ArticlePDF AvailableLiterature Review

Laser and Non-laser Light Sources for Photodynamic Therapy

Authors:

Abstract and Figures

Photodynamic therapy (PDT) is an anticancer combination therapy, which requires a photosensitiser, which tends to accumulate preferentially in the tumour, and light. Historically large, complex lasers have been used to carry out PDT treatment. Nowadays there is a wide range of coherent and non-coherent sources that can be used. This paper considers the important characteristics of light sources for PDT, including dye lasers pumped by argon or metal vapour lasers and frequency-doubled Nd:YAG lasers. Non-laser sources including tungsten filament, xenon arc, metal halide and fluorescent lamps are also discussed. New exciting developments such as LEDs and femtosecond lasers are also reviewed. The relative merits of laser and non-laser sources are critically examined.
Content may be subject to copyright.
Lasers Med Sci 2002, 17:173–186
Ownership and Copyright
© 2002 Springer-Verlag London Limited
Laser and Non-laser Light Sources for
Photodynamic Therapy
L. Brancaleon
1,2
and H. Moseley
1,2
1
Department of Medical Physics,
2
The Photobiology Unit, Ninewells Hospital and Medical School, Dundee, UK
Abstract. Photodynamic therapy (PDT) is an anticancer combination therapy, which requires a photo-
sensitiser, which tends to accumulate preferentially in the tumour, and light. Historically large, complex
lasers have been used to carry out PDT treatment. Nowadays there is a wide range of coherent and
non-coherent sources that can be used. This paper considers the important characteristics of light sources for
PDT, including dye lasers pumped by argon or metal vapour lasers and frequency-doubled Nd:YAG lasers.
Non-laser sources including tungsten filament, xenon arc, metal halide and fluorescent lamps are also
discussed. New exiciting developments such as LEDs and femtosecond lasers are also reviewed. The relative
merits of laser and non-laser sources are critically examined.
Keywords: Cancer treatment; Light sources; Photodynamic therapy (PDT); Photosensitisers
INTRODUCTION
Photodynamice therapy (PDT) is a treatment
modality available for palliation or eradi-
cation of several cancers. PDT involves the use
of a photoactive drug (photosensitiser) and
light (typically visible or infrared) [1,2]. Upon
absorption of light, the photosensitiser (PS)
initiates chemical reactions that lead to the
direct or indirect production of cytotoxic
species such as radicals and singlet oxygen
[3,4]. The reaction of the cytotoxic species with
subcellular organelles and macromolecules
(proteins, DNA, etc) lead to apoptosis and/or
necrosis of the cells hosting the PS. The
preferential accumulation of PSs in cancer
cells [5,6] (which in many cases can be signifi-
cant) combined with the localised delivery of
light to the tumour, results in the selective
destruction of the cancerous lesion [5]. Com-
pared to other traditional anticancer thera-
pies, PDT does not involve generalised destruc-
tion of healthy cells. In addition to direct cell
killing, PDT can also act on the vasculature,
reducing blood flow to the tumour causing its
necrosis [7,8]. In particular cases it can be used
as a less invasive alternative to surgery [9–12].
Since PDT depends on localised light deliv-
ery, it can be applied only to tumours that can
be reached by light either directly or through
an optical fibre. E$cient PDT is limited, how-
ever, by the penetration of light into the tissue,
which confines the treatment to superficial
cancer [13]. Despite its e$cacy, the application
of PDT in humans is still relatively experimen-
tal and for the treatment of the same type of
tumours protocols can vary considerably.
Quite obviously the light source and light
delivery are two of the fundamental aspects in
PDT. The choice of light source for PDT can be
dictated by the location of the tumour, by the
light dose delivered and by the choice of pho-
tosensitiser. Lasers and lamps have both been
employed to perform PDT and the superiority
of one source over the other has not been
demonstrated, therefore the use of lasers or
lamps depends on the specific application.
Although PDT has been traditionally per-
formed using lasers, the availability of broad-
band sources (lamps) is challenging the use of
lasers where light can be directly delivered to
the tumour (skin, oral cavity, etc.) without the
need to couple the source to an optical fibre.
This paper reviews the characteristics of the
light sources presently employed in PDT and
their preferred application. PDT research is
widespread and includes studies at various
stages of clinical trials, animal studies, in vitro
Correspondence to: Dr H. Moseley, University of Dundee,
The Photobiology Unit, Ninewells Hospital and Medical
School, Dundee, DD1 9SY, UK.
investigation, cellular studies, etc. We will not
review light sources used for scientific
research in PDT where the choice of sources
for PDT is somewhat less constrained, and we
will rather concentrate on sources used for
PDT in human subjects. Although PDT has
been the subject of many reviews, to date there
has not been a comprehensive description of
the light sources available to scientists and
clinicians.
PHOTOSENSITISERS AND MECHANISMS OF
PDT
In PDT, absorption of light by the PS initiates
chemical reactions that produce transient
phototoxic compounds. The mechanism of pro-
duction of these transient species has been
thoroughly described elsewhere [1,4,14,15].
Briefly photodynamic mechanisms proceed
from the first excited single state (S
1
)ofthe
photosensitiser produced by the absorption of
a photon elsewhere [14]. From S
1
the molecule
either loses an electron to originate a radical
cation (PS
.
) or quickly relaxes into the first
excited triplet state (T
1
). Both PS
.
and T
1
have a relatively long lifetime and can interact
with molecular oxygen to generate highly
reactive compounds such as peroxides and sin-
glet oxygen. The reaction that proceeds via PS
.
is normally called Type I and the one that
proceeds from T
1
is called Type II [3]. The
species produced are very reactive and can
induce oxidative stress of the cell hosting the
PS with consequent cell death via apoptosis
and/or necrosis [16,17].
The high reactivity of radicals and singlet
oxygen also produces photobleaching of the
PS. This occurs when the PS itself reacts with
the transient species to undergo reversible or
irreversible chemical reactions that lead to
the creation of photoproducts. Such photo-
products have di#erent absorption character-
istics (di#erent extinction coe$cient and/or
di#erent absorption maxima) [18,19]. Photo-
bleaching is normally regarded as a limiting
factor in PDT as it depletes the amount of PS
available during treatment. However, the role
of photobleaching in PDT is not fully under-
stood and it may also produce some benefits
such as increase the penetration of PDT into
the tumour as a result of the change in absorp-
tion at the excitation wavelength. Moreover
the role of photoproducts on PDT in vivo has
not been investigated.
The number of photosensitisers undergoing
various stages of clinical trials is large and
includes mostly various types of tetrapyrrolic
rings such as porphyrin derivatives [20–25],
phthalocyanines [15,26] and chlorins [25,27,28].
These compounds are all characterised by a
large absorption band between 400 and 430 nm
(Soret Band) [15,19] and smaller absorption
bands (Q-bands) above 550 nm [15,19]. Q-bands
above 600 nm are normally targeted for PDT
purposes; they retain high quantum yields for
Type I or II reactions and at the same time light
above 600 nm penetrates deeper into the tissue.
The absolute penetration of light depends on
the optical characteristics of the tissue, and
the geometry of light delivery [29–32]. The op-
tical penetration depth (OPD) is defined as the
depth at which the intensity of the propagating
light is attenuated approximately 37% (1/e) of
its initial value (at the air/tissue interface)
[13]. For instance in brain tissue the OPD at
635 nm is 800 m whereas in the bladder it is
4 mm [32]. Moreover, according to most tissue
modelling, light in the 600–700 nm region of the
spectrum penetrates 50–200% more than light
in the 400–500 nm region [33,34]. When the
optical properties of photosensitisers are also
considered in determining tissue penetration
[13] then penetration at 630 nm (for instance)
is 3–4 times larger than penetration at 400–
420 nm where the absorption coe$cient of pho-
tosensitisers is much larger. As a result PDT is
usually performed at wavelengths longer than
620 nm so that a larger volume of diseased
(cancerous) tissue can be treated.
These requirements have pushed the devel-
opment of light sources for PDT mostly (there
are exceptions as are described below) towards
outputs in the red region of the spectrum and,
with the advent of newer generation of PSs,
towards the near-infrared where penetration of
the incident radiation is even larger.
Despite the many studies performed using
di#erent photosensitisers, only a few have
reached the stage of advanced human clinical
trial or even FDA approval for clinical use.
These drugs include Photofrin
, Levulan
,
Foscan
and Visudyne
. The characteristics
of these photosensitisers are summarised in
Table 1. Photofrin
, Foscan
and Visudyne
are porphyrin or chlorins (two forms of
tetrapyrrolic rings) and are administered
systemically by intravenous injection and
have been used mostly for malignant or
premalignant lesions of internal organs such
as brain [35], head and neck [9], bladder [5],
174 L. Brancaleon and H. Moseley
lung [6], etc., and occasionally for malignan-
cies of the oral cavity [12]. Visudyne has been
used for treatment of age-related macular
degeneration [36] although its application to
other lesions (e.g. actinic keratosis) is under
investigation. Levulan
is the commercial
name for 5-aminolaevulinic acid (ALA), which
is a precursor of protoporphyrin IX (PPIX)
which is a clinically useful photosensitiser
[24]. By supplying ALA to cells it is possible to
overcome the negative feedback mechanisms
in the synthesis of haem [24] and accumulate
PPIX well above the physiological concen-
tration. Unlike the other PSs, ALA can be
administered topically and orally and is the
preferred choice for superficial lesions in skin
[23,37] and oral cavity [38] and it has been used
in oesophageal and stomach malignancies and
dysplasia [39]. Investigations and clinical
trials are ongoing to study the benefits of
ALA-esters [23]. These molecules introduce the
benefit of a long lipophilic chain attached to
ALA which increases penetration into tissues
and through the stratum corneum (particu-
larly important for ALA-PDT in skin) and their
metabolism still leads to the formation of
elevated intracellular levels of PPIX.
Naturally, most of the light sources for PDT
application have developed to optimise the
output near the absorption wavelengths
reported in Table 1. Moreover the tendency of
agencies, such as the US Food and Drug
Administration, has been to approve not just
the drug but the drug and the light source to
be used for its optical excitation.
PDT LIGHT SOURCES
Lasers
Historically argon lasers and metal vapour
lasers (see below) were the initial choice for
PDT. These lasers combined several character-
istics such as high power output, the possi-
bility of pumping dye lasers that would in
turn give access to the wavelength region for
excitation of porphyrins and easy coupling to
optical fibres for use with endoscopes.
Argon Lasers and Argon-pumped Dye
Lasers
Argon lasers-pumping dye lasers are among
the most popular devices for PDT treatment.
Laser dyes (such as rhodamine B, rhodamine
101 and sulphorhodamine 640) can be chosen
with absorption at one of the two main
emission lines of the argon (488 nm and
514 nm) and emission in the 600–650 nm region
to match the absorption of porphyrins
(Table 1). These lasers require a high level of
technical support. For instance, because the
argon laser beam has a narrow cross-section,
the alignment with the dye module is critical
and tends to require regular re-adjustments.
Argon lasers provide high irradiance at the
emission lines (up to 1 W/cm
2
) (Table 2). The
output of the dye laser pumped by the main
argon lines is in the range 10–500 mW/cm
2
despite the intrinsic losses of the dye laser. The
spectral output of the dye laser has a band-
width of 5–10 nm. The fluence rates reported in
Table 2 are su$cient to deliver e#ective PDT
by both direct or fibre-mediated irradiation.
Argon-pumped dye lasers coupled to an optical
fibre have been used in primary lung cancer
[40,41], oral precancer [38], oesophagus [42]
and bladder cancers [5]. The core of the fibre is
variable depending on the site treated and may
be 200–600 m. In many applications, a di#user
is fixed at the end of the fibre to allow uniform
irradiation within a lumen or tumour. The
direct expanded and attenuated beam of
argon-pumped dye lasers has also been used
for PDT of superficial skin cancer [10] and
for PDT of vulval neoplasia [43]. The direct
Table 1. Summary of the main characteristics of the most common commercially available photosensitisers
Commercial name Chemical definition
Absorption
maximum
Delivery
Photofrin Mixture of di-hematoporphyrin esters and ethers 630 nm Systemic
Foscan Meta-tetrahydroxyphenylchlorin (m-THPC) 652 nm Systemic
Visudyne Benzoporphyrin derivative 690 nm Systemic
Levulan 5-Aminolaevulinic acid (ALA)
converted into protoporphyrin IX (PPIX)
635 nm Oral/topical
Laser and Non-laser Light Sources for Photodynamic Therapy 175
Table 2. Types of lasers available for clinical PDT
Wavelength(s) Bandwidth Irradiance
Pulse
duration
Light
delivery
Argon laser 488 and 514.5 nm Monochorom 0.5–1 W/cm
2
CW Direct or optical fibre
Dye laser pumped by argon laser 500–750 nm (depending on the dye) 5–10 nm 10–200 mW/cm
2
CW Direct or optical fibre
Metal vapour laser UV or visible (depending on metal) Monochrom Up to 10 W/cm
2
10–50 ns quasi-CW Direct or optical fibre
Dye laser pumped by
metal vapour laser
500–750 nm (depending on the dye) 5–10 nm 10–500 mW/cm
2
10–50 ns quasi-CW Direct or optical fibre
Solid state For a Nd:Yag 1064, 532, 355, 266 nm Monochrom Up to 10 W/cm
2
10 ps–30 ns quasi-cw Direct or optical fibre
Dye laser pumped by
solid state laser
400–750 nm (depending on dye) 5–10 nm 10–500 mW/cm
2
10 ps–30 ns quasi-cw Direct or optical fibre
Solid state optical
parametric oscillator
250–2000 nm Monochrom Up to 1 W/cm
2
10 ps–30 ns Direct or optical fibre
Semiconductor diode lasers 600–950 nm Monochrom Up to 700 mW/cm
2
CW Optical fibre
176 L. Brancaleon and H. Moseley
monochromatic emission of the argon laser has
not been widely employed in PDT because of
the lower penetration into the tissue of the 488
and 514 nm wavelengths. However, prelimi-
nary animal experiments are ongoing using
the line at 514 nm [44]. As a final note we
would like to point out that argon lasers (and
argon-pumped dye lasers) are especially indi-
cated for endoscopic PDT because the output
beam has a very small cross-section (<1 mm)
and can readily be coupled to optical fibres.
Conversely argon lasers are not the most con-
venient choice in typically large skin or oral
lesions where its use involves the addition of a
beam expander which can become cumbersome
and reduce the fluence rate.
Metal Vapour-pumped Dye Laser
These lasers have also been (and still are) a
popular choice for PDT particularly among
European investigators. Unlike truly CW
argon lasers, metal vapour lasers are normally
pulsed, with pulsewidth ranging from 10 to
50 ns (Table 2) and pulse rates of 1 KHz. Such a
high repetition rate makes the source quasi-
continuous for clinical purposes. The pump
beam (in the UV or visible depending on the
metal mixture) provides high primary output
power that can be used to pump tunable
dye lasers, which in turn give access to the
spectral region where porphyrins absorb.
Metal vapour-pumped dye lasers are able to
deliver light at irradiance up to several
hundred mW/cm
2
(Table 2). These lasers can be
coupled to optical fibres and used for endo-
scopic PDT such as in oral precancer [12], in
head and neck cancer [9], the oesophagus [45],
lung [11], bladder [5]. The bandwidth of the dye
laser is the same as that of argon-pumped dye
lases. Because of their large beam cross-
section (typically 1–3 cm
2
), the metal vapour
laser can be applied for PDT of large lesions
such as those occurring in the skin [46,47]
without the need to use a beam expander. As
with argon laser systems these lasers do
require a good level of technical support.
However, because of the large cross-section of
the pump-laser beam, alignment with the dye
module is not as critical as for argon lasers.
Solid State Lasers
Solid state lasers such as Nd:YAG lasers are a
more recent development in laser technology
and can be applied in PDT similarly to argon
and metal vapour lasers. They o#er more
compact design than the previous lasers with
obvious advantages for laboratory or clinical
use. They are normally pulsed at higher
rates (MHz) and shorter pulsewidths (sub-
nanosecond). These lasers normally emit a
fundamental line in the near infrared (e.g. for
Nd:YAG at 1064 nm). The output from the
fundamental line has energies of up to seven
J/pulse. Pulses from Q-switched NdYAG lasers
are 5–10 ns which translates into large peak
energies and irradiances (Table 2) and can
e$ciently undergo frequency doubling to give
lines in the visible (532 nm for Nd:YAG) or in
the UV (266 nm for Nd:YAG) with energies of
up to 50 mJ/pulse. The frequency-doubled out-
put can then be used to pump a dye laser and
obtain high power output in the region of
porphyrin absorption with the same bandwidth
as for the other laser-pumped dye lasers.
Optical Parametric Oscillators Lasers
Optical parametric oscillator (OPO) lasers can
be used for irradiation of up to several hun-
dred mW/cm
2
. OPO are solid state-based
pulsed lasers that via frequency doubling and
wave-mixing give access to a large number of
monochromatic wavelengths from the UV to
the near-IR region of the spectrum [48]. Wave-
length tunability and fluence rate for PDT are
easily obtainable with these lasers (Table 2).
Solid state lasers have been applied for PDT
of skin lesions, oesophageal cancer [39,49],
oral precancer and cancer [9,39], lung [50] and
bladder [5]. A potential future advantage of
solid state lasers compared to argon and metal
vapour is the possibility of using the near
infrared monochromatic fundamental emission
of these lasers. Indeed one of the possible
developments of PDT is the synthesis of photo-
sensitisers that can be excited in the near
infrared where radiation would penetrate
deeper into tissues and extend PDT treat-
ment to less superficial tumours. Solid state
lasers would then be one of the main sources
available.
Diode Lasers
These lasers represent a potential major
breakthrough in the widespread clinical use of
PDT. Lasers made with semiconductors are
extremely compact (Fig. 1a) yet retain high
output (Table 2). They are extremely versatile
as they can be used in CW mode or be pulsed
Laser and Non-laser Light Sources for Photodynamic Therapy 177
(picosecond to millisecond). Their bandwidth
is typically 6 nm; the power supply is also
compact and they are normally air-cooled.
These lasers are very attractive for clinical use
as they are easy to operate and portable for use
in laboratory and clinical settings. They are
normally coupled to optical fibres and are ideal
for endoscopic PDT. Diode lasers have been
used in the treatment of a variety of lesions
in the skin, oral cavity and in the eye [51],
pituitary adenomas [35], and also to treat age-
related macular degeneration [36]. The fibre
output can also be expanded for use in large
lesions in the skin. At present diode lasers tend
to o#er only a single output wavelength, which
limits their versatility. However, systems are
being developed that will allow interchange-
able laser modules with multiple wavelengths.
We foresee that in the future these lasers will
be more widely used for PDT.
Lamps
Lasers are not the only option for PDT. In
clinical settings especially, several PDT
sources now use filtered output high power
lamps (Fig. 1b,c). General maintenance of
lamps is normally easier and cheaper. In com-
parison with lasers, lamps emit a much wider
spectral output. Because of the broad emission
spectrum of lamps, a combination of narrow-
band, longpass and shortpass filters are often
required. Narrowband filters select the
irradiation wavelength within 10 nm, longpass
filters help to cut high-power UV radiation
associated with the lamp output and shortpass
filters are usually necessary to cut IR emission
from the filament which could cause heating of
the treated area and may also damage the
optics of the lamp. Although the combination
of PDT and hyperthermia (due to IR radiation)
has been suggested [52], in skin this is avoided
as hyperthermia is associated with higher
levels of pain.
Therefore, instead of a high intensity mono-
chromatic source they produce high intensity
over a larger spectral range. The superiority
of monochromatic over broadband light
delivery has not been demonstrated for PDT.
The e#ectiveness of light sources depends
on several factors, some of which can be
simplified under the concept of ‘total e#ective
fluence rate’ [53] which combines incident
spectral irradiance, tissue transmission and
the absorption properties of the photo-
sensitiser. Dosimetry is important if meaning-
ful comparisons are to be made between di#er-
ent light sources. If a laser is being used, then
the wavelength is clearly identified. This
becomes more complex with a broadband light
source and in this case the spectral irradiance
should always been given. What is required is
the e#ective photodynamic dose. Total e#ec-
tive fluence rate indicates that, for instance,
light in the green region of the spectrum may
be more e#ective within a depth of 2 mm,
beyond which red light appears to be superior
for PDT. Lamps are portable and easy to use.
They deliver light over a large area and can be
(a)
(b)
(c)
Fig. 1. (a) Example of operating diode laser (Diomed 630);
(b) example of operating tungsten filament lamp (Photocure);
(c) example of metal halide lamp (Waldmann PDT 1200).
178 L. Brancaleon and H. Moseley
coupled to large cross-section light guides and
are therefore suitable for the treatment of
large superficial lesions. Conversely the out-
put of a lamp cannot be easily coupled into
small optical fibres without greatly limiting
their power output. For these reasons the
use of lamps has been limited to skin lesion
and they have not been used for endoscopic
PDT.
Tungsten Filament Quartz Halogen Lamps
These are essentially incandescent sources
where the temperature of the tungsten fila-
ment is raised to approximately 3000K. At this
temperature there is a considerable amount of
optical radiation emitted from UV to near-IR.
The use of these lamps for PDT was introduced
by Pottier and Kennedy [54] who conducted
animal experiments using the filtered output of
a slide projector. Since then, commercial appli-
cations of the concept have been developed.
These lamps can deliver up to 250 mW/cm
2
over a wide spectrum (350–850 nm) (Table 3). A
single wavelength can be selected using com-
binations of long-pass, and narrowband filters.
The output of these sources can be coupled
into a liquid light guide (up to 1 cm in
diameter) or expanded to several cm
2
. These
lamps have been mainly employed for topical
ALA-PDT (i.e. PDT performed after topical
application of 5-aminolaevulinic acid) in
which the targeted photosensitiser is proto-
porphyrin IX whose absorption maximum is at
635 nm (see above). A representative spectral
output, recorded using a calibrated double-
grating spectroradiometer, is shown in Fig. 2.
Xenon Arc Lamps
In these lamps, radiation is provided by an
electrical arc that forms between the elec-
trodes in the presence of Xenon vapour. They
are another possible light source for PDT.
They are characterised by a broad spectral
emission (300–1200 nm) and by high output (up
to 8 W for direct exposure and up to 1 W using
a liquid light guide) leading to potential
fluence rates of several hundred mW/cm
2
.
Combination of band-pass and narrow-band
filters can be used to eliminate IR radiation
(consequently heat) and to select irradiation
wavelengths within 60 nm (Table 3).
Although many of these light sources were
assembled and used by research laboratories
[23], a number of them are also available com-
mercially. These sources have mostly been
used for PDT of non-melanoma skin cancer
and other skin disorders [23,55–58].
Metal Halide Lamps
These lamps comprise a mixture of mercury
and metal halide vapour that is ignited by an
electrical discharge. This produces a broad
emission spectrum superimposed on a series of
emission lines that depend on the gas used to
fill the bulb. These light sources are another
example of broadband lamps that can be used
to perform PDT. The emission spectrum from
one of the commercially available sources,
recorded using a calibrated double-grating
spectroradiometer, is shown in Fig. 3. Selec-
tion of the waveband can be achieved with
filters to obtain spectral output in the 590–
720 nm range (Fig. 3). The irradiance range is
between 10 and 250 mW/cm
2
and the treat-
ment area is potentially large (up to 20 cm in
diameter) (Table 3). Unlike xenon arc and
tungsten filament lamps, these sources are
always used for direct exposure rather than
coupled into a large core liquid light guide.
Like the other broadband sources, metal
halide lamps have been mostly employed in
PDT of superficial lesions such as non-
melanoma skin cancer [47,59], vulval intraepi-
thelial neoplasia [60,61]. Although most of
these lamps have been home built [47], exam-
ples of commercial metal halide lamps can also
be found.
Phosphor-coated Sodium Lamp
The working principle of this lamp is similar
to the metal halide lamps where an electric
Fig. 2. Emission spectrum of the Photocure PDT Lamp
(Tungsten filament lamp). The spectrum was recorded in our
laboratory using a calibrated double-monochromator spectro-
radiometer.
Laser and Non-laser Light Sources for Photodynamic Therapy 179
Table 3. Types of lamps available for clinical PDT
Wavelength(s) Bandwidth Irradiance Light delivery
Tungsten filament 400–1100 nm 10–100 nm (depending on filters used) Up to 250 mW/cm
2
or
typically up to 1.8 mW/cm
2
/nm
Direct or via
liquid light guide
Xenon arc 300–1200 nm 10–100 nm (depending on filters used) Up to 300 mW/cm
2
or
typically up to 3 mW/cm
2
/nm
Normally liquid light guide
Metal halide Depending on the metal,
lines between 250–730 nm
(can be phosphor coated)
10–100 nm (depending on filters used) Up to 250 mW/cm
2
or
typically 1.2 mW/cm
2
/nm
Direct or liquid light guide
Sodium (phosphor coated) 590–670 nm 10–80 nm (depending on filters) Up to 100 mW/cm
2
Direct illumination
Fluorescent 400–450 nm Approximately 30 nm Up to 10 mW/cm
2
Direct illumination
180 L. Brancaleon and H. Moseley
discharge is produced in the presence of
sodium vapour. The surface of the bulb is
coated with phosphors that absorb the sodium
lines and emit in a di#erent region acting
similarly
to the dye lasers described previously. The
spectral output includes wavelengths in the
590–670 nm region and the intensity is in
the 25–100 mW/cm
2
range (Table 3). Similarly
to the metal halide lamp, the area illuminated
is large (up to 100 cm
2
) and it can be used to
perform PDT of skin lesions [62].
Fluorescent Lamps
These sources represent a di#erent approach
to PDT. As discussed above, light sources for
PDT were developed to emit in the 600–700 nm
region as in this region the photosensitisers
currently used in therapy or clinical trials
have one of their absorption maxima and light
penetrates deeper into the tissue allowing
treatment of thicker lesions. As described
earlier photosensitisers have a more intense
absorption band in the region between 400 and
450 nm (Soret Band). Fluorescent lamps have
been developed to match this region of the
spectrum. Higher absorption coe$cients of
the PS produce equal or higher e$ciency of the
photodynamic e#ect with a lower concen-
tration of the drug in the tissue (smaller
amount of drug administered either topically
or systematically). Treatment, however, is
limited to very superficial skin lesions since
the penetration of light between 400 and
450 nm is approximately 300–400 m. Fluor-
escent lamps for PDT have a maximum near
4175 nm and bandwidth of 30 nm. The power
output is only 10 mW/cm
2
but their ease of use
can be attractive for clinical settings. Their
spectral characteristics match the absorption
of protoporphyrin IX and their use has been
limited to topical ALA-PDT of superficial skin
lesions [63].
Laser vs. Lamp
As briefly discussed earlier, to date there has
not been a thorough comparison between
lasers and lamps in treating the same type of
tumours in vivo and only sporadic studies have
been reported [47,64,65]. We believe that such
a comparison is fundamental for the develop-
ment of PDT, and in our group, studies are
ongoing on this particular topic. Lasers
provide a monochromatic, very powerful
source of light that can reduce the time neces-
sary to deliver the final PDT dose. Because
they are monochromatic the choice of laser
wavelength becomes crucial as it must be
matched with the often narrow absorption
band of the photosensitiser (see Table 1) with
the result that one laser can only be used in
combination with one (or a limited number)
PS. On the other hand, lamps provide a broad
range of wavelengths at reduced fluence rates.
Since most investigators limit fluence rates to
relatively low values of 100–300 mW/cm
2
,to
avoid thermal e#ects, the use of lamps does not
necessarily produce a dramatic increase in the
time required for the treatment. Because of
their broad emission, lamps can be used in
combination with several PSs with di#erent
absorption maxima within the emission
spectrum of the lamp. So, the same lamp could
be used for PDT with Foscan, Photofrin or
ALA (Tables 1 and 2). Moreover, lamps
normally also excite the region where photo-
products absorb. Although the role of photo-
products in PDT is unclear, it is possible that
some additional PDT e#ect can be obtained by
photoproducts themselves. Lasers at present
are the only possible light source to treat
malignancies located in sites that can be
reached only with optical fibres. Beam quality,
dedicated optical accessories and power out-
put are among the characteristics that make
lasers the only real choice if light has to be
coupled to an optical fibre with cores smaller
than 500 m in diameter. Because of the possi-
bility of using light di#users of di#erent shapes
and microlenses to produce uniform collimated
Fig. 3. Emission spectrum of the Waldmann PDT 1200
Lamp (metal halide lamp). The spectrum was recorded in our
laboratory using a calibrated double-monochromator spectro-
radiometer.
Laser and Non-laser Light Sources for Photodynamic Therapy 181
beams, lasers are also suitable for use in direct
illumination of lesions located in accessible
sites (such as skin or oral cavity). Lamps on
the other hand cannot be used in combination
with small optical fibres because of the poor
beam quality, large beam size and small power
density. They can, however, be used direct or
coupled to a liquid light guides of between 5
and 10 mm in diameter. Moreover, compared to
lasers, lamps are normally less expensive and
more user friendly. Because of their character-
istics lamps are well suited for treatment of
accessible lesions especially for larger skin
lesions (with or without the use of liquid light
guides).
Other Sources
With constant advancement in photonics tech-
nology, new sources are constantly developed
and will be available in the near future for
large-scale use in PDT.
Light Emitting Diodes (LED)
In the past few years the development of LED
has advanced them to a stage where their use
in phototherapy (and PDT in particular) is
possible. LED would o#er several advantages
for clinical and laboratory use. The choice of
emission wavelength ranges from UVA
(350 nm) to near infrared (1100 nm). The band-
width is 5–10 nm and the power output can
provide up to 150 mW/cm
2
over an area of
approximately 20 cm
2
(Table 3). The power out-
put can still be a limiting factor in their
widespread use for PDT, however further
improvement in their technology could
improve this aspect. Two major characteristics
in favour of the use of LED are price and
versatility. LED are inexpensive (in compari-
son with all the other sources described so far),
therefore they can be arranged in arrays to
irradiate large areas. They can be powered by
batteries, making them totally and easily port-
able. Moreover, they can be arranged in di#er-
ent geometric combination to compensate for
di$cult anatomic areas (non-melanoma skin
cancer for instance tends to occur in the face
and the head where large curvatures may
reduce the e$cacy of other light delivery
systems). Prototypes for the use of LED in
phototherapy and PDT are currently under
development.
Femtosecond Solid State Lasers
The use of femtosecond lasers has been
proposed for possible two-photon PDT.
Femtosecond lasers are presently used for
two-photon excitation in several advanced
research areas (microscopy and spectroscopy
[66]). Two-photon excitation is based on the
observation that when the incident light is
characterised by a high photon density two
photons of equal energy can be simultaneously
absorbed by a chromophore [67] to excite an
electron to an energy level that is equal to the
sum of the energy of the two absorbed photons.
Therefore, for instance, the excitation of
porphyrins in the 400–450 nm region can be
obtained using light of high photon density in
the 800–900 nm region. To obtain such high
photon density, however, the laser has to emit
a large number of photons during a very short
pulse and the emitted photons have to be
focused into a very small volume. These char-
acteristics match the current femtosecond
solid state lasers (Table 4), that have high
energy, very short pulses and can be focused
into very small (1–5 m
3
) volumes creating
the necessary photon density to produce two
photon absorption. It has been shown that
two-photon excitation can be achieved in
tissues and in vivo [66]. The advantage intro-
duced by these lasers is related to the exci-
tation wavelength. The incident radiation
would be in the near infrared and as a result
its penetration into the tissue would be much
Table 4. Examples of new sources potentially useful for clinical PDT in the near future
Wavelength(s) Bandwidth Irradiance
Pulse
duration
Light
delivery
Solid state lasers
for two photon
PDT
Near infrared Monochorom 1 W in a volume
of approximately
5–10 m
3
0.1–10 ps Direct, scanned
over the lesion
LED Visible and infrared 5–10 nm Up to 150 mW/cm
2
CW Direct
182 L. Brancaleon and H. Moseley
larger [33,34,66] and would increase the depth
of PDT. Despite this attractive advantage
especially for the treatment of less superficial
lesions, their current limitations outweigh the
advantages. Femtosecond lasers are di$cult to
maintain and to operate and would require
additional specialised personnel. More import-
ant, however, is the necessity of scanning the
beam over the lesions. Because two-photon
absorption occurs within a very small volume
the treatment of a lesion would require raster
scanning the incident radiation over the lesion
with additional technical problems and longer
treatment time.
CONCLUSIONS
The choice of photosensitisers for PDT is still
limited compared to the choice of laser sources
available for treatment and the process for
approval of new photosensitisers has proven
extremely lengthy. Nonetheless, the wide-
spread availability of potentially useful light
sources means that PDT is no longer limited to
centres with high technical expertise; rather it
is now a treatment option, which may be ex-
ploited more widely. Historically, large com-
plex lasers were required for PDT, limiting its
use to centres that could provide the necessary
technical support. In clinical settings these
lasers have been replaced by reliable, easy-to-
use light sources which no longer require com-
plex technologies and expensive maintenance.
This means that PDT can become widely avail-
able as a realistic treatment option and it is
likely that the interest in this therapeutic
modality will increase. The targeted organ,
photosensitiser, reliability, ease of use, cost
and space are the most important variables
that need to be considered in a clinical setting.
In the past decade there has been renewed
interest in the development of both laser and
non-laser light sources for PDT and more
patients will be able to benefit from this
treatment.
ACKNOWLEDGEMENTS
The authors would like to thank the Barbara Stewart
Scottish Laser Centre Trust for Cancer.
APPENDIX
In this appendix we will summarise some of the makers of
light sources for PDT. We acknowledge the fact that light
sources with the proper characteristics for PDT can be
assembled by the investigators in laboratory or clinical
settings, nonetheless we list commercially available
sources specifically designed for clinical PDT. This review
does not intend to advertise one product rather than
another and does not state the superiority of one particu-
lar product. The list is as complete as possible (to our
knowledge) and is limited to companies that advertise
their product for specific PDT application. We apologise if
some manufacturers are not listed.
Lasers
Argon and argon-pumped dye lasers
Coherent Inc., Santa Clara, CA, USA
(www.coherentinc.com)
Spectra-Physics, Mountain View, CA, USA
(www.spectraphysics.com)
Metal vapour and metal vapour-pumped dye lasers
Oxford Lasers, Oxford, UK
(www.oxfordlasers.com)
Solid state
Laserscope, San Jose, CA, USA
(www.laserscope.com)
Coherent Inc., Santa Clara, CA, USA
(www.coherentinc.com)
Diode Lasers
Diomed, Cambridge, UK and Andover, MA,
USA (www.diomed-lasers.com)
Applied Optronics Corp., South Plainfield,
NJ, USA (www.appliedoptronicscorp.com)
Coherent Inc., Santa Clara, CA, USA
(www.coherentinc.com)
Oxford Optronix, Oxford, UK
(www.oxfordshire.co.uk)
Ceramoptec, Bonn, Germany
(www.ceramoptec.com)
Carl Zeiss, Oberkochen, Germany
(www.zeiss.de)
Two-photon technology
Coherent Inc., Santa Clara, CA, USA
(www.coherentinc.com)
Lamps
Tungsten filament
MBG Technologies (Lumacare),
Newport Beach, CA, USA
(www.mbgtech.com or www.lumacare.com)
Photocure, Oslo, Norway
(www.photocure.com)
Xenon arc
Photo Therapeutics, Altrincham, UK
(www.phototherapeutics.co.uk)
ESC Medical Systems, Yokneam, Israel
(www.escmed.com)
Metal halide
Waldmann, Villingen-Schwenningen,
Germany (www.waldmann.com)
Phosphor-coated sodium
Medeikonos, Goteborg, Sweden
(www.medeikonos.com)
Fluorescent
Dusa Pharmaceuticals, Wilmington, MA,
USA (www.dusapharma.com)
Laser and Non-laser Light Sources for Photodynamic Therapy 183
LED
PRP Optoelectronics, Towcester, UK
(www.prpopto.co.uk)
REFERENCES
1. Dougherty TJ. Photodynamic therapy. Photochem
Photbiol 1993;58:895–900.
2. Henderson BW, Dougherty TJ. How does photo-
dynamic therapy work? Photochem Photobiol 1992;
55:145–57.
3. Girotti AW. Photosensitized oxidation of cholesterol
in biological systems: reaction pathways, cyto-
toxic e#ects and defense mechanisms. J Photochem
Photobiol 1992;13:105–18.
4. Nonell S, Redmond RW. On the determination of
quantum yields for singlet molecular oxygen photo-
sensitization. J Photochem Photobiol B: Biol 1994;
22:171–2.
5. Pope AJ, Bown SG. Photodynamic therapy. Br J Urol
1991;68:1–9.
6. Lam S. Photodynamic therapy of lung cancer. Semin
Oncol 1994;21:15–19.
7. Fingar VH. Vascular e#ects of photodynamic therapy.
Clin Laser Med Surg 1996;14:323–8.
8. Fingar VH, Kik PK, Haydon PS, Cerrito PB, Tseng M,
Abang E et al. Analysis of acute vascular damage
after photodynamic therapy using Benzoporphyrin
derivative (BPD). Br J Cancer 1999;79:1702–8.
9. Dilkes MG, DeJode ML, Rowntree-Taylor A,
McGilligan JA, Kenyon GS, McKelvie P. m-THPC
photodynamic therapy for head and neck cancer.
Lasers Med Sci 1996;11:23–9.
10. Edell ES, Cortese DA. Photodynamic therapy in
the management of early superficial squamous cell
carcinoma as an alternative to surgical resection.
Chest 1992;102:1319–22.
11. Furuse K, Fukuoka M, Kato H, Horai T, Kubota K,
Kodama N et al. A prospective phase II study on
photodynamic therapy with photophrin II for centrally
located early stage lung cancer. J Clin Oncol
1993;11:1852–7.
12. Grant WE, Hopper C, Speight PM, Path MRC,
Macrobert AJ, Bown SG. Photodynamic therapy of
malignant and premalignant lesions in patients with
‘field cancerization’ of the oral cavity. J Laryngol Otol
1993;107:1140–5.
13. Profio AE, Doiron DR. Transport of light in tissue
in photodynamic therapy. Photochem Photobiol 1987;
46:591–9.
14. Girotti AW. Photodynamic lipid peroxidation in bio-
logical systems. Photochem Photobiol 1990;51:497–509.
15. Reddi E, Jori G. Steady-state and time-resolved
spectroscopic studies of photodynamic sensitizers:
porphyrins and phtalocyanines. Rev Chem Interm
1988;10:241–68.
16. Ahmad N, Feyes DK, Agarwal R, Mukhtar H. Photo-
dynamic therapy results in induction of WAF1/CIP1/
P21 leading to cell cycle arrest and apoptosis. Proc
Natl Acad Sci 1998;95:6977–82.
17. LaMuraglia GM, Schiereck J, Heckenkamp J, Nigri G,
Waterman P, Leszcynski D et al. Photodynamic
therapy induces apoptosis in intimal hyperplastic
arteries. Am J Pathol 2000;157:867–75.
18. Konig K, Schneckenburger H, Ruck A, Steiner R.
In vivo photoproduct formation during PDT with
ALA-induced endogenous porphyrin. J Photochem
Photobiol B: Biol 1993;18:287–90.
19. Aveline BM, Hasan T, Redmond RW. The e#ects of
aggregation protein binding and cellular incorpor-
ation on the photophysical properties of benzo-
porphyrin derivative monoacid ring A (BPDMA).
J Photochem Photobiol B: Biol 1995;30:161–9.
20. Bilsel O, Buchler JW, Hammerschmitt P, Rodriguez J,
Holten D. Electronic states and (,*) absorption and
emission characteristics of strongly coupled porphyrin
dimers: sandwich complexes of Hf
IV
and Zr
IV
. Chem
Phys Lett 1991;182:415–21.
21. Candide C, Morliere P, Maziere JC, Goldstein S,
Santus R, Dubertret L et al. In vitro interaction of the
photoactive anticancer porphyrin derivative photofrin
II with low density lipoprotein, and its delivery to
cultured human fibroblasts. FEBS 1986;207:133–8.
22. Faustino MAF, Neves MGPMS, Cavaleiro JAS,
Neumann M, Brauer HD, Jori G. Part 2. Meso-
tetraphenylporphyrin dimer derivatives as potential
photosensitizers in photodynamic therapy. Photochem
Photobiol 2000;72:217–25.
23. Gerscher S, Connely JP, Gri$ths J, Brown SB,
MacRobert AJ, Wong G et al. Comparison of the
pharmacokinetics and phototoxicity of protoporphyrin
IX metabolized from 5-aminolevulinic acid and
two derivatives in human skin in vivo. Photochem
Photobiol 2000;72:569–74.
24. Kennedy JC, Pottier RH. Endogenous protoporphyrin
IX, a clinically useful photosensitiser for photo-
dynamic therapy. J Photochem Photobiol B: Biol
1992;14:275–92.
25. Roeder B, Wabnitz H. Time-resolved fluorescence
spectroscopy of hematoporphyrin, mesoporphyrin,
pheophorbide a and chlorin e
6
in ethanol and aqueous
solutions. J Photochem Photobiol B: Biol 1987;1:
103–13.
26. Cuomo V, Jori G, Rihter B, Kenney ME, Rodgers
MAJ. Liposome-delivered Si(IV)-naphtalocyanine as a
photodynamic sensitiser for experimental tumours:
pharmacokinetic and phototherapeutic studies. Br J
Cancer 1990;62:966–70.
27. Pogue BW, Redmond RW, Trivedi N, Hasan T. Photo-
physical properties of tin ethyl etiopurpurin I (SnET
2
)
and tin octaethylbenzochlorin (SnOEBC) in solution
and bound to albumin. Photochem Photobiol 1998;
68:809–15.
28. Soukos NS, Hamblin MR, Hasan T. The e#ect of
change on cellular uptake and phototoxicity of poly-
lysine chlorin
e6
conjugates. Photochem Photobiol
1997;65:723–9.
29. Gardner CM, Jacques SL, Welch AJ. Fluorescence
spectroscopy of tissue: recovery of intrinsic fluor-
escence from measured fluorescence. Appl Opt
1996;35:1780–92.
30. Richards-Kortum R, Rava RP, Fitzmaurice M, Tong
LL, Ratli# NB, Feld MS. A one-layer model of laser-
induced fluorescence for diagnosis of disease in human
tissue: application to atherosclerosis. IEEE Trans
Biomed Eng 1989;36:1222–32.
31. Richards-Kortum R, Sevick-Muraca E. Quantitative
optical spectroscopy for tissue diagnosis. Annu Rev
Phys Chem 1996;47:555–606.
32. Shackley DC, Whitehurst C, Moore JV, George NJR,
Betts CD, Clarke NW. Light penetration in bladder
184 L. Brancaleon and H. Moseley
tissue: implication for the intravescical photodynamic
therapy of bladder tumours. BJU Int 2000l86:638–43.
33. Keijzer M, Richards-Kortum R, Jacques SL, Feld MS.
Fluorescence spectroscopy of turbid media: auto-
fluorescence of the human aorta. Appl Opt 1989;
28:4286–92.
34. Wu J, Feld MS, Rava RP. Analytical model for extract-
ing intrinsic fluorescence in turbid media. Appl Opt
1993;32:3585–95.
35. Marks PV, Belchetz PE, Saxena A, Igbaseimokumo U,
Thomson S, Nelson M et al. E#ect of photodynamic
therapy on recurrent pituitary adenomas: clinical
phase I/II an early report. Br J Neurosurg 2000;
14:317–25.
36. Fong DS. Photodynamic therapy with verteporfin for
age-related macular degeneration. Ophtalmology 2000;
107:2314–17.
37. Rhodes LE, Tsoukas MM, Anderson RR, Kollias N.
Iontophoretic delivery of ALA provides a quantitative
model for ALA pharmacokinetics and PPIX photo-
toxicity in human skin. J Invest Dermatol 1997;108:87–
91.
38. Kubler A, Haase T, Rheinwald M, Barth T, Muhling J.
Treatment of oral leukoplakia by topical application
of 5-aminolevulinic acid. Int J Oral Maxillofac Surg
1998;27:466–9.
39. Gossner L, May A, Sroka R, Stolte M, Hahn EG, Ell C.
Photodynamic destruction of high grade dysplasia and
early carcinoma of the esophagus after oral admin-
istration of 5-aminolevulinic acid. Cancer 1999;
86:1921–7.
40. Okunaka T, Harubimi K, Konaka C, Kawate N, Bon-
aminio A, Yamamoto H et al. Photodynamic therapy
for multiple primary bronchogenic carcinoma. Cancer
1991;68:253–8.
41. Marijnissen JPA, Baas P, Beck JF, van Moll JH, van
Zandwijk N, Star WM. Pilot study on light dosimetry
for endobronchial photodynamic therapy. Photochem
Photobiol 1993;58:92–9.
42. Krishnadath KK, Wang KK, Taniguchi K, Sebo TJ,
Buttar NS, Anderson MA et al. Persistent genetic
abnormalities in Barrett’s esophagus after photo-
dynamic therapy. Gastroenterology 2000;119:624–30.
43. Hillemanns P, Untch M, Danneker C, Baumgartner R,
Stepp H, Diebold J et al. Photodynamic therapy
of vulvar intraepithelial neoplasia using
5-aminolevulinic acid. Int J Cancer 2000;85:649–53.
44. van der Veen N, Hebeda KM, de Bruijn HS, Star WM.
Photodynamic e#ectiveness and vasoconstriction in
hairless mouse skin after topical 5-aminolevulinic
acid and single- or two-fold illumination. Photochem
Photobiol 1999;70:921–9.
45. Ackroyd R, Brown NJ, Davis MF, Stephenson TJ,
Marcus SL, Stoddard CJ et al. Photodynamic therapy
for dysplastic Barrett’s oesophagus: a prospective,
double blind, randomised, placebo controlled trial. Gut
2000;47:612–17.
46. Itoh Y, Ninomiya Y, Tajima S, Ishibash A. Photody-
namic therapy for acne vulgaris with topical
5-aminolevulinic acid. Arch Dermatol 2000;136:1093–5.
47. Soler AM, Angell-Petersen E, Warloe T, Tausio J,
Steen HB, Moan J et al. Photodynamic therapy
of superficial basal cell carcinoma with
5-aminolaevulinic acid with dimethylsulfoxide and
ethylendiaminetetraacetic acid: a comparison of two
light sources. Photobiol 2000;71:724–9.
48. Ebrahimzadeh M, Dunn MH. In: Bass M (ed) Hand-
book of Optics Vol IV, New York: McGraw-Hill,
2000:22.1–22.72.
49. van de Boogert J, van Staveren HJ, de Bruin RWF,
de Rooij FWM, Edixhoven-Bosdijk A, Siersema PD
et al. Fractionated illumination in oesophageal ALA-
PDT: e#ect on ferrochelatase activity. J Photochem
Photobiol B: Biol 2000;56:53–60.
50. Shah SK, Ost D. Photodynamic therapy. A case series
demonstrating its role in patients receiving mechan-
ical ventilation. Chest 2000;118:1419–23.
51. Karamata B, Sickenberg M, van den Bergh H. A fibre
optic light distributor for the preventive photo-
dynamic therapy of secondary cataract. Laser Med Sci
2000;15:238–45.
52. Tsujino I, Anderson GS, Sieber F. Postirradiation
hyperthermia selectively potentiates the merocyanine
540-sensitized photoinactivation of small cell lung
cells. Photochem Photobiol 2001;73:191–8.
53. Moseley H. Total e#ective fluence: a useful concept
in photodynamic therapy. Lasers Med Sci 1996;11:139–
43.
54. Kennedy JC, Marcus SL, Pottier RH. Photodynamic
therapy (PDT) and photodiagnosis (PD) using endog-
enous photosensitization induced by 5-aminolevulinic
acid (ALA): mechanisms and clinical results. J Clin
Laser Med Surg 1996;14:289–304.
55. Morton CA, Whitehurst C, Moseley H, McColl JH,
Moore JV, MacKie RM. Comparison of photodynamic
therapy with cryotherapy in the treatment of Bowen’s
disease. Br J Dermatol 1996;135:766–71.
56. Morton CA, Whitehurst C, Moore JV, MacKie RM.
Comparison of red and green light in the treatment of
Bowen’s disease by photodynamic therapy. Br J Der-
matol 2000;143:767–72.
57. Thissen MR, Neumann MH, Schouten LJ. A system-
atic review of treatment modalities for primary
basal cell carcinomas. Arch Dermatol 1999;135:1177–
83.
58. Thissen MRTM, Schroeter CA, Neumann HAM.
Photodynamic therapy with delta-aminolaevulinic
acid for nodular basal cell carcinomas using
prior debulking technique. Br J Dermatol
2000;142:338–9.
59. van den Akker JTHM, de Bruijn HS, Beijersbergen
van Henegouwen GM, Star WM, Sterenborg HJCM.
Protoporphyrin IX fluorescence kinetics and localiz-
ation after topical application of ALA pentyl ester
and ALA on hairless mouse skin with UVB-induced
early skin cancer. Photochem Photobiol 2000;72:399–
406.
60. Kurwa HA, Barlow RJ, Neill S. Single-episode
photodynamic therapy and vulval intraepithelial neo-
plasia type III resistant to conventional therapy. Br J
Dermatol 2000;143:1040–2.
61. Henta T, Itoh Y, Kobayashi M, Ninomiya Y, Ishibashi
A. Photodynamic therapy for inoperable vulval
Padget’s disease using delta-aminolevulinic acid:
successful management of a large skin lesion. Br J
Dermatol 2999;141:347–9.
62. Wennberg AM, Gudmundson F, Stenquist B,
Ternesten A, Molne L, Rosen A et al. In vivo detection
of basal cell carcinoma using imaging spectroscopy.
Acta Derm Venereol 1999;79:54–61.
63. Bissonnette R, Shapiro J, Zeng H, Mclean DI, Liu H.
Topical photodynamic therapy with 5-aminolaevulinic
Laser and Non-laser Light Sources for Photodynamic Therapy 185
acid does not induce hair regrowth in patients with
extensive alopecia areata. Br J Dermatol 2000;
143:1032–5.
64. Stringer MR. Problems associated with the use of
broad-band illumination sources for photodynamic
therapy. Phys Med Biol 1995;40:1733–4.
65. Szeimies RM, Rudiger H, Baumler W, Heine A,
Landthaler M. A possible new incoherent lamp for
photodynamic treatment of superficial skin lesions.
Acta Derm Venereol 1994;74:117–19.
66. Masters BR, So PTC, Gratton E. Multiphoton exci-
tation fluorescence microscopy and spectroscopy of in
vivo human skin. Biophys J 1997;72:2405–12.
67. Konig K. Multiphoton microscopy in life sciences.
J Microsc Oxford 2000;200:83–104.
Paper received 18 June 2001;
accepted 6 December 2001.
186 L. Brancaleon and H. Moseley
... Cervical cancer remains a major global health problem and is the fourth most common cancer in women worldwide [1,2,3,4]. The central role of persistent human papillomavirus (HPV) in the development of cervical cancer is now well established [5,6]. ...
... The basic principle of PDT is the unique interaction of three key components: a photosensitizer, speci c wavelengths of light, and molecular oxygen. When these components come together, they trigger a cascade of events that culminate in the selective destruction of tumor cells while sparing the surrounding healthy tissue [1,12]. ...
... Light sources activate photosensitizers that emit wavelengths corresponding to the selected photosensitizer's absorption spectra. As a rule, lasers or light-emitting diodes (LEDs) are used [1]. The wavelength and light intensity are carefully selected according to the optimal absorption properties of the photosensitizer. ...
Article
Photodynamic therapy (PDT) has shown promise as a modality for the treatment of cervical cancer caused by the human papillomavirus (HPV). This review provides a comprehensive examination of the role of PDT in overcoming the challenges presented by conventional treatments for cervical cancer. Beginning with an overview of the relationship between cervical cancer and HPV infection, the review introduces the principles of PDT, its mechanism of action, and its potential as an innovative treatment strategy. The review highlights preclinical studies in animal models that demonstrate the efficacy of PDT in targeting HPV-infected cervical cells and provide mechanistic insights into its cytotoxic effects. We reviewed clinical studies and case reports highlighting the potential of PDT as an alternative or adjunctive treatment option. Challenges and limitations, including depth of light penetration, photosensitizer specificity, and standardization of protocols, will be discussed in the context of potential side effects and comparison with conventional treatments. Future directions include ongoing research, combination therapies with immunotherapy or targeted agents, advances in photosensitizer development, and personalized approaches. The advancement of PDT promises to change the landscape of HPV-associated cervical cancer treatment by providing a targeted, personalized, and minimally invasive approach.
... Another positive aspect is related to the use in large areas, without the need for coupling to fibers (Brancaleon;Moseley, 2002). ...
... Another positive aspect is related to the use in large areas, without the need for coupling to fibers (Brancaleon;Moseley, 2002). ...
Chapter
Full-text available
With the emergence of several strains of antibiotic-resistant bacteria, there has been a demand for new therapeutic alternatives to combat the growth of microbial pathogens. PACT (Photodynamic Antimicrobial Chemotherapy) is a therapeutic modality, inspired by the treatment of tumors, that aims to kill microorganisms by photodamage, by employing the combination of a photosensitizer and light irradiation, usually of the wavelength of visible light. Thus, PACT can be applied in different contexts, with Dentistry as one of the greatest applications, but PACT can be used in the treatment of skin infections and even disinfection of surfaces in hospital and industrial environments. This paper addresses historical and technical aspects of PACT, in terms of mechanism of action, sensitive microorganisms and future perspectives. This document is the result of the final work of the discipline "Special Topics in Pharmacology II", prepared by graduate students in Natural Products, Synthetics, Bioactives, Cellular and Molecular Biology.
... Light of particular bandwidth which comes in the absorption band of the photosensitizer is absorbed by th light sources used for pho tungsten-halogen lamps; gaseous discharge lamps like xenon lamps or metal halide lamps, lasers, light emitting diodes were use have some drawbacks like most narrow down their bandwidth. These filters may be long band filters [10]. Halogen lamps have heating problems so need extra filters to avoid heating effect. ...
Chapter
Agriculture and Food Science Book series aims to bring together leading academic scientists, researchers and research scholars to publish their experiences and research results on all aspects of Agriculture and Food Science. It also provides a premier interdisciplinary platform for researchers, practitioners and educators to present and discuss the most recent innovations, trends, and concerns as well as practical challenges encountered and solutions adopted in the fields of Agriculture and Food Science. High quality research contributions describing original and unpublished results of conceptual, constructive, empirical, experimental, or theoretical work in all areas of Agriculture and Food Science are cordially invited for publication.
... In recent years, many established studies have enumerated various merits of ideal photosensitizers [26], [28], [50]. An ideal PS needs to be hydrophilic in nature for easy absorption, non-toxic till exposed to light, and activated by an appropriate wavelength by a tunable laser light source [51]. More importantly, a good PS should generate a good photodynamic outcome based on its cellular localization and selectivity [52], [53]. ...
Article
Full-text available
In recent years, photodynamic therapy (PDT) has been accepted as an alternative option for the treatment of a wide spectrum of human ailments. It is a minimally invasive treatment that involves the interaction of a non-toxic photosensitizer. In PDT, combining photosensitizing (PS) agent that absorbs specified wavelength of light, which in turn produces free radical molecules to eliminate unwanted cells and tissues. The photosensitization process is activated by the light-induced excitation of molecules within the tissue. Bioactive principles acquired from plants documented as nature-inspired potential photosensitizers with varied properties against microbes, insects, or tumor cells. PDT is a promising method for removing diverse types of cancers but needs to be recognized in therapy as conventional chemotherapy. At present, natural compounds with PS properties are being continuously unearthed and identified. As of now, hundreds of photosensitive drugs or drug leads identified from natural sources with reduced or no toxicity to healthy tissues and no side effects encourage investigators to pursue natural PS for PDT. Although existing PS was developed years back, only a handful of them are engaged in human clinical applications. The main classes of natural photosensitizers discussed in this review are chlorophylls (hypocrellin A and B), hypericin, chlorins (Chlorin e6), and other emerging ones such as curcumin. Hence, the present review aimed to explore the efficacious PS properties of a few herbal-derived PS, preferably the potential ones in terms of specificity, and mechanism of action, inducing less or no toxicity to normal cells but their other medicinal applications.
... In general, the PDT effect achieved by excitation with narrow-spectrum light sources is stronger, all other conditions being equal in control. For isolated, small lesions the laser is more effective as a light source, for multiple, large lesions the LED light with a large spot is more suitable [84,85]. Lasers have been used in a wide variety of PDT applications [86]. ...
Article
Full-text available
Carbon dots (CDs) are novel carbon-based nanomaterials that have been used as photosensitizer-mediated photodynamic therapy (PDT) in recent years due to their good photosensitizing activity. Photosensitizers (PSs) are main components of PDT that can produce large amounts of reactive oxygen species (ROS) when stimulated by light source, which have the advantages of low drug resistance and high therapeutic efficiency. CDs can generate ROS efficiently under irradiation and therefore have been extensively studied in disease local phototherapy. In tumor therapy, CDs can be used as PSs or PS carriers to participate in PDT and play an extremely important role. In bacterial infectious diseases, CDs exhibit high bactericidal activity as CDs are effective in disrupting bacterial cell membranes leading to bacterial death upon photoactivation. We focus on recent advances in the therapy of cancer and bacteria with CDs, and also briefly summarize the mechanisms and requirements for PSs in PDT of cancer, bacteria and other diseases. We also discuss the role CDs play in combination therapy and the potential for future applications against other pathogens. Graphical Abstract
... When treating diseases other than cancer, the PDT light source's light intensity can reach up to 50 mW/cm 2 (40 J/cm 2 ), while it can reach up to 200 mW/cm 2 (150 J/cm 2 ) when treating cancer-related diseases. Due to interactions with skin chromophores such melanin and haemoglobin, red light or infrared light can penetrate the skin surface [22]. Some of the light sources that have positive impacts on PDT are Xenon arc lamps, metal halide lamps, lasers, and LEDs. ...
Article
Full-text available
Toothache is the prevalent dental disorder that affect men. Common activities including speaking, chewing, and smiling are negatively impacted by tooth problems.Particularly, serious issues are brought on by dental illnesses such periodontitis, dental caries, gingivostomatitis, gingivitis, etc.Bacteria are the main contributor to oral disorders that affect various parts of our mouths. Photodynamic therapy, a promising minimally invasive treatment option that has long been used to treat microvascular illnesses, cancer, and other conditions, has recently been concentrated on the treatment of dental issues and is seeing an exponential rise in global publications. Light with a specified wavelength, typically from a laser light source, is used in photodynamic treatment.When the photosensitizers are excited by a particular laser light, they combine with the molecular oxygen to produce singlet oxygen, which may be targeted to the diseased site and is site-specific and selective.Rapid bacterial eradication, minimum or no resistance development, safety of surrounding host tissue, and maintenance of normal microflora are all features of photodynamic therapy for dental problems.The review of photosensitizers, light sources, singlet oxygen, and upconversion nanoparticles is condensed in this work. The main dental conditions that impact the oral cavity have been covered, along with the various photodynamic therapy treatment modalities. Clinical trials on photosensitizer-loaded formulations for dental applications also has been discussed.
... An additional method for early diagnosis and treatment of malignant and potentially malignant lesions of the oral cavity involves the use of photosensitizing substances, irradiated by light through laser and nonlaser devices; the most commonly used device is the diode laser because it is small and inexpensive [138,139]. Siddiqui et al. recently proposed a portable and easy-to-use fiber-coupled LED system that uses a smartphone fluorescence imaging device [140]. Regarding photosensitizing agents, the use of Photofrin ® , administered intravenously, has been suggested for the treatment of oral cancer at any stage [141]. ...
Article
Full-text available
Oral Potentially Malignant Disorder (OPMD) is a significant concern for clinicians due to the risk of malignant transformation. Oral Squamous Cell Carcinoma (OSCC) is a common type of cancer with a low survival rate, causing over 200,000 new cases globally each year. Despite advancements in diagnosis and treatment, the five-year survival rate for OSCC patients remains under 50%. Early diagnosis can greatly improve the chances of survival. Therefore, understanding the development and transformation of OSCC and developing new diagnostic methods is crucial. The field of oral medicine has been advanced by technological and molecular innovations, leading to the integration of new medical technologies into dental practice. This study aims to outline the potential role of non-invasive imaging techniques and molecular signatures for the early detection of Oral Malignant and Potentially Malignant Disorders.
Chapter
Topical photodynamic therapy (PDT) is widely used as an effective and well‐tolerated treatment for field change actinic keratosis, Bowen disease and superficial basal cell carcinoma. There is a strong evidence base to support its use and guidelines, which summarise the evidence and clinical utility of this therapy. There are also standards available for guidance with respect to establishing PDT services in dermatology and emphasis is placed on the availability of PDT for practitioners affiliated with a skin cancer multi‐disciplinary team. Developments in photosensitisers and light delivery, such as through the increasing application of daylight PDT, have greatly improved the tolerance, acceptability and uptake of PDT, such that this should be widely available through dermatology services. Further refinements in photosensitiser and light delivery will be expected to continue to improve PDT outcomes and the uptake of this invaluable therapeutic approach.
Article
Full-text available
Simple Summary Photodynamic therapy (PDT) has emerged as a versatile treatment option for head and neck disorders. The purpose of our review was to provide an overview of the use of PDT therapy in a variety of diseases. The use of PDT therapy has recently emerged as an innovative and effective method of cancer treatment. However, more and more studies are showing that it is also promising for treating a variety of non-cancerous lesions, providing a minimally invasive alternative for certain conditions. The same following principles apply: the application of a photosensitizer, exposure to light, and then the destruction of abnormal cells. In our work, we highlighted the importance of photodynamic therapy and its valuable treatment as an option for a variety of diseases of the head and neck area, not just cancer. With this, we want to encourage researchers to more widely use photodynamic therapy in head and neck disorders. Abstract Photodynamic therapy (PDT) has emerged as a promising modality for the treatment of various diseases. This non-invasive approach utilizes photosensitizing agents and light to selectively target and destroy abnormal cells, providing a valuable alternative to traditional treatments. Research studies have explored the application of PDT in different areas of the head. Research is focusing on a growing number of new developments and treatments for cancer. One of these methods is PDT. Photodynamic therapy is now a revolutionary, progressive method of cancer therapy. A very important feature of PDT is that cells cannot become immune to singlet oxygen. With this therapy, patients can avoid lengthy and costly surgeries. PDT therapy is referred to as a safe and highly selective therapy. These studies collectively highlight the potential of PDT as a valuable therapeutic option in treating the head area. As research in this field progresses, PDT may become increasingly integrated into the clinical management of these conditions, offering a balance between effectiveness and minimal invasiveness.
Article
A 74-year-old woman with extensive inoperable vulval extramammary Paget's disease who had lymph node and pulmonary metastases, was treated with etoposide (VP16) 100 mg and 5000 cGy electron beam irradiation, which reduced the lesion by 60%. Following these interventions, the residual lesion was successfully treated with repeated photodynamic therapy using delta-aminolaevulinic acid, which achieved a nearly complete remission.
Article
BACKGROUND AND AIMS—Photodynamic therapy (PDT) is a treatment in which cell damage is achieved by the action of light on a photosensitising agent. We have assessed the potential use of PDT in the ablation of Barrett's oesophagus. METHODS—Thirty six patients with dysplastic Barrett's oesophagus receiving acid suppression medication with omeprazole were randomised to receive oral 5-aminolaevulinic acid (ALA) 30 mg/kg or placebo, followed four hours later by laser endoscopy. Follow up endoscopy was performed at one, six, 12, and 24 months. RESULTS—Of 18 patients in the ALA group, a response was seen in 16 (median decrease in area in the treated region 30%; range 0-60%). In the placebo group, a decrease in area of 10% was observed in two patients with no change in 16 (median 0%; range 0-10%; treatment v placebo, p<0.001). No dysplasia was seen in the columnar epithelium within the treatment area of any patient in the PDT group. However, in the placebo group, persistent low grade dysplasia was found in 12 patients (p<0.001). There were no short or long term major side effects. The effects of treatment were maintained for up to 24 months. CONCLUSIONS—This is the first randomised controlled trial of PDT for Barrett's oesophagus. It demonstrates that ALA induced PDT can provide safe and effective ablation of low grade dysplastic epithelium. Keywords: Barrett's oesophagus; photodynamic therapy; aminolaevulinic acid; protoporphyrin IX
Article
Photodynamic therapy (PDT), which utilizes a photoactivated drug (photosensitizer) to destroy malignant or certain other undesirable tissue, is approved by health agencies worldwide, mainly for the treatment of early- and late-stage lung cancer and premalignant and obstructive esophageal cancer. The clinical and scientific development of PDT began in the early 1970s at Roswell Park Cancer Institute in Buffalo, NY. This paper presents the history of this process at Roswell Park.
Article
Tumor destruction in photodynamic therapy is the result of the combination of direct cellular toxicity and damage to tumor microvasculature. These phenomena appear to be caused by tissue interactions with toxic oxygen compounds which are formed when light interacts with photosensitizing agents. Although injury to cell membranes, mitochondria and the nucleus have been noted, such injuries by themselves tend to be sublethal and cannot totally account for the effectiveness of PDT. The mechanism of effect of PDT on the vasculature has not been fully investigated. The vascular effects are believed to involve both intravascular and perivascular phenomena. Platelet aggregation appears to be an early event. Changes to the endothelium, and smooth muscle contraction as well as increased capillary permeability have also been observed during therapy. Initial experiments using'cyclooxygenase inhibitors indicate that arachidonic acid metabolites are active elements in producing the vascular phase of the therapeutic response and that these microvasculature effects appear to be critical to permanent tumor destruction.
Article
Cholesterol resembles other unsaturated lipids in being susceptible to peroxidative degradation when exposed to a sensitizing agent, exciting light of suitable wavelength and molecular oxygen. Selected hydroperoxides of cholesterol can be used as relatively convenient and reliable indicators of primary photochemical mechanisms, allowing a distinction to be made between free radical-mediated and singlet oxygen-mediated reactions. When generated in cell membranes, hydroperoxides of cholesterol and other lipids can have deleterious effects on membrane structure and function. Such damage may be exacerbated if these photoproducts undergo one-electron reduction to oxylradicals which in turn initiate chain peroxidation reactions. Cells can resist these effects by using a membrane-based glutathione peroxidase to catalyze the two-electron reduction and detoxification of lipid hydroperoxides. Recent advances in our understanding of cholesterol photo-oxidation from the standpoints of (a) mechanistic information, (b) cytotoxicity and (c) cytoprotection are discussed in this article.
Article
The incidence of basal cell carcinomas (BCCs) is still increasing, and there is a demand for an easy, effective and selective non-invasive treatment such as topical photodynamic therapy (PDT). Twenty-three patients with 24 nodular BCCs were treated once with delta-aminolaevulinic acid (δ-ALA) PDT (100 mW cm−2, 120 J/cm2) 3 weeks after prior debulking of the BCCs. Three months after PDT, all lesions were surgically excised and histopathologically evaluated for residual tumour. Twenty-two (92%) of the 24 nodular BCCs showed a complete response on clinical and histopathological examination. PDT for superficially abraded nodular BCCs with topically applied δ-ALA and the VersaLight® as light source is an easy, effective and safe therapy, with excellent cosmetic results and no serious side-effects, in cases where non-surgical treatment of BCCs is indicated.
Article
The tissue photosensitizer protoporphyrin IX (PpIX) is an immediate precursor of heme in the biosynthetic pathway for heme. In certain types of cells and tissues, the rate of synthesis of PpIX is determined by the rate of synthesis of 5-aminolevulinic acid (ALA), which in turn is regulated via a feedback control mechanism governed by the concentration of free heme. The presence of exogenous ALA bypasses the feedback control, and thus may induce the intracellular accumulation of photosensitizing concentrations of PpIX. However, this occurs only in certain types of cells and tissues. The resulting tissue-specific photosensitization provides a basis for using ALA-induced PpIX for photodynamic therapy. The topical application of ALA to certain malignant and non-malignant lesions of the skin can induce a clinically useful degree of lesion-specific photosensitization. Superficial basal cell carcinomas showed a complete response rate of approximately 79% following a single exposure to light. Recent preclinical studies in experimental animals and human volunteers indicate that ALA can induce a localized tissue-specific photosensitization if administered by intradermal injection. A generalized but still quite tissue-specific photosensitization may be induced if ALA is administered by either subcutaneous or intraperitoneal injection or by mouth. This opens the possibility of using ALA-induced PpIX to treat tumors that are too thick or that lie too deep to be accessible to either topical or locally injected ALA.