Open access peer-reviewed chapter

Reactivity and Applications of Singlet Oxygen Molecule

Written By

Celia María Curieses Andrés, José Manuel Pérez de la Lastra, Celia Andrés Juan, Francisco J. Plou and Eduardo Pérez-Lebeña

Submitted: 04 May 2023 Reviewed: 31 May 2023 Published: 22 October 2023

DOI: 10.5772/intechopen.112024

From the Edited Volume

Reactive Oxygen Species - Advances and Developments

Edited by Rizwan Ahmad

Chapter metrics overview

98 Chapter Downloads

View Full Metrics

Abstract

Reactive oxygen species (ROS) are molecules produced in living organisms, in the environment, and in various chemical reactions. The main species include, among others, singlet oxygen (1O2), the superoxide anion radical (•O2−), the hydroxyl radical (HO•), and the hydroperoxyl radical (HOO•). In general, the reactivity of 1O2 is lower than that of HO• but even higher than that of •O2−. Singlet oxygen is the lowest energy excited state of molecular oxygen, but it is also a highly reactive species, which can initiate oxidation reactions of biomolecules such as amino acids, proteins, nucleic acids, and lipids, either by a direct reaction or by the induction of ROS. Singlet oxygen is a highly reactive electrophilic species that reacts with electron-rich molecules and is related to several types of pathologies. To inhibit the oxidation of biomolecules with this species, some substances act as antioxidants by performing a quenching effect. In this chapter, aspects such as its physicochemical properties, methods of generation and detection, as well as the reactivity of this molecule are detailed.

Keywords

  • singlet oxygen
  • photochemistry
  • phototherapy
  • photosensitizers
  • singlet oxygen generation methods

1. Introduction: Singlet oxygen (1O21Δg or 1O2*)

Oxygen is the most abundant element in the earth’s crust, mainly in its gaseous biatomic molecule form, constituting 21% by volume of dry air. It is one of the most biologically important elements because of the type and number of reactions in which it participates, providing the thermodynamic force necessary for the metabolism of all higher organisms. Molecular oxygen in its basal state has an open-shell electronic configuration, with two electrons of equal spin occupying different degenerate molecular orbital. It is only capable of accepting one electron at a time during a redox (radical-type) reaction, reacting slowly with most organic molecules. It has two excited states of singlet multiplicity. The lower energy state is designated as O2(1Δg) or 1O2, and the electron distribution in 1Δg has antiparallel spins, the two electrons occupy the same orbital with opposite spins, therefore the spin restriction does not exist and it is able to accept two electrons at a time, thus increasing its oxidative capacity compared to 3O2. The next excited singlet level is represented by the symbol O2(1Σg+), Figure 1.

Figure 1.

Electronic configuration molecular oxygen and the first two states mentioned.

Because of the high energy, low stability, and shorter lifetimes of the O2 state (1Σg+), the term “singlet oxygen” commonly refers to the 1O2 state. Singlet oxygen, or the first excited state of oxygen, is a highly oxidizing species that is generated by a light-activated compound, called a photosensitizer, and appears to play a significant role in solution reactions. It is found at 22.5 Kcal/mol above the basal state. This species emits phosphorescent light in the infrared region (1270 nm). The higher energy state (1Sg+), on the other hand, is rapidly deactivated at 1Dg and therefore will not be as reactive. The two energetically closest electronically excited states are singlet states, whose spectroscopic notations are 1Δg and 1Δg+. Singlet oxygen (1O2) is an electronically excited species of molecular oxygen (O2) and participates in numerous oxidation reactions as an activated species and plays a key role in many biological processes, Figure 2.

Figure 2.

Energy diagram for O2. The singlet states of oxygen are 156.9 and 94.2 kJ/mol higher in energy than its ground triplet state.

The transition between the triplet ground state (3Σg) and the first excited singlet (1Δg) is spin, symmetry, and parity forbidden; therefore, direct excitation of the ground state to form singlet oxygen is very unlikely, and gas-phase singlet oxygen has an extremely long lifetime (72 min) [1]; however, interaction with solvents reduces the lifetime, depending on the solvent and varies from 10−3 to 10−6 s [2], with the shortest lifetime observed in water [1].

From the values in Table 1, it can be concluded that some of the solvent characteristics that influence the lifetime are the number of C-H and O-H bonds and the presence of halogen and deuterium atoms [2].

Solventτ(μs)Solventτ(μs)
H2O3.1CH3CN77.1
CH3OH9.5CH2Cl299
C6H1423.4D2O68
C6H630.0C6D6681
(CH3)2CO51.2(CD3)2CO992

Table 1.

Lifetime of 1O2 in different solvents.

Advertisement

2. Singlet oxygen chemistry

Because 1O2 is a molecule in an electronically excited state, it is very unstable with respect to its ground state, and once generated it can undergo various spontaneous processes to deactivate it. Such deactivation can occur by different routes, Figure 3.

Figure 3.

Singlet oxygen deactivation pathways. The molecule that interacts with O2(1Δg) is commonly called quencher (Q).

The non-radiative deactivation process of 1O2 (kd) involves the transfer of electronic energy to the vibrational levels of the solvent. The radioactive deactivation process of 1O2 (kdr) occurs when 1O2 decays to its ground state emitting phosphorescent radiation of 1270 nm. The value of this rate constant depends on the medium.

The deactivation processes by molecules in the medium can be either physical (kq) or chemical (kr). Physical deactivation of 1O2 occurs mainly by energy transfer mechanisms. Molecules with several conjugated double bonds, such as long-chain polyenes, quinones, dyes, and transition metal complexes, are some examples where this type of deactivation occurs [3].

The reactive deactivation of singlet oxygen, as the name suggests, refers to its chemical reaction with a wide variety of organic molecules. Singlet oxygen is a highly reactive electrophilic species and has the ability to rapidly attack organic compounds. It is about 1000 times more reactive than the basal state of oxygen. This high reactivity is simply because many of the substances with which it reacts are in the singlet basal state, so the reaction is a singlet-singlet reaction, more likely than a triplet-singlet reaction, as it should be with oxygen in its basal state. As singlet molecular oxygen possesses a pair of electrons with opposite spins in the highest occupied molecular orbital, they give O2(1Δg) dienophilic properties, which explains its significant reactivity towards electron-rich organic molecules, particularly those with conjugated double bonds [4]. These reactions are electrophilic [4π + 2π], [2π + 2π], and ene-type additions, leading to the formation of allylic hydroperoxides, dioxetanes, or endoperoxides [5, 6, 7, 8].

Singlet oxygen also oxidizes sulfur, selenium, phosphorus, and nitrogen compounds, Figure 4. Singlet oxygen-mediated oxidation reactions have been explored with a variety of organic molecules containing heteroatoms.

Figure 4.

Singlet oxygen-mediated photooxidations: (a) cycloaddition [2 + 2], (b) [4 + 2] cyclo addition, (c) 1,3-ene addition, and hetero atom oxidation. In all these reactions the primary products formed can undergo rearrangements to give a wide range of stable oxidized products.

Advertisement

3. Singlet oxygen production

There are three main sources for generating singlet oxygen: photochemical, chemical, and in vivo reactions, Figure 5.

Figure 5.

Sources of singlet oxygen.

3.1 Chemical generation

There are many chemical reactions, one of the best-known examples being the reaction of NaClO with H2O2. The decomposition reaction of H2O2 in the presence of sodium hypochlorite or sodium hypobromite is the oldest reaction [9] for the formation of singlet oxygen, Figure 6.

Figure 6.

Singlet oxygen generation from H2O2 by NaOCl.

1O2 is formed from the chlorine peroxide anion (ClOO), which is an intermediate in the above reaction. The best results are achieved when H2O2 decomposition is performed using a MoO42− catalyst [10, 11, 12, 13], Figure 7.

Figure 7.

Singlet oxygen generation from H2O2 by the [MoO4]2− catalyst.

Among the inorganic peroxides, calcium peroxide diperoxohydrate (CaO2-2H2O2), which is easily prepared by treating CaCl2 or Ca(OH)2 with H2O2 is of special interest for its ability to release 1O2. In the first step, Ca(OH)2 is converted into CaO2-8H2O, which takes up two H2O2 molecules giving CaO2-2H2O2. By thermolysis at moderate temperature, the latter peroxide decomposes into 1O2 and CaO2-8H2O. This catalytic cycle would operate until complete disproportionation of H2O2 and leave CaO2-8H2O as the final product [14], Figure 8.

Figure 8.

Singlet oxygen generation from CaO2-2H2O2.

Singlet oxygen can also be generated chemically by the decomposition of ozonides of triphenylphosphite [15], Figure 9.

Figure 9.

Singlet oxygen generation from ozonides of tri-phenyl-phosphites.

1O2 production from acyl peroxides such as benzoyl peroxide and lauryl peroxide [16, 17], Figure 10.

Figure 10.

Singlet oxygen generation from benzoyl peroxide.

By decomposition of aromatic endoperoxides, singlet oxygen is also generated chemically. Certain dienes (anthracenes, naphthalenes, and pyridones) react with O2(3Σg-) via sensitized photooxygenation to produce endoperoxides, which by heating dissociate and regenerate the starting hydrocarbon and 1O2. The thermal stability of endoperoxides depends on the structure of the aromatic or heteroaromatic hydrocarbon. These organic endoperoxides are called reversible molecular singlet oxygen transporters. The lifetime of the endoperoxide precursors depends on the temperature and chemical structure of the dienes. Derivatives of naphthalenes, anthracenes, and pyridones are among the most studied examples that can undergo reversible endoperoxide formation for this purpose, Figure 11. All carriers are based on the reversible [4 + 2] cycloaddition of 1O2 to a diene and are therefore generated by photooxygenation.

Figure 11.

Generation of singlet oxygen based on the formation and decomposition of endoperoxides.

3.2 Photochemical generation

The absorption of light by a photosensitizer in its ground state (1Sens) transforms it into an excited singlet state (1Sens*). This short-lived state is rapidly converted through intersystem crossover into a more stable and longer-lived species, the excited triplet state (3Sens*). The photosensitizer in its triple-excited state has the ability to transfer its energy to the 3O2 dissolved in the medium. As a consequence of this transfer, the sensitizer is regenerated in its basal state, and the O2 remains in its singlet excited state, Figure 12.

Figure 12.

The most widely used method in the generation of O2(1Δg) is photosensitization.

Photosensitization represents the most convenient, safe, easy, and environmentally friendly method for 1O2 generation. The generation of singlet oxygen through photosensitization has been widely exploited in photodynamic therapy, environmental remediation, and synthesis [18].

3.3 Possible production of singlet oxygen in vivo

Some of the possible biological sources of 1O2 generation are shown in Figure 13.

Figure 13.

Generation of singlet oxygen in biological systems.

Possible biological sources of O2(1Δg) include (i) reactions catalyzed by peroxidases (myeloperoxidase) or oxygenases (lipoxygenases) [19, 20], (ii) recombination of peroxyl radicals can lead to the release of 1O2 as a result of the decomposition of transient tetroxides according to Russell’s concerted mechanism (Russell reaction) [6, 21], (iii) oxidation with ozone of amino acids, peptides, and proteins, (iv) hydrogen peroxide reactions with hypochlorite or peroxynitrite, (v) ozone oxidation of amino acids, peptides, and proteins [22], (vi) hydrogen peroxide reactions with hypochlorite or peroxynitrite [23, 24], (vii) thermolysis of endoperoxides [25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35], (vii) in vitro photodynamic processes involving type-II photosensitization reactions using suitable dyes [36, 37, 38], and (viii) UV irradiation of aromatic amino acids in proteins and immunoglobulins.

Advertisement

4. Applications

1O2 has been gaining much attention due to its pivotal role in a wide variety of chemical and biological processes, for example, plant signaling, organic synthesis, oxidation of food and beverages, or photodynamic therapy [39, 40], Figure 14.

Figure 14.

The following sections outline different applications that have been developed over the last decades.

4.1 Photodynamic therapy (PDT)

Photodynamic therapy (PDT) has received increasing attention for the treatment of cancer and other diseased tissues as this methodology allows noninvasive, selective, and localized destruction of tumor cells with reduced side effects. PDT consists of the administration of a photosensitizer, which is selectively accumulated in certain cells or tissues so that with subsequent irradiation and in the presence of oxygen, it triggers photooxidation of biological materials and subsequent cell death.

PDT is composed of three components, Figure 15. Each of these components alone lacks curative properties. The contribution of each depends on the type and dose of PS, the time between administration and exposure to light, the intensity of light, and the concentration of oxygen generated.

Figure 15.

PDT components.

PDT combines light, molecular oxygen, and a photosensitizer (PS) for the production of reactive oxygen species (ROS) such as singlet oxygen (1O2) and free radicals that induce oxidative stress and eventually cell death. PDT was the first example of a drug-device combination approved by the Food and Drug Administration (FDA) for the production of reactive oxygen species (ROS) such as singlet oxygen (1O2) and free radicals that induce oxidative stress and eventually cell death [41]. PDT was the first example of a drug-device combination approved by the Food and Drug Administration (FDA) [42].

4.2 Light sources

Most Ps are activated with red light between 630 and 700 nm, which has a tissue penetrating power between 0.5 cm (at 630 nm) and 1.5 cm (at 700 nm) [43]. Lasers and light-emitting diodes (LEDs) coupled to flexible fiber optic devices are used as light sources. Depending on the depth of the pathologies to be treated, LEDs can be implanted at the end of different catheters to form screens for irradiation of large areas. The wavelength required will be determined by the PS with which it is to be associated and applied in the most focal way possible on the tissue to be treated. However, the most commonly used light sources in photodynamic treatments are lasers.

4.3 Photosensitizing agent in photodynamic therapy

A PS is defined as a compound capable of absorbing light and subsequently triggering a photophysical or photochemical reaction in response to it. For clinical use, it must meet a number of characteristics as summarized in Figure 16 [39, 44].

Figure 16.

PS properties.

The PS most suitable for PDT can be divided into two main groups: porphyrinoid or non-porphyrinoid derivatives [44, 45, 46, 47], Figure 17. Porphyrinoid-based derivatives are the most widely used in PDT applications as their extended π-systems confer unique photochemical characteristics that are valuable for their use as photosensitizers [48]. Within the porphyrinoid PS, they are usually classified into first-, second-, and third-generation photosensitizers depending on their evolution [49]. Natural porphyrins and their derivatives constitute the first generation of PS. Most of the PS under investigation for the treatment of cancer and other diseases is based on the tetrapyrrole core that includes porphyrins, chlorins, bacterio-chlorins, phthalocyanines, and texaphyrins. These molecules have been chosen for their low toxicity in the absence of light to mammalian and animal cells and for their tumor-localizing properties. PS that have been studied for their ability to kill microorganisms are halogenated xanthenes such as rose bengal (RB), chlorinated poly-L-lysine-chlorinacorinate, and phenothiazines such as toluidine blue O (TBO) and methylene blue [50] and poly-L-lysine-chlorinate conjugates [51].

Figure 17.

Types of photosensitizers.

4.4 First-generation photosensitizers

First-generation PSs are a complex mixture of compounds among which hematoporphyrin (HpD) and photofrin (R) derivatives are the representatives of the first-generation Fs. Photofrin is one of the few photosensitizers approved for the treatment of early and advanced esophageal and lung cancer by the US FDA and by various health agencies worldwide. It is now being extended to the therapy of different oncological pathologies of the head, neck, abdomen, thorax, brain, bowel, cervical, skin, and breast Photofrin® (as HPD) causes persistent photosensitization of the skin. This required avoidance of intense sunlight for at least 30 days after drug administration. HpDs are a complex mixture of monomers, dimers, and oligomers, linked together by the formation of ethers, esters, and carbon-carbon bonds [44], Figure 18.

Figure 18.

First-generation porphyrin derivatives.

4.5 Second-generation photosensitizers

Second-generation photosensitizers have been developed since the late 1980s. They are pure compounds of known chemical structure with absorption maxima at wavelengths above 630 nm, high molar extinction coefficients, quantitative 1O2 formation, and a reduction in the side effects and undesirable properties of the first generation. The major disadvantage of these is that they are highly lipophilic, which hinders their bioavailability, favors self-aggregation, and poses a challenge for the development of a pharmaceutical form for intravenous administration.

4.5.1 Porphyrin photosensitizers

Porphyrin derivatives such as meta-tetra-(hydroxyphenyl)-porphyrin (m-THPP, Foscan, or Temoporphyrin) and 5,10,15,20-tetrakis(4-sulfonatophenyl)-21H-23H-porphyrin (TPPS4). These compounds are pure, extremely potent structures, which are photoactivated at wavelengths longer than those corresponding to HpD, with higher molar extinction coefficients. This group also includes macrocyclic porphyrin derivatives such as chlorines, bacterio-chlorines, phthalocyanines, porphyrins, and texaphyrins. Due to the potency of the second-generation Fs, the drug doses and light intensity required to obtain a photofrin(R)-like response are up to 100 times lower, Figure 19.

Figure 19.

(a) Second-generation porphyrin derivatives and (b) Second-generation porphyrin macrocycles.

4.5.2 Non-porphyrin photosensitizers

The development of non-porphyrin PSs, Figure 20, for application in photodynamic therapy lags considerably behind the evolution of porphyrin derivatives. Cationic compounds, such as phenothiazines, have high molar extinction coefficients between 600 and 800 nm, including methylene blue and toluidine blue. Methylene blue has been applied in the clinical treatment of basal cell carcinoma and Kaposi’s sarcoma. In addition, there are in vitro assays of adenocarcinoma, bladder carcinoma, and cervical tumor cells. Toluidine blue has been extensively studied for the inactivation of various pathogenic microorganisms. Both thiazine derivatives are currently being evaluated for the treatment of chronic periodontitis. Literature has shown that the mitochondrial membrane potential of tumor cells is more negative than that of normal cells so that cationic PSs are retained mainly in the cells to be treated. The bacterial cell wall has anionic components that readily interact with cationic drugs.

Figure 20.

Chemical structure of non-porphyrin photosensitizers.

Rose bengal has shown good results in the photodynamic treatment of breast carcinoma and metastatic melanoma. 4,5-dibromorhodamine methyl ester is effective against graft-versus-host disease, resulting in the destruction of lymphocytes by apoptosis.

Hypericin is a natural anthraquinone with defined photochemical properties and shows selectivity for tumor cells. Merocyanin 540 is the representative of the cyanins used in the treatment of leukemia and neuroblastoma.

4.5.3 Endogenous photosensitizer: precursors

Protoporphyrin IX (PpIX) is an endogenous photosensitizer. Under normal conditions, PpIX is present at an extremely low concentration to produce a photosensitizing reaction. To stimulate the synthesis of protoporphyrin IX, it is necessary to administer an excess of amino levulinic acid (ALA) and its derivatives, ALA-methyl ester (MAL) and ALA hexyl ester (HAL), thus stimulating the synthesis of PpIX, which can accumulate in damaged tissue and be applied in the photodynamic treatment of different pathologies. It has been used in the treatment of actinic keratosis, carcinoma, basal cell carcinoma, Bowen’s disease, and bladder cancer. PDT with ALA and its derivatives, ALA-methyl ester (MAL) and ALAhexyl ester (HAL), is an approved treatment for a number of malignant and premalignant conditions (Figure 21).

Figure 21.

Chemical structure of Protoporphyrin IX and of ALA MAL and HAL.

4.6 Third-generation photosensitizers

Third-generation photosensitizers are called nano photosensitizers and are made up of a combination of PSs and different vector systems. They are proposed as a solution to overcome the problems of low water solubility of the first- and second-generation Ps as most of them have an aromatic character.

Numerous delivery strategies have been evaluated with the aim of achieving a highly selective and effective therapy. Among the drug carrier systems used in the development of third-generation PSs, applicable in PDT are polymeric nanoparticles, solid lipid nanoparticles, nanostructured lipid carriers, gold nanoparticles, hydrogels, liposomes, liquid crystals, dendrimers, and cyclodextrin.

4.6.1 Liposomes

One of the most recent and successful applications of liposomes is to serve as delivery vehicles for photodynamic agents to improve their solubility, specificity, bioavailability, and tendency to aggregate. Therefore, liposomal photosensitizers as third-generation formulations have shown great potential to increase the efficacy of photodynamic cancer therapy and consist of concentric vesicles formed by one or more concentric natural or synthetic phospholipid bilayers enclosing an aqueous compartment. Nontoxic, highly biocompatible, and biodegradable, these delivery systems have allowed increasing the activity of Ps in vitro and in vivo, reducing their toxicity and improving their biodistribution [52].

4.6.2 Micelles

Micelles are monolayer structures formed by amphiphilic surfactants, also called amphipathic surfactants, which are molecules that have one hydrophilic end, that is they are soluble in water, and one hydrophobic end, which means that they repel water.

Polymeric micelles have been widely used in recent years due to their higher stability compared to micellar systems based on conventional surfactants. They are prepared by association of copolymers dispersed in an aqueous medium, forming particles with diameters of less than 100 nm [53]. Polymeric micelles are presented as an alternative for the delivery of hydrophobic Ps and have the following advantages (i) simple preparation, (ii) efficient drug loading without chemical modification, (iii) controlled release [54], and (iv) no side effects of skin photosensitivity [55].

4.6.3 Solid nanoparticles

Solid nanoparticles are a promising new tool for drug delivery in PDT [56, 57]. The advantages over second-generation Ps can be summarized as (i) ability to deliver a large amount of drug to target cells, (ii) prevent degradation in the biological environment, (iii) incorporate multiple components as contrast agents, and (iv) incorporate of different ligands in order to increase selectivity [58].

The solid nanoparticles, most commonly used in the development of third-generation Ps, are polymeric, silica, and gold nanoparticles [57, 59]. Silica nanoparticles have numerous advantages over polymeric systems such as their nontoxic nature, inert, and stable properties. As they are nonbiodegradable, particle size control is necessary to ensure their elimination from the body via the kidney and to avoid their uptake by the reticuloendothelial system. Due to their high porosity, silica nanoparticles have a high PS loading capacity and contribute significantly to the efficacy of PDT.

Gold nanoparticles increase the field of incident light around them, which could increase the excitation efficiency of the PS they carry [60]. The vehiculization of porphyrins, phthalocyanines, and thiazines on gold nanoparticles increased (i) blood circulation time, (ii) selective delivery to tumor tissues, (iii) 1O2 generation yield, and (iv) photodynamic activity of free PSs [61].

Advertisement

5. Mechanism of action of photosensitizers

Most PS in their ground state has two electrons with opposite spins. Light absorption leads to the transfer of an electron to a higher energy orbital. This excited PS is very unstable and emits this excess energy in the form of fluorescence and/or heat. Another possibility is that an excited PS can undergo cross-system crossover to form a triplet state. The photosensitizer in the triplet state can decay without radiation to the ground state or transfer its energy to molecular oxygen. This step leads to the formation of singlet oxygen, and the reaction is called a type-II process [62]. A type-I process can also occur [63], Figure 22.

Figure 22.

Dynamic singlet oxygen therapy: Sensitizer diagram.

In the type-I process, electron transfer occurs between the activated PS and the surrounding molecules leading to the release of free radicals. These radicals are highly active and interact with oxygen-producing endogenous molecules to produce anion superoxide, hydroxyl radical, hydrogen peroxide, and free radicals. These ROS cause damage to the integrity of cell membranes and their internal structures.

The type-II process, this is the interaction of PS with oxygen. For this reaction to take place, PS must be in its triplet form, activating oxygen to its active or singlet form (1O2), which allows it to interact with a large number of substances directly such as amino acids or lipids. The short half-life of this singlet oxygen (<0.4 milliseconds) means that its diffusion range is limited to 45 nm in the cellular medium, and destruction is only limited to the intracellular structures, it can access [64, 65].

Advertisement

6. Clinical applications of the PDT

PDT is a minimally invasive outpatient therapeutic modality that has low toxicity and can be applied repeatedly at the site of action. In oncology, it can be combined with chemotherapy, ionizing radiation, or surgery. PDT is an approved technique for the treatment of some types of cancer and for antimicrobial therapies.

6.1 Photodynamic therapy of cancer

Photosensitizers (PS) can be used in various types of tumors. PS has been used as therapeutic agent for more than a century. The clinical use of eosin in the treatment of skin cancer can be given as an example of the first applications of PDT. The first objective was to treat skin tumors by using topical eosin [66, 67].

According to the World Health Organization (WHO), cancer is a group of cellular diseases characterized by unregulated cell growth. In the early stages, it is localized in healthy tissues and spreads to neighboring tissues or even to other organs of the body through metastatic cells. The idea of using photodynamic therapy (PDT) as a new treatment strategy was suggested in the early twentieth century. PDT has fewer side effects and toxicity than chemotherapy and/or radiotherapy. PDT is a particularly attractive alternative to conventional antitumor drugs due to its fundamental specificity and selectivity. This excitation causes the photosensitizer to generate singlet oxygen and other reactive oxygen species. PDT has been used in several types of cancer, including non-melanoma skin cancer, bladder cancer, esophageal cancer, head and neck cancer, and non-small cell lung cancer (NSCLC), Figure 23.

Figure 23.

PDT antitumor treatments.

The PDT clinical procedure involves four stages, as shown in Figure 24.

  1. Ps is administered to the patient, usually topically or intravenously.

  2. Between 3- and 96-hours elapse before the affected area is illuminated. During this time, the drug is distributed throughout the body and selectively localized in the affected cells.

  3. The affected area is irradiated locally using an appropriate irradiation source for photodynamic treatment.

  4. A series of intracellular reactions are triggered, which give rise to the formation of ROS that produce irreversible biological damage in the treated area, leading to cell death. Cell destruction can be triggered by different pathways [44, 68].

Figure 24.

Representation of an antitumor treatment with PDT.

6.2 Non-oncological applications of PDT

PDT has been mainly focused on cancer treatment, but in recent years, new applications in different fields are being evaluated, Figure 25.

Figure 25.

Non-oncological applications of PDT.

Esthetics: Photo depilation is based on using a laser that irradiates at an appropriate wavelength (depending on the color of the hair and the color of the skin) to permanently remove the hair. The action of topically applied ALA is currently being studied to improve the photo-depilation technique.

Dermatology: PDT is used to treat non-oncological dermatological diseases such as psoriasis, vascular malformations, and acne [69, 70, 71].

Inactivation of bacteria: The increasing resistance of bacteria to antibiotics has led to the development of alternative antimicrobial techniques. Certain bacteria can be inactivated by being illuminated after an incubation period with certain porphyrins and phthalocyanines. This new application may be of great use in the treatment of internal cavity cleaning and in the treatment of oral conditions.

Virus inactivation: Photoinactivation of viruses in human blood as a method to sterilize blood and blood products for transfusion [72].

Ophthalmology: Age-related macular degeneration (AMD) is a leading cause of blindness and is due to the rapid abnormal growth of blood vessels in the retina. The results obtained by treating this disease with PDT show a hopeful future.

Arteriosclerosis: The possibility of treating this disease with PDT is based on the fact that atheroma plaques in damaged arteries retain higher concentrations of porphyrins than the normal vascular wall [69].

Gynecology: PDT is an alternative to hysterectomy for women with dysfunctional uterine bleeding [73, 74].

Advertisement

Funding information

This research was funded by Agencia Canaria de Investigación, Innovación y Sociedad de la Información (ACIISI) del Gobierno de Canarias, Project ProID2020010134, Caja Canarias, Project 2019SP43, the Spanish Ministry of Economy and Competitiveness (Grant PID2019-105838RB-C31) and the State Plan for Scientific, Technical Research and Innovation 2021–2023 from the Spanish Ministry of Science and Innovation (project PLEC2022-009507) and the State Plan for Scientific, Technical Research and Innovation 2021–2023 from the Spanish Ministry of Science and Innovation (project PLEC2022-009507).

References

  1. 1. Wilkinson F, Helman WP, Ross AB. Rate constants for the decay and reactions of the lowest electronically excited singlet state of molecular oxygen in solution. An expanded and revised compilation. Journal of Physical and Chemical Reference Data. 1995;24:663-677. DOI: 10.1063/1.555965
  2. 2. Schweitzer C, Schmidt R. Physical mechanisms of generation and deactivation of singlet oxygen. Chemical Reviews. 2003;103:1685-1758. DOI: 10.1021/cr010371d
  3. 3. Volman DH, Hammond GS, Gollnick K. Advances in Photochemistry. Hoboken, NJ, USA: John Wiley & Sons Inc.; 2009
  4. 4. Frimer AA. The reaction of singlet oxygen with olefins: the question of mechanism. Chemical Reviews. 1979;79:359-387
  5. 5. Cadet J, Di Mascio P. Peroxides in biological systems. Patai’s Chemistry of Functional Groups. 2009. DOI: 10.1002/9780470682531.pat0357
  6. 6. Miyamoto S, Martinez GR, Medeiros MH, Di Mascio P. Singlet molecular oxygen generated from lipid hydroperoxides by the Russell mechanism: studies using 18O-labeled linoleic acid hydroperoxide and monomol light emission measurements. Journal of the American Chemical Society. 2003;125:6172-6179
  7. 7. Miyamoto S, Nantes IL, Faria PA, Cunha D, Ronsein GE, Medeiros MH, et al. Cytochrome c-promoted cardiolipin oxidation generates singlet molecular oxygen. Photochemical & Photobiological Sciences. 2012;11:1536-1546
  8. 8. Miyamoto S, Martinez GR, Rettori D, Augusto O, Medeiros MH, Di Mascio P. Linoleic acid hydroperoxide reacts with hypochlorous acid, generating peroxyl radical intermediates and singlet molecular oxygen. Proceedings of the National Academy of Sciences. 2006;103:293-298
  9. 9. Murray RW. Chemical Sources of Singlet Oxygen. 1979
  10. 10. Aubry J. Search for singlet oxygen in the decomposition of hydrogen peroxide by mineral compounds in aqueous solutions. Journal of the American Chemical Society. 1985;107:5844-5849
  11. 11. Aubry J, Cazin B. Chemical sources of singlet oxygen. 2. Quantitative generation of singlet oxygen from hydrogen peroxide disproportionation catalyzed by molybdate ions. Inorganic Chemistry. 1988;27:2013-2014
  12. 12. Niu Q, Foote C. Singlet molecular oxygen generation from the decomposition of sodium peroxotungstate and sodium peroxomolybdate. Inorganic Chemistry. 1992;31:3472-3476
  13. 13. Nardello V, Marko J, Vermeersch G, Aubry J. 90Mo NMR and kinetic studies of peroxomolybdic intermediates involved in the catalytic disproportionation of hydrogen peroxide by molybdate ions. Inorganic Chemistry. 1995;34:4950-4957
  14. 14. Nardello V, Aubry J-M, Briviba K, Sies H. Identification of the precursor of singlet oxygen (1 O 2, 1 Δ g) involved in the disproportionation of hydrogen peroxide catalyzed by calcium hydroxide. Chemical Communications. 1998:599-600. DOI: 10.1039/A708161H
  15. 15. CaMinade AM, Khatib FE, Koenig M, Aubry JM. Ozonides de phosphite source d'oxygène singulet: rendement, mécanisme. Canadian Journal of Chemistry. 1985;63:3203-3209
  16. 16. Danen WC, Arudi RL. Generation of singlet oxygen in the reaction of superoxide anion radical with diacyl peroxides. Journal of the American Chemical Society. 1978;100:3944-3945
  17. 17. Khan AU, Kasha M. Singlet molecular oxygen in the Haber-Weiss reaction. Proceedings of the National Academy of Sciences. 1994;91:12365-12367
  18. 18. DeRosa MC, Crutchley RJ. Photosensitized singlet oxygen and its applications. Coordination Chemistry Reviews. 2002;233:351-371
  19. 19. Kanofsky J, Axelrod B. Singlet oxygen production by soybean lipoxygenase isozymes. Journal of Biological Chemistry. 1986;261:1099-1104
  20. 20. Klebanoff SJ. Myeloperoxidase: friend and foe. Journal of Leukocyte Biology. 2005;77:598-625
  21. 21. Russell GA. Deuterium-isotope effects in the autoxidation of aralkyl hydrocarbons. mechanism of the interaction of peroxy radicals1. Journal of the American Chemical Society. 1957;79:3871-3877
  22. 22. Kanofsky J, Sima P. Singlet oxygen production from the reactions of ozone with biological molecules. Journal of Biological Chemistry. 1991;266:9039-9042
  23. 23. Di Mascio P, Bechara EJ, Medeiros MH, Briviba K, Sies H. Singlet molecular oxygen production in the reaction of peroxynitrite with hydrogen peroxide. FEBS Letters. 1994;355:287-289
  24. 24. Martinez GR, Di Mascio P, Bonini MG, Augusto O, Briviba K, Sies H, et al. Peroxynitrite does not decompose to singlet oxygen (1ΔgO2) and nitroxyl (NO−). Proceedings of the National Academy of Sciences. 2000;97:10307-10312
  25. 25. Moureu C, Dufraisse C, Dean P. An organic peroxide separable from the rubrene peroxide. Comptes Rendus Hebdomadaires des Séances de l'Académie des Sciences. 1926;182:1584-1587
  26. 26. Dufraisse C, Velluz L. The labile union of carbon and peroxide which spontaneously dissociates in the cold. Comptes Rendus Hebdomadaires des Séances de l'Académie des Sciences. 1939;208:1822-1824
  27. 27. Mano CM, Prado FM, Massari J, Ronsein GE, Martinez GR, Miyamoto S, et al. Excited singlet molecular O2 (1Δg) is generated enzymatically from excited carbonyls in the dark. Scientific Reports. 2014;4:5938
  28. 28. Wasserman HH, Larsen DL. Formation of 1, 4-endoperoxides from the dye-sensitized photo-oxygenation of alkyl-naphthalenes. Journal of the Chemical Society, Chemical Communications. 1972:253-254. DOI: 10.1039/C39720000253
  29. 29. Di Mascio P, Sies H. Quantification of singlet oxygen generated by thermolysis of 3, 3′-(1, 4-naphthylene) dipropionate endoperoxide. Monomol and dimol photoemission and the effects of 1, 4-diazabicyclo [2.2.2] octane. Journal of the American Chemical Society. 1989;111:2909-2914
  30. 30. Pierlot C, Aubry J-M, Briviba K, Sies H, Di Mascio P. [1] Naphthalene endoperoxides as generators of singlet oxygen in biological media. Methods in Enzymology. 2000;319:3-20
  31. 31. Saito I, Matsuura T, Inoue K. Formation of superoxide ion from singlet oxygen. Use of a water-soluble singlet oxygen source. Journal of the American Chemical Society. 1981;103:188-190
  32. 32. Ravanat J-L, Di Mascio P, Martinez GR, Medeiros MH, Cadet J. Singlet oxygen induces oxidation of cellular DNA. Journal of Biological Chemistry. 2000;275:40601-40604
  33. 33. Ravanat J-L, Sauvaigo S, Caillat S, Martinez GR, Medeiros MHG, Mascio PD, Favier A, Cadet J. Singlet oxygen-mediated damage to cellular DNA determined by the comet assay associated with DNA repair enzymes. 2004;385(1):17-20. DOI: 10.1515/BC.2004.003
  34. 34. Martinez GR, Ravanat J-L, Medeiros MH, Cadet J, Di Mascio P. Synthesis of a naphthalene endoperoxide as a source of 18O-labeled singlet oxygen for mechanistic studies. Journal of the American Chemical Society. 2000;122:10212-10213
  35. 35. Rolim JP, De-Melo MA, Guedes SF, Albuquerque-Filho FB, De Souza JR, Nogueira NA, et al. The antimicrobial activity of photodynamic therapy against Streptococcus mutans using different photosensitizers. Journal of Photochemistry and Photobiology B: Biology. 2012;106:40-46
  36. 36. Yoon HY, Koo H, Choi KY, Lee SJ, Kim K, Kwon IC, et al. Tumor-targeting hyaluronic acid nanoparticles for photodynamic imaging and therapy. Biomaterials. 2012;33:3980-3989
  37. 37. Bäumler W, Regensburger J, Knak A, Felgenträger A, Maisch T. UVA and endogenous photosensitizers–the detection of singlet oxygen by its luminescence. Photochemical & Photobiological Sciences. 2012;11:107-117
  38. 38. BW, H. How does photodynamic therapy work? Photochemistry and Photobiology. 1991;55:145-157
  39. 39. Agostinis P, Berg K, Cengel KA, Foster TH, Girotti AW, Gollnick SO, et al. Photodynamic therapy of cancer: an update. CA: A Cancer Journal for Clinicians. 2011;61:250-281. DOI: 10.3322/caac.20114
  40. 40. Kim C, Meskauskiene R, Apel K, Laloi C. No single way to understand singlet oxygen signalling in plants. EMBO Reports. 2008;9:435-439
  41. 41. Ding H, Yu H, Dong Y, Tian R, Huang G, Boothman DA, et al. Photoactivation switch from type II to type I reactions by electron-rich micelles for improved photodynamic therapy of cancer cells under hypoxia. Journal of Controlled Release. 2011;156:276-280
  42. 42. Celli JP, Spring BQ, Rizvi I, Evans CL, Samkoe KS, Verma S, et al. Imaging and photodynamic therapy: Mechanisms, monitoring, and optimization. Chemical Reviews. 2010;110:2795-2838
  43. 43. Veerendra NR, Rekha RK, Chandana G, Sangeeta S. Photodynamic therapy. Indian Journal of Dental Advancements. 2009;1:46-51
  44. 44. Ormond AB, Freeman HS. Dye Sensitizers for Photodynamic Therapy. Materials (Basel). Mar 6 2013;6(3):817-840. DOI: 10.3390/ma6030817. PMID: 28809342; PMCID: PMC5512801
  45. 45. Detty MR, Gibson SL, Wagner SJ. Current clinical and preclinical photosensitizers for use in photodynamic therapy. Journal of Medicinal Chemistry. 2004;47:3897-3915
  46. 46. O’Connor AE, Gallagher WM, Byrne AT. Porphyrin and nonporphyrin photosensitizers in oncology: preclinical and clinical advances in photodynamic therapy. Photochemistry and Photobiology. 2009;85:1053-1074
  47. 47. Yano S, Hirohara S, Obata M, Hagiya Y, Ogura S-I, Ikeda A, et al. Current states and future views in photodynamic therapy. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2011;12:46-67
  48. 48. Dąbrowski JM, Pucelik B, Regiel-Futyra A, Brindell M, Mazuryk O, Kyzioł A, et al. Engineering of relevant photodynamic processes through structural modifications of metallotetrapyrrolic photosensitizers. Coordination Chemistry Reviews. 2016;325:67-101. DOI: 10.1016/j.ccr.2016.06.007
  49. 49. Allison RR, Downie GH, Cuenca R, Hu XH, Childs CJ, Sibata CH. Photosensitizers in clinical PDT. Photodiagnosis and Photodynamic Therapy. 2004;1:27-42. DOI: 10.1016/s1572-1000(04)00007-9
  50. 50. Usacheva MN, Teichert MC, Biel MA. Comparison of the methylene blue and toluidine blue photobactericidal efficacy against gram-positive and gram-negative microorganisms. Lasers in Surgery and Medicine. 2001;29:165-173. DOI: 10.1002/lsm.1105
  51. 51. Hamblin MR, O'Donnell DA, Murthy N, Rajagopalan K, Michaud N, Sherwood ME, et al. Polycationic photosensitizer conjugates: effects of chain length and Gram classification on the photodynamic inactivation of bacteria. The Journal of Antimicrobial Chemotherapy. 2002;49:941-951. DOI: 10.1093/jac/dkf053
  52. 52. Lehner R, Wang X, Marsch S, Hunziker P. Intelligent nanomaterials for medicine: carrier platforms and targeting strategies in the context of clinical application. Nanomedicine. 2013;9:742-757. DOI: 10.1016/j.nano.2013.01.012
  53. 53. Ni K, Luo T, Nash GT, Lin W. Nanoscale Metal-Organic Frameworks for Cancer Immunotherapy. Accounts of Chemical Research. 2020;53:1739-1748. DOI: 10.1021/acs.accounts.0c00313
  54. 54. Li L, Huh KM. Polymeric nanocarrier systems for photodynamic therapy. Biomaterials Research. 2014;18:19. DOI: 10.1186/2055-7124-18-19
  55. 55. Gibot L, Lemelle A, Till U, Moukarzel B, Mingotaud AF, Pimienta V, et al. Polymeric micelles encapsulating photosensitizer: structure/photodynamic therapy efficiency relation. Biomacromolecules. 2014;15:1443-1455. DOI: 10.1021/bm5000407
  56. 56. Bechet D, Couleaud P, Frochot C, Viriot ML, Guillemin F, Barberi-Heyob M. Nanoparticles as vehicles for delivery of photodynamic therapy agents. Trends in Biotechnology. 2008;26:612-621. DOI: 10.1016/j.tibtech.2008.07.007
  57. 57. Calixto GM, Bernegossi J, de Freitas LM, Fontana CR, Chorilli M. Nanotechnology-based drug delivery systems for photodynamic therapy of cancer: A review. Molecules. 2016;21:342. DOI: 10.3390/molecules21030342
  58. 58. Huang YY, Sharma SK, Dai T, Chung H, Yaroslavsky A, Garcia-Diaz M, et al. Can nanotechnology potentiate photodynamic therapy? Nanotechnology Reviews. 2012;1:111-146. DOI: 10.1515/ntrev-2011-0005
  59. 59. Schwiertz J, Wiehe A, Gräfe S, Gitter B, Epple M. Calcium phosphate nanoparticles as efficient carriers for photodynamic therapy against cells and bacteria. Biomaterials. 2009;30:3324-3331. DOI: 10.1016/j.biomaterials.2009.02.029
  60. 60. Lucky SS, Soo KC, Zhang Y. Nanoparticles in photodynamic therapy. Chemical Reviews. 2015;115:1990-2042. DOI: 10.1021/cr5004198
  61. 61. Deda DK, Araki K. Nanotechnology, light and chemical action: an effective combination to kill cancer cells. Journal of the Brazilian Chemical Society. 2015;26:2448-2470
  62. 62. Foote CSM, of photosensitized oxidation. There are several different types of photosensitized oxidation which may be important in biological systems. Science. 1968;162(3857):963-970
  63. 63. Fritsch C, Ruzicka T. Fluorescence diagnosis and photodynamic therapy in dermatology from experimental state to clinic standard methods. Journal of Environmental Pathology, Toxicology and Oncology: Official Organ of the International Society for Environmental Toxicology and Cancer. 2006;25(1–2):425-439
  64. 64. Moan JE. Properties for optimal PDT sensitizers. Journal of Photochemistry and Photobiology. B, Biology. 1990;5(3–4):521-524
  65. 65. Ochsner MIF. Photophysical and photobiological processes in the photodynamic therapy of tumours. Journal of Photochemistry and Photobiology. B, Biology. 1997;39(1):1-18
  66. 66. Dolmans D, Fukumura D, Jain RK. Photodynamic therapy for cancer. Nature Reviews Cancer. 2003;3:380-387
  67. 67. Juarranz Á, Jaen P, Sanz-Rodríguez F, Cuevas J, González S. Photodynamic therapy of cancer. Basic principles and applications. Clinical and Translational Oncology. 2008;10:148-154
  68. 68. Niculescu A-G, Grumezescu AM. Photodynamic Therapy—An Up-to-Date Review. Applied Sciences. 2021;11:3626
  69. 69. Fisher AMR, Murphree AL, Gomer CJ. Clinical and preclinical photodynamic therapy. Lasers in Surgery and Medicine. 1995;17:2
  70. 70. Braathen LR. Photodynamic therapy: When and why? Journal of the European Academy of Dermatology and Venereology. 2002;16:1063
  71. 71. Fialová P, Vašků, V. PDT and its possible use in dermatology. Dermatol. praxi. 2017;11:166
  72. 72. Ben-hur E, Horowitz B. Advances in photochemical approaches for blood sterilization. Photochemistry and Photobiology. 1995;62:383
  73. 73. Wyss P, Tadir Y, Tromberg BJ, Haller U, editors. Photomedicine in Gynecology and Reproduction. S. Karger AG; 2000. DOI: 10.1159/isbn.978-3-318-00453-3. ISBN (print): 978-3-8055-6905-7. ISBN (electronic): 978-3-318-00453-3
  74. 74. Van Vugt DA, Krzemien A, Roy BN, et al. Photodynamic Endometrial Ablation in the Nonhuman Primate. Reprod. Sci. 2000;7:125-130. DOI: 10.1177/107155760000700208

Written By

Celia María Curieses Andrés, José Manuel Pérez de la Lastra, Celia Andrés Juan, Francisco J. Plou and Eduardo Pérez-Lebeña

Submitted: 04 May 2023 Reviewed: 31 May 2023 Published: 22 October 2023