[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,297)

Search Parameters:
Keywords = PDT

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 24194 KiB  
Case Report
Antimicrobial Photodynamic Therapy Combined with Photobiomodulation Therapy in Teeth with Asymptomatic Apical Periodontitis: A Case Series
by Francisco Rubio, Josep Arnabat-Domínguez, Eulàlia Sans-Serramitjana, Camila Saa, Kinga Grzech-Leśniak and Pablo Betancourt
Appl. Sci. 2024, 14(20), 9341; https://doi.org/10.3390/app14209341 (registering DOI) - 14 Oct 2024
Viewed by 333
Abstract
Introduction: Apical periodontitis (AP) is an inflammatory disease of the periapical tissues that is often asymptomatic and diagnosed through radiographic examination. A challenge in traditional endodontics is disinfection of the root canal system (RCS), which anatomically presents numerous variations, often leading to persistent [...] Read more.
Introduction: Apical periodontitis (AP) is an inflammatory disease of the periapical tissues that is often asymptomatic and diagnosed through radiographic examination. A challenge in traditional endodontics is disinfection of the root canal system (RCS), which anatomically presents numerous variations, often leading to persistent infections. Antimicrobial photodynamic therapy (aPDT) and photobiomodulation therapy (PBMT) offer promising adjuncts, due to their antimicrobial and tissue-healing properties. Objective: The aim of this article was to report five cases of teeth with pulp necrosis and asymptomatic apical periodontitis (AAP) treated with aPDT and PBMT protocols. Materials and Methods: Five cases of pulp necrosis and AAP were treated with conventional endodontic therapy supplemented with aPDT and PBMT. The treatment protocol included chemomechanical preparation (CMP), aPDT using a 660 nm diode laser (DL) with methylene blue (MB) as a photosensitizer (5 min pre-irradiation time), and PBMT using a 940 nm DL. Treatment results were evaluated through cone-beam computed tomography (CBCT)-based evaluation over 1 year of clinical follow-up. Results: All cases showed significant bone regeneration and tissue healing, demonstrating the efficacy of the combination of aPDT and PBMT. Post-operative pain did not occur in any of the patients, suggesting a possible analgesic effect of PBMT. Conclusions: The combination of aPDT and PBMT in endodontic therapy promoted tissue recovery and improved the prognosis of AAP. Further research and randomized control trials are needed to optimize treatment protocols and evaluate the long-term effects. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of mechanisms of antimicrobial photodynamic therapy: (<b>A</b>) When light is absorbed, the PS is activated, shifting from its ground state to an excited singlet state. This excited state may undergo inter-system crossing, where the spin quantum number changes, resulting in a lower-energy, longer-lasting triplet state. The triplet state can then react through one or both of the oxygen-dependent mechanisms, known as Type I and Type II photoprocesses; (<b>B</b>) in the type I pathway, electron transfer reactions occur from the triplet PS state, involving a substrate to form radical ions (free radicals), which can then interact with oxygen to generate cytotoxic species such as superoxide, hydroxyl radicals, and lipid radicals. In contrast, in the Type II pathway, energy transfer from the triplet PS state to ground-state molecular oxygen occurs, producing singlet oxygen which can oxidize biological molecules such as proteins, nucleic acids, and lipids, leading to cytotoxic outcomes.</p>
Full article ">Figure 2
<p>Schematic of the mechanisms of action and clinical effects of photobiomodulation: (<b>A</b>) Cytochrome C oxidase (CCO), the primary chromophore located in complex IV of the mitochondrial respiratory chain, plays a critical role in PBMT. When CCO absorbs light (particularly in the red and near-infrared wavelengths), it leads to the photodissociation of nitric oxide (NO). This process reduces oxidative stress and subsequently boosts the production of reactive oxygen species (ROS), ATP, and calcium ions (Ca<sup>2+</sup>). As a result of increased ATP and protein synthesis, the growth factor response within cells and tissue occurs. These changes promote the desired biological responses, including anti-inflammatory, analgesic, and healing effects, while also enhancing cellular processes such as differentiation, proliferation, and migration. For these effects to occur, light must be absorbed by chromophores, which are interrelated molecules (e.g., enzymes, cell membranes, and extracellular substances) that have the capacity to absorb light. (<b>B</b>) Extraoral application of laser (red/infrared) directed toward the apical area of the affected tooth. By irradiating this area as a supplement to the endodontic treatment, accelerated bone tissue regeneration can be achieved.</p>
Full article ">Figure 3
<p>Three-dimensional CBCT images of tooth 1.2. (<b>A</b>–<b>C</b>) Coronal, sagittal, and axial pre-treatment sections. The CBCT examination showed an osteolytic lesion of dimensions 14 × 11 × 9 mm, causing expansion and fenestration of the vestibular table. (<b>D</b>–<b>F</b>) Coronal, sagittal, and axial post-laser treatment images show evident bone tissue gain, including recovery of the vestibular bone table.</p>
Full article ">Figure 4
<p>Intracanal disinfection with aPDT was conducted utilizing a DL with the following specifications: wavelength 660 ± 10 nm, power 100 mW. (<b>A</b>) MB inside the root canal and pulp chamber; (<b>B</b>) the laser used for all cases was therapy EC; (<b>C</b>) activation of MB using laser light inside the root canal; (<b>D</b>) the PBMT treatment entailed employing a wavelength of 940 ± 10 nm, one spot measuring 0.2 cm<sup>2</sup>, and a power of 0.1 W.</p>
Full article ">Figure 5
<p>CBCT images of teeth 1.1 and 2.1: (<b>A</b>–<b>C</b>) Coronal, sagittal, and axial pre-treatment images show an osteolytic lesion measuring 11 × 10 × 8 mm, resulting in expansion and fenestration of the vestibular table and (<b>D</b>–<b>F</b>) coronal, sagittal, and axial post-treatment images, which evidence significant bone regeneration in the affected area.</p>
Full article ">Figure 6
<p>CBCT images of tooth 1.2: (<b>A</b>–<b>C</b>) Coronal, Sagittal, and axial pre-treatment images revealed an osteolytic lesion measuring 10 × 10 × 8 mm, resulting in expansion and fenestration of the vestibular table and (<b>D</b>–<b>F</b>) coronal, sagittal, and axial post-treatment images. Notice the gain of bone tissue throughout the affected area.</p>
Full article ">Figure 7
<p>CBCT images of tooth 3.3. Coronal, axial, and transverse CBCT views at pre-treatment (<b>A</b>–<b>C</b>) and post-treatment (<b>D</b>–<b>F</b>). The apical lesion initially measured 10 × 9 × 8 mm. A clear recovery of trabecular and buccal cortical bone can be observed, when comparing images C and F, after using conventional endodontic therapy combined with laser therapy in both its aPDT and PBMT modalities.</p>
Full article ">Figure 8
<p>CBCT images of tooth 2.7. Coronal, sagittal, and axial CBCT images before (<b>A</b>–<b>C</b>) and after treatment (<b>D</b>–<b>F</b>). The three-dimensional CBCT study reveals an osteolytic lesion measuring 9 × 9 × 6 mm at the apical level, affecting the mesiobuccal and distobuccal roots. Comparing images C and F, it can be observed that, after 1 year of laser therapy, the lesion had satisfactorily resolved.</p>
Full article ">Figure 9
<p>Control periapical X-rays of the five cases treated with conventional non-surgical endodontics supplemented with aPDT and PBMT: (<b>A</b>) Case 1, Tooth 1.2. An apical puff can be observed. (<b>B</b>) Case 2, Teeth 1.1 and 2.1. The root canals appear fully obturated. (<b>C</b>) Case 3, Tooth 1.2. A straight root canal is completely sealed. (<b>D</b>) Case 4, Tooth 3.3. The tooth is obturated in the cervical, middle, and apical thirds. (<b>E</b>) Case 5, Tooth 2.7. Endodontic obturation of the mesiobuccal, distobuccal, and palatal canals is visible. In all cases, bone tissue regeneration can be observed at the apical level.</p>
Full article ">
9 pages, 1213 KiB  
Article
Choroidal and Choriocapillaris Changes after Photodynamic Therapy and Subthreshold Micropulse Laser Treatment for Central Serous Chorioretinopathy
by Maria Ludovica Ruggeri, Marta Di Nicola, Marzia Passamonti, Carolina Lorenzi, Alberto Quarta, Rodolfo Mastropasqua and Lisa Toto
Medicina 2024, 60(10), 1674; https://doi.org/10.3390/medicina60101674 - 12 Oct 2024
Viewed by 266
Abstract
Background and Objectives: The aim of the present study is to analyze choroidal and choriocapillaris structural and functional changes in eyes affected by Central serous chorioretinopathy after Photodynamic Therapy (PDT) and Subthreshold Micropulse laser (SML) treatment. Materials and Methods: Forty-two eyes [...] Read more.
Background and Objectives: The aim of the present study is to analyze choroidal and choriocapillaris structural and functional changes in eyes affected by Central serous chorioretinopathy after Photodynamic Therapy (PDT) and Subthreshold Micropulse laser (SML) treatment. Materials and Methods: Forty-two eyes of forty-two patients were analyzed in this observational study. Twenty-four patients underwent SML treatment, whereas eighteen patients were treated with PDT. Examinations were performed at baseline and after 3 months of treatment. Main outcome measures were: Best corrected visual acuity (BCVA), central macular thickness (CMT), central choroidal thickness (CCT), pigment epithelial detachment (PED) presence and maximum height (PEDMH), and choroidal vascularity index (CVI) measured by means of Spectralis HRA + OCT (Heidelberg Engineering, Heidelberg, Germany) Optical coherence tomography (OCT) and choriocapillaris flow voids (CCFV) measured on Optical Coherence Tomography Angiography (OCT-A) platform PLEX Elite 9000 device (Carl Zeiss Meditec Inc., Dublin, CA, USA). Results: Changes in BCVA were registered in both groups over time (p < 0.001). Structural changes in terms of reduced CMT and PED presence were noted in the two groups at follow-up (p < 0.001 and p = 0.001, respectively). Structural and functional choroidal changes were found in the two groups with reduced CCT and CVI over time (p = 0.004 and p = 0.007, respectively), with significant differences between the two groups for CVI parameter (p = 0.001). CCFV increased over time in the PDT group and decreased in the SML group. Conclusions: PDT and SML are effective approaches in CSC eyes and are able to improve structural and functional parameters over time. Choroidal and choriocapillaris parameters are promising biomarkers able to monitor disease course, showing greater impact of PDT on choroid-choriocapillaris complex over time. Full article
(This article belongs to the Special Issue Retinal Diseases: Clinical Presentation and Novel Treatments)
Show Figures

Figure 1

Figure 1
<p>Percentage of presence of PED at baseline and at 3 months of treatment in two groups. McNemar test <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Choroidal changes in SML group before (<b>A</b>) and after treatment (<b>B</b>) measured by CVI. Changes in CVI before (<b>C</b>) and after (<b>D</b>) PDT are reported.</p>
Full article ">
16 pages, 5222 KiB  
Review
Application and Challenge of Metalloporphyrin Sensitizers in Noninvasive Dynamic Tumor Therapy
by Jiacheng Ouyang, Dan Li, Lizhen Zhu, Xiaoyuan Cai, Lanlan Liu, Hong Pan and Aiqing Ma
Molecules 2024, 29(20), 4828; https://doi.org/10.3390/molecules29204828 - 11 Oct 2024
Viewed by 428
Abstract
Dynamic tumor therapies (mainly including photodynamic therapy (PDT) and sonodynamic therapy (SDT)) offer new approaches to cancer treatment. They are often characterized by their noninvasive nature, high selectivity, and low toxicity. Sensitizers are crucial for dynamic therapy. Developing efficient sensitizers with good biocompatibility [...] Read more.
Dynamic tumor therapies (mainly including photodynamic therapy (PDT) and sonodynamic therapy (SDT)) offer new approaches to cancer treatment. They are often characterized by their noninvasive nature, high selectivity, and low toxicity. Sensitizers are crucial for dynamic therapy. Developing efficient sensitizers with good biocompatibility and controllability is an important aim in dynamic therapy. Porphyrins and metalloporphyrins attract great attention due to their excellent photophysical properties and low cytotoxicity under non-light. Compared to porphyrins, metalloporphyrins show greater potential for dynamic therapy due to their enhanced photochemical and photophysical properties after metal ions coordinate with porphyrin rings. This paper reviews some metalloporphyrin-based sensitizers used in photo/sonodynamic therapy and combined therapy. In addition, the probable challenges and bottlenecks in clinical translation are also discussed. Full article
(This article belongs to the Special Issue Study on Synthesis and Photochemistry of Dyes)
Show Figures

Figure 1

Figure 1
<p>Typical structures, i.e., (<b>A</b>) pyrrole, (<b>B</b>) porphyrin, consisting of four pyrrole rings joined by methane bridges, and (<b>C</b>) metalloporphyrin (M = Mn, Fe, Cu, Zn, and so on).</p>
Full article ">Figure 2
<p>The possible mechanisms of PDT (<b>A</b>) and SDT (<b>B</b>).</p>
Full article ">Figure 3
<p>The photophysical properties and effect of different metal chelation on PoP<sup>41</sup> and Por<sup>42</sup> ligands. (<b>A</b>) Chemical structure of the HPPH lipids (PoP) examined. (<b>B</b>) Fluorescence emission spectra of an equal concentration of indicated PoP liposomes in phosphate-buffered saline. (<b>C</b>) Singlet oxygen generation was assessed indirectly by examining the increase in green fluorescence of a singlet oxygen sensor before and after laser irradiation. (<b>D</b>) Chemical structure of ZnPor, PdPor, and PtPor. (<b>E</b>) The fluorescence intensity detection at the characteristic peak of DCFH (525 nm) irradiated by different complexes as a function of irradiation time.</p>
Full article ">Figure 4
<p>Different metalloporphyrins for enhanced antitumor PDT through assistant reaction and function. (<b>A</b>) TiOP-loaded liposome nanosystem (FA−TiOPs) to photocatalyze H<sub>2</sub>O and H<sub>2</sub>O<sub>2</sub> for antitumor PDT (a) Preparation of FA–TiOPs. (b) Mechanism of in situ supplying ROS under FA–TiOPs photocatalysis [<a href="#B48-molecules-29-04828" class="html-bibr">48</a>]. (<b>B</b>) The enhanced ROS generation evaluation induced by FA-TiOPs under light irradiation. (<b>C</b>) The structure of ZnP1 and ZnP2 [<a href="#B49-molecules-29-04828" class="html-bibr">49</a>]. (<b>D</b>) Scheme of photothermal-assistant PDT-based ZnP2. (<b>E</b>) Scheme of self-supply CuS and (Cu)HMME for photothermal-assistant PDT. I: GSH triggered H<sub>2</sub>S generation; II: The H<sub>2</sub>S triggered CuS production and the resealse of metalloporphyrin ((Cu)HMMe) [<a href="#B50-molecules-29-04828" class="html-bibr">50</a>].</p>
Full article ">Figure 5
<p>Imaging-guided precise tumor treatment with metalloporphyrin-based PDT. (<b>A</b>) The structure of <sup>68</sup>Ga–porphyrin complex [<a href="#B53-molecules-29-04828" class="html-bibr">53</a>]. (<b>B</b>) MR imaging of <sup>68</sup>Ga–porphyrin in tumor site. (<b>C</b>) Schematic illustration of the preparation of Gd–PNPs and their application in FL/MR imaging-guided PDT [<a href="#B55-molecules-29-04828" class="html-bibr">55</a>]. (<b>D</b>) Schematic illustrating the co-assembly of the nanocomposites by ZnTPP and GdTPP for MR/FL bimodal imaging-guided PDT [<a href="#B56-molecules-29-04828" class="html-bibr">56</a>].</p>
Full article ">Figure 5 Cont.
<p>Imaging-guided precise tumor treatment with metalloporphyrin-based PDT. (<b>A</b>) The structure of <sup>68</sup>Ga–porphyrin complex [<a href="#B53-molecules-29-04828" class="html-bibr">53</a>]. (<b>B</b>) MR imaging of <sup>68</sup>Ga–porphyrin in tumor site. (<b>C</b>) Schematic illustration of the preparation of Gd–PNPs and their application in FL/MR imaging-guided PDT [<a href="#B55-molecules-29-04828" class="html-bibr">55</a>]. (<b>D</b>) Schematic illustrating the co-assembly of the nanocomposites by ZnTPP and GdTPP for MR/FL bimodal imaging-guided PDT [<a href="#B56-molecules-29-04828" class="html-bibr">56</a>].</p>
Full article ">Figure 6
<p>The SDT effects of some metalloporphyrins with different metal centers and porphyrin ligands. (<b>A</b>) Scheme illustrating the synthesis of MTTP complexes with different metal centers and the corresponding nanocomplexes of HSA for antitumor SDT [<a href="#B73-molecules-29-04828" class="html-bibr">73</a>]. (<b>B</b>) Structural illustration of Mn-PpIX-based sonosensitizer for antitumor SDT [<a href="#B74-molecules-29-04828" class="html-bibr">74</a>]. (<b>C</b>) Schematic illustration of the synthesis of FA–L–CuPP [<a href="#B75-molecules-29-04828" class="html-bibr">75</a>]. (<b>D</b>) The mechanism of Cu(II)NS and Cu(I)NS for SDT [<a href="#B76-molecules-29-04828" class="html-bibr">76</a>].</p>
Full article ">Figure 7
<p>Tumor microenvironment adjustment relevant to SDT effect. (<b>A</b>) MnP-mediated SDT-ICD antitumor [<a href="#B78-molecules-29-04828" class="html-bibr">78</a>]. (<b>B</b>) Schematic of preparation and oxygen-enhanced SDT based on DOX/Mn-TPPS@RBCs [<a href="#B79-molecules-29-04828" class="html-bibr">79</a>]. (<b>C</b>) Oxygen-enhanced SDT based on Fe–porphyrin [<a href="#B80-molecules-29-04828" class="html-bibr">80</a>]. (<b>D</b>) Intracellular NADH oxidization for enhanced SDT effect [<a href="#B81-molecules-29-04828" class="html-bibr">81</a>].</p>
Full article ">
19 pages, 4897 KiB  
Article
Photodynamic Therapy against Colorectal Cancer Using Porphin-Loaded Arene Ruthenium Cages
by Suzan Ghaddar, Aline Pinon, Manuel Gallardo-Villagran, Jacquie Massoud, Catherine Ouk, Claire Carrion, Mona Diab-Assaf, Bruno Therrien and Bertrand Liagre
Int. J. Mol. Sci. 2024, 25(19), 10847; https://doi.org/10.3390/ijms251910847 - 9 Oct 2024
Viewed by 412
Abstract
Colorectal cancer (CRC) is the third most common cancer in the world, with an ongoing rising incidence. Despite secure advancements in CRC treatments, challenges such as side effects and therapy resistance remain to be addressed. Photodynamic therapy (PDT) emerges as a promising modality, [...] Read more.
Colorectal cancer (CRC) is the third most common cancer in the world, with an ongoing rising incidence. Despite secure advancements in CRC treatments, challenges such as side effects and therapy resistance remain to be addressed. Photodynamic therapy (PDT) emerges as a promising modality, clinically used in treating different diseases, including cancer. Among the main challenges with current photosensitizers (PS), hydrophobicity and low selective uptake by the tumor remain prominent. Thus, developing an optimal design for PS to improve their solubility and enhance their selective accumulation in cancer cells is crucial for enhancing the efficacy of PDT. Targeted photoactivation triggers the production of reactive oxygen species (ROS), which promote oxidative stress within cancer cells and ultimately lead to their death. Ruthenium (Ru)-based compounds, known for their selective toxicity towards cancer cells, hold potential as anticancer agents. In this study, we investigated the effect of two distinct arene-Ru assemblies, which lodge porphin PS in their inner cavity, and tested them as PDT agents on the HCT116 and HT-29 human CRC cell lines. The cellular internalization of the porphin-loaded assemblies was confirmed by fluorescence microscopy. Additionally, significant photocytotoxicity was observed in both cell lines after photoactivation of the porphin in the cage systems, inducing apoptosis through caspase activation and cell cycle progression disruptions. These findings suggest that arene-Ru assemblies lodging porphin PS are potent candidates for PDT of CRC. Full article
(This article belongs to the Special Issue New Molecular Aspects of Colorectal Cancer)
Show Figures

Figure 1

Figure 1
<p>Schematic description of the cellular mechanisms induced by <b>PS⸦M1</b> and <b>PS⸦M2</b>, evaluated in this study. (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 29 June 2024)).</p>
Full article ">Figure 2
<p>Chemical structure of porphin (<b>PS</b>) and the arene-ruthenium assemblies <b>M1</b> and <b>M2</b>, together with the host–guest systems <b>PS⸦M</b>.</p>
Full article ">Figure 3
<p>Phototoxicity of arene-Ru porphin <b>PS</b> assemblies on human CRC cell lines. (<b>A</b>) HCT116 and (<b>B</b>) HT-29 cell lines were cultured in RPMI medium for 24 h. After 24 h, cells were treated or not with <b>PS⸦M1</b>, <b>PS⸦M2</b>, <b>M1</b>, or <b>M2</b>. Illumination (630 nm, 75 J/cm<sup>2</sup>) of the cells occurred 24 h after treatment, and the cell viability for all the conditions was determined 12 h, 24 h, and 48 h post-illumination. Data are represented as a mean ± SEM of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001. (<b>C</b>) Graphical representation of the IC<sub>50</sub> values (nM) of <b>PS⸦M1</b> and <b>PS⸦M2</b> determined by MTT assay on HCT116 and HT-29 cell lines. Data are represented as a mean ± SEM of three independent experiments.</p>
Full article ">Figure 3 Cont.
<p>Phototoxicity of arene-Ru porphin <b>PS</b> assemblies on human CRC cell lines. (<b>A</b>) HCT116 and (<b>B</b>) HT-29 cell lines were cultured in RPMI medium for 24 h. After 24 h, cells were treated or not with <b>PS⸦M1</b>, <b>PS⸦M2</b>, <b>M1</b>, or <b>M2</b>. Illumination (630 nm, 75 J/cm<sup>2</sup>) of the cells occurred 24 h after treatment, and the cell viability for all the conditions was determined 12 h, 24 h, and 48 h post-illumination. Data are represented as a mean ± SEM of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001. (<b>C</b>) Graphical representation of the IC<sub>50</sub> values (nM) of <b>PS⸦M1</b> and <b>PS⸦M2</b> determined by MTT assay on HCT116 and HT-29 cell lines. Data are represented as a mean ± SEM of three independent experiments.</p>
Full article ">Figure 4
<p>Intracellular ROS production was evaluated in (<b>A</b>) HCT116 and (<b>B</b>) HT-29 human CRC cell lines immediately after illumination (630 nm, 75 J/cm<sup>2</sup>) using DCFDA staining. Cells were analyzed using flow cytometry. Quantification of the intensity of fluorescence emitted due to DCF formation is correlated to the level of ROS generation. Data are represented as a mean ± SEM of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Detection of the cellular internalization of <b>PS⸦M1</b> and <b>PS⸦M2</b> in HCT116 (<b>A</b>) and HT-29 (<b>B</b>) cell lines by confocal microscopy. The cells were seeded into incubation chambers and cultured for 24 h. The cells were then treated with the compounds, and the fluorescence was measured by confocal microscopy (laser Zeiss LSM 510 Meta―×1000). The internalization was processed using the ImageJ image-processing software (version 1.54f). White scale bar = 20 μm.</p>
Full article ">Figure 6
<p>Cell cycle distribution analysis on HCT116 and HT-29 cell lines after the photoactivation of <b>PS⸦M1</b> and <b>PS⸦M2</b>. Cells were seeded for 24 h in a culture medium before the treatment or not with the assemblies at their determined IC<sub>50</sub> concentrations. After 24 h of treatment, cells were illuminated or not with red light at 630 nm and 75 J/cm<sup>2</sup>. Cells were collected at 12 h, 24 h, and 48 h post-illumination for analysis by flow cytometry using PI staining. Images of the cell cycle distribution on (<b>A</b>) HCT116 cells at 24 h post-illumination and (<b>B</b>) HT-29 cell line at 48 h post-illumination are represented. Histograms representing the percentage cell numbers at each phase of the cell cycle on (<b>C</b>) HCT116 and (<b>D</b>) HT-29 cell lines are displayed as a mean ± SEM of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Apoptosis due to photoactivation was assessed on HCT116 and HT-29 cell lines. Cells were seeded and incubated for 24 h then treated or not with <b>PS⸦M1</b> and <b>PS⸦M2</b> assemblies at IC<sub>50</sub> concentrations. Cells were either illuminated (630 nm, 75 J/cm<sup>2</sup>) or not after 24 h of treatment, and then, they were collected at 12 h, 24 h, and 48 h post-illumination. The collected cells were stained with Annexin V- FITC and PI, and their state was revealed by flow cytometry. Representative data from flow cytometry for the HCT116 cell line at 24 h post-illumination (<b>A</b>) and HT-29 at 48 h post-illumination (<b>B</b>) are displayed. Histograms represent the viable and apoptotic cell percentages of the treated HCT116 (<b>C</b>), and HT-29 (<b>D</b>) cells subjected to prior illumination. Data are represented as a mean ± SEM of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7 Cont.
<p>Apoptosis due to photoactivation was assessed on HCT116 and HT-29 cell lines. Cells were seeded and incubated for 24 h then treated or not with <b>PS⸦M1</b> and <b>PS⸦M2</b> assemblies at IC<sub>50</sub> concentrations. Cells were either illuminated (630 nm, 75 J/cm<sup>2</sup>) or not after 24 h of treatment, and then, they were collected at 12 h, 24 h, and 48 h post-illumination. The collected cells were stained with Annexin V- FITC and PI, and their state was revealed by flow cytometry. Representative data from flow cytometry for the HCT116 cell line at 24 h post-illumination (<b>A</b>) and HT-29 at 48 h post-illumination (<b>B</b>) are displayed. Histograms represent the viable and apoptotic cell percentages of the treated HCT116 (<b>C</b>), and HT-29 (<b>D</b>) cells subjected to prior illumination. Data are represented as a mean ± SEM of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Protein expression was evaluated by Western blot. HCT116 (<b>A</b>) and HT-29 (<b>B</b>) cells were seeded and incubated for 24 h and then were treated or not at IC<sub>50</sub> values of <b>PS⸦M1</b> or <b>PS⸦M2</b>. After treatment, cells were either illuminated (630 nm, 75 J/cm<sup>2</sup>) or not and collected 24 h and 48 h post-illumination. Proteins were extracted, and the level of protein expression of the different conditions were revealed. β-actin was used as a loading control. Representative images are shown.</p>
Full article ">Figure 9
<p>DNA fragmentation in HCT116 (<b>A</b>) and HT-29 (<b>B</b>) cells analyzed from cytosol extracts using ELISA assay. After seeding the cells for 24 h, followed by treatment, the cells were illuminated (630 nm, 75 J/cm<sup>2</sup>) or not. After 12 h, 24 h, and 48 h post-illumination, the cells were collected, and the level of DNA fragmentation was analyzed. Histograms are represented as a mean ± SEM of at least three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
24 pages, 6303 KiB  
Article
Optimization of the Treatment of Squamous Cell Carcinoma Cells by Combining Photodynamic Therapy with Cold Atmospheric Plasma
by Sigrid Karrer, Petra Unger, Nina Spindler, Rolf-Markus Szeimies, Anja Katrin Bosserhoff, Mark Berneburg and Stephanie Arndt
Int. J. Mol. Sci. 2024, 25(19), 10808; https://doi.org/10.3390/ijms251910808 - 8 Oct 2024
Viewed by 401
Abstract
Actinic keratosis (AK) is characterized by a reddish or occasionally skin-toned rough patch on sun-damaged skin, and it is regarded as a precursor to squamous cell carcinoma (SCC). Photodynamic therapy (PDT), utilizing 5-aminolevulinic acid (ALA) along with red light, is a recognized treatment [...] Read more.
Actinic keratosis (AK) is characterized by a reddish or occasionally skin-toned rough patch on sun-damaged skin, and it is regarded as a precursor to squamous cell carcinoma (SCC). Photodynamic therapy (PDT), utilizing 5-aminolevulinic acid (ALA) along with red light, is a recognized treatment option for AK that is limited by the penetration depth of light and the distribution of the photosensitizer into the skin. Cold atmospheric plasma (CAP) is a partially ionized gas with permeability-enhancing and anti-cancer properties. This study analyzed, in vitro, whether a combined treatment of CAP and ALA-PDT may improve the efficacy of the treatment. In addition, the effect of the application sequence of ALA and CAP was investigated using in vitro assays and the molecular characterization of human oral SCC cell lines (SCC-9, SCC-15, SCC-111), human cutaneous SCC cell lines (SCL-1, SCL-2, A431), and normal human epidermal keratinocytes (HEKn). The anti-tumor effect was determined by migration, invasion, and apoptosis assays and supported the improved efficacy of ALA-PDT in combination with CAP. However, the application sequence ALA-CAP–red light seems to be more efficacious than CAP-ALA–red light, which is probably due to increased intracellular ROS levels when ALA is applied first, followed by CAP and red light treatment. Furthermore, the expression of apoptosis- and senescence-related molecules (caspase-3, -6, -9, p16INK4a, p21CIP1) was increased, and different genes of the junctional network (ZO-1, CX31, CLDN1, CTNNB1) were induced after the combined treatment of CAP plus ALA-PDT. HEKn, however, were much less affected than SCC cells. Overall, the results show that CAP may improve the anti-tumor effects of conventional ALA-PDT on SCC cells. Whether this combined application is successful in treating AK in vivo has to be carefully examined in follow-up studies. Full article
(This article belongs to the Special Issue Molecular Aspects of Photodynamic Therapy)
Show Figures

Figure 1

Figure 1
<p>Onset and progression of squamous cell carcinoma (SCC). Early actinic keratosis (AK) is mainly caused by chronic ultraviolet (UV) light exposure. UV radiation induces a mutation of the tumor suppressor gene <span class="html-italic">tp53</span>, which is considered to be one cause of the development of AK. As a result, atypical keratinocytes proliferate in an uncontrolled manner, and apoptosis is reduced. AK is regarded as a potential precursor of squamous cell carcinoma (SCC). (red arrow: reduction; green arrow: induction). Created with BioRender.com.</p>
Full article ">Figure 2
<p>(<b>a</b>) Metabolic activity curves of cutaneous SCC cell lines (SCL-1, SCL-2, A431), oral SCC cell lines (SCC-9, SCC-15, SCC-111), and HEKn cells were generated 24 h after ALA incubation for 3 h followed by red light treatment (100 J/cm<sup>2</sup>; 160 mW/cm<sup>2</sup>). The ALA concentration at which approximately 70–80% of cells are metabolically active was defined as the MA<sub>75</sub> value and served as the individual ALA concentration for cell culture examinations. (<b>b</b>) The numbers of living cells of (a–c) cutaneous SCC cells (SCL-1, SCL-2, A431), (d–f) oral SCC cells (SCC-9, SCC-15, SCC-111), and (g) HEKn cells were determined 24 h after treatments (CAP, ALA-PDT, CAP-ALA–red light, ALA-CAP–red light) and in untreated cells (ctrl.) using Acridine Orange (AO)/Propidium Iodide (PI) staining and were analyzed using LUNA-FL™ in an automated fluorescence cell counting mode The results are the means of a single experiment performed in duplicate.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) Metabolic activity curves of cutaneous SCC cell lines (SCL-1, SCL-2, A431), oral SCC cell lines (SCC-9, SCC-15, SCC-111), and HEKn cells were generated 24 h after ALA incubation for 3 h followed by red light treatment (100 J/cm<sup>2</sup>; 160 mW/cm<sup>2</sup>). The ALA concentration at which approximately 70–80% of cells are metabolically active was defined as the MA<sub>75</sub> value and served as the individual ALA concentration for cell culture examinations. (<b>b</b>) The numbers of living cells of (a–c) cutaneous SCC cells (SCL-1, SCL-2, A431), (d–f) oral SCC cells (SCC-9, SCC-15, SCC-111), and (g) HEKn cells were determined 24 h after treatments (CAP, ALA-PDT, CAP-ALA–red light, ALA-CAP–red light) and in untreated cells (ctrl.) using Acridine Orange (AO)/Propidium Iodide (PI) staining and were analyzed using LUNA-FL™ in an automated fluorescence cell counting mode The results are the means of a single experiment performed in duplicate.</p>
Full article ">Figure 3
<p>Determination of DCF in arbitrary units [a.u.] as indicator of intracellular ROS levels in untreated SCC and HEKn cells using fluorogenic DCFH-DA assay. Intracellular ROS levels in cutaneous SCC cells (SCL-1, SCL-2, A431) and oral SCC cells (SCC-9, SCC-15, SCC-11) in comparison to non-tumorigenic HEKn cells. The results are the means of three independent measurements. Statistical analysis: Ordinary one-way ANOVA with Tukey’s multiple comparison test was carried out to compare the means of HEKn and SCC cell lines. **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>The determination of DCF as an indicator of intracellular ROS levels in treated SCC and HEKn cells using a fluorogenic DCFH-DA assay. DCF (x-fold to ctrl.) was determined in cutaneous SCC cell lines (<b>a</b>) SCL-1, (<b>b</b>) SCL-2, and (<b>c</b>) A431; in oral SCC cell lines (<b>d</b>) SCC-9, (<b>e</b>) SCC-15 and (<b>f</b>) SCC-111; and in (<b>g</b>) normal HEKn after CAP treatment, ALA-PDT and after combined treatments (CAP-ALA–red light, ALA-CAP–red light). Statistical analysis: Ordinary one-way ANOVA with Bonferroni’s multiple comparison test was carried out to compare the mean of ALA-PDT with the results of the combined CAP-ALA–red light and ALA-CAP–red light treatment. * <span class="html-italic">p</span> ≤ 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Migration and invasion of SCC cells after conventional ALA-PDT and after combined treatments (CAP-ALA–red light, ALA-CAP–red light). (<b>a</b>) Boyden Chamber migration assay. Exemplary overview of stained 8 µm pore filter membranes of migrated cutaneous SCC cells (SCL-1, SCL-2, A431) and oral SCC cells (SCC-9, SCC-15, SCC-111). (<b>b</b>) Boyden Chamber invasion assay. Exemplary overview of stained 8 µm pore filter membranes of matrigel-invaded cutaneous SCC cells (SCL-1, SCL-2, A431) and oral SCC cells (SCC-9, SCC-15, SCC-111). Quantification and statistical examination of (<b>c</b>) migrated and (<b>d</b>) invaded SCC cells. Each experiment was carried out in triplicates. Of each experiment, at least five representative images per group were taken at 20-fold magnification, and the cells per field of view were counted and summarized. Statistical analysis: Ordinary one-way ANOVA with Bonferroni’s multiple comparison test was used with * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001 to indicate the mean differences within the conventional ALA-PDT and the combined treatments (CAP-ALA–red light, ALA-CAP–red light).</p>
Full article ">Figure 5 Cont.
<p>Migration and invasion of SCC cells after conventional ALA-PDT and after combined treatments (CAP-ALA–red light, ALA-CAP–red light). (<b>a</b>) Boyden Chamber migration assay. Exemplary overview of stained 8 µm pore filter membranes of migrated cutaneous SCC cells (SCL-1, SCL-2, A431) and oral SCC cells (SCC-9, SCC-15, SCC-111). (<b>b</b>) Boyden Chamber invasion assay. Exemplary overview of stained 8 µm pore filter membranes of matrigel-invaded cutaneous SCC cells (SCL-1, SCL-2, A431) and oral SCC cells (SCC-9, SCC-15, SCC-111). Quantification and statistical examination of (<b>c</b>) migrated and (<b>d</b>) invaded SCC cells. Each experiment was carried out in triplicates. Of each experiment, at least five representative images per group were taken at 20-fold magnification, and the cells per field of view were counted and summarized. Statistical analysis: Ordinary one-way ANOVA with Bonferroni’s multiple comparison test was used with * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001 to indicate the mean differences within the conventional ALA-PDT and the combined treatments (CAP-ALA–red light, ALA-CAP–red light).</p>
Full article ">Figure 5 Cont.
<p>Migration and invasion of SCC cells after conventional ALA-PDT and after combined treatments (CAP-ALA–red light, ALA-CAP–red light). (<b>a</b>) Boyden Chamber migration assay. Exemplary overview of stained 8 µm pore filter membranes of migrated cutaneous SCC cells (SCL-1, SCL-2, A431) and oral SCC cells (SCC-9, SCC-15, SCC-111). (<b>b</b>) Boyden Chamber invasion assay. Exemplary overview of stained 8 µm pore filter membranes of matrigel-invaded cutaneous SCC cells (SCL-1, SCL-2, A431) and oral SCC cells (SCC-9, SCC-15, SCC-111). Quantification and statistical examination of (<b>c</b>) migrated and (<b>d</b>) invaded SCC cells. Each experiment was carried out in triplicates. Of each experiment, at least five representative images per group were taken at 20-fold magnification, and the cells per field of view were counted and summarized. Statistical analysis: Ordinary one-way ANOVA with Bonferroni’s multiple comparison test was used with * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001 to indicate the mean differences within the conventional ALA-PDT and the combined treatments (CAP-ALA–red light, ALA-CAP–red light).</p>
Full article ">Figure 6
<p>Annexin V/PI double-staining assay was performed 48 h after different treatments (CAP, ALA-PDT, CAP-ALA–red light, and ALA-CAP–red light) and compared to untreated control cells (ctrl.). Apoptosis (early and late apoptosis), necrosis, and the number of live cells were analyzed in oral SCC cell lines (<b>a</b>) SCC-9, (<b>b</b>) SCC-15 and (<b>c</b>) SCC-111; cutaneous SCC cell lines (<b>d</b>) SCL-1, (<b>e</b>) SCL-2 and (<b>f</b>) A431; and (<b>g</b>) HEKn cells. The graphs present the percentage (mean ± SD) of the cells in the region among the total cells from three independent experiments in duplicate. Statistical analysis: Ordinary one-way ANOVA with Bonferroni’s multiple comparison test was used with * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001 to indicate the mean differences within the conventional ALA-PDT and the combined treatment groups (CAP-ALA–red light and ALA-CAP–red light).</p>
Full article ">Figure 7
<p>Expression of apoptosis- and senescence-related molecules on the mRNA level after different treatments (CAP, ALA-PDT, CAP-ALA–red light and ALA-CAP–red light) and in untreated cells (ctrl.). (<b>a</b>–<b>c</b>) mRNA expression (caspase-3, -7, -9, p16<sup>INK4a</sup>, p21<sup>CIP1</sup>) in cutaneous SCC cell lines (SCL-1, SCL-2, A431), (<b>d</b>–<b>f</b>) in oral SCC cell lines, and (<b>g</b>) HEKn was measured 24 h after treatment. Statistical analysis: Ordinary one-way ANOVA with Bonferroni’s multiple comparison test was carried out to compare the mean of conventional ALA-PDT with combined treatments (CAP-ALA–red light, ALA–CAP-red light). * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 8
<p>Expression of genes of the junctional network after different treatments (CAP, ALA-PDT, CAP-ALA–red light, or ALA-CAP–red light) and in untreated cells (ctrl.). (<b>a</b>–<b>c</b>) mRNA expression (CLDN1, ZO-1, CTNNB1, CX31) in cutaneous SCC cell lines (SCL-1, SCL-2, A431), (<b>d</b>–<b>f</b>) in oral SCC cell lines, and (<b>g</b>) in HEKn was measured 24 h after treatment. Statistical analysis: Ordinary one-way ANOVA with Bonferroni’s multiple comparison test was carried out to compare the mean of conventional ALA-PDT with combined treatments (CAP-ALA–red light, ALA-CAP–red light). * <span class="html-italic">p</span> ≤ 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 8 Cont.
<p>Expression of genes of the junctional network after different treatments (CAP, ALA-PDT, CAP-ALA–red light, or ALA-CAP–red light) and in untreated cells (ctrl.). (<b>a</b>–<b>c</b>) mRNA expression (CLDN1, ZO-1, CTNNB1, CX31) in cutaneous SCC cell lines (SCL-1, SCL-2, A431), (<b>d</b>–<b>f</b>) in oral SCC cell lines, and (<b>g</b>) in HEKn was measured 24 h after treatment. Statistical analysis: Ordinary one-way ANOVA with Bonferroni’s multiple comparison test was carried out to compare the mean of conventional ALA-PDT with combined treatments (CAP-ALA–red light, ALA-CAP–red light). * <span class="html-italic">p</span> ≤ 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
23 pages, 2243 KiB  
Review
Combining Photodynamic Therapy and Targeted Drug Delivery Systems: Enhancing Mitochondrial Toxicity for Improved Cancer Outcomes
by J. P. Jose Merlin, Anine Crous and Heidi Abrahamse
Int. J. Mol. Sci. 2024, 25(19), 10796; https://doi.org/10.3390/ijms251910796 - 8 Oct 2024
Viewed by 1105
Abstract
Cancer treatment continues to be a substantial problem due to tumor complexities and persistence, demanding novel therapeutic techniques. This review investigates the synergistic potential of combining photodynamic therapy (PDT) and tailored medication delivery technologies to increase mitochondrial toxicity and improve cancer outcomes. PDT [...] Read more.
Cancer treatment continues to be a substantial problem due to tumor complexities and persistence, demanding novel therapeutic techniques. This review investigates the synergistic potential of combining photodynamic therapy (PDT) and tailored medication delivery technologies to increase mitochondrial toxicity and improve cancer outcomes. PDT induces selective cellular damage and death by activating photosensitizers (PS) with certain wavelengths of light. However, PDT’s efficacy can be hampered by issues such as poor light penetration and a lack of selectivity. To overcome these challenges, targeted drug delivery systems have emerged as a promising technique for precisely delivering therapeutic medicines to tumor cells while avoiding off-target effects. We investigate how these technologies can improve mitochondrial targeting and damage, which is critical for causing cancer cell death. The combination method seeks to capitalize on the advantages of both modalities: selective PDT activation and specific targeted drug delivery. We review current preclinical and clinical evidence supporting the efficacy of this combination therapy, focusing on case studies and experimental models. This review also addresses issues such as safety, distribution efficiency, resistance mechanisms, and costs. The prospects of further research include advances in photodynamic agents and medication delivery technology, with a focus on personalized treatment. In conclusion, combining PDT with targeted drug delivery systems provides a promising frontier in cancer therapy, with the ability to overcome current treatment limits and open the way for more effective, personalized cancer treatments. Full article
Show Figures

Figure 1

Figure 1
<p>The mechanism of photodynamic therapy (PDT). Light absorption allows the photocatalyst to transition from a singlet energy state (S<sup>0</sup>) to an excited singlet state (S<sup>1</sup>) when exposed to light with a wavelength corresponding to the photothermal absorption (PS). The stored energy in the PS decays to a useful state, the excited triplet state (T<sup>1</sup>), but some of the energy is dissipated as an emission quantum. Cancer cells produce reactive oxygen species due to PDT, using light, oxygen, and PS to induce apoptosis. Meaning of the different colored arrows: green—Electronic transition, cyan—Fluorescence, blue—Internal convention, green—Intersystem crossing, purple—Phosphorescence.</p>
Full article ">Figure 2
<p>The mechanism of apoptosis. The primary stage of the apoptotic process is driven by mitochondria, which permits a variety of apoptogens to approach the cytosol and initiate procaspases. When cytochrome c (cyt c) is released, a multiprotein complex called the apoptosome is formed. This component cleaves procaspase-9 to activate caspase-9, which in turn stimulates the production of other inhibitors of apoptosis. To release cyt c and complex 1 (Apaf-1) from mitochondria and form the procaspase-9 complex that activates caspase-9, the tumor suppressor protein (p53) is also involved in the emergence of the mitochondrial apoptosome. In contrast, Bcl-2 can bind to Bax/Bak and sequester BH3 proteins, which inhibits this process of cell death and promotes cell survival.</p>
Full article ">Figure 3
<p>The mechanism of autophagy. The activation of the photosensitizer in the mitochondria results in a decrease in ATP synthesis and an increase in reactive oxygen species production. The energy-sensing AMPK activates ULK1, which starts autophagy. Photodynamic therapy can also activate NFκB, an autophagy mechanism involved in protein, lipid, and nucleotide production, to initiate lysosome formation and autophagy. Autophagy can be transcriptionally controlled by mitochondrial photooxidation via MAPK, CHOP, and HIF-1α.</p>
Full article ">
9 pages, 232 KiB  
Article
Effect of Clinicopathological Characteristics on the Outcomes of Topical 5-Aminolevulinic Acid Photodynamic Therapy in Patients with Cervical High-Grade Squamous Intraepithelial Lesions (HSIL/CIN2): A Retrospective Cohort Study
by Yingting Wei, Jing Niu, Liying Gu, Zubei Hong, Zhouzhou Bao and Lihua Qiu
Biomedicines 2024, 12(10), 2255; https://doi.org/10.3390/biomedicines12102255 - 3 Oct 2024
Viewed by 648
Abstract
Background: Minimally-invasive 5-aminolevulinic acid photodynamic therapy (ALA-PDT) is used for treating cervical high-grade squamous intraepithelial lesions (HSIL/CIN2). The purpose of this study was to analyze the factors affecting the efficacy of ALA-PDT in the treatment of cervical HSIL/CIN2 in order to guide physicians [...] Read more.
Background: Minimally-invasive 5-aminolevulinic acid photodynamic therapy (ALA-PDT) is used for treating cervical high-grade squamous intraepithelial lesions (HSIL/CIN2). The purpose of this study was to analyze the factors affecting the efficacy of ALA-PDT in the treatment of cervical HSIL/CIN2 in order to guide physicians in making appropriate treatment decisions. Methods: A retrospective study including 69 female patients with pathologically diagnosed HSIL/CIN2 was conducted. Patients were given six doses of 20% ALA-PDT at 7–14-day intervals. Cytology, HPV testing, colposcopy, and pathology were performed before treatment and at 6-month follow-up after treatment to assess efficacy. The main outcome of this study was the regression of HSIL/CIN2 and the clearance of high-risk HPV (hrHPV) infection after ALA-PDT treatment. Clinicopathological characteristics were collected to analyze the factors affecting the effectiveness of ALA-PDT treatment for HSIL/CIN2. Results: Between the successful and failed lesion regression group, there was a significant difference in sleeping disorders (p < 0.05). Between the successful and failed hrHPV clearance group, no statistically significant factors were found. With sensitivity values of 0.556 and 0.778, respectively, multivariate analysis showed that current smoking and sleeping disorders were independent prognostics of failure in lesion regression after ALA-PDT treatment. Conclusions: Smoking and sleep disorders were independent risk factors for failure in HSIL/CIN2 regression following ALA-PDT, suggesting the need for careful consideration of ALA-PDT for patients with these conditions. Full article
(This article belongs to the Special Issue Photodynamic Therapy (3rd Edition))
16 pages, 2654 KiB  
Article
Effects of Zinc Phthalocyanine Photodynamic Therapy on Vital Structures and Processes in Hela Cells
by Jakub Hosik, Barbora Hosikova, Svatopluk Binder, Rene Lenobel, Marketa Kolarikova, Lukas Malina, Hanna Dilenko, Katerina Langova, Robert Bajgar and Hana Kolarova
Int. J. Mol. Sci. 2024, 25(19), 10650; https://doi.org/10.3390/ijms251910650 - 3 Oct 2024
Viewed by 355
Abstract
This work presents results on the efficiency of newly designed zinc phthalocyanine-mediated photodynamic therapy of both tumoral and nontumoral cell models using the MTT assay. Further detailed examinations of mechanistic and cell biological effects were focused on the HELA cervical cancer cell model. [...] Read more.
This work presents results on the efficiency of newly designed zinc phthalocyanine-mediated photodynamic therapy of both tumoral and nontumoral cell models using the MTT assay. Further detailed examinations of mechanistic and cell biological effects were focused on the HELA cervical cancer cell model. Here, ROS production, changes in the mitochondrial membrane potential, the determination of genotoxicity, and protein changes determined by capillary chromatography and tandem mass spectrometry with ESI were analyzed. The results showed that, in vitro, 5 Jcm−2 ZnPc PDT caused a significant increase in reactive oxygen species. Still, except for superoxide dismutase, the levels of proteins involved in cell response to oxidative stress did not increase significantly. Furthermore, this therapy damaged mitochondrial membranes, which was proven by a more than 70% voltage-dependent channel protein 1 level decrease and by a 65% mitochondrial membrane potential change 24 h post-therapy. DNA impairment was assessed by an increased level of DNA fragmentation, which might be related to the decreased level of DDB1 (decrease in levels of more than 20% 24 h post-therapy), a protein responsible for maintaining genomic integrity and triggering the DNA repair pathways. Considering these results and the low effective concentration (LC50 = 30 nM), the therapy used is a potentially very promising antitumoral treatment. Full article
(This article belongs to the Special Issue Photodynamic Therapy and Photodetection, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>The dependence of tumoral and nontumoral cell viability on the concentrations of the zinc phthalocyanine photosensitizer. The dependence of the cellular viability on the concentration of the zinc phthalocyanine photosensitizer was determined by measuring the enzyme activity of living cells using the MTT test. The irradiation dose used was 5 Jcm<sup>−2</sup>. The control represents the irradiated cells without the photosensitizer (the negative control), and its value is set at 100%. Data are presented as ±SD from three independent measurements. The results were considered statistically significant when <span class="html-italic">p</span> &lt; 0.05 and indicated by an asterisk symbol.</p>
Full article ">Figure 2
<p>ROS production and protein level changes after 5 Jcm<sup>−2</sup> ZnPc photodynamic therapy. The dependence of the HeLa cells’ ROS production on the concentration of the zinc phthalocyanine photosensitizer measured immediately after irradiation: (<b>A</b>) singlet oxygen production; (<b>B</b>) ROS production after type I of PDT reaction. The control represents the irradiated cells without the photosensitizer (the negative control). Data are presented as ±SD from three independent measurements. Results were considered statistically significant when <span class="html-italic">p</span> &lt; 0.05 and indicated in the graphs by an asterisk symbol. (<b>C</b>) Changes in the proteins involved in oxidation stress reduction depending on incubation time after the therapy. The relative quantification was calculated as the proportion of protein in the control cell sample compared to the protein level in the treated cell sample. P32119 (UniProt identificator)—peroxiredoxin 2, P30044—peroxiredoxin 5, P30041—peroxiredoxin 6, P00441—superoxide dismutase [Cu-Zn], P10599—thioredoxin.</p>
Full article ">Figure 3
<p>Changes at the mitochondrial level after 5 Jcm<sup>-2</sup> ZnPc PDT. (<b>A</b>) Protein level changes depending on incubation time after the therapy (OPA1 (UniProt ID O60313—dynamin-like 120 kDa protein); VDAC (P21796—voltage-dependent anion-selective channel protein 1)). The relative quantification was calculated as the proportion of protein in the control cell sample compared to the protein level in the treated cell sample. (<b>B</b>) The change in HeLa cells’ mitochondria membrane potential using LC50 concentration of zinc phthalocyanine photosensitizer. The control represents the irradiated cells without a photosensitizer (the negative control). Data are presented as ±SD from three independent measurements. Results were considered statistically significant when <span class="html-italic">p</span> &lt; 0.05 and indicated in the graphs by an asterisk symbol.</p>
Full article ">Figure 4
<p>Changes at the DNA level after 5 Jcm<sup>−2</sup> ZnPc PDT. (<b>A</b>) Changes in the protein level depending on the incubation period after the therapy. The relative quantification was calculated as the proportion of protein in the control cell sample compared to the protein level in the treated cell sample. Q16531—DNA damage-binding protein 1; P09874—poly [ADP-ribose] polymerase 1; P12004—proliferating cell nuclear antigen. (<b>B</b>) DNA damage evaluated 24 h after therapy by comet assay using LC50 concentration of zinc phthalocyanine photosensitizer. The control represents the irradiated cells without a photosensitizer (the negative control).). Data are presented as ±SD from three independent measurements. Results were considered statistically significant when <span class="html-italic">p</span> &lt; 0.05 and indicated in the graphs by an asterisk symbol.</p>
Full article ">Figure 5
<p>The seven largest groups of the degraded proteins in terms of cell localization (G0 terms) identified in samples with 0, 4, and 24 h incubation time after in vitro 5 Jcm<sup>−2</sup> ZnPc PDT at the LC50 concentration on HeLa cells. Degraded proteins were identified by comparing the measured sample to the control cells without photosensitizer and irradiation. The proteins identified were sorted using David Bioinformatic Resources 6.8 when <span class="html-italic">p</span> &lt; 0.0001 [<a href="#B44-ijms-25-10650" class="html-bibr">44</a>,<a href="#B45-ijms-25-10650" class="html-bibr">45</a>].</p>
Full article ">Figure 6
<p>The six largest groups of the degraded proteins in terms of molecular function (G0 terms) identified 24 h after in vitro 5 Jcm<sup>−2</sup> ZnPc PDT at the LC50 concentration on HeLa cells. Degraded proteins were identified by comparing the measured sample to the control cells without photosensitizer and irradiation and sorted using David Bioinformatic Resources 6.8 when <span class="html-italic">p</span> &lt; 0.0001 [<a href="#B44-ijms-25-10650" class="html-bibr">44</a>,<a href="#B45-ijms-25-10650" class="html-bibr">45</a>].</p>
Full article ">Figure 7
<p>The dependence of tumoral HeLa and nontumoral BJ cells’ viability on the concentrations of the zinc phthalocyanine photosensitizer after sensitizer incorporation into the liposomes. The cellular viability was determined by measuring the enzyme activity of living cells using the MTT test. The irradiation dose used was 5 Jcm<sup>−2</sup>. The liposomal content was released from liposomes using ultrasound (3 Wcm<sup>−2</sup>, 60 s, pulse mode). The control represents the irradiated and sonicated cells without the photosensitizer (the negative control). Data are presented as ±SD from three independent measurements.</p>
Full article ">Figure 8
<p>The chemical structure of the zinc phthalocyanine (ZnPc) used in the study. The chemical formula of ZnPc: 2,3,9,10,16,17,23,24-Octakis[(2-(triethylammonio)ethyl)sulfanyl]phthalocyaninato]zinc(II) Octaiodide [<a href="#B52-ijms-25-10650" class="html-bibr">52</a>].</p>
Full article ">
28 pages, 1342 KiB  
Review
Photodynamic Therapy Review: Past, Present, Future, Opportunities and Challenges
by Yaran Allamyradov, Justice ben Yosef, Berdimyrat Annamuradov, Mahmood Ateyeh, Carli Street, Hadley Whipple and Ali Oguz Er
Photochem 2024, 4(4), 434-461; https://doi.org/10.3390/photochem4040027 - 1 Oct 2024
Viewed by 532
Abstract
Photodynamic therapy (PDT) is a medical treatment that utilizes photosensitizing agents, along with light, to produce reactive oxygen species that can kill nearby cells. When the photosensitizer is exposed to a specific wavelength of light, it becomes activated and generates reactive oxygen that [...] Read more.
Photodynamic therapy (PDT) is a medical treatment that utilizes photosensitizing agents, along with light, to produce reactive oxygen species that can kill nearby cells. When the photosensitizer is exposed to a specific wavelength of light, it becomes activated and generates reactive oxygen that can destroy cancer cells, bacteria, and other pathogenic micro-organisms. PDT is commonly used in dermatology for treating actinic keratosis, basal cell carcinoma, and other skin conditions. It is also being explored for applications in oncology, such as treating esophageal and lung cancers, as well as in ophthalmology for age-related macular degeneration. In this study, we provide a comprehensive review of PDT, covering its fundamental principles and mechanisms, as well as the critical components for its function. We examine key aspects of PDT, including its current clinical applications and potential future developments. Additionally, we discuss the advantages and disadvantages of PDT, addressing the various challenges associated with its implementation and optimization. This review aims to offer a thorough understanding of PDT, highlighting its transformative potential in medical treatments while acknowledging the areas requiring further research and development. Full article
Show Figures

Figure 1

Figure 1
<p>PDT applications.</p>
Full article ">Figure 2
<p>Jablonski diagram for a photosensitizer compound illustrating the PDT mechanism.</p>
Full article ">Figure 3
<p>Chemical structure of porphyrin (C<sub>20</sub>H<sub>14</sub>N<sub>4</sub>).</p>
Full article ">Figure 4
<p>Different light sources for PDT.</p>
Full article ">
20 pages, 688 KiB  
Systematic Review
Hydrogels Associated with Photodynamic Therapy Have Antimicrobial Effect against Staphylococcus aureus: A Systematic Review
by Ricardo S. Moura, João Pedro R. Afonso, Diego A. C. P. G. Mello, Renata Kelly Palma, Iransé Oliveira-Silva, Rodrigo F. Oliveira, Deise A. A. P. Oliveira, Dante B. Santos, Carlos Hassel M. Silva, Orlando A. Guedes, Giuseppe Insalaco and Luís V. F. Oliveira
Gels 2024, 10(10), 635; https://doi.org/10.3390/gels10100635 - 30 Sep 2024
Viewed by 427
Abstract
Staphylococcus aureus (S. aureus) is a Gram-positive bacterium that causes infections ranging from mild superficial cases to more severe, potentially fatal conditions. Many photosensitisers used in photodynamic therapy are more effective against superficial infections due to limitations in treating deeper tissue [...] Read more.
Staphylococcus aureus (S. aureus) is a Gram-positive bacterium that causes infections ranging from mild superficial cases to more severe, potentially fatal conditions. Many photosensitisers used in photodynamic therapy are more effective against superficial infections due to limitations in treating deeper tissue infections. Recently, attention to this bacterium has increased due to the emergence of multidrug-resistant strains, which complicate antibiotic treatment. As a result, alternative therapies, such as antimicrobial photodynamic therapy (PDT), have emerged as promising options for treating non-systemic infections. PDT combines a photosensitiser (PS) with light and oxygen to generate free radicals that destroy bacterial structures. This systematic review evaluates the effectiveness of PDT delivered via different types of hydrogels in treating wounds, burns, and contamination by S. aureus. Following PRISMA 2020 guidelines, a bibliographic search was conducted in PubMed, Web of Science, and Scopus databases, including articles published in English between 2013 and 2024. Seven relevant studies were included, demonstrating evidence of PDT use against S. aureus in in vitro and in vivo studies. We concluded that PDT can effectively complement antimicrobial therapy in the healing of wounds and burns. The effectiveness of this technique depends on the PS used, the type of hydrogel, and the lesion location. However, further in vivo studies are needed to confirm the safety and efficacy of PDT delivered via hydrogels. Full article
Show Figures

Figure 1

Figure 1
<p>Flow diagram of the current systematic review conducted according to the Preferred Reporting Items for Systematic Reviews and Meta-analysis (PRISMA) guidelines.</p>
Full article ">Figure 2
<p>Mechanism of photodynamic therapy combined with hydrogel for combatting <span class="html-italic">Staphylococcus aureus</span>. Created with BioRender.com.</p>
Full article ">
44 pages, 5021 KiB  
Review
Capped Plasmonic Gold and Silver Nanoparticles with Porphyrins for Potential Use as Anticancer Agents—A Review
by Nthabeleng Hlapisi, Sandile P. Songca and Peter A. Ajibade
Pharmaceutics 2024, 16(10), 1268; https://doi.org/10.3390/pharmaceutics16101268 - 28 Sep 2024
Viewed by 669
Abstract
Photothermal therapy (PTT) and photodynamic therapy (PDT) are potential cancer treatment methods that are minimally invasive with high specificity for malignant cells. Emerging research has concentrated on the application of metal nanoparticles encapsulated in porphyrin and their derivatives to improve the efficacy of [...] Read more.
Photothermal therapy (PTT) and photodynamic therapy (PDT) are potential cancer treatment methods that are minimally invasive with high specificity for malignant cells. Emerging research has concentrated on the application of metal nanoparticles encapsulated in porphyrin and their derivatives to improve the efficacy of these treatments. Gold and silver nanoparticles have distinct optical properties and biocompatibility, which makes them efficient materials for PDT and PTT. Conjugation of these nanoparticles with porphyrin derivatives increases their light absorption and singlet oxygen generation that create a synergistic effect that increases phototoxicity against cancer cells. Porphyrin encapsulation with gold or silver nanoparticles improves their solubility, stability, and targeted tumor delivery. This paper provides comprehensive review on the design, functionalization, and uses of plasmonic silver and gold nanoparticles in biomedicine and how they can be conjugated with porphyrins for synergistic therapeutic effects. Furthermore, it investigates this dual-modal therapy’s potential advantages and disadvantages and offers perspectives for future prospects. The possibility of developing gold, silver, and porphyrin nanotechnology-enabled biomedicine for combination therapy is also examined. Full article
(This article belongs to the Section Drug Targeting and Design)
Show Figures

Figure 1

Figure 1
<p>Different gold nanoparticles and their sizes.</p>
Full article ">Figure 2
<p>Porphyrins used for the potential treatment of cancer: (<b>a</b>) 5,10,15,20 tetrakis(pyridyl)porphyrin; (<b>b</b>) meso-tetrakis-(4-sulfonatophenyl)porphyrin; (<b>c</b>) tetra-3-carboxyphenyl porphyrin; (<b>d</b>) meso-tetrakis-(morpholine)porphyrin; (<b>e</b>) 5,10,15,20 tetrakis(dimethylaniline)porphyrin; (<b>f</b>) Tetrakis(benzene-1,2-diol)porphyrin.</p>
Full article ">Figure 3
<p>Singlet oxygen production by the photosensitizing process [<a href="#B62-pharmaceutics-16-01268" class="html-bibr">62</a>].</p>
Full article ">Figure 4
<p>Mechanism of PTT.</p>
Full article ">Figure 5
<p>Steps generally used for functionalizing nanoparticles.</p>
Full article ">Figure 6
<p>The effect of AuNP size on their biological behavior. Smaller-sized nanoparticles result in a high clearance from the liver and the blood–brain barrier.</p>
Full article ">Figure 7
<p>Scheme that shows how a photosensitizer is activated and the production of singlet oxygen leading to cell death.</p>
Full article ">Figure 8
<p>Porphyrin surface-protected ligands (<b>a</b>) and (<b>b</b>) ligand.</p>
Full article ">Figure 9
<p>Porphyrin–alkanethiols. (<b>a</b>) tetraphenyl porphyrin-5-yl) phenyl) acetamide, (<b>b</b>), 5,10,15 3,5-di-tert-butyl(phenyl) porphyrin-5-yl)phenyl)acetamide.</p>
Full article ">Figure 10
<p>Porphyrin–alkanethiol post-synthesis loading on 1-dodecanethiolate-protected AuNPs.</p>
Full article ">Figure 11
<p>Illustration of Au@PH2TPP_SAM.</p>
Full article ">Figure 12
<p>(<b>a</b>) Chemical structures of poly(isobutylene-alt-maleic-anhydrite) and (<b>b</b>) poly(isobutylene-alt-maleic-acid ammonium salt).</p>
Full article ">
20 pages, 3354 KiB  
Article
Iron Metabolism in Aminolevulinic Acid-Photodynamic Therapy with Iron Chelators from the Thiosemicarbazone Group
by Robert Gawecki, Patrycja Rawicka, Marta Rogalska, Maciej Serda and Anna Mrozek-Wilczkiewicz
Int. J. Mol. Sci. 2024, 25(19), 10468; https://doi.org/10.3390/ijms251910468 - 28 Sep 2024
Viewed by 323
Abstract
Iron plays a crucial role in various metabolic processes. However, the impact of 5-aminolevulinic acid (ALA) in combination with iron chelators on iron metabolism and the efficacy of ALA-photodynamic therapy (PDT) remain inadequately understood. This study aimed to examine the effect of thiosemicarbazone [...] Read more.
Iron plays a crucial role in various metabolic processes. However, the impact of 5-aminolevulinic acid (ALA) in combination with iron chelators on iron metabolism and the efficacy of ALA-photodynamic therapy (PDT) remain inadequately understood. This study aimed to examine the effect of thiosemicarbazone derivatives during ALA treatment on specific genes related to iron metabolism, with a particular emphasis on mitochondrial iron metabolism genes. In our study, we observed differences depending on the cell line studied. For the HCT116 and MCF-7 cell lines, in most cases, the decrease in the expression of selected targets correlated with the increase in protoporphyrin IX (PPIX) concentration and the observed photodynamic effect, aligning with existing literature data. The Hs683 cell line showed a different gene expression pattern, previously not described in the literature. In this study, we collected an extensive analysis of the gene variation occurring after the application of novel thiosemicarbazone derivatives and presented versatile and effective compounds with great potential for use in ALA-PDT. Full article
(This article belongs to the Section Molecular Endocrinology and Metabolism)
Show Figures

Figure 1

Figure 1
<p>Iron metabolism in a cell considering the molecular targets studied in this paper (created in BioRender). PPIX—protoporphyrin IX, FECH—ferrochelatase, HO-1—heme oxygenase, FTMT—ferritin, FXN—frataxin, ICSU—iron-sulfur cluster assembly enzyme, MFRN1/2—mitoferrin 1/2, ABCB8—ATP-binding cassette subfamily B member 8.</p>
Full article ">Figure 2
<p>The structures of the TSCs and Cp94 examined in this study.</p>
Full article ">Figure 3
<p>Accumulation of PPIX after treatment with 5-ALA, TSCs, Cp94, and their combinations on (<b>A</b>): HCT116, (<b>B</b>): MCF-7, and (<b>C</b>): Hs683 cell lines. The red dotted line on the graph depicts the fluorescence intensity of PPIX following treatment with 5-ALA alone. The data are presented as the means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Absorption titration spectra of selected compounds (50 μM) with iron (III) ions in water: (<b>A</b>)—Cp94; (<b>B</b>)—TSC-34; (<b>C</b>)—TSC-109; (<b>D</b>)—TSC-113; (<b>E</b>)—TSC-116. Blue arrows indicate a decrease in band intensity for the test compound, while red arrows highlight an increase in band intensity for the complex. Measurements were performed at room temperature 4 h after the samples had been prepared.</p>
Full article ">Figure 5
<p>Phototoxic effect after treatment with 5-ALA, TSCs, Cp94, and their combinations on (<b>A</b>): HCT116, (<b>B</b>): MCF-7, and (<b>C</b>): Hs683 cell lines. The data are presented as means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Expression of FECH in the (<b>A</b>): HCT116, (<b>B</b>): MCF-7, and (<b>C</b>): Hs683 cell lines. The data are presented as means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Expression of HO-1 in the (<b>A</b>): HCT116, (<b>B</b>): MCF-7, and (<b>C</b>): Hs683 cell lines. The data are presented as means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 8
<p>Expression of FTMT in the (<b>A</b>): HCT116, (<b>B</b>): MCF-7, and (<b>C</b>): Hs683 cell lines. The data are presented as means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 9
<p>Expression of FXN in the (<b>A</b>): HCT116, (<b>B</b>): MCF-7, and (<b>C</b>): Hs683 cell lines. The data are presented as means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 10
<p>Expression of MFRN 1 and 2 genes on the (<b>A</b>,<b>D</b>): HCT116, (<b>B</b>,<b>E</b>): MCF-7 and (<b>C</b>,<b>F</b>): Hs683 cell lines. The data are presented as means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 11
<p>Expression of DMT1 in the (<b>A</b>): HCT116, (<b>B</b>): MCF-7, and (<b>C</b>): Hs683 cell lines. The data are presented as means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 12
<p>Expression of ABCB8 in the (<b>A</b>): HCT116, (<b>B</b>): MCF-7, and (<b>C</b>): Hs683 cell lines. The data are presented as means ± standard deviation from three independent experiments and analyzed using one-way ANOVA with Tukey’s post hoc test, indicating significance levels as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
16 pages, 4847 KiB  
Article
Activity of Hydrophilic, Biocompatible, Fluorescent, Organic Nanoparticles Functionalized with Purpurin-18 in Photodynamic Therapy for Colorectal Cancer
by Rayan Chkair, Justine Couvez, Frédérique Brégier, Mona Diab-Assaf, Vincent Sol, Mireille Blanchard-Desce, Bertrand Liagre and Guillaume Chemin
Nanomaterials 2024, 14(19), 1557; https://doi.org/10.3390/nano14191557 - 26 Sep 2024
Viewed by 572
Abstract
Photodynamic therapy (PDT) is a clinically approved, non-invasive therapy currently used for several solid tumors, triggering cell death through the generation of reactive oxygen species (ROS). However, the hydrophobic nature of most of the photosensitizers used, such as chlorins, limits the overall effectiveness [...] Read more.
Photodynamic therapy (PDT) is a clinically approved, non-invasive therapy currently used for several solid tumors, triggering cell death through the generation of reactive oxygen species (ROS). However, the hydrophobic nature of most of the photosensitizers used, such as chlorins, limits the overall effectiveness of PDT. To address this limitation, the use of nanocarriers seems to be a powerful approach. From this perspective, we have recently developed water-soluble and biocompatible, fluorescent, organic nanoparticles (FONPs) functionalized with purpurin-18 and its derivative, chlorin p6 (Cp6), as new PDT agents. In this study, we aimed to investigate the induced cell death mechanism mediated by these functionalized nanoparticles after PDT photoactivation. Our results show strong phototoxic effects of the FONPs[Cp6], mediated by intracellular ROS generation, and subcellular localization in HCT116 and HT-29 human colorectal cancer (CRC) cells. Additionally, we proved that, post-PDT, the FONPs[Cp6] induce apoptosis via the intrinsic mitochondrial pathway, as shown by the significant upregulation of the Bax/Bcl-2 ratio, the activation of caspases 9, 3, and 7, leading poly-ADP-ribose polymerase (PARP-1) cleavage, and DNA fragmentation. Our work demonstrates the photodynamic activity of these nanoparticles, making them promising candidates for the PDT treatment of CRC. Full article
Show Figures

Figure 1

Figure 1
<p>Phototoxic effects of FONPs on human CRC cell lines. HCT116 (<b>A</b>), and HT-29 (<b>B</b>) cells were seeded in 96-well plates for 24 h, then cells were treated or not with FONPs at different concentrations for 24 h. Cells were illuminated or not with red light (650 nm, 38 mW/cm<sup>2</sup>). The cytotoxic effects were then monitored 48 h following illumination by MTT assay. Data are shown as mean ± SEM (n = 3); ns: not significant relative to the control group.</p>
Full article ">Figure 2
<p>Cellular internalization of FONPs[Cp6] on HCT116 (<b>A</b>), HT-29 (<b>B</b>), and HEK-293 (<b>C</b>) cell lines. Cells were grown for 24 h in chamber slides coated with a type I collagen (3 mg/mL) with acetic acid (20 mM) gel. Cells were then treated with FONPs[Cp6] at IC<sub>50</sub> concentrations for HCT116 and HT-29 and 1 or 10 µg/mL for HEK-293 cells. After 24 h, red fluorescence was assessed by confocal microscopy (Zeiss LSM880 confocal microscopeJena, Germany). Co-localization was analyzed using the ImageJ software. White scale bars represent 50 µm.</p>
Full article ">Figure 3
<p>Cellular localization of FONPs[Cp6] on HCT116 (<b>A</b>) and HT-29 (<b>B</b>) CRC cell lines. Cells were seeded in chamber slides for 24 h prior to exposure to FONPs[Cp6] at IC<sub>50</sub> concentrations. After 24 h, cells were co-treated with MitoTracker, LysoTracker, or ER-Tracker organelle probes. The images on the right of each panel correspond to the zoomed-in merged images (yellow box). Localization was determined by confocal microscopy, and co-localization was assessed using the ImageJ software. Scale bar = 50 µm.</p>
Full article ">Figure 4
<p>ROS generation by the FONPs[Cp6] in HCT116 (<b>A</b>,<b>C</b>,<b>D</b>) and HT-29 (<b>B</b>,<b>C</b>,<b>E</b>) CRC cell lines. For all experiments, cells were grown for 24 h. Then, cells were treated with the FONPs[Cp6] at IC<sub>50</sub> concentrations and illuminated or not. Intracellular ROS generation was quantified by flow cytometry using DCFDA staining directly after PDT. An increased fluorescence intensity, resulting from increased 2′,7′-dichlorofluorescein (DCF) formation, causes a shift to the right. Hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>) was used as a positive control (<b>A</b>,<b>B</b>). Results of flow cytometry analyses are shown in (<b>C</b>). The confocal microscopy images of DCFDA-labeled HCT116 (<b>D</b>) and HT-29 (<b>E</b>) CRC cells were captured immediately after PDT illumination. Scale bar = 20 µm. Data are represented as mean ± SEM (n = 3). ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001, relative to the control.</p>
Full article ">Figure 5
<p>Apoptosis effects of the FONPs[Cp6] on HCT116 (<b>A</b>,<b>C</b>) and HT-29 (<b>B</b>,<b>D</b>) CRC cells. Cells were grown 24 h before exposure or not to the FONPs[Cp6] at IC<sub>50</sub> concentrations. Then, cells were illuminated or not. Twenty-four and forty-eight hours after PDT illumination, cells were stained with Annexin V-FITC and PI, and cell apoptosis was measured by flow cytometry. (<b>A</b>,<b>B</b>) The results are represented in a scatter plot as four quadrants: living (lower left), early apoptotic (lower right), late apoptotic (upper right), and necrotic (upper left) cells. (<b>C</b>,<b>D</b>) The percentage of each cell population is expressed as mean ± SEM (n = 3). A t-test was used to compare each population in the treated group to its corresponding population in the control. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, relative to the control.</p>
Full article ">Figure 6
<p>Effects of photoactivation of the FONPs[Cp6] on caspases-3/7 activity and protein expression in HCT116 (<b>A</b>,<b>C</b>,<b>D</b>) and HT-29 (<b>B</b>,<b>E</b>,<b>F</b>) CRC cells. Cells were cultured for 24 h, then treated with the FONPs[Cp6] at IC<sub>50</sub> concentrations for 24 h, then illuminated. (<b>A</b>,<b>B</b>) Quantification of caspases-3/7 activity. Directly after PDT of cells co-treated with caspases-3/7 green reagent, fluorescence over time was monitored by Incucyte S3 live-cell imaging system. The ratio of green fluorescent cells to total cells was quantified using IncuCytesoftware version 2023A Rev1). (<b>C</b>–<b>F</b>) The expression of the key apoptotic proteins was assessed by western blotting 24 and 48 h after illumination. β-actin served as a loading control. Bands were quantified by densitometry using ImageJ software, and the Bax/Bcl-2 ratio was then calculated relative to β-actin. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001, relative to the control; ns: not significant.</p>
Full article ">Figure 7
<p>Effects of photoactivation of the FONPs[Cp6] on DNA fragmentation in HCT116 (<b>A</b>) and HT-29 (<b>B</b>) CRC cells. Cells were grown for 24 h before being treated or not with the FONPs[Cp6] at IC<sub>50</sub> concentrations. After 24 h of incubation, cells were then illuminated or not. DNA fragmentation 24 and 48 h post-illumination was quantified from cytosol extracts by ELISA assay. The results are expressed as n-fold compared to the control. Data are expressed as mean ± SEM (n = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, compared to the control, or # <span class="html-italic">p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01, compared to 24 h data.</p>
Full article ">
16 pages, 2597 KiB  
Article
Enhanced Photodynamic Therapy Efficacy through Solid Lipid Nanoparticle of Purpurin-18-N-Propylimide Methyl Ester for Cancer Treatment
by Sooho Yeo, Huiqiang Wu, Il Yoon, Hye-Soo Kim, Young Kyu Song and Woo Kyoung Lee
Int. J. Mol. Sci. 2024, 25(19), 10382; https://doi.org/10.3390/ijms251910382 - 26 Sep 2024
Viewed by 323
Abstract
Photodynamic therapy (PDT) is an innovative cancer treatment that utilizes light. When light irradiates, purpurin-18-N-propylimide methyl ester (P18 N PI ME) generates reactive oxygen species that destroy cancer cells. The hydrophobic nature of P18 N PI ME presents challenges regarding its aggregation in [...] Read more.
Photodynamic therapy (PDT) is an innovative cancer treatment that utilizes light. When light irradiates, purpurin-18-N-propylimide methyl ester (P18 N PI ME) generates reactive oxygen species that destroy cancer cells. The hydrophobic nature of P18 N PI ME presents challenges regarding its aggregation in the body, which can affect its effectiveness. This study aimed to enhance the bioavailability and effectiveness of cancer treatment by synthesizing P18 N PI ME and formulating P18 N PI ME-loaded solid lipid nanoparticles (SLNs). The efficacy of PDT was estimated using the 1,3-diphenylisobenzofuran (DPBF) assay and photocytotoxicity tests on the HeLa (human cervical carcinoma) and A549 (human lung carcinoma) cell lines. The P18 N PI ME-loaded SLNs demonstrated particle sizes in the range of 158.59 nm to 248.43 nm and zeta potentials in the range of –15.97 mV to –28.73 mV. These SLNs exhibited sustained release of P18 N PI ME. DPBF analysis revealed enhanced PDT effects with SLNs containing P18 N PI ME compared with standalone P18 N PI MEs. Photocytotoxicity assays indicated toxicity under light irradiation but no toxicity in the dark. Furthermore, the smallest-sized formulation exhibited the most effective photodynamic activity. These findings indicate the potential of P18 N PI ME-loaded SLNs as promising strategies for PDT in cancer therapy. Full article
(This article belongs to the Special Issue Anticancer Drug Discovery Based on Natural Products)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Schematic representation of the synthesis of P18 N PI ME from P18ME; (<b>B</b>) <sup>1</sup>H-NMR spectrum of P18 N PI ME (500 MHz, CDCl<sub>3</sub>, 25 °C, TMS).</p>
Full article ">Figure 2
<p>UV–vis spectrum and the calibration curve of P18 N PI ME. (<b>A</b>) Specificity data for P18 N PI ME, placebo (no P18 N PI ME), and P18 N PI ME-loaded SLNs in MeOH at 25 °C. (<b>B</b>) Linearity data for the standard solution of P18 N PI ME in MeOH.</p>
Full article ">Figure 3
<p>Particle characterization of P18 N PI ME-loaded SLNs manufactured using various components. (<b>A</b>) Particle size and polydispersity index (PDI) and (<b>B</b>) zeta potential. Results are presented as means ± standard deviation from three independent experiments (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>(<b>A</b>) Loading efficiency (LE) and (<b>B</b>) loading amount (LA) of P18 N PI ME in the formulations. Results are expressed as means ± standard deviation from three independent experiments (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Cumulative release percentage profiles of P18 N PI ME from SLNs in the release medium, determined using the dialysis bag method. Results are presented as means ± standard error from three independent experiments (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 6
<p>Photostability assessment of P18 N PI ME solution, using the percentage of non-degraded P18 N PI ME in both the solution and SLNs before and after LED irradiation. Results are expressed as means ± standard deviations of three independent experiments (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>Reduction rate (%) of DPBF absorbance at 418 nm for P18 N PI ME with or without SLNs after light exposure (total light dose of 2 J/cm<sup>2</sup>; exposure time of 15 min). Statistical significance of the difference in DPBF degradation between P18 N PI ME and formulation is indicated by a single asterisk (<span class="html-italic">p</span> &lt; 0.05) or double asterisks (<span class="html-italic">p</span> &lt; 0.01). Results are shown as means ± standard deviation for triplicates (<span class="html-italic">n</span> = 3). NC: DPBF (1,3-diphenylisobenzofuran); PC: MB (methylene blue).</p>
Full article ">Figure 8
<p>Cytotoxicity of P18 N PI ME solution and formulations F1, F2, F3, and F4 against (<b>A</b>) HeLa and (<b>B</b>) A549 cell lines. The WST assay was used for the estimation of viability. Results are exhibited as means ± standard deviation for triplicates (<span class="html-italic">n</span> = 3).</p>
Full article ">
15 pages, 1500 KiB  
Review
Singlet Oxygen in Photodynamic Therapy
by Shengdong Cui, Xingran Guo, Sen Wang, Zhe Wei, Deliang Huang, Xianzeng Zhang, Timothy C. Zhu and Zheng Huang
Pharmaceuticals 2024, 17(10), 1274; https://doi.org/10.3390/ph17101274 - 26 Sep 2024
Viewed by 598
Abstract
Photodynamic therapy (PDT) is a therapeutic modality that depends on the interaction of light, photosensitizers, and oxygen. The photon absorption and energy transfer process can lead to the Type II photochemical reaction of the photosensitizer and the production of singlet oxygen (1 [...] Read more.
Photodynamic therapy (PDT) is a therapeutic modality that depends on the interaction of light, photosensitizers, and oxygen. The photon absorption and energy transfer process can lead to the Type II photochemical reaction of the photosensitizer and the production of singlet oxygen (1O2), which strongly oxidizes and reacts with biomolecules, ultimately causing oxidative damage to the target cells. Therefore, 1O2 is regarded as the key photocytotoxic species accountable for the initial photodynamic reactions for Type II photosensitizers. This article will provide a comprehensive review of 1O2 properties, 1O2 production, and 1O2 detection in the PDT process. The available 1O2 data of regulatory-approved photosensitizing drugs will also be discussed. Full article
(This article belongs to the Special Issue Photodynamic Therapy 2023)
Show Figures

Figure 1

Figure 1
<p>The lowest electronic state energy level transitions and molecular orbital schematics of the oxygen molecule. The configuration of the molecular orbitals of the <sup>1</sup>Δ<sub>g</sub> can be described as follows: O<sub>2</sub>KK(2σ<sub>g</sub>)<sup>2</sup>(2σ<sub>u</sub>)<sup>2</sup>(3σ<sub>g</sub>)<sup>2</sup>(1π<sub>u</sub>)<sup>4</sup>(<math display="inline"><semantics> <msubsup> <mrow> <mn>1</mn> <mo>π</mo> </mrow> <mi mathvariant="normal">g</mi> <mo>+</mo> </msubsup> </semantics></math>)(<math display="inline"><semantics> <msubsup> <mrow> <mn>1</mn> <mo>π</mo> </mrow> <mi mathvariant="normal">g</mi> <mo>+</mo> </msubsup> </semantics></math>).</p>
Full article ">Figure 2
<p>Simplified Jablonski energy level diagram of Type I and Type II photochemical reaction.</p>
Full article ">
Back to TopTop