[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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

Journals

Article Types

Countries / Regions

Search Results (149)

Search Parameters:
Keywords = cytosolic delivery

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 4965 KiB  
Article
T14diLys/DOPE Liposomes: An Innovative Option for siRNA-Based Gene Knockdown?
by Sophie Meinhard, Frank Erdmann, Henrike Lucas, Maria Krabbes, Stephanie Krüger, Christian Wölk and Karsten Mäder
Pharmaceutics 2025, 17(1), 25; https://doi.org/10.3390/pharmaceutics17010025 - 27 Dec 2024
Viewed by 456
Abstract
Background/Objectives: Bringing small interfering RNA (siRNA) into the cell cytosol to achieve specific gene silencing is an attractive but also very challenging option for improved therapies. The first step for successful siRNA delivery is the complexation with a permanent cationic or ionizable compound. [...] Read more.
Background/Objectives: Bringing small interfering RNA (siRNA) into the cell cytosol to achieve specific gene silencing is an attractive but also very challenging option for improved therapies. The first step for successful siRNA delivery is the complexation with a permanent cationic or ionizable compound. This protects the negatively charged siRNA and enables transfection through the cell membrane. The current study explores the performance of the innovative, ionizable lipid 2-Tetradecylhexadecanoic acid-(2-bis{[2-(2,6-diamino-1-oxohexyl)amino]ethyl}aminoethyl)-amide (T14diLys), in combination with 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE), for siRNA delivery and the impact of the production method (sonication vs. extrusion) on the particle properties. Methods: Liposomes were produced either with sonication or extrusion and characterized. The extruded liposomes were combined with siRNA at different N/P ratios and investigated in terms of size zeta potential, encapsulation efficiency, lipoplex stability against RNase A, and knockdown efficiency using enhanced green fluorescent protein (eGFP)-marked colon adenocarcinoma cells. Results: The liposomes prepared by extrusion were smaller and had a narrower size distribution than the sonicated ones. The combination of siRNA and liposomes at a nitrogen-to-phosphate (N/P) ratio of 5 had optimal particle properties, high encapsulation efficiency, and lipoplex stability. Gene knockdown tests confirmed this assumption. Conclusions: Liposomes produced with extrusion were more reproducible and provided enhanced particle properties. The physicochemical characterization and in vitro experiments showed that an N/P ratio of 5 was the most promising ratio for siRNA delivery. Full article
(This article belongs to the Special Issue Drug Nanocarriers for Pharmaceutical Applications)
Show Figures

Figure 1

Figure 1
<p>Impact of the manufacturing process on Z-Average and Polydispersity index (PDI) between T14diLys:DOPE (1:2) liposome (0.05 µg/µL) preparation with extrusion and sonication in 10 mM MES buffer pH 6.5, three samples per batch with three measurements each.</p>
Full article ">Figure 2
<p>(<b>A</b>) Impact of storage time on size and polydispersity for 28 days of T14diLys:DOPE (1:2) liposomes (0.05 µg/µL) prepared with extrusion in 10 mM MES buffer pH 6.5, n = 3. The error bars are within the limits of the symbols. (<b>B</b>) Results of zeta potential of three T14diLys:DOPE (1:2) liposome (0.6 µg/µL) samples of two batches in 10 mM MES buffer pH 6.5, n = 3.</p>
Full article ">Figure 3
<p>Impact of N/P ratio on particle size, polydispersity (<b>A</b>), and zeta potential (<b>B</b>) in 10 mM MES buffer pH 6.5 of T14diLys:DOPE (1:2) lipoplexes (5 ng/µL siRNA per sample), n = 3. Red dots correspond to the mean value of the three black dots, which represents the PDI.</p>
Full article ">Figure 4
<p>Separation of lipoplex preparations in agarose gel electrophoresis as a function of different N/P ratios: The complexation efficiency of siRNA by T14diLys:DOPE (1:2). Lipoplex = LPX. At NP ratios &gt; 1, no band visible.</p>
Full article ">Figure 5
<p>TEM images of T14diLys:DOPE (1:2) lipoplexes in 10 mM MES buffer pH 6.5 after adding uranyl acetate for negative staining at N/P 2 (<b>A</b>), N/P 3 (<b>B</b>)<b>,</b> N/P 4 (<b>C</b>), and N/P 5 (<b>D</b>)—scale bar in (<b>A</b>,<b>B</b>) represents 500 nm—and (<b>C</b>,<b>D</b>) 250 nm Cryo-TEM images of T14diLys:DOPE (1:2) lipoplexes in 10 mM MES buffer pH 6.5 at N/P 2 (<b>E</b>), N/P 3 (<b>F</b>), N/P 4 (<b>G</b>), and N/P 5 (<b>H</b>)—scale bar in (<b>E</b>–<b>H</b>) represents 200 nm. Black arrows indicate for lipoplexes, white arrows for uncomplexed liposomes.</p>
Full article ">Figure 6
<p>Agarose gel electrophoresis chromatogram of T14diLys:DOPE (1:2) lipoplexes (LPX) at N/P 5 after incubation with RNase A (R-A) and release of stable siRNA out of lipoplex with Heparin after RNase A and RNase Inhibitor (R-IH) treatment. The red boxes highlight the important bands.</p>
Full article ">Figure 7
<p>(<b>A</b>) Boxplot of eGFP expression level of (N/P 2–5) T14diLys:DOPE (1:2) lipoplexes at different N/P ratios after 72 h incubation (100 nM siRNA); (NC) negative control = scrambled siRNA + Lipofectamine2000 and (PC) positive control = eGFP siRNA with Lipofectamine2000 and (UT) untreated eGFP-DLD1 cells. eGFP expression was analyzed as duplicates. The experiment was repeated 3 times independently. Untreated eGFP-DLD1 cells are set to 100%. (<b>B</b>) Fluorescence images with Cytation 5 after 72 h from transfection of eGFP-DLD1 cells, A1 eGFP siRNA with Lipofectamine2000, and scrambled siRNA with Lipofectamine2000 (A2) with Lipofectamine2000, A3 untreated eGFP-DLD1 cells, A4 wildtype DLD1 cells. B1–4 T14 diLys:DOPE (1:2) lipoplexes N/P 2–5, C1–4 liposome amount N/P 2–5. 200 × zoom, scale bars = 100 µm. (<b>C</b>): Flow cytometry data as dot plots (<b>C1</b>–<b>C4</b>) and histograms (<b>C5</b>–<b>C8</b>) from (<b>C1</b> + <b>C5</b>): wildtype DLD1 cells, (<b>C2</b> + <b>C6</b>): eGFP-DLD1 cells, (<b>C3</b> + <b>C7</b>): eGFP-DLD1 cells with Lipofectamine2000 + siRNA, (<b>C4</b> + <b>C6</b>): eGFP-DLD1 cells with T14diLys:DOPE (1:2) + siRNA N/P 5. M7, 4, and 1 represent the GFP intensity in (<b>C5</b>–<b>C8</b>).</p>
Full article ">Figure 8
<p>eGFP expression level of eGFP-DLD1 cells at different N/P ratios (100 nM siRNA) after 72 h incubation with and without sterile filtration of T14diLys:DOPE (1:2) liposomes before transfection using 0.20 µm regenerative cellulose filter membrane. The (NC) negative control = scrambled siRNA + Lipofectamine2000 and (PC) positive control = eGFP-siRNA were combined with Lipofectamine2000, (UT) untreated eGFP-DLD1 cells. eGFP expression was analyzed as duplicates. The experiment was repeated 3 times independently. Untreated eGFP-DLD1 cells are set to 100%.</p>
Full article ">Figure 9
<p>eGFP expression level at increasing N/P ratios (100 nM siRNA) of T14diLys:DOPE (1:2) lipoplexes after 72 h incubation; UT = untreated eGFP-DLD1 cells.</p>
Full article ">Figure 10
<p>Results of cytotoxicity of T14diLys/DOPE (1:2) lipoplexes at N/P 5 and 10 with siRNA concentrations of 1, 2.8, and 10 µg/mL after 24 h and 96 h incubation in NHDF and 3T3 cells. (<b>A</b>) NHDF after 24 h, (<b>B</b>) NHDF after 96 h, (<b>C</b>) 3T3 after 24 h, (<b>D</b>) 3T3 after 96 h. The vital control contains just cells and medium representing the 100% vital cells, and the positive control includes also Triton X100, which represents the most toxic effect possible. Results are presented as mean ± SD, n = 4.</p>
Full article ">
22 pages, 1150 KiB  
Review
Endosomal Escape and Nuclear Localization: Critical Barriers for Therapeutic Nucleic Acids
by Randall Allen and Toshifumi Yokota
Molecules 2024, 29(24), 5997; https://doi.org/10.3390/molecules29245997 - 19 Dec 2024
Viewed by 1126
Abstract
Therapeutic nucleic acids (TNAs) including antisense oligonucleotides (ASOs) and small interfering RNA (siRNA) have emerged as promising treatment strategies for a wide variety of diseases, offering the potential to modulate gene expression with a high degree of specificity. These small, synthetic nucleic acid-like [...] Read more.
Therapeutic nucleic acids (TNAs) including antisense oligonucleotides (ASOs) and small interfering RNA (siRNA) have emerged as promising treatment strategies for a wide variety of diseases, offering the potential to modulate gene expression with a high degree of specificity. These small, synthetic nucleic acid-like molecules provide unique advantages over traditional pharmacological agents, including the ability to target previously “undruggable” genes. Despite this promise, several biological barriers severely limit their clinical efficacy. Upon administration, TNAs primarily enter cells through endocytosis, becoming trapped inside membrane-bound vesicles known as endosomes. Studies estimate that only 1–2% of TNAs successfully escape endosomal compartments to reach the cytosol, and in some cases the nucleus, where they bind target mRNA and exert their therapeutic effect. Endosomal entrapment and inefficient nuclear localization are therefore critical bottlenecks in the therapeutic application of TNAs. This review explores the current understanding of TNA endosomal escape and nuclear transport along with strategies aimed at overcoming these challenges, including the use of endosomal escape agents, peptide-TNA conjugates, non-viral delivery vehicles, and nuclear localization signals. By improving both endosomal escape and nuclear localization, significant advances in TNA-based therapeutics can be realized, ultimately expanding their clinical utility. Full article
(This article belongs to the Section Chemical Biology)
Show Figures

Figure 1

Figure 1
<p>Uptake and intracellular trafficking of TNAs. Following endocytosis, TNAs become encapsulated inside early endosomes (EE) which undergo maturation to multivesicular bodies (MVBs) and late endosomes (LEs). Non-productive pathways (red) do not permit TNAs to reach their intracellular targets. Such pathways include recycling to the plasma membrane, retention in depot endosomes, or enzymatic degradation in lysosomes. Productive pathways (green) allow the successful escape of TNAs into the cytosol to interact with mRNA targets, or eventually the nucleus when targeting pre-mRNA. A small portion (1–2%) of freely delivery TNAs escape endosomes during trafficking primarily from MVBs and LEs.</p>
Full article ">Figure 2
<p>Traditional endosomal escape strategies. (<b>A</b>) Cationic amphiphilic small molecules (CADs) such as chloroquine enter endosomes buffering changes in pH. The resulting proton-sponge effect induces osmotic swelling and endosomal rupture, allowing TNAs to escape. Other small molecules may directly interact with endosomes causing membrane destabilization. (<b>B</b>) Peptide-mediated endosomal escape can be facilitated by biomimetic or cell-penetrating peptides. Interaction between cationic peptides and the anionic endosomal membrane causes fusion, membrane destabilization, or pore formation. (<b>C</b>) Non-viral delivery vehicle-mediated endosomal escape can be achieved using lipid nanoparticles. Cationic or ionizable lipids facilitate fusion with the endosomal membrane, allowing TNA release.</p>
Full article ">Figure 3
<p>Alternative strategies to the endosomal escape problem. (1) Direct cytosolic entry of TNAs which can be facilitated by peptide or nanoparticle-mediated delivery. Avoiding endocytosis and subsequent endosomal entrapment allows free translocation to the nucleus. (2) Inhibition of endosomal recycling can be accomplished through small molecules such as NP3.47. Preventing the exocytosis of internalized TNAs provides increased potential for endosomal escape events. (3) Inhibition of endo-Golgi retrograde transport with small molecules such as Retro-1. The mechanism of action remains unclear but may increase the retention of TNAs in endosomes, increasing the probability for escape. (4) Inhibition of endo-lysosomal fusion with molecules such as SH-BC-893. Preventing the degradation of TNAs entrapped in endosomes increases their cytosolic quantity, improving treatment efficacy.</p>
Full article ">Figure 4
<p>Mechanism nuclear localization signal (NLS) internalization. TNA-NLS peptide conjugates interact with importin-α and β in the cytosol which facilitate active transport through the nuclear pore complex (NPC). Following nuclear entry, the binding of RanGTP causes the dissociation of the complex. The free TNA-NLS is now capable of binding to target pre-mRNA in the nucleus.</p>
Full article ">
26 pages, 2674 KiB  
Review
Guardians of the Genome: Iron–Sulfur Proteins in the Nucleus
by Lorena Novoa-Aponte, Andres Leon-Torres and Caroline C. Philpott
Inorganics 2024, 12(12), 316; https://doi.org/10.3390/inorganics12120316 - 6 Dec 2024
Viewed by 605
Abstract
Iron–sulfur (Fe-S) clusters are essential cofactors found in many proteins in the mitochondria, cytosol, and nucleus of the cell. These versatile cofactors may undergo reversible oxidation–reduction reactions to enable electron transfers; they may be structural and confer stability to a folded protein; they [...] Read more.
Iron–sulfur (Fe-S) clusters are essential cofactors found in many proteins in the mitochondria, cytosol, and nucleus of the cell. These versatile cofactors may undergo reversible oxidation–reduction reactions to enable electron transfers; they may be structural and confer stability to a folded protein; they may be regulatory and transduce an iron signal that alters the function or stability of a recipient protein. Of the nearly 70 proteins described in mammalian cells that bind Fe-S clusters, about half localize exclusively or partially to the nucleus, where they are required for DNA replication and repair, telomere maintenance, transcription, mitosis, and cell cycle control. Most nuclear Fe-S cluster proteins interact with DNA, including DNA polymerases, primase, helicases, and glycosylases. However, the specific roles of the clusters in the enzymatic activities of these proteins and their interplay with DNA remain a matter of debate. Defects in the metallation of nuclear Fe-S proteins cause genome instability and alter the regulation of cell division and proliferation, which are hallmarks of various genetic diseases and cancers. Here, we provide an inventory of the nuclear Fe-S cluster-binding proteins and discuss cluster types, binding sites, the process of cluster acquisition, and the potential roles of the cluster in the function of the proteins. However, many questions remain unresolved. We highlight critical gaps in our understanding of cluster delivery to nuclear client proteins, the potential for cluster repair, and the mechanistic roles that clusters play in these enzymes. Taken together, this review brings the focus to the nucleus of the human cell as a hotspot for Fe-S cluster proteins and aims to inspire new research on the roles of iron in DNA metabolism and the maintenance of genome integrity. Full article
(This article belongs to the Special Issue Iron-Sulfur Clusters: Assembly and Biological Roles)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A comparison of solvent-exposed and buried Fe-S clusters in proteins. The structures (left) and close-ups of the Fe-S binding domains (square, right). (<b>A</b>) X-ray structure of human xanthine oxidase (XDH) showing the two [2Fe-2S] clusters deeply buried within the protein structure (PDB 2CKJ (<a href="https://www.rcsb.org/structure/2CKJ" target="_blank">https://www.rcsb.org/structure/2CKJ</a>, accessed on 30 September 2024)). (<b>B</b>) Cryo-EM structure of the human ribosome recycling factor ABCE1 showing the two [4Fe-4S] clusters situated in solvent-exposed loops at the N-terminal domain (PDB 7A09 (<a href="https://www.rcsb.org/structure/7A09" target="_blank">https://www.rcsb.org/structure/7A09</a>, accessed on 1 October 2024)). The clusters in ABCE1 are prone to degradation [<a href="#B102-inorganics-12-00316" class="html-bibr">102</a>]. Each cluster in ABCE1 is coordinated by three nearby Cys residues and one distant Cys residue in the primary sequence, suggesting a structural role for these clusters [<a href="#B103-inorganics-12-00316" class="html-bibr">103</a>]. (<b>C</b>) Cryo-EM structure of human DNA primase subunits, PRIM1 (green) and PRIM2 (blue) (PDB 8QJ7 (<a href="https://www.rcsb.org/structure/8QJ7" target="_blank">https://www.rcsb.org/structure/8QJ7</a>, accessed on 2 October 2024)). The [4Fe-4S] cluster bound to PRIM2 locates at the junction of the N-terminal and C-terminal domains close to the protein surface. PRIM1 is proposed to serve as an adaptor for the recognition of PRIM2 by the CTC to facilitate its metallation. (<b>D</b>) Cryo-EM structure of the yeast elongator complex proteins ELP3 (green) and ELP4 (yellow) (PDB 8ASV (<a href="https://www.rcsb.org/structure/8ASV" target="_blank">https://www.rcsb.org/structure/8ASV</a>, accessed on 2 October 2024)). The radical SAM enzyme ELP3 has a [4Fe-4S](Cys)<sub>3</sub> cluster, which is exposed to solvent via two channels [<a href="#B103-inorganics-12-00316" class="html-bibr">103</a>]. ELP4 is proposed to serve as an adaptor for the recognition of ELP3 by the CTC to facilitate its metallation.</p>
Full article ">Figure 2
<p>The essential roles of nuclear Fe-S cluster-binding proteins in DNA replication, telomere maintenance, and mitosis. (<b>A</b>) A representation of a eukaryotic DNA replication fork. The mini-chromosome maintenance (MCM2-7) helicase complex unwinds DNA at replication origins, separating the double helix to generate the leading and lagging strands. All replicative DNA polymerases, Pol α, δ, and ε, bind a [4Fe-4S] cluster. DNA replication initiates with an RNA primer (orange) synthesized by DNA primase, with the regulatory subunit PRIM2 also binding a [4Fe-4S] cluster. Pol ε further extends the primer on the leading strand, while Pol δ does so on the lagging strand. The proliferating cell nuclear antigen sliding clamp PCNA enhances the processivity of DNA polymerases and facilitates DNA repair. The [4Fe-4S] cluster DNA helicases FANCJ and DDX11 unwind G4s to prevent stalling of the replisome. DNA lesions obstruct replication, triggering monoubiquitylation of PCNA, which recruits the DNA damage bypass DNA polymerase Pol ζ. The catalytic subunit of Pol ζ also binds a [4Fe-4S] cluster. Finally, the [4Fe-4S] cluster nuclease–helicase DNA2 cleaves exposed single-stranded DNA ends from Okazaki fragments and stalled replication forks. (<b>B</b>) Under Fe-sufficient conditions, the MCM helicase moves in opposite directions from the activated origin of replication. This results in the formation of two functional replication forks, enabling DNA synthesis. Under iron-depleted conditions, apo NCOA4 accumulates in the nucleus and binds to the helicase subunit MCM7, hindering MCM helicase activity and inactivating replication origins. This ensures that DNA synthesis only occurs when there is a sufficient pool of metallated Fe-S enzymes, thereby maintaining genomic integrity and preventing replication under suboptimal conditions. (<b>C</b>) The [4Fe-4S] cluster helicase RTEL1 unwinds G4s and R-loops at the telomeric repeats, facilitating telomere replication by telomerase and preventing telomere shortening. (<b>D</b>) During the late stages of mitosis (metaphase, anaphase, and telophase), the nuclear [4Fe-4S] cluster mitotic factor KIF4A binds the arms of condensed chromosomes. As a kinesin, KIF4A moves along microtubules to mobilize cargoes. KIF4A localizes at the central spindle, accumulating at the spindle midzone and midbody. The close-up image on the bottom illustrates the colocalization of the CIA targeting factors MMS19 and CIAO2B with KIF4A during mitosis, suggesting in situ metallation of KIF4A during function.</p>
Full article ">Figure 3
<p>The Fe-S cluster proteins within the six major DNA repair pathways: base excision repair (BER), nucleotide excision repair (NER), mismatch repair (MMR), homologous recombination (HR), non-homologous end joining (NHEJ), and translesion DNA synthesis (TLS). A schematic representation of DNA lesions and the corresponding repair pathways. The key proteins involved in each pathway are organized according Knijnenburg et al. [<a href="#B172-inorganics-12-00316" class="html-bibr">172</a>]. DNA repair proteins that bind [4Fe-4S] clusters are shown in a blue square at the bottom. * While DDX11 and EXO5 are not formally categorized under HR, their functions align with this repair mechanism.</p>
Full article ">
19 pages, 4810 KiB  
Article
Endosomal pH, Redox Dual-Sensitive Prodrug Micelles Based on Hyaluronic Acid for Intracellular Camptothecin Delivery and Active Tumor Targeting in Cancer Therapy
by Huiping Zhang, Liang Li, Wei Li, Hongxia Yin, Huiyun Wang and Xue Ke
Pharmaceutics 2024, 16(10), 1327; https://doi.org/10.3390/pharmaceutics16101327 - 14 Oct 2024
Viewed by 1449
Abstract
Background: CPT is a pentacyclic monoterpene alkaloid with a wide spectrum of antitumor activity. Its clinical application is restricted due to poor water solubility, instability, and high toxicity. We developed a new kind of multifunctional micelles to improve its solubility, reduce the side [...] Read more.
Background: CPT is a pentacyclic monoterpene alkaloid with a wide spectrum of antitumor activity. Its clinical application is restricted due to poor water solubility, instability, and high toxicity. We developed a new kind of multifunctional micelles to improve its solubility, reduce the side effecs, and obtain enhanced antitumor effects. Methods: We constructed HA-CPT nano-self-assembly prodrug micelles, which combined the advantages of pH-sensitivity, redox-sensitivity, and active targeting ability to CD44 receptor-overexpressing cancer cells. To synthesize dual sensitive HA-CPT conjugates, CPT was conjugated with HA by pH-sensitive histidine (His) and redox-sensitive 3,3′-dithiodipropionic acid (DTPA). In vitro, we studied the cellular uptake and antitumor effect for tumor cell lines. In vivo, we explored the bio-distribution and antitumor effects of the micelles in HCT 116 tumor bearing nude mice. Results: The dual-sensitive and active targeting HA-His-ss-CPT micelles was proved to be highly efficient in CPT delivery by the in vitro cellular uptake study. The HA-His-ss-CPT micelles escaped from endosomes of tumor cells within 4 h after cellular uptake due to the proton sponge effect of the conjugating His and then quickly released CPT in the cytosol by glutathione (GSH). In mice, HA-His-ss-CPT micelles displayed efficient tumor accumulation and conspicuous inhibition of tumor growth. Conclusions: The novel, dual-sensitive, active targeting nano-prodrug micelles exhibited high efficiency in drug delivery and cancer therapy. This “all in one” drug delivery system can be realized in an ingenious structure and avoid intricate synthesis. This construction strategy can illume the design of nanocarriers responding to endogenous stimuli in tumors. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Chemical structure of endosomal pH and redox dual-sensitive amphiphilic CPT prodrugs and construction of HHSC micelles; (<b>B</b>) schematic illustration of self-active targeting HHSC micelles for endosomal escape, GSH-triggered release of CPT and significant antitumor activity.</p>
Full article ">Figure 2
<p>The synthesis route of HA-His-ss-CPT (HHSC).</p>
Full article ">Figure 3
<p>1H NMR spectrum of HHSC.</p>
Full article ">Figure 4
<p>Effect of pH on particle size (<b>A</b>) and zeta potential (<b>B</b>) of HHSC micelles. All data represent mean ± SD (n = 3); (<b>C</b>) The intensity ratios I339/I333 from pyrene excitation spectra of HHSC micelles; (<b>D</b>) Percentage of CPT released from HHSC micelles in PBS buffer at pH 7.4 containing different concentrations of GSH and at pH 5.0 containing 40 mM GSH. Error bars indicate SD (n = 3).</p>
Full article ">Figure 5
<p>Intracellular tracking of HHSC micelles incubated with HCT116 cells for 0.5 h and 4 h. Lysotracker was used to stain the endosomes of the cell. SYTOX was used to stain the nuclei of the cell. Scale bars = 50 μm.</p>
Full article ">Figure 6
<p>Cellular uptake and intracellular localization of HHSC micelles, HC micelles, and free CPT in HCT116 cells by CLSM at 0.5 h (<b>A</b>) and 4 h (<b>B</b>). Scale bars = 50 μm.</p>
Full article ">Figure 7
<p>Flow cytometry profiles (<b>A</b>,<b>B</b>) and fluorescence intensity (<b>C</b>) of HCT 116 cells treated with free CPT, HC micelles, and HHSC micelles for 0.5 h and 4 h. (<b>D</b>) Fluorescence intensity of HCT 116 cells treated with HHSC micelles with or without free HA for 4 h. Error bars indicate SD (n = 3). ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>In vitro cytotoxicities of HHSC micelles, HC micelles, and CPT against HCT116 tumor cells.</p>
Full article ">Figure 9
<p>(<b>A</b>) Time-dependent in vivo NIRF images of HCT 116 tumor-bearing mice after intravenous injection of free DiR or DiR-labeled HC micelles or HHSC micelles; (<b>B</b>) ex vivo images of the main organs and tumors excised at 96 h post intravenous injection; (<b>C</b>) semi-quantitive analysis of the mean fluorescence intensity in the main organs and tumors excised at 96 h post intravenous injection. Error bars indicate SD (n = 3). *: <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 10
<p>In vivo therapeutic efficacy of HCT116 tumor-bearing nude mice after intravenous injection of saline and different CPT forms at an equivalent dose of 5 mg/kg. (<b>A</b>) Tumor growth curves (***: <span class="html-italic">p</span> &lt; 0.001), (<b>B</b>) body weight changes, (<b>C</b>) photos of tumors from different treatment groups excised on day 10, (<b>D</b>) tumor weights (****: <span class="html-italic">p</span> &lt; 0.0001). All data represented mean ± SD (n = 5).</p>
Full article ">
12 pages, 2878 KiB  
Article
Fusogenic Liposomes for the Intracellular Delivery of Phosphocreatine
by Okhil K. Nag, Eunkeu Oh and James B. Delehanty
Pharmaceuticals 2024, 17(10), 1351; https://doi.org/10.3390/ph17101351 - 10 Oct 2024
Viewed by 1041
Abstract
Background/Objective: Maintaining intracellular adenosine triphosphate (ATP) levels is essential for numerous cellular functions, including energy metabolism, muscle contraction, and nerve impulse transmission. ATP is primarily synthesized in mitochondria through oxidative phosphorylation. It is also generated in the cytosol under anaerobic conditions using phosphocreatine [...] Read more.
Background/Objective: Maintaining intracellular adenosine triphosphate (ATP) levels is essential for numerous cellular functions, including energy metabolism, muscle contraction, and nerve impulse transmission. ATP is primarily synthesized in mitochondria through oxidative phosphorylation. It is also generated in the cytosol under anaerobic conditions using phosphocreatine (PCr) as a phosphate donor to adenosine diphosphate. However, the intracellular delivery of exogenous PCr is challenging as it does not readily cross the plasma membrane. This complicates the use of PCr as a therapeutic agent to maintain energy homeostasis or to treat conditions like cerebral creatine deficiency syndrome (CDS), which results from defective creatine transporters. Methods: This study employs the use of fusogenic liposomes to deliver PCr directly into the cytosol, bypassing membrane impermeability issues. We engineered various 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP)-based fusogenic liposomes, incorporating phospholipids such as 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE) in combination with phospholipid-aromatic dye components to facilitate membrane fusion and to enhance the delivery of the PCr cargo. Liposomal formulations were co-loaded with membrane-impermeable chromophores and PCr and studied on live cells using confocal microscopy. Conclusions: We demonstrated the successful intracellular delivery of these agents and observed a 23% increase in intracellular ATP levels in cells treated with PCr-loaded liposomes. This increase was not observed with free PCr, confirming the effectiveness of the liposome-based delivery system. Additionally, cell viability assays showed minimal toxicity from the liposomes. Our results indicate that fusogenic liposomes are a promising method for the delivery of PCr (and potentially other cell-impermeable therapeutic agents) to the cellular cytosol. The approach demonstrated here could be advantageous for treating energy-related disorders and improving cellular energy homeostasis. Full article
Show Figures

Figure 1

Figure 1
<p>Physical characterization of fusogenic liposomes for PCr delivery to cells. (<b>A</b>–<b>D</b>) Dynamic light scattering data showing the distribution of the hydrodynamic diameter of the various liposomal preparations. (<b>E</b>) Table summarizing the average diameter, polydispersity index (PDI), and zeta potential of the liposome formulations.</p>
Full article ">Figure 2
<p>Intracellular delivery of calcein with DOPE/DOTAP/Rhod-PE liposomes. (<b>A</b>) DIC and confocal images of live HEK 293T/17 cells after 20 min of incubation with DOPE/DOTAP/Rhod-PE/calcein. Rhod-PE (red) membranous signal, confirming the fusion of the liposomal lipids with the plasma membrane coupled with the intracellular delivery of the cell-impermeable fluorophore, calcein (green). (<b>B</b>) Time-dependent images of the same cells after adding free calcein to the media. The images show the extracellular accumulation of calcein (squares) over 30 min while the intracellular calcein signal remains constant (circles). The scale bar is 20 µm. (<b>C</b>) Quantification of the images in (<b>B</b>) showing the time-dependent extracellular accumulation of calcein (squares) while the intracellular calcein signal remains constant (circles).</p>
Full article ">Figure 3
<p>Intracellular delivery of DAPI with DOPE/DOTAP/NBD-PE liposomes. (<b>A</b>) Time-resolved imaging of HEK 293T/17 cells shows clear NBD-PE staining of the plasma membrane, confirming successful liposome–plasma membrane fusion and DAPI staining of nuclei, demonstrating DAPI’s delivery to the cytosol and nucleus. The quantification of cellular NBD-PE (<b>B</b>) and DAPI (<b>C</b>) staining shows that while the NBD-PE staining plateaus at ~15 min, the DAPI signal continues to increase over the 30 min imaging window. The scale bar is 20 µm.</p>
Full article ">Figure 4
<p>PCr-loaded fusogenic liposomes mediate increased ATP production in HEK 293T/17 cells. (<b>A</b>) Micrograph showing the increased red fluorescence of the ATP-Red probe in cells 30 min after incubation with DOPE/DOTAP/NBD-PE/PCr liposomes. The white arrows denote specific cells exhibiting increased fluorescence. The scale bar is 20 µm. (<b>B</b>) Time-resolved quantification of increased ATP production in PCr(+) and PCr(−) cells over 30 min period. (<b>C</b>) Overall increase in ATP production in cells incubated with PCr(+) liposomes without and with DAPI. Controls were cells incubated with liposomes containing no PCr and free PCr. Data are plotted as average (±SD) change in the fluorescence intensity from 0 min to 30 min, collected from ~60 cells in two separate experiments (****, <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 5
<p>Cellular viability of HEK 293T/17 cells incubated with liposomal formulations. Cellular viability of HEK 293T/17 cells incubated with various liposome formulations with an increasing concentration (0.37–6 pm). Viability was determined using MTS colorimetric assay after liposomes were incubated on the cells and the cells were allowed to proliferate for 72 h. Average (n = 4 ± SEM) cell viability for different concentrations of liposomes were plotted as a percentage of control cells (not treated).</p>
Full article ">Scheme 1
<p>Schematic of fusogenic liposomal delivery of phosphocreatine (PCr) to the cellular cytosol. (<b>A</b>) Structures of the phospholipids and cargos (calcein, DAPI, and PCr) used for fusogenic liposome synthesis and labeling. DOTAP and DOPE are cationic and nonionic unsaturated lipids, respectively. The fluorescent phospholipids Rhod-PE and NBD-PE facilitate both fusion to the plasma membrane and the tracking of the fusion of liposomal membrane to the plasma membrane. (<b>B</b>,<b>C</b>) An illustration of calcein, DAPI, and PCr interactions with the plasma membrane when delivered from the bulk solution (<b>B</b>) and loaded into the fusogenic liposomes (<b>C</b>). Due to their charge or amphipathic properties, free calcein, DAPI, and PCr do not cross the plasma membrane. Fusogenic liposomes loaded with calcein, DAPI, and PCr are delivered into the cytosol and further traverse to subcellular locations. The PCr delivered into the cytosol via fusogenic liposomes drives ATP production.</p>
Full article ">
20 pages, 4599 KiB  
Review
Recent Advances in Lipid Nanoparticles and Their Safety Concerns for mRNA Delivery
by Jialiang Wang, Yaopeng Ding, Kellie Chong, Meng Cui, Zeyu Cao, Chenjue Tang, Zhen Tian, Yuping Hu, Yu Zhao and Shaoyi Jiang
Vaccines 2024, 12(10), 1148; https://doi.org/10.3390/vaccines12101148 - 8 Oct 2024
Cited by 3 | Viewed by 6318
Abstract
Introduction: The advent of lipid nanoparticles (LNPs) as a delivery platform for mRNA therapeutics has revolutionized the biomedical field, particularly in treating infectious diseases, cancer, genetic disorders, and metabolic diseases. Recent Advances in Therapeutic LNPs: LNPs, composed of ionizable lipids, phospholipids, cholesterol, and [...] Read more.
Introduction: The advent of lipid nanoparticles (LNPs) as a delivery platform for mRNA therapeutics has revolutionized the biomedical field, particularly in treating infectious diseases, cancer, genetic disorders, and metabolic diseases. Recent Advances in Therapeutic LNPs: LNPs, composed of ionizable lipids, phospholipids, cholesterol, and polyethylene glycol (PEG) lipids, facilitate efficient cellular uptake and cytosolic release of mRNA while mitigating degradation by nucleases. However, as synthetic entities, LNPs face challenges that alter their therapeutic efficacy and safety concerns. Toxicity/Reactogenicity/Immunogenicity: This review provides a comprehensive overview of the latest advancements in LNP research, focusing on preclinical safety assessments encompassing toxicity, reactogenicity, and immunogenicity. Summary and Outlook: Additionally, it outlines potential strategies for addressing these challenges and offers insights into future research directions for enhancing the application of LNPs in mRNA therapeutics. Full article
(This article belongs to the Special Issue Biotechnologies Applied in Vaccine Research)
Show Figures

Figure 1

Figure 1
<p>Lipid nanoparticles for organ-targeting mRNA delivery and their safety concerns for mRNA delivery. (Figure created with BioRender.com).</p>
Full article ">Figure 2
<p>(<b>a</b>) Preparation of SORT LNPs using multiple technical methods for tissue-specific mRNA delivery. Reproduced from ref. [<a href="#B29-vaccines-12-01148" class="html-bibr">29</a>]. Copyright Nature Publishing Group. (<b>b</b>) In vivo editing of lung stem cells for durable gene correction in mice. Reproduced from ref. [<a href="#B30-vaccines-12-01148" class="html-bibr">30</a>]. Copyright AAAS. (<b>c</b>) Spleen SORT LNP-generated in situ CAR T cells extend survival in a mouse model of lymphoreplete B cell lymphoma. Reproduced from ref. [<a href="#B31-vaccines-12-01148" class="html-bibr">31</a>]. Copyright Wiley-VCH. (<b>d</b>) Expanding RNAi to kidneys, lungs, and spleen via SORT siRNA LNPs. Reproduced from ref. [<a href="#B28-vaccines-12-01148" class="html-bibr">28</a>]. Copyright Wiley-VCH. (<b>e</b>) LNP-mediated drug delivery to the brain. Reproduced from ref. [<a href="#B40-vaccines-12-01148" class="html-bibr">40</a>]. Copyright Elsevier. (<b>f</b>) Bone-marrow-homing LNPs for genome editing in diseased and malignant hematopoietic stem cells. Reproduced from ref. [<a href="#B26-vaccines-12-01148" class="html-bibr">26</a>]. Copyright Nature Publishing Group.</p>
Full article ">Figure 3
<p>(<b>a</b>) Enhancing CRISPR/Cas gene editing through modulating cellular mechanical properties for cancer therapy. Illustration of triple loading of FAK siRNA, Cas9 mRNA, and sgRNA into 5A2-SC8 LNPs. (<b>b</b>) Dendrimer LNPs encapsulating FAK siRNA, Cas9 mRNA, and targeted sgRNAs could exhibit enhanced penetration into tumors with increased gene editing of PD-L1 for improved cancer therapy. (<b>c</b>) Representative fluorescence microscopy images at 0 h (top) and 48 h (bottom) and quantification of GFP fluorescence intensity. Reproduced from ref. [<a href="#B44-vaccines-12-01148" class="html-bibr">44</a>]. Copyright Nature Publishing Group.</p>
Full article ">Figure 4
<p>(<b>a</b>) Endosomal escape: A bottleneck for LNP-mediated therapeutics. Reproduced from ref. [<a href="#B48-vaccines-12-01148" class="html-bibr">48</a>]. Copyright NAS. (<b>b</b>) Rational design of cationic lipids for siRNA delivery. Reproduced from ref. [<a href="#B11-vaccines-12-01148" class="html-bibr">11</a>]. Copyright Nature Portfolio. (<b>c</b>) Nanomechanical action opens endo-lysosomal compartments. Reproduced from ref. [<a href="#B32-vaccines-12-01148" class="html-bibr">32</a>]. Copyright Nature Portfolio.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) A LNP platform incorporating trehalose glycolipid for exceptional mRNA vaccine safety. Reproduced from ref. [<a href="#B33-vaccines-12-01148" class="html-bibr">33</a>]. Copyright KeAi Communications Co., Ltd. (<b>c</b>) Polysarcosine-functionalized LNPs for therapeutic mRNA delivery. Reproduced from ref. [<a href="#B34-vaccines-12-01148" class="html-bibr">34</a>]. Copyright ACS.</p>
Full article ">Figure 6
<p>(<b>a</b>) Carrier-free mRNA vaccine induces robust immunity against SARS-CoV-2 in mice and non-human primates without systemic reactogenicity. Reproduced from ref. [<a href="#B65-vaccines-12-01148" class="html-bibr">65</a>]. Copyright Cell Press. (<b>b</b>) Biomimetic noncationic LNPs for mRNA delivery. The red dashed circle, the NC-TNP binds to mRNA via hydrogen bonds. Reproduced from ref. [<a href="#B35-vaccines-12-01148" class="html-bibr">35</a>]. Copyright NAS. (<b>c</b>) LNP and NC-TNP complexed with mRNA encoding EGFP incubated with DC2.4 cells to evaluate the EGFP protein expression by confocal microscopy images. Reproduced from ref. [<a href="#B35-vaccines-12-01148" class="html-bibr">35</a>]. Copyright NAS. (<b>d</b>) Schematic illustration of the experiment design. C57BL/6 mice were subcutaneously or intramuscularly injected with c-LNP, i-LNP, and NC-TNP. After different times, the skin samples from the injection site were collected for analysis. Reproduced from ref. [<a href="#B35-vaccines-12-01148" class="html-bibr">35</a>]. Copyright NAS.</p>
Full article ">Figure 7
<p>(<b>a</b>) MyD88-dependent and TRIF-dependent pathways signaling through TLRs activates intracellular signaling cascades that lead to nuclear translocation of AP-1 and NF-κB or IRF3. (<b>b</b>) Following IL-6 binding, the signal is transduced by a receptor to activate the JAKs, which then activate STATs. STATs are dephosphorylated in the nucleus, leading to the activation of downstream cytokines. Reproduced from ref. [<a href="#B68-vaccines-12-01148" class="html-bibr">68</a>]. Copyright Impact Journals.</p>
Full article ">Figure 8
<p>(<b>a</b>) Adjuvant lipidoid-substituted LNPs augment the immunogenicity of SARS-CoV-2 mRNA vaccines. Reproduced from ref. [<a href="#B36-vaccines-12-01148" class="html-bibr">36</a>]. Copyright Nature Publishing Group. (<b>b</b>) The role of nanoparticle format and route of administration on self-amplifying mRNA vaccine potency. Reproduced from ref. [<a href="#B75-vaccines-12-01148" class="html-bibr">75</a>]. Copyright Elsevier.</p>
Full article ">
15 pages, 4287 KiB  
Article
Targeted Delivery of STING Agonist via Albumin Nanoreactor Boosts Immunotherapeutic Efficacy against Aggressive Cancers
by Zhijun Miao, Xue Song, Anan Xu, Chang Yao, Peng Li, Yanan Li, Tao Yang and Gang Shen
Pharmaceutics 2024, 16(9), 1216; https://doi.org/10.3390/pharmaceutics16091216 - 17 Sep 2024
Viewed by 1346
Abstract
Background: Activating the cytosolic innate immune sensor, the cGAS-STING pathway, holds great promise for enhancing antitumor immunity, particularly in combination with immune checkpoint inhibitors (ICIs). However, the clinical application of STING agonists is often hindered by poor tumor accumulation, limited cellular uptake, and [...] Read more.
Background: Activating the cytosolic innate immune sensor, the cGAS-STING pathway, holds great promise for enhancing antitumor immunity, particularly in combination with immune checkpoint inhibitors (ICIs). However, the clinical application of STING agonists is often hindered by poor tumor accumulation, limited cellular uptake, and rapid clearance. To address these challenges, we developed a human serum albumin (HSA)-based nanoreactor system for the efficient delivery of the STING agonist SR-717, aiming to improve its antitumor efficacy. Methods: Using a biomineralization technique, we encapsulated SR-717 within HSA nanocages to form SH-NPs. These nanoparticles were characterized in terms of size, stability, and cellular uptake, and their ability to activate the STING pathway was assessed in both in vitro and in vivo models, including freshly isolated human renal tumor tissues. In vivo antitumor efficacy was evaluated in a murine renal tumor model, and immune responses were measured. Results: SH-NPs exhibited enhanced stability, efficient cellular uptake, and superior tumor accumulation compared to free SR-717. They robustly activated the STING pathway, as evidenced by increased phosphorylation of TBK1 and IRF3, along with elevated IFN-β production. Additionally, SH-NPs reshaped the immunosuppressive tumor microenvironment, promoting T-cell-mediated immunity and improving the therapeutic efficacy of checkpoint blockade in murine models. The validation in human renal tumor tissues further highlighted their potential for clinical translation. Importantly, SH-NPs were well tolerated with minimal systemic toxicity. Conclusions: This study underscores the potential of HSA-based nanoparticles for the targeted delivery of STING agonists, effectively enhancing antitumor immunity and improving cancer immunotherapy outcomes. SH-NPs offer a promising solution to the limitations of current STING agonists in clinical settings. Full article
(This article belongs to the Section Nanomedicine and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>Characterization of SH−NPs. TEM image (<b>a</b>) and hydrodynamic size distribution (<b>b</b>) of SH−NPs; (<b>c</b>) hydrodynamic size of SH−NPs stored at 4 °C for 7 days; (<b>d</b>) eta potential of SH−NPs; (<b>e</b>) accumulative release of SR−717 from free SR−717 and SH−NPs in pH 7.4 phosphate buffer, pH 5.0 acetate buffer, and pH 5.0 acetate buffer containing 10 μg mL<sup>−1</sup> of CB.</p>
Full article ">Figure 2
<p>Cellular behaviors of SH−NPs. Expression of downstream proteins of cGAS−STING signaling pathway including p−TBK1 and p−IRF3 (<b>a</b>) and statistics of protein relative gray value (<b>b</b>); (<b>c</b>) concentration of IFN−β secreted by cells treated with free SR−717 and SH−NPs; the expression of CD80 and CD86 on the surface of DC2.4 after administration (<b>d</b>) and statistics of maturation ratio (<b>e</b>); (<b>f</b>−<b>i</b>) expression and the fluorescence intensity of CD206 (<b>f</b>,<b>g</b>) and CD86 (<b>h</b>,<b>i</b>) on the surface of macrophages after incubation with free SR−717 and SH−NPs. Statistical differences: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 (<a href="#pharmaceutics-16-01216-f002" class="html-fig">Figure 2</a>b,c,e,g,i).</p>
Full article ">Figure 3
<p>Tumor targeting and in vivo antitumor efficacy of SH−NPs. In vivo fluorescence imaging of renal tumor bearing mice receiving SH−NPs at different time points (<b>a</b>) and the fluorescence intensity of tumor region (<b>b</b>); (<b>c</b>) tumor accumulation amount of SR−717 in the renal tumor bearing mice at 12 h post−injection of SH−NPs; (<b>d</b>) timeline schedule of treatment of renal tumor bearing mice; (<b>e</b>,<b>f</b>) tumor growth profiles (<b>e</b>) and survival curve (<b>f</b>) of the mice bearing renal tumor treated with PBS, free SR−717, aPD−L1, SH−NPs and SH−NPs/aPD−L1. Statistical differences: *** <span class="html-italic">p</span> &lt; 0.001 (<a href="#pharmaceutics-16-01216-f003" class="html-fig">Figure 3</a>b,c,e).</p>
Full article ">Figure 4
<p>In vivo STING activation and immune responses. (<b>a</b>,<b>b</b>) Expression (<b>a</b>) and the relative protein gray value (<b>b</b>) of p−TBK1 and p−IRF3 in tumor tissue from the mice bearing renal tumors that were treated with PBS, free SR−717, aPD−L1, SH−NPs, and SH−NPs/aPD−L1 at 72 h post−injection; (<b>c</b>−<b>f</b>) concentration of IFN−β (<b>c</b>), CXCL−10 (<b>d</b>), IL−6 (<b>e</b>), TNF−α (<b>f</b>) in tumor tissue from the same set of mice, assessed at the same time point; (<b>g</b>−<b>i</b>) quantification of tumor−infiltrating CTLs (<b>g</b>,<b>i</b>) and NK cells (<b>h</b>) in tumor tissue; (<b>j</b>,<b>k</b>) quantification of T−cells (<b>j</b>) and Tregs (<b>k</b>) in tumor tissue; (<b>i</b>) quantification of matured dendritic cells inside tumor−draining lymph nodes; (<b>m</b>,<b>n</b>) quantification of M2−like TAMs (<b>m</b>) and M1−like TAMs (<b>n</b>) in tumor tissue; (<b>o</b>) quantification of MDSCs in tumor tissue. Statistical differences: * <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 (<a href="#pharmaceutics-16-01216-f004" class="html-fig">Figure 4</a>b−h,j−o), ns-not significant (<a href="#pharmaceutics-16-01216-f004" class="html-fig">Figure 4</a>b–e).</p>
Full article ">Figure 5
<p>STING activation in freshly isolated human renal tumor tissues. (<b>a</b>−<b>c</b>) Concentration of IFN−β CD14b<sup>+</sup>CD68<sup>+</sup>CD86<sup>+</sup> TAMs (<b>b</b>) and CD14b<sup>+</sup>CD68<sup>+</sup>CD206<sup>+</sup> TAMs (<b>c</b>) in human renal tumor tissues treated with intratumoral injection of PBS, free SR−717, and SH−NPs. Statistical differences: *** <span class="html-italic">p</span> &lt; 0.001, ns-not significant (<a href="#pharmaceutics-16-01216-f005" class="html-fig">Figure 5</a>a−c).</p>
Full article ">Scheme 1
<p>The synthesis process of using human serum albumin (HSA) as a single−molecule nanoreactor to encapsulate the STING agonist SR−717, and the subsequent process of immune activation in tumor immunotherapy.</p>
Full article ">
22 pages, 2852 KiB  
Article
Upgrading Mitochondria-Targeting Peptide-Based Nanocomplexes for Zebrafish In Vivo Compatibility Assays
by Rúben Faria, Eric Vivès, Prisca Boisguérin, Simon Descamps, Ângela Sousa and Diana Costa
Pharmaceutics 2024, 16(7), 961; https://doi.org/10.3390/pharmaceutics16070961 - 20 Jul 2024
Viewed by 1224
Abstract
The lack of effective delivery systems has slowed the development of mitochondrial gene therapy. Delivery systems based on cell-penetrating peptides (CPPs) like the WRAP (tryptophan and arginine-rich peptide) family conjugated with a mitochondrial targeting sequence (MTS) have emerged as adequate carriers to mediate [...] Read more.
The lack of effective delivery systems has slowed the development of mitochondrial gene therapy. Delivery systems based on cell-penetrating peptides (CPPs) like the WRAP (tryptophan and arginine-rich peptide) family conjugated with a mitochondrial targeting sequence (MTS) have emerged as adequate carriers to mediate gene expression into the mitochondria. In this work, we performed the PEGylation of WRAP/pDNA nanocomplexes and compared them with previously analyzed nanocomplexes such as (KH)9/pDNA and CpMTP/pDNA. All nanocomplexes exhibited nearly homogeneous sizes between 100 and 350 nm in different environments. The developed complexes were biocompatible and hemocompatible to both human astrocytes and lung smooth muscle cells, ensuring in vivo safety. The nanocomplexes displayed mitochondria targeting ability, as through transfection they preferentially accumulate into the mitochondria of astrocytes and muscle cells to the detriment of cytosol and lysosomes. Moreover, the transfection of these cells with MTS–CPP/pDNA complexes produced significant levels of mitochondrial protein ND1, highlighting their efficient role as gene delivery carriers toward mitochondria. The positive obtained data pave the way for in vivo research. Using confocal microscopy, the cellular internalization capacity of these nanocomplexes in the zebrafish embryo model was assessed. The peptide-based nanocomplexes were easily internalized into zebrafish embryos, do not cause harmful or toxic effects, and do not affect zebrafish’s normal development and growth. These promising results indicate that MTS–CPP complexes are stable nanosystems capable of internalizing in vivo models and do not present associated toxicity. This work, even at an early stage, offers good prospects for continued in vivo zebrafish research to evaluate the performance of nanocomplexes for mitochondrial gene therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Cellular viability of human astrocyte cells ((<b>A</b>) 24 h, (<b>B</b>) 48 h) and lung smooth muscle cells ((<b>C</b>) 24 h, (<b>D</b>) 48 h) after incubation with naked pND1 and the 20% PEG–MTS–WRAP1/pND1 (PEG–MTS–W1/pND1), 20% PEG–MTS–WRAP5/pND1 (PEG–MTS–W5/pND1) and MTS–(KH)<sub>9</sub>/pND1 nanocomplexes formulated at N/P ratio of 5 (pND1 = 1 µg). Non-transfected cells were used as a positive control (Control (+)) and cells treated with ethanol were used as a negative control (Control (−)). Data were analyzed by one-way ANOVA with Bonferroni’s multiple comparison test (ns—non-significant (<span class="html-italic">p</span> &gt; 0.05); * <span class="html-italic">p</span> ˂ 0,05; **** <span class="html-italic">p</span> ˂ 0.0001).</p>
Full article ">Figure 2
<p>In vitro hemolysis assay using rat red blood cells (RBCs), which were incubated with 20% PEG–MTS–WRAP1/pND1, 20% PEG–MTS–WRAP5/pND1, and MTS–(KH)<sub>9</sub>/pND1 (1 µg of pND1, N/P ratio = 5). The negative control was incubated with PBS pH 7.4, while in the positive control, RBCs were incubated with Triton X-100 (1%) to provoke hemolysis. The hemolysis percentages were calculated according to Formula (1). Data are presented as mean (%) ± SD (n = 3). Data were analyzed by one-way ANOVA with Bonferroni’s multiple comparison test (ns—non-significant (<span class="html-italic">p</span> &gt; 0.05); **** <span class="html-italic">p</span> ˂ 0.0001).</p>
Full article ">Figure 3
<p>Quantification of FITC fluorescence intensity ((a.u)/µg Protein) in the lysosomes, cytosol, and mitochondria of human astrocyte cells (<b>A</b>) and lung smooth muscle cells (<b>B</b>), after 24 h of transfection with 20% PEG–MTS–WRAP1/pND1 (PEG–MTS–W1/pND1), 20% PEG–MTS–WRAP5/pND1 (PEG–MTS–W5/pND1) and MTS–(KH)<sub>9</sub>/pND1 systems. All complexes were formulated with an N/P ratio = 5 (pND1 = 1 µg). Untreated cells and naked pND1 stained with FITC were used as controls. Data were analyzed by two-way ANOVA with Bonferroni’s multiple comparison test (ns—non-significant (<span class="html-italic">p</span> &gt; 0.05); * <span class="html-italic">p</span> ˂ 0.05; ** <span class="html-italic">p</span> ˂ 0.01; *** <span class="html-italic">p</span> ˂ 0.001; **** <span class="html-italic">p</span> ˂ 0.0001).</p>
Full article ">Figure 4
<p>Quantification of ND1 protein levels (ng/mL) in human astrocyte cells (<b>A</b>) and lung smooth muscle cells (<b>B</b>), after 48 h of transfection with 20% PEG–MTS–WRAP1/pND1 (PEG–MTS–W1/pND1), 20% PEG–MTS–WRAP5/pND1 (PEG–MTS–W5/pND1), and MTS–(KH)<sub>9</sub>/pND1 systems (pND1 = 1 µg for all). All complexes were formulated with an N/P ratio = 5. Data were analyzed by one-way ANOVA with Bonferroni’s multiple comparison tests (** <span class="html-italic">p</span> = 0.0041 (<b>A</b>) and 0.0015 (<b>B</b>),**** <span class="html-italic">p</span> ˂ 0.0001).</p>
Full article ">Figure 5
<p>Evaluation of the ability of CPP-based nanocomplexes for ZF embryo transfection. Representative confocal images of ZF embryos expressing the GFP protein (green signal) transfected with different CPP-based complexes encapsulating Alexa594-labelled pND1. (<b>A</b>) 20% PEG–MTS–WRAP1/pND1 with 2 µg, (<b>B</b>) 20% PEG–MTS–WRAP1/pND1 with 1 µg, (<b>C</b>) 20% PEG–MTS–WRAP1/pND1 with 0.5 µg, (<b>D</b>) 20% PEG–WRAP5/pND1 with 2 µg, and (<b>E</b>) MTS–(KH)<sub>9</sub>/pND1 with 2 µg imaged after 24 h incubation. Peptide nanocomplexes were formulated at an N/P ratio of 5 using the indicated final plasmid concentrations. Untransfected ZF embryos were used as control (<b>F</b>). Bars represent 100 µm.</p>
Full article ">Figure 6
<p>Assessment of the toxicity of the 20% PEG–MTS–WRAP1/pND1 (<b>A</b>), 20% PEG–MTS–WRAP5/pND1 (<b>B</b>), and MTS–(KH)<sub>9</sub>/pND1 (<b>C</b>) nanocomplexes (N/P ratio = 5) in ZF embryos. Toxicity was assessed through the average size of the embryos (µm) and their survival (/12) after 48 h of incubation. Non-transfected embryos were used as a control group. All nanocomplexes were tested at three different amounts (1, 2, and 5 µg).</p>
Full article ">
14 pages, 3143 KiB  
Article
Trigonometric Bundling Disulfide Unit Starship Synergizes More Effectively to Promote Cellular Uptake
by Lei Wang, Dezhi Wang, Wenzhuo Lei, Tiantian Sun, Bei Gu, Han Dong, Yosuke Taniguchi, Yichang Liu and Yong Ling
Int. J. Mol. Sci. 2024, 25(14), 7518; https://doi.org/10.3390/ijms25147518 - 9 Jul 2024
Cited by 1 | Viewed by 829
Abstract
A small molecule disulfide unit technology platform based on dynamic thiol exchange chemistry at the cell membrane has the potential for drug delivery. However, the alteration of the CSSC dihedral angle of the disulfide unit caused by diverse substituents directly affects the effectiveness [...] Read more.
A small molecule disulfide unit technology platform based on dynamic thiol exchange chemistry at the cell membrane has the potential for drug delivery. However, the alteration of the CSSC dihedral angle of the disulfide unit caused by diverse substituents directly affects the effectiveness of this technology platform as well as its own chemical stability. The highly stable open-loop relaxed type disulfide unit plays a limited role in drug delivery due to its low dihedral angle. Here, we have built a novel disulfide unit starship based on the 3,4,5-trihydroxyphenyl skeleton through trigonometric bundling. The intracellular delivery results showed that the trigonometric bundling of the disulfide unit starship effectively promoted cellular uptake without any toxicity, which is far more than 100 times more active than that of equipment with a single disulfide unit in particular. Then, the significant reduction in cell uptake capacity (73–93%) using thiol erasers proves that the trigonometric bundling of the disulfide starship is an endocytosis-independent internalization mechanism via a dynamic covalent disulfide exchange mediated by thiols on the cell surface. Furthermore, analysis of the molecular dynamics simulations demonstrated that trigonometric bundling of the disulfide starship can significantly change the membrane curvature while pushing lipid molecules in multiple directions, resulting in a significant distortion in the membrane structure and excellent membrane permeation performance. In conclusion, the starship system we built fully compensates for the inefficiency deficiencies induced by poor dihedral angles. Full article
(This article belongs to the Section Materials Science)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Brief description of disulfide unit performance between structure and corresponding dihedral angle. (<b>B</b>) The trigonometric bundling disulfide unit starship strategy used.</p>
Full article ">Figure 2
<p>The cytotoxicity of disulfide unit starship using CCK-8 assay for intermediate <b>7a’</b> and <b>7b’</b>.</p>
Full article ">Figure 3
<p>CLSM images of A549 cells after incubation with 1 μM SS1/SS3-FITC probe together with DAPI to stain the nuclei for (<b>A</b>,<b>B</b>); Quantitative analysis of fluorescence intensity and colocalization analysis are shown in (<b>C</b>,<b>D</b>), respectively. Scatterplot analysis of the merge image in the last view with the white arrow symbol is shown in (<b>E</b>). Scar bar: 10 µm.</p>
Full article ">Figure 4
<p>CLSM images of A549 cells after 1 h incubation with different concentrations of <b>SS1/SS3-FITC</b> probe together with DAPI to stain the nuclei for (<b>A</b>,<b>B</b>). Quantitative analysis of fluorescence intensity is shown in (<b>C</b>). Scar bar: 10 µm.</p>
Full article ">Figure 5
<p>CLSM images of A549 cells that were preincubated with inhibitors (NEM, SIA, and DTNB, 1.2 mM, 0.5 h) and then incubated with an <b>SS3-FITC</b> probe (1 μM, 1 h) together with DAPI to stain the nuclei for (<b>A</b>,<b>B</b>). Quantitative analysis of fluorescence intensity is shown in (<b>C</b>). Scar bar: 10 µm.</p>
Full article ">Figure 6
<p>(<b>A</b>) The membrane distortion induced by <b>6a</b> and <b>6b</b>. The local membrane thickness maps were shown in the top panel (color bar unit: nm). (<b>B</b>) The molecular details sounded by <b>6a</b> and <b>6b</b> were shown in the bottom panel (the head groups and tail groups of lipid molecules were colored in sea blue and white; the carbon, oxygen, nitrogen, and sulfur atoms in <b>6a</b> and <b>6b</b> were colored in green, red, blue, and yellow, respectively). (<b>C</b>) The lipid order parameters of (<span class="html-italic">S</span><sub>cd</sub>) of the palmitic chain (up) and oleic chain (down) in lipid molecules. The carbon index started from the carboxyl group of the fatty acid chain. The legend “control” indicates that nothing inserts in membrane.</p>
Full article ">Scheme 1
<p>Synthesis of <b>SS1/3-FITC</b> probe (<b>8a</b> and <b>8b</b>).</p>
Full article ">
21 pages, 3027 KiB  
Article
The JMU-SalVac-System: A Novel, Versatile Approach to Oral Live Vaccine Development
by Andreas Iwanowitsch, Joachim Diessner, Birgit Bergmann and Thomas Rudel
Vaccines 2024, 12(6), 687; https://doi.org/10.3390/vaccines12060687 - 20 Jun 2024
Viewed by 1463
Abstract
Salmonella enterica Serovar Typhi Ty21a (Ty21a) is the only licensed oral vaccine against typhoid fever. Due to its excellent safety profile, it has been used as a promising vector strain for the expression of heterologous antigens for mucosal immunization. As the efficacy of [...] Read more.
Salmonella enterica Serovar Typhi Ty21a (Ty21a) is the only licensed oral vaccine against typhoid fever. Due to its excellent safety profile, it has been used as a promising vector strain for the expression of heterologous antigens for mucosal immunization. As the efficacy of any bacterial live vector vaccine correlates with its ability to express and present sufficient antigen, the genes for antigen expression are traditionally located on plasmids with antibiotic resistance genes for stabilization. However, for use in humans, antibiotic selection of plasmids is not applicable, leading to segregational loss of the antigen-producing plasmid. Therefore, we developed an oral Ty21a-based vaccine platform technology, the JMU-SalVac-system (Julius-Maximilians-Universität Würzburg) in which the antigen delivery plasmids (pSalVac-plasmid-series) are stabilized by a ΔtyrS/tyrS+-based balanced-lethal system (BLS). The system is made up of the chromosomal knockout of the essential tyrosyl-tRNA-synthetase gene (tyrS) and the in trans complementation of tyrS on the pSalVac-plasmid. Further novel functional features of the pSalVac-plasmids are the presence of two different expression cassettes for the expression of protein antigens. In this study, we present the construction of vaccine strains with BLS plasmids for antigen expression. The expression of cytosolic and secreted mRFP and cholera toxin subunit B (CTB) proteins as model antigens is used to demonstrate the versatility of the approach. As proof of concept, we show the induction of previously described in vivo inducible promoters cloned into pSalVac-plasmids during infection of primary macrophages and demonstrate the expression of model vaccine antigens in these relevant human target cells. Therefore, antigen delivery strains developed with the JMU-SalVac technology are promising, safe and stable vaccine strains to be used against mucosal infections in humans. Full article
(This article belongs to the Special Issue Advances in Oral Vaccine Development)
Show Figures

Figure 1

Figure 1
<p>Genetic map of pSalVac base vector (<b>top</b>) and the inserts (<b>bottom</b>). The base vector contains the <span class="html-italic">hlyA</span> secretion cluster including a cloning site (A-site) for the insertion of secreted antigens such as CTB-FLAG. Another independent cloning site (B-site) is located downstream of the <span class="html-italic">hlyA</span> cluster. A kanamycin resistance cassette flanked by FRT-sites can be excised via Flippase expression and TyrS is expressed from a lacI-derivative promoter for balance-lethal stabilization in the Δ<span class="html-italic">tyrS</span> host. Figure created with BioRender.com.</p>
Full article ">Figure 2
<p>Schematic representation of the genomic <span class="html-italic">tyrS</span> locus of Ty21a during generation of BLS-stabilized JMU-SalVac strains. First, the <span class="html-italic">tyrS</span> locus of the licensed Ty21a strain was replaced with a <span class="html-italic">tyrS</span>-CmR cassette flanked by FRT sites. After transformation with one of the TyrS complementing pSalVac-plasmids, the <span class="html-italic">tyrS</span> cassette is deleted by expression of yeast flippase encoded on pCP20, leaving behind a short nucleotide sequence with one FRT site. Figure created with BioRender.com.</p>
Full article ">Figure 3
<p>Maintenance of BLS plasmids and effect on growth behavior and antigen expression. (<b>A</b>) In vitro plasmid stability of non-stabilized (Ty21a ActxB_B0 KanR, Ty21a pMKhly-ctxB) and BLS-stabilized (BLS ActxB_B0, BLS-ActxB_B0 KanR) strains. For details, see <a href="#sec2-vaccines-12-00687" class="html-sec">Section 2</a>. (<b>B</b>) endpoint plasmid retention of three independent replicates of the plasmid stability assay. (<b>C</b>) Growth curves of WT Ty21a and BLS-stabilized vaccine strains (<span class="html-italic">N</span> = 3). (<b>D</b>) Area under the curve analysis (AUC) of growth curves of different <span class="html-italic">S.</span> Typhi/plasmid combinations. Data were obtained from three independent replicates (error bars represent standard deviation, <span class="html-italic">N</span> = 3) (<b>E</b>) Western blot analysis of BLS-ActxB_B0, whole cell lysate (<b>left</b>) and supernatants (<b>right</b>). Approx. 7.5 × 10<sup>5</sup> bacteria per well (for whole cell lysate, WCL) and 5 µg total protein per well (supernatant) were analyzed by SDS-PAGE (10% gel) before transfer to PVDF. Detection of CTB-FLAG was performed with anti-FLAG primary antibody and TyrS-6xHis with anti-6xHis primary antibody. (<b>F</b>) Western blot analysis before (Day 0) and after (Day 6) plasmid stability assay.</p>
Full article ">Figure 4
<p>In vivo inducible promoters in BLS-stabilized plasmids are active in Ty21a-infected RAW246.7 macrophages. Immunofluorescence images (<b>A</b>) and flow cytometry analysis (<b>B</b>) of RAW246.7 cells infected with Ty21a strains with in vivo inducible promoters driving mRFP expression. Empty vector (BLS-A0_B0) was used as negative control (e.v.) for 2h. Extracellular bacteria were killed with 15 µg/mL gentamycin for 30 min and further 2.5 h (<b>B</b>) or 3.5 h (<b>A</b>) with 1 µg/mL gentamycin. Cells were stained with fixable viability stain (B only), fixed, permeabilized and stained with anti-Typhi primary antibody and Alexa488 secondary before analysis. (<b>B</b>) All live cells were gated for infection ((<b>first</b>) panel), mRFP expression ((<b>middle</b>) panel), or sub-gated for infected cells with mRFP signal ((<b>right</b>) panel). Data were obtained from five independent replicates (error bars represent standard deviation, <span class="html-italic">N</span> = 5).</p>
Full article ">Figure 5
<p>In vivo inducible promoters in BLS-stabilized plasmids are active in Ty21a-infected hMDMs. Immunofluorescence images (<b>A</b>) and flow cytometry analysis (<b>B</b>) of human-monocyte-derived macrophages (hMDMs) infected at MOI10 with BLS strains with in vivo inducible promoters driving mRFP expression. BLS-A0_B0 was used as negative control (e.v.). hMDMs were infected with Ty21a strains expressing mRFP under the control of P<sub>asr</sub> (A0_BP<sub>asr</sub>mRFP) or P<sub>pagC</sub> (A0_BP<sub>pacC</sub>mRFP) for 2 h. Extracellular bacteria were killed with 15 µg/mL gentamycin for 30 min and further 2.5 h with 1 µg/mL gentamycin. Cells were stained with fixable viability stain ((<b>B</b>) only), fixed, permeabilized and stained with anti-Typhi primary antibody and Alexa488 secondary before analysis. (<b>B</b>) All live cells were gated for infection ((<b>first</b>) panel), mRFP expression ((<b>middle</b>) panel), or sub-gated for infected cells with mRFP signal ((<b>right</b>) panel). Data were obtained from three independent replicates (error bars represent standard deviation, <span class="html-italic">N</span> = 3).</p>
Full article ">Figure 6
<p>Simultaneous expression of two antigens and intracellular replication of BLS strains in hMDMs. BLS strain constitutively expressing CTB-FLAG and the dual reporter system mRFP and arabinose inducible GFP (BLS-ActxB_BDR) visualized by (<b>A</b>) western blot analysis of whole cell lysate and TCA precipitated supernatant, (<b>B</b>) growth curve analysis and (<b>C</b>) hMDM infection. (<b>A</b>) Approx. 7.5 × 10<sup>5</sup> bacteria per well (for whole cell lysate, WCL) and 5µg precipitated supernatant were analyzed on a 12% SDS-PAGE before transfer to PVDF. Detection of CTB-FLAG was performed with anti-FLAG primary antibody. GFP+ denotes induction with 0.2% arabinose. (<b>B</b>) Growth curves of WT Ty21a (circles) and BLS-stabilized vaccine strains BLS-A0_B0 (boxes) and BLS-ActxB_BDR (triangles) (<span class="html-italic">N</span> = 3) (<b>C</b>) hMDMs were infected for 2 h with arabinose induced <span class="html-italic">S.</span> Typhi BLS-ActxB_BDR. Extracellular bacteria were killed with 15 µg/mL gentamycin for 30 min and further 2.5 h with 1 µg/mL gentamycin and mounted with ProLong mounting medium containing NucBlue.</p>
Full article ">
12 pages, 1720 KiB  
Communication
Hemoglobin Binding to the Red Blood Cell (RBC) Membrane Is Associated with Decreased Cell Deformability
by Gregory Barshtein, Leonid Livshits, Alexander Gural, Dan Arbell, Refael Barkan, Ivana Pajic-Lijakovic and Saul Yedgar
Int. J. Mol. Sci. 2024, 25(11), 5814; https://doi.org/10.3390/ijms25115814 - 27 May 2024
Cited by 4 | Viewed by 2503
Abstract
The deformability of red blood cells (RBCs), expressing their ability to change their shape as a function of flow-induced shear stress, allows them to optimize oxygen delivery to the tissues and minimize their resistance to flow, especially in microcirculation. During physiological aging and [...] Read more.
The deformability of red blood cells (RBCs), expressing their ability to change their shape as a function of flow-induced shear stress, allows them to optimize oxygen delivery to the tissues and minimize their resistance to flow, especially in microcirculation. During physiological aging and blood storage, or under external stimulations, RBCs undergo metabolic and structural alterations, one of which is hemoglobin (Hb) redistribution between the cytosol and the membrane. Consequently, part of the Hb may attach to the cell membrane, and although this process is reversible, the increase in membrane-bound Hb (MBHb) can affect the cell’s mechanical properties and deformability in particular. In the present study, we examined the correlation between the MBHb levels, determined by mass spectroscopy, and the cell deformability, determined by image analysis. Six hemoglobin subunits were found attached to the RBC membranes. The cell deformability was negatively correlated with the level of four subunits, with a highly significant inter-correlation between them. These data suggest that the decrease in RBC deformability results from Hb redistribution between the cytosol and the cell membrane and the respective Hb interaction with the cell membrane. Full article
Show Figures

Figure 1

Figure 1
<p>Variability in the content of membrane-bound Hb-subunits (expressed by (Ln (LFQ)). Statistical analysis was carried out for 15 samples of healthy adult RBCs.</p>
Full article ">Figure 2
<p>Correlation between the level of the RBC membrane-bound β, α, and δ subunits (Ln (LFQ)) and the cell deformability (AER); for statistical analysis, see <a href="#ijms-25-05814-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 3
<p>Scheme of cell-flow analyzer CFA.</p>
Full article ">Figure 4
<p>A. Image of RBC field in the CFA flow-chamber under flow-induced shear stress of 3.0 Pa. B. Highly deformable cell (ER = 2.3). C. Non-deformable cell (ER = 1).</p>
Full article ">
14 pages, 8878 KiB  
Article
Investigating Internalization of Reporter-Protein-Functionalized Polyhedrin Particles by Brain Immune Cells
by Krishma A. K. Parwana, Priyapreet Kaur Gill, Runyararo Njanike, Humphrey H. P. Yiu, Chris F. Adams, Divya Maitreyi Chari and Stuart Iain Jenkins
Materials 2024, 17(10), 2330; https://doi.org/10.3390/ma17102330 - 14 May 2024
Cited by 1 | Viewed by 1342
Abstract
Achieving sustained drug delivery to the central nervous system (CNS) is a major challenge for neurological injury and disease, and various delivery vehicles are being developed to achieve this. Self-assembling polyhedrin crystals (POlyhedrin Delivery System; PODS) are being exploited for the delivery of [...] Read more.
Achieving sustained drug delivery to the central nervous system (CNS) is a major challenge for neurological injury and disease, and various delivery vehicles are being developed to achieve this. Self-assembling polyhedrin crystals (POlyhedrin Delivery System; PODS) are being exploited for the delivery of therapeutic protein cargo, with demonstrated efficacy in vivo. However, to establish the utility of PODS for neural applications, their handling by neural immune cells (microglia) must be documented, as these cells process and degrade many biomaterials, often preventing therapeutic efficacy. Here, primary mouse cortical microglia were cultured with a GFP-functionalized PODS for 24 h. Cell counts, cell morphology and Iba1 expression were all unaltered in treated cultures, indicating a lack of acute toxicity or microglial activation. Microglia exhibited internalisation of the PODS, with both cytosolic and perinuclear localisation. No evidence of adverse effects on cellular morphology was observed. Overall, 20–40% of microglia exhibited uptake of the PODS, but extracellular/non-internalised PODS were routinely present after 24 h, suggesting that extracellular drug delivery may persist for at least 24 h. Full article
(This article belongs to the Special Issue Synthesis, Assembly and Applications of Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>PODS are co-crystals of polyhedrin and a cargo protein—in this instance GFP—and were readily imaged by microscopy. (<b>a</b>) Schematic indicating co-crystalline lattice structure of GFP-PODS. (<b>b</b>) Merged phase contrast and fluorescence micrograph illustrating consistent morphologies of GFP-PODS and limited aggregation (inset shows phase contrast micrograph alone). (<b>c</b>) Merged phase contrast and fluorescence micrograph illustrating size range of GFP-PODS and their tendency to ‘stand’ flush to a flat surface, presenting a squared upper surface. (<b>d</b>) Transmission electron micrograph of intracellular PODS. PODS were found in the cytosol, amongst the organelles. Arrow shows a tangential section of a cell nucleus, arrowhead points to a mitochondrion and * indicates vacuole-like structures. Scale bar: 1 μm. (<b>e</b>) Scanning electron micrograph of GFP-PODS, showing cuboid morphology; scale bar: 2 μm. (<b>f</b>) Graph of PODS side length; error bars indicate standard deviation.</p>
Full article ">Figure 2
<p>Microglia incubated with GFP-PODS did not exhibit acute toxicity. Similar cell numbers and morphologies were observed in control and GFP-PODS-treated microglial cultures. Low magnification phase contrast micrographs of (<b>a</b>) control and (<b>b</b>) PODS-treated microglial cultures. Higher magnification merged fluorescence micrographs of (<b>c</b>) control and (<b>d</b>) PODS-treated microglial cultures (insets show Iba1 staining alone). (<b>e</b>) Graph indicating similar cell counts in control and GFP-PODS-treated cultures. Data normalised to average of all control counts; no significant difference (<span class="html-italic">p</span> = 0.838), two-tailed unpaired <span class="html-italic">t</span>-test, <span class="html-italic">n</span> = 4.</p>
Full article ">Figure 3
<p>Detailed morphological analyses found no differences between control and PODS-treated cultures, and Iba1 expression was similar. (<b>a</b>) Merged fluorescence micrograph of control culture showing lack of green fluorescence and ramified Iba1+ microglia. (<b>b</b>) Merged fluorescence micrograph of Iba1+ microglia in PODS-treated culture. Note intracellular GFP-PODS distributed throughout the cytosol. Cell morphologies are similar to those in control cultures, with similar quantities and dimensions of processes. Occasional instances of extensive PODS uptake were observed, as shown in (<b>c</b>–<b>f</b>): counterpart phase contrast, merged fluorescence, red channel fluorescence and green channel fluorescence micrographs, respectively. Note amoeboid morphology. (<b>g</b>) Graph showing similar intensity of Iba1 expression in control and PODS-treated microglial cultures. Graphs comparing various cellular morphometrics: (<b>h</b>) area, (<b>i</b>) perimeter, (<b>j</b>) Feret’s (max) diameter, (<b>k</b>) Feret’s min diameter, (<b>l</b>) Feret’s aspect ratio and (<b>m</b>) solidity. All graphs show no significant differences from two-tailed unpaired <span class="html-italic">t</span>-tests, <span class="html-italic">n</span> = 4; <span class="html-italic">p</span>-values: (<b>g</b>) 0.285, (<b>h</b>) 0.801, (<b>i</b>) 0.292, (<b>j</b>) 0.648, (<b>k</b>) 0.718, (<b>l</b>) 0.226 and (<b>m</b>) 0.429. Scale bars: 10 µm.</p>
Full article ">Figure 4
<p>PODS showed both cytosolic and perinuclear localisation and occasionally distorted the shape of the nucleus. (<b>a</b>–<b>d</b>) Micrographs showing Iba1+ microglia with intracellular GFP-PODS, both perinuclear and cytosolic, including particle distant to nucleus (red arrow): (<b>a</b>) merged fluorescence, (<b>b</b>) red channel only, (<b>c</b>) green channel only and (<b>d</b>) blue channel only, with phase contrast inset. White arrow indicates flattened edge of nucleus, which coincides with flat face of PODS particle. (<b>e</b>–<b>i</b>) Micrographs of Iba1+ microglia, each with a perinuclear PODS particle: green, merged, phase contrast, phase contrast-blue merge and blue. White arrow indicates angled indent in nucleus edge, coinciding with corner of particle. Orange arrow indicates similar dimple in region without PODS particle. Such observations left doubt as to whether PODS were genuinely displacing the nuclear envelope. (<b>j</b>–<b>m</b>) Fluorescence micrographs of Iba1+ microglial cell: merged, green, blue and red. White arrow indicates flattened edge of nucleus, co-localised with flat edge of PODS particle. <a href="#app1-materials-17-02330" class="html-app">Supplementary Materials: video file showing z stack microscopy of this cell and particle</a>. (<b>n</b>) Graph indicating percentage of Iba1+ microglial cells that exhibited PODS uptake within four separate cultures. Error bars indicate SEM.</p>
Full article ">Figure 5
<p>Possible fates for intra- and extracellular PODS particles. This schematic offers speculation on possible PODS fate, depending on whether PODS are internalised by cells, and then whether there is degradation of the polyhedrin and whether intracellular release of cargo would result in secretion of the cargo into extracellular space, or whether the cargo may be sequestered or also subject to intracellular degradation. Prolonged resistance to immune cell uptake (extracellular PODS) would be beneficial for extracellular drug release, as would cellular uptake followed by drug secretion (green background). However, microglial sequestration of PODS without cargo release, or with degradation of cargo, would prevent drug delivery (blue background). Finally, the worst-case scenario would be cellular clearance followed by degradation, resulting in toxic breakdown products leading to cytotoxicity, possibly to the extent of cell death (orange background).</p>
Full article ">
17 pages, 3702 KiB  
Review
Failure of Autophagy in Pompe Disease
by Hung Do, Naresh K. Meena and Nina Raben
Biomolecules 2024, 14(5), 573; https://doi.org/10.3390/biom14050573 - 13 May 2024
Viewed by 2247
Abstract
Autophagy is an evolutionarily conserved lysosome-dependent degradation of cytoplasmic constituents. The system operates as a critical cellular pro-survival mechanism in response to nutrient deprivation and a variety of stress conditions. On top of that, autophagy is involved in maintaining cellular homeostasis through selective [...] Read more.
Autophagy is an evolutionarily conserved lysosome-dependent degradation of cytoplasmic constituents. The system operates as a critical cellular pro-survival mechanism in response to nutrient deprivation and a variety of stress conditions. On top of that, autophagy is involved in maintaining cellular homeostasis through selective elimination of worn-out or damaged proteins and organelles. The autophagic pathway is largely responsible for the delivery of cytosolic glycogen to the lysosome where it is degraded to glucose via acid α-glucosidase. Although the physiological role of lysosomal glycogenolysis is not fully understood, its significance is highlighted by the manifestations of Pompe disease, which is caused by a deficiency of this lysosomal enzyme. Pompe disease is a severe lysosomal glycogen storage disorder that affects skeletal and cardiac muscles most. In this review, we discuss the basics of autophagy and describe its involvement in the pathogenesis of muscle damage in Pompe disease. Finally, we outline how autophagic pathology in the diseased muscles can be used as a tool to fast track the efficacy of therapeutic interventions. Full article
Show Figures

Figure 1

Figure 1
<p>mTORC1 and AMPK signaling. A diagram shows the position of the proteins analyzed in Pompe disease. Under the nutrient-replete condition, mTORC1 is recruited to the lysosome where it is activated by GTP-bound Rheb. Activated mTORC1 suppresses autophagy through phosphorylation-dependent inhibition of the ULK1 complex and the transcription factors TFEB/TFE3; mTORC1 phosphorylates 4E-BP1 (translation repressor protein) and S6K, thereby stimulating the initiation of protein synthesis. Under nutrient deprivation, AMPK promotes autophagy initiation through positive phosphorylation of ULK1. AMPK also promotes autophagy indirectly by inhibiting mTORC1 activity through phosphorylation of TSC2; mTORC1 is displaced from the lysosome (inactivation), and TFEB/TFE3 translocate from cytoplasm to nucleus to promote the transcription of autophagy/lysosome-related genes.</p>
Full article ">Figure 2
<p>Confocal microscopy of live unstained muscle fibers freshly isolated from a GFP-LC3:KO mouse (ex vivo analysis). The image shows typical autophagic buildup (green area) which disrupts muscle architecture. Bar: 20 μm.</p>
Full article ">Figure 3
<p>Intravital microscopy of limb muscles of untreated and treated GFP-LC3:KO mice. Gastrocnemius muscles of GFP-LC3:KO mice were imaged at different ages to monitor the progression of autophagic pathology. Inset shows ring-shaped autophagosomes. The right panel shows the elimination of autophagic buildup in a treated GFP-LC3:KO mouse. The IVM imaging was performed two months after a single intravenous administration of an AAV9 vector expressing human <span class="html-italic">GAA</span> transgene into a 5-month-old animal. Bars: 30 μm.</p>
Full article ">
25 pages, 3699 KiB  
Article
A Conditionally Activated Cytosol-Penetrating Antibody for TME-Dependent Intracellular Cargo Delivery
by Carolin Sophie Dombrowsky, Dominic Happel, Jan Habermann, Sarah Hofmann, Sasi Otmi, Benny Cohen and Harald Kolmar
Antibodies 2024, 13(2), 37; https://doi.org/10.3390/antib13020037 - 2 May 2024
Cited by 1 | Viewed by 3458
Abstract
Currently, therapeutic and diagnostic applications of antibodies are primarily limited to cell surface-exposed and extracellular proteins. However, research has been conducted on cell-penetrating peptides (CPP), as well as cytosol-penetrating antibodies, to overcome these limitations. In this context, a heparin sulfate proteoglycan (HSPG)-binding antibody [...] Read more.
Currently, therapeutic and diagnostic applications of antibodies are primarily limited to cell surface-exposed and extracellular proteins. However, research has been conducted on cell-penetrating peptides (CPP), as well as cytosol-penetrating antibodies, to overcome these limitations. In this context, a heparin sulfate proteoglycan (HSPG)-binding antibody was serendipitously discovered, which eventually localizes to the cytosol of target cells. Functional characterization revealed that the tested antibody has beneficial cytosol-penetrating capabilities and can deliver cargo proteins (up to 70 kDa) to the cytosol. To achieve tumor-specific cell targeting and cargo delivery through conditional activation of the cell-penetrating antibody in the tumor microenvironment, a single-chain Fc fragment (scFv) and a VL domain were isolated as masking units. Several in vitro assays demonstrated that fusing the masking protein with a cleavable linker to the cell penetration antibody results in the inactivation of antibody cell binding and internalization. Removal of the mask via MMP-9 protease cleavage, a protease that is frequently overexpressed in the tumor microenvironment (TME), led to complete regeneration of binding and cytosol-penetrating capabilities. Masked and conditionally activated cytosol-penetrating antibodies have the potential to serve as a modular platform for delivering protein cargoes addressing intracellular targets in tumor cells. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic overview of the (<b>A</b>) masking capabilities in non-cancerous tissue, in comparison to (<b>B</b>) activation and restoration of binding capabilities after TME-associated MMP-9 cleavage, and (<b>C</b>) the pathway of the activated cytosol-penetrating antibody after HSPG binding on the cell surface. The antibody is released from HSPG through receptor-mediated endocytosis, followed by internalization and subsequent release. Decrease in pH promotes endosomal escape of the cytosol-penetrating antibody with attached cargo, which is, in this instance, C-terminally truncated <span class="html-italic">P. aeruginosa</span> exotoxin. This figure was created using BioRender.com.</p>
Full article ">Figure 2
<p>HeLa cell binding and internalization of CPAb, CPAb (118S–121S) and trastuzumab in HeLa and SKBR-3 cells. A total of 10,000 events per sample were recorded with the CytoFLEX S cytometer. TAMRA-labeled antibodies and internalized pH-dye labeled antibodies were detected in the PE channel. (<b>A</b>) In the heparin competition assay, HeLa cells were treated with 300 units/mL heparin sodium salt (HS) prior to antibody treatment. Either 2 µM TAMRA-coupled CPAb or trastuzumab, serving as a negative control, were added. (<b>B</b>) For investigation of the binding motif, TAMRA-coupled CPAb (118S–121S), CPAb (as positive control) or trastuzumab (as negative control) were added to HeLa cells. Fluorescence was measured 8 h after treatment. The internalization into SKBR-3 cells of (<b>C</b>) CPAb (118S–121S), (<b>D</b>) CPAb, (<b>E</b>) Trastuzumab and (<b>F</b>) HerT4 using a pH-dependent dye was analyzed after 24 h. The determined dye-to-antibody ratios ranged from 6.7–6.9. Untreated SKBR-3 cells were used as negative controls.</p>
Full article ">Figure 3
<p>CLSM images as brightfield images or TAMRA fluorescence images of HeLa wildtype cells with (<b>A</b>) 1.5 µM CPAb-TAMRA, (<b>B</b>) 1.5 µM CPAb (118S–121S)-TAMRA, or (<b>C</b>) 1.5 µM trastuzumab-TAMRA. (<b>D</b>,<b>E</b>) show 2-fold magnifications of cells treated with CPAb-TAMRA or trastuzumab-TAMRA, respectively. The white scale bar is equivalent to 50 µm. Fluorescence images were generated using ImageJ 1.53c.</p>
Full article ">Figure 4
<p>CLSM images as brightfield images or GFP fluorescence images of HeLa wildtype cells (<b>A</b>) untreated or treated with (<b>B</b>) 1 µM CPAb-GFP, (<b>C</b>) 1 µM HerT4-GFP, (<b>D</b>) 1 µM CPAb-(118S–121S)-GFP, and (<b>E</b>) Trastuzumab-GFP. The white scale bar is equivalent to 50 µm. Fluorescence images were generated using ImageJ 1.53c.</p>
Full article ">Figure 5
<p>Proliferation assays of CPAb and HerT4, in comparison to several negative controls on HeLa, SKBR-3 and A-431 cells. The antibodies, coupled or uncoupled, were added to the cells for 72 h in a dilution series ranging from 0.2 to 500 nM. The resulting data points are shown as mean and error bars that represent standard deviation derived from experimental duplicates or triplicates. EC<sub>50</sub> values were determined from variable slope four-parameter fitting using GraphPad Prism 10.1.0 (316). (<b>A</b>) The negative controls trastuzumab and mutated variant CPAb (118S–121S) were tested, coupled, and uncoupled on HeLa cells. (<b>B</b>) The cytosol-penetrating antibodies CPAb and HerT4 were analyzed as PE<sub>cat</sub> conjugate, resulting in EC<sub>50</sub> values of 200 nM or 300 nM, respectively. (<b>C</b>) Cytosol-penetration of coupled and uncoupled CPAb-PE<sub>cat</sub> and HerT4-PE<sub>cat</sub> were further tested in comparison with internalizing trastuzumab-PE<sub>cat</sub>, using SKBR-3 cells, and not resulting in a determinable EC<sub>50</sub> value. (<b>D</b>) Proliferation assay of CPAb-PE<sub>cat</sub> and HerT4-PE<sub>cat</sub> in A-431 cells.</p>
Full article ">Figure 6
<p>NanoBiT assay for cytosol-penetrating studies of CPAb-HiBiT<sub>2</sub> and HerT4-HiBiT<sub>2</sub>, in comparison to HiBiT peptide in different concentrations (500 nM, 250 nM, and 125 nM). The concentration of the HiBiT peptide was adjusted to the amount of peptide per antibody, resulting in 2-fold higher concentrations. The results were shown as mean values, with error bars representing the standard deviation resulting from experimental duplicates. One-way ANOVA with Tukey’s multiple comparisons tests (with <span class="html-italic">p</span> value style GP: 0.1234 (ns), 0.0002 (***), and &lt;0.0001 (****)) were used to display the significance level (with definition of statistical significance: <span class="html-italic">p</span> &lt; 0.05). Statistical analysis was performed in GraphPad Prism 10.1.0 (316).</p>
Full article ">Figure 7
<p>CLSM images of the brightfield or GFP fluorescence channels of HeLa cells treated with 1 µM antibody–GFP conjugates. (<b>A</b>) CPAb (118S–121S)-GFP, (<b>B</b>) CPAb-GFP, (<b>C</b>) S4-CPAb-GFP, (<b>D</b>) S4-CPAb-GFP (MMP-9 cleaved), (<b>E</b>) S5-CPAb-GFP and (<b>F</b>) S5-CPAb-GFP (MMP-9 cleaved) were incubated for 8 h with HeLa cells and subsequent washing and fixation with 4% PFA. For the GFP fluorescence imaging, the laser with 488 nm was utilized. The scale bar is equivalent to 50 µm. Fluorescence images were generated using ImageJ 1.53c.</p>
Full article ">Figure 8
<p>Cytosol-penetration capabilities of the masked S4-CPAb (MMP-9 cleaved and uncleaved). (<b>A</b>) Proliferation assay of S4-CPAb-PE<sub>cat</sub> (MMP-9 cleaved und untreated), in comparison to unconjugated PE<sub>cat</sub> and S4-CPAb in HeLa cells at different concentrations (0.2–500 nM). The resulting data points, shown as mean and error bars, represent standard deviation derived from experimental duplicates. EC<sub>50</sub> values were determined from variable slope four-parameter fitting using GraphPad Prism 10.1.0 (316). (<b>B</b>) NanoBiT<sup>®</sup> assay of S4-CPAb-HiBiT<sub>2</sub> (MMP-9 cleaved und untreated), in comparison to CPAb-HiBiT<sub>2</sub> and CPAb (118S–121S)-HiBiT<sub>2</sub>. The resulting data points, shown as mean and error bars, represent standard deviation derived from experimental duplicates. An unpaired, two-tailed <span class="html-italic">t</span>-test (with <span class="html-italic">p</span> value style GP: 0.1234 (ns), 0.0332 (*), 0.0021 (**)) was used to display the significance level (with definition of statistical significance: <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Proliferation assay for determination of cytosol-penetration capabilities of the masked S4-CPAb-PE in HeLa (MMP-9 negative) and A-431 (MMP-9 positive) cells at different concentrations (0.2–500 nM). The resulting data points, shown as mean and error bars, represent standard deviation derived from experimental duplicates using GraphPad Prism 10.1.0 (316).</p>
Full article ">
20 pages, 2311 KiB  
Article
LDLR-Mediated Targeting and Productive Uptake of siRNA-Peptide Ligand Conjugates In Vitro and In Vivo
by Baptiste Broc, Karine Varini, Rose Sonnette, Belinda Pecqueux, Florian Benoist, Maxime Masse, Yasmine Mechioukhi, Géraldine Ferracci, Jamal Temsamani, Michel Khrestchatisky, Guillaume Jacquot and Pascaline Lécorché
Pharmaceutics 2024, 16(4), 548; https://doi.org/10.3390/pharmaceutics16040548 - 17 Apr 2024
Cited by 1 | Viewed by 2324
Abstract
Small RNA molecules such as microRNA and small interfering RNA (siRNA) have become promising therapeutic agents because of their specificity and their potential to modulate gene expression. Any gene of interest can be potentially up- or down-regulated, making RNA-based technology the healthcare breakthrough [...] Read more.
Small RNA molecules such as microRNA and small interfering RNA (siRNA) have become promising therapeutic agents because of their specificity and their potential to modulate gene expression. Any gene of interest can be potentially up- or down-regulated, making RNA-based technology the healthcare breakthrough of our era. However, the functional and specific delivery of siRNAs into tissues of interest and into the cytosol of target cells remains highly challenging, mainly due to the lack of efficient and selective delivery systems. Among the variety of carriers for siRNA delivery, peptides have become essential candidates because of their high selectivity, stability, and conjugation versatility. Here, we describe the development of molecules encompassing siRNAs against SOD1, conjugated to peptides that target the low-density lipoprotein receptor (LDLR), and their biological evaluation both in vitro and in vivo. Full article
(This article belongs to the Special Issue Peptide-Based Carriers for Drug Delivery)
Show Figures

Figure 1

Figure 1
<p>General design of siSOD1-peptide conjugates. (<b>A</b>) Detail of the stabilization scheme of siSOD1 duplex. Chemical modifications: green dot = 2’-Ome; blue dot = 2’-F; red line = PS; P = phosphate; VP = vinylphosphonate. (<b>B</b>) Detailed structure of the reactive amine at the 3′-end of the sense strand (SS) for further conjugation. N6 = 6-carbon aliphatic arm ending with an amine group. 5-LC-NU = 5-Aminohexylacrylamino-Uridine, modified uridine with an aminohexylacrylamine arm at position 5. (<b>C</b>) Detailed structure of the 5′-end of the antisens strand (AS). P = standard phosphate; (E)-VP = modified and metabolically stable vinylphosphonate with a double bond in E configuration. (<b>D</b>) Detail of the VH4127 peptide sequence (cyclo[(D)-Cys-Met-Thz-Arg-Leu-Arg-Gly-Pen]) and scheme of its structure. Disulfide cyclization occurred between the penicillamine and cysteine side chains. VH4127 binding affinity to LDLR: Kd = 40.1 nM, surface plasmon resonance (SPR): Biochip NiHC1000m; mode MCK).</p>
Full article ">Figure 2
<p>Synthesis strategies and characterization methods of siSOD1-peptide conjugates. (<b>A</b>) Production process of siSOD1-peptide conjugates. A constraint alkyne (DBCO) was introduced on the N6 modification at the 3′ end of siSOD1 sense strand. An azide function was incorporated in the VH4127 peptide sequence during SPPS in the form of azidolysine (Lys(N<sub>3</sub>) or K(N<sub>3</sub>)) and spaced from it with a PEG2. (<b>B</b>) Alternative synthesis strategy of siSOD1-peptide conjugates. The siSOD1-peptide conjugate was obtained through direct amidation between the N6 modification at the 3′ end of siSOD1m sense strand and the free carboxylic acid of peptide VH4127. (<b>C</b>) LC/MS characterization of siSOD1-peptide conjugates. Buffer A: HFIP 12.5 mM and DIEA 4 mM in H<sub>2</sub>O; Buffer B: HFIP 12.5 mM and DIEA 4 mM in MeOH. Flow rate was 0.3 mL/min and column temperature set at 65 °C. Detection was performed at 260 and 214 nm. MS analysis in negative mode and spectra was deconvoluted.</p>
Full article ">Figure 3
<p>Gene-silencing potency of unconjugated siRNAs (5′P-siSOD1 and 5′VP-siSOD1) and conjugated siRNAs (siSOD1-31, -32, -33, -34, -35, -36, -37) using transfection and free uptake. (<b>A</b>) Transfection of unconjugated siSOD1 and siSOD1-peptide conjugates on Neuro-2a cells at 30 nM. After 24 h of incubation mSOD1 mRNA levels were quantified using RT-qPCR. (<b>B</b>) Free uptake experiments of unconjugated siSOD1 and siSOD1-peptide conjugates on Neuro-2a cells at 1 µM. After 3 days at 37 °C, mSOD1 mRNA levels were quantified using RT-qPCR. Each dot corresponds to the mean value obtained in independent experiments. (<b>C</b>) Free uptake experiments of siSOD1-31, -35, -36, -37 conjugates with their respective negative controls siSOD1-31Sc, -35Sc, -36Sc, -37Sc on Neuro-2a cells at 1 µM. After 3 days at 37 °C, mSOD1 mRNA levels were quantified using RT-qPCR. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>In vivo gene-silencing potency of unconjugated siRNAs (5′P-siSOD1 and 5′VP-siSOD1) and conjugated siRNAs (siSOD1-31, -32, -33, -34, -35, -36, -37) in mice liver. Mice were injected (i.v. lateral tail vein) with unconjugated siRNAs and siSOD1-peptide conjugates in PBS (15 mg/kg). Seven days post-administration mice were euthanized and perfused with saline solution (0.9% NaCl). Organs were collected to assess the SOD1 mRNA levels using RT-qPCR. Each dot corresponds to one mouse. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
Back to TopTop