ArticlePDF Available

Photodynamic treatment of adenoviral vectors with visible light: An easy and convenient method for viral inactivation

Authors:
  • Qia Consultancy & Training / DKFZ Heidelberg

Abstract and Figures

Recombinant adenovirus vectors are popular tools for gene transfer and gene therapy. However biosafety constraints require that all handling of the vectors and vector-containing samples is restricted to dedicated containment laboratories, unless they had undergone a validated virus-inactivation procedure, which decontaminates the samples from any active virus. In this study we evaluated the feasibility of photodynamic treatment (PDT) with visible light to inactivate recombinant adenovirus vectors in biological samples, with minimum associated effects on other biological activities. Several photosensitizers were tested for their capacity to inactivate a model human adenovirus vector, AdCMVLuc, upon illumination. Four photosensitizers (methylene blue (MB), rose bengal (RB), uroporphyrin (UP) and aluminum phthalocynine tetrasulphonate (AIPcS4)) could inactivate the adenovirus, as measured by expression of the luciferase reporter gene and by plaque assay. Of these, MB demonstrated to be the most effective sensitizer in phosphate-buffered saline (PBS), giving > 7 log10 inactivation of the adenovirus. DNA isolated from MB- and light-treated virions was inefficient as a template for transcription. Furthermore, Southern blot analysis revealed fragmentation of the viral DNA. Based on its preference for DNA, MB is suited for adenovirus inactivation in blood plasma. Spiking experiments in which AdCMVLuc was added to plasma samples demonstrated a reduction (> 4 log10-fold) of reporter gene expression to almost background levels. In contrast to MB, photodynamic treatment with RB, UP or AIPcS4 did not lead to DNA damage. Although alterations of the viral capsid could not be detected, the binding pattern of the particles to target cells was significantly changed. Taken together, our data demonstrate that PDT is an efficient, convenient and useful method for the inactivation of adenovirus vectors in biological samples.
Content may be subject to copyright.
Gene Therapy (1999) 6, 873–881
1999 Stockton Press All rights reserved 0969-7128/99 $12.00
http://www.stockton-press.co.uk/gt
Photodynamic treatment of adenoviral vectors with
visible light: an easy and convenient method for viral
inactivation
FHE Schagen, ACE Moor, SC Cheong, SJ Cramer, H van Ormondt, AJ van der Eb,
TMAR Dubbelman and RC Hoeben
Department of Molecular Cell Biology, Leiden University Medical Center, Leiden, The Netherlands
Recombinant adenovirus vectors are popular tools for sensitizer in phosphate-buffered saline (PBS), giving
7
gene transfer and gene therapy. However biosafety con- log
10
inactivation of the adenovirus. DNA isolated from MB-
straints require that all handling of the vectors and vector- and light-treated virions was inefficient as a template for
containing samples is restricted to dedicated containment transcription. Furthermore, Southern blot analysis revealed
laboratories, unless they had undergone a validated virus- fragmentation of the viral DNA. Based on its preference
inactivation procedure, which decontaminates the samples for DNA, MB is suited for adenovirus inactivation in blood
from any active virus. In this study we evaluated the feasi- plasma. Spiking experiments in which AdCMVLuc was
bility of photodynamic treatment (PDT) with visible light to added to plasma samples demonstrated a reduction (
4
inactivate recombinant adenovirus vectors in biological log
10
-fold) of reporter gene expression to almost back-
samples, with minimum associated effects on other biologi- ground levels. In contrast to MB, photodynamic treatment
cal activities. Several photosensitizers were tested for their with RB, UP or AlPcS
4
did not lead to DNA damage.
capacity to inactivate a model human adenovirus vector, Although alterations of the viral capsid could not be
AdCMVLuc, upon illumination. Four photosensitizers detected, the binding pattern of the particles to target cells
(methylene blue (MB), rose bengal (RB), uroporphyrin (UP) was significantly changed. Taken together, our data
and aluminum phthalocynine tetrasulphonate (AlPcS
4
)) demonstrate that PDT is an efficient, convenient and useful
could inactivate the adenovirus, as measured by method for the inactivation of adenovirus vectors in
expression of the luciferase reporter gene and by plaque biological samples.
assay. Of these, MB demonstrated to be the most effective
Keywords:
photo-inactivation; photodynamic treatment; blood plasma
Introduction
In the past decade, recombinant adenovirus vectors have
become very popular tools for in vivo gene transfer into
animals and more recently, for gene therapy purposes in
patients. Many vaccination and gene therapy studies
have borne out the relative safety of adenoviral vectors
for human application.
1–3
However, biosafety constraints
require that all handling of vectors and vector-containing
samples or materials is restricted to dedicated contain-
ment laboratories, unless the material has been exempted
and permitted to be released from the biosafety contain-
ment. Such a release is possible only if one has met two
requirements: the material is produced under GMP con-
ditions and the vector lot to be released is characterized
extensively, viz free of replication-competent adeno-
viruses.
Obviously, this cannot be achieved for all vector virus
lots used in a research setting. Thus, adequate and vali-
dated adenovirus inactivation procedures are needed in
Correspondence: RC Hoeben, Applied Virology Group, Department of
Molecular Cell Biology, Wassenaarseweg 72, 2333 AL Leiden, The
Netherlands
Received 12 August 1998; accepted 21 December 1998
order to decontaminate biological samples fully to allow
transfer to non-containment research facilities. In this
study, we have evaluated the use of photodynamic treat-
ment with photosensitizers which can be excited by vis-
ible white light, for the inactivation of adenoviruses in
general and in the context of plasma with the mainte-
nance of biological activities in particular.
So far, photodynamic inactivation has been proven to
be a powerful method for inactivating enveloped viruses,
such as murine retroviral vectors, human immunodefi-
ciency viruses (HIV-1 and -2), hepatitis-B and -C and ves-
icular stomatitis virus (VSV).
4–8
The procedures were also
effective in the presence of blood or blood plasma. In this
study, we demonstrate that also a non-enveloped virus,
the adenovirus, can be effectively inactivated with PDT.
One of the agents tested, MB, can be used specifically to
degrade the viral DNA. Our data show that PDT can be
an efficient, convenient and useful technique for the inac-
tivation of adenovirus vectors in biological samples.
Beside this decontamination purpose, PDT with MB can
also be used to study the effect of invading adenoviral
particles on recipient cells, since the photodynamically
inactivated adenoviruses are internalized, but have lost
the ability to express viral genes.
Photodynamic inactivation of adenovirus
FHE Schagen
et al
874
Results
Viral inactivation
Photoinactivation of human adenoviruses was studied
using AdCMVLuc, a recombinant adenovirus that carries
the firefly luciferase gene as a reporter. Initially, six
photosensitizers were tested: methylene blue (MB),
protoporphyrin (PP), rose bengal (RB), uroporphyrin
(UP), aluminum phthalocyanine (AlPc) and aluminum
phthalocyanine tetrasulfonate (AlPcS
4
). The effectiveness
was assessed by exposing HepG2 cells to the treated
virus samples and measuring the expression of the
reporter gene. From this series of photosensitizers, MB,
RB, UP and AlPcS
4
tremendously reduced the reporter
gene expression, without affecting the viability of the
cells (data not shown). To optimize the exposure con-
ditions, various sensitizer concentrations and illumi-
nation periods were tested at a light intensity of
106 mW/cm
2
. As shown in Figure 1, all four photosensiti-
zers were able to inactivate AdCMVLuc upon illumi-
nation, while no effect was seen when the samples were
protected from light. The degree of inactivation was
Figure 1 Luciferase activity as a measure of AdCMVLuc infectivity after PDT with MB, RB, UP or AlPcS
4
. HepG2 cells were infected with photoinactiv-
ated AdCMVLuc virus, lysed after 2 days and luciferase activity was measured. Open symbols represent samples exposed to light, filled symbols depict
samples shielded from light. Infectivity of AdCMVLuc after treatment with (a) 2.7
m
MB (circles), 1.3
m
MB (squares), 0.7
m
MB (triangles) and
0.3
m
MB (diamonds); (b) 20
m
RB (circles), 10
m
RB (squares), 5
m
RB (triangles) and 2
m
RB (diamonds); (c) 50
m
UP (circles), 20
m
UP (squares), 10
m
UP (triangles) and 2
m
UP (diamonds); (d) 50
m
AlPcS
4
(circles), 20
m
AlPcS
4
(squares), 10
m
AlPcS
4
(triangles) and
1
m
AlPcS
4
(diamonds). The background activity in the luciferase assay was 200 light units.
clearly dependent on the photosensitizer used, its con-
centration and on the applied illumination. A concen-
tration of 10
8
p.f.u./ml AdCMVLuc, giving a luciferase
activity of 5 × 10
7
light units was completely inactivated
upon a 1-min illumination in 2.7 m MB (Figure 1a). The
same holds for 1.3 m MB, although here a 10-min illumi-
nation period was necessary for complete inactivation.
With either 10 m or 20 m of the photosensitizer RB, a
5 log
10
decrease in luciferase activity was reached only
after an illumination period of 20 min (see Figure 1b).
Figure 1c depicts the results obtained with UP: a 50 m
concentration and 30-min illumination were needed for
complete inactivation. Although AlPcS
4
can significantly
inactivate AdCMVLuc, even upon treatment with 50 m
ALPcS
4
and a 30-min illumination inactivation is
incomplete.
To verify that the luciferase data reflect the viral titer,
plaque assays were performed in parallel (Table 1). The
results of the plaque assays were in good agreement with
the results of the luciferase assay. Inactivation was com-
plete, when 10
8
p.f.u./ml AdCMVLuc were treated with
MB, RB or UP, whereas 50 m AlPcS
4
and 30-min illumi-
Photodynamic inactivation of adenovirus
FHE Schagen
et al
875
Table 1 AdCMVLuc titers after photodynamic treatment
Photosensitizer Concentration Illumination AdCMVLuc titer
(
m
) period (min) (p.f.u./ml)
MB 1.3 0 4 × 10
7
5ND
0.7 5 7 × 10
2
10 ND
RB 20 0 4 × 10
7
51× 10
2
10 10 ND
10 1 × 10
3
20 ND
UP 50 0 4 × 10
7
10 6 × 10
2
30 ND
AlPcS
4
50 0 5 × 10
7
20 5 × 10
5
30 3 × 10
3
ND, not detectable (10
2
).
nation were insufficient for inactivating 10
8
p.f.u./ml
completely. In the unilluminated control samples, the
virus titer decreased no more than two-fold with each of
the photosensitizers.
DNA integrity
Adenoviruses are double-stranded DNA viruses coated
by an icosahedral protein capsid. Consequently, the pho-
tosensitizers can exert their inactivating effect at the viral
DNA or the protein capsid. To study the integrity of the
adenoviral genome, DNA was isolated from illuminated
virions and analyzed by transfection and Southern
blotting.
AdCMVLuc was treated with the various photosensiti-
zers during illumination periods sufficient to inactivate
the virus completely. Subsequently, the intactness of the
isolated viral DNA was evaluated by means of the
luciferase expression in 911 cells. As depicted in Figure 2,
MB-based photoinactivation abolished the luciferase
Figure 2 Luciferase activity upon transfection of DNAfrom photoinactiv-
ated AdCMVLuc. 911 Cells were transfected with 2.1 ng AdCMVLuc
DNA isolated from virus particles, which had been illuminated for 15 min
in the case of MB (1.3
m
), 20 min in the case of RB (20
m
) or 30 min
in the case of UP (50
m
) or AlPcS
4
(50
m
). Black bars represent the
unexposed samples, white bars the exposed ones.
expression completely. This is in contrast with the other
photosensitizers, which hardly affected, if at all, the
capacity of the DNA to serve as template for transcrip-
tion. Transfection with RB- or UP-treated unilluminated
DNA samples resulted even in a slightly lower luciferase
activity than their illuminated counterparts. The controls
without photosensitizer did not show any effect of light
exposure. Similar results were obtained in HepG2 cells
(data not shown).
The integrity of the extracted DNA from light-treated
samples was further studied by Southern blot analysis.
As shown in Figure 3, all DNA samples yielded a sharp
banding pattern as a result of the PstI digestion, except
the DNA isolated from virus photoinactivated with
1.3 m MB and 15-min illumination. This sample showed
a smear, indicative of random DNA fragmentation.
Electromicroscopic analysis
To study whether the phototreatment led to gross disrup-
tion of the virions, aliquots of 5.0 × 10
8
p.f.u. AdCMVLuc
were photoinactivated and the viral particles were vis-
ualized by electron microscopy. No major differences
were seen between the protein capsids or overall struc-
tures of illuminated or unilluminated virus particles or
between any of the four photosensitizers (see Figure 4).
Close inspection of MB and illuminated versus unillumi-
nated virions showed however a slight effect on the
organized structure in the capsid, although its regular
icosahedral form is completely preserved (Figure 4a and
b). Apart from complete virions, additional structures can
be seen in the AlPcS
4
-exposed samples (Figure 4c and d).
These structures, possibly capsid fragments, are present
in similar quantities in illuminated and unilluminated
samples and therefore, are presumably not a result of the
activated photosensitizers.
Binding pattern
To assess changes in the binding to and/or entry into the
cell, binding assays using
3
H-labelled AdCMVLuc were
performed. Upon addition of unilluminated radio-
labelled virus to HeLa cells, approximately 40% of the
Figure 3 Southern blot analysis of photodynamically inactivated
AdCMVLuc. PDT conditions were as indicated. Viral DNA was isolated,
digested with PstI and 1.0 ng was size-fractionated. Hybridization was
performed with PstI-fragmented pJM17. The Figure shown is representa-
tive of several blots.
Photodynamic inactivation of adenovirus
FHE Schagen
et al
876
Figure 4 Electron micrographs of adenovirus particles negatively stained with uranyl acetate. 5 × 10
8
p.f.u. AdCMVLuc treated with 2.7
m
MB and
15 min illumination (a), or treated with 50
m
AlPcS
4
and 30 min illumination (c). Control particles similarly treated but shielded from illumination
are for MB represented by (b) and for AlPcS
4
by (d). One unit of the scale bar is 114 nm in (a), (b) and 141 nm in (c) and (d).
radio-label was found bound to the cells, of which
approximately 75% was internalized (Figure 5). This
binding and especially the internalized fraction was low
(0–2%) in CHO cells, a cell line which lacks the primary
receptor and serves as negative control.
9,10
Photodynamic
treatment caused a dramatic change in the binding pat-
tern. Photoinactivation induced by RB, UP or AlPcS
4
,
resulting in complete inactivation of
3
H-labelled
AdCMVLuc (data not shown), reduced the internalized
fraction from about 30% of the total radioactivity to 0–5%.
At the same time the external-bound, but noninternalized
fraction increased from approximately 15% to 45% for UP
and AlPcS
4
, leaving the total amount of attached radiola-
bel almost unaffected. This may be a result of aspecific
binding due to an effect on the primary receptor–fiber
complex formation, or more likely, a result of modifi-
Photodynamic inactivation of adenovirus
FHE Schagen
et al
877
Figure 5 Binding pattern of adenovirus to HeLa or CHO cells after pho-
toinactivation by MB (1.3
m
), RB (20
m
), UP (50
m
) or AlPcS
4
(50
m
). An amount of 1.8 × 10
6
p.f.u.
3
H-AdCMVLuc was phototreated
with the illumination period (min) as indicated. Black/light gray bars rep-
resent the binding to HeLa cells with the black part representing the tryp-
sin-sensitive fraction and the light gray part representing the internalized,
trypsin-resistant fraction. Dark gray/white bars (right hand bar in each
group) represent the binding to CHO cells, with the dark gray part as
trypsin-sensitive fraction and the white part as trypsin-resistant fraction.
Binding fractions are depicted as percentages of total radioactivity in
3
H-
AdCMVLuc used.
cation of the pentonbase, which is involved in the
initiation of internalization. The increase in the external-
bound fraction was less evident for the RB sample.
After a 1-min illumination in the presence of MB, the
binding pattern of adenovirus to HeLa cells did not
change significantly. Only a slight reduction of the
internalized fraction is observed. This is fully consistent
with the hypothesis that damage induced by MB is
established at the DNA level.
Effectivity of photoinactivation in blood plasma
To study whether this phototreatment could be useful for
biological samples, 2 × 10
8
p.f.u./ml AdCMVLuc in
90% (vol/vol) rat plasma and each of the four photo-
sensitizers were illuminated for 30 min with
106 mWatt/cm
2
white light (Figure 6). In the case of non-
Figure 6 Luciferase activity as a measure of the AdCMVLuc infectivity
after photoinactivation of 10
7
p.f.u. in a background of approximately 90%
rat blood plasma. Photoinactivated samples were used for infection of
HepG2 cells. White bars represent the samples exposed to 106 mW/cm
2
white light for a period of 30 min, black bars are treated similarly, but
were protected from illumination. The concentrations used are: 2.7
m
MB, 40
m
RB, 100
m
UP and 100
m
AlPcS
4
.
exposed samples and in the controls without photosensit-
izer, the infection with 10
7
p.f.u. resulted in a stable
luciferase activity of approximately 10
7
light units. In the
illuminated samples only MB caused a 4 log
10
reduction
in luciferase activity, whereas the others left the infec-
tivity almost unaffected: phototreatment with RB, UP or
AlPcS
4
yielded only a two- to three-fold reduction.
Side-effects of the photodynamic treatment
To study the effect of PDT on plasma proteins two para-
meters were evaluated. Firstly, luciferase was added to
rat plasma and its activity was measured to reflect the
integrity of the protein. Secondly, factor VIII levels in
human blood plasma were analyzed. As depicted in
Figure 7a, application of PDT with 30-min illumination
in 2.7 m MB resulted in a 14% decrease of luciferase
activity compared with illumination without photosensit-
izer. RB, UP and AlPcS
4
were much more pronounced in
their protein inactivation: 93%, 97% and 98%, respect-
ively. This is in accordance with the effect of the treat-
ment determined by measurement of the factor VIII lev-
els. After a 15-min illumination in 2.7 m MB we could
still detect 75% of the amount of factor VIII protein. Fac-
tor VIII was, as expected, much more susceptible to the
other photosensitizers. PDT with a 20-min illumination
in 40 m RB, a 30-min illumination in 50 m AlPcS
4
or
Figure 7 PDT induced side-effects evaluated by (a) luciferase activity or
(b) cell viability. (a) In the background of 70% blood plasma, extracts
from AdCMVLuc-infected cells were exposed to a 30-min illumination
without photosensitizer (control) or in presence of 2.7
m
MB, 40
m
RB, 100
m
UP or 100
m
AlPcS
4
. The luciferase activity is shown as
percentage of the control. (b)
3
H-thymidine incorporation of HepG2 cells,
incubated for 48 h in the indicated MB concentrations. White bars repes-
ent the MB samples, which were illuminated for a 30-min period before
they were added to the cells. Black bars represent samples, which were
protected from illumination.
Photodynamic inactivation of adenovirus
FHE Schagen
et al
878
50 m UP reduced the factor VIII levels to undetectable
levels (1%; data not shown).
As MB is the peferred photosensitizer for adenovirus
inactivation in plasma samples, the effect of MB on the
viability of cells was assessed by measuring the incorpor-
ation of
3
H-thymidine. After a 48-h exposure to 2.7 m
MB of HepG2 cells, there was no effect detected on the
replication (Figure 7b), demonstrating that MB is not
affecting the viability of the cells.
Discussion
So far, photodynamic treatment has been used mainly for
inactivation of enveloped viruses. Although in some
studies inactivation of nonenveloped viruses has been
reported, the inactivation of nonenveloped viruses like
adenovirus, poliovirus, or parvovirus,
5,11–15
has never
been very efficient. In this study, we have analyzed sev-
eral sensitizers for their capacity to inactivate adenovirus
and determined their target site. White light plus low
concentrations of MB sufficed to inactivate completely
10
8
p.f.u./ml AdCMVLuc in PBS. RB and UP achieved
the same effect, but at higher concentrations, whereas
AlPcS
4
affected adenovirus to a lesser extent at all con-
ditions tested.
So far, other groups have never been able to show any
photoinactivating effect of AlPcS
4
on nonenveloped
viruses, such as encephalomyocarditis virus (EMCV) or
adenovirus.
6,7,12
These studies however, used less strin-
gent conditions and were performed on internalized
adenovirus and in the case of EMCV, in the presence of
human plasma.
7,12
This has possibly introduced too many
competitors for AlPcS
4
-induced viral damage, as con-
firmed by our AlPcS
4
data, which showed complete loss
of inactivation in a background of rat plasma. This is sup-
ported by the work of Malik et al,
16
who demonstrated
that protein damage (viz crosslinks) was responsible for
the loss of infectivity of HSV after PDT with several
phthalocyanines.
RB-induced inactivation of influenza and VSV was
reported to be correlated with their inability to fuse with
the host cell membrane.
17,18
After phototreatment, both
viruses showed cross-linking of viral membrane proteins,
which are involved in this fusion. Although the use of
uroporphyrin for virus inactivation has never been
reported, the photodynamic effects of porphyrins, includ-
ing uroporphyrin, on isolated cells, cell membranes, sub-
cellular organelles and on proteins of cytosol and plasma
have been investigated extensively.
19–21
All these studies
indicate that porphyrins induce photodynamic modifi-
cation of proteins.
In approximately 90% blood plasma the photosensitiz-
ers RB, UP and AlPcS
4
only cause a two- to three-fold
reduction in viral infectivity versus a more than 10
5
-fold
reduction in PBS. Apparently, the large amounts of
plasma protein quench the effect on the adenovirus, con-
firming the hypothesis of a protein basis of adenoviral
inactivation by these sensitizers.
The capacity of MB to photoinactivate enveloped
viruses is well documented and the molecular mech-
anism for this photoinactivation has been studied by sev-
eral groups.
5,22,23
It is generally accepted that MB strongly
associates with DNA and upon illumination, induces the
formation of 8-hydroxyguanine (8-OHG).
24–26
Single-
stranded nicks were observed as result of this 8-OHG for-
mation.
24
The data presented here strongly indicate a
preference of MB for DNA, as it was the only sensitizer
tested that completely impaired the functional integrity
of the viral DNA after a 5-min exposure (Figure 2). Upon
extended illumination (15 min), Southern analysis
revealed DNA fragmentation. This suggests that forma-
tion of 8-OHG is sufficient for inactivation, which is in
agreement with the photoinactivation of M13, a non-
enveloped bacteriophage with a single-stranded circular
DNA genome.
22
Abe and Wagner observed a correlation
between the inactivating effect of MB and the number of
8-OHG piperidine-labile sites. Moreover, they described
that higher illumination doses introduced more piperi-
dine-labile sites, as shown by DNA smearing.
Although the inactivating capacity of MB diminished
in the presence of 90% rat plasma, PDT still induced a
4 log
10
-fold reduction of the luciferase activity to almost
background levels. Despite the stringency of the con-
ditions, we were still able to detect 86% of luciferase
activity and 75% of blood clotting factor VIII in blood
plasma, as compared with untreated controls. Factor VIII
is known to be labile
27
and constitutes therefore an appro-
priate model to study the effects that MB and light
exposure has on biological samples, as confirmed by
Lambrecht et al.
5
This group showed that factor VIII is
even less affected at milder conditions. In this perspective
MB appears to be a promising agent for the decontami-
nation of blood or blood products, especially as this pho-
tosensitizer has been reported to induce no genotoxicity
in vivo
10
and did not show any effect on the viability of
HepG2s.
As has been described by Cotten et al
28
defective
virions can be used to facilitate the entry of ligand-coated
DNA particles into cells. In the latter study, inactivation
of adenovirus was based on UV exposure in the presence
of psoralens.
29
A 5-log
10
reduction in virus titer was gen-
erated with the use of 110 g/ml 8-methoxypsoralen (8-
MOP) and 280 g/ml 4-aminomethyl-4,5,8-trimethylp-
soralen (AMT), while DNA delivery capacity was main-
tained. Psoralen-derivatives are however, reported to be
mutagenic, which makes their removal mandatory and
their use laborious.
30
In contrast to psoralen, MB is not
reported to have any mutagenic or genotoxic effect. This
makes the use of MB together with the short white-light
illumination very easy and efficient. The plasmid-
delivering capacity of MB-inactivated adenovirus has not
been investigated and although the virions were shown
to be internalized, aspecific protein modification and con-
comitant loss of endosomolytic properties cannot be
excluded completely.
Materials and methods
Cell culture and virus production
The Ad5 E1-transformed human embryonic retina cell
line 911
31
and the hepatoma cell line HepG2 were grown
in Dulbecco’s modified Eagle medium (DMEM) sup-
plemented with 10% fetal calf serum (FCS), antibiotics
and 3 g/l glucose in a 5% CO
2
atmosphere at 37°C.
Incorporation of
3
H-thymidine in chromosomal DNA
was assessed on HepG2 cells, which were grown on a
six-well plate. At a 40% density, the medium was
replaced by 2 ml DMEM with 1.0% dialyzed FCS and
a MB concentrate. They were incubated at 37°C for 3 h,
Photodynamic inactivation of adenovirus
FHE Schagen
et al
879
whereafter 10 Ci
3
H-thymidine was added. After 48 h,
the medium was removed and the cells were washed
twice with PBS. Finally the cells were detached from the
dish by a trypsin treatment. The
3
H-thymidine content
of the resulting fractions was measured by liquid scintil-
lation counting.
The recombinant adenovirus AdCMVLuc (defective for
the E1 region), which was kindly provided by Dr J Herz
(University of Texas, Southwestern Medical Centre,
Dallas, TX, USA) contains the firefly luciferase gene
under control of the cytomegalovirus immediate–early
promoter.
32
This virus was propagated in the comp-
lementing cell line 911 and isolated as described by
Stratford-Perricaudet and Perricaudet.
33
Briefly, 48 h after
infection, detached cells were harvested and collected in
1 ml phosphate-buffered saline (PBS) containing 1%
horse serum (HS). This was followed by three rounds of
freeze-thawing, two steps of CsCl density-gradient puri-
fication and dialysis to remove the CsCl from the purified
virus suspension. The resulting AdCMVLuc stock was
stored as 50–100 l aliquots in 10% glycerol at 80°C.
3
H-thymidine-labeled AdCMVLuc was produced in
911 cells. A nearly confluent 911 monolayer in a 10-cm
dish was infected by AdCMVLuc in 1 ml 2% HS in PBS
at a multiplicity of infection (MOI) of 5. After 45 min at
room temperature, the inoculum was replaced by 10 ml
of fresh DMEM containing 1% dialyzed FCS. Seven hours
later, this was replaced again by 10 ml fresh DMEM now
containing 0.75% dialyzed FCS and 100 Ci
3
H-
thymidine. One day later the medium was enriched with
2% dialyzed FCS. The medium was collected 48 h after
infection, mixed with an equal volume of 20% PEG-6000
(BDH, Poole, UK) and incubated overnight at 4°C. The
PEG-6000
3
H-AdCMVLuc precipitate was sedimented by
centrifugation at 2800 g for 20 min at 4°C. The resulting
pellet was resolved in 1 × TD buffer (25 mm Tris, 137 mm
NaCl, 5 mm KCl, 0.73 mm Na
2
HPO
4
, 0.9 mm CaCl
2
,
0.5 mm MgCl
2
pH 7.45) and loaded on to a CsCl block-
gradient, consisting of a 2 ml heavy CsCl cushion
(1.45 g/cm
3
CsCl/10 mm Tris (pH 8.1), 1 mm EDTA;
refractive index (RI) 1.374) and a 3 ml light CsCl cushion
(1.20 g/cm
3
CsCl/10 mm Tris (pH 8.1), 1 mm EDTA; RI,
1.355). The centrifugation was performed in a SW-41
rotor at 130 000 g for 1 h at 17°C. The resulting virus band
was dialyzed and stored at 80°C in aliquots containing
10% glycerol.
AdCMVLuc infectivity assay
After PDT, the infectivity of AdCMVLuc was assessed
via the expression level of the luciferase reporter gene
and by determination of the titer. In the luciferase assay,
PDT samples containing photosensitizer and the equival-
ent of 10
7
p.f.u. AdCMVLuc, were adjusted to 200 l with
PBS and added to HepG2 cells in six-well plates. After a
20-min incubation at room temperature, the inoculum
was replaced by 2 ml fresh DMEM containing 2% HS.
Two days after infection, cells were lysed in 200 l lysis
buffer (25 mm Tris-phosphate, pH 7.8, 2 mm DTT, 2 mm
1,2-diaminocyclohexane-N,N,N,N-tetraacetic acid, 10%
glycerol, 1% Triton X-100), the cell debris was sedimented
and supernatant was collected as whole lysate. Luciferase
activity was determined in 10 l of whole lysate by
addition of luciferase assay reagent (Promega, Madison,
WI, USA). After a 10-s incubation, the produced light was
measured in a luminometer (Lumat LB 9501, Berthold,
Bad Wildbad, Germany).
The adenovirus titers were determined in plaque
assays, as described by Fallaux et al.
31
Briefly, PDT
samples were serially diluted in 2 ml of DMEM contain-
ing 2% HS (heat-inactivated at 56°C for 30 min) and
added to confluent 911 cells in six-well plates. After 2 h
of incubation at 37°C, the inoculum was aspirated, and
the cells were overlaid by F15 minimum essential
medium (MEM) containing 12.3 mm MgCl
2
, 0.0025% l-
glutamine, 2% HS, 0.85% agarose (Sigma, St Louis, MO,
USA) and 20 mm HEPES pH 7.4. Plaques were scored
7–10 days after infection.
Photosensitizers
The photosensitizer MB was purchased from GT Baker
(Deventer, The Netherlands) and was stored as a 1.3-mm
stock solution at 4°C. RB was obtained from BDH (Poole,
UK) and AlPcS
4
and UP from Porphyrin Products
(Logan, UT, USA). The latter were stored as 1.0 mm stock
solutions at 4°C. All stocks were prepared in 50 mm
sodium-phosphate buffer (pH 7.4), filter-sterilized
(0.22 m filter) and stored in the dark.
Photoinactivation
Photoinactivations were performed, unless indicated
otherwise, on 10
7
p.f.u. AdCMVLuc in 100 l PBS. All
illuminations were done at a constant temperature of
20°C using a slide projector with a 150 W Xenophot HLX
64640 lamp as light source. The dose rate of white-light
irradiation was 106 mW/cm
2
, as measured with a Gentec
(Saint Foy, CA, USA) TMP-310 photometer. Controls
were prepared and incubated identically, but were shi-
elded from light by wrapping them in aluminum foil.
Analysis of viral DNA
The DNA integrity of the photoinactivated virions was
tested by transfection of the viral DNA into HepG2 and
911 cells and by Southern blot analysis. After photoinac-
tivation of 3 × 10
8
p.f.u. AdCMVLuc, aliquots of the
exposed virus were mixed with 5 g carrier DNA and
400 g proteinase K, the volume adapted to 100 l with
PBS and incubated overnight at 48°C to degrade the viral
capsid. The DNA was further purified by phenol/
chloroform (1:1) extraction and ethanol precipitation and
2.1 ng was used to transfect HepG2 or 911 cells in six-
well plates. DNA transfections were performed with 0.5
mg/ml DEAE-dextran in 1 × TBS (25 mm Tris Cl, 137 mm
NaCl, 5 mm KCl, 0.7 mm CaCl
2
, 0.5 mm MgCl
2
, 0.6 mm
Na
2
HPO
4
; pH 7.4).
Other aliquots of the exposed virus were used for
Southern blot analysis. DNA was isolated essentially as
described above, with 1 g tRNA as carrier. The DNA
was subsequently digested with PstI, and 1 ng of each
sample was size-fractionated by electrophoresis on a 1.2%
agarose gel. The fractionated DNA was transferred on to
Hybond N+ (Amersham, UK) with 0.4 m NaOH/0.6 m
NaCl as the transfer solution. The resulting blot was
hybridized with the radio-labeled PstI-digested plasmid
pJM17 (containing the entire adenovirus genome) as
probe. Labelling was performed according to the ran-
dom-primer method using -
32
P-dATP, random hexanu-
cleotides and Klenow DNA polymerase. Finally, the
Southern blot was exposed to Fuji-RX film at 80°C.
Photodynamic inactivation of adenovirus
FHE Schagen
et al
880
Electron microscopy
Aliquots of 5.0 × 10
8
p.f.u. AdCMVLuc were exposed to
light in 2.7 m MB for 0 and 15 min, in 20 m RB for 0
and 20 min and in either 50 m AlPcS
4
or 50 m UP for
0 and 30 min. All samples were fixed for 5 min in equal
volumes of 1% glutaraldehyde at room temperature.
Hereafter, virus samples were loaded on to carbon-
coated grids, negatively stained with 3–7% uranyl acetate
and analyzed with a Philips electron microscope
(Eindhoven, The Netherlands).
Binding characteristics of adenovirus
The general procedure of the virion-attachment assay
was outlined by Wohlfart et al.
34
60–70% confluent mono-
layers of either HeLa or CHO cells were preincubated
with 1% BSA for 10 min at 20°C.
35
Samples of 1.6 × 10
7
p.f.u./ml
3
H-thymidine-labeled virions were photoinac-
tivated, added to the monolayer at a MOI of 1–2 and
incubated for 120 min at 37°C. Subsequently, the inocu-
lum was aspirated and the cells were washed twice with
ice-cold 2% HS in PBS to remove the unbound virus.
Finally, the monolayers were incubated in the presence
of trypsin, until the cells detached from the dish. This cell
suspension was centrifuged at 150 g for 5 min and div-
ided into the cell pellet, representing the trypsin-insensi-
tive adenovirus fraction (the internalized virions) and the
supernatant, representing the trypsin-sensitive fraction
(the external bound, but non-internalized virions).
36
Radioactivity of all fractions was measured by liquid
scintillation counting and the percentage of attached or
internalized virus was calculated.
Acknowledgements
We are grateful to Ferry Spies and Hans van der Meulen
for the electron microscopy and Diana van den Wollen-
berg for excellent technical assistance. We also want to
thank Nico van Tilburg for the analysis of the plasma
samples. This investigation was financially supported by
the Dutch Organisation for Scientific Research (NWO),
project No. 901–01–096, and the Dutch ‘Praeventiefonds’,
grant No. 28–2178–1.
References
1 Takafuji ET, Gaydos JC, Allen RG, Top FH Jr. Simultaneous
administration of live, enteric-coated adenovirus types 4, 7 and
21 vaccines: safety and immunogenicity. J Infect Dis 1979; 140:
48–53.
2 Boulikas T. Status of gene therapy in 1997: molecular mech-
anisms, disease targets, and clinical applications. Gene Ther Mol
Biol 1998; 1: 1–172.
3 Imler JL. Adenovirus vectors as recombinant viral vaccines. Vac-
cine 1995; 13: 1143–1151.
4 Ben-Hur E et al. Photodynamic inactivation of retroviruses by
phthalocyanines: the effects of sulphonation, metal ligand and
fluoride. J Photochem Photobiol B 1992; 13: 145–152.
5 Lambrecht B, Mohr H, Knuver Hopf J, Schmitt H. Photoinactiv-
ation of viruses in human fresh plasma by phenothiazine dyes
in combination with visible light. Vox Sang 1991; 60: 207–213.
6 Horowitz B et al. Inactivation of viruses in blood with aluminum
phthalocynine derivatives. Transfusion 1991; 31: 102–108.
7 Horowitz B et al. Inactivation of viruses in red cell and platelet
concentrates with aluminum phtalocyanine (AlPc) sulfonates.
Blood Cells 1992; 18: 141–150.
8 Ben-Hur E et al. The photodecontamination of cellular blood
components: mechanisms anduse of photosensitization in trans-
fusion medicine. Transfus Med Rev 1996; X: 15–22.
9 Stevenson SC et al. Human adenovirus serotypes 3 and 5 bind
to diferent cellular receptors via the fiber head domain. J Virol
1997; 69: 2850–2857.
10 Wagner SJ et al. Mammalian genotoxicity assessment of methyl-
ene blue in plasma: implications for virus inactivation. Trans-
fusion 1995; 35: 407–413.
11 Dewilde A et al. Virucidal activity of pure singlet oxygen gener-
ated by thermolysis of a water-soluble naphtalene endoperox-
ide. J Photochem Photobiol B 1996; 36: 23–29.
12 Smetana Z et al. Photodynamic inactivation of herpes viruses
with phthalocyanine derivatives. J Photochem Photobiol B 1994;
22: 37–43.
13 Matthews JL et al. Photodynamic therapy of viral contaminants
with potential for blood banking applications. Transfusion 1988;
28: 81–83.
14 Sieber F et al. Antiviral activity of merocyanine 540. Photochem
Photobiol 1987; 46: 707–711.
15 O’Brien JM et al. Evaluation of merocyanine 540-sensitized pho-
toirradiation as a means to inactivate enveloped viruses in blood
products. J Lab Clin Med 1990; 116: 439–447.
16 Malik Z et al. Alteration in herpes simplex virus proteins follow-
ing photodynamic teatment with phtalocyanines. Photochem Pho-
tobiol 1996; 63 (Suppl.): 59S–60S (Abstr. TAM-A6).
17 Lenard J, Vanderoef R. Photoinactivation of influenza virus
fusion and infectivity by rose bengal. Photochem Photobiol 1993;
58: 527–531.
18 Lenard J, Rabson A, Vanderoef R. Photodynamic inactivation of
infectivity of human immunodeficiency virus and other envel-
oped viruses using hypericin and rose bengal: inhibition of
fusion and syncytia formation. Proc Natl Acad Sci USA 1993; 90:
156–162.
19 Vincent S, Muller-Eberhard U. Effects of porphyrins on proteins
of cytosol and plasma. In vitro photo-oxidation and crosslinking
of proteins by naturally occurring and synthetic porphyrins. J
Lab Clin Med 1987; 110: 475–482.
20 Vincent S, Holeman B, Cully BC, Muller-Eberhard U. Porphyrin-
induced photodynamic cross-linking of hepatic heme-binding
proteins. Life Sci 1986; 38: 365–372.
21 Sandberg S, Romslo I. Porphyrin-sensitized photodynamic dam-
age of isolated rat liver mitochondria. Biochim Biophys Acta 1980;
593: 187–195.
22 Abe H, Wagner SJ. Analysis of viral DNA, protein and envelope
damage after methylene blue, phthalocyanine derivative or mer-
ocyanine 540 photosensitization. Photochem Photobiol 1995; 61:
402–409.
23 Schnipper LE, Lewin AA, Swartz M, Crumpacker CS. Mech-
anisms of photodynamic inactivation of herpes simplex viruses:
comparison between methylene blue, light plus electricity, and
hematoporphyrin plus light. J Clin Invest 1980; 65: 432–438.
24 Schneider JE Jr et al. Methylene blue and rose bengal photoinac-
tivation of RNA bacteriophages: comparative studies of 8-oxog-
uanine formation in isolated RNA. Arch Biochem Biophys 1993;
301: 91–97.
25 Floyd RA, West MS, Eneff KL, Schneider JE. Methylene blue
plus light mediates 8-hydroxyguanine formation in DNA. Arch
Biochem Biophys 1989; 273: 106–111.
26 Wang J, Hogan M, Austin RH. DNA motions in the nucleosome
core particle. Proc Natl Acad Sci USA 1982; 79: 5896–5900.
27 Noe DA. A mathematical model of coagulation factor VIII kin-
etics. Haemostasis 1996; 26: 289–303.
28 Cotten M et al. High-efficiency receptor-mediated delivery of
small and large (48 kilobase) gene constructs using the endo-
some-disruption activity of defective or chemically inactivated
adenovirus particles. Proc Natl Acad Sci USA 1992; 89: 6094–6098.
29 Cotten M et al. Psoralen treatment of adenovirus particles elim-
inates virus replication and transcription while maintaining the
endosomolytic activity of the virus capsid. Virology 1994; 205:
254–261.
30 Wagner SJ et al. Determination of residual 4-aminomethyl-
4,5,8-trimethylpsoralen and mutagenicity testing following pso-
ralen plus UVA treatment of platelet suspensions. Photochem
Photobiol 1993; 57: 819–824.
Photodynamic inactivation of adenovirus
FHE Schagen
et al
881
31 Fallaux FJ et al. Characterization of 911: a new helper cell line
for the titration and propagation of early region 1-deleted aden-
oviral vectors. Hum Gene Ther 1996; 7: 215–222.
32 Herz J, Gerard RD. Adenovirus-mediated transfer of low den-
sity lipoprotein receptor gene acutely accelerates cholesterol
clearance in normal mice. Proc Natl Acad Sci USA 1993; 90:
2812–2816.
33 Stratford-Perricaudet LD, Perricaudet M. Gene transfer into ani-
mals the promise of adenovirus. In: Cohen-Haguenauer O,
Boiron M (eds). Human Gene Transfer. Inserm/John Libbey Euro-
text: Paris, 1991, pp 51–61.
34 Wohlfart CEG, Svensson UK, Everitt E. Interaction between
HeLa cells and adenovirus type 2 virions neutralized by differ-
ent antisera. J Virol 1985; 56: 896–903.
35 Blixt Y, Varga MJ, Everitt E. Enhancement of intracellular
uncoating of adenovirus in HeLa cells in the presence of benzyl
alcohol as a membrane fluidizer. Arch Virol 1993; 129: 265–277.
36 Greber UF, Willetts M, Webster P, Helenius A. Stepwise dis-
mantling of adenovirus 2 during entry into cells. Cell 1993; 75:
477–486.
... However, normalized ΦX174 inactivation rates for metalloporphyrins PdT4 and PdC14 were about an order of magnitude faster than previous reports (~10 −5 L cm 2 µmol −1 mJ −1 ), indicating significant improvement over TMPyP-mediated inactivation. Schagen et al. [67] reported nearly as rapid inactivation of recombinant adenovirus via methylene blue/halogen light inactivation, with a low concentration (1.3 µM) of methylene blue resulting in a greater normalized inactivation rate (~10 −4 L cm 2 µmol −1 mJ −1 ). However, the quantification of recombinant adenovirus via green fluorescent protein expression [67], rather than infectious cultural assay, may undermine direct comparison. ...
... Schagen et al. [67] reported nearly as rapid inactivation of recombinant adenovirus via methylene blue/halogen light inactivation, with a low concentration (1.3 µM) of methylene blue resulting in a greater normalized inactivation rate (~10 −4 L cm 2 µmol −1 mJ −1 ). However, the quantification of recombinant adenovirus via green fluorescent protein expression [67], rather than infectious cultural assay, may undermine direct comparison. ...
Article
Full-text available
Effective broad-spectrum antiviral treatments are in dire need as disinfectants and therapeutic alternatives. One such method of disinfection is photodynamic inactivation, which involves the production of reactive oxygen species from dissolved oxygen in response to light-stimulated photosensitizers. This study evaluated the efficacy of functionalized porphyrin compounds for photodynamic inactivation of bacteriophages as human virus surrogates. A blue-light light emitting diode (LED) lamp was used to activate porphyrin compounds in aqueous solution (phosphate buffer). The DNA bacteriophages ΦX174 and P22 were more resistant to porphyrin TMPyP photodynamic inactivation than RNA bacteriophage fr, with increasing rates of inactivation in the order: ΦX174 << P22 << fr. Bacteriophage ΦX174 was therefore considered a resistant virus suitable for the evaluation of three additional porphyrins. These porphyrins were synthesized from TMPyP by inclusion of a central palladium ion (PdT4) and/or the addition of a hydrophobic C14 chain (PdC14 or C14). While the inactivation rate of bacteriophage ΦX174 via TMPyP was similar to previous reports of resistant viruses, ΦX174 inactivation increased by a factor of approximately 2.5 using the metalloporphyrins PdT4 and PdC14. The order of porphyrin effectiveness was TMPyP < C14 < PdT4 < PdC14, indicating that both Pd2+ ligation and C14 functionalization aided virus inactivation.
... For nonenveloped viruses, the photoinactivation effects depend mainly on damages in the protein capsid and/or loosening of protein-DNA interaction (Egyeki et al., 2003). So far, photodynamic inactivation has been proven to be a powerful method for inactivating enveloped viruses, such as murine retroviral vectors (Ben-Hur et al., 1992), human immunodeficiency viruses (HIV-1 and -2) (Schagen et al., 1999;Vzorov et al., 2002), herpes simplex viruses (Silva et al., 2005;Tome et al., 2007), hepatitis-B (Wagner et al., 2001) and vesicular stomatitis virus (Horowitz et al., 1991) and also for the inactivation of nonenveloped viruses, like the adenovirus (Schagen et al., 1999), hepatitis A virus (Casteel et al., 2004), human papilloma virus (Wainwight, 2004) and T7 (Egyeki et al., 2003), lambda (Kasturi and Plaz, 1992) and MS2 (Casteel et al., 2004). Antifungal photoinactivation: Fungi are much more complex targets than bacteria. ...
... For nonenveloped viruses, the photoinactivation effects depend mainly on damages in the protein capsid and/or loosening of protein-DNA interaction (Egyeki et al., 2003). So far, photodynamic inactivation has been proven to be a powerful method for inactivating enveloped viruses, such as murine retroviral vectors (Ben-Hur et al., 1992), human immunodeficiency viruses (HIV-1 and -2) (Schagen et al., 1999;Vzorov et al., 2002), herpes simplex viruses (Silva et al., 2005;Tome et al., 2007), hepatitis-B (Wagner et al., 2001) and vesicular stomatitis virus (Horowitz et al., 1991) and also for the inactivation of nonenveloped viruses, like the adenovirus (Schagen et al., 1999), hepatitis A virus (Casteel et al., 2004), human papilloma virus (Wainwight, 2004) and T7 (Egyeki et al., 2003), lambda (Kasturi and Plaz, 1992) and MS2 (Casteel et al., 2004). Antifungal photoinactivation: Fungi are much more complex targets than bacteria. ...
Article
Full-text available
Global aquaculture production in 2012 touched new high of 90.4 million tonnes including 66.6 million tonnes of food fish and 23.8 million tonnes of aquatic algae providing 19.2 kg per capita food fish suppy. Aquaculture is reported to suffer heavy production and financial losses due to fish infections caused by microbial pathogens. Therefore in order to make aquaculture industry more sustainable, effective strategies to control fish infections are urgently needed. Antimicrobial Photodynamic Therapy (aPDT) is an emerging, low-cost anti-microbial approach to the treatment of locally occurring infections and also for the treatment of aquaculture water and waste waters. Already proven effective in various medical and clinical applications, it utilizes three vital components: a photosensitizing agent (PS), a light source of an appropriate wave length and oxygen. aPDT has got a potential of being a preferred choice over antibiotics in aquaculture systems because of its non-target specificity, few side effects, lack of the pathogenicity reversal and re-growth of the micro-organism after treatment and the lack of development of resistance mechanisms. The technique has been proved effective in vitro against bacteria (including drug-resistant strains), yeasts, fungi, viruses, parasites and even the stubborn biofilms. Although preliminary results indicate that this technology has a high potential to disinfect waters in aquaculture system and also in hatcheries and seed production units, but it clearly needs more deep knowledge and multi-dimenstional approach.
... This is a critically important advantage of this treatment approach. Moreover, there is potential to address many other pathogens contracted via the nasal and oral mucosa, including influenza, the common cold, respiratory syncytial virus, adenovirus infections and possibly pox viruses, using these combined protocols [108][109][110]. ...
Article
Full-text available
Coronavirus disease 2019 (COVID-19) caused by SARS-CoV-2 virus was first recognized in late 2019 and remains a significant threat. We therefore assessed the use of local methylene blue photodynamic viral inactivation (MB-PDI) in the oral and nasal cavities, in combination with the systemic anti-viral, anti-inflammatory and antioxidant actions of orally ingested methylene blue (MB) and photobiomodulation (PBM) for COVID-19 disease. The proposed protocol leverages the separate and combined effects of MB and 660nm red light emitted diode (LED) to comprehensively address the pathophysiological sequelae of COVID-19. A total of eight pilot subjects with COVID-19 disease were treated in the Bahamas over the period June 2021–August 2021, using a remote care program that was developed for this purpose. Although not a pre-requisite for inclusion, none of the subjects had received any COVID-19 vaccination prior to commencing the study. Clinical outcome assessment tools included serial cycle threshold measurements as a surrogate estimate of viral load; serial online questionnaires to document symptom response and adverse effects; and a one-year follow-up survey to assess long-term outcomes. All subjects received MB-PDI to target the main sites of viral entry in the nose and mouth. This was the central component of the treatment protocol with the addition of orally ingested MB and/or PBM based on clinical requirements. The mucosal surfaces were irradiated with 660 nm LED in a continuous emission mode at energy density of 49 J/cm2 for PDI and 4.9 J/cm2 for PBM. Although our pilot subjects had significant co-morbidities, extremely high viral loads and moderately severe symptoms during the Delta phase of the pandemic, the response to treatment was highly encouraging. Rapid reductions in viral loads were observed and negative PCR tests were documented within a median of 4 days. These laboratory findings occurred in parallel with significant clinical improvement, mostly within 12–24 h of commencing the treatment protocol. There were no significant adverse effects and none of the subjects who completed the protocol required in-patient hospitalization. The outcomes were similarly encouraging at one-year follow-up with virtual absence of “long COVID” symptoms or of COVID-19 re-infection. Our results indicate that the protocols may be a safe and promising approach to challenging COVID-19 disease. Moreover, due its broad spectrum of activity, this approach has the potential to address the prevailing and future COVID-19 variants and other infections transmitted via the upper respiratory tract. Extensive studies with a large cohort are warranted to validate our results.
... Type II mechanism involving generation of 1 O 2 has been shown to be the predominant mechanism of bacteria/virus inactivation (Costa et al., 2012). Comprehensive reviews highlighting the various bacteria (Hamblin and Hasan, 2004;Liu et al., 2015) and viruses (Costa et al., 2012) (Lenard et al., 1993), dengue virus (Huang et al., 2004), vaccinia virus (Turner and Kaplan, 1968), adenovirus (Schagen et al., 1999), enterovirus (Wong et al., 2010) etc. have all been reported to be inactivated by 1 O 2 . In addition, it was showed that herpesvirus and influenza virus (enveloped type) and adenovirus and poliovirus (unenveloped type) were inactivated by 1 O 2 using a spray of Rose Bengal (RB) solution as mist and air scavenging system (Masaoka et al., 2013). ...
Article
Full-text available
Airborne infectious diseases such as the new Coronavirus 2019 (COVID-19) pose serious threat to human health. Indoor air pollution is a problem of global environmental concern as well. Singlet oxygen (¹O2) is a reactive oxygen species that plays important role in bacteria/virus inactivation and pollutant degradation. In this study, we found that commercially available filters typically deployed in air purifier and air conditioning units, impregnated with Rose Bengal (RB) as a ¹O2 sensitizer, can be used for heterogeneous gas-phase generation of ¹O2. It was confirmed that irradiation of the RB filter under oxygen gas stream produced ¹O2, which was measured using furfuryl alcohol trapping method followed by HPLC analysis. It was also observed that the amount of ¹O2 generated increases as the light intensity increased. Similarly, the sensitizer loading also positively influenced the ¹O2 generation. The heterogeneous gas-phase generation of ¹O2 can find potential applications in air purifier and air conditioning units for the purpose of bacteria/virus inactivation and/or pollutant degradation thereby improving indoor air quality.
... VHSV can infect many fish species, including olive flounder, leading to economic losses (Meyers and Winton, 1995;Isshiki et al., 2001;Skall et al., 2005;Kim et al., 2009). In general, a virus can be readily inactivated by conventional disinfectants (Schagen et al., 1999;Oye and Rimstad, 2001;Szeimies, 2003). However, chemotherapeutic agents are currently unavailable to treat viral diseases in any fish. ...
... 304,307 This IgG binding is critical, Apart from blood product disinfection phthalocyanines have been proposed and tested for the inactivation of adenoviral vectors in the context of biosafety in laboratories using recombinant adenovirus vectors gene transfer and gene therapy. 308 A few publications have addressed the targets of antiviral PDI with phthalocyanines. 87,309 For aluminum phthalocyanine tetrasulfonate it was elucidated by quenching experiments that 1 O 2 plays the dominant role for VSV inactivation. ...
Article
Full-text available
Photodynamic therapy (PDT) is a well-established treatment option in the treatment of certain cancerous and pre-cancerous lesions. Though best-known for its application in tumor therapy, historically the photodynamic effect was first demonstrated against bacteria at the beginning of the 20th century. Today, in light of spreading antibiotic resistance and the rise of new infections, this photodynamic inactivation (PDI) of microbes, such as bacteria, fungi, and viruses, is gaining considerable attention. This review focuses on the PDI of viruses as an alternative treatment in antiviral therapy, but also as a means of viral decontamination, covering mainly the literature of the last decade. The PDI of viruses shares the general action mechanism of photodynamic applications: the irradiation of a dye with light and the subsequent generation of reactive oxygen species (ROS) which are the effective phototoxic agents damaging virus targets by reacting with viral nucleic acids, lipids and proteins. Interestingly, a light-independent antiviral activity has also been found for some of these dyes. This review covers the compound classes employed in the PDI of viruses and their various areas of use. In the medical area, currently two fields stand out in which the PDI of viruses has found broader application: the purification of blood products and the treatment of human papilloma virus manifestations. However, the PDI of viruses has also found interest in such diverse areas as water and surface decontamination, and biosafety.
Article
The control of pests and vector-borne diseases (VDBs) are considered public health issues Worldwide. Among the control techniques and pesticides used so far, photodynamic inactivation (PDI) has been shown as an eco-friendly, low cost, and efficient approach to eliminate pests and VDBs. PDI is characterized using a photosensitizing molecule, light and molecular oxygen (O 2 ) resulting in production of reactive oxidative species which can promote the oxidation of biomolecules on pests and vectors. Herein, we review the past 51 years (1970–2021) regarding the use of photo pesticides, reporting the most important parameters for the protocol applied, the results obtained, and limitations. Moreover, we described the mechanism of action of the PDI, main classes of photopesticides used so far as well as the cell death mechanism resulting from the photodynamic action.
Article
Background Herpes Simplex Virus Type 1 (HSV-1) is one of the most widespread infections that can effect the orofacial region. Recurrent infection is considered a life-long oral health problem, leading to pain, discomfort, and social restriction due to esthetic features when active. Effective therapies are needed. This study aimed to compare photodynamic therapy (PDT), Topical Acyclovir (AC), and the association of both in the healing process and self-reported symptomologies of HSV-1 recurrences. Methods Patients were randomly assigned into 3 groups (n = 25): PDT (low-power laser, 660 nm, 40 mW, 120 J/cm², 4.8 J, 120 s per point) and methylene blue (0.005 %) as photosensitizer; AC (5%); PDT + AC.Data concerning lesion size, healing time, and self-reported healing parameters, such as pain, tingling, and edema were taken every day up to complete healing for all studied groups. Results There was no significant difference in healing time and pain between groups. AC group showed a significant minor reduction of the lesion compared to the AC-PDT group on day 1. Regarding edema and tingling, the comparison of treatments showed a statistical difference only on day 1, where PDT showed better results. Conclusion With all the limitations of this study, it can be concluded that only on day 1 PDT showed positive effects in the treatment of herpes lesions in comparison to AC.
Article
The COVID-19 pandemic has updated research on inactivation of coronaviruses with physical, chemical, and physical-chemical treatments. The review focuses on the inactivation of coronaviruses including the recent SARS-CoV-2 and viruses of other groups with analogous structure using optical radiation. The antiviral effects of different optical ranges from vacuum ultraviolet to infrared radiation are described in terms of the mechanisms of virus photoinactivation, sensitive molecular targets and efficacy. Direct and photosensitized damaging effects of light on the viral molecular structures are considered. Information on the applied pathogen photoinactivation technologies and the advantages of light sources for future applications is provided.
Article
Full-text available
Antimicrobial violet-blue light is an emerging technology designed for enhanced clinical decontamination and treatment applications, due to its safety, efficacy and ease of use. This systematised review was designed to compile the current knowledge on the antimicrobial efficacy of 380-480 nm light on a range of healthcare and food related pathogens including vegetative bacteria, bacterial endospores, fungi and viruses. Data was compiled from 79 studies, with the majority focussing on wavelengths in the region of 405 nm. Analysis indicated that Gram positive and negative vegetative bacteria are the most susceptible organisms, whilst bacterial endospores, viruses and bacteriophage are the least. Evaluation of the dose required for a 1 log10 reduction of key bacteria compared to population, irradiance and wavelength indicated that microbial titre and light intensity had little effect on the dose of 405 nm light required, however linear analysis indicated organisms exposed to longer wavelengths of violet-blue light, may require greater doses for inactivation. Additional research is required to ensure this technology can be used effectively, including: investigating inactivation of multidrug-resistant organisms, fungi, viruses and protozoa; further knowledge about the photodynamic inactivation mechanism of action; the potential for microbial resistance; and the establishment of a standardised exposure methodology. This article is protected by copyright. All rights reserved.
Article
Full-text available
One limit to successful receptor-mediated gene delivery is the exit of the endocytosed material from the endosome. We demonstrate here the delivery of marker genes to tissue culture cells using a modification of the receptor-mediated gene delivery technique that exploits the endosomolytic activity of defective adenovirus particles. In particular, greater than 90% of the transfected-cell population is found to express a beta-galactosidase gene, and, most importantly, this high level of expression can be obtained with psoralen-inactivated virus particles. Furthermore, because the delivered gene is not carried within the genome of the adenovirus particle, the size constraints are relieved, and we can, therefore, show the delivery of a 48-kilobase cosmid DNA molecule.
Article
The safety of and the immune response to simultaneous administration of live, enteric-coated adenovirus type 4 (ADV-4), type 7 (ADV-7), and type 21 (ADV-21) vaccines were studied. Volunteers (476 men), randomly assigned to four study groups, received three vaccines (ADV-4, ADV-7, and ADV-21), two vaccines (ADV4 and ADV-7), one vaccine (ADV-21), or no vaccine (placebo). Subjects were observed for three weeks, and no side effects due to vaccination occurred. The percentages of susceptible subjects (those entering the study with a neutralizing antibody titer of >1:2 to each vaccine virus received) who seroconverted to ADV-4 were similar in both groups that received ADV-4 vaccine (78% of 77 subjects and 74% of 76). However, in the group that received three vaccines, only 62% of 77 subjects seroconverted to ADV-7, compared with 79% of 76 in the group that received two vaccines (P < 0.05). Only 58% of 77 subjects in the three-vaccine group seroconverted to ADV-21, compared with 69% of 59 in the group that received one vaccine (P > 0.1).
Article
The photodynamic inactivation of retroviruses was investigated using aluminium and zinc phthalocyanine (Pc) derivatives. The N2 retrovirus packaged in either of the two murine cell lines, Psi2 and PA317, was used as a model for enveloped viruses. AlPc derivatives were found to be more effective photodynamically for inactivation of the viruses than the corresponding ZnPc derivatives. Sulphonation of the Pc macrocycle reduced its photodynamic activity progressively for both AlPc and ZnPc. Fluoride at 5 mM during light exposure completely protected viruses against inactivation by AlPc. In the presence of F-, inactivation by the sulphonated derivatives AlPcS1 and AlPcS4 was reduced 2.5- and twofold respectively. In a biological membrane (erythrocyte ghosts), F- had no significant effect on AlPcS4-sensitized lipid peroxidation. Under similar conditions, cross-linking of spectrin monomers in ghosts is drastically inhibited (E. Ben-Hur and A. Orenstein, Int. J. Radiat. Biol., 60 (1991) 293-301). Since Pc derivatives do not inactivate non-enveloped viruses, it is hypothesized that inactivation occurs by photodynamic damage to envelope protein(s). Substitution of sulphonic acid residues reduces the binding of Pc derivatives to the envelope protein(s), thereby diminishing their photodynamic efficacy and the ability of F- to modify it.
Article
Aluminum phthalocyanine tetrasulfonates (AIPcS) are photoactive compounds with absorption maxima at 665-675 nm. The inactivation of viruses (vesicular stomatitis virus, VSV; human immunodeficiency virus, HIV) added to either whole blood or red blood cell concentrates (RBCC) and platelet concentrates (PC) on treatment with tetrasulfonated AIPc (AIPcS4) was evaluated. Treatment of RBCC with 10 microM AIPcS4 and 44 J/cm2 visible light resulted in the inactivation of greater than or equal to 10(5.5) infectious doses (TCID50) of cell-free VSV, greater than or equal to 10(5.6) TCID50 of cell-associated VSV, and greater than or equal to 10(4.7) TCID50 of cell-free sindbis virus. Both greater than or equal to 10(4.2) TCID50 of cell-free and greater than or equal to 10(3.6) TCID50 of cell-associated forms of HIV were also shown to be inactivated. Encephalomyocarditis virus, used as a model for nonenveloped viruses, was not inactivated. Equivalent virus kill with Photofrin II required a substantially higher concentration of dye and longer exposure to visible light. Following AIPcS4 treatment, red cell integrity was well maintained as judged by the low level (less than 2%) of hemoglobin release immediately following treatment and on subsequent storage, by measurements of erythrocyte osmotic fragility, and by the normal recovery and circulatory survival on infusion of treated, autologous red blood cells in baboons. Treatment of PC with 10 microM AIPcS4 and 44 J/cm2 visible light also resulted in effective virus kill (greater than or equal to 10(5.5) TCID50) of VSV; however, both the rate and extent of platelet aggregation in response to collagen addition declined by at least 50%. Based on these results, further characterization of AIPcS4-treated RBCC is justified.
Article
We developed a photodynamic method to inactivate viruses in human fresh plasma. Single plasma bags were illuminated with visible light in the presence of low doses of phenothiazine dyes like methylene blue or toluidine blue. By this treatment the infectivity of different enveloped viruses including the causative agent of AIDS, HIV-1, was completely removable from the plasma. Non enveloped viruses, however, proved to be more stable. The activities of clotting factors and other plasma proteins were only slightly decreased. There was no indication that the procedure led to important structural modifications of plasma proteins. The dyes are photodynamically active at concentrations much lower than those at which they are therapeutically used as antidots in the treatment of methemoglobinemia.
Article
The inactivation of viruses added to whole blood and a red cell concentrate with aluminum phthalocyanine and its sulfonated derivatives was studied. A cell-free form of vesicular stomatitis virus (VSV), used as a model, was completely inactivated (greater than 10(4) infectious units; TCID50) on treatment of whole blood with 10 microM (10 mumol/L) aluminum phthalocyanine chloride (AIPs) and visible light dosage of 88 to 176 J per cm2. At 44 J per cm2, complete VSV inactivation was achieved on raising the concentration of AIPc to 25 microM (25 mumol/L). Results at least as good were achieved on similar treatment of a red cell concentrate. Also inactivated were a cell-associated form of VSV and both cell-free and cell-associated forms of human immunodeficiency virus; encephalomyocarditis virus, used as a model for non-lipid-enveloped viruses, was not inactivated by this procedure. This inactivation of cell-free VSV suggests that a similar degree of inactivation could be achieved with a lower concentration of the sulfonated forms of aluminum phthalocyanine. Throughout the above studies, red cell integrity was well maintained, as judged by the absence of hemoglobin release (less than or equal to 2%) during the course of treatment or on subsequent storage. Red cell osmotic fragility was decreased on treatment of whole blood with AIPc. This study indicates that AIPc may be a promising method for the inactivation of viruses in cellular blood products.
Article
Dry or wet heat, solvents, and detergents combined with ultraviolet irradiation provide effective means of sterilizing soluble blood products such as albumin or factor VIII. For obvious reasons, these procedures are not applicable to cellular blood components. We have recently shown that simultaneous exposure to the photosensitizer, merocyanine 540 (MC 540) and white light rapidly inactivates the Friend erythroleukemia virus complex and Friend virus-transformed cells, but causes relatively little damage to pluripotent hematopoietic stem cells. In this communication, we show that several lipid-enveloped human pathogenic viruses are also highly susceptible to MC 540-sensitized photoirradiation, and we report on an initial evaluation of the ability of MC 540-sensitized photoirradiation to sterilize blood products.
Article
Three types of hepatic proteins, a heme-binding Z protein, a mixture of the glutathione S-transferases and a cytochrome P450 isozyme, were shown to be susceptible to photodynamic cross-linking and loss in antigenicity by naturally occurring porphyrins. At 50 microM, uroporphyrin caused the most and protoporphyrin the least photodecomposition. Hemopexin, a specific serum heme carrier, was photodecomposed but no cross-linking was detected. Heme and scavengers of singlet oxygen partially prevented protein photodecomposition.
Article
We examined the photodynamic effects of porphyrins, known photosensitizers, on proteins of cytosol and plasma that bind them and are implicated in their transport. Their susceptibility to photodecomposition by porphyrins was found to be higher than that of proteins with low or no affinity for tetrapyrroles. Inhibition of porphyrin binding by the addition of equimolar amounts of heme had no effect, indicating that protein photodecomposition may be induced, in part, by free or nonspecifically bound porphyrins. HBP, a heme-binding Z protein of liver cytosol, exhibited the highest susceptibility of all proteins tested, including glutathione S-transferases, albumin, hemopexin, and apotransferrin. HBP was extensively photo-oxidized, as evidenced by a decrease in its antigenicity and electrophoretic mobility, and it was cross-linked by naturally occurring porphyrins as well as by the synthetic tin-protoporphyrin and hematoporphyrin derivative. The water-soluble singlet oxygen scavengers L-histidine (50 mmol/L) and sodium azide (100 mmol/L) completely prevented the photodynamic effects of uroporphyrin (100 mumol/L) on HBP. Hydroxyl radical scavengers such as manitol and benzoate were partially effective, whereas water-insoluble singlet oxygen scavengers such as beta-carotene were totally ineffective. Preferential inhibition of cross-linking over other photodynamic effects of uroporphyrin was consistent with previous reports that cross-linking occurs subsequently to amino acid oxidation.