Mechanisms of triggering antitumor immunity: from photodynamic effects to immunogenic cell death

Cover Page


Cite item

Full Text

Open Access Open Access
Restricted Access Access granted
Restricted Access Subscription or Fee Access

Abstract

The development of cancer immunotherapies provides hope to millions of patients for better clinical outcomes after tumor treatment. Thus, investigating the fundamental mechanisms of antitumor immunity activation is an urgent task. Numerous studies have outlined the effect of immunogenic cell death (ICD) on cancer cells, and this outcome is both sophisticated and simple. Different stimuli can cause ICD; however, photodynamic exposure has been proven to be an effective inducer of programed death on par with radiotherapy. The link between triggering a photodynamic response in cancer cells and triggering ICD has been poorly described in experimental works. The question of which molecular cascades are activated after photodynamic therapy (PDT) irradiation and the way damage-associated molecular patterns (DAMPs) are released is intriguing. Much is known about reactive oxygen species generation and endoplasmic reticulum stress but little about the Golgi apparatus. Photosensitizers of different types can exert different effects, including completely nonimmunogenic ones.

This review describes the cascades that link the induction of cell death by photodynamic exposure and the immunogenic pattern of DAMP release. The photosensitizers that have shown potential as ICD inducers and the different pathways of programed death that occur during PDT exposure are also discussed.

Full Text

INTRODUCTION

The treatment of cancer is a significant biomedical challenge. Despite therapeutic strategies and clinical studies [1–5], cancer remains a severe illness that necessitates the creation of innovative, patient-specific therapies. Disease manifestations arise from the uncontrolled replication of tumor stem cells that exhibit unlimited self-renewal potential, insensitivity to growth inhibitory signals and high heterogeneity, and ability to invade and evade apoptosis [6]. These acquired characteristics confer invincibility upon cancerous cells and enable them to commandeer immune components of the organism to shield the tumor site from physiological immune assault [7, 8]. This leads to the wasting away of the patient; thus, the discovery of cancer immunotherapy has become a crucial breakthrough in cancer treatment [9].

Cancer immunotherapy is a method of treating cancer using the host immune system to control and eliminate cancer cells. The idea of training the immune system to recognize and combat cancer is sophisticated and straightforward. Nevertheless, researchers encounter challenges in reprograming immune cells because they have already attained the endpoint of proliferation and differentiation, and manipulating stem cell components may result in uncontrolled autoimmune reactivity. Therefore, the process of altering immune cell function can be arduous [10, 11]. In 2005, G. Kroemer et al. introduced a novel mode of cell death, called immunogenic apoptosis, which was later referred to as immunogenic cell death (ICD). This discovery sparked a new research field that aims to develop anticancer treatments based on the ICD concept, which involves inducing tumor cell death by enabling the immune system to recognize and respond to them [12]. The functions and properties of immune cells do not require alteration. However, modifying the modes of tumor cell death is a relatively straightforward task.

Immunogenic cell death is a regulated cell death modality characterized by the release of damage-associated molecular patterns (DAMPs), followed by the release of different cellular molecules, stress-related reactions, and a strong adaptive immunity response. The immunogenicity of such death is caused by the adjuvant signal formed by DAMP molecules and tumor antigens. DAMP profiles include extracellular ATP, calreticulin (CRT), heat shock proteins, nuclear high mobility group box 1 (HMGB1), annexin 1, and some members of the interleukin family [13–17]. In the last few decades, researchers have attempted to determine the usage of ICD in fundamental and clinical studies for cancer treatment [18–27]. Photodynamic therapy (PDT) along with several immunogenic dyes can be used as an ICD inducer [28–36]. These are produced by cell death during PDT, with PDT-treated cells breaking down into antigens that are easily taken up by antigen-presenting cells. Thus, this photoinduced tumor death facilitates the resetting of the immune system components to combat the tumor.

PDT-related immunogenicity is considered more potent when it generates concentrated endoplasmic reticulum (ER) stress by reactive oxygen species (ROS) rather than as an incidental or additional effect of ER stress, such as with some ICD inducers, for example, mitoxantrone and oxaliplatin. Specifically, PDT has fewer molecular elements that mediate danger signaling pathways and consequently aids in DAMP turnover [13, 37]. As a result, PDT has higher effectiveness and stability in inducing tumor immunogenicity than other ICD inducers. However, no molecular mechanism has been described to establish a connection between photodynamic damage and immunogenic release of DAMPs [34, 38]. Most research groups commonly describe ICD without specifying the type of apoptosis or other ICD pathways that it undergoes at the molecular level.

This review offers a succinct overview of the molecular pathways responsible for ICD development by photodynamic effects. The traditional effects of photodynamics are described, and which photosensitizers are more likely to produce immunogenicity are investigated. This review also discusses the activation cascades in diverse cellular structures. Furthermore, the mechanisms underlying the oxidative stress caused by PDT at the subcellular and molecular levels are elucidated. Finally, we postulated that photodynamic exposure can induce a range of regulated and ICD modalities.

POTENTIAL OF PHOTOSENSITIZERS FROM VARIOUS GENERATIONS TO INDUCE IMMUNOGENIC CELL DEATH

Photodynamic therapy relies on the use of a photosensitive compound, known as a photosensitizer, which accumulates specifically in cancerous tissues. Photosensitizer molecules absorb light of a specific wavelength, initiating a process that destroys proteins, lipids, and other cell components. This selective destruction leads to the elimination of pathological cells.

Currently, one of the commonly used photosensitizer classifications encompasses the physical and chemical properties of three photoactive dye generations [39, 40]. First-generation photosensitizers are dyes based on the porphyrin framework: a hematoporphyrin derivative (HpD) and photofrin II (a purified form of HpD) [39, 41]. These agents exhibit a marked response following PDT; however, they impose considerable restrictions for clinical acceptance as effective antitumor therapy. Limitations include hydrophobicity, limited penetration depth, aggregation in water, low singlet oxygen emission, low tumor selectivity, high therapeutic dosage, prolonged half-life, and skin photosensitivity [42, 43].

The abovementioned restrictions on first-generation photosensitizers have led to wide-ranging studies to improve the efficacy of photosensitizer molecules by altering the peripheral functionality of porphyrin [44] or by direct modification of the porphyrin core [45]. The addition of hydrophilic radicals [46–50] to pyrrole rings resulted in increased water solubility and overall molecular stability. Nonetheless, better water solubility increases renal clearance, followed by lower bioavailability and assimilability of the photosensitizer. Skin photosensitivity persists when second-generation photosensitizers are used [43]. Only several second-generation photosensitizers are under clinical trials or approved for clinical use in cancer treatment [51–54].

Finding an appropriate photosensitizer is a protracted process. Advanced third-generation photosensitizers unify the conjugation of second-generation photosensitizers with targeting agents (e.g., carbohydrates, antibodies, amino acids, and peptides) or a combination of delivery carriers (nanocapsules, nanoparticles, liposomes, protein dots, and micelles) with encapsulated derivatives of second-generation photosensitizers [55]. Third-generation photosensitizer research aims to improve the physical and chemical characteristics of photoactive drugs, resulting in higher tumor affinity, ROS generation, and reduced adverse effects on the surrounding normal tissue. Nanoparticle delivery vehicles can play a role in multiple steps of immune system activation to suppress cancer. Nanoparticle-based therapeutics can induce tumor cell death and, in turn, increase neoantigen release from the tumor. Nanoparticles can be used to improve antigen presentation and T-cell activation [55–59].

Investigation into ICD, stimulated in cancer cells by PDT, represents a nascent subfield of ICD research. The application of this field is presently broadening the possibilities of cancer therapy. Interestingly, all three generations of photosensitizers can induce ICD, opening up a new area of research and potentially providing a new avenue for evaluating previously described photosensitizers for ICD-PDT. Thus, we provide a comprehensive table containing details regarding photosensitizers that can induce ICD in vitro, supported by an in vivo demonstration of activating specific antitumor immune responses (Supplement 1, Table 1) [60–122].

In photodynamic treatment-induced ICD, recent scientific studies have primarily described second- and third-generation photosensitizers [34, 67, 84]. However, the immunogenicity of certain first-generation photosensitizers remains under investigation [91, 109]. Elaborating on a photoactive drug is a complex and multistage process that requires a thorough understanding of the chemical and physical properties of photosensitizers, their interaction with target tumor cells, and their potential for clinical use in cancer control [43, 123].

Numerous research teams have focused on developing third-generation photosensitizers, which are first- or second-generation photosensitizers that have been augmented with nanocarriers such as nanoparticles, liposomes, and nanocapsules. This trend is primarily explained by the enhancement of critical features, including increased tumor affinity and improved photoactive agent stability [55, 100, 123]. All three generations can induce cell death, which activates dendritic cells in vitro and produces specific immune responses in vivo following different treatment modalities. Consequently, these photosensitizers have the potential for further research into patient-specific cancer treatment [73, 93, 104].

Definitively determining the most efficient generation of photosensitizers or selecting a single drug is impossible, given the evidence of comparable activation of dendritic cells or the emergence of antitumor immunity in each documented case. However, not only the in vitro and in vivo efficacy on animal models but also the requirements for an accepted photosensitizer, such as high water solubility, tumor specificity, and low or absence of cytotoxicity to healthy tissue, must be considered [124–126]. This presents prospects for further investigation into the identification of photoactive drugs that are both immunogenic and meet the requirements for clinical use.

PHOTOSENSITIZER ACCUMULATION IS A CRUCIAL STEP IN THE INITIATION OF CELL DEATH

Photosensitizers can percolate into cells by endocytosis or diffusion. Penetration and localization types depend on the size of the molecules, polarity, total charge, distribution of charged groups, and ability to form hydrogen bonds. Hydrophobic photosensitizers mainly accumulate in the plasma membrane and the membranes of cellular organelles because they easily penetrate the lipid phase of the membranes; however, the reverse process is impeded. Hydrophilic photosensitizers enter the cell mainly through endocytosis. Amphiphilic substances best enter the cell, are distributed in it, and photosensitize [127]. These terms define further the localization of photosensitizers and the way they accumulate in the intracellular space. The accumulation in the body critically indicates the place of ROS formation as a photodynamic irritation effect.

After photosensitizers accumulate in cells, it is irritated by the light of the visible spectrum. After light absorption, the photosensitizer in its ground state is activated to the short-lived excited singlet state, after which it loses its energy. This occurs by emitting light during PDT or by generating heat during photothermal therapy. The excited singlet state may also undergo intersystem crossing to form the relatively long-lived excited triplet state. The excited triplet state can then undergo two types of reactions with surrounding molecules. Type I and II photochemical reactions can be simultaneously performed, and the ratio between them depends mainly on the type of photosensitizers used, substrate concentrations, and availability of oxygen. Because of the photodynamic reaction, various molecular mechanisms are activated, leading to different cell death modalities [36]. Although these primary ROS are short-lived, sample evidence shows that PDT induces prolonged oxidative stress in PDT-treated cells. PostPDT oxidative stress stems from (per)oxidized reaction products such as lipids and proteins that have a longer lifetime and, in addition to acutely generated ROS, depletion of intracellular antioxidants and, hence, further exacerbation of already perturbed intracellular redox homeostasis [128]. The first type of reaction often leads to the formation of superoxide ions by transferring one electron to an oxygen molecule. These ions do not act as active ions in biological systems but can produce H2O2, which is easily absorbed by cell membranes. At high concentrations, H2O2 can react with superoxide molecules through the formation of hydroxyl radicals, which can ionize molecules with low activation energy. In reaction II, the photosensitizer molecule transforms into excited singlet oxygen by transitioning from the triplet state to the ground state and transferring energy to the oxygen molecule. Singlet oxygen as a molecule without charge can spread into the cytoplasm and biological membranes. Type I and type II reactions, as direct effects of PDT, occur in parallel, depending on the photosensitizer type and oxygen concentration. For more commonly used photosensitizers, the type II reaction is the dominant process [129].

The localization of the photosensitizer is significant for PDT-based ICD. Moreover, its localization during irradiation in cellular compartments largely determines where oxidative damage will occur because of the short lifetime and limited diffusion capacity of ROS (1O2) [130, 131]. If the photosensitizer accumulates in membranes, the cell dies from the rupture of membranes because of necrosis [132]. If the photosensitizer accumulates in the nucleus, the main consequence of PDT is DNA damage [133]. In this case, ICD can be initiated; however, it will mostly depend on p53, BAX, and caspase-8. As mentioned above, this makes the immunogenic pathway difficult for tumors with p53 mutations, which are present in the cells of several human tumors [134]. Photodamage of the mitochondria releases the products they contain, for example, the release of cytochrome c into the cytoplasm, which triggers apoptosis (Fig. 1) [135]. The most optimal localizations of photosensitizers are ER, AG, lysosomes, and mitochondria. In this case, oxidative stress occurs, which quickly leads to an internal pathway that triggers apoptosis [136–138]. Today, PDT-based ICDs, which directly cause growth stress in these cell organoids, are hypericin PDT [139], photosens PDT, photoditazine PDT [34], and methylene blue PDT [140].

 

Fig. 1. Main mechanisms triggering immunogenic cell death under photodynamic exposure. The localization of the photosensitizer in the endoplasmic reticulum leads to misfolding of proteins, which triggers three pathways of endoplasmic reticulum stress. The accumulation of reactive oxygen species in the ER during photodynamic therapy also leads to the loss of calcium homeostasis and its outflow into the cytosol or mitochondria, including through mitochondria-associated membranes. When a photosensitizer is localized in the mitochondria, large amounts of reactive oxygen species accumulate in organelles, and their integrity is subsequently lost, accompanied by the release of numerous apoptotic factors. The accumulation of photosensitizers and reactive oxygen species formation in lysosomes lead to the loss of lysosomal calcium and the convergence of ATG5 to tATG5, followed by the triggering of mitochondrial apoptosis. MMP — mitochondrial membrane potential; Bsl-XL — B-cell lymphoma-extra large. See text for designations.

Рис. 1. Основной механизм запуска иммуногенной клеточной смерти вследствие фотодинамического воздействия. Локализация фотосенсибилизатора в эндоплазматическом ретикулуме приводит к неправильному сворачиванию белков, что запускает три пути стресса эндоплазматического ретикулума. Накопление активных форм кислорода при фотодинамической терапии в эндоплазматическом ретикулуме также приводит к снижению гомеостаза кальция и его выходу в цитозоль или митохондрии, в том числе через ассоциированные с митахондриями мембраны. При локализации фотосенсибилизатора в митохондриях в органеллах накапливается большое количество активных форм кислорода и впоследствии происходит нарушение их целостности, сопровождающееся высвобождением ряда апоптотических факторов. При накоплении фотосенсибилизаторов и появлении активных форм кислорода в лизосомах происходит потеря лизосомального кальция и конверсия ATG5 в tATG5, после чего запускается митохондриальный апоптоз. MMP — мембранный потенциал митохондрий; Bsl-XL — B-клеточная лимфома экстракрупного размера. Обозначения см. в тексте.

 

ROLE OF REACTIVE OXYGEN SPECIES IN OXIDATIVE AND ENDOPLASMIC RETICULUM STRESS

Reactive oxygen species are produced in each aerobic cell and are involved in some metabolic processes [141]. ROS generation depends on several enzymes, such as nicotinamide adenine dinucleotide phosphate oxidase (NADPH, which converts electrons into molecular oxygen), xanthine oxidoreductase, and peroxidases, as well as mitochondria containing electron transport systems [141]. However, cells have antioxidant defense systems, including catalase, superoxide dismutase (SOD), glutathione (GSH) peroxidase, and other exogenous substances with antioxidant properties that can significantly prevent ROS production by direct action on radical chain reactions and detoxification enzymes [141, 142]. However, under prolonged and powerful oxidative stress, the balance between ROS and the oxidative system is disrupted, homeostasis/proteostasis may not be restored, and cell death mechanisms, such as apoptosis, are activated [143].

In tumor cells, high levels of intracellular ROS are a frequent phenomenon, which may be caused by their increased metabolic activity, mitochondrial dysfunction due to constant hypoxic conditions, peroxisome activity, constant presence of growth factors, and oncogene activity. In addition, the activities of NADPH oxidase, cyclooxygenase, and lipoxygenase, which are ROS sources, are increased in tumor cells. ROS can promote tumor proliferation and mutagenic transformation by affecting intracellular signaling pathways that control proliferation, survival, and stimulation [144].

Reactive oxygen species formation is an important feature of PDT-induced cell death, particularly ICD. ROS release in this situation is the standard defense mechanism [145–147], which some authors call mitochondrial DAMPs [17, 148, 149]. Oxidative stress emerges not only after PDT; it is also caused by chemotherapy [150], radiotherapy [151], oncolytic viruses [152], and other anticancer treatments [153–156]. In this case, ROS are formed as in the case of specially stimulated exposure (Hyp-PDT), so it is synthesized by the cell itself in response to stress (anthracyclines) [133]. Primary ROS are formed when a photosensitizer is irradiated. These primary ROS are short-lived. Oxidative stress after PDT occurs because of (re)oxidized reaction products, such as lipids and proteins, which have a longer lifetime. In addition to ROS accumulation, intracellular antioxidants are depleted, which further exacerbates the already disturbed intracellular redox homeostasis [128].

The ER is one of the key centers of oxidative stress during the induction of ICD-based PDT [157]. The ER is highly sensitive to stresses that perturb cellular energy levels, the redox state, or Ca2+ concentration. ER disturbance reduces the protein-folding capacity in the ER, which results in the accumulation and aggregation of unfolded proteins, a condition referred to as ER stress [158]. To avert the harmful effects of ER stress, cells have evolved various protective strategies, collectively termed unfolded protein response (UPR) (See Fig. 1). This concerted and complex cellular cascade is mediated through three ER transmembrane receptors: pancreatic ER kinase (PKR)-like ER kinase (PERK), activating transcription factor 6 (ATF6), and inositol-requiring enzyme 1 (IRE1). Normally, all three ER stress receptors are maintained in an inactive state through their association with the ER chaperone 78 kDa glucose-regulated protein (GRP78). On accumulation of unfolded proteins, GRP78 dissociates from the three receptors, which leads to their activation and triggers UPR [158]. The UPR is a prosurvival response that reduces the accumulation of unfolded proteins and restores normal ER functioning. On the aggregation of unfolded proteins, GRP78 dissociates from the three ER stress receptors, PERK, ATF6, and IRE1, allowing their activation. The receptors are activated sequentially, with PERK being the first and rapidly followed by ATF6, whereas IRE1 is activated last. Active PERK blocks general protein synthesis by phosphorylating eukaryotic initiation factor 2α (eIF2α) that enables activating transcription factor 4 (ATF4) translation, which occurs through an alternative, eIF2α-independent translation pathway. ATF4, being a transcription factor, translocates to the nucleus and induces the transcription of genes required to restore ER homeostasis [159]. Important targets driven by ATF4 are transcriptional factor C/EBP homologous protein (CHOP), growth arrest and DNA damage-inducible 34 (GADD34), and ATF3 [160]. ATF3 contemporaneously regulates Hsp40 transcription, which is a staple DAMP for ICD [128]. In addition, PERK provides effective interaction between ER and mitochondrial membranes through ROS and Ca2+ transportation and plays a crucial role in the mitochondria and ER contact site stabilization and MAM (mitochondria-associated membranes) integrity [138]. ATF6 is activated by limited proteolysis after its translocation from the ER to the Golgi apparatus (GA). Active ATF6 is also a transcription factor that regulates the expression of ER chaperones and X box-binding protein 1, another transcription factor (XBP1) [158]. ATF6 is also responsible for some DAMP transcription, including CRT, HMGB1, and HSP90 [128].

To achieve its active form, XBP1 must undergo mRNA splicing, which is performed by IRE1. Spliced XBP1 protein translocates to the nucleus and controls the transcription of chaperones, cochaperones, and PERK inhibitor P58IPK, and genes involved in protein degradation. This concerted action restores ER function by blocking further buildup of client proteins, enhancing the folding capacity, and initiating degradation of protein aggregates CHOP and C/EBP homologous protein. Under conditions of prolonged oxidative stress, homeostasis/proteostasis may not be restored; instead, mechanisms of cell death, such as apoptosis, are activated [159]. Interestingly, ER stress can lead not only to apoptosis but also to paraptosis [161, 162]. However, this mechanism is still poorly understood in the context of the IСD path.

An important feature of ER and GA is the presence of Ca-related proteins. Specific and limited PDI oxidation by endoplasmic reticulum disulphide oxidase 1α (ERO1α) is essential for avoiding ER hyperoxidation. Under normal physiological conditions, the ER forms an oxidizing environment, and Ca2+ stores are filled, allowing the proper function of the various chaperones. During ER stress, ERO1α is upregulated in a CHOP-dependent manner, leading to ER hyperoxidation. Because the conformation of the third luminal loop of inositol trisphosphate receptor (IP3R) depends on the oxidation state, hyperoxidation could disrupt the interaction between ERp44 and IP3R, causing IP3R hypersensitivity, increased calcium release, and induction of apoptosis [162].

Calcium from ER cisternae flows mainly through calcium release channels as IP3R. This channel is accumulated in MAM, which are associated with the mitochondrial outer membrane. Calcium ions from the cytoplasm enter the mitochondria through voltage-dependent anion channels or calcium uniporters. High levels of calcium stimulate respiratory chain activity, leading to higher ROS levels. ROS can further target ER-based calcium channels, leading to the increased release of calcium and further increased ROS levels. Increased ROS and calcium load can open mitochondrial permeability transition pores, resulting in the release of proapoptotic factors [163]. Furthermore, calcium unleashed from the ER binds and activates several MAM receptors, such as adenine nucleotide translocase (ANT), on the mitochondrial membrane. ANT activation leads to depolarization and membrane bursts followed by proapoptotic factors, cytochrome c, and other cell death-dependent molecules [164].

Calcium ions activate the calcium-dependent enzyme calpain, which then triggers caspase-12, which is located in the ER. This enzyme mobilizes to the cytosol and subsequently initiates the apoptotic cascade via interactions with caspase-9 and caspase-3 [165]. Moreover, ROS formation not only in the ER but also in the GA and lysosomes can trigger this mechanism [166]. In this case, calcium is involved in oxidative stress in both the mitochondria and EP. The constancy of the calcium current can be disrupted both at the onset of oxidative stress in the mitochondria and ER, which in total will lead to the same consequences.

In the mitochondria, ROS typically fulfill numerous crucial roles, such as contributing to glucose oxidation and synthesis of fatty acids, amino acids, and hormones. Furthermore, the mitochondria serve as a hub for ROS signaling during apoptosis and innate immune responses. Concurrently, the mitochondria produce ROS and possess an influential antioxidant mechanism to counteract ROS-mediated destruction, comprising mitochondrial superoxide dismutase (SOD2) and a thioredoxin system composed of thioredoxin-2 (Trx2), thioredoxin reductase-2, and peroxidase 3 [167]. Various methods by which prolonged opening of the pores during mitochondrial permeability transition (MPT) and the permeability of the outer mitochondrial membrane can destroy the inner and outer mitochondrial membranes. This reaction is an initial response to oxidative stress and is a crucial stage in cell death. In general, MPT pores become open and remain open after mitochondrial damage, which is linked to compromised calcium levels and oxidative stress [148]. In PDT, oxidative stress triggers a sequence of events resulting in decreased permeability of the inner mitochondrial membrane, thereby regulating the MPT pore complex and disrupting the membrane potential. Although photosensitizers may accumulate outside the mitochondria, these events primarily occur within this cellular compartment [168]. When singlet oxygen accumulates directly in the mitochondrial membrane, cytochrome c is released. This binds to apoptotic protease activating factor 1 (APAF-1) and procaspase-9, forming a complex that activates caspase-3 [143, 169]. Furthermore, the disintegration of the mitochondrial membrane releases a crucial DAMP, ATP [148, 170].

Lysosomal photodamage results in lysosomal calcium depletion, activating the calpain protease. In addition, it facilitates ATG5 cleavage, an apoptosis-related protein, into a fragment named tATG5. Subsequently, tATG5 attaches to the mitochondrial membrane and triggers a series of events that ultimately culminate in the loss of cytochrome c, catalyst for apoptosis (See Fig. 1) [171]. When the lysosomal membrane is permeabilized, proteases such as cathepsins B and D are released. They subsequently activate Bid proteolytically, leading to mitochondrial apoptosis and release [133, 169].

NOT JUST APOPTOSIS

Apoptosis and necrosis are the predominant cell death modes triggered by PDT. Apoptosis is an energy-dependent cell death modality that is regulated. Several morphological changes characterize it, including chromatin condensation, cytoplasmic reduction, plasma membrane blebbing, and phagocytic activity [172, 173]. Necrosis is an unregulated cell death type induced by external damage [173], characterized by cellular swelling along with chromatin condensation, eventually leading to cellular and nuclear lysis with subsequent inflammation [174]. In addition, some cell death pathways such as autophagy, mitotic catastrophe, cellular senescence, and programed necrosis have been described and summarized in other reviews [175, 176]. The extent to which photodynamic treatment can induce ferroptosis remains unclear [34].

Because the tumor microenvironment is naturally poor in oxygen, the reactions leading to ROS formation under PDT are usually rapidly attenuated. A recent study reported that the presence of Fe(III) in nanoparticles derived from the initial hem supports the Fenton redox reaction, which provides oxygen as the starting material for 1O2 formation. Furthermore, Fe(III) can react with the antioxidant capabilities of GSH, resulting in its deactivation. Inhibiting cystine uptake via the cystine/glutamate antiporter through p53 activation limits the production of intracellular GSH, ultimately shielding tumor cells from ferroptosis [177, 178]. GPX4 plays a crucial role in the reduction of phospholipid hydroperoxides (PLOOH) to phospholipid alcohols. Furthermore, GSH is essential for the normal physiological operation of GPX4. Intracellular synthesis of GSH is catalyzed by glutamate–cysteine ligase. Therefore, cysteine, which is taken up as cystine by the cystine/glutamate antiporter (xCT) system, has been described as the rate-limiting amino acid in GSH synthesis. Therefore, depriving cells of cysteine by inhibiting the xCT system also indirectly contributes to GPX4 inhibition. Consequently, GPX4 inactivation can lead to PLOOH accumulation, resulting in cell membrane damage and ferroptotic death [177, 178].

PDT-ACTIVATED IMMUNOGENIC CELL DEATH PATHWAYS

We were interested in the relationship between photodynamic exposure and ICD induction. Several studies have convincingly demonstrated the link and identified the pathways. Chen et al. described the mechanism of apoptosis induced by methylene blue-mediated photodynamic therapy (MB-PDT) in B16F1 cells. Western blotting showed the activation of caspase-3-dependent apoptosis 6 h after MB-PDT. Furthermore, immunoblot analysis showed a rapid increase in caspase-9 and cytochrome c release, indicating the activation of the intrinsic pathway. Apoptosis occurs via the intrinsic mitochondrial pathway through cytochrome c release and subsequent caspase-9 activation. Cell death depends on mitochondrial stress [179]. Intrinsic apoptosis is activated by intracellular and extracellular stresses. The cascade is initiated by mitochondrial membrane permeabilization and subsequent release of cytochrome c into the cytosol. The released cytochrome c binds to APAF-1, and procaspase-9 enters a mature state and forms the apoptosome. Caspase-9 cleaves procaspase-3. Thus, apoptosis-inducing factor released into the cytosol rapidly enters the nucleus and promotes nuclear DNA degradation in a caspase-independent manner [180]. In another study using MB-PDT, Lu et al. described apoptosis in HeLa cells. The detection of cytochrome c and poly (ADP-ribose) polymerase suggests the occurrence of intrinsic apoptosis [181].

Xie et al. showed that translocator protein-targeted photodynamic therapy (TSPO-PDT) on MC38 cells induces immunogenic apoptosis as a mode of ICD [182]. Caspase-3 was also involved in intrinsic apoptosis. The authors also analyzed the association between apoptosis and DAMP release, following the “golden standard” for ICD indication [183], and found high levels of CRT and heat shock protein 70 (HSP70). Interestingly, the effects of TSPO-PDT were measured in the context of ER stress. The upregulation of ER stress markers was observed at different time points, indicating the ability of TSPO-PDT to induce ER stress. A high rate of dendritic cell maturation was observed [182].

Panzarini et al. described several cell death pathways after RBAc-PDT [184]. In this study, the cell type was differentiated by quantifying caspase-9, -8, -3, and cleaved caspase-12. The highest levels of caspase-9 were observed immediately after PDT. This process correlated with the onset of caspase-3 and cytochrome c release, and early appearance of apoptotic cells at the 12-h time point. At 8 h after PDT, caspase-9 levels were lower, whereas caspase-3 levels were still high at 72 h after PDT, indicating further activation of the apoptotic pathway at different time points [182, 184].

The key role of caspases-3 and -9 in the induction of immunogenic apoptosis after PDT is evident. Cytochrome c represents the initiation of the intrinsic pathway of apoptosis, which appears to be associated with ATP release and nuclear membrane disassembly with the release of HMGB1. This mechanism requires further discussion and refinement.

DISCUSSION

Cells have various pathways that undergo rapid ICD during photodynamic exposure. ROS emergence activates cascades interconnected with ER and GA, resulting in transcriptional changes that lead to the release of DAMPs and cell death. In 2012, this unique property was applied in the development of a technology to treat mouse glioblastoma with immunogenic dying cells [123]. PDT-related ICD is more effective if it is subjected to focused straight ROS-provided ER stress but not secondary or side ER stress effects. Moreover, targeting the ER allows the immediate triggering of molecular cascades leading to the transcription and emission of DAMPS.

Endoplasmic reticulum is not the sole initiator of ICD in organoids. By combining two photosensitizers according to [185], a pathway that harms both the lysosomes and mitochondria can be triggered. By activating a lysosome-driven pathway, photofrin enables the benzoporphyrin derivative to accumulate in the mitochondria, ultimately resulting in tumor reduction and better therapeutic outcomes. Both lysosome- and mitochondria-dependent pathways can induce ICD.

Whether there is a standard approach for inducing ICD with PDT for treating murine and human tumor cells remains unclear. Many studies have focused on activating death pathways that encourage an immune response in tumor cells, and positive findings suggest that ICD triggered by photoimmunotherapy is an area of significant investigation. Notably, certain tumors, such as mammary carcinoma 4T1, melanoma B16F10, and colon adenocarcinoma CT26, exhibit increased immunogenic potential for immunogenic PDT [73–79]. ICD-PDT is used for the treatment of highly challenging and life-threatening tumors, such as glioma, which are usually resistant to traditional therapies [30, 80].

Further research should concentrate on discovering novel ICD inducers that have clinical potential. In addition, the integration of PDT as a powerful activator of this process should be explored. Previous studies have shown that PDT can induce a mixed form of cell death, combining apoptosis and ferroptosis. Furthermore, ROS affect the expression of PDL1 in tumor cells, increasing therapeutic effectiveness [186, 187]. PDL1 is a protein that belongs to the group of T-cell suppressor receptors. Its function involves inhibiting apoptotic T-cell immunological responses by binding to PDL1 on the surface of tumor cells [188].

In conclusion, research in the last three decades has increasingly proved that PDT is a powerful strategy for ICD induction. However, identifying the molecular mechanisms of the relationship between PDT and ICD may help identify effective inducers and modulate the immunogenicity of the therapy.

CONCLUDING REMARKS

In conclusion, the research of the last three decades have increasingly proved the idea that PDT is a powerful strategy for immunogenic cell death induction. However, uncovering the molecular mechanisms of the relationship between PDT and immunogenic cell death may help to uncover effective inducers and modulate the immunogenicity of the therapy.

ADDITIONAL INFORMATION

Funding source. The study was supported by the Russian Science Foundation grant No. 22-25-00716, https://rscf.ru/project/22-25-00716/

Competing interests. The authors declare that they have no competing interests.

Authors’ contribution. E.E. Sleptsova, T.S. Redkin — literature review, writing the text and editing the article; M.O. Saviuk — editing the article; V.D. Turubanova — conceptualization and administration of the study, writing the text and editing the article.

ДОПОЛНИТЕЛЬНАЯ ИНФОРМАЦИЯ

Источник финансирования. Исследование было поддержано грантом Российского научного фонда № 22-25-00716, https://rscf.ru/project/22-25-00716/

Конфликт интересов. Авторы декларируют отсутствие явных и потенциальных конфликтов интересов, связанных с публикацией настоящей статьи.

Вклад авторов. Е.Е. Слепцова, Т.С. Редькин — обзор литературы, написание текста и редактирование статьи; М.О. Савюк — редактирование статьи; В.Д. Турубанова — концептуализация и администрирование исследования, написание текста и редактирование статьи.

ДОПОЛНИТЕЛЬНЫЕ МАТЕРИАЛЫ

Приложение 1

 

Table 1. ICD-inducing photosensitizers

Таблица 1. Фотосенсибилизаторы, индуцирующие иммуногенную клеточную смерть

Photosensitizer (PS)

PS generation and type of additional structure

Tumor cell line

Intracellular localization of PS and possible cell death modality

ICD markers and signs of adaptive immunity activation

References

Indocyanine green (ICG)

1st

HT-29, T84, human colorectal adenocarcinoma; human colon cancer

ER, GA, lysosomes

Overproduced ROS

[60, 61]

3rd, ICG@Fe/FeO–PPP NCs

KB, epidermal carcinoma

Overproduced ROS

[62]

3rd, Pt-ICG/PES and Mg-ICG/PES

4T1, murine mammary carcinoma

CRT, HMGB-1, ATP; BMDCs maturation (CD11c+CD80+CD86+) 26.02±0.98% for Pt-ICG/PES and 24.47±0.54% for Mg-ICG/PES of total DCs population

[63]

3rd, Dox/ICG BSA NPs,

T/Dox-ICG NPs

B16F10, murine melanoma

SGC7901/VCR, gastric cancer

CRT, HSP 70, HSP 90 exposition

[64, 65]

Oxaliplatin

3rd, OXP/ICG/PEP OI_NP

ID8, ovarian cancer

CRT, HMGB-1, ATP, ROS;

Cytotoxic response of T lymphocytes to tumor cells

[66]

Benzophenazine photosensitizer OR141

3rd

SCC7, murine squamous cell carcinoma; A43, human squamous cell carcinoma

ER

HMGB1, ATP, Annexin V, Annexin A1; MHC-I on cancer cell surface; BMDCs maturation (MHC-II+, CD80+, CD86+) more than 50% of whole BMDCs population

[67, 68]

PIO-based PSs

3rd, PIO + anti-PD-L1

HeLa, human cervical cancer

Predominantly ER

GRP78 emission; CRT, HMGB1; F4/80 macrophages, CD4+, CD8+ T cells in tumor; INF-α, IFN-γ, and IL-6 in serum

[67, 69]

(3-bromopropyl) trimethylammonium bromide

3rd, Cu SAZs nanozyme (PDA NPs)

CT26, murine colon adenocarcinoma

Overproduced ROS

[70]

Semiconducting polymer (SP) nanoparticles

3rd, SPSS NPs

CT26, murine colon adenocarcinoma

Lysosomes

Overproduced ROS; CRT HMGB1, HSP70, ATP; ecto-CRT, IFN-γ, TNF-α, granzyme B in tumor; CD8+ T cells in tumor and blood

[71]

Boron difluoride dipyrromethene (BODIPY)-vadimezan conjugate (BDPVDA)

3rd, PVB NPs

4T1, breast cancer (mimics human)

ER

Overproduced ROS

[72]

Heptamethine cyanine small molecules containing with 8 eight carbon atoms of N-alkyl side chains

2nd

4T1, breast cancer (mimics human)

CRT, HMGB-1; DC (CD11c+CD86+ and CD11c+CD80+) maturation; TNF-α, IL-6, IL-12p70 in blood

[73]

Ferrocene-containing Ir(III) photosensitizer (IrFc1)

2nd

4T1, breast cancer (mimics human)

Lysosomes

Overproduced ROS; CRT, HMGB-1, АТP; ferroptosis (GPX4), changing mitochondrial morphology; CD8+ and CD4+ T cells, DC in tumor; IL-12p70, IL-6, TNF-α, IL-10 in blood

[74]

Redaporfin (5,10,15,20-tetrakis (2,6-difluoro-3-N-methylsulfamoylphenyl) bacteriochlorin)

3rd

CT26, murine colon adenocarcinoma

ER and GA

IL-2, IL-4, IL-6, IFN-γ, TNF- α, IL-17A и IL-10 in serum; granulocytes, neutrophils, CD4+ and CD8+ Т cells in blood; T cells infiltration in tumor

[75, 76]

Aluminum-phthalocyanine chloride

2nd, AlPcNE (aluminum-phthalocyanine nanoemulsion)

B16F10, murine melanoma

CRT, HMGB1, АТP

[35, 77]

Photosensitizer (PS)

PS generation and type of additional structure

Tumor cell line

Intracellular localization of PS and possible cell death modality

ICD markers and signs of adaptive immunity activation

References

Aluminum phthalocyanine disulfonate

2nd

Rat glioma (F98 and BT4C)

Limited tumor development in experimental groups (histology)

[78, 79]

Zinc(II)phthalocyanine

2nd + anti-PD-L1 therapy

B16F10, murine melanoma

ER, mitochondria and lysosomes

TNF-α, IFN-γ and IL-6; CD8+ T and CD4+ T cells in tumor

[80, 81]

3rd, DH@ZnPc NPs

TC1, murine cervical cells; HeLa human cervical

cells

Early (1 h) — lysosomes, late (24 h) — mitochondria

CRT, HMGB1; pyroptosis (caspase-1, NLRP3, IL-1β, IL18)

[82]

Silicon(IV) phthalocyanine

2nd

CT26, murine colon adenocarcinoma

CRT, HMGB-1, АТP, ERp57, vasostatin, HSP90 CXCL1, CXCL2, CXCL3, CXCL10, CXCL12, CXCL13

[83, 84]

Photosens

2nd

GL261, murine glioma; MCA205, murine fibrosarcoma

ER and GA

CRT, HMGB-1, АТP; induced DC maturation; IL-6 in serum

[34, 78]

Porphyrin-cisplatin conjugate

3rd, NP@Pt-1

CT26, murine colon adenocarcinoma

Cytosol

CRT, HMGB-1, АТP; mature DCs (CD11C+, CD80+, CD86+) and CD8+ T cells in tumor

[25]

Aluminum-phthalocyanine

2nd, aluminum-phthalocyanine nanoemulsion

B16F10, murine melanoma

CRT, HMGB-1, АТP;

[35, 78]

CT26, murine colon adenocarcinoma; 4T1, murine mammary adenocarcinoma

CRT, HSP70, HSP90, HMGB-1; IL-1β

[78, 85]

M-chlorin e6

2nd

4T1, murine mammary adenocarcinoma

ER

CRT; increased proportion of CD80+ CD86+ macrophages

[86–88]

3rd, PEG-Chlorin e6 (Ce6)-Fe2+ -Gossypol NPs

CRT, HMGB1, АТP; IL-6, IFN-γ, IL-12p70, TNF-α;

DC maturation 80%; overproduced ROS

[89]

3rd, Ce6/MLT@SAB NPs

Overproduced ROS CRT, АТP; increased percentage of CD80+ CD86+ cells

[90]

Hematoporphyrin

1st

C6, rat glioma

Mitochondria

DC maturation (42.8 CD80+, 50.82 OX-6+); IL-12, IL-10, TNF-α; co-culturing with spleen cells cytotoxicity 95.50±0.016%, 90.20±0.024%

[54, 91, 92]

3rd, HMME/R837@Lip

4T1, murine mammary adenocarcinoma; CT26, murine colon adenocarcinoma

Prophylactic vaccination: TNF-a, IL-6 и IL-12p70; 15% DC maturation (HMME@Lip), 21% DC maturation (HMME/R837@Lip); CD45+ leukocytes in tumor;

Therapeutic vaccination: CD4+ CD8+ T cells in spleen; IFN-γ, TNF-α

[93]

IRdye700DX

2nd, NIR-PIT: Antibody+IR700

3T3-HER2, transfected fibroblasts

MC38, murine colorectal carcinoma

Lysosomes, membranes

CRT, HSP70, HSP90, HMGB1, ATP; induced DC maturation (CD80, CD86, CD40, HLA-DR, IL-12)

[94–96]

2nd

MC38, murine colorectal carcinoma

A431, human epidermis carcinoma

CD8+,CD4+ T-cells, Treg in tumor

[97–99]

Glycoconjugated chlorin (G-chlorin)

3rd

CT26, murine colon adenocarcinoma

CRT, HMGB1; promoting macrophage phagocytosis; protect mice against tumour in prophylactic vaccination model

[100, 101]

 

Photosensitizer (PS)

PS generation and type of additional structure

Tumor cell line

Intracellular localization of PS and possible cell death modality

ICD markers and signs of adaptive immunity activation

References

Photodithazine

2nd

GL261, murine glioma; MCA205, murine fibrosarcoma

ER and GA

CRT, ATP, HMGB1; induced DC maturation; IL-6

[34]

2-[1-Hexyloxyethyl]-2-devinyl pyropheophorbide (HPPH)

3rd, pRNVs/HPPH/IND

B16F10, murine melanoma; MC38, murine colorectal carcinoma

ER

CRT; overproduced ROS; CD8+, CD4+ T-cells in tumor; apoptosis

[102]

Ru PS class

2nd

B16F10, murine melanoma

Overproduced ROS; HMGB1, АТP; IL6, CXCL10, TLR3

[103]

DPA-TPE-SCP

3rd, peptide dots

4T1, murine mammary adenocarcinoma

Overproduced ROS, CRT HMGB1, HSP70, ATP; DC (CD80+ CD86+, CD11c+) in lymph nodes; CD8+ T cells and Treg (Foxp3+, CD4+, CD3+) in tumor

[104]

Cyanoarylporphyrazines

2nd

MCA205, murine fibrosarcoma

GA and lysosomes (pz I), ER and lysosomes (pz III)

HMGB-1, АТP; induced DC maturation

[105, 106]

GL261, murine glioma

HMGB-1, АТP; smaller tumor size in experimental groups

[105–108]

Methylene blue

1st

H1299 (P53 null) and A549 (P53 wt), human non-small cell lung carcinoma;

NSCLC, tumour organoids

CRT, MHCII, ICAM-1; CD8+ cells cytotoxic activity

[109, 110]

Rose Bengal acetate

2nd

HeLa, human cervical cancer

Perinuclear

CRT, HSP70, HSP90, ATP; TNF-a,

TGF-b and IL-10; macrophage recruiting;

[111, 112]

4,4',4'',4'''-(porphyrin-5,10,15,20-tetrayl)tetrakis(N-(2-((4-methylphenyl)sulfonamido)ethyl)benzamide)

3rd, PEG-s-s-1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[amino(polyethylene glycol)-2000] (Ds-sP) NPs

4T1, murine mammary adenocarcinoma

ER

Overproduced ROS, CRT, HMGB-1; DC (CD80+CD86+) maturation; TNF-α, IL-12p40; CD8+ T cells in tumor

[113]

2-(5-(4-(bis(phenyl)amino)benzylidene)-4-oxo-3-phenylthiazolidin-2- ylidene)malononitrile (TPA-DCR)

3rd, TPA-DCR NPs

4T1, murine mammary adenocarcinoma

Overproduced ROS, CRT HMGB1, HSP70, ATP; TNF-α; increase of M1 macrophage proportion; CD8+, CD4+ T-cells in tumor

[114]

5-aminolevulinic acid (5-ALA)

2nd

MC38, murine colorectal carcinoma; HCT116, human colorectal carcinoma; B16F10, murine melanoma

CRT, HSP70, HSP90, HMGB1, ATP; induced DC maturation; CD8+ T-cells in spleen; IFN-γ, TNF-α; DCs and CD8+ T-cells in tumor

[115, 116]

Hypericin

2nd

AY27, rat transitional cell carcinoma of the urinary bladder; CT26, murine colon adenocarcinoma

HMGB1, ATP, CRT

[117, 118]

T24, transitional cell carcinoma; CT26, murine colon adenocarcinoma; MEF, murine fibroblasts

HSP70, CRT

[117, 119]

GL261, murine glioma

CRT, HSP70, HSP90, HMGB1, ATP; CD3+ T-cells, CD4+ T-cells, CD8+ T-cells, TH1 cells, CTLs, Th17, Tregs; stimulated DC maturation

[117, 120]

8-Methoxypsoralen (8-MOP)

2nd

B16, murine melanoma; YUCOT, human melanoma; MC38, murine colorectal carcinoma

HMGB1, ATP, IFN I; tumor engulfment by monocytes

[121]

Photosensitizer (PS)

PS generation and type of additional structure

Tumor cell line

Intracellular localization of PS and possible cell death modality

ICD markers and signs of adaptive immunity activation

References

Polyethylene glycol-poly(methyl methyacrylate-co-2-amino ethyl methacrylate (thiol/amine))-poly 2-(dimethylamino)ethyl methacrylate (PEG-P(MMA-co-AEMA (SH/NH2)-PDMA)

2nd

MC38, murine colorectal carcinoma

Overproduced ROS; CRT

HMGB1; induced DC maturation; IL-6, IL-12, TNF-α

[122]

×

About the authors

Ekaterina E. Sleptsova

National Research Lobachevsky State University of Nizhny Novgorod

Email: ees222@list.ru
ORCID iD: 0009-0002-3825-1270
Russian Federation, Nizhny Novgorod

Tikhon S. Redkin

National Research Lobachevsky State University of Nizhny Novgorod

Email: big.t.nsdav@outlook.com
ORCID iD: 0009-0001-0835-5143
Russian Federation, Nizhny Novgorod

Mariia O. Saviuk

National Research Lobachevsky State University of Nizhny Novgorod; Ghent University

Email: mariia.saviuk@ugent.be
ORCID iD: 0000-0002-1333-470X
Russian Federation, Nizhny Novgorod; Ghent, Belgium

Victoria D. Turubanova

National Research Lobachevsky State University of Nizhny Novgorod

Author for correspondence.
Email: vikaturu@mail.ru
ORCID iD: 0000-0002-4648-0738
SPIN-code: 8262-6560

PhD, Cand. Sci. (Biology)

Russian Federation, Nizhny Novgorod

References

  1. Boulikas T, Vougiouka M. Recent clinical trials using cisplatin, carboplatin and their combination chemotherapy drugs (review). Oncol Rep. 2004;11(3):559–595.
  2. Annovazzi L, Mellai M, Schiffer D. Chemotherapeutic drugs: DNA damage and repair in glioblastoma. Cancers (Basel). 2017;9(6):57. doi: 10.3390/cancers9060057
  3. Silasi DA, Alvero AB, Rutherford TJ, et al. Phenoxodiol: pharmacology and clinical experience in cancer monotherapy and in combination with chemotherapeutic drugs. Expert Opin Pharmacother. 2009;10(6):1059–1067. Corrected and republished from: Expert Opin Pharmacother. 2009;10(8):1387. doi: 10.1517/14656560902837980
  4. Peng M, Darko KO, Tao T, et al. Combination of metformin with chemotherapeutic drugs via different molecular mechanisms. Cancer Treat Rev. 2017;54:24–33. doi: 10.1016/j.ctrv.2017.01.005
  5. Romiti A, Cox MC, Sarcina I, et al. Metronomic chemotherapy for cancer treatment: a decade of clinical studies. Cancer Chemother Pharmacol. 2013;72(1):13–33. doi: 10.1007/s00280-013-2125-x
  6. Hanahan D. Hallmarks of cancer: new dimensions. Cancer Discov. 2022;12(1):31–46. doi: 10.1158/2159-8290.CD-21-1059
  7. Mahoney DJ, Stojdl DF. Molecular pathways: multimodal cancer-killing mechanisms employed by oncolytic vesiculoviruses. Clin Cancer Res. 2013;19(4):758–763. doi: 10.1158/1078-0432.CCR-11-3149
  8. Ning Y, Shen K, Wu Q, et al. Tumor exosomes block dendritic cells maturation to decrease the T cell immune response. Immunol Lett. 2018;199:36–43. doi: 10.1016/j.imlet.2018.05.002
  9. Galluzzi L, Kepp O, Hett E, et al. Immunogenic cell death in cancer: concept and therapeutic implications. J Transl Med. 2023;21(1):162. doi: 10.1186/s12967-023-04017-6
  10. Placek K, Schultze JL, Aschenbrenner AC. Epigenetic reprogramming of immune cells in injury, repair, and resolution. J Clin Invest. 2019;129(8):2994–3005. doi: 10.1172/JCI124619
  11. Guo Y, Xie YQ, Gao M, et al. Metabolic reprogramming of terminally exhausted CD8+ T cells by IL-10 enhances anti-tumor immunity. Nat Immunol. 2021;22(6):746–756. doi: 10.1038/s41590-021-00940-2
  12. Casares N, Pequignot MO, Tesniere A, et al. Caspase-dependent immunogenicity of doxorubicin-induced tumor cell death. J Exp Med. 2005;202(12):1691–1701. doi: 10.1084/jem.20050915
  13. Krysko DV, Garg AD, Kaczmarek A, et al. Immunogenic cell death and DAMPs in cancer therapy. Nat Rev Cancer. 2012;12(12):860–875. doi: 10.1038/nrc3380
  14. Golden EB, Apetoh L. Radiotherapy and immunogenic cell death. Semin Radiat Oncol. 2015;25(1):11–17. doi: 10.1016/j.semradonc.2014.07.005
  15. Fucikova J, Kepp O, Kasikova L, et al. Detection of immunogenic cell death and its relevance for cancer therapy. Cell Death Dis. 2020;11(11):1013. doi: 10.1038/s41419-020-03221-2
  16. Garg AD, Dudek-Peric AM, Romano E, Agostinis P. Immunogenic cell death. Int J Dev Biol. 2015;59(1-3):131–140. doi: 10.1387/ijdb.150061pa
  17. Galluzzi L, Vitale I, Warren S, et al. Consensus guidelines for the definition, detection and interpretation of immunogenic cell death. J Immunother Cancer. 2020;8(1):e000337. Corrected and republished from: J Immunother Cancer. 2020;8(1). doi: 10.1136/jitc-2019-000337
  18. Inoue H, Tani K. Multimodal immunogenic cancer cell death as a consequence of anticancer cytotoxic treatments. Cell Death Differ. 2014;21(1):39–49. doi: 10.1038/cdd.2013.84
  19. Sen S, Won M, Levine MS, et al. Metal-based anticancer agents as immunogenic cell death inducers: the past, present, and future. Chem Soc Rev. 2022;51(4):1212–1233. doi: 10.1039/d1cs00417d
  20. Mardi A, Shirokova AV, Mohammed RN, et al. Biological causes of immunogenic cancer cell death (ICD) and anti-tumor therapy; combination of oncolytic virus-based immunotherapy and CAR T-cell therapy for ICD induction. Cancer Cell Int. 2022;22(1):168. doi: 10.1186/s12935-022-02585-z
  21. Zhang Z, Yue Y, Li Q, et al. Design of calixarene-based ICD inducer for efficient cancer immunotherapy. Adv Funct Mater. 2023;33(16). doi: 10.1002/adfm.202213967
  22. Zhang Y, Sultonova RD, You SH, et al. The anticancer effect of PASylated calreticulin-targeting L-ASNase in solid tumor bearing mice with immunogenic cell death-inducing chemotherapy. Biochem Pharmacol. 2023;210:115473. doi: 10.1016/j.bcp.2023.115473
  23. Zhang M, Jin X, Gao M, et al. A self-reporting fluorescent salicylaldehyde–chlorambucil conjugate as a type-II ICD inducer for cancer vaccines. Adv Mater. 2022;34(36):e2205701. doi: 10.1002/adma.202205701
  24. Song N, Park M, Kim N, et al. Tumor-targeting oxidative stress nanoamplifiers as anticancer nanomedicine with immunostimulating activity. Biomater Sci. 2022;10(21):6160–6171. doi: 10.1039/d2bm00601d
  25. Song H, Cai Z, Li J, et al. Light triggered release of a triple action porphyrin-cisplatin conjugate evokes stronger immunogenic cell death for chemotherapy, photodynamic therapy and cancer immunotherapy. J Nanobiotechnology. 2022;20(1):329. doi: 10.1186/s12951-022-01531-5
  26. Wang W, Yang X, Li C, et al. Immunogenic cell death (ICD) of murine H22 cells induced by lentinan. Nutr Cancer. 2022;74(2):640–649. doi: 10.1080/01635581.2021.1897632
  27. Lv S, Sylvestre M, Song K, Pun SH. Development of D-melittin polymeric nanoparticles for anti-cancer treatment. Biomaterials. 2021;277:121076. doi: 10.1016/j.biomaterials.2021.121076
  28. Olek M, Machorowska-Pieniążek A, Czuba ZP, et al. Effect of hypericin-mediated photodynamic therapy on the secretion of soluble TNF receptors by oral cancer cells. Pharmaceutics. 2023;15(4):1279. doi: 10.3390/pharmaceutics15041279
  29. Kumar Das D. Synthetic hypericin photodynamic therapy for cutaneous T-Cell lymphoma. Oncology Times. 2022;44(20):1, 7. doi: 10.1097/01.COT.0000892568.02297.9b
  30. Dong X, Zeng Y, Zhang Z, et al. Hypericin-mediated photodynamic therapy for the treatment of cancer: a review. J Pharm Pharmacol. 2021;73(4):425–436. doi: 10.1093/jpp/rgaa018
  31. Kataoka H, Nishie H, Tanaka M, et al. Potential of photodynamic therapy based on sugar-conjugated photosensitizers. J Clin Med. 2021;10(4):841. doi: 10.3390/jcm10040841
  32. Lu Y, Sun W, Du J, et al. Immuno-photodynamic therapy (IPDT): organic photosensitizers and their application in cancer ablation. JACS Au. 2023;3(3):682–699. doi: 10.1021/jacsau.2c00591
  33. Lin X, Mao D, Bai R. Comparison of three commercial photosensitizers for efficiency of inducing immunogenic cell death in anti-tumor immunotherapy. Nan Fang Yi Ke Da Xue Xue Bao. 2022;42(12):1791–1798. (In Chinese). doi: 10.12122/j.issn.1673-4254.2022.12.06
  34. Turubanova VD, Balalaeva IV, Mishchenko TA, et al. Immunogenic cell death induced by a new photodynamic therapy based on photosens and photodithazine. J Immunother Cancer. 2019;7(1). doi: 10.1186/s40425-019-0826-3
  35. Morais JAV, Almeida LR, Rodrigues MC, et al. The induction of immunogenic cell death by photodynamic therapy in B16F10 cells in vitro is effected by the concentration of the photosensitizer. Photodiagnosis Photodyn Ther. 2021;35:102392. doi: 10.1016/j.pdpdt.2021.102392
  36. Alzeibak R, Mishchenko TA, Shilyagina NY, et al. Targeting immunogenic cancer cell death by photodynamic therapy: past, present and future. J Immunother Cancer. 2021;9(1):e001926. Corrected and republished from: J Immunother Cancer. 2021;9(10). doi: 10.1136/jitc-2020-001926
  37. Garg AD, Krysko DV, Verfaillie T, et al. A novel pathway combining calreticulin exposure and ATP secretion in immunogenic cancer cell death. EMBO J. 2012;31(5):1062–1079. doi: 10.1038/emboj.2011.497
  38. Vedunova M, Turubanova V, Vershinina O, et al. DC vaccines loaded with glioma cells killed by photodynamic therapy induce Th17 anti-tumor immunity and provide a four-gene signature for glioma prognosis. Cell Death Dis. 2022;13(12):1062. doi: 10.1038/s41419-022-05514-0
  39. Allison RR, Downie GH, Cuenca R, et al. Photosensitizers in clinical PDT. Photodiagnosis Photodyn Ther. 2004;1(1):27–42. doi: 10.1016/S1572-1000(04)00007-9
  40. Escudero A, Carrillo-Carrión C, Castillejos MC, et al. Photodynamic therapy: photosensitizers and nanostructures. Mater Chem Front. 2021;5(10):3788–3812. doi: 10.1039/D0QM00922A
  41. Spikes JD. New trends in photobiology. J Photochem Photobiol B. 1990;6(3):259–274. doi: 10.1016/1011-1344(90)85096-f
  42. Dabrowski JM, Arnaut LG. Photodynamic therapy (PDT) of cancer: from local to systemic treatment. Photochem Photobiol Sci. 2015;14(10):1765–1780. doi: 10.1039/c5pp00132c
  43. Park J, Lee YK, Park IK, Hwang SR. Current limitations and recent progress in nanomedicine for clinically available photodynamic therapy. Biomedicines. 2021;9(1):85. doi: 10.3390/biomedicines9010085
  44. Josefsen LB, Boyle RW. Photodynamic therapy: novel third-generation photosensitizers one step closer? Br J Pharmacol. 2008;154(1):1–3. doi: 10.1038/bjp.2008.98
  45. Wall MH Jr, Basu P, Buranda T, et al. Photoinduced electron transfer in covalently linked oxomolybdenum (V) porphyrin systems. Inorg Chem. 1997;36(25):5676–5677. doi: 10.1021/ic9705711
  46. Shakhova M, Loginova D, Meller A, et al. Photodynamic therapy with chlorin-based photosensitizer at 405 nm: numerical, morphological, and clinical study. J Biomed Opt. 2018;23(9):1–9. doi: 10.1117/1.JBO.23.9.091412
  47. Hak A, Ali MS, Sankaranarayanan SA, et al. Chlorin e6: a promising photosensitizer in photo-based cancer nanomedicine. ACS Appl Bio Mater. 2023;6(2):349–364. doi: 10.1021/acsabm.2c00891
  48. Li X, Zheng BD, Peng XH, et al. Phthalocyanines as medicinal photosensitizers: developments in the last five years. Coord Chem Rev. 2019;379:147–160. doi: 10.1016/j.ccr.2017.08.003
  49. Roguin LP, Chiarante N, García Vior MC, Marino J. Zinc(II) phthalocyanines as photosensitizers for antitumor photodynamic therapy. Int J Biochem Cell Biol. 2019;114:105575. doi: 10.1016/j.biocel.2019.105575
  50. Zhou L, Zou M, Xu Y, et al. Nano drug delivery system for tumor immunotherapy: next-generation therapeutics. Front Oncol. 2022;12:864301. doi: 10.3389/fonc.2022.864301
  51. Baskaran R, Lee J, Yang SG. Clinical development of photodynamic agents and therapeutic applications. Biomater Res. 2018;22(1):25. doi: 10.1186/s40824-018-0140-z
  52. Foley P. Clinical efficacy of methyl aminolevulinate (Metvix®) photodynamic therapy. J Dermatolog Treat. 2003;14 Suppl. 3:15–22. doi: 10.1080/jdt.14.s3.15.22
  53. Yoon I, Li JZ, Shim YK. Advance in photosensitizers and light delivery for photodynamic therapy. Clin Endosc. 2013;46(1):7–23. doi: 10.5946/ce.2013.46.1.7
  54. O’Connor AE, Gallagher WM, Byrne AT. Porphyrin and nonporphyrin photosensitizers in oncology: preclinical and clinical advances in photodynamic therapy. Photochem Photobiol. 2009;85(5):1053–1074. doi: 10.1111/j.1751-1097.2009.00585.x
  55. Mfouo-Tynga IS, Dias LD, Inada NM, Kurachi C. Features of third generation photosensitizers used in anticancer photodynamic therapy: review. Photodiagnosis Photodyn Ther. 2021;34:102091. doi: 10.1016/j.pdpdt.2020.102091
  56. Lim S, Park J, Shim MK, et al. Recent advances and challenges of repurposing nanoparticle-based drug delivery systems to enhance cancer immunotherapy. Theranostics. 2019;9(25):7906–7923. doi: 10.7150/thno.38425
  57. Mitchell MJ, Billingsley MM, Haley RM, et al. Engineering precision nanoparticles for drug delivery. Nat Rev Drug Discov. 2021;20(2):101–124. doi: 10.1038/s41573-020-0090-8
  58. Wu P, Han J, Gong Y, et al. Nanoparticle-based drug delivery systems targeting tumor microenvironment for cancer immunotherapy resistance: current advances and applications. Pharmaceutics. 2022;14(10):1990. doi: 10.3390/pharmaceutics14101990
  59. Yi Y, Yu M, Li W, et al. Vaccine-like nanomedicine for cancer immunotherapy. J Control Release. 2023;355:760–778. doi: 10.1016/j.jconrel.2023.02.015
  60. Costa RA, Farah ME, Freymüller E, et al. Choriocapillaris photodynamic therapy using indocyanine green. Am J Ophthalmol. 2001;132(4):557–565. doi: 10.1016/s0002-9394(01)01138-2
  61. Onda N, Kimura M, Yoshida T, Shibutani M. Preferential tumor cellular uptake and retention of indocyanine green for in vivo tumor imaging. Int J Cancer. 2016;139(3):673–682. doi: 10.1002/ijc.30102
  62. Wang Z, Ju Y, Ali Z, et al. Near-infrared light and tumor microenvironment dual responsive size-switchable nanocapsules for multimodal tumor theranostics. Nat Commun. 2019;10(1):4418. doi: 10.1038/s41467-019-12142-4
  63. Zhao H, Xu J, Feng C, et al. Tailoring aggregation extent of photosensitizers to boost phototherapy potency for eliciting systemic antitumor immunity. Adv Mater. 2022;34(8):e2106390. doi: 10.1002/adma.202106390
  64. Heshmati Aghda N, Torres Hurtado S, Abdulsahib SM, et al. Dual photothermal/chemotherapy of melanoma cells with albumin nanoparticles carrying indocyanine green and doxorubicin leads to immunogenic cell death. Macromol Biosci. 2022;22(2):e2100353. doi: 10.1002/mabi.202100353
  65. Gao H, Bai Y, Chen L, et al. Self-assembly nanoparticles for overcoming multidrug resistance and imaging-guided chemo-photothermal synergistic cancer therapy. Int J Nanomedicine. 2020;15:809–819. doi: 10.2147/IJN.S232449
  66. Xie W, Zhu S, Yang B, et al. The destruction of laser-induced phase-transition nanoparticles triggered by low-intensity ultrasound: an innovative modality to enhance the immunological treatment of ovarian cancer cells. Int J Nanomedicine. 2019;14:9377–9393. doi: 10.2147/IJN.S208404
  67. Doix B, Trempolec N, Riant O, Feron O. Low photosensitizer dose and early radiotherapy enhance antitumor immune response of photodynamic therapy-based dendritic cell vaccination. Front Oncol. 2019;9:811. doi: 10.3389/fonc.2019.00811
  68. Pinto A, Mace Y, Drouet F, et al. A new ER-specific photosensitizer unravels 1O2-driven protein oxidation and inhibition of deubiquitinases as a generic mechanism for cancer PDT. Oncogene. 2016;35(30):3976–3985. doi: 10.1038/onc.2015.474
  69. Yu H, Chen B, Huang H, et al. AIE-active photosensitizers: manipulation of reactive oxygen species generation and applications in photodynamic therapy. Biosensors (Basel). 2022;12(5):348. doi: 10.3390/bios12050348
  70. Huang C, Zhang T, Li Y, et al. Type-I AIE photosensitizer triggered cascade catalysis system for tumor targeted therapy and postoperative recurrence suppression. Chemical Engineering Journal. 2022;446(3):136381. doi: 10.1016/j.cej.2022.136381
  71. Qi J, Jia S, Kang X, et al. Semiconducting polymer nanoparticles with surface-mimicking protein secondary structure as lysosome-targeting chimaeras for self-synergistic cancer immunotherapy. Adv Mater. 2022;34(31):e2203309. doi: 10.1002/adma.202203309
  72. Chen D, Yu Q, Huang X, et al. A highly-efficient type i photosensitizer with robust vascular-disruption activity for hypoxic-and-metastatic tumor specific photodynamic therapy. Small. 2020;16(23):e2001059. Corrected and republished from: Small. 2023;19(32):e2305388. doi: 10.1002/smll.202001059
  73. Luo S, Luo X, Wang X, et al. Tailoring multifunctional small molecular photosensitizers to in vivo self-assemble with albumin to boost tumor-preferential accumulation, NIR imaging, and photodynamic/photothermal/immunotherapy. Small. 2022;18(27):e2201298. doi: 10.1002/smll.202201298
  74. Ling YY, Wang WJ, Hao L, et al. Self-amplifying iridium(III) photosensitizer for ferroptosis-mediated immunotherapy against transferrin receptor-overexpressing cancer. Small. 2022;18(49):e2203659. doi: 10.1002/smll.202203659
  75. Lobo ACS, Gomes-da-Silva LC, Rodrigues-Santos P, et al. Immune responses after vascular photodynamic therapy with redaporfin. J Clin Med. 2019;9(1):104. doi: 10.3390/jcm9010104
  76. Gomes-da-Silva LC, Zhao L, Bezu L, et al. Photodynamic therapy with redaporfin targets the endoplasmic reticulum and Golgi apparatus. EMBO J. 2018;37(13):e98354. doi: 10.15252/embj.201798354
  77. Calori IR, Tedesco AC. Lipid vesicles loading aluminum phthalocyanine chloride: formulation properties and disaggregation upon intracellular delivery. J Photochem Photobiol B. 2016;160:240–247. doi: 10.1016/j.jphotobiol.2016.03.050
  78. Chilakamarthi U, Giribabu L. Photodynamic therapy: past, present and future. The Chemical Record. 2017;17(8):775–802. doi: 10.1002/tcr.201600121
  79. Madsen SJ, Christie C, Huynh K, et al. Limiting glioma development by photodynamic therapy-generated macrophage vaccine and allo-stimulation: an in vivo histological study in rats. J Biomed Opt. 2018;23(2):1–7. doi: 10.1117/1.JBO.23.2.028001
  80. Zheng BY, Li SZ, Zheng BD, et al. Phthalocyanine-based photosensitizers combined with anti-PD-L1 for highly efficient photodynamic immunotherapy. Dyes and Pigments. 2021;185(5):108907. doi: 10.1016/j.dyepig.2020.108907
  81. Chen J, Wang Y, Fang Y, et al. Improved photodynamic anticancer activity and mechanisms of a promising zinc(II) phthalocyanine-quinoline conjugate photosensitizer in vitro and in vivo. Biomed Opt Express. 2020;11(7):3900–3912. doi: 10.1364/BOE.394186
  82. Wang H, Jing G, Niu J, et al. A mitochondria-anchored supramolecular photosensitizer as a pyroptosis inducer for potent photodynamic therapy and enhanced antitumor immunity. J Nanobiotechnology. 2022;20(1):513. doi: 10.1186/s12951-022-01719-9
  83. Zhang Y, Cheung YK, Ng DKP, Fong WP. Immunogenic necroptosis in the anti-tumor photodynamic action of BAM-SiPc, a silicon(IV) phthalocyanine-based photosensitizer. Cancer Immunol Immunother. 2021;70(2):485–495. doi: 10.1007/s00262-020-02700-x
  84. Baron ED, Malbasa CL, Santo-Domingo D, et al. Silicon phthalocyanine (pc 4) photodynamic therapy is a safe modality for cutaneous neoplasms: results of a phase 1 clinical trial. Lasers Surg Med. 2010;42(10):728–735. doi: 10.1002/lsm.20984
  85. Rodrigues MC, de Sousa Júnior WT, Mundim T, et al. Induction of Immunogenic cell death by photodynamic therapy mediated by aluminum-phthalocyanine in nanoemulsion. Pharmaceutics. 2022;14(1):196. doi: 10.3390/pharmaceutics14010196
  86. Li Y, Yu Y, Kang L, Lu Y. Effects of chlorin e6-mediated photodynamic therapy on human colon cancer SW480 cells. Int J Clin Exp Med. 2014;7(12):4867–4876.
  87. Kimura Y, Aoki H, Soyama T, et al. Photodynamic therapy using mannose-conjugated chlorin e6 increases cell surface calreticulin in cancer cells and promotes macrophage phagocytosis. Med Oncol. 2022;39(5):82. doi: 10.1007/s12032-022-01674-3
  88. Korbelik M. Role of cell stress signaling networks in cancer cell death and antitumor immune response following proteotoxic injury inflicted by photodynamic therapy. Lasers Surg Med. 2018;50(5):491–498. doi: 10.1002/lsm.22810
  89. Zhang Z, Sang W, Xie L, et al. Polyphenol-based nanomedicine evokes immune activation for combination cancer treatment. Angew Chem Int Ed Engl. 2021;60(4):1967–1975. doi: 10.1002/anie.202013406
  90. Liu H, Hu Y, Sun Y, et al. Co-delivery of bee venom melittin and a photosensitizer with an organic–inorganic hybrid nanocarrier for photodynamic therapy and immunotherapy. ACS Nano. 2019;13(11):12638–12652. doi: 10.1021/acsnano.9b04181
  91. Shixiang Y, Xi S, Junliang L, et al. Antitumor efficacy of a photodynamic therapy-generated dendritic cell glioma vaccine. Med Oncol. 2011;28 Suppl. 1:S453-S461. doi: 10.1007/s12032-010-9713-y
  92. Yow CM, Chen JY, Mak NK, et al. Cellular uptake, subcellular localization and photodamaging effect of Temoporfin (mTHPC) in nasopharyngeal carcinoma cells: comparison with hematoporphyrin derivative. Cancer Lett. 2000;157(2):123–131. doi: 10.1016/s0304-3835(00)00453-5
  93. Yue W, Chen L, Yu L, et al. Checkpoint blockade and nanosonosensitizer-augmented noninvasive sonodynamic therapy combination reduces tumour growth and metastases in mice. Nat Commun. 2019;10(1):2025. doi: 10.1038/s41467-019-09760-3
  94. Ogawa M, Tomita Y, Nakamura Y, et al. Immunogenic cancer cell death selectively induced by near infrared photoimmunotherapy initiates host tumor immunity. Oncotarget. 2017;8(6):10425–10436. doi: 10.18632/oncotarget.14425
  95. Wakiyama H, Furusawa A, Okada R, et al. Increased immunogenicity of a minimally immunogenic tumor after cancer-targeting near infrared photoimmunotherapy. Cancers (Basel). 2020;12(12):3747. doi: 10.3390/cancers12123747
  96. Nakajima K, Ogawa M. Phototoxicity in near-infrared photoimmunotherapy is influenced by the subcellular localization of antibody-IR700. Photodiagnosis Photodyn Ther. 2020;31:101926. doi: 10.1016/j.pdpdt.2020.101926
  97. Maruoka Y, Furusawa A, Okada R, et al. Near-infrared photoimmunotherapy combined with CTLA4 checkpoint blockade in syngeneic mouse cancer models. Vaccines (Basel). 2020;8(3):528. doi: 10.3390/vaccines8030528
  98. Yamashita S, Kojima M, Onda N, Shibutani M. In vitro comparative study of near-infrared photoimmunotherapy and photodynamic therapy. Cancers (Basel). 2023;15(13):3400. doi: 10.3390/cancers15133400
  99. Nagaya T, Friedman J, Maruoka Y, et al. Host immunity following near-infrared photoimmunotherapy is enhanced with PD-1 checkpoint blockade to eradicate established antigenic tumors. Cancer Immunol Res. 2019;7(3):401–413. doi: 10.1158/2326-6066.CIR-18-0546
  100. Kataoka H, Nishie H, Hayashi N, et al. New photodynamic therapy with next-generation photosensitizers. Ann Transl Med. 2017;5(8):183. doi: 10.21037/atm.2017.03.59
  101. Tanaka M, Kataoka H, Yano S, et al. Immunogenic cell death due to a new photodynamic therapy (PDT) with glycoconjugated chlorin (G-chlorin). Oncotarget. 2016;7(30):47242–47251. doi: 10.18632/oncotarget.9725
  102. Yang W, Zhang F, Deng H, et al. Smart nanovesicle-mediated immunogenic cell death through tumor microenvironment modulation for effective photodynamic immunotherapy. ACS Nano. 2020;14(1):620–631. doi: 10.1021/acsnano.9b07212
  103. Konda P, Lifshits LM, Roque JA 3rd, et al. Discovery of immunogenic cell death-inducing ruthenium-based photosensitizers for anticancer photodynamic therapy. Oncoimmunology. 2020;10(1):1863626. doi: 10.1080/2162402X.2020.1863626
  104. Jia S, Gao Z, Wu Z, et al. Sonosensitized aggregation-induced emission dots with capacities of immunogenic cell death induction and multivalent blocking of programmed cell death-ligand 1 for amplified antitumor immunotherapy. CCS Chemistry. 2022;4(2):501–514. doi: 10.31635/ccschem.021.202101458
  105. Turubanova VD, Mishchenko TA, Balalaeva IV, et al. Novel porphyrazine-based photodynamic anti-cancer therapy induces immunogenic cell death. Sci Rep. 2021;11(1):7205. doi: 10.1038/s41598-021-86354-4
  106. Shestakova LN, Lyubova TS, Lermontova SA, et al. Comparative analysis of Tetra(2-naphthyl)tetracyano-porphyrazine and its iron complex as photosensitizers for anticancer photodynamic therapy. Pharmaceutics. 2022;14(12):2655. doi: 10.3390/pharmaceutics14122655
  107. Redkin TS, Sleptsova EE, Turubanova VD, et al. Dendritic cells pulsed with tumor lysates induced by tetracyanotetra(aryl)porphyrazines-based photodynamic therapy effectively trigger anti-tumor immunity in an orthotopic mouse glioma model. Pharmaceutics. 2023;15(10):2430. doi: 10.3390/pharmaceutics15102430
  108. Balalaeva IV, Mishchenko TA, Turubanova VD, et al. Cyanoarylporphyrazines with high viscosity sensitivity: a step towards dosimetry-assisted photodynamic cancer treatment. Molecules. 2021;26(19):5816. doi: 10.3390/molecules26195816
  109. Valančiūtė A, Mathieson L, O’Connor RA, et al. Phototherapeutic induction of immunogenic cell death and CD8+ T cell-granzyme b mediated cytolysis in human lung cancer cells and organoids. Cancers (Basel). 2022;14(17):4119. doi: 10.3390/cancers14174119
  110. Lim DJ. Methylene blue-based nano and microparticles: fabrication and applications in photodynamic therapy. Polymers (Basel). 2021;13(22):3955. doi: 10.3390/polym13223955
  111. Panzarini E, Inguscio V, Fimia GM, Dini L. Rose Bengal acetate photodynamic therapy (RBAc-PDT) induces exposure and release of damage-associated molecular patterns (DAMPs) in human HeLa cells. PLoS One. 2014;9(8):e105778. doi: 10.1371/journal.pone.0105778
  112. Alibasha A, Ghosh M. Recent developments of porphyrin photosensitizers in photodynamic therapy. ChemRxiv. 2023.
  113. Deng H, Zhou Z, Yang W, et al. Endoplasmic reticulum targeting to amplify immunogenic cell death for cancer immunotherapy. Nano Lett. 2020;20(3):1928–1933. doi: 10.1021/acs.nanolett.9b05210
  114. Yang G, Lu SB, Li C, et al. Type I macrophage activator photosensitizer against hypoxic tumors. Chem Sci. 2021;12(44):14773–14780. doi: 10.1039/d1sc04124j
  115. Sun Z, Zhao M, Wang W, et al. 5-ALA mediated photodynamic therapy with combined treatment improves anti-tumor efficacy of immunotherapy through boosting immunogenic cell death. Cancer Lett. 2023;554:216032. doi: 10.1016/j.canlet.2022.216032
  116. Mahmoudi K, Garvey KL, Bouras A, et al. 5-aminolevulinic acid photodynamic therapy for the treatment of high-grade gliomas. J Neurooncol. 2019;141(3):595–607. doi: 10.1007/s11060-019-03103-4
  117. Skalkos D, Gioti E, Stalikas CD, et al. Photophysical properties of Hypericum perforatum L. extracts — novel photosensitizers for PDT. J Photochem Photobiol B. 2006;82(2):146–151. doi: 10.1016/j.jphotobiol.2005.11.001
  118. Garg AD, Elsen S, Krysko DV, et al. Resistance to anticancer vaccination effect is controlled by a cancer cell-autonomous phenotype that disrupts immunogenic phagocytic removal. Oncotarget. 2015;6(29):26841–26860. Corrected and republished from: Oncotarget. 2018;9(49):29284. doi: 10.18632/oncotarget.4754
  119. Garg AD, Krysko DV, Vandenabeele P, Agostinis P. Hypericin-based photodynamic therapy induces surface exposure of damage-associated molecular patterns like HSP70 and calreticulin. Cancer Immunol Immunother. 2012;61(2):215–221. Corrected and republished from: Cancer Immunol Immunother. 2018;67(7):1179–1180. doi: 10.1007/s00262-011-1184-2
  120. Garg AD, Vandenberk L, Koks C, et al. Dendritic cell vaccines based on immunogenic cell death elicit danger signals and T cell-driven rejection of high-grade glioma. Sci Transl Med. 2016;8(328):328ra27. doi: 10.1126/scitranslmed.aae0105
  121. Tatsuno K, Yamazaki T, Hanlon D, et al. Extracorporeal photochemotherapy induces bona fide immunogenic cell death. Cell Death Dis. 2019;10(8):578. doi: 10.1038/s41419-019-1819-3
  122. Yang W, Zhu G, Wang S, et al. In Situ dendritic cell vaccine for effective cancer immunotherapy. ACS Nano. 2019;13(3):3083–3094. doi: 10.1021/acsnano.8b08346
  123. Calixto G, Bernegossi J, de Freitas L, et al. Nanotechnology-based drug delivery systems for photodynamic therapy of cancer: a review. Molecules. 2016;21(3):342. doi: 10.3390/molecules21030342
  124. Kudinova NV, Berezov TT. Photodynamic therapy of cancer: search for ideal photosensitizer. Biochem Mosc Suppl B Biomed Chem. 2010;4(1):95–103. doi: 10.1134/S1990750810010129
  125. MacRobert AJ, Bown SG, Phillips D. What are the ideal photoproperties for a sensitizer? Ciba Found Symp. 1989;146:4–16. doi: 10.1002/9780470513842.ch2
  126. Simões JCS, Sarpaki S, Papadimitroulas P, et al. Conjugated photosensitizers for imaging and PDT in cancer research. J Med Chem. 2020;63(23):14119–14150. doi: 10.1021/acs.jmedchem.0c00047
  127. Uzdensky AB. Cellular and molecular mechanisms of photodynamic therapy. Saint Petersburg: Nauka; 2010. (In Russ). EDN: QLXTXH
  128. Broekgaarden M, Weijer R, van Gulik TM, et al. Tumor cell survival pathways activated by photodynamic therapy: a molecular basis for pharmacological inhibition strategies. Cancer Metastasis Rev. 2015;34(4):643–690. doi: 10.1007/s10555-015-9588-7
  129. Sobhani N, Samadani AA. Implications of photodynamic cancer therapy: an overview of PDT mechanisms basically and practically. J Egypt Natl Canc Inst. 2021;33(1):34. doi: 10.1186/s43046-021-00093-1
  130. Dougherty TJ, Gomer CJ, Henderson BW, et al. Photodynamic therapy. J Natl Cancer Inst. 1998;90(12):889–905. doi: 10.1093/jnci/90.12.889
  131. Kessel D. Subcellular targets for photodynamic therapy: implications for initiation of apoptosis and autophagy. J Natl Compr Canc Netw. 2012;10 Suppl. 2:S56–S59. doi: 10.6004/jnccn.2012.0177
  132. Hsieh YJ, Wu CC, Chang CJ, Yu JS. Subcellular localization of Photofrin® determines the death phenotype of human epidermoid carcinoma A431 cells triggered by photodynamic therapy: when plasma membranes are the main targets. J Cell Physiol. 2003;194(3):363–375. doi: 10.1002/jcp.10273
  133. Iulianna T, Kuldeep N, Eric F. The Achilles’ heel of cancer: targeting tumors via lysosome-induced immunogenic cell death. Cell Death Dis. 2022;13(5):509. doi: 10.1038/s41419-022-04912-8
  134. Béroud C, Soussi T. p53 gene mutation: software and database. Nucleic Acids Res. 1998;26(1):200–204. doi: 10.1093/nar/26.1.200
  135. Kessel D. Critical PDT theory III: events at the molecular and cellular level. Int J Mol Sci. 2022;23(11):6195. doi: 10.3390/ijms23116195
  136. Sinha K, Das J, Pal PB, Sil PC. Oxidative stress: the mitochondria-dependent and mitochondria-independent pathways of apoptosis. Arch Toxicol. 2013;87(7):1157–1180. doi: 10.1007/s00204-013-1034-4
  137. Marchi S, Patergnani S, Missiroli S, et al. Mitochondrial and endoplasmic reticulum calcium homeostasis and cell death. Cell Calcium. 2018;69:62–72. doi: 10.1016/j.ceca.2017.05.003
  138. Verfaillie T, Rubio N, Garg AD, et al. PERK is required at the ER-mitochondrial contact sites to convey apoptosis after ROS-based ER stress. Cell Death Differ. 2012;19(11):1880–1891. doi: 10.1038/cdd.2012.74
  139. Garg AD, Krysko D V., Vandenabeele P, Agostinis P. The emergence of phox-ER stress induced immunogenic apoptosis. Oncoimmunology. 2012;1(5):786–788. doi: 10.4161/onci.19750
  140. de Almeida DRQ, dos Santos AF, Wailemann RAM, et al. Necroptosis activation is associated with greater methylene blue-photodynamic therapy-induced cytotoxicity in human pancreatic ductal adenocarcinoma cells. Photochem Photobiol Sci. 2023;22(4):729–744. doi: 10.1007/s43630-022-00347-4
  141. Chen X, Shi C, He M, et al. Endoplasmic reticulum stress: molecular mechanism and therapeutic targets. Signal Transduct Target Ther. 2023;8(1):352. doi: 10.1038/s41392-023-01570-w
  142. Zheng L, Li Z, Wang Z, et al. The oxidative stress, mitochondrial pathway apoptosis, and the antagonistic effects of chrysophanol in SH-SY5Y cells via DTPP-induced photodynamic therapy. Photodiagnosis Photodyn Ther. 2022;40:103062. doi: 10.1016/j.pdpdt.2022.103062
  143. Donohoe C, Senge MO, Arnaut LG, Gomes-da-Silva LC. Cell death in photodynamic therapy: from oxidative stress to anti-tumor immunity. Biochim Biophys Acta Rev Cancer. 2019;1872(2):188308. doi: 10.1016/j.bbcan.2019.07.003
  144. Fiaschi T, Chiarugi P. Oxidative stress, tumor microenvironment, and metabolic reprogramming: a diabolic liaison. Int J Cell Biol. 2012;2012:762825. doi: 10.1155/2012/762825
  145. Averbeck D, Rodriguez-Lafrasse C. Role of mitochondria in radiation responses: epigenetic, metabolic, and signaling impacts. Int J Mol Sci. 2021;22(20):11047. doi: 10.3390/ijms222011047
  146. Xu Y, Shen J, Ran Z. Emerging views of mitophagy in immunity and autoimmune diseases. Autophagy. 2020;16(1):3–17. doi: 10.1080/15548627.2019.1603547
  147. Picca A, Calvani R, Coelho-Junior HJ, Marzetti E. Cell death and inflammation: the role of mitochondria in health and disease. Cells. 2021;10(3):537. doi: 10.3390/cells10030537
  148. Banoth B, Cassel SL. Mitochondria in innate immune signaling. Transl Res. 2018;202:52–68. doi: 10.1016/j.trsl.2018.07.014
  149. West AP, Shadel GS, Ghosh S. Mitochondria in innate immune responses. Nat Rev Immunol. 2011;11(6):389–402. doi: 10.1038/nri2975
  150. Zhou J, Wang G, Chen Y, et al. Immunogenic cell death in cancer therapy: present and emerging inducers. J Cell Mol Med. 2019;23(8):4854–4865. doi: 10.1111/jcmm.14356
  151. Zhu M, Yang M, Zhang J, et al. Immunogenic cell death induction by ionizing radiation. Front Immunol. 2021;12:705361. doi: 10.3389/fimmu.2021.705361
  152. Kaufman HL, Kohlhapp FJ, Zloza A. Oncolytic viruses: a new class of immunotherapy drugs. Nat Rev Drug Discov. 2015;14(9):642–662. Corrected and republished from: Nat Rev Drug Discov. 2016;15(9):660. doi: 10.1038/nrd4663
  153. Li W, Yang J, Luo L, et al. Targeting photodynamic and photothermal therapy to the endoplasmic reticulum enhances immunogenic cancer cell death. Nat Commun. 2019;10(1):3349. doi: 10.1038/s41467-019-11269-8
  154. Li Z, Chu Z, Yang J, et al. Immunogenic cell death augmented by manganese zinc sulfide nanoparticles for metastatic melanoma immunotherapy. ACS Nano. 2022;16(9):15471–15483. doi: 10.1021/acsnano.2c08013
  155. Guo J, Yu Z, Sun D, et al. Two nanoformulations induce reactive oxygen species and immunogenetic cell death for synergistic chemo-immunotherapy eradicating colorectal cancer and hepatocellular carcinoma. Mol Cancer. 2021;20(1):10. doi: 10.1186/s12943-020-01297-0
  156. Ecker V, Stumpf M, Brandmeier L, et al. Targeted PI3K/AKT-hyperactivation induces cell death in chronic lymphocytic leukemia. Nat Commun. 2021;12(1):3526. doi: 10.1038/s41467-021-23752-2
  157. Garg AD, Agostinis P. ER stress, autophagy and immunogenic cell death in photodynamic therapy-induced anti-cancer immune responses. Photochem Photobiol Sci. 2014;13(3):474–487. doi: 10.1039/c3pp50333j
  158. Szegezdi E, Logue SE, Gorman AM, Samali A. Mediators of endoplasmic reticulum stress-induced apoptosis. EMBO Rep. 2006;7(9):880–885. doi: 10.1038/sj.embor.7400779
  159. Hetz C. The unfolded protein response: controlling cell fate decisions under ER stress and beyond. Nat Rev Mol Cell Biol. 2012;13(2):89–102. doi: 10.1038/nrm3270
  160. Sano R, Reed JC. ER stress-induced cell death mechanisms. Biochimica et Biophysica Acta. 2013;1833(12):3460–3470. doi: 10.1016/j.bbamcr.2013.06.028
  161. Sperandio S, de Belle I, Bredesen DE. An alternative, nonapoptotic form of programmed cell death. Proc Natl Acad Sci U S A. 2000;97(26):14376–14381. doi: 10.1073/pnas.97.26.14376
  162. Kessel D, Obaid G, Rizvi I. Critical PDT theory II: current concepts and indications. Photodiagnosis Photodyn Ther. 2022;39:102923. doi: 10.1016/j.pdpdt.2022.102923
  163. Görlach A, Bertram K, Hudecova S, Krizanova O. Calcium and ROS: a mutual interplay. Redox Biol. 2015;6:260–271. doi: 10.1016/j.redox.2015.08.010
  164. Bauer TM, Murphy E. Role of mitochondrial calcium and the permeability transition pore in regulating cell death. Circ Res. 2020;126(2):280–293. doi: 10.1161/CIRCRESAHA.119.316306
  165. Carreras-Sureda A, Pihán P, Hetz C. Calcium signaling at the endoplasmic reticulum: fine-tuning stress responses. Cell Calcium. 2018;70:24–31. doi: 10.1016/j.ceca.2017.08.004
  166. Xie D, Wang Q, Wu G. Research progress in inducing immunogenic cell death of tumor cells. Front Immunol. 2022;13:1017400. doi: 10.3389/fimmu.2022.1017400
  167. Chen Y, Zhou Z, Min W. Mitochondria, oxidative stress and innate immunity. Front Physiol. 2018;9:1487. doi: 10.3389/fphys.2018.01487
  168. Hayes JD, Dinkova-Kostova AT, Tew KD. Oxidative stress in cancer. Cancer Cell. 2020;38(2):167–197. doi: 10.1016/j.ccell.2020.06.001
  169. Maharjan PS, Bhattarai HK. Singlet oxygen, photodynamic therapy, and mechanisms of cancer cell death. J Oncol. 2022;2022:7211485. doi: 10.1155/2022/7211485
  170. Krysko O, Løve Aaes T, Bachert C, et al. Many faces of DAMPs in cancer therapy. Cell Death Dis. 2013;4(5):e631. doi: 10.1038/cddis.2013.156
  171. Yousefi S, Perozzo R, Schmid I, et al. Calpain-mediated cleavage of Atg5 switches autophagy to apoptosis. Nat Cell Biol. 2006;8(10):1124–1132. doi: 10.1038/ncb1482
  172. Zhang A, Wu Y, Lai HWL, Yew DT. Apoptosis — a brief review. Neuroembryology. 2004;3(1):47–59. doi: 10.1159/000085404
  173. D’Arcy MS. Cell death: a review of the major forms of apoptosis, necrosis and autophagy. Cell Biol Int. 2019;43(6):582–592. doi: 10.1002/cbin.11137
  174. Van Cruchten S, Van den Broeck W. Morphological and biochemical aspects of apoptosis, oncosis and necrosis. Anat Histol Embryol. 2002;31(4):214–223. doi: 10.1046/j.1439-0264.2002.00398.x
  175. Kepp O, Tesniere A, Schlemmer F, et al. Immunogenic cell death modalities and their impact on cancer treatment. Apoptosis. 2009;14(4):364–375. doi: 10.1007/s10495-008-0303-9
  176. Galluzzi L, Vitale I, Vacchelli E, Kroemer G. Cell death signaling and anticancer therapy. Front Oncol. 2011;1:5. doi: 10.3389/fonc.2011.00005
  177. Richardson DR, Ponka P. The molecular mechanisms of the metabolism and transport of iron in normal and neoplastic cells. Biochim Biophys Acta. 1997;1331(1):1–40. doi: 10.1016/s0304-4157(96)00014-7
  178. Ursini F, Maiorino M, Valente M, et al. Purification from pig liver of a protein which protects liposomes and biomembranes from peroxidative degradation and exhibits glutathione peroxidase activity on phosphatidylcholine hydroperoxides. Biochim Biophys Acta. 1982;710(2):197–211. doi: 10.1016/0005-2760(82)90150-3
  179. Chen Y, Zheng W, Li Y, et al. Apoptosis induced by methylene-blue-mediated photodynamic therapy in melanomas and the involvement of mitochondrial dysfunction revealed by proteomics. Cancer Sci. 2008;99(10):2019–2027. doi: 10.1111/j.1349-7006.2008.00910.x
  180. Agostinis P, Breyssens H, Buytaert E, Hendrickx N. Regulatory pathways in photodynamic therapy induced apoptosis. Photochem Photobiol Sci. 2004;3(8):721–729. doi: 10.1039/b315237e
  181. Lu Y, Jiao R, Chen X, et al. Methylene blue-mediated photodynamic therapy induces mitochondria-dependent apoptosis in HeLa cell. J Cell Biochem. 2008;105(6):1451–1460. doi: 10.1002/jcb.21965
  182. Xie Q, Li Z, Liu Y, et al. Translocator protein-targeted photodynamic therapy for direct and abscopal immunogenic cell death in colorectal cancer. Acta Biomater. 2021;134:716–729. doi: 10.1016/j.actbio.2021.07.052
  183. Humeau J, Lévesque S, Kroemer G, Pol JG. Gold standard assessment of immunogenic cell death in oncological mouse models. Methods Mol Biol. 2019;1884:297–315. doi: 10.1007/978-1-4939-8885-3_21
  184. Panzarini E, Inguscio V, Dini L. Timing the multiple cell death pathways initiated by Rose Bengal acetate photodynamic therapy. Cell Death Dis. 2011;2(6):e169. doi: 10.1038/cddis.2011.51
  185. Cincotta L, Szeto D, Lampros E, et al. Benzophenothiazine and benzoporphyrin derivative combination phototherapy effectively eradicates large murine sarcomas. Photochem Photobiol. 1996;63(2):229–237. doi: 10.1111/j.1751-1097.1996.tb03019.x
  186. Guo R, Li Y, Wang Z, et al. Hypoxia-inducible factor-1α and nuclear factor-κB play important roles in regulating programmed cell death ligand 1 expression by epidermal growth factor receptor mutants in non-small-cell lung cancer cells. Cancer Sci. 2019;110(5):1665–1675. doi: 10.1111/cas.13989
  187. Noman MZ, Desantis G, Janji B, et al. PD-L1 is a novel direct target of HIF-1α, and its blockade under hypoxia enhanced MDSC-mediated T cell activation. J Exp Med. 2014;211(5):781–790. doi: 10.1084/jem.20131916
  188. Jiang X, Wang J, Deng X, et al. Role of the tumor microenvironment in PD-L1/PD-1-mediated tumor immune escape. Mol Cancer. 2019;18(1):10. doi: 10.1186/s12943-018-0928-4

Supplementary files

Supplementary Files
Action
1. JATS XML
2. Fig. 1. Main mechanisms triggering immunogenic cell death under photodynamic exposure. The localization of the photosensitizer in the endoplasmic reticulum leads to misfolding of proteins, which triggers three pathways of endoplasmic reticulum stress. The accumulation of reactive oxygen species in the ER during photodynamic therapy also leads to the loss of calcium homeostasis and its outflow into the cytosol or mitochondria, including through mitochondria-associated membranes. When a photosensitizer is localized in the mitochondria, large amounts of reactive oxygen species accumulate in organelles, and their integrity is subsequently lost, accompanied by the release of numerous apoptotic factors. The accumulation of photosensitizers and reactive oxygen species formation in lysosomes lead to the loss of lysosomal calcium and the convergence of ATG5 to tATG5, followed by the triggering of mitochondrial apoptosis. MMP — mitochondrial membrane potential; Bsl-XL — B-cell lymphoma-extra large. See text for designations.

Download (1MB)

Copyright (c) 2024 Eco-Vector

Creative Commons License
This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License.

СМИ зарегистрировано Федеральной службой по надзору в сфере связи, информационных технологий и массовых коммуникаций (Роскомнадзор).
Регистрационный номер и дата принятия решения о регистрации СМИ: 

This website uses cookies

You consent to our cookies if you continue to use our website.

About Cookies