[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (409)

Search Parameters:
Keywords = targeting photosensitizer

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
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 459
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 ">
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 1161
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 ">
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 626
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 ">
25 pages, 3059 KiB  
Article
Discovery of Benzopyrone-Based Candidates as Potential Antimicrobial and Photochemotherapeutic Agents through Inhibition of DNA Gyrase Enzyme B: Design, Synthesis, In Vitro and In Silico Evaluation
by Akram Abd El-Haleem, Usama Ammar, Domiziana Masci, Sohair El-Ansary, Doaa Abdel Rahman, Fatma Abou-Elazm and Nehad El-Dydamony
Pharmaceuticals 2024, 17(9), 1197; https://doi.org/10.3390/ph17091197 - 11 Sep 2024
Viewed by 604
Abstract
Bacterial DNA gyrase is considered one of the validated targets for antibacterial drug discovery. Benzopyrones have been reported as promising derivatives that inhibit bacterial DNA gyrase B through competitive binding into the ATP binding site of the B subunit. In this study, we [...] Read more.
Bacterial DNA gyrase is considered one of the validated targets for antibacterial drug discovery. Benzopyrones have been reported as promising derivatives that inhibit bacterial DNA gyrase B through competitive binding into the ATP binding site of the B subunit. In this study, we designed and synthesized twenty-two benzopyrone-based derivatives with different chemical features to assess their antimicrobial and photosensitizing activities. The antimicrobial activity was evaluated against B. subtilis, S. aureus, E. coli, and C. albicans. Compounds 6a and 6b (rigid tetracyclic-based derivatives), 7a-7f (flexible-linker containing benzopyrones), and 8a-8f (rigid tricyclic-based compounds) exhibited promising results against B. subtilis, S. aureus, and E. coli strains. Additionally, these compounds demonstrated photosensitizing activities against the B. subtilis strain. Both in silico molecular docking and in vitro DNA gyrase supercoiling inhibitory assays were performed to study their potential mechanisms of action. Compounds 8a-8f exhibited the most favorable binding interactions, engaging with key regions within the ATP binding site of the DNA gyrase B domain. Moreover, compound 8d displayed the most potent IC50 value (0.76 μM) compared to reference compounds (novobiocin = 0.41 μM and ciprofloxacin = 2.72 μM). These results establish a foundation for structure-based optimization targeting DNA gyrase inhibition with antibacterial activity. Full article
(This article belongs to the Special Issue In Silico and In Vitro Screening of Small Molecule Inhibitors)
Show Figures

Figure 1

Figure 1
<p>Diagrammatic illustration of the molecular structure of DNA gyrase enzyme and the function of each component. DNA gyrase is a tetrameric enzyme composed of two GyrA and two GyrB subunits (GyrA2GyrB2). Structurally, this complex is organized through three sets of “gates”, whose sequential opening and closing facilitate the direct transfer of DNA segments and the introduction of two negative supercoils. The N-gates are shaped by the ATPase domains of the GyrB subunits. The binding of two ATP molecules triggers dimerization, leading to the closure of these gates. Conversely, hydrolysis causes their opening. The catalytic center responsible for DNA cleavage and reunion is situated in the DNA gates formed by all gyrase subunits. On the other hand, the C-gates are formed by the GyrA subunits [<a href="#B13-pharmaceuticals-17-01197" class="html-bibr">13</a>,<a href="#B14-pharmaceuticals-17-01197" class="html-bibr">14</a>].</p>
Full article ">Figure 2
<p>The reported antimicrobial agents (DNA gyrase inhibitors, (<b>A</b>)) and active photochemotherapeutic furobenzopyrones (<b>B</b>).</p>
Full article ">Figure 3
<p>The rational design of first-in-class benzopyrone-based candidates as dual antimicrobial and photochemotherapeutic agents. A group of structural modifications and developments has been applied to the newly designed benzopyrone-based derivatives. We used both reported benzopyrone derivatives (novobiocin and clorobiocin) and benzothiazole derivative (<b>II</b>) as lead compounds to build our compounds. The fused bicyclic benzopyrone ring was selected to be the central scaffold in our designed derivatives. The phenyl alkyl amide group of benzothiazole derivative (<b>II</b>) was replaced with rigid benzopyrone isostere to maintain the bicyclic structure and incorporate a HBA group in order to provide H-bond interactions with the key amino acid residues within the ATP active site of DNA gyrase B. The fused five-membered thiazole ring of benzothiazole derivative (<b>II</b>) was replaced with its isostere, furan ring, to mimic the photochemotherapeutic agents. Accordingly, a novel series of rigid-based derivatives were designed with the angular furobenzopyrone scaffold in their core structures. An additional phenyl ring was incorporated in fused form (tetracyclic-based derivatives, <b>6a</b> and <b>6b</b>) and substituted form (tricyclic-based compounds, <b>8a</b>-<b>8f</b>) to evaluate accessibility into the active site of the gyrase B domain. Additional series with flexible linkers between the benzopyrone core scaffold and the distal phenyl ring were designed (<b>7a</b>-<b>7f</b>) to evaluate the effect of HBA groups on the binding affinity into the active site.</p>
Full article ">Figure 4
<p>Photosensitizing activity of tested compounds against <span class="html-italic">B. subtilis</span> expressed as minimum inhibitory concentrations (MIC, mg mL<sup>−1</sup>) using xanthotoxin as standard. MIC of tested compounds before and after irradiation with UV lamp (360 nm) for 20 min.</p>
Full article ">Figure 5
<p>3D binding interactions of generated conformational clusters of the tested compounds (<b>6b</b>, <b>7a</b>, and <b>8d</b>) and the reference ligand (<b>II</b>) within the ATP binding site of the <span class="html-italic">N</span>-terminal domain of gyrase B subunit (GyrB24kDa, PDB ID: 6YD9). (<b>A</b>), the reference ligand (<b>II</b>), showed a key H-binding interaction with Arg136 (2.1 Å) and hydrophobic interactions with the valine-rich region (Val43, Val71, Val120, and Val167); (<b>B</b>), the rigid tetracyclic benzopyrone derivative (<b>6b</b>), showed H-bond interaction with Gly77 (2.2 Å). The fused phenyl ring could not access the valine-rich hydrophobic pocket (left region) to exhibit key hydrophobic interactions; (<b>C</b>), the flexible-spacer-containing compound (<b>7a</b>), showed a key H-bond interaction with Gly77 (2.0 Å). The rotation of the distal phenyl ring disrupted the coplanarity of the structure, bringing the distal phenyl ring away from the valine-rich hydrophobic pocket; (<b>D</b>), the rigid tricyclic benzopyrone derivative (<b>8e</b>), showed the highest docking score and the greatest number of binding interactions among the tested compounds and the reference ligand (<b>II</b>). It displayed a strong H-bond interaction with Thr165 (2.1 Å). The distal phenyl ring, along with the 4-Me group, was oriented to fully access the valine-rich hydrophobic pocket.</p>
Full article ">Figure 6
<p>General molecular interactions between the designed compounds (rigid tricyclic derivatives, <b>8d</b>-<b>8f</b>) and the ATP binding site of the gyrase B subunit. R = H (<b>8d</b>), Me (<b>8e</b>), OMe (<b>8f</b>). The diagram illustrates the key regions within the ATP binding site. The central benzopyrone core scaffold exhibited a number of hydrophobic interactions with the hydrophobic surface (black). The HBA (CO) of benzopyrone showed strong HB interactions with Gly77 or Thr165 (&lt;2.3 Å, purple). The phenyl ring is positioned at the solvent-exposed area, showing a number of hydrophobic interactions with the key amino acid residues at this region (green). The distal phenyl ring displayed different hydrophobic interactions with the valine-rich hydrophobic pocket (red). The rigid angular furan ring forced the distal phenyl ring into the valine-rich hydrophobic pocket, modulating the coplanarity of the designed compounds.</p>
Full article ">Scheme 1
<p>Reagents and conditions. <b>a</b>, 3-chlorobutan-2-one, acetone, K<sub>2</sub>CO<sub>3</sub>, reflux, 24 h; <b>b</b>, NaOH, iPrOH, reflux, 4 h; <b>c</b>, 2-chlorocyclohexanone, acetone, K<sub>2</sub>CO<sub>3</sub> reflux, 24 h; <b>d</b>, NaOH, iPrOH, reflux, 4 h; <b>e</b>, DDQ, benzene, reflux, 20 h.</p>
Full article ">Scheme 2
<p>Reagents and conditions. <b>a</b>, appropriate phenacyl bromide, acetone, K<sub>2</sub>CO<sub>3</sub>, reflux, 24 h; <b>b</b>, Na, EtOH, 2 h.</p>
Full article ">
20 pages, 19943 KiB  
Article
Polyamine Derived Photosensitizer: A Novel Approach for Photodynamic Therapy of Cancer
by Hao Deng, Ke Xie, Liling Hu, Xiaowen Liu, Qingyun Li, Donghui Xie, Fengyi Xiang, Wei Liu, Weihong Zheng, Shuzhang Xiao, Jun Zheng and Xiao Tan
Molecules 2024, 29(17), 4277; https://doi.org/10.3390/molecules29174277 - 9 Sep 2024
Viewed by 624
Abstract
Polyamines play a pivotal role in cancer cell proliferation. The excessive polyamine requirement of these malignancies is satisfied through heightened biosynthesis and augmented extracellular uptake via the polyamine transport system (PTS) present on the cell membrane. Meanwhile, photodynamic therapy (PDT) emerges as an [...] Read more.
Polyamines play a pivotal role in cancer cell proliferation. The excessive polyamine requirement of these malignancies is satisfied through heightened biosynthesis and augmented extracellular uptake via the polyamine transport system (PTS) present on the cell membrane. Meanwhile, photodynamic therapy (PDT) emerges as an effective anti-cancer treatment devoid of drug resistance. Recognizing these intricacies, our study devised a novel polyamine-derived photosensitizer (PS) for targeted photodynamic treatment, focusing predominantly on pancreatic cancer cells. We synthesized and evaluated novel spermine-derived fluorescent probes (N2) and PS (N3), exhibiting selectivity towards pancreatic cancer cells via PTS. N3 showed minimal dark toxicity but significant phototoxicity upon irradiation, effectively causing cell death in vitro. A significant reduction in tumor volume was observed post-treatment with no pronounced dark toxicity using the pancreatic cancer CDX mouse model, affirming the therapeutic potential of N3. Overall, our findings introduce a promising new strategy for cancer treatment, highlighting the potential of polyamine-derived PSs in PDT. Full article
(This article belongs to the Special Issue Advances in Fluorescent Probe Technology)
Show Figures

Figure 1

Figure 1
<p>The uptake of N2 into PANC-1 and hTERT-HPNE cells. The excitation and emission wavelength of N1 (<b>A</b>) and N2 (<b>B</b>) were found at 420 nm and 530 nm; (<b>C</b>) Fluorescent property of N2 in different solvent: (a) ethanol; (b) DMSO; (c) DMEM and (d) water; (<b>D</b>) The toxicity of N2 in PANC-1 and hTERT-HPNE cells; (<b>E</b>) Fluorescence observed in PANC-1 cells after treated with 1 and 2 µM of N2 for 3, 6, 12 and 24 h, normal exposure time (200×); Fluorescence observed in hTERT-HPNE cells after treated with 1 and 2 µM of N2 for 3, 6, 12 and 24 h, normal exposure time (<b>F</b>), double exposure time (<b>G</b>) (200×); Fluorescence observed in PANC-1 cells after treated with 1 and 2 µM of N1 for 3, 6, 12 and 24 h, normal exposure time (<b>H</b>), 7 fold exposure time (<b>I</b>) (200×). * <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 2
<p>The uptake of N2 into PANC-1 cells relied on PTS. (<b>A</b>) The rate of uptake of N2 into PANC-1 cells increased by increasing N2 concentration until PTS saturation; (<b>B</b>) The uptake of N2 (2 µM) into PANC-1 cells was inhibited by competition of polyamines (10 µM) (Spm: spermine; Put: putrescine; Spd: spermidine; Mix: a mixture of putrescine, spermidine, and spermine); (<b>C</b>) Uptake of N2 (1 and 2 µM) observed in PANC-1 (normal exposure time) and hTERT-HPNE (5 fold exposure time) cells in presence of a different concentration of DFMO (0, 0.31, 0.93 and 2.78 mM) (200×); (<b>D</b>) Evaluation of uptake of N2 (2 µM) into PANC-1 using spectrofluorometry in presence of different concentration of DFMO (0, 0.31, 0.93 and 2.78 mM). DFMO promoted the uptake of N2 into PANC-1 cells. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>The uptake of N3 into PANC-1 cells through PTS. (<b>A</b>) The excitation and emission wavelength of N3 were found at 540 nm and 560 nm; (<b>B</b>) the production rates of singlet oxygen by N3 were increased in an irradiation time and concentration-dependent manner; (<b>C</b>) evaluation of dark cytotoxicity of N3 in PANC-1 cells; (<b>D</b>) the uptake of N3 (0, 1, 2, 4 and 8 µM) into PANC-1 cells for 3, 6, 12 and 24 h (200×); (<b>E</b>) evaluation of uptake of N3 into PANC-1 cells using flow cytometry in presence of different concentrations of DFMO (0, 0.93, 2.78 and 8.33 mM). DFMO promoted the uptake of N3 into PANC-1 cells. * <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 4
<p>The PDT effect of N3 in PANC-1 cells was in a concentration-dependent manner. (<b>A</b>) The PDT was performed in PANC-1 cells with 0, 10, 20, 40, and 80 nM of N3 or N2, N3 exhibited cytotoxicity in a concentration-dependent manner, N2 had no influence on PANC-1 cells, irradiation energy was 6.5 w/m<sup>2</sup>; (<b>B</b>) observation of cell nucleus size and cell count of PANC-1 cells with different concentration of N3 using DAPI dye (200×), irradiation energy was 6.5 w/m<sup>2</sup>; (<b>C</b>) evaluation of total ROS in PANC-1 cells, irradiation energy was 1.28 w/m<sup>2</sup>; (<b>D</b>) observation of ROS content in PANC-1 cells in presence of N3. Green fluorescence was ROS detected by DCFH-DA probe (200×), irradiation energy was 6.5 w/m<sup>2</sup>. Irradiation wavelength: 550 ± 50 nm, * <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 5
<p>The PDT effect of N3 in CDX mouse model. Mice were divided into the N3 group (tumor-bearing mice, treated with N3 and irradiated every 2 days in the first 7 days), DMSO group (tumor-bearing mice, provided DMSO and irradiated as the N3 group), and N3 control group (normal nude mice without tumor-bearing, treated with N3 and irradiated as the N3 group). (<b>A</b>) The tumors on mice in the N3 group and DMSO group on the 1st day and 22nd day. Red arrows indicate tumors (<span class="html-italic">n</span> = 10); (<b>B</b>) The tumor volume between the N3 group and DMSO group; (<b>C</b>) the tumor weight in the N3 group and DMSO group on the 22nd day; (<b>D</b>) the body weight of mice in DMSO group, N3 group, and the N3 control group at 22nd day; (<b>E</b>) the HE assays of tumors in DMSO group and the N3 group (200×). Irradiation condition: 532 nm, 1.25 w/cm<sup>2</sup>. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Scheme 1
<p>Synthetic routes of N1 (<b>A</b>), N2 (<b>B</b>), and N3 (<b>C</b>).</p>
Full article ">
17 pages, 7708 KiB  
Article
Self-Assembled Nanoparticles of Silicon (IV)–NO Donor Phthalocyanine Conjugate for Tumor Photodynamic Therapy in Red Light
by Kadireya Aikelamu, Jingya Bai, Qian Zhang, Jiamin Huang, Mei Wang and Chunhong Zhong
Pharmaceutics 2024, 16(9), 1166; https://doi.org/10.3390/pharmaceutics16091166 - 4 Sep 2024
Viewed by 636
Abstract
The combination of photodynamic therapy (PDT) and pneumatotherapy is emerging as one of the most effective strategies for increasing cancer treatment efficacy while minimizing side effects. Photodynamic forces affect nitric oxide (NO) levels as activated photosensitizers produce NO, and NO levels in the [...] Read more.
The combination of photodynamic therapy (PDT) and pneumatotherapy is emerging as one of the most effective strategies for increasing cancer treatment efficacy while minimizing side effects. Photodynamic forces affect nitric oxide (NO) levels as activated photosensitizers produce NO, and NO levels in the tumor and microenvironment directly impact tumor cell responsiveness to PDT. In this paper, 3-benzenesulfonyl-4-(1-hydroxy ether)-1,2,5-oxadiazole-2-oxide NO donor–silicon phthalocyanine coupling (SiPc–NO) was designed and prepared into self-assembled nanoparticles (SiPc–NO@NPs) by precipitation method. By further introducing arginyl-glycyl-aspartic acid (RGD) on the surface of nanoparticles, NO-photosensitizer delivery systems (SiPc–NO@RGD NPs) with photo-responsive and tumor-targeting properties were finally prepared and preliminarily evaluated in terms of their formulation properties, NO release, and photosensitizing effects. Furthermore, high reactive oxygen species (ROS) generation efficiency and high PDT efficiency in two breast cancer cell lines (human MCF-7 and mouse 4T1) under irradiation were also demonstrated. The novel SiPc–NO@RGD NPs show great potential for application in NO delivery and two-photon bioimaging-guided photodynamic tumor therapy. Full article
(This article belongs to the Section Drug Targeting and Design)
Show Figures

Figure 1

Figure 1
<p>Schematic of the preparation, characterization and anticancer effect of SiPc-NO@RGD NPs.</p>
Full article ">Figure 2
<p>Synthesis routes of the SiPc–NO. (a) m-CPBA, DCM, r.t., 1 h; (b) CH<sub>3</sub>COOH, nitric acid, reflux, 1.5 h; (c) 1,3-propanediol, NaOH, THF, r.t., 15 min; (d) DTDP, EDC, DMAP, DCM, r.t., 48 h; (e) toluene, reflux, 36 h.</p>
Full article ">Figure 3
<p>(<b>A</b>) The TEM result and appearances of SiPc–NO@RGD NPs. (<b>B</b>) Stability of SiPc–NO@RGD NPs in water stored at room temperature for 30 days. (<b>C</b>) UV spectrum of SiPc–NO self-assembled nanoparticles. (<b>D</b>) Fluorescence spectra of SiPc–NO self-assembled nanoparticles. (<b>E</b>) NO release from SiPc–NO self-assembled nanoparticles. (<b>F</b>) Reactive oxygen determination of SiPc–NO@RGD NPs.</p>
Full article ">Figure 4
<p>Cytotoxicity of SiPc–NO self-assembled nanoparticles. Significance identification: *, <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 5
<p>(<b>A</b>) Cellular uptake of NPs in 4T1 cells detected by CLSM. (<b>B</b>) Cellular uptake of NPs in MCF-7cells detected by CLSM. (<b>C</b>) Cellular uptake of NPs in 4T1 cells detected by flow cytometry. (<b>D</b>) Cellular uptake of NPs in MCF-7 cells detected by flow cytometry.</p>
Full article ">Figure 6
<p>(<b>A</b>) SiPc–NO self-assembled nanoparticles affect NO content in 4T1 cells. (<b>B</b>) SiPc–NO self-assembled nanoparticles affect NO content in MCF-7 cells.</p>
Full article ">Figure 7
<p>(<b>A</b>) SiPc–NO self-assembled nanoparticles affect reactive oxygen species content in 4T1 cells. (<b>B</b>) SiPc–NO self-assembled nanoparticles affect reactive oxygen species content in MCF-7 cells.</p>
Full article ">Figure 8
<p>(<b>A</b>) SiPc–NO self-assembled nanoparticles inhibit the healing ability of 4T1 cells. (<b>B</b>) SiPc–NO self-assembled nanoparticles inhibit the healing ability of MCF-7 cells. Significance identification: **, <span class="html-italic">p</span> &lt; 0.01; ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
14 pages, 2411 KiB  
Article
Reactive Oxygen Species-Regulated Conjugates Based on Poly(jasmine) Lactone for Simultaneous Delivery of Doxorubicin and Docetaxel
by Jyoti Verma, Vishal Kumar, Carl-Eric Wilen, Jessica M. Rosenholm and Kuldeep K. Bansal
Pharmaceutics 2024, 16(9), 1164; https://doi.org/10.3390/pharmaceutics16091164 - 3 Sep 2024
Viewed by 714
Abstract
In cancer therapy, it is essential to selectively release cytotoxic agents into the tumor to prevent the adverse effects associated with anticancer drugs. Thus, in this study, a stimuli-sensitive polymer–drug conjugate was synthesized for selective drug release. Doxorubicin (DOX) and docetaxel (DTX) were [...] Read more.
In cancer therapy, it is essential to selectively release cytotoxic agents into the tumor to prevent the adverse effects associated with anticancer drugs. Thus, in this study, a stimuli-sensitive polymer–drug conjugate was synthesized for selective drug release. Doxorubicin (DOX) and docetaxel (DTX) were conjugated onto novel poly(jasmine lactone) based copolymer via a thioketal (TK) linker. In addition, a photosensitizer (chlorin e6) was attached to the polymer, which served as a reactive oxygen species generator to cleave the TK linker. The conjugate is readily self-assembled into micelles less than 100 nm in size. Micelles demonstrate a notable increase in their ability to cause cell death when exposed to near-infrared (NIR) light on MDA-MB-231 breast cancer cells. The increase in cytotoxicity is higher than that observed with the combination of free DOX and DTX. The accumulation of DOX in the nucleus after release from the micelles (laser irradiation) was also confirmed by confocal microscopy. In the absence of light, micelles did not show any toxicity while the free drugs were found toxic irrespective of the light exposure. The obtained results suggest the targeted drug delivery potential of micelles regulated by the external stimuli, i.e., NIR light. Full article
(This article belongs to the Special Issue Functional Nanomaterials for Drug Delivery in Photodynamic Therapy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Particle size distribution by volume and intensity percentage determined by DLS and TEM image to confirm size and shape of micelles (300 µg/mL concentration). Inset: picture of the formulation.</p>
Full article ">Figure 2
<p>(<b>A</b>) Absorbance changes of DPBF at wavelength 410 nm with DPBF added before the laser irradiation (<b>B</b>) Absorbance changes of DPBF at wavelength 410 nm with DBPF added after the laser irradiation (n = 3).</p>
Full article ">Figure 3
<p>Percentage cell proliferation inhibition study of DOX, DTX, mixture of DOX-DTX, Ce6, and PJL-DOX-DTX micelles (equivalent to 240, 360, 480 ng mL<sup>−1</sup> for Ce6) in the presence and absence of laser light with or without media replacement on MDA-MB-231 cell lines. ns, <span class="html-italic">p</span> ≥ 0.05, **, <span class="html-italic">p</span> &lt; 0.01, ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>(<b>A</b>) DCFH-DA fluorescence intensity post treatment with Ce6 micelles and free Ce6 in MDA MB-231 cells after 3 h incubation and laser irradiation. (<b>B</b>) Percentage cell proliferation inhibition study of DOX, DTX, mixture of DOX-DTX, Ce6, and PJL-DOX-DTX micelles (2.5, 5, 10 μg mL<sup>−1</sup> with respect to DOX) on Vero H10 cell lines after 48 h incubation. ns, <span class="html-italic">p</span> ≥ 0.05, ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>CLSM images of PJL-DOX-DTX micelles internalized in MDA-MB-231 cells after 12 h incubation with or without laser treatment. Cells were treated with 400 ng mL<sup>−1</sup> equivalent concentration of DOX. Cell nucleus was stained with DAPI (blue), while DOX inherent red fluorescence was used to detect PJL-DOX-DTX micelles. Scale bar—20 µm (laser) and 10 µm (without laser).</p>
Full article ">Scheme 1
<p>Conjugation of doxorubicin, docetaxel, and chlorin e6 to the mPEG-b-PJL-OH polymer to obtain mPEG-b-PJL-Ce6-TK-DOX-DTX (PJL-DOX-DTX). DOX-TK and DTX-TK were synthesized separately before addition to the reaction. (N-(3-Dimethylaminopropyl)-N′-ethylcarbodiimide (EDC), 4-(Dimethylamino) pyridine (DMAP).</p>
Full article ">
44 pages, 3570 KiB  
Review
Applicability of Quantum Dots in Breast Cancer Diagnostic and Therapeutic Modalities—A State-of-the-Art Review
by Dominika Kunachowicz, Karolina Kłosowska, Natalia Sobczak and Marta Kepinska
Nanomaterials 2024, 14(17), 1424; https://doi.org/10.3390/nano14171424 - 31 Aug 2024
Viewed by 941
Abstract
The increasing incidence of breast cancers (BCs) in the world population and their complexity and high metastatic ability are serious concerns for healthcare systems. Despite the significant progress in medicine made in recent decades, the efficient treatment of invasive cancers still remains challenging. [...] Read more.
The increasing incidence of breast cancers (BCs) in the world population and their complexity and high metastatic ability are serious concerns for healthcare systems. Despite the significant progress in medicine made in recent decades, the efficient treatment of invasive cancers still remains challenging. Chemotherapy, a fundamental systemic treatment method, is burdened with severe adverse effects, with efficacy limited by resistance development and risk of disease recurrence. Also, current diagnostic methods have certain drawbacks, attracting attention to the idea of developing novel, more sensitive detection and therapeutic modalities. It seems the solution for these issues can be provided by nanotechnology. Particularly, quantum dots (QDs) have been extensively evaluated as potential targeted drug delivery vehicles and, simultaneously, sensing and bioimaging probes. These fluorescent nanoparticles offer unlimited possibilities of surface modifications, allowing for the attachment of biomolecules, such as antibodies or proteins, and drug molecules, among others. In this work, we discuss the potential applicability of QDs in breast cancer diagnostics and treatment in light of the current knowledge. We begin with introducing the molecular and histopathological features of BCs, standard therapeutic regimens, and current diagnostic methods. Further, the features of QDs, along with their uptake, biodistribution patterns, and cytotoxicity, are described. Based on the reports published in recent years, we present the progress in research on possible QD use in improving BC diagnostics and treatment efficacy as chemotherapeutic delivery vehicles and photosensitizing agents, along with the stages of their development. We also address limitations and open questions regarding this topic. Full article
(This article belongs to the Special Issue Advances in the Investigation of Semiconductor Quantum Dots)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) General histological and (<b>b</b>) molecular classification of breast cancers. Created with Servier Medical Art (<a href="https://smart.servier.com/" target="_blank">https://smart.servier.com/</a>).</p>
Full article ">Figure 2
<p>(<b>a</b>) Cancer stem cells (CSCs) undergo either symmetric division, generating two new stem cells, which is known as self-renewal, or asymmetric division, where one CSC and one non-stem cancer cell is formed. These non-stem cancer cells further undergo differentiation, giving rise to various populations. (<b>b</b>) A heterogeneous tumor microenvironment (TME) consists of a variety of cells, including tumor-associated macrophages (TAMs), tumor-infiltrating lymphocytes (TILs), cancer-associated fibroblasts (CAFs), tumor-associated adipocytes (TAAs), and different populations of cancerous cells. (<b>c</b>) Clonal evolution of cancer cells. All of these—heterogeneity, presence, features of cancer stem cells, and clonal evolution—contribute to cancer progression in a number of ways. Created with Servier Medical Art, on a license CC BY 4.0 (<a href="https://smart.servier.com/" target="_blank">https://smart.servier.com/</a>).</p>
Full article ">Figure 3
<p>Summary of current treatment and diagnostic modalities in breast cancer with their drawbacks. Created with Servier Medical Art (<a href="https://smart.servier.com/" target="_blank">https://smart.servier.com/</a>).</p>
Full article ">Figure 4
<p>Features of quantum dots. Created with Servier Medical Art (<a href="https://smart.servier.com/" target="_blank">https://smart.servier.com/</a>).</p>
Full article ">Figure 5
<p>TEM images of (<b>a</b>) CdSe, (<b>b</b>) CdTe, (<b>c</b>) HgS, and (<b>d</b>) CQDs. Reproduced with permission from (<b>a</b>) Ref. [<a href="#B93-nanomaterials-14-01424" class="html-bibr">93</a>], American Chemical Society; (<b>b</b>) Ref. [<a href="#B94-nanomaterials-14-01424" class="html-bibr">94</a>], American Chemical Society; (<b>c</b>) Ref. [<a href="#B95-nanomaterials-14-01424" class="html-bibr">95</a>], American Chemical Society; (<b>d</b>) Ref. [<a href="#B96-nanomaterials-14-01424" class="html-bibr">96</a>] reproduced under the Creative Commons CC BY license.</p>
Full article ">Figure 6
<p>In a selective, specific detection of breast cancer biomarkers, such as those listed in this chapter, proteins and non-coding RNA, QD–antibodies, QD–aptamers, or QD–ssDNA systems are used. Specific interaction leads to signal generation, which is further detected and analyzed. To perform a successful detection or quantification, one component—a detected or detection molecule—is immobilized. Created with Servier Medical Art (<a href="https://smart.servier.com/" target="_blank">https://smart.servier.com/</a>).</p>
Full article ">Figure 7
<p>Schematic representation of chemotherapeutic release from QD–drug conjugates. Upon internalization, which may occur through macropinocytosis, caveolin-dependent, clathrin-dependent, or clathrin-/caveolin-independent endocytosis, an internal or external stimulus (e.g., pH, enzyme, redox conditions, ultrasound, magnetic field) triggers drug release from the complex. The free drug, according to its mechanism of action, causes different events, leading to cancer cell death. Created with Servier Medical Art (<a href="https://smart.servier.com/" target="_blank">https://smart.servier.com/</a>).</p>
Full article ">
16 pages, 4083 KiB  
Article
Multi-Sensitive Au NCs/5-FU@Carr-LA Composite Hydrogels for Targeted Multimodal Anti-Tumor Therapy
by Chunxia Qi, Ang Li, Baoming Wu and Peisan Wang
Molecules 2024, 29(17), 4051; https://doi.org/10.3390/molecules29174051 - 27 Aug 2024
Viewed by 492
Abstract
Multifunctional targeted drug delivery systems have been explored as a novel cancer treatment strategy to overcome limitations of traditional chemotherapy. The combination of photodynamic therapy and chemotherapy has been shown to enhance efficacy, but the phototoxicity of traditional photosensitizers is a challenge. In [...] Read more.
Multifunctional targeted drug delivery systems have been explored as a novel cancer treatment strategy to overcome limitations of traditional chemotherapy. The combination of photodynamic therapy and chemotherapy has been shown to enhance efficacy, but the phototoxicity of traditional photosensitizers is a challenge. In this study, we prepared a multi-sensitive composite hydrogel containing gold nanoclusters (Au NCs) and the temperature-sensitive antitumor drug 5-fluorourac il (5-FU) using carboxymethyl cellulose (Carr) as a dual-functional template. Au NCs were synthesized using sodium borohydride as a reducing agent and potassium as a promoter. The resulting Au NCs were embedded in the Carr hydrogel, which was then conjugated with lactobionic acid (LA) as a targeting ligand. The resulting Au NCs/5-FU@Carr-LA composite hydrogel was used for synergistic photodynamic therapy (PDT), photothermal therapy (PTT), and chemotherapy. Au NCs/5-FU@Carr-LA releases the drug faster at pH 5.0 due to the acid sensitivity of the Carr polymer chain. In addition, at 50 °C, the release rate of Au NCs/5-FU@Carr-LA is 78.2%, indicating that the higher temperature generated by the photothermal effect is conducive to the degradation of Carr polymer chains. The Carr hydrogel stabilized the Au NCs and acted as a matrix for drug loading, and the LA ligand facilitated targeted delivery to tumor cells. The composite hydrogel exhibited excellent biocompatibility and synergistic antitumor efficacy, as demonstrated by in vitro and in vivo experiments. In addition, the hydrogel had thermal imaging capabilities, making it a promising multifunctional platform for targeted cancer therapy. Full article
(This article belongs to the Special Issue Hydrogels: Preparation, Characterization, and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The SEM image of Au NCs/5-FU@Carr-LA; (<b>B</b>) the TEM image of Au NCs. (The insets in (<b>B</b>) present the lattice spacing of Au NCs.) (<b>C</b>) UV-Vis-NIR spectra and (<b>D</b>) FTIR spectra of (a) Carr, (b) Au NCs@Carr, and (c) Au NCs@Carr-LA, respectively. (<b>E</b>) The zeta potential diagram of different samples in the medium. (<b>F</b>) Zeta potentials of different samples.</p>
Full article ">Figure 2
<p>(<b>A</b>) The digital image of Au NCs@Carr hydrogel; sol-gel conversion of hydrogel. (<b>B</b>) Temperature curves of different samples irradiated with 808 nm laser at 1 W/cm<sup>2</sup> for 10 min; the insets show the corresponding thermal images of PBS (a), Carr (b), Au NCs@Carr (c), and Au NCs/5-FU@Carr-LA (d). (<b>C</b>) The temperature change curve of Au NCs/5-FU@Carr-LA with irradiation for five cycles.</p>
Full article ">Figure 3
<p>(<b>A</b>) UV-Vis spectra of Au NCs@Carr-LA loaded with 5-FU under different conditions. (The insets in (<b>A</b>) present the standard curve of the 5-FU.) (<b>B</b>) In vitro drug release profiles for the release of 5-FU from Au NCs/5-FU@Carr-LA in the different temperatures (37 °C, 50 °C) and pHs (5.0, 7.4) of the buffer solution.</p>
Full article ">Figure 4
<p>Changes of O<sub>2</sub> generation with time from the decomposition of H<sub>2</sub>O<sub>2</sub> catalyzed by different samples at pH (<b>A</b>) 6.5 and (<b>B</b>) 5.5.</p>
Full article ">Figure 5
<p>Photodynamic effects of (<b>A</b>) DPBF, (<b>B</b>) Carr, (<b>C</b>) Au NCs@Carr, and (<b>D</b>) Au NCs/5-FU@Carr-LA; (<b>E</b>) Au NCs/5-FU@Carr-LA+ H<sub>2</sub>O<sub>2</sub>; (<b>F</b>) the absorbance change curves of different sample dispersions at 410 nm.</p>
Full article ">Figure 6
<p>(<b>A</b>) Viability of HepG2 cells incubated with different concentrations of samples under different conditions, respectively; (<b>B</b>) cytotoxicity assays of (a) HepG2 and (b) BEAS-2B cells incubated with various concentrations of Au NCs/5-FU@Carr-LA with laser irradiation.</p>
Full article ">Figure 7
<p>Fluorescence microscopy images of HepG2 cells incubated with (<b>A1</b>,<b>A2</b>) control groups (without samples), (<b>B1</b>,<b>B2</b>) Carr, (<b>C1</b>,<b>C2</b>) Au NCs@Carr, (<b>D1</b>,<b>D2</b>) Au NCs/5-FU@Carr-LA, (<b>A1</b>,<b>B1</b>,<b>C1</b>,<b>D1</b>) without laser irradiation, and (<b>A2</b>,<b>B2</b>,<b>C2</b>,<b>D2</b>) with laser irradiation (808 nm, 1 W/cm<sup>2</sup>), respectively.</p>
Full article ">Figure 8
<p>In vivo antitumor effect of samples under different conditions. (<b>A</b>) Time-dependent body weight change curves, (<b>B</b>) the tumor weight change histograms, (<b>C</b>) tumor-volume curves, and (<b>D</b>) the survival rates of the ICR mice treated with different samples. (<b>E</b>) Biodistribution of Au in main organs and tumor after intravenous injection of Au NCs/5-FU@Carr-LA for 12 h. (<b>F</b>) The physical image of the tumor (L represents laser).</p>
Full article ">Scheme 1
<p>Potassium borohydride was used as a reducing agent of chloro-auric acid to prepare gold nanoclusters (Au NCs), and, at the same time, it was used as a potassium source to induce carrageenan to form a double-sensitive hydrogel, and nanoparticles containing Au NCs and the temperature-sensitive anticancer drug 5-flouoruracil (5-FU) were prepared. Au NCs/5-FU@Carr-LA nanogel particles can be used for the study of the synergistic anti-tumor effect of phototherapy and chemotherapy, which can specifically recognize the glycoprotein lactose acid molecule (LA) with galactose residue as a targeting agent.</p>
Full article ">
35 pages, 5422 KiB  
Review
A Review of the Efficacy of Nanomaterial-Based Natural Photosensitizers to Overcome Multidrug Resistance in Cancer
by Jagadeesh Rajaram, Lokesh Kumar Mende and Yaswanth Kuthati
Pharmaceutics 2024, 16(9), 1120; https://doi.org/10.3390/pharmaceutics16091120 - 24 Aug 2024
Viewed by 902
Abstract
Natural photosensitizers (PS) are compounds derived from nature, with photodynamic properties. Natural PSs have a similar action to that of commercial PSs, where cancer cell death occurs by necrosis, apoptosis, and autophagy through ROS generation. Natural PSs have garnered great interest over the [...] Read more.
Natural photosensitizers (PS) are compounds derived from nature, with photodynamic properties. Natural PSs have a similar action to that of commercial PSs, where cancer cell death occurs by necrosis, apoptosis, and autophagy through ROS generation. Natural PSs have garnered great interest over the last few decades because of their high biocompatibility and good photoactivity. Specific wavelengths could cause phytochemicals to produce harmful ROS for photodynamic therapy (PDT). However, natural PSs have some shortcomings, such as reduced solubility and lower uptake, making them less appropriate for PDT. Nanotechnology offers an opportunity to develop suitable carriers for various natural PSs for PDT applications. Various nanoparticles have been developed to improve the outcome with enhanced solubility, optical adsorption, and tumor targeting. Multidrug resistance (MDR) is a phenomenon in which tumor cells develop resistance to a wide range of structurally and functionally unrelated drugs. Over the last decade, several researchers have extensively studied the effect of natural PS-based photodynamic treatment (PDT) on MDR cells. Though the outcomes of clinical trials for natural PSs were inconclusive, significant advancement is still required before PSs can be used as a PDT agent for treating MDR tumors. This review addresses the increasing literature on MDR tumor progression and the efficacy of PDT, emphasizing the importance of developing new nano-based natural PSs in the fight against MDR that have the required features for an MDR tumor photosensitizing regimen. Full article
(This article belongs to the Special Issue Development of Novel Tumor-Targeting Nanoparticles, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of mechanisms contributing to the development of MDR in cancer cells. Cancer cells have developed a wide range of mechanisms to fight against various therapeutical drugs. Abc transporters, alterations in the apoptosis pathway, drug inactivation through cellular metabolism, and mutations in cellular and drug targets, an enhanced DNA repair mechanism, are well-known mechanisms. Altered mechanism in the cancer cell nullifies the effect of drug by changes in various properties which denoted as drug*.</p>
Full article ">Figure 2
<p>Schematic illustration of the molecular mechanism of PDT. The Jablonski diagram illustrates the transitions of PSs from the ground state to the excited state, both for type I and type II reactions. This transition occurs when the PSs absorb light of a particular wavelength, resulting in the generation of free radicals that can cause cellular damage.</p>
Full article ">Figure 3
<p>(<b>A</b>) Schematic illustrations of the chemical structure of curcumin and its derivative C086 and the preparation of C086@Hsa nanoparticles and the process of photodynamic therapy. C086, a derivative of curcumin, was loaded into the albumin-based nanoparticles by a simple self-assembly method. (<b>B</b>) The formation of self assembled curcumin derivative loaded Human serum albumin for PDT against cancer cells “Reprinted with permission from Ref. [<a href="#B147-pharmaceutics-16-01120" class="html-bibr">147</a>]. 2021, Elsevier B.V.”.</p>
Full article ">Figure 4
<p>A diagrammatic representation of the synthetic process of DOX@PLA@Au−PEG−MnP NPs. Adapted with permission from Ref. [<a href="#B167-pharmaceutics-16-01120" class="html-bibr">167</a>].</p>
Full article ">Figure 5
<p>A schematic illustration of the synthesis process of PGCA@PA nanoparticles and the mechanism by which they induce cell death using light exposure. The acid-responsive polygalactose-co-polycinnamaldehyde polyprodrug was combined with cinnamaldehyde (CA) and PS pheophorbide A (PA). The galactose receptor present in cancer cells is responsible for the internal breakdown of PGCA NP and the generation of ROS levels upon exposure to light. This ultimately leads to the apoptosis of cancer cells. “Reprinted with permission from Ref. [<a href="#B120-pharmaceutics-16-01120" class="html-bibr">120</a>]. 2020. Elsevier B.V.”.</p>
Full article ">Figure 6
<p>Schematic illustrations of HB and silver nanoparticles loaded with PLGA nanoparticles that were prepared for this study. The focus was on investigating the phototoxicity and singlet oxygen production of HB under light irradiation. To achieve this, polylactide-co-glycolide nanoparticle formulations were developed by incorporating HB and nanosilver enhances the toxicity against cancer cells. Reprinted with permission from Ref. [<a href="#B178-pharmaceutics-16-01120" class="html-bibr">178</a>]. 2017. Elsevier B.V.</p>
Full article ">Figure 7
<p>Schematic illustration of effect of Pec@Ag on the generation of the triplet excited state of Rf (<sup>3</sup>Rf*) by the presence of pectin-coated silver nanoparticles (Pec@AgNP) due to the formation of a complex between Rf and Pec@AgNP (Rf-Pec@AgNP). Pec@AgNP* represents the formation of superoxide radicals when irradiated with light. These radical are toxic against cancer cells. Reprinted with permission from Ref. [<a href="#B92-pharmaceutics-16-01120" class="html-bibr">92</a>] 2018. American Chemical Society.</p>
Full article ">Figure 8
<p>Schematic illustration of hypericin-loaded transferrin nanoparticles and their mechanism of action by inducing PP2A-mediated BMI1 ubiquitination/degradation under light irradiation. “Adapted with permission from Ref. [<a href="#B192-pharmaceutics-16-01120" class="html-bibr">192</a>] 2020. American Chemical Society”.</p>
Full article ">Figure 9
<p>Schematic illustration of the preparation of CS-PGA-5-ALA nanoparticles and their phototoxicity under light illustration. The nanocomplexes were prepared by using chitosan as a polycation and alginic and polygalacturonic acid as polyanions through coacervation. “Reprinted with permission from Ref. [<a href="#B126-pharmaceutics-16-01120" class="html-bibr">126</a>]. 2017. Elsevier B.V.”.</p>
Full article ">Figure 10
<p>Schematic illustration of CD assembly derived from pheophytins for FL imaging and synergistic PDT of cancer. Pheophytin is a type of Mg-free chlorophyll derivative used for the preparation of CDs by using a microwave method and self-assembled with DSPE-mPEG2000 for efficient <sup>1</sup>O<sub>2</sub> generation and efficient fluorescence (FL) imaging. Reprinted with permission from Ref. [<a href="#B99-pharmaceutics-16-01120" class="html-bibr">99</a>]. 2019. Wiley.</p>
Full article ">Figure 11
<p>Schematic illustration of the synthesis of FBC nanogel and the mechanism of the PDT performance of the FBC nanogel. A bacteriochlorin analogue, tetra fluorophenyl bacteriochlorin, was developed by the one-step reduction of tetra fluorophenyl porphyrin (TFPP). A biocompatible FBC nanogel could be directly formed by blending FBC with SH–PEG–SH. “Adapted with permission from Ref. [<a href="#B127-pharmaceutics-16-01120" class="html-bibr">127</a>]”.</p>
Full article ">Figure 12
<p>Schematic illustration of the administration of 5-aminolevulinic acid and <sup>64</sup>Cu-DOTA-trastuzumab based nanocarrier into HER2-overexpressing cancer cells. The nanocarrier generats <sup>1</sup>O<sub>2</sub> via Cerenkov luminescence energy transfer (CLET) results in cancer cell death. Reprinted with permission from Ref. [<a href="#B239-pharmaceutics-16-01120" class="html-bibr">239</a>]. 2019. Elsevier B.V.</p>
Full article ">
23 pages, 4636 KiB  
Article
Synergistic Photothermal Therapy and Chemotherapy Enabled by Tumor Microenvironment-Responsive Targeted SWCNT Delivery
by Shuoye Yang, Jiaxin Liu, Huajian Yuan, Qianqian Cheng, Weiwei Shen, Yanteng Lv, Yongmei Xiao, Lu Zhang and Peng Li
Int. J. Mol. Sci. 2024, 25(17), 9177; https://doi.org/10.3390/ijms25179177 - 23 Aug 2024
Viewed by 629
Abstract
As a novel therapeutic approach, photothermal therapy (PTT) combined with chemotherapy can synergistically produce antitumor effects. Herein, dithiodipropionic acid (DTDP) was used as a donor of disulfide bonds sensitive to the tumor microenvironment for establishing chemical bonding between the photosensitizer indocyanine green amino [...] Read more.
As a novel therapeutic approach, photothermal therapy (PTT) combined with chemotherapy can synergistically produce antitumor effects. Herein, dithiodipropionic acid (DTDP) was used as a donor of disulfide bonds sensitive to the tumor microenvironment for establishing chemical bonding between the photosensitizer indocyanine green amino (ICG-NH2) and acidified single-walled carbon nanotubes (CNTs). The CNT surface was then coated with conjugates (HD) formed by the targeted modifier hyaluronic acid (HA) and 1,2-tetragacylphosphatidyl ethanolamine (DMPE). After doxorubicin hydrochloride (DOX), used as the model drug, was loaded by CNT carriers, functional nano-delivery systems (HD/CNTs-SS-ICG@DOX) were developed. Nanosystems can effectively induce tumor cell (MCF-7) death in vitro by accelerating cell apoptosis, affecting cell cycle distribution and reactive oxygen species (ROS) production. The in vivo antitumor activity results in tumor-bearing model mice, further verifying that HD/CNTs-SS-ICG@DOX inhibited tumor growth most significantly by mediating a synergistic effect between chemotherapy and PTT, while various functional nanosystems have shown good biological tissue safety. In conclusion, the composite CNT delivery systems developed in this study possess the features of high biocompatibility, targeted delivery, and responsive drug release, and can achieve the efficient coordination of chemotherapy and PTT, with broad application prospects in cancer treatment. Full article
(This article belongs to the Special Issue Natural Products and Synthetic Compounds for Drug Development 2.0)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Characterization results of various SWCNT carriers or DDS. (<b>A</b>) TEM images. (<b>B</b>) <sup>1</sup>H Nuclear magnetic resonance spectra (a: DTDP, b: DTDPA, c: NH<sub>2</sub>-PEG-OH, d: NH<sub>2</sub>-PEG-SS-COOH). (<b>C</b>) Particle size determination results. (<b>D</b>) Zeta potential determination results. (<b>E</b>) FT-IR spectra (a: DTDP, b: DTDPA, c: NH<sub>2</sub>-PEG-OH, d: NH<sub>2</sub>-PEG-SS-COOH, e: CNTs, f: CNTs-SS). (<b>F</b>) UV-vis spectra (a: DOX, b: HD/CNTs-SS-ICG@DOX, c: HD/CNTs-SS@DOX, d: ICG, e: HD/CNTs-SS). (<b>G</b>) Raman spectra. (<b>H</b>) XRD spectra (In (<b>G</b>,<b>H</b>) a: SWCNTs, b: CNTs, c: CNTs-SS, d: HD/CNTs-SS).</p>
Full article ">Figure 2
<p>In vitro release of ICG and DOX from SWCNT DDS. (<b>A</b>) Cumulative release of ICG from HD/CNTs-SS-ICG at pH 5.0 (a: 10 mM GSH, b: 5 mM GSH, c: 2 mM GSH, d: no GSH). (<b>B</b>) Cumulative release of DOX from HD/CNTs-SS-ICG@DOX (a: 45 °C, pH 5.0; b: 25 °C, pH 5.0; c: 45 °C, pH 7.4; d: 25 °C, pH 7.4). (<b>C</b>) Cumulative release of DOX from various DDS at pH 5.0. (<b>D</b>) Cumulative release of DOX from various DDS at pH 7.4. (In (<b>C</b>,<b>D</b>) a: HD/CNTs-SS-ICG@DOX, b: CNTs-SS@DOX, c: CNTs@DOX, d: SWCNTs@DOX).</p>
Full article ">Figure 3
<p>Photothermal performance verification of ICG and various nanosystems. (<b>A</b>) Thermal images of different concentrations of ICG under 5 W/cm<sup>2</sup> of NIR irradiation for 5 min. (<b>B</b>) Thermal images of HD/CNTs-SS-ICG with different power densities of NIR irradiation for 5 min. (<b>C</b>) Temperature curves under NIR irradiation, ICG = 2.2108 μg/mL (a: HD/CNTs-SS-ICG + 6 W/cm<sup>2</sup>, b: HD/CNTs-SS-ICG + 5 W/cm<sup>2</sup>, c: HD/CNTs-SS-ICG + 4 W/cm<sup>2</sup>, d: PBS). (<b>D</b>) Temperature curves of different samples under NIR irradiation of 5 W/cm<sup>2</sup> for 5 min. a: ICG = 2.2108 μg/mL, b: ICG = 1.9535 μg/mL and c: ICG = 0 μg/mL, (a: HD/CNTs-SS-ICG, b: HD/CNTs-SS-ICG@DOX, c: HD/CNTs-SS, d: PBS). (<b>E</b>) Photothermal conversion property of HD/CNTs-SS-ICG in the heating–cooling cycle.</p>
Full article ">Figure 4
<p>Cytotoxicity and biosafety assessment results of various SWCNT carriers or DDSs. (<b>A</b>) Cytotoxicity of various SWCNT carriers. (<b>B</b>) Cell viability of various DDSs towards IPEC-1 cells. (<b>C</b>,<b>E</b>) Hemolytic property evaluation of various SWCNTs (a: 1% Triton, b: SWCNTs, c: CNTs, d: CNTs-SS, e: ICG, f: CNTs-SS-ICG). (<b>D</b>) Cell viability of various DDSs towards MCF-7 cells. d: ICG = 2.000 μg/mL, e/f: ICG = 2.2108 μg/mL, and g: ICG = 1.9535 μg/mL (In (<b>B</b>,<b>D</b>) a: NS, b: DOX, c: HD/CNTs-SS@DOX, d: ICG, e: HD/CNTs-SS-ICG, f: HD/CNTs-SS-ICG+NIR, and g: HD/CNTs-SS-ICG@DOX+NIR, with NS (a), chemotherapy (c), and PTT (f) as controls). * <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 5
<p>Fluorescence images of MCF-7 cells treated with various SWCNT DDSs. (<b>A</b>) 4 h; (<b>B</b>) 24 h. (<b>C</b>) Fluorescence intensity analysis results of MCF-7 cells treated with various SWCNT DDSs. (<b>D</b>) Time-dependent quantitative uptake of different DDSs analyzed by FCM. ROS fluorescence images of MCF-7 cells induced by different treatments. (<b>E</b>) 4 h; (<b>F</b>) 24 h. (<b>G</b>) ROS fluorescence intensity analysis results of MCF-7 cells. Control-1: DOX, Control-2: ICG. * <span class="html-italic">p</span> &lt; 0.01; ** <span class="html-italic">p</span> &lt; 0.005; *** <span class="html-italic">p</span> &lt; 0.0003.</p>
Full article ">Figure 6
<p>Apoptosis scatter diagrams of MCF-7 cells after different treatments. (<b>A</b>) 16 h; (<b>B</b>) 48 h (a: NS, b: DOX, c: HD/CNTs-SS@DOX, d: ICG+NIR, e: HD/CNTs-SS-ICG+NIR, f: HD/CNTs-SS-ICG@DOX+NIR). Q1: Necrotic cells, Q2: Late stage apoptotic cells, Q3: Early stage apoptotic cells, Q4: Cells that have not undergone apoptosis. Living cells typically display low blue/low red colors. Apoptotic cells are displayed as high blue/low red colors. Necrotic cells appear as low blue/high red colors. (<b>C</b>) Apoptosis rate analysis results of MCF-7 cells after different treatments. (<b>D</b>) Apoptosis and necrosis rates analysis results of MCF-7 cells after different treatments. ** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Cycle distribution diagrams of MCF-7 cells after different treatments. (<b>A</b>) 16 h; (<b>B</b>) 48 h (a: NS, b: DOX, c: HD/CNTs-SS@DOX, d: ICG+NIR, e: HD/CNTs-SS-ICG+NIR, f: HD/CNTs-SS-ICG@DOX+NIR). G1: Early stage of DNA synthesis, G2: Late stage of DNA synthesis, S: DNA synthesis period. The green peak represents G1 phase, the blue peak represents G2 phase, and the yellow peak represents S phase. Cycle distribution variation in MCF-7 cells after different treatments. (<b>C</b>) 16 h; (<b>D</b>) 48 h.</p>
Full article ">Figure 8
<p>(<b>A</b>) Thermal images in mice treated with different samples irradiated by NIR. NIR irradia-tion time was set at 0–5 min. (<b>B</b>) Temperature curves of different samples under NIR irradiation. (<b>C</b>) HE staining images of tumors and main organ slices collected from mice in different groups.</p>
Full article ">Figure 9
<p>(<b>A</b>) Treatment procedure. (<b>B</b>) Tumor growth curves of mice after different treatments. (<b>C</b>) Body weight variation in mice after different treatments. (<b>D</b>) Tumor weight of mice after different treatments. (<b>E</b>) Tumor inhibition rate of mice in different treatment groups. (<b>F</b>) Tumor images of mice after 15 days of treatment. (<b>G</b>) Tumor variation in mice in different treatment groups assessed every 3 days up to 12 days. (In (<b>D</b>,<b>E</b>) a: NS, b: DOX, c: HD/CNTs-SS@DOX, d: HD/CNTs-SS-ICG+NIR, e: HD/CNTs-SS-ICG@DOX+NIR).</p>
Full article ">Scheme 1
<p>Schematic diagram of the preparation procedure and antitumor mechanism of HD/CNTs-SS-ICG@DOX DDS. A drug delivery system (HD/CNTs-SS-ICG@DOX) was developed by conjugating ICG-NH<sub>2</sub> on the surface of CNTs through S-S, loading chemotherapy drug DOX inside the lumen, and finally coating the surface with targeting agent HD. The drug delivery system can release drugs in the TME to realize the synergistic antitumor effect by mediating chemotherapy and PTT.</p>
Full article ">
26 pages, 8300 KiB  
Article
Adipocyte-Targeted Nanocomplex with Synergistic Photothermal and Pharmacological Effects for Combating Obesity and Related Metabolic Syndromes
by Yuanyuan Zhang, Xiaojiao Zeng, Fan Wu, Xiaopeng Yang, Tingting Che, Yin Zheng, Jie Li, Yufei Zhang, Xinge Zhang and Zhongming Wu
Nanomaterials 2024, 14(16), 1363; https://doi.org/10.3390/nano14161363 - 19 Aug 2024
Viewed by 1136
Abstract
Obesity is a global epidemic which induces a multitude of metabolic disorders. Browning of white adipose tissue (WAT) has emerged as a promising therapeutic strategy for promoting weight loss and improving associated metabolic syndromes in people with obesity. However, current methods of inducing [...] Read more.
Obesity is a global epidemic which induces a multitude of metabolic disorders. Browning of white adipose tissue (WAT) has emerged as a promising therapeutic strategy for promoting weight loss and improving associated metabolic syndromes in people with obesity. However, current methods of inducing white adipose tissue browning have limited applicability. We developed a nanocomplex pTSL@(P+I), which is a temperature-sensitive liposome (TSL) surface-conjugated with an adipocyte-targeting peptide (p) and loaded with both browning-promoting agents (P) and photosensitizing agents (I). This nanocomplex exhibits adipocyte targeting, as well as synergistic pharmacological and photothermal properties to promote browning. pTSL@(P+I) effectively upregulates UCP1 and COX5B expression by activating the transcription axis of PPARγ/PGC1α and HSF1/PGC1α, thereby promoting white adipose tissue browning and reducing obesity. This novel nanocomplex exhibited a uniform spherical shape, with an average diameter of approximately 200 nm. Additionally, the nanocomplexes exhibited remarkable photothermal properties and biocompatibility. Further, when adipocytes were treated with pTSL@(P+I), their triglyceride content decreased remarkably and intracellular mitochondrial activity increased significantly. When applied to diet-induced obesity (DIO) mice, the nanocomplex exhibited significant efficacy, demonstrating a notable 14.4% reduction in body weight from the initial measurement, a decreased fat/lean mass ratio of 20.8%, and no statistically significant disparities (p > 0.05) in associated side effects when compared to the control group. In summary, implementation of the targeted nanocomplex pTSL@(P+I) to enhance energy expenditure by stimulating white adipose tissue browning offers a promising therapeutic approach for the treatment of obesity and related metabolic syndromes. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) <sup>1</sup>H NMR spectra of DSPE-PEG2000-Mal in CDCl<sub>3</sub> and DSPE-PEG2000-Pep in CD<sub>4</sub>O. (<b>B</b>) TEM images of targeted nanocomplex pTSL@(P+I). (<b>C</b>) Histogram of the particle size distribution of pTSL@(P+I) measured by DLS. (<b>D</b>) Potential distribution of the targeted nanocomplex measured by DLS. (<b>E</b>) UV–visible-near infrared (UV–VIS-NIR) spectrum of pTSL@(P+I). (<b>F</b>) Particle size stability of pTSL@(P+I) at 25 °C and 37 °C measured by DLS. Photothermal conversion curves (<b>G</b>) and photothermal images (<b>H</b>) of different concentrations of pTSL@(P+I) after irradiation with 808-nm NIR light (1.0 W/cm<sup>2</sup>) for various durations. (<b>I</b>) Photothermal conversion curves of pTSL@(P+I) at a concentration of 500 μg/mL under varying powers and durations when exposed to 808-nm NIR light. (<b>J</b>) Cumulative release curves of Piog from pTSL@(P+I) at 37 °C and 45 °C, respectively. (<b>K</b>) Hemolysis of different concentrations of pTSL@(P+I). (<b>L</b>) 3T3-L1 cell and differentiated 3T3-L1 cell survival rates measured by CCK-8 method after treatment with different concentrations of pTSL@(P+I) for 48 h.</p>
Full article ">Figure 2
<p>(<b>A</b>) After labeling TSL@(P+I) and pTSL@(P+I) with FITC fluorescent probes, CLSM images of differentiated 3T3-L1 cells following co-incubation with TSL@(P+I), pTSL@(P+I), and anti-prohibitin + pTSL@(P+I) for 1 h (left) or 3 h (right), respectively. (the DAPI-stained nucleus as blue fluorescence, the FITC-labeled nanocomplex as green fluorescence) (<b>B</b>) Quantification of the average fluorescence intensity (FI) of FITC in 3T3-L1 cells from each experimental group in (<b>A</b>) using Image J 1.8.0 software. (<b>C</b>) Schematic diagram of cell culture and intervention process. Oil Red O-stained images of 3T3-L1 cells induced differentiation and intervened in each group (<b>D</b>,<b>E</b>) and quantification of lipid droplets after extraction with isopropanol (<b>F</b>). (<b>G</b>) Quantification of the number of lipid droplets per cell following various interventions (<span class="html-italic">n</span> = 5). Concentrations of lipid triglyceride (TAG) (<b>H</b>) and free fatty acid (FFA) (<b>I</b>) in differentiated 3T3-L1 cells subjected to various interventions. Visualization of mitochondrial fluorescence (<b>J</b>) in differentiated 3T3-L1 cells following various interventions and quantification of the average fluorescence intensity (FI) (<b>K</b>) in cells from each experimental group in (<b>J</b>) using Image J 1.8.0 software. (the DAPI-stained nucleus as blue fluorescence, the Mito-Tracker as red fluorescence) * <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; ns = not significant. Statistical significance was assessed via two-way ANOVA (<b>B</b>) and one-way ANOVA (<b>F</b>–<b>K</b>).</p>
Full article ">Figure 3
<p>(<b>A</b>) Protein expression levels of PGC1α, HSF1, PPARγ, UCP1, and COX5B were assessed via Western blot analysis in each group of 3T3-L1 cells following induction of differentiation and intervention. (<b>B</b>–<b>F</b>) Quantitative analysis of the corresponding protein levels in (<b>A</b>) was conducted using Image J 1.8.0 software. The levels of inflammatory factors IL-1β (<b>G</b>) and TNF-α (<b>H</b>) were measured by ELISA in each group of 3T3-L1 cells following induction of differentiation and intervention. * <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; ns = not significant. Statistical significance was assessed via one-way ANOVA.</p>
Full article ">Figure 4
<p>(<b>A</b>) A schematic diagram illustrating the in vivo experiments conducted on mice (HI: Latin for hypodermic injection). (<b>B</b>) Representative effect plots depicting the outcomes observed in each group of diet-induced obesity mice following a 4-week treatment. Assessment of daily changes in body weight (<b>C</b>) and dietary food intake (<b>D</b>) of mice (<span class="html-italic">n</span> = 5) in each intervention group. (<b>E</b>) Analysis of the weight ratio of fat/lean at the experiment’s endpoint using a small animal body fat analyzer (<span class="html-italic">n</span> = 5). Typical anatomical drawings (<b>F</b>) and corresponding weights (<b>G</b>) of eWAT, iWAT and BAT of mice (<span class="html-italic">n</span> = 5) in each experimental group at the endpoint. (<b>H</b>) After subcutaneous injection of nanocomplexes into both sides of the iWAT of mice in each group, infrared thermal imaging images of the left inguinal region of mice were irradiated by 808 nm NIR light (0.5 W/cm<sup>2</sup>) at different times. (<b>I</b>) Quantitative analysis of photothermal conversion in the left inguinal area of mice in the PBS, pTSL@I, and pTSL@(P+I) groups at different times with irradiation of 808 nm NIR light (0.5 W/cm<sup>2</sup>) in (H) (<span class="html-italic">n</span> = 3). (<b>J</b>) Quantitative analysis of left and right groin temperature before and after laser irradiation of left groin in pTSL@I and pTSL@(P+I) groups in (H) (<span class="html-italic">n</span> = 3). * <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; ns = not significant. Statistical significance was assessed via one-way ANOVA.</p>
Full article ">Figure 5
<p>Glucose measurements and area under the curve of glucose tolerance test (GTT) (<b>A</b>,<b>B</b>) and insulin tolerance test (ITT) (<b>C</b>,<b>D</b>) in each group of diet-induced obesity mice after treatment (<span class="html-italic">n</span> = 5). The 24 h mean activity counts (<b>E</b>), mean and time-point O<sub>2</sub> consumption (<b>F</b>,<b>G</b>), mean and time-point CO<sub>2</sub> production (<b>H</b>,<b>I</b>), and mean respiratory quotient (RQ) (<b>J</b>) in mice analyzed by physiological metabolic cages (<span class="html-italic">n</span> = 3). Body temperatures of diet-induced obesity mice (<b>K</b>) and area under the curve (<b>L</b>) in each group during a 4 h cold tolerance test before the end of the experiment (<span class="html-italic">n</span> = 3). * <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; ns = not significant. Statistical significance was assessed via one-way ANOVA.</p>
Full article ">Figure 6
<p>(<b>A</b>–<b>C</b>) Levels of lipid triglyceride (TAG), TC, and LDL-C in each group of mice (<span class="html-italic">n</span> = 5) at the endpoint of the experiment. Protein expression levels of PGC1α, HSF1, PPARγ, UCP1, and COX5B in iWAT after intervention were detected using Western blotting in each group of mice (<b>D</b>) and quantitative analysis of corresponding protein levels by Image J 1.8.0 software (<b>E</b>–<b>I</b>). * <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; ns = not significant. Statistical significance was assessed via one-way ANOVA.</p>
Full article ">Figure 7
<p>Levels of iWAT inflammatory cytokines IL-1β (<b>A</b>) and TNF-α (<b>B</b>) in mice measured by ELISA. (<b>C</b>) H&amp;E staining of typical iWAT and eWAT of mice in each group after treatment. (<b>D</b>) Morphometric analysis of adipocyte cell sizes (<span class="html-italic">n</span> = 100 data points) in (<b>C</b>). Uncoupling protein 1 (UCP-1) labeled immunohistochemical staining (<b>E</b>) and quantification of UCP-1 expression (<b>F</b>) in (<b>E</b>) using Image J 1.8.0 software and UCP-1 labeled immunofluorescence images (<b>G</b>) and quantification of UCP-1 expression (<b>H</b>) in (<b>G</b>) using Image J 1.8.0 software of typical iWAT and eWAT of mice in each group after treatment. * <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; ns = not significant. Statistical significance was assessed via one-way ANOVA.</p>
Full article ">Scheme 1
<p>(<b>A</b>) Composition and structure of targeted nanocomplex pTSL@(P+I). (<b>B</b>) NIR light-activated targeted nanocomplex for white adipose tissue browning to enhance anti-obesity effects in mice.</p>
Full article ">Scheme 2
<p>Anti-obesity mechanisms of pTSL@(P+I)-mediated photothermal and pharmacological synergistic therapy.</p>
Full article ">
14 pages, 3308 KiB  
Article
Optimization of a Modular Nanotransporter Design for Targeted Intracellular Delivery of Photosensitizer
by Rena T. Alieva, Alexey V. Ulasov, Yuri V. Khramtsov, Tatiana A. Slastnikova, Tatiana N. Lupanova, Maria A. Gribova, Georgii P. Georgiev and Andrey A. Rosenkranz
Pharmaceutics 2024, 16(8), 1083; https://doi.org/10.3390/pharmaceutics16081083 - 18 Aug 2024
Viewed by 757
Abstract
Modular nanotransporters (MNTs) are drug delivery systems for targeted cancer treatment. As MNTs are composed of several modules, they offer the advantage of high specificity and biocompatibility in delivering drugs to the target compartment of cancer cells. The large carrier module brings together [...] Read more.
Modular nanotransporters (MNTs) are drug delivery systems for targeted cancer treatment. As MNTs are composed of several modules, they offer the advantage of high specificity and biocompatibility in delivering drugs to the target compartment of cancer cells. The large carrier module brings together functioning MNT modules and serves as a platform for drug attachment. The development of smaller-sized MNTs via truncation of the carrier module appears advantageous in facilitating tissue penetration. In this study, two new MNTs with a truncated carrier module containing either an N-terminal (MNTN) or a C-terminal (MNTC) part were developed by genetic engineering. Both new MNTs demonstrated a high affinity for target receptors, as revealed by fluorescent-labeled ligand-competitive binding. The liposome leakage assay proved the endosomolytic activity of MNTs. Binding to the importin heterodimer of each truncated MNT was revealed by a thermophoresis assay, while only MNTN possessed binding to Keap1. Finally, the photodynamic efficacy of the photosensitizer attached to MNTN was significantly higher than when attached to either MNTC or the original MNTs. Thus, this work reveals that MNT’s carrier module can be truncated without losing MNT functionality, favoring the N-terminal part of the carrier module due to its ability to bind Keap1. Full article
Show Figures

Figure 1

Figure 1
<p>Principal scheme of modular nanotransporter-mediated targeted intracellular delivery of photosensitizers (PS). Modular nanotransporters recognize target cells binding to cell-specific plasma membrane receptors via a ligand module. Following receptor internalization, modular nanotransporters exit endosomes with the help of an endosomolytic module and then can deliver their cargo to the target intracellular compartment (e.g., cell nucleus). PS—photosensitizer; ROS—reactive oxygen species.</p>
Full article ">Figure 2
<p>Binding of MNTs to EGFR of A431 cells. (<b>A</b>) Binding of Alexa647-MNT<sub>1</sub> to A431 cells; (<b>B</b>) and (<b>C</b>) represent concurrent binding of Alexa647-MNT<sub>1</sub> to MNT<sub>N</sub> and MNT<sub>C</sub>, respectively. The data of typical experiments performed in three repetitions is presented. The data are presented as average values ± SEM.</p>
Full article ">Figure 3
<p>MNT-induced dye leakage from phosphatidylcholine liposomes loaded with the fluorescent dye calcein to the fluorescence self-quenching concentration. The data are mean values ± SEM.</p>
Full article ">Figure 4
<p>The interaction of MNT<sub>S</sub> with Keap1-Cy3 assessed by thermophoresis. Dependences of relative fluorescence intensities (fluorescence intensity before the start of thermophoresis is taken as 100%) at 20 s after the start of thermophoresis on the concentration of the MNT<sub>C</sub> (<b>A</b>), MNT<sub>N</sub> (<b>B</b>), and MNT<sub>F</sub> (<b>C</b>) at a constant concentration of the Keap1-Cy3. Standard errors (SEs) of relative fluorescence intensities are shown (<span class="html-italic">n</span> = 12).</p>
Full article ">Figure 5
<p>The interaction of MNTs with α/β importin complex labeled with Cy3 assessed by thermophoresis. Dependences of relative fluorescence intensities (fluorescence intensity before the start of thermophoresis is taken as 100%) at 20 s after the start of thermophoresis on the concentration of MNT<sub>N</sub> (<b>A</b>) and MNT<sub>C</sub> (<b>B</b>) at a constant concentration of the importin complex. Standard errors (SE) of relative fluorescence intensities are shown (<span class="html-italic">n</span> = 7–9).</p>
Full article ">Figure 6
<p>Photocytotoxicity of MNT<sub>F</sub>-chlorin <span class="html-italic">e</span><sub>6</sub> (<b>A</b>), MNTc-chlorin <span class="html-italic">e</span><sub>6</sub> (<b>B</b>), MNT<sub>N</sub>-chlorin <span class="html-italic">e</span><sub>6</sub> (<b>C</b>), and free chlorin <span class="html-italic">e</span><sub>6</sub> (<b>D</b>) on A431 cells. Experimental data of six experiments (circles) fitted using the four-parameter logistic sigmoid curves (lines) are given. EC<sub>50</sub> values derived from sigmoid model approximation for each conjugate (<b>E</b>). Error bars represent standard errors (<span class="html-italic">n</span> = 6).</p>
Full article ">
29 pages, 1811 KiB  
Review
New Insights Concerning Phytophotodermatitis Induced by Phototoxic Plants
by Cristina Grosu (Dumitrescu), Alex-Robert Jîjie, Horaţiu Cristian Manea, Elena-Alina Moacă, Andrada Iftode, Daliana Minda, Raul Chioibaş, Cristina-Adriana Dehelean and Cristian Sebastian Vlad
Life 2024, 14(8), 1019; https://doi.org/10.3390/life14081019 - 16 Aug 2024
Viewed by 980
Abstract
The present review explores the underlying mechanisms of phytophotodermatitis, a non-immunologic skin reaction triggered by certain plants followed by exposure to ultraviolet radiation emitted by sunlight. Recent research has advanced our understanding of the pathophysiology of phytophotodermatitis, highlighting the interaction between plant-derived photosensitizing [...] Read more.
The present review explores the underlying mechanisms of phytophotodermatitis, a non-immunologic skin reaction triggered by certain plants followed by exposure to ultraviolet radiation emitted by sunlight. Recent research has advanced our understanding of the pathophysiology of phytophotodermatitis, highlighting the interaction between plant-derived photosensitizing compounds (e.g., furanocoumarins and psoralens) and ultraviolet light leading to skin damage (e.g., erythema, fluid blisters, edema, and hyperpigmentation), identifying these compounds as key contributors to the phototoxic reactions causing phytophotodermatitis. Progress in understanding the molecular pathways involved in the skin’s response to these compounds has opened avenues for identifying potential therapeutic targets suitable for the management and prevention of this condition. The review emphasizes the importance of identifying the most common phototoxic plant families (e.g., Apiaceae, Rutaceae, and Moraceae) and plant species (e.g., Heracleum mantegazzianum, Ruta graveolens, Ficus carica, and Pastinaca sativa), as well as the specific phytochemical compounds responsible for inducing phytophototoxicity (e.g., limes containing furocoumarin have been linked to lime-induced photodermatitis), underscoring the significance of recognizing the dangerous plant sources. Moreover, the most used approaches and tests for accurate diagnosis such as patch testing, Wood’s lamp examination, or skin biopsy are presented. Additionally, preventive measures such as adequate clothing (e.g., long-sleeved garments and gloves) and treatment strategies based on the current knowledge of phytophotodermatitis including topical and systemic therapies are discussed. Overall, the review consolidates recent findings in the field, covering a diverse array of phototoxic compounds in plants, the mechanisms by which they trigger skin reactions, and the implications for clinical management. By synthesizing these insights, we provide a comprehensive understanding of phytophotodermatitis, providing valuable information for both healthcare professionals and researchers working to address this condition. Full article
(This article belongs to the Section Plant Science)
Show Figures

Figure 1

Figure 1
<p>Schematic representation depicting Type I and Type II phototoxic response.</p>
Full article ">Figure 2
<p>Schematic representation of the developmental stages of phytophotodermatitis. The figures in the diagram were created using the Flaticon platform and Servier Medical Art (licensed under Creative Commons Attribution 3.0 Unported License).</p>
Full article ">
14 pages, 2351 KiB  
Article
Synthesis of Coumarin-Based Photosensitizers for Enhanced Antibacterial Type I/II Photodynamic Therapy
by Min Ma, Lili Luo, Libing Liu, Yuxuan Ding, Yixuan Dong and Bing Fang
Molecules 2024, 29(16), 3793; https://doi.org/10.3390/molecules29163793 - 10 Aug 2024
Viewed by 628
Abstract
Photodynamic therapy (PDT) is an effective method for treating microbial infections by leveraging the unique photophysical properties of photosensitizing agents, but issues such as fluorescence quenching and the restricted generation of reactive oxygen species (ROS) under hypoxic conditions still remain. In this study, [...] Read more.
Photodynamic therapy (PDT) is an effective method for treating microbial infections by leveraging the unique photophysical properties of photosensitizing agents, but issues such as fluorescence quenching and the restricted generation of reactive oxygen species (ROS) under hypoxic conditions still remain. In this study, we successfully synthesized and designed a coumarin-based aggregation-induced emission luminogen (AIEgen), called ICM, that shows a remarkable capacity for type I ROS and type II ROS generation. The 1O2 yield of ICM is 0.839. The ROS it produces include hydroxyl radicals (HO) and superoxide anions (O2•−), with highly effective antibacterial properties specifically targeting Staphylococcus aureus (a Gram-positive bacterium). Furthermore, ICM enables broad-spectrum fluorescence imaging and exhibits excellent biocompatibility. Consequently, ICM, as a potent type I photosensitizer for eliminating pathogenic microorganisms, represents a promising tool in addressing the threat posed by these pathogens. Full article
(This article belongs to the Special Issue Multifunctional Nanomaterials for Bioapplications, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Normalized UV–vis spectra of 10 μM ICM in H<sub>2</sub>O/DMSO; (<b>B</b>) normalized fluorescence spectra of 10 μM ICM in H<sub>2</sub>O/DMSO; (<b>C</b>) normalized fluorescence spectra in different solvents (excitation at 430 nm); (<b>D</b>) fluorescence spectra of ICM in DMSO solvent with different water fractions; (<b>E</b>) amplified fluorescence spectra at 654 nm; (<b>F</b>) line chart of relative fluorescence intensity (I/I<sub>0</sub>) at 500 nm and 654 nm, with I<sub>0</sub> at 90% or 0% water fractions; (<b>G</b>) zeta potential of ICM; (<b>H</b>) DLS data of ICM.</p>
Full article ">Figure 2
<p>(<b>A</b>) Fluorescence chart for the detection of total ROS with 50 nM ICM and DCFH under 5 mW/cm<sup>2</sup> white light irradiation, with measurements taken every minute; (<b>B</b>) detection of <sup>1</sup>O<sub>2</sub> generation with enhancement of SOSG for 0.5 μM ICM under 5 mW/cm<sup>2</sup> light irradiation; (<b>C</b>) detection of O<sub>2</sub><sup>•−</sup> generation with the decomposition of DHR123 for 0.5 μM ICM under 5 mW/cm<sup>2</sup> light irradiation; (<b>D</b>) detection of HO<sup>•</sup> with 0.5 μM ICM and HPF upon 5 mW/cm<sup>2</sup> white light irradiation; (<b>E</b>) calculated frontier molecular orbitals, with the ΔE<sub>S1T1</sub> value for ICM used in the Gaussian 09 program package at the B3LPY/6-31G (d, p) level.</p>
Full article ">Figure 3
<p>(<b>A</b>) <span class="html-italic">S. aureus</span> and different concentrations of ICM on agar plates under light (5 mW/cm<sup>2</sup>) and dark conditions; (<b>B</b>) survival rate graph of <span class="html-italic">S. aureus</span> treated with different concentrations of ICM; (<b>C</b>) SEM images of <span class="html-italic">S. aureus</span> incubated with various concentrations of ICM (scale bar: 1 μm); (<b>D</b>) zeta potential of <span class="html-italic">S. aureus</span> treated with different concentrations of ICM; (<b>E</b>) CLSM images of ICM interacting with different pathogenic microorganisms (scale bar: 10 μm).</p>
Full article ">Scheme 1
<p>(<b>A</b>) Molecular structures of ICM. (<b>B</b>) Illustration of photophysical mechanisms of Type I and Type II photosensitizers and antibacterial application of ICM.</p>
Full article ">Scheme 2
<p>The synthetic route of ICM.</p>
Full article ">
Back to TopTop