflammatory cytokines, which enhance virus-specific
host cell lysis by CD8+ T-cells and stimulate the MCH
upregulation on APC cells, that further helps in activating
antibody production by B-cells. Other interleukines
secreted by Th2 type CD4+ T-cells promote this B-cell
activation (Salusto et al., 1998). The host responsiveness
to virus and the outcome of the infection have been
observed to furthermore depend on the status of the host
cell cycle. Several studies have revealed actively dividing
cells, like T-cells, to be more susceptible to virus and to
efficiently produce viral progeny (Liu et al., 2000b; Molina
et al., 1992). Quiescent cells (in G0 phase) on the other
hand do not support viral multiplication, but may harbor
infective viruses thus creating persistence (Feuer et al.,
2002). These theories may partially explain the individual
susceptibility of cells to infection. Some studies further
indicate several viruses capable of forcing the host cell
into a favored phase of cell cycle (Op De Beeck and
Caillet-Fauquet, 1997; Swanton and Jones, 2001). Viral
mRNA, containing an internal ribosome entry site (IRES)
instead of the eukaryotic 5' cap, may also redirect the cel-
lular translation pattern and favor viral protein synthesis,
particularly in enterovirus-infected cells where cap-de-
pendent translation is specifically inhibited (Kuyumcu-
Martinez et al., 2002; Marissen et al., 2000).
Infections and clinical evidence
Enteroviruses are predominantly transmitted by the fecal-
oral route into the human body, although transmission via
either upper respiratory tract or the conjunctiva of the eye
is also possible. They normally replicate in the respiratory
and gastrointestinal mucosa, where the infection may
remain subclinical or result in mild symptoms. In a pro-
portion of cases, the virus spreads through the lympha-
tics into the circulation, and after a brief viraemic phase
may establish secondary replication sites in specific
tissues and organs. Polioviruses may proverbially enter
the central nervous system, replicate in the motor neu-
rons and in about 1% of cases with the most virulent
strains result in flaccid paralysis (Melnick, 1996). Coxsac-
kie A viruses typically induce diseases with mucosal and
skin lesions (e.g. herpangina, hand, foot, and mouth dis-
ease) (Bendig and Fleming, 1996; Itagaki et al., 1983;
Seddon and Duff, 1971). In addition to the milder dis-
orders, coxsackie B viruses are also associated with
more severe and possibly chronic diseases including
meningitis, myopericarditis, epidemic pleurodynia and
T1D (Beck et al., 1990; Gauntt and Huber, 2003; Muir
and van Loon, 1997; Rotbart, 1995; Tracy et al., 2000).
They are together with echoviruses responsible for se-
vere neonatal viral infections (Chiou et al., 1998; Sawyer,
1999).
Sharma et al. 047
Enteroviruses Vs T1D
Enterovirus infections could initiate or accelerate beta cell
damage many years before the clinical onset of T1D.
Seroepidemiological studies exist to support this at least
after early CVB and rubella exposures (Hyöty et al.,
1995; Lönnrot et al., 1998, 2000; Salminen et al., 2003).
Some mini case reports show evidence for intrauterine
viral exposure, a maternal enterovirus infection, as a risk
factor for future development of T1D (Dahlquist et al.,
1995, Hyöty et al., 1995). In a larger birth cohort study
(Funchtenbusch et al., 2001) contradictory results were
obtained. A recent survey reveals that while the incidence
of T1D has doubled within the last 40 years in Finland,
the incidence of enterovirus infections has simultaneously
decreased (Karvonen et al., 1999; Viskari et al., 2002).
Despite the fact that the maternal enterovirus antibody
levels have decreased (Viskari et al., 2002), the low
maternal antibody status could boost the viral exposure
of the fetus/newborn and the rarer incidence furthermore
exposes the child to catch the infection later, often in the
absence of protecting maternal antibodies.
This peculiar phenomenon could partially explain the
increase-ing number of patients with type I diabetes and
supports the role of enteroviruses as possible triggers.
Coxsackie B viruses have long been associated with the
pathogenesis of T1D (Andreoletti et al., 1997; Banatvala
et al., 1985; Yoon, 1990). Groundbreaking findings were
made in 1979. When the autopsy specimens of a 10-
year-old patient with diabetes showed lymphocytic infiltra-
tion of the islets and beta cell necrosis which lead to the
detection of a diabetogenic variant of CVB-4 from the
pancreatic cultures (Yoon et al., 1979). This very sero-
type was further observed to induce T1D in mouse. How-
ever, in a series of 88 pancreas-specimens from patients
who died soon after onset of T1D, coxsackie-viruses
were not detected in the islets (Foulis et al., 1990).
Therefore, it should be considered unlikely that an acute
CVB infection in pancreatic islets would frequently asso-
ciate with the onset of T1D. However, this does not
exclude the possibility that CVB could have been present
in the islets at an earlier stage of T1D pathogenesis. In
recent seroepidemiological studies several enterovirus
serotypes have been shown to share the association with
T1D (Frisk et al., 1992; Helfand et al., 1995; Roivainen et
al.,1998; Otonkoski 2000; Vreugdenhil et al., 2000). In
fact, it has been proposed that any serotype could poten-
tially possess diabetogenicity. Furthermore, aggre-ssive,
cytolytic isolates can develop from originally benign ente-
rovirus serotypes (Roivainen et al., 2002).
Mechanisms of beta cell obliteration
T1D is an immune-mediated disease in which a specific
immune response to islet beta cells is induced. In the
animal model of T1D, the NOD mouse, diabetes may be
transferred only with cells of the immune system. Also in
048 Sci. Res. Essays
Figure 3. The putative pathways involved in virus induced beta cell death. A potential
course of the process: Virus infects the beta cell and stimulates intracellular IFN/ ß and
MHC I production and the innate arm (IFN, nonspecific antibodies and NK cells) of the
hosting immune system. The resulting initial beta cell damage provokes the activation and
infiltration of macrophages and T cells, which secrete cytokines and finally stimulate
antibody production. The produced cytokines, interferons, nitric oxide, ROS and perforin
contribute to beta cell death through apoptosis or necrosis.
humans T1D has been adoptively transferred via bone-
marrow transplantation from a diabetic donor. However,
there are several mechanisms through which a virus
could induce immune mediated destruction of islet beta-
cells, as shown in Figure 3.
Cell lysis
Lytic cell death is an aggressive, rapid process featured
by breakdown of the cell's structural integrity. It equals
necrosis, which is used generally to describe cell death of
aggressive nature. As mentioned earlier viruses often
induce host cell pyknosis in the initial phase of infection
(Koch and Koch, 1985), which may in later stages turn
into lysis. In a recent study (Roivainen et al., 2002) a
large group of prototype enteroviruses from different ge-
netic subgroups were characterized according to their
cytolytic activity. Echovirus 6, 7, 11, CVA 13, CVB 1, 3, 4,
5, 6 and poliovirus type 1/Mahoney were detected to
cause beta cell lysis. Furthermore an echovirus 9 isolate,
from a 6-week-old baby with acute onset T1D was found
with destructive features unlike its corresponding proto-
type virus. These variations of outcomes might have to
do with donor-related characteristics, thus the possibility
that a certain HLA-type or antibody-positivity could confer
resistance or susceptibility for destructive beta cell death
remains.
Programmed cell death (Apoptosis)
In order to maintain an efficient multiplication environ-
ment, viruses may prevent or delay host cell death
through various routes. Especially the abilities to block
p53 tumor suppressor -dependent and Bax/Bak/Bik
dependent apoptosis are well characterized (Afonso et
al., 1996; Bargonetti et al., 1992; Chen et al., 1996; Hen-
derson et al., 1993; Dobner et al., 1996; Scheffner et al.,
1993). On the other hand the same pathways may be
used in the opposite direction to induce host cell apop-
osis and safely release new viral particles in enveloped
bodies to be engulfed by other target cells (Debbas and
White, 1993; Sastry et al., 1996; Westendorp et al.,
1995).
Additionally, by this mechanism the inflammatory
/immune reactions following lytic cell death or free virus
exposure are avoided and contact with neutralizing anti-
bodies is prevented, thus potentiating the survival of the
virus and uninterrupted dissemination of infection (Glied-
man et al., 1975; Jeurissen et al., 1992). Concerning non-
enveloped viruses, into which category enteroviruses be-
long, the mechanisms of exit from a dying host cell are
still obscure. Although cell lysis is thought to play a major
role, a recent report presents host cell apoptosis as a
potential new explanation. It would result in the formation
of membrane-bound bodies enabling viral exit (Teodoro
and Branton, 1997). In the group of picornaviruses, apop-
tosis has been extensively studied by the model of CVB3
induced myocarditis, during which persistence often
develops and apoptotic myocardial death is evident
(Carthy et al., 1998; Huber et al., 1999b). Also poliovirus
infection or individual poliovirus proteases alone are
reported to be able to induce apoptotic death in selected
cell types (Barco et al., 2000; Girard et al., 1999;
Goldstaub et al., 2000; Lopez-Guerrero et al., 2000;
Tolskaya et al., 1995).
Considering development of T1D, virus infections are
regarded as the leading environmental triggers and apop-
tosis the leading form of beta cell death (Eizirik and
Mandrup-Poulsen, 2001). Their possible association is
difficult to study in clinical materials. Experimentally,
recent studies on rat islets stressed with synthetic double
stranded RNA, a general product of viral replication;
reveal enhanced susceptibility to cytokine induced islet
cell apoptosis by Fas-FasL interaction (Liu et al., 2002).
Virus infections generally induce an inflammatory reac-
tion mediated in part by proinflammatory cytokines. In
addition to their direct proapoptotic potential, cytokines
may also launch the expression of MHC class I mole-
cules in beta cells followed by their appearance on plas-
ma membranes, which exposes beta cells to T cell me-
diated death through Fas or perforin dependent pathways
(Kim et al., 2002; Seewaldt et al., 2000).
Enteroviruses as key players of autoimmunity in T1D
Molecular mimicry
Molecular mimicry is based on a sequential and structural
homology between a foreign antigen and a host protein
enabling an immunologic attack against the pathogen to
cross-react with the host molecule (Davies, 1997). Gluta-
mic acid decarboxylase GAD65, an enzyme synthesizing
the inhibitory neurotransmitter gamma amino butyric acid
(GABA), is one of the important islet cell auto-anti-gens in
T1D. The sequences of GAD65 and the non-structural
protein 2C of CVB-like enteroviruses were shown to
share a similar motif (PEVKEK) and to bind the same
groove in a MHC-molecule in an exactly similar position
(Kaufman et al., 1992). In later studies with synthetic pep-
tides, the binding was restricted to HLA-DR3 molecule
(Vreugdenhil et al., 1998). Furthermore, GAD65-reactive
T-cells are present and circulating in both T1D patients
and healthy normal people, with the difference of func-
tioning like pre-activated memory cells in patients and
thus being highly more susceptible to enter clonal expan-
sion. These data have given evidence for molecular
mimicry between enteroviral and self-fac-tors as a possi-
ble mechanism of beta cell destruction, although contra-
dictory observations have also been made (Atkinson and
Maclaren, 1994; Horwitz et al., 1998; Marttila et al., 2001;
Schloot et al., 2001). Accordingly, molecular mimicry
could also exist between GAD and cytomegalovirus on T-
cell level (Roep et al., 2002).
Local inflammatory destruction and bystander
progression
The model of locally (intra-pancreatically) launched beta
cell autoimmunity and destruction has been widely stu-
Sharma et al. 049
died by Horwitz et al. (1999). Their recent report con-
cluded that the primary role of an enterovirus infection
(pancreatrophic CVB4) in the pathogenesis of T1D is to
specifically infect and damage beta cells leading to
release of sequestered islet antigen and stimulation of a
local inflammatory response. The intra-pancreatic antigen
presenting cells introduce the antigen to the population of
resting beta-cell-autoreactive T-cells resulting in initiation
of the disease process. Unless a critical threshold level of
these precursor autoreactive T-cells exist, the destructive
process will not be triggered (Horwitz et al., 1998). They
also observed the critical role of beta cell damage as the
driving force of virus induced T1D, since a non-specific
cytokine-attack did not precipitate T1D in mice with a
diabetogenic T cell repertoire, while exposure to the beta-
cell toxin streptozotozin was able to do it (Horwitz et al.,
2002). CVB infections have previously been reported to
lead to necrotic cell death in both rodent and human exo-
crine pancreatic tissue (Arnesjo et al., 1976; Lansdown,
1976; Ozsvar et al., 1992; Ramsingh, 1997a; Vella et al.,
1992). The inflammatory mediators produced during this
infective process have been speculated to activate the
bystanding beta-cell-autoreactive T-cell pool and thus
possibly accelerate T1D development (Serreze et al.,
2000).
Oxidative stress
Oxidative metabolites Vs apoptosis
Oxygen derived reactive metabolites (reactive oxygen
species, ROS) appear abundantly in the human body
(Figure 4). They form as a consequence of incomplete
reduction of molecular oxygen by the mitochondrial respi-
ratory (electron-transport) chain (Fernandez-Checa et al.,
1998) or through some reactions of cellular metabolism
by oxidizing enzymes including xanthine oxidase, P450
mono-oxygenase, cyclooxygenase, lipoxygenase, mono-
amine oxidase etc (Boveris and Chance, 1973; Forman,
1982; Maeda et al., 1999; Siraki et al., 2002). The most
common intracellular ROS molecules include superoxide
anion (O
2
-
), hydrogen peroxide (H
2
O
2
) and hydroxyl radi-
cal (-OH) (Fridovich, 1998; DiGuiseppi and Fridovich,
1984). ROS react with biological molecules such as pro-
teins, lipids, carbohydrates and DNA threatening their
integrity and exposing the organism to toxic, mutagenic
and carcinogenic assaults (Stadtman and Levine, 2000;
Steinberg, 1997; Marnett, 2000). Low molecular weight
GTPases Ras and Rac1, for example, have been shown
to be ROS targets leading to altered signal transduction
(Irani et al., 1997; Sundaresan et al., 1996). Except for -
OH, which is mostly noxious, ROS also have beneficial
functions mainly as stimulators and mediators of intra-cel-
lular signaling cascades (Krieger-Brauer and Kather,
1992; Lo and Cruz, 1995; Ohba et al., 1994; Sundaresan
et al., 1995). They also function in microbial killing during
infection, for example through inactivating viral enzymes
by nitrosylation (Adams et al., 1990; Babior, 1978a,b;
050 Sci. Res. Essays
Figure 4. The major pathways of ROS generation and antioxidant defense
(modified from Jacobson MD 1996).
Colasanti et al., 1999). Their properties often shift into
pathologic in response to increasing concentrations,
which might be due to the metabolic state (Di Meo and
Venditti, 2001; Gambelunghe et al., 2001), pH, oxygen
partial pressure (pO
2
) and ADP availability. Very com-
monly, overproduction of ROS is triggered by nonphysio-
logical states like inflammation or exogenous toxins,
which stimulate phagocytic cells to produce ROS. In a
variety of disease processes associated with either acute
or chronic inflammation, increased oxidative stress ap-
pears to be involved. Depending on the surrounding
milieu, ROS molecules may also react together. Speci-
fically, in a reaction between O
2
-
and NO highly cytotoxic
peroxynitrite, ONOO
-
is formed (Beckman and Koppenol,
1996), which may further react with carbon dioxide (CO
2
)
to form peroxocarboxylate (ONOOCO
2
-
). H
2
O
2
may react
with H
+
+ Cl
-
or Cu
+
/Fe
2+
forming either highly reactive
HOCl (hypochlorus acid) or a hydroxyl radical, respect-
tively (Babior, 2000).
Highlighting the dual role of ROS in both physiological
and disease states, recent data show concentration de-
pendent stimulation of either cell growth or death. Mode-
rate levels of pro-oxidants may promote mitogenic stimuli
(Burdon and Rice-Evans, 1989) e.g. by affecting kinase
or proto-oncogene activities (Cerutti and Trump, 1991;
Larsson and Cerutti, 1988). At slightly increased levels
ROS are reported at various different experimental
conditions to be able to induce apoptotic cell death either
directly or in combination with antioxidant, mainly gluta-
thione (GSH), depletion (Hampton and Orrenius, 1997;
Lennon et al., 1991; Macho et al., 1997). The stimulatory
mechanisms often involve exposing to or triggering the
mitochondrial membrane permeability transition which
leads to the release of apoptogenic factors and thus
stimulation of the downstream apoptotic cascade (Arm-
strong and Jones,2002; Armstrong et al., 2002; Costan-
tini et al., 1996;. Datta et al., 2002; Dypbukt et al., 1994;
Fleury et al., 2002; Petronilli et al., 1994; Ueda et al.,
2002; Wei and Lee, 2002) Also, HIV-infection mediated
apoptosis of CD4+ T-cells is preceded by ROS produc-
tion and antioxidant depletion. This process has recently
been associated with p53, NF-B and AP-1, pro-apoptotic
redox-active factors, that also support expression of viral
genes and pro-inflammatory cytokines (Perl et al., 2002).
p53, independent of the activating factor, may further
stimulate ROS production and induce apoptosis (Sawada
et al., 2001; Li et al., 1999).
Moreover, studies evidencing antioxidants' capacity to
inhibit apoptosis and recent observations on antiapoptotic
molecules with antioxidative properties further argue for a
role for ROS in stimulating apoptosis (Kelso et al., 2001;
Melnick, 1996; Sato et al., 2002). Additionally, oxidants
may direct an apoptotic process into a necrotic one by
inactivating caspases or impairing the mitochondrial
energy production resulting in subsequent ATP depletion
(Leist et al., 1999; Samali et al., 1999). Although possible
in the other direction, overwhelming oxidative stress
usually always leads to necrosis leaving no possibilities
to inhibit or re-regulate the process towards apoptosis.
Antioxidative system
In an attempt to prevent ROS-mediated damage, cells
have developed an antioxidative defense system (Benzie,
2000). The main goal is to maintain the intracellular
milieu in a reduced and stable state. This antioxidative
machinery consists of several components including
heme- or thiol-based enzymatic systems and non-enzy-
matic scavengers. The most ubiquitous and abundant of
them is the glutathione (GSH) system (Anderson, 1998;
Deneke and Fanburg, 1989). GSH is formed from the
aminoacids glutamate, cysteine and glycine in two ATP-
dependent enzymatic reactions by gamma-glutamyl-
cysteine synthetase and glutathione synthetase (Griffith
and Mulcahy, 1999). GSH itself functions mainly as a
sulfhydryl buffer and helps to detoxify xenobiotics in con-
jugation reactions catalyzed by glutathione S-transferase.
Most importantly glutathione peroxidases, selenocys-
teine-containing enzymes, use GSH as substrate in the
elimination of e.g. hydroxyl radical, peroxynitrite, hydro-
peroxides and reactive electrophiles. In these reactions
two GSH-molecules are oxidized to GSSG, which is
converted back to reduced GSH in an NADPH-dependent
reaction either by glutathione reductase or the thioredoxin
system, described below. Furthermore, as a thiol-con-
taining reductant, GSH maintains so-called thiol-enzymes
in their catalytically active forms, and low molecular
weight antioxidants, vitamins C and E in their biologically
active forms (Fridovich, 1999; Havivi et al., 1991; Living-
stone and Davis, 2007). Other major antioxidants include
catalase (CAT) and superoxide dismutase (SOD). As
referred to by its name, the latter inhibits ROS-mediated
damage by scavenging O
2
-
(Fridovich, 1995, 1999). Two
isoenzymes exist: an essential mitochondrial Mn-SOD
and a less essential cytosolic Cu/Zn-SOD. They meta-
bolize two O2
-
molecules to O
2
and H
2
O
2
; thus another
ROS is formed. The general and ubiquitous mechanism
to remove H
2
O
2
is by peroxisomal heme-containing CAT,
which converts two H
2
O
2
molecules to O
2
and 2 x H
2
O
(Kirkman and Gaetani, 1984; Kirkman et al., 1999). More-
over, CAT detoxifies e.g. phenols and alcohols via cou-
pled reactions with H
2
O
2.
Another thiol-containing and
ubiquitous reducing enzyme system is the thioredoxin
(Trx) machinery. Trx functions as a hydrogen donor for
other catalytic enzymes (e.g. glutathione peroxidase) and
reduces disulfide bonds of diverse proteins (Holmgren,
1984). It protects particularly against peroxide induced
stress (Nakamura et al., 1994; Spector et al., 1988) and
has also been reported to have anti-apoptotic power
Sharma et al. 051
(Saitoh et al., 1998). Like the GSH system, oxidized Trx
is converted back to reduce form by a flavoenzyme thio-
redoxin reductase (TrxR) in an NADPH-dependent reac-
tion (Holmgren, 1985). In addition to Trx, TrxR reduces
lipid peroxides, diverse antioxidative selenium containing
compounds and converts vitamin C back to its active
reduced form.
A distinct, non-enzymatic group of antioxidants of low
molecular weight include vitamins E and C, several sele-
nium-containing compounds, lipoic acid, and ubiqui-
nones. Ascorbic acid (AA; vitamin C) and alpha-toco-
pherol (alpha-TOH; vitamin E) constitute the major water-
soluble and lipid-soluble small antioxidants, respectively
(Buettner, 1993). As alpha-TOH reduces e.g. peroxyl
radicals by its OH group and converts into tocopheroxyl
(chromanoxyl) radical, vitamin C, as its main function,
reduces it back to alpha-TOH, the active vitamin E (Sies
et al., 1992). Additionally, vitamin C acts as an electron
donor for some transmembrane enzymes with oxidore-
ductase activity (May et al., 1995). As previously men-
tioned, TrxR reduces oxidized vitamin C (Mendiratta et
al., 1998), but its major recycler is GSH (May et al.,
1996). Heme (Fe protoporphyrin IX) is an integral protein
to life delivering oxygen into cells. It circulates incorpo-
rated in hemoproteins and occurs unbound (free) only in
pathologies sometimes associated with tissue accumula-
tion. Free heme as such is a powerful oxidative molecule
and thus a system for its degradation exists. It consists of
an oxidative stress–inducible protein HO-1 (heme oxy-
genase-1, HSP32) and the constitutive isozyme HO-2,
which catalyze the oxidation of heme to biologically active
molecules: free iron, a gene regulator (Ferris et al. 1999),
biliverdin, an antioxidant and carbon monoxide, a heme
ligand (Maines, 1997). Uric acid is produced in liver and
represents the end product of the purine metabolism
reaction chain. The enzyme xanthine oxidase catalyses
the last two reactions from hypoxanthine through xanthine
to uric acid as byproducts of which two molecules of H
2
O
2
are created. Uric acid is considered a powerful antioxi-
dant (Ames et al., 1981) especially because of its poten-
cy in peroxynitrite scavenging (Balavoine and Geletii,
1999; Regoli and Winston, 1999; Whiteman et al., 2002)
but also in its ability to chelate transition metal ions and
stabilize reactive hypochlorus acid and hydroxyl radical
(Becker, 1993). On the other hand, its increased produc-
tion indicates increased H
2
O
2
production, which may
further react with peroxynitrite and form new aggressive
oxidative metabolites (Santos et al., 1999; Skinner et al.,
1998; Vasquez-Vivar et al., 1996) including alloxan and a
nitrite derivative capable of NO release. These both are
known for their direct or indirect beta cell toxicity (Lenzen
and Panten, 1988). NO may further stimulate peroxy-
nitrate production by reacting with superoxide, thus con-
tributing to beta cell injury (Suarez-Pinzon et al., 1997,
2001). However, there is evidence that uric acid might
further react with and scavenge also these newly formed
radicals (Kooy et al., 1994; Whiteman and Halliwell,
052 Sci. Res. Essays
1996), which implies to a dual role for this molecule in
controlling/affecting the cellular redox status.
Redox status in beta cells Vs T1D
It is well recognized that type 2 diabetes is a progressive
condition with insulin production tending to fall with time
as the beta cell mass fails and the cells synthesize less
insulin. Hyperglycaemia is recognized to have toxic
effects on beta cell, so-called ‘glucotoxicity’ leading to
reduced insulin gene expression, impaired insulin secre-
tion and ultimately cell death (Robertson, 2004). Recently
it has been hypothesized that chronic oxidative stress as
a consequence of hyperglycaemia is an important me-
chanism for glucotoxicity (Robertson et al., 2003). The
mechanisms whereby glucose can lead to the production
of relative oxidative species (ROS) in beta cells are well
defined. Beta cells appear to be vulnerable to oxidative
stress as they contain relatively low levels of glutathione
peroxidase (GPx) and other protective enzymes com-
pared to other cells (Grankvist et al., 1981). Moreover,
studies on pancreatic islets and beta cell lines observed
the cells to be incapable of increasing their antioxidant
enzyme expression in response to cellular stress induced
by exposure to glucose (Tiedge et al., 1997). There is
evidence that ROS are also involved in the pathogenesis
of peripheral insulin resistance (Haber et al., 2003). Nitro-
sative stress is also thought to contribute to the
pathogenesis of beta cell apoptosis (Bast et al., 2002).
Over time these factors increase the burden on beta
cells, promoting the development of diabetes.
Nevertheless, cellular damage in diabetes is associated
with various biochemical pathways. There is some evi-
dence of trace element and vitamin deficiencies in dia-
betes which may contribute to beta cell damage (Aruoma,
1998; Havivi et al., 1991), concomitant with oxi-dative
mechanisms participating in the pathogenesis of both
beta cell destruction at the prediabetic period and of
vascular damage and endothelial dysfunction during the
development of further disease complications. According
to recent knowledge, the latter are mostly due to
increased mitochondrial production of superoxide anion
stimulated by prolonged hyperglycemia (Nishikawa et al.,
2000). The earlier mentioned NF-B, an important media-
tor of cytokine induced pathways of beta cell damage, is
also redox-sensitive and assumably also activated by the
minor amounts of ROS produced at the initial stage of
beta cell destruction. NF-B is further reported to
increase the expression level of iNOS (among other
genes) and then through NO-signaling to down regulate
the expression levels of beta cell specific genes including
Pdx-1, Glut-2, and Isl-1 and also to stimulate the produc-
tion of cytokines and chemotactic agents amplifying the
production of ROS. This cascade then leads to beta cell
dysfunction and death through apoptosis or necrosis
depending on the overall stress (Cardozo et al., 2001; Ho
and Bray, 1999). A chain of events resembling this is also
considered to characterize insulitis, during which T-cells
and macrophages invade the islets and act as sources of
inflammatory mediators and ROS (Rabinovitch et al.,
1996b). Many studies have discovered the general anti-
oxidative capacity to be defective in diabetic patients
(Maxwell et al., 1997; Santini et al., 1997; Tsai et al.,
1994). Depending on the design of the study, decreased
levels of vitamin C, vitamin E, uric acid, GSH and GSH-
related enzymes have been detected in diabetics com-
pared to control subjects (Courderot-Masuyer et al.,
2000; Hoeldtke et al., 2002; Jain and McVie, 1994; Marra
et al., 2002; Maxwell et al., 1997; Ruiz et al., 1999;
Seghieri et al., 1998; Seghrouchni et al., 2002; Sharma et
al., 2000). Controversial results also exist and imply the
need for critical evaluation of the markers measured to
assess oxidative stress (Leinonen et al., 1998;
VanderJagt et al., 2001; Vessby et al., 2002). Importantly,
beta cells themselves have a poorer antioxidative
defense system in comparison to other cell types. Speci-
fically, the expression levels and activities of antioxidants
are low and the adaptive properties to increase antioxi-
dant enzyme production during stress are limited, thus
rende-ring beta cells extremely vulnerable to oxidative
damage (Grankvist et al., 1981; Lenzen et al., 1996;
Tiedge et al., 1997). Further evidence for these defects
have been obtained from substitution studies on insulin
producing cells showing effective protection against
ROS- or cytokine mediated cell death by a cocktail of
antioxidants namely CAT+Gpx+SOD (Loetz et al.,2000;
Tiedge et al., 1998). Beta cells' own properties are also
considered to affect their survival, exemplified by exagge-
rated ROS-sensitivity during low glucose levels and slow
mitochon- drial metabolic rate. Thus, normal kinetics of
glucose metabolism may protect beta cells against ROS-
induced damage (Pipeleers et al., 2001).
The role of Trx has been studied in nonobese diabetic
mice overexpressing Trx specifically in beta cells. The
incidences of both spontaneous autoimmune and drug-
(streptozotocin, a ROS generating agent) induced T1D
were reduced (Hotta et al., 1998). Also after major injury
(partial pancreatectomy) or transplantation, antioxidative
molecules have been shown to increase beta-cell survival
through e.g. attenuated apoptotic beta cell death and
increased viability (Ribeiro et al., 2003; Laybutt et al.,
2002; Bottino et al., 2002; Gunther et al., 2002; Pileggi et
al., 2001). Overall, the existing data indicates oxidative
molecules to be important mediators of beta cell damage
and thus suggests a role for antioxidant supplementation
in diabetes. Due to the complexity and non-specificity of
the machinery controlling redox status, this balance is of
crucial importance.
Interaction between oxidative stress and Virus
infections:
Malnutrition is known to result in defective and inefficient
antioxidative capacity and aggravated exposure to oxida-
tive stress triggered by various pathogens, e.g. viruses
(Sofic et al., 2002). According to the traditional view, mal-
nutrition could predispose an individual to infections by
weakening the host's immune system and thus allowing
the pathogen to multiply and disseminate in the orga-
nism. Today it is generally known that many virus infec-
tions (e.g. influenzavirus induced pneumonia, HIV, cox-
sackievirus myocarditis) exert many kinds of oxidative
stress in the host (Maeda and Akaike, 1998; Schwarz;
1996; Xie et al., 2002b). Phagocytes become activated
and produce ROS (Peterhans et al., 1987) and pro-oxida-
tive cytokines (TNF-, IL-1, which further poten-tiate
oxidative and other viral damage through e.g. increased
virus multiplication, impaired mitochondrial function and
reactive iron accumulation (Polla et al., 1996; Schulze-
Osthoff et al., 1992; Schreck et al., 1992; Klempner et al.,
1978). Various studies have demonst-rated the effect of
these cytokines to provoke ROS, especially NO, produc-
tion within the islets, both in macrophages and beta cells
(Rabinovitch and Suarez-Pinzon, 1998; Eizirik et al.,
1996; Rabinovitch et al., 1996a; Mandrup-Poulsen et al.,
1990). Similarly, the pathogenesis of influenza virus
induced pneumonia has been reported to be mostly due
to IFN- mediated NOS and iNOS activation resulting in
NO and further peroxy-nitrate production (Akaike et al.,
1996). The previously mentioned antioxidant-substitution
studies confirm ROS production to mediate IL-1+ TNF-
+ IFN- stimulated beta cell damage (Lortz et al., 2000).
These interactions are further strengthened by the know-
ledge of NF-B as a mediator of both ROS and cytokine
induced apoptosis and the fact that also ROS me-
diate p53- dependent apoptosis, the pathway com-
monly used by viruses (Datta et al., 2002; Armstrong et
al., 2002). The exposure to radical stress is further inten-
sified by virus-mediated impairment of host's antioxidative
defenses through decreasing concentrations of several
ROS scavengers (Allard et al., 1998; Bannister et al.,
1986; Hennet et al., 1992; Staal et al., 1992; Xie et al.,
2002a). Many in vitro studies have characterized this
phenomenon and observed reparative effects by supple-
menting antioxidants. In vivo the effect of antioxidant
therapy is usually tested as a supplement to some speci-
fic antiviral treatment, because of the antioxidants weak
efficiency alone. In some combinations, antioxidative sup-
plement has resulted in improved outcome of the specific
therapy for influenzavirus and HIV infections, although
the specific mechanisms of protection still remain poorly
understood (Allard et al., 1998; Oda et al., 1989). Con-
cerning the immune system, ROS may act as immuno-
modulators allowing or activating T-cell proliferation, an
essential phenomenon in cell-mediated immune res-
ponse. This idea is based on observations on the ability
of several antioxidants to directly inhibit T-cell prolife-
ration or the activation of transcription factors involved in
T-cell activation (Chaudhri et al., 1986; Hunt, 1994). On
the other hand, the fatality of HIV infection lies on the exact
opposite: increased apoptosis of CD4+ T-cells potentiated
or possibly even triggered by the changed redox status
Sharma et al. 053
(Banki et al., 1998; Romero-Alvira and Roche, 1998).
At last but not least, it has been shown that the interact-
tions extend further malnutrition and deficiency of antioxi-
dants are capable of affecting not only the host, but also
the pathogen by increasing its virulence. In a series of
experiments Beck et al. (1997) first observed that an
avirulent strain of CVB3 (3/0), which did not cause any
damage in normally fed control mice, established mode-
rate myocardial lesions in both selenium and vitamin E
deficient mice. When this avirulent strain was further pas-
saged in a selenium or vitamin E deficient mouse and
then reinoculated into a control mouse, severe myocar-
dial damage was provoked. This was shown to be due to
six point mutations in the viral genome, which changed
an originally avirulent strain into virulent (Beck, 1997). As
an example of another virus, ROS are reported to acti-
vate the binding of NF-B to the viral promoter region of
HIV resulting in increased viral replication and production
of Tax protein. Tax again stimulates NF-B (Baruchel and
Wainberg, 1992) thus creating a vicious cycle for the
benefit of the dissemination of the infection.
ACKNOWLEDGEMENTS
An institutional research promotion grant to the Depart-
ment of Biotechnology, College of Engineering and Tech-
nology, Moradabad, U.P., India is duly acknowledged.The
authors are grateful to Prof. R.M. Dubey (Managing
Director), Prof. B.N. Basu (Director, Acade-mics) and
Prof. B. N. Kaul (Director, Administra-tion) of CET, IFTM,
Moradabad, U.P., India for providing the necessary facili-
ties and encouragement. Besides, Prof. R.B. Singh,
Halberg Hospital & Research Institute, Moradabad, U.P.,
India is thanked for critical and valuable comments on the
manuscript.
REFERENCES
Adachi K, Muraishi A, Seki Y, Yamaki K, Yoshizuka M (1996).
Coxsackievirus B3 genomes detected by polymerase chain reaction:
evidence of latent persistency in the myocardium in experimental
murine myocarditis. Histol. Histopathol. 11: 587-596.
Afonso CL, Neilan JG, Kutish GF, Rock DL (1996). An African swine
fever virus Bc1-2 homolog, 5-HL, suppresses apoptotic cell death. J.
Virol. 70; 4858-4863.
Agol VI, Belov GA, Bienz K, Egger D, Kolesnikova MS, Raikhlin NT,
Romanova LI, Smirnova EA Tolskaya EA (1998). Two types of death
of poliovirus-infected cells: caspase involvement in the apoptosis but
not cytopathic effect. Virolology 252, 343-353.
Ahmad S, Alsayed Y, Druker BJ, Platanias LC (1997). The Type I
interferon receptor mediates tyrosine phosphorylation of the CrkL
adaptor protein. J. Biol. Chem. 272: 29991-29994.
Akaike T, Noguchi Y, Ijiri S, Setoguchi K, Suga M, Zheng YM,
Dietzschold B, Maeda H (1996). Pathogenesis of influenza virus-
induced pneumonia: involvement of both nitric oxide and oxygen
radicals. Proc. Natl. Acad. Sci. USA 93: 2448-2453.
Allard JP, Aghdassi E, Chau J, Salit I, Walmsley S (1998) Oxidative
stress and plasma antioxidant micronutrients in humans with HIV
infection. Am. J. Clin. Nutr. 67: 143-147.
Ames BN, Cathcart R, Schwiers E, Hochstein P (1981). Uric acid
provides an antioxidant defense in humans against oxidant- and
radical-caused aging and cancer: a hypothesis. Proc. Natl. Acad. Sci.
USA. 78: 6858-6862.
054 Sci. Res. Essays
Anastasi E, Santangelo C, Bulotta A, Dotta F, Argenti B, Mincione C,
Gulino A, Maroder M, Perfetti R, Di Mario U (2005). The acquisition
of an insulin secreting phenotype by HGF treated rat pancreatic
ductal cells (ARIP) is associated with the development of
susceptibility to cytokine – induced apoptosis J. Mole. Endocrinol. 34:
367-376.
Anderson ME (1998). Glutathione: an overview of biosynthesis and
modulation. Chem. Bio. Interact. 111-112, 1-14.
Andreoletti L, Hober D, Becquart P, Belaich S, Copin MC, Lambert V,
Wattre P (1997). Experimental CVB3-induced chronic myocarditis in
two murine strains: evidence of interrelationships between virus
replication and myocardial damage in persistent cardiac infection. J.
Med. Virol. 52: 206-214.
Armitage RJ (1994). Tumor necrosis factor receptor superfamily
members and their ligands.Curr Opin. Immunol. 6: 407-413.
Armstrong JS, Jones DP (2002). Glutathione depletion enforces the
mitochondrial permeability transition and causes cell death in Bcl-2
overexpressing HL60 cells. FASEB J. 16: 1263-1265.
Armstrong JS, Steinauer KK, Hornung B, Irish JM, Lecane P, Birrell
GW, Peehl DM, Knox SJ (2002). Role of glutathione depletion and
reactive oxygen species generation in apoptotic signaling in a human
B lymphoma cell line. Cell Death Differ. 9: 252-263.
Arnesjo B, Eden T, Ihse I, Nordenfelt E, Ursing B (1976). Enterovirus
infections in acute pancreatitis - a possible etiological connection.
Scand. J. Gastroenterol. 11: 645-649.
Aruoma OI (1998). Free radicals, oxidative stress and antioxidants in
human health and disease. J. Am. Chem. Soc. 75: 448-456.
Ashkenazi A, Dixit VM (1998). Death receptors: signaling and modula-
lation. Science 281: 1305-1308.
Atkinson MA, Maclaren NK (1994). The pathogenesis of insulin-
dependent diabetes mellitus. N. Engl. J. Med. 331: 1428-1436.
Augstein P, Wachlin G, Berg S, Bahr J, Salzsieder C, Hehmke B,
Heinke P, Salzsieder E (2003). Surface and intracellular Fas
expression associated with cytokine- induced apoptosis in rodent islet
and insulinoma cells. J. Mol. Endocrinol. 30: 163-171.
Babior BM (1978a). Oxygen-dependent microbial killing by phagocytes
(first of two parts). N. Engl. J. Med. 298: 659-668.
Babior BM (1978b). Oxygen-dependent microbial killing by phagocytes
(second of two parts). N Engl. J. Med. 298: 721-725.
Babior BM (2000). Phagocytes and oxidative stress. Am. J. Med. 109:
33-44. 245, 229-240.
Bach JF (1994). Insulin-dependent diabetes mellitus as an autoimmune
disease. Endocr. Rev. 15: 516-542.
Baker SJ, Reddy EP (1998). Modulation of life and death by the TNF
receptor superfamily. Oncogene 17: 3261-3270.
Balavoine GG, Geletii YV (1999). Peroxynitrite scavenging by different
antioxidants. Part I: convenient assay. Nitric Oxide 3: 40-54.
Banatvala JE, Bryant J, Schernthaner G, Borkenstein M, Schober E,
Brown D, De Silva LM, Menser MA, Silink M (1985). Coxsackie B,
mumps, rubella, and cytomegalovirus specific IgM responses in
patients with juvenile-onset insulin-dependent diabetes mellitus in
Britain, Austria, and Australia. Lancet 1: 1409-1412.
Banki K, Hutter E, Gonchoroff NJ, Perl A (1998). Molecular ordering in
HIV-induced apoptosis. Oxidative stress, activation of caspases, and
cell survival are regulated by transaldolase. J. Biol. Chem. 273:
11944-11953.
Balachandran S, Roberts PC, Kipperman T, Bhalla KN, Compans RW,
Archer DR, Barber GN (2000). Alpha/Beta Interferons Potentiate
Virus-Induced Apoptosis through Activation of the FADD/Caspase-8
Death Signaling Pathway. J. Virol. 74: 1513-1523.
Bannister WH, Federici G, Heath JK, Bannister JV (1986). Antioxidant
systems in tumour cells: the levels of antioxidant enzymes, ferritin,
and total iron in a human hepatoma cell line. Free Radic Res.
Commun. 1: 361-367.
Bargonetti J, Reynisdottir I, Friedman PN, Prives C (1992). Site-specific
binding of wild-type p53 to cellular DNA is inhibited by SV40 T
antigen and mutant Genes Dev. 53(6): 1886-1898.
Baruchel S, Wainberg MA (1992). The role of oxidative stress in
disease progression in individuals infected by the human
immunodeficiency virus. J. Leukoc. Biol. 52: 111-114.
Bast A, Wolf G, Oberbaumer I, Walther R (2002). Oxidative and
nitrosative stress induces peroxiredoxins in pancreatic beta cells.
Diebeteologia 45: 867-876.
Beck MA, Chapman NM, McManus BM, Mullican JC, Tracy S. (1990)
Secondary enterovirus infection in the murine model of myocarditis.
Pathologic and immunologic aspects. Am. J. Pathol. 136: 669-681.
Beck MA (1997). Increased virulence of coxsackievirus B3 in mice due
to vitamin E or selenium deficiency. J. Nutr. 127: 966S-970S.
Becker BF (1993). Towards the physiological function of uric acid. Free
Radic. Biol. Med. 14: 615-631.
Beckman JS, Koppenol WH (1996). Nitric oxide, superoxide, and
peroxynitrite: the good, the bad, and ugly. Am. J. Physiol. 271:
C1424-1437.
Bendelac A, Carnaud C, Boitard C, Bach JF (1987). Syngeneic transfer
of autoimmune diabetes from diabetic NOD mice to healthy
neonates. Requirement for both L3T4+ and Lyt- 2+ T cells. J Exp.
Med. 166: 823-832.
Bendig JW, Fleming DM (1996). Epidemiological, virological, and
clinical features of an epidemic of hand, foot, and mouth disease in
England and Wales. Commun. Dis. Rep. CDR. Rev. 6: R81-86.
Bennett ST, Lucassen AM, Gough SC, Powell EE, Undlien DE,
Pritchard LE, Merriman ME, Kawaguchi Y, Dronsfield MJ, Pociot F
(1995). Susceptibility to human type 1 diabetes at IDDM2 is
determined by tandem repeat variation at the insulin gene
minisatellite locus. Nat. Genet. 9: 284-292.
Benzie IF (2000). Evolution of antioxidant defence mechanisms. Eur. J.
Nutr. 39: 53-61.
Bergelson JM, Mohanty JG, Crowell RL, St John NF, Lublin DM,
Finberg RW (1995). Coxsackievirus B3 adapted to growth in RD cells
binds to decay- accelerating factor (CD55). J. Virol. 69: 1903-1906.
Bergelson JM, Cunningham JA, Droguett G, Kurt-Jones EA, Krithivas A,
Hong JS, Horwitz MS, Crowell RL, Finberg RW (1997). Isolation of a
common receptor for Coxsackie B viruses and adenoviruses 2 and 5.
Sci. 275: 1320-1323.
Bergelson JM, Krithivas A, Celi L, Droguett G, Horwitz MS, Wickham T,
Crowell RL, Finberg RW (1998). The murine CAR homolog is a
receptor for coxsackie B viruses and adenoviruses. J. Virol. 72: 415-
419.
Bottino R, Balamurugan AN, Bertera S, Pietropaolo M, Trucco M,
Piganelli JD (2002). Preservation of human islet cell functional mass
by anti-oxidative action of a novel SOD mimic compound. Diabetes
51: 2561-2567.
Bose S, Banerjee AK (2003). Innate immune response against
nonsegmented negative strand RNA viruses. J. Interferon Cytokine
Res. 23: 401-412.
Buettner GR (1993). The pecking order of free radicals and
antioxidants: lipid peroxidation, alphatocopherol, and ascorbate.
Arch. Biochem. Biophys. 300: 535-543.
Buja LM, Eigenbrodt ML, Eigenbrodt EH (1993). Apoptosis and
necrosis. Basic types and mechanisms of cell death. Arch. Pathol.
Lab. Med. 117: 1208-1214.
Burdon RH, Rice-Evans C (1989). Free radicals and the regulation of
mammalian cell proliferation. Free Radic. Res. Commun. 6: 345-358.
Campbell RD, Trowsdale J (1993). Map of the human MHC. Immunol.
Today 14: 349-352.
Cardozo AK, Heimberg H, Heremans Y, Leeman R, Kutlu B, Kruhoffer
M, Orntoft T, Eizirik DL (2001). A comprehensive analysis of
cytokine-induced and nuclear factor-kappa B- dependent genes in
primary rat pancreatic beta-cells. J. Biol. Chem. 276: 48879-48886.
Carthy CM, Granville DJ, Watson KA, Anderson DR, Wilson JE, Yang
D, Hunt DW, McManus BM (1998). Caspase activation and specific
cleavage of substrates after coxsackievirus B3-induced cytopathic
effect in HeLa cells. J. Virol. 72: 7669-7675.
Cerutti PA, Trump BF (1991). Inflammation and oxidative stress in
carcinogenesis. Cancer Cells 3: 1-7.
Chakrabarti D, Huang X, Beck J, Henrich J, McFarland N, James RF,
Stewart TA (1996). Control of islet intercellular adhesion molecule-1
expression by interferon-alpha and hypoxia. Diabetes 45: 1336-1343.
Chaudhri G, Clark IA, Hunt NH, Cowden WB, Ceredig R (1986). Effect
of antioxidants on primary alloantigen-induced T cell activation and
proliferation. J. Immunol. 137: 2646-2652.
Chehadeh W, Kerr-Conte J, Pattou F, Alm G, Lefebvre J, Wattre P,
Hober D (2000). Persistent infection of human pancreatic islets by
coxsackievirus B is associated with alpha interferon synthesis in beta
cells. J. Virol. 74: 10153-10164.
Chen G, Branton PE, Yang E, Korsmeyer SJ, Shore GC (1996).
Adenovirus E1B 19- Da death suppressor protein interacts with Bax
but not with Bad. J. Biol. Chem. 271: 24221- 24225.
Chiou CC, Liu WT, Chen SJ, Soong WJ, Wu KG, Tang RB, Hwang B
(1998). Coxsackievirus B1 infection in infants less than 2 months of
age. Am. J. Perinatol. 15: 155-159.
Colasanti M, Persichini T, Venturini G, Ascenzi P (1999). S-nitrosylation
of viral proteins: molecular bases for antiviral effect of nitric oxide.
IUBMB Life 48: 25-31.
Conaldi PG, Biancone L, Bottelli A, De Martino A, Camussi G, Toniolo A
(1997a). Distinct pathogenic effects of group B coxsackieviruses on
human glomerular and tubular kidney cells. J. Virol. 71: 9180-9187.
Conaldi PG, Serra C, Mossa A, Falcone V, Basolo F, Camussi G, Dolei
A, Toniolo (1997b). Persistent infection of human vascular endothelial
cells by group B coxsackieviruses. J. Infect. Dis. 175: 693-696.
Dahlquist G, Blom L, Holmgren G, Hagglof B, Larsson Y, Sterky G, Wall
S (1985). The epidemiology of diabetes in Swedish children 0-14
years--a six-year prospective study. Diabetologia 28: 802-808.
Dahlquist GG, Ivarsson S, Lindberg B, Forsgren M (1995). Maternal
enteroviral infection during pregnancy as a risk factor for childhood
IDDM. A population-based case-control study. Diabetes 44: 408-413.
Darnell Jr JE, Kerr IM, Stark GR (1994). Jak-STAT pathways and
transcriptional activation in response to IFNs and other extracellular
proteins. Science 264: 1415–1420.
Datta K, Babbar P, Srivastava T, Sinha S, Chattopadhyay P (2002).
p. 53 dependent apoptosis in glioma cell lines in response to
hydrogen peroxide induced oxidative stress. Int. J. Biochem. Cell
Biol. 34: 148- 157.
Davies JL, Kawaguchi Y, Bennett ST, Copeman JB, Cordell
HJ,Pritchard LE, Reed PW, Gough SC, Jenkins SC, Palmer SM,
Balfour KM, Rowe BR, Farral M, Barnett AH, Bain SC, Todd JA
(1994). A genome-wide search for human type 1 diabetes
susceptibility genes. Nature 371: 130-136.
Davies JM (1997). Molecular mimicry: can epitope mimicry induce
autoimmune disease? Immunol. Cell Biol. 75: 113-126.
Debbas M, White E (1993). Wild-type p53 mediates apoptosis by E1A,
which is inhibited by E1B. Genes Dev. 7: 546-554.
Delaney CA, Pavlovic D, Hoorens A, Pipeleers DG, Eizirik DL (1997).
Cytokines induce deoxyribonucleic acid strand breaks and apoptosis
in human pancreatic islet cells. Endocrinol.138: 2610-2614.
Deneke SM, Fanburg BL (1989). Regulation of cellular glutathione. Am.
J. Physiol. 257: L163- 173.
Di Meo S, Venditti P (2001). Mitochondria in exercise-induced oxidative
stress. Biol. Signals Recept 10: 125-140.
DiGuiseppi J, Fridovich I (1984). The toxicology of molecular oxygen.
Crit. Rev. Toxicol. 12: 315-342.
Dobner T, Horikoshi N, Rubenwolf S, Shenk T (1996). Blockage by
adenovirus E4orf6 of transcriptional activation by the p53 tumor
suppressor. Science 272: 1470-1473.
Dorman JS, Bunker CH (2000). HLA-DQ locus of the human leukocyte
antigen complex and type 1 diabetes mellitus: a HuGE review.
Epidemiol. Rev. 22: 218-227.
Dypbukt JM, Ankarcrona M, Burkitt M, Sjoholm A, Strom K, Orrenius S,
Nicotera P (1994). Different prooxidant levels stimulate growth,
trigger apoptosis, or produce necrosis of insulin-secreting RINm5F
cells. The role of intracellular polyamines. J. Biol. Chem. 269: 30553-
30560.
Eizirik DL, Flodstrom M, Karlsen AE, Welsh N (1996). The harmony of
the pheres: inducible nitric oxide synthase and related genes in
pancreatic beta cells. Diabetologia 39: 875-890.
Eizirik DL, Mandrup-Poulsen T (2001). A choice of death--the signal-
transduction of immunemediated beta-cell apoptosis. Diabetologia
44: 2115-2133.
Fehsel K, Kolb-Bachofen V, Kroncke KD (2003). Necrosis Is the
Predominant Type of Islet Cell Death During Development of Insulin-
Dependent Diabetes Mellitus in BB Rats. Lab Invest 83: 549-559.
Fernandez-Checa JC, Garcia-Ruiz C, Colell A, Morales A, Mari M,
Miranda M, Ardit E (1998). Oxidative stress: role of mitochondria and
protection by glutathione. Biofactors 8: 7-11.
Ferris CD, Jaffrey SR, Sawa A, Takahashi M, Brady SD, Barrow RK,
Tysoe SA, Wolosker H, Baranano DE, Dore S, Poss KD, Snyder SH
Sharma et al. 055
(1999). Haem oxygenase-1 prevents cell death by regulating cellular
iron. Nat. Cell Biol. 1: 152-157.
Feuer R, Mena I, Pagarigan R, Slifka MK, Whitton JL (2002). Cell cycle
status affects coxsackievirus replication, persistence, and reactivation
in vitro. J. Virol. 76: 4430-4440.
Fleury C, Mignotte B, Vayssiere JL (2002). Mitochondrial reactive
oxygen species in cell deathsignaling. Biochimie 84: 131-141.
Forman HJ (1982). Free radicals in biology. Pyor, W.A., editor.
Academic Press New York. p. 65-90.
Foulis AK, Farquharson MA, Cameron SO, McGill M, Schonke H,
Kandolf R (1990). A search for the presence of the enteroviral capsid
protein VP1 in pancreases of patients with type 1 (insulin-dependent)
diabetes and pancreases and hearts of infants who died of
coxsackieviral myocarditis. Diabetologia. 33: 290-8.
Fridovich I (1995). Superoxide radical and superoxide dismutases.
Annu. Rev. Biochem. 64: 97-112.
Fridovich I (1998). Oxygen toxicity: a radical explanation. J. Exp. Biol.
201(Pt 8): 1203-1209.
Fridovich I (1999). Fundamental aspects of reactive oxygen species, or
what's the matter with oxygen? Ann. NY Acad. Sci. 893: 13-18.
Frisk G, Nilsson E, Tuvemo T, Friman G, Diderholm H (1992). The
possible role of Coxsackie A and echo viruses in the pathogenesis of
type I diabetes mellitus studied by IgM analysis. J. Infect. 24: 13-22.
Froelich CJ, Dixit VM, Yang X (1998). Lymphocyte granule-mediated
apoptosis: matters of viral mimicry and deadly proteases. Immonol.
today 19: 30-36.
Gauntt C, Huber S (2003). Coxsackievirus experimental heart diseases.
Front Biosci 8, E23-35. Gepts, W. (1965) Pathologic anatomy of the
pancreas in juvenile diabetes mellitus. Diabetes 14: 619-633.
Girard S, Couderc T, Destombes J, Thiesson D, Delpeyroux F, Blondel
B (1999). Poliovirus induces apoptosis in the mouse central nervous
system. J. Virol. 73: 6066-6072.
Goldstaub D, Gradi A, Bercovitch Z, Grosmann Z, Nophar Y, Luria S,
Sonenberg N, Kahana C (2000). Poliovirus 2A protease induces apo-
ptotic cell death. Mol. Cell Biol. 20: 1271-1277.
Grankvist K, Marklund SL, Taljedal IB (1981). CuZn-superoxide
dismutase, Mn-superoxide dismutase, catalase and glutathione
peroxidase in pancreatic islets and other tissues in the mouse.
Biochem. J. 199: 393-398.
Griffith OW, Mulcahy RT (1999). The enzymes of glutathione synthesis:
gammaglutamylcysteine synthetase. Adv Enzymol Relat. Areas Mol.
Biol. 73: 209-267.
Haber CA, Lam TK, Yu Z (2003). N-acetylcysteine and taurine prevent
hyperglycaemia-induced insulin resistance in vivo possible role of
oxidative stress. Am. J. Physiol. Endocrinol. Metab. 285: 744-753.
Hampton MB, Orrenius S (1997). Dual regulation of caspase activity by
hydrogen peroxide: implications for apoptosis. FEBS Lett 414: 552-
556.
Hanafusa T, Sugihara S, Fujino-Kurihara H, Miyagawa J, Miyazaki A,
Yoshioka T, Yamada K, Nakajima H, Asakawa H, Kono N (1988).
Induction of insulitis by adoptive transfer with L3T4+Lyt2- T-
lymphocytes in T-lymphocyte-depleted NOD mice. Diabetes 37: 204-
208.
Havivi E, Baron H, Reshef A, Stein P, Raz I (1991). Vitamins and trace
metals status in non insulin dependent diabetes. Int. J. Vitam. Nutr.
Res. 61: 328-333.
Heimberg H, Nagy ZP, Somers G, De Leeuw I, Schuit FC (1992).
Complementation of HLA-DQA and -DQB genes confers
susceptibility and protection to insulin-dependent diabetes mellitus.
Hum. Immunol. 33: 10-17.
Heimberg H, Heremans Y, Jobin C, Leemans R, Cardozo AK, Darville
M, Eizirik DL (2001). Inhibition of cytokine-induced NF-kappaB
activation by adenovirus- mediated expression of a NF-kappaB
super-repressor prevents beta-cell apoptosis. Diabetes 50: 2219-
2224.
Hengartner MO (2000). The biochemistry of apoptosis. Nature 407:
770-776.
Hirsch T, Marchetti P, Susin SA, Dallaporta B, Zamzami N, Marzo I,
Geuskens M, Kroemer G (1997). The apoptosis-necrosis paradox.
Apoptogenic proteases activated after mitochondrial permeability
transition determine the mode of cell death. Oncogene 15: 1573-
1581.
056 Sci. Res. Essays
Hoeldtke RD, Bryner KD, McNeill DR, Hobbs GR, Riggs JE, Warehime
SS, Christie I, Ganser G Van Dyke K (2002). Nitrosative stress, uric
Acid, and peripheral nerve function in early type 1 diabetes. Diabetes
51: 2817-2825.
Holmgren A (1985). Thioredoxin. Annu. Rev. Biochem. 54: 237-271.
Hoorens A, Stange G, Pavlovic D, Pipeleers D (2001). Distinction
between interleukin-1-induced necrosis and apoptosis of islet cells.
Diabetes 50: 551-557.
Horwitz MS, Bradley LM, Harbertson J, Krahl T, Lee J, Sarvetnick N
(1998). Diabetes induced by Coxsackie virus: initiation by bystander
damage and not molecular mimicry. NatMed 4: 781-785.
Horwitz MS, Sarvetnick N (1999). Viruses, host responses, and
autoimmunity. Immunol. Rev.169: 241-53.
Horvitz HR (1999). Genetic control of programmed cell death in the
nematode Caenorhabditis elegans. Cancer Res. 59: 1701s-1706s.
Horwitz MS, Ilic A, Fine C, Rodriguez E, Sarvetnick N (2002). Presented
antigen from damaged pancreatic beta cells activates autoreactive T
cells in virus-mediated autoimmune diabetes. J. Clin. Invest. 109: 79-
87.
Hotta M, Tashiro F, Ikegami H, Niwa H, Ogihara T, Yodoi J, Miyazaki J
(1998). Pancreatic beta cell-specific expression of thioredoxin, an
antioxidative and antiapoptotic protein, prevents autoimmune and
streptozotocin-induced diabetes. J. Exp. Med. 188: 1445-1451.
Hunt N (1994). Redox mechanisms in T-cell activation. In: “Oxidative
Stress, Cell Activation and Viral Infection” (eds. C. Pasquier et al.),
Birkhäuser Verlag Basel, Switzerland. pp. 237-251.
Hyypiä T, Hovi T, Knowles NJ, Stanway G (1997). Classification of
enteroviruses based on molecular and biological properties. J. Gen.
Virol. 78: 1-11.
Itagaki A, Ishihara J, Mochida K, Ito Y, Saito K, Nishino Y, Koike
S,Kurimura T (1983). A clustering outbreak of hand, foot, and mouth
disease caused by Coxsackie virus A10. Microbiol. Immunol. 27: 929-
935.
Jarpe AJ, Hickman MR, Anderson JT, Winter WE, Peck AB (1990-
1991). Flow cytometric enumeration of mononuclear cell populations
infiltrating the islets of Langerhans in prediabetic NOD mice:
development of a model of autoimmune insulitis for type I diabetes.
Reg. Immunol. 3: 305-317.
Jeurissen SH, Wagenaar F, Pol JM, van der Eb AJ, Noteborn MH
(1992). Chicken anemia virus causes apoptosis of thymocytes after in
vivo infection and of cell lines after in vitro infection. J. Virol. 66:
7383-7388.
Kadowaki N, Antonenko S, Lau JY, Liu YJ (2000). Natural interferon
alpha/beta-producing cells link innate and adaptive immunity. J. Exp.
Med. 192: 219-226.
Kagi D, Vignaux F, Ledermann B, Burki K, Depraetere V, Nagata S,
Hengartner H, Golstein P (1994). Fas and perforin pathways as major
mechanisms of T cell-mediated cytotoxicity. Sci. 265: 528-530.
Karvonen M, Pitkaniemi J, Tuomilehto J (1999). The onset age of type 1
diabetes in Finnish children has become younger. The Finnish
Childhood Diabetes Registry Group. Diabetes Care. 22: 1066-70.
Kaufman DL, Erlander MG, Clare-Salzler M, Atkinson MA, Maclaren
NK, Tobin AJ (1992). Autoimmunity to two forms of glutamate
decarboxylase in insulin- dependent diabetes mellitus. J. Clin. Invest.
89: 283-292.
Kerr JF, Wyllie AH, Currie AR (1972). Apoptosis: a basic biological
phenomenon with wide ranging implications in tissue kinetics. Br. J.
Cancer 26: 239-257.
Kim KA, Kim S, Chang I, Kim GS, Min YK, Lee MK, Kim KW, Lee MS
(2002). IFN gamma/TNF alpha synergism in MHC class II induction:
effect of nicotinamide on MHC class II expression but not on islet-cell
apoptosis. Diabetologia 45: 385-393.
Kirkman HN, Gaetani GF (1984). Catalase: a tetrameric enzyme with
four tightly bound molecules of NADPH. Proc. Natl. Acad. Sci. USA.
81: 4343-4347.
Kitamura N, Semler BL, Rothberg PG, Larsen GR, Adler CJ, Dorner AJ,
Emini EA, Hanecak R, Lee JJ, van der Werf S, Anderson CW,
Wimmer E (1981). Primary structure, gene organization and
polypeptide expression of poliovirus RNA. Nature 291: 547-553.
Koch F, Koch G (1985). Morphological alterations of the host cell as an
essential basis for poliovirus replication. In "The Molecular Biology of
Poliovirus" (eds. F.Koch and G.Koch), Springer-Verlag, Wien-New
York. pp. 226-266
Korant BD (1973). Cleavage of poliovirus-specific polypeptide
aggregates. J. Virol. 12: 556-563.
Kuyumcu-Martinez NM, Joachims M, Lloyd RE (2002). Efficient
cleavage of ribosome associated poly(A)-binding protein by ente-
rovirus 3C protease. J. Virol. 76: 2062-74.
Lansdown AB (1976). Pathological changes in the pancreas of mice
following infection with coxsackie B viruses. Br. J. Exp. Pathol. 57:
331-338.
Larsson R, Cerutti P (1988). Oxidants induce phosphorylation of
ribosomal protein S6. J. BiolChem. 263: 17452-17458.
Lemasters JJ (1999). V. Necrapoptosis and the mitochondrial
permeability transition: shared pathways to necrosis and apoptosis.
Am. J. Physiol. 276: G1-6.
Lennon SV, Martin SJ, Cotter TG (1991). Dose-dependent induction of
apoptosis in human tumour cell lines by widely diverging stimuli. Cell
Prolif 24: 203-214.
Lenzen S, Panten U (1988). Alloxan: history and mechanism of action.
Diabetologia 31: 337-342.
Levy-Marchal C, Patterson C, Green A (1995). Variation by age group
and seasonality at diagnosis of childhood IDDM in Europe. The
EURODIAB ACE Study Group. Diabetologia 38: 823-830.
Liu D, Cardozo AK, Darville MI, Eizirik DL (2002). Double-stranded RNA
cooperates with interferon-gamma and IL-1 beta to induce both
chemokine expression and nuclear factorkappa B-dependent
apoptosis in pancreatic beta-cells: potential mechanisms for viral-
induced insulitis and beta-cell death in type 1 diabetes mellitus.
Endocrinology 143: 1225-1234.
Liu X, Zou H, Slaughter C, Wang X (1997). DFF, a heterodimeric
protein that functions downstream of caspase-3 to trigger DNA
fragmentation during apoptosis. Cell 89: 175-84.
Livingstone C, Davis J (2007). Review: Targeting therapeutics against
glutathione depletion in diabetes and its complications. Br. J. Diab.
Vasc. Dis. 7: 258-265.
Lopez-Guerrero JA, Alonso M, Martin-Belmonte F, Carrasco L (2000).
Poliovirus induces apoptosis in the human U937 promonocytic cell
line. Virology 272: 250-256.
Lortz S, Tiedge M, Nachtwey T, Karlsen AE, Nerup J, Lenzen S (2000).
Protection of insulin-producing RINm5F cells against cytokine-
mediated toxicity through overexpression of antioxidant enzymes.
Diabetes 49: 1123-1130.
Maeda H, Sawa T, Yubisui T, Akaike T (1999). Free radical generation
from heterocyclic amines by cytochrome b5 reductase in the
presence of NADH. Cancer Lett. 143: 117-121.
Maines MD (1997). The heme oxygenase system: a regulator of second
messenger gases. Annu. Rev. Pharmacol. Toxicol. 37: 517-554.
Mandrup-Poulsen T (1996). The role of interleukin-1 in the
pathogenesis of IDDM. Diabetologia 39: 1005-1029.
Marselli L, Marchetti P, Tellini C, Giannarelli R, Lencioni C, Del Guerra
S, Lupi R, Carmellini M, Mosca F, Navalesi R (2000). Lymphokine
release from human lymphomononuclear cells after co-culture with
isolated pancreatic islets: effects of islet species, long-term culture,
and monocyte-macrophage cell removal. Cytokine 12: 503-505.
Maxwell SR, Thomason H, Sandler D, Leguen C, Baxter MA, Thorpe
GH, Jones AF, Barnett AH (1997) Antioxidant status in patients with
uncomplicated insulin-dependent and non-insulin-dependent diabetes
mellitus. Eur. J. Clin. Invest. 27: 484-490.
May JM, Qu ZC, Whitesell RR (1995). Ascorbate is the major electron
donor for a transmembrane oxidoreductase of human erythrocytes.
Biochim. Biophys. Acta. 1238: 127-136.
May JM, Qu ZC, Whitesell RR, Cobb CE (1996). Ascorbate recycling in
human erythrocytes: role of GSH in reducing dehydroascorbate. Free
Radic. Biol. Med. 20: 543-551.
Melnick JL (1996). Current status of poliovirus infections. Clin.
Microbiol. Rev. 9: 293-300.
Mendiratta S, Qu ZC, May JM (1998). Enzyme-dependent ascorbate
recycling in human erythrocytes: role of thioredoxin reductase. Free
Radic. Biol. Med. 25: 221-228.
Miller BJ, Appel MC, O'Neil JJ, Wicker LS (1988). Both the Lyt-2+ and
L3T4+ T cell subsets are required for the transfer of diabetes in
nonobese diabetic mice. J. Immunol. 140: 52-58.
Molina TJ, Kishihara K, Siderovski DP, van Ewijk W, Narendran A,
Timms E, Wakeham A, Paige CJ, Hartmann KU, Veillette A, Davidson
D, Mak TW (1992). Profound block in thymocyte development in mice
lacking p56lck. Nature 357: 161-164.
Moriwaki M, Itoh N, Miyagawa J, Yamamoto K, Imagawa A, Yamagata
K, Iwahashi H, Nakajima H, Namba M, Nagata S, Hanafusa T,
Matsuzawa Y (1999). Fas and Fas ligand expression in inflamed
islets in pancreas sections of patients with recent-onset Type I
diabetes mellitus. Diabetologia 42: 1332-1340.
Muir P, van Loon AM (1997). Enterovirus infections of the central
nervous system. Intervirol. 40: 153-166.
Nicholson DW, Thornberry NA (1997). Caspases: killer proteases.
Trends Biochem. Sci. 22: 299-306.
Nishikawa T, Edelstein D, Du XL, Yamagishi S, Matsumura T, Kaneda
Y, Yorek MA, Beebe D, Oates PJ, Hammes HP, Giardino I, Brownlee
M (2000). Normalizing mitochondrial superoxide production blocks
three pathways of hyperglycaemic damage. Nature 404: 787-790.
Oda T, Akaike T, Hamamoto T, Suzuki F, Hirano T, Maeda H (1989).
Oxygen radicals in influenza-induced pathogenesis and treatment
with pyran polymer-conjugated SOD. Science 244: 974-976.
Ohba M, Shibanuma M, Kuroki T, Nose K (1994). Production of
hydrogen peroxide by transforming growth factor-beta 1 and its
involvement in induction of egr-1 in mouse osteoblastic cells. J. Cell
Biol. 126: 1079-1088.
Op De Beeck A, Caillet-Fauquet P (1997). Viruses and the cell cycle.
Prog. Cell Cycle Res. 3: 1-19.
Opavsky MA, Martino T, Rabinovitch M, Penninger J, Richardson C,
Petric M, Trinidad C, Butcher, L, Chan J, Liu PP (2002). Enhanced
ERK-1/2 activation in mice susceptible to coxsackievirus- induced
myocarditis. J. Clin. Invest. 109: 1561-1569.
Padaiga Z, Tuomilehto J, Karvonen M, Podar T, Brigis G, Urbonaite B,
Kohtamaki K, Lounamaa R, Tuomilehto-Wolf E, Reunanen A (1997).
Incidence trends in childhood onset IDDM in four countries around
the Baltic sea during 1983-1992. Diabetologia 40: 187-192.
Perl A, Gergely P Jr Puskas F, Banki K (2002). Metabolic switches of T-
cell activation and apoptosis. Antioxid Redox Signal. 4: 427-443.
Peterhans E, Grob M, Burge T, Zanoni R (1987). Virus-induced
formation of reactive oxygenintermediates in phagocytic cells. Free
Radic. Res. Commun. 3: 39-46.
Petronilli V, Costantini P, Scorrano L, Colonna R, Passamonti S,
Bernardi P (1994). The voltage sensor of the mitochondrial
permeability transition pore is tuned by the oxidationreduction state of
vicinal thiols. Increase of the gating potential by oxidants and its
reversal by reducing agents. J. Biol. Chem. 269: 16638-16642.
Pileggi A, Molano RD, Berney T, Cattan P, Vizzardelli C, Oliver R,
Fraker C, Ricordi C, Pastori RL, Bach FH, Inverardi L (2001). Heme
oxygenase-1 induction in islet cells results in protection from
apoptosis and improved in vivo function after transplantation.
Diabetes 50: 1983-1991.
Pinkoski MJ, Waterhouse NJ, Heibein JA, Wolf BB, Kuwana T,
Goldstein JC, Newmeyer DD, Bleackley RC, Green DR (2001).
Granzyme B-mediated apoptosis proceeds predominantly through a
Bcl-2-inhibitable mitochondrial pathway. J. Biol. Chem. 276: 12060-
12067.
Pipeleers D, Hoorens A, Marichal-Pipeleers M, Van de Casteele M,
Bouwens L, Ling Z (2001). Role of pancreatic beta-cells in the
process of beta-cell death. Diabetes 50(1): S52-57.
Platanias LC, Uddin S, Domanski P, Colamonici OR (1996). Differences
in signaling between IFNs and : IFN selectively induces the
interaction of the and L subunits of the Type I IFNR. J. Biol. Chem.
271: 23630–23633.
Pociot F, McDermott MF (2002). Genetics of type 1 diabetes mellitus.
Genes Immun 3: 235-249.
Polla BS, Jacquier-Sarlin MR, Kantengwa S, Mariethoz E, Hennet T,
Russo-Marie F, Cossarizza A (1996). TNF alpha alters mitochondrial
membrane potential in L929 but not in TNF alpha-resistant L929.12
cells: relationship with the expression of stress proteins, annexin 1
and superoxide dismutase activity. Free Radic. Res. 25: 125-131.
Rabinovitch A, Suarez-Pinzon WL, Strynadka K, Lakey JR, Rajotte RV
(1996b). Human pancreatic islet beta-cell destruction by cytokines
involves oxygen free radicals and aldehyde production. J. Clin.
Endocrinol. Metab. 81: 3197-3202.
Rabinovitch A, Suarez-Pinzon WL (1998). Cytokines and their roles in
Sharma et al. 057
pancreatic islet beta-cell destruction and insulin-dependent diabetes
mellitus. Biochem. Pharmacol. 55: 1139-1149.
Racaniello VR (2001). Picornaviridae: the viruses and their replication.
In: Knipe DM, Howley PM, Griffin DE, Lamb RA, Martin MA, Roizman
B, Straus SE (eds.) Fields Virol., 4th ed, Lippincott Williams and
Wilkins,Philadelphia, USA. pp. 665-672.
Ramsingh AI (1997b). Coxcakieviruses and pancreatitis. Front. Biosci.
2: 53-62.
Ribeiro MM, Klein D, Pileggi A, Molano RD, Fraker C, Ricordi C.
Inverardi L, Pastori RL (2003). Hemeoxygenase-1 fused to a TAT
peptide transduces and protects pancreatic beta-cells. Biochem.
Biophys. Res. Commun.305: 876-881.
Robertson RP (2004). Chronic oxidative stress as a central mechanism
for glucose toxicity in pancreatic islet beta cells in diabetes. J. Biol.
Chem. 279: 42351-42354.
Robertson RP, Harmon J, Tran PO, Tanaka Y, Takahashi H (2003).
Glucose toxicity in pancreatic islets beta cells: type-2 diabetes, good
radicals gone bad and the glutathione connection. Diabetes 52: 581-
587.
Rodriguez J, Lazebnik Y (1999). Caspase-9 and APAF-1 form an
activeholoenzyme. Genes Dev. 13: 3179-3184.
Roep BO, Hiemstra HS, Schloot NC, De Vries RR, Chaudhuri A, Behan
PO, Drijfhout JW (2002). Molecular mimicry in type 1 diabetes:
immune cross-reactivity between islet autoantigen and human
cytomegalovirus but not Coxsackie virus. Ann. NY Acad. Sci. 958:
163-165.
Roivainen M, Piirainen L, Hovi T, Virtanen I, Riikonen T, Heino J,
Hyypia T (1994). Entry of coxsackievirus A9 into host cells: specific
interactions with alpha v beta 3 integrin, the vitronectin receptor.
Virology 203: 357-365.
Roivainen M, Ylipaasto P, Savolainen C, Galama J, Hovi T, Otonkoski T
(2002). Functional impairment and killing of human beta cells by
enteroviruses: the capacity is shared by a wide range of serotypes,
but the extent is a characteristic of individual virus strains.
Diabetologia 45: 693-702.
Romero-Alvira D, Roche E (1998). The keys of oxidative stress in
acquired immune deficiency syndrome apoptosis. Med. Hypotheses
51: 169-173.
Ronco LV, Karpova AY, Vidal M, Howley PM (1998). Human
papillomavirus 16 E6 oncoprotein binds to interferon regulatory
factor-3 and inhibits its transcriptional activity. Genes Dev. 12: 2061-
2072.
Rotbart HA (2002). Treatment of picornavirus infections. Antiviral Res.
53: 83-98.
Saitoh M, Nishitoh H, Fujii M, Takeda K, Tobiume K, Sawada Y,
Kawabata M, Miyazono K, Ichijo H (1998). Mammalian thioredoxin is
a direct inhibitor of apoptosis signal regulating kinase (ASK) 1. EMBO
J. 17: 2596-2606.
Samali A, Nordgren H, Zhivotovsky B, Peterson E, Orrenius S (1999). A
comparative study of apoptosis and necrosis in HepG2 cells: oxidant-
induced caspase inactivation leads to necrosis. Biochem. Biophys.
Res. Commun. 255: 6-11.
Santini SA, Marra G, Giardina B, Cotroneo P, Mordente A, Martorana
GE, Manto A, Ghirlanda G (1997). Defective plasma antioxidant
defenses and enhanced susceptibility to lipid peroxidation in
uncomplicated IDDM. Diabetes 46: 1853-1858.
Santos CX, Anjos EI, Augusto O (1999). Uric acid oxidation by
peroxynitrite: multiple reactions, free radical formation, and
amplification of lipid oxidation. Arch. Biochem. Biophys. 372: 285-
294.
Sawada M, Nakashima S, Kiyono T, Nakagawa M, Yamada J,
Yamakawa H, Banno Y, Shinoda J, Nishimura Y, Nozawa Y, Sakai N
(2001). Regulates ceramide formation by neutral sphingomyelinase
through reactive oxygen species in human glioma cells. Oncogene
20: 1368-1378.
Schloot NC, Willemen SJ, Duinkerken G, Drijfhout JW, de Vries RR,
Roep BO (2001). Molecular mimicry in type 1 diabetes mellitus
revisited: T-cell clones to GAD65 peptides with sequence homology
to Coxsackie or proinsulin peptides do not crossreact with
homologous counterpart. Hum. Immunol. 62: 299-309.
Schranz DB, Lernmark A (1998). Immunology in diabetes: an update.
Diabetes Metab. Rev. 14: 3-29.
058 Sci. Res. Essays
Schwarz KB (1996). Oxidative stress during viral infection: A review.
Free Radic. Biol. Med. 21: 641-649.
Seghieri G, Martinoli L, di Felice M, Anichini R, Fazzini A, Ciuti M, Miceli
M, Gaspa L, Franconi F (1998). Plasma and platelet ascorbate pools
and lipid peroxidation in insulindependent diabetes mellitus. Eur. J.
Clin. Invest. 28: 659-663.
Sies H, Stahl W, Sundquist AR (1992). Antioxidant functions of vitamins.
Vitamins E and C, beta-carotene, and other carotenoids. Ann. NY.
Acad. Sci. 669: 7-20.
Skinner KA, White CR, Patel R, Tan S, Barnes S, Kirk M, Darley-Usmar
V, Parks DA (1998). Nitrosation of uric acid by peroxynitrite.
Formation of a vasoactive nitric oxide donor. J. Biol. Chem. 273:
24491-24497.
Slee EA, Adrain C, Martin SJ (2001). Executioner caspase-3, -6, and -7
perform distinct, nonredundant roles during the demolition phase of
apoptosis. J. Biol. Chem. 276: 7320–7326.
Sofic E, Rustembegovic A, Kroyer G, Cao G (2002). Serum
antioxidantcapacity in neurological, psychiatric, renal diseases and
cardiomyopathy. J. Neural Transm. 109: 711- 719.
Somersalo O (1955). Ann. Paediatr. Fenn. 1: 239-249.
Spector A, Yan GZ, Huang RR, McDermott MJ, Gascoyne PR, Pigiet V
(1988). The effect of H2O2 upon thioredoxin-enriched lens epithelial
cells. J. Biol. Chem. 263: 4984-4990.
Stadtman ER, LevineRL (2000). Protein oxidation. Ann. NY Acad. Sci.
899: 191-208.
Suarez-Pinzon WL, Szabo C, Rabinovitch A (1997). Development of
autoimmune diabetes in NOD mice is associated with the formation
of peroxynitrite in pancreatic islet beta-cells. Diabetes 46: 907-911.
Sumida T, Furukawa M, Sakamoto A, Namekawa T, Maeda T, Zijlstra
M, Iwamoto I, Koike T, Yoshida S, Tomioka H, Taniguchi M (1994).
kPrevention of insulitis and diabetes in beta 2- microglobulin-deficient
non-obese diabetic mice. Int. Immunol. 6: 1445-1449.
Sundaresan M, Yu ZX, Ferrans VJ, Sulciner DJ, Gutkind JS, Irani K,
Goldschmidt-Clermont PJ Finkel T (1996). Regulation of reactive-
oxygen-species generation in fibroblasts by Rac1. Biochem. J. 318
(Pt 2): 379-382.
Susin SA, Lorenzo HK, Zamzami N, Marzo I, Snow BE, Brothers GM,
Mangion J, Jacotot E, Costantini P, Loeffler M, Larochette N,
Goodlett DR, Aebersold R, Siderovski DP, Penninger JM Kroemer G
(1999). Molecular characterization of mitochondrial apoptosis-
inducing factor. Nature 397: 441-446.
Swanton C, Jones N (2001). Strategies in subversion: de-regulation of
the mammalian cell cycle by viral gene products. Int. J. Exp. Pathol.
82: 3-13.
Tanaka N, Sato M, Lamphier MS, Nozawa H, Oda E, Noguchi S,
Schreiber RD, Tsujimoto Y, Taniguchi T (1998). Type I interferons
are essential mediators of apoptotic death in virally infected cells.
Genes Cells 3: 29-37.
Tartaglia LA, Ayres TM, Wong GH, Goeddel DV (1993). A novel domain
within the 55 kd TNF receptor signals cell death. Cell 74: 845-853.
Teodoro JG, Branton PE (1997). Regulation of apoptosis by viral gene
products. J. Virol. 71: 1739-1746.
Thorsby E, Ronningen KS (1993). Particular HLA-DQ molecules play a
dominant role in determining susceptibility or resistance to type 1
(insulin-dependent) diabetes mellitus. Diabetologia 36: 371-377.
Tiedge M, Lortz S, Munday R, Lenzen S (1998). Complementary action
of antioxidant enzymes in the protection of bioengineered insulin-
producing RINm5F cells against the toxicity of reactive oxygen
species. Diabetes 47: 1578-1585.
Tolskaya EA, Romanova LI, Kolesnikova MS, Ivannikova TA, Smirnova
EA, Raikhlin NT, Agol VI (1995). Apoptosis-inducing and apoptosis-
preventing functions of poliovirus. J. Virol. 69: 1181-1189.
Tracy S, Hofling K, Pirruccello S, Lane PH, Reyna SM, Gauntt CJ
(2000). Group B coxsackievirus myocarditis and pancreatitis:
connection between viral virulence phenotypes in mice. J. Med. Virol.
62: 70-81.
Tsai EC, Hirsch IB, Brunzell JD, Chait A (1994). Reduced plasma
peroxyl radical trapping capacity and increased susceptibility of LDL
to oxidation in poorly controlled IDDM. Diabetes 43: 1010-1014.
Tuomilehto J, Lounamaa R, Tuomilehto-Wolf E, Reunanen A, Virtala E,
Kaprio EA, Akerblom HK (1992). Epidemiology of childhood diabetes
mellitus in Finland—background of a nationwide study of type 1
(insulin-dependent) diabetes mellitus. The Childhood Diabetes in
Finland (DiMe) Study Group. Diabetologia 35: 70-76.
VanderJagt DJ, Harrison JM, Ratliff DM, Hunsaker LA, Vander Jagt DL
(2001). Oxidative stress indices in IDDM subjects with and without
long-term diabetic complications. Clin. Biochem. 34: 265-270.
Vaux DL, Weissman IL, Kim SK (1992). Prevention of programmed cell
death in Caenorhabditis elegans by human bcl-2. Sci. 258: 1955-
1957.
Vessby J, Basu S, Mohsen R, Berne C, Vessby B (2002). Oxidative
stress and antioxidant status in type 1 diabetes mellitus. J. Int. Med.
251: 69-76.
Viskari, H.R., Roivainen, M., Reunanen, A., Pitkaniemi, J., Sadeharju,
K., Koskela, P., Hovi,T., Leinikki P, Vilja P, Tuomilehto J, Hyoty H
(2002). Maternal first-trimester enterovirus infection and future risk of
type 1 diabetes in the exposed fetus. Diabetes. 51: 2568-2571.
Vreugdenhil GR, Geluk A, Ottenhoff TH, Melchers WJ, Roep BO,
Galama JM (1998). Molecular mimicry in diabetes mellitus: the
homologous domain in coxsackie B virus protein 2C and islet
autoantigen GAD65 is highly conserved in the coxsackie B-like
enteroviruses and binds to the diabetes associated HLA-DR3
molecule. Diabetologia 41: 40-46.
Whiteman M, Halliwell B (1996). Protection against peroxynitrite-
dependent tyrosine nitration and alpha 1-antiproteinase inactivation
by ascorbic acid. A comparison with other biological antioxidants.
Free Radic. Res. 25: 275-283.
Wicker LS, Leiter EH, Todd JA, Renjilian RJ, Peterson E, Fischer PA,
Podolin PL, Zijlstra M, Jaenisch R, Peterson LB (1994). Beta 2-
microglobulin-deficient NOD mice do not develop insulitis or diabetes.
Diabetes 43: 500-504.
Wong FS, Visintin I, Wen L, Flavell RA, Janeway CA Jr (1996). CD8 T
cell clones from young nonobese diabetic (NOD) islets can transfer
rapid onset of diabetes in NOD mice in the absence of CD4 cells. J.
Exp. Med. 183: 67-76.
Xie B, Zhou JF, Lu Q, Li CJ, Chen P (2002a). Oxidative stress in
patients with acute coxsackie virus myocarditis. Biomed. Environ. Sci.
15: 48-57.
Yoon JW, Austin M, Onodera T, Notkins AL (1979). Isolation of a virus
from the pancreas of a child with diabetic ketoacidosis. N. Engl. J.
Med. 300: 1173-1179.
Zou H, Li Y, Liu X, Wang X (1999). An APAF-1.cytochrome c multimeric
complex is a functional apoptosome that activates procaspase-9. J.
Biol. Chem. 274: 11549-11556.