[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (385)

Search Parameters:
Keywords = adeno-associated virus

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 2484 KiB  
Article
Limb Perfusion Delivery of a rAAV1 Alpha-1 Antitrypsin Vector in Non-Human Primates Is Safe but Insufficient for Therapy
by Debora Pires-Ferreira, Darcy Reil, Qiushi Tang, Meghan Blackwood, Thomas Gallagher, Allison M. Keeler, Jessica A. Chichester, Kristin K. Vyhnal, Jane A. Lindborg, Janet Benson, Dongtao Fu, Terence R. Flotte and Alisha M. Gruntman
Genes 2024, 15(9), 1188; https://doi.org/10.3390/genes15091188 - 10 Sep 2024
Viewed by 280
Abstract
Background/Objectives: α-1 antitrypsin (AAT) deficiency is an inherited, genetic condition characterized by reduced serum levels of AAT and increased risk of developing emphysema and liver disease. AAT is normally synthesized primarily in the liver, but muscle-targeting with a recombinant adeno-associated virus (rAAV) vector [...] Read more.
Background/Objectives: α-1 antitrypsin (AAT) deficiency is an inherited, genetic condition characterized by reduced serum levels of AAT and increased risk of developing emphysema and liver disease. AAT is normally synthesized primarily in the liver, but muscle-targeting with a recombinant adeno-associated virus (rAAV) vector for α-1 antitrypsin (AAT) gene therapy has been used to minimize liver exposure to the virus and hepatotoxicity. Clinical trials of direct intramuscular (IM) administration of rAAV1-hAAT have demonstrated its overall safety and transgene expression for 5 years. However, the failure to reach the therapeutic target level after 100 large-volume (1.5 mL) IM injections of maximally concentrated vector led us to pursue a muscle-targeting approach using isolated limb perfusion. This targets the rAAV to a greater muscle mass and allows for a higher total volume (and thereby a higher dose) than is tolerable by multiple direct IM injections. Limb perfusion has been shown to be feasible in non-human primates using the rAAV1 serotype and a ubiquitous promoter expressing an epitope-tagged AAT matched to the host species. Methods: In this study, we performed a biodistribution and preclinical safety study in non-human primates with a clinical candidate rAAV1-human AAT (hAAT) vector at doses ranging from 3.0 × 1012 to 1.3 × 1013 vg/kg, bracketing those used in our clinical trials. Results: We found that limb perfusion delivery of rAAV1-hAAT was safe and showed a biodistribution pattern similar to previous studies. However, serum levels of AAT obtained with high-dose limb perfusion still reached only ~50% of the target serum levels. Conclusions: Our results suggest that clinically effective AAT gene therapy may ultimately require delivery at doses between 3.5 × 1013–1 × 1014 vg/kg, which is within the dose range used for approved rAAV gene therapies. Muscle-targeting strategies could be incorporated when delivering systemic administration of high-dose rAAV gene therapies to increase transduction of muscle tissues and reduce the burden on the liver, especially in diseases that can present with hepatotoxicity such as AAT deficiency. Full article
(This article belongs to the Special Issue Gene Therapy for Childhood Diseases)
Show Figures

Figure 1

Figure 1
<p>Maps of AAV vector constructs. Linear maps of the AAV gene cassettes used for this study. AAV2-ITR = inverted terminal repeat sequence from AAV2, CMVe = CMV immediate early enhancer, ACTpro = chicken β-actin promoter, Hybrid IVS = hybrid intron with upstream portion from chicken β-actin and downstream portion from rabbit β-globin, hSERPINA1= human <span class="html-italic">SERPINA1</span> coding sequence, rhSERPINA1-myc = rhesus <span class="html-italic">SERPINA1</span> coding sequence with c-myc epitope tag fusion, pA = SV40 polyadenylation signal.</p>
Full article ">Figure 2
<p>Vector genome biodistribution at injection site and muscles. Vector biodistribution was measured in the rectus femoris muscle taken from a biopsy at post-injection day 28 and at the injection site as well as in ipsilateral and contralateral gastrocnemius and rectus femoris muscles at post-injection day 90 from tissue samples collected during necropsy. Results are expressed as genome copies per diploid genome. Individual results are shown and lines indicate means.</p>
Full article ">Figure 3
<p>Vector genome biodistribution in tissues. Vector biodistribution was measured in the lymph node, lung, spleen, heart, liver, and gonads at post-injection day 90 from tissue samples collected during necropsy. Results are expressed as the number of vector genome copies per diploid genome. Individual results are shown and lines indicate means.</p>
Full article ">Figure 4
<p>AAT serum levels. ELISA assays were performed on serum samples from NHPs that received the vector carrying the human AAT gene (rAAV1-CB-hAAT), the c-myc-AAT gene (rAAV1-CB-rhAATmyc), or vehicle control. Expression levels of hAAT or c-myc were quantified from serum samples collected prior to vector injection (day 0) and at 14, 30, 45, 60, 75, and 90 days post-injection. Data are shown as mean ± standard error of the mean. Two technical replicates were performed of the hAAT samples and four technical replicates were performed of the c-myc samples.</p>
Full article ">Figure 5
<p>Immunohistochemical staining of muscle biopsy tissue. Skeletal muscle was collected by biopsy on day 28 post-injection, sections were immunohistochemically stained for hAAT, and the positive staining was quantified as a percentage of the total area. Tissue sections and quantification results are shown. Samples from animal 2002 were not able to be processed and the result is indicated as N/A. Three tissue sections from animal 3002 (mean ± standard error of the mean are shown) and one tissue section from the other animals were analyzed.</p>
Full article ">Figure 6
<p>Immunohistochemical staining of muscle tissue sections. Skeletal muscle tissue was collected during necropsy at day 90 post-injections, sections were stained immunohistochemically stained for hAAT, and the positive staining was quantified as a percentage of the total area. Representative tissue sections and quantification results are shown. Two tissue sections were analyzed from each animal.</p>
Full article ">
13 pages, 3832 KiB  
Article
Thioredoxin-Interacting Protein’s Role in NLRP3 Activation and Osteoarthritis Pathogenesis by Pyroptosis Pathway: In Vivo Study
by Ruba Altahla and Xu Tao
Metabolites 2024, 14(9), 488; https://doi.org/10.3390/metabo14090488 - 7 Sep 2024
Viewed by 332
Abstract
Thioredoxin-interacting protein (TXNIP) has been involved in oxidative stress and activation of the NOD-like receptor protein-3 (NLRP3) inflammasome, directly linking it to the pyroptosis pathway. Furthermore, pyroptosis may contribute to the inflammatory process in osteoarthritis (OA). The purpose of this study was to [...] Read more.
Thioredoxin-interacting protein (TXNIP) has been involved in oxidative stress and activation of the NOD-like receptor protein-3 (NLRP3) inflammasome, directly linking it to the pyroptosis pathway. Furthermore, pyroptosis may contribute to the inflammatory process in osteoarthritis (OA). The purpose of this study was to investigate the role of TXNIP in activating the NLRP3 inflammasome through the pyroptosis pathway in an OA rat model. Destabilization of the medial meniscus (DMM) was induced in the OA model with intra-articular injections of adeno-associated virus (AAV) overexpressing (OE) or knocking down (KD) TXNIP. A total of 48 healthy rats were randomly divided into six groups (N = 8 each). During the experiment, the rats’ weights, mechanical pain thresholds, and thermal pain thresholds were measured weekly. Morphology staining, micro-CT, 3D imaging, and immunofluorescence (IF) staining were used to measure the expression level of TXNIP, and ELISA techniques were employed. OE-TXNIP-AAV in DMM rats aggravated cartilage destruction and subchondral bone loss, whereas KD-TXNIP slowed the progression of OA. The histological results showed that DMM modeling and OE-TXNIP-AAV intra-articular injection caused joint structure destruction, decreased anabolic protein expression, and increased catabolic protein expression and pyroptosis markers. Conversely, KD-TXNIP-AAV slowed joint degeneration. OE-TXNIP-AVV worsened OA by accelerating joint degeneration and damage, while KD-TXNIP-AAV treatment had a protective effect. Full article
(This article belongs to the Section Cell Metabolism)
Show Figures

Figure 1

Figure 1
<p>Behavior assessment. (<b>A</b>) The process diagram of SD rats receiving SHAM/DMM surgery and intra-articular AAV injection was created using Biorender.com; (<b>B</b>) the body weight of rats; (<b>C</b>) the mechanical pain assessment quantifications; (<b>D</b>) the thermal pain assessment quantifications. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>The CT-3D images of rats and histology staining. (<b>A</b>) The CT-3D images of rats after SHAM/DMM surgery and KD-NC, KD-TXNIP, OE-NC, and OE-TXNIP intra-articular injection; (<b>B</b>) the CT-3D quantification of the bone volume/tissue volume fraction (BV/TV), trabecular number (Tb.N), trabecular separation (Tb.Sp), and trabecular thickness (Tb.Th); (<b>C</b>) the hematoxylin and eosin; safranin O/fast staining quantification; (<b>D</b>) the OARSI score, white scalebar = 1 mm, Yellow arrow show the injury region, ns: non-significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, and N = 8.</p>
Full article ">Figure 3
<p>IHC images of rats. (<b>A</b>) Col II and MMP13 IHC images of rats after SHAM/DMM surgery and KD-NC, KD-TXNIP, OE-NC, and OE-TXNIP intra-articular injection of articular cartilage; (<b>B</b>) IHC quantification, (red scalebar = 200 μm), ns: non-significant, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, N = 8.</p>
Full article ">Figure 4
<p>IHC images of rats. (<b>A</b>) Caspase-1, NLRP3, and gasdermin D IHC images of rats after SHAM/DMM surgery and KD-NC, KD-TXNIP, OE-NC, and OE-TXNIP intra-articular injection of articular cartilage; (<b>B</b>) IHC quantification (red scalebar = 200 μm), ns: non-significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, N = 8.</p>
Full article ">Figure 5
<p>Immunofluorescence staining images of rats. (<b>A</b>) TXNIP IF images of rats after SHAM/DMM surgery and KD-NC, KD-TXNIP, OE-NC, and OE-TXNIP intra-articular injection of articular cartilage; (<b>B</b>) IF quantification (green scalebar = 200 μm), ns: non-significant, * <span class="html-italic">p</span> &lt; 0.05, **** <span class="html-italic">p</span> &lt; 0.0001, N = 8.</p>
Full article ">Figure 6
<p>Serum concentration quantification in rats after SHAM/DMM surgery and KD-NC, KD-TXNIP, OE-NC, and OE-TXNIP intra-articular injection of articular cartilage. (<b>A</b>) IL-1β; (<b>B</b>) IL-18, ns: non-significant, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, N = 8.</p>
Full article ">
2 pages, 136 KiB  
Correction
Correction: Lopez-Gordo et al. Natural Adeno-Associated Virus Serotypes and Engineered Adeno-Associated Virus Capsid Variants: Tropism Differences and Mechanistic Insights. Viruses 2024, 16, 442
by Estrella Lopez-Gordo, Kyle Chamberlain, Jalish Mahmud Riyad, Erik Kohlbrenner and Thomas Weber
Viruses 2024, 16(9), 1366; https://doi.org/10.3390/v16091366 - 28 Aug 2024
Viewed by 231
Abstract
There was an update regarding the affiliation for Kyle Chamberlain [...] Full article
(This article belongs to the Special Issue Viral Receptors and Tropism)
16 pages, 2172 KiB  
Article
Adeno-Associated Virus (AAV)-Delivered Exosomal TAT and BiTE Molecule CD4-αCD3 Facilitate the Elimination of CD4 T Cells Harboring Latent HIV-1
by Xiaoli Tang, Huafei Lu, Patrick M. Tarwater, David L. Silverberg, Christoph Schorl and Bharat Ramratnam
Microorganisms 2024, 12(8), 1707; https://doi.org/10.3390/microorganisms12081707 - 18 Aug 2024
Viewed by 825
Abstract
Combinatorial antiretroviral therapy (cART) has transformed HIV infection from a death sentence to a controllable chronic disease, but cannot eliminate the virus. Latent HIV-1 reservoirs are the major obstacles to cure HIV-1 infection. Previously, we engineered exosomal Tat (Exo-Tat) to reactivate latent HIV-1 [...] Read more.
Combinatorial antiretroviral therapy (cART) has transformed HIV infection from a death sentence to a controllable chronic disease, but cannot eliminate the virus. Latent HIV-1 reservoirs are the major obstacles to cure HIV-1 infection. Previously, we engineered exosomal Tat (Exo-Tat) to reactivate latent HIV-1 from the reservoir of resting CD4+ T cells. Here, we present an HIV-1 eradication platform, which uses our previously described Exo-Tat to activate latent virus from resting CD4+ T cells guided by the specific binding domain of CD4 in interleukin 16 (IL16), attached to the N-terminus of exosome surface protein lysosome-associated membrane protein 2 variant B (Lamp2B). Cells with HIV-1 surface protein gp120 expressed on the cell membranes are then targeted for immune cytolysis by a BiTE molecule CD4-αCD3, which colocalizes the gp120 surface protein of HIV-1 and the CD3 of cytotoxic T lymphocytes. Using primary blood cells obtained from antiretroviral treated individuals, we find that this combined approach led to a significant reduction in replication-competent HIV-1 in infected CD4+ T cells in a clonal in vitro cell system. Furthermore, adeno-associated virus serotype DJ (AAV-DJ) was used to deliver Exo-Tat, IL16lamp2b and CD4-αCD3 genes by constructing them in one AAV-DJ vector (the plasmid was named pEliminator). The coculture of T cells from HIV-1 patients with Huh-7 cells infected with AAV-Eliminator viruses led to the clearance of HIV-1 reservoir cells in the in vitro experiment, which could have implications for reducing the viral reservoir in vivo, indicating that Eliminator AAV viruses have the potential to be developed into therapeutic biologics to cure HIV-1 infection. Full article
(This article belongs to the Special Issue Viral Diseases: Current Research and Future Directions)
Show Figures

Figure 1

Figure 1
<p>Cytotoxic T lymphocytes (CTLs) from healthy donor blood clear U1 cells. U1 cells were cocultured with CTLs from healthy donor blood at target:effector = 1:40 in RPMI1640 medium containing 10% fetal bovine serum and 100 units penicillin/100 μg streptomycin/mL. Twenty-four hours later, cells and supernatant were separated by centrifugation. Cell pellets were lysed with Pierce IP Lysis Buffer. The p24 levels in supernatants and cell lysates were measured using an Alliance HIV-1 ELISA Kit (PerkinElmer Inc.) following the manufacturer’s instructions. The OD450nm readouts are shown as relative p24 levels. (<b>A</b>) CTLs kill U1 cells, leading to the release of p24 into culture medium. (<b>B</b>) CTLs kill U1 cells, leading to cell loss and total intracellular p24 level decrease.</p>
Full article ">Figure 2
<p>Autologous cytotoxic T cells fail to effectively eliminate CD4+ T cells harboring reactivated HIV-1. Five million T cells from the blood of cART-treated HIV-1-infected individuals were cultured in RPMI 1640 medium supplemented with penicillin-streptomycin, L-glutamine, 0.1 nM IL-7, 1 μM tenofovir, 1 μM nevirapine, 1 μM emtricitabine and treated with control exosomes (Exo-C) or Exo-Tat exosomes for 4 days. On day 5, half of the T cells were used to check the activation of latent HIV-1 by measuring the intracellular HIV-1 mRNA level using RT-qPCR (<b>A</b>). Another half of the T cells were cocultured with MOLT-4/CCR5 cells and irradiated PBMCs in an HIVE assay medium containing 2.5 μg/mL of Phytohemagglutinin (PHA) and 60 U/mL of IL-2 for an additional 14 days. The culture medium was changed every 3 days. The final culture supernatants were used for measuring p24 using Simoa Technology with an analytical sensitivity of 0.0074 pg/mL (<b>B</b>).</p>
Full article ">Figure 3
<p>Construction and expression of CD4-αCD3. (<b>A</b>) Schematic structure of CD4-αCD3. The domains 1 and 2 of CD4 are fused to the single-chain variable fragment of anti-CD3 via a linker. An HA-tag is attached to the C-terminus of the fused protein for the convenience of measuring the expression level of the fusion protein. (<b>B</b>) Expression of CD4-αCD3 in HEK293T cells. An empty vector pAAV-MCS (EV) or expression vector pAAV-CD4-αCD3 was transfected into HEK293T cells, respectively. The protein expression level of CD4-αCD3 was determined by Western blot. (<b>C</b>) Secretion of CD4-αCD3 into culture medium of MOLT-4 cells. MOLT-4 cells were transfected with EV or pAAV-CD4-αCD3. Forty-eight hours post-transfection, the supernatants were collected and precipitated with anti-HA rabbit monoclonal antibody (sepharose beads conjugate). The precipitates were used for Western blot to measure levels of secreted CD4-αCD3.</p>
Full article ">Figure 4
<p>CD4αCD3 mediates the killing of U1 cells by cytotoxic T lymphocytes. U1 cells and CTLs from healthy donor blood were cocultured at target:effector = 1:40 in RPMI1640 medium plus control solution (control) or CD4αCD3 solution. Twenty-four hours later, cells and supernatant were separated by centrifugation. Cell pellets were lysed with Pierce IP Lysis Buffer. The p24 levels in supernatants and cell lysates were measured using an Alliance HIV-1 ELISA Kit (PerkinElmer Inc.) following the manufacturer’s instructions. (<b>A</b>) CD4αCD3 facilitates the killing of U1 cells by CTLs leading to further increase in extracellular p24 level. (<b>B</b>) CD4αCD3 facilitates the killing of U1 cells by CTLs leading to further cell loss and total intracellular p24 level decrease.</p>
Full article ">Figure 5
<p>CD4-αCD3 mediates the elimination of HIV-1-infected resting CD4+ T cells ex vivo. (<b>A</b>) CD4-αCD3 mediates the elimination of CD4+ T cells with HIV-1 reactivated by PMA/I. Five million T cells from the blood of cART-treated HIV-1-infected individuals were cultured in control or CD4-αCD3 medium supplemented with penicillin-streptomycin, L-glutamine, 0.1 nM IL-7, 1 μM tenofovir, 1 μM nevirapine, 1 μM emtricitabine and treated with solvent control or PMA/I for 18 h. The cells were washed three times with regular RPMI 1640 culture medium to remove PMA/I. Half of the T cells were cocultured with 2 million MOLT-4/CCR5 cells and irradiated PBMCs in an HIVE assay medium containing 2.5 μg/mL of PHA and 60 U/mL of IL-2 for a further 14 days. The culture medium was changed every 3 days. The final culture supernatants were used for measuring p24 levels using Simoa technology. (N = 4, <span class="html-italic">p</span> &lt; 0.01). (<b>B</b>) The combination of Exo-Tat and CD4-αCD3 eliminates latent HIV-1 reservoir ex vivo. The experimental procedure was the same as mentioned in <a href="#microorganisms-12-01707-f005" class="html-fig">Figure 5</a>A, except that the T cells were treated with Exo-C exosomes or Exo-Tat exosomes instead of solvent control or PMA/I for 4 days. On day 5, half of the T cells were cocultured with 2 million MOLT-4/CCR5 cells and irradiated PBMCs in an HIVE assay medium containing 2.5 μg/mL of PHA and 60 U/mL of IL-2 for a further 14 days. The final culture supernatants were used for measuring p24 levels using Simoa technology(N = 5, <span class="html-italic">p</span> &lt; 0.0001). (<b>C</b>) LRAs PMA/I or Exo-Tat reactivate latent HIV-1. T cells treated with or without PMA/I or Exo-Tat were used to detect intracellular HIV-1 mRNA level using RT-qPCR on theViiA 7Real-Time PCR System.</p>
Full article ">Figure 6
<p>AAV delivered protein expression in vitro. Molt-4/CCR5 cells were infected with AAV-Eliminator at MOI = 1 × 10<sup>4</sup>. Three days post inoculation, cells and supernatant were separated by centrifugation. The cell pellet was used to prepare cell lysate using IP lysis buffer. Twenty μL of cell lysate were used for Western blot to determine the expression levels of HA-tagged proteins. The supernatant was used to purify exosomes. Twenty μL of exosomes were used for checking the expression levels of HA-tagged proteins by Western blot. The exosome-free supernatant was used to measure CD4-αCD3 expression level by immunoprecipitation-Western blot method. (<b>A</b>) Exo-Tat, IL16lamp2b and CD4-αCD3 expressed in AAV-Eliminator-infected Molt-4 cells. (<b>B</b>) Exo-Tat and IL16lamp2b were detected in the exosomes purified from the supernatant of AAV-Eliminator-infected Molt-4 cells. (<b>C</b>) CD4-αCD3 was detected in the exosome-free.</p>
Full article ">Figure 7
<p>AAV delivered protein expression in vivo. AAV-DJ-GFP viruses (control) or AAV-Eliminator viruses (Eliminator) were injected intravenously into Balb/cJ mice via tail veins (1 × 10<sup>12</sup> GC/mouse in 100 μL PBS). Thirty-one days post injection, the mice were euthanized with overdose isoflurane and various tissues were taken out for Western blot, immunohistochemistry (IHC) or H&amp;E staining. (<b>A</b>) Western blot showing HA-tagged proteins expressed in various tissues. (<b>B</b>) IHC staining showing HA-tagged proteins express in various tissues (Brown color). (<b>C</b>) H&amp;E staining showing AAV delivered target proteins do not change the structures of the tissues.</p>
Full article ">Figure 8
<p>AAV-Eliminator reduces latent HIV-1 reservoir in T cells isolated from HIV-1-infected, cART-treated patients. One million Huh-7 cells were infected with control AAV-GFP (Control) or AAV-Eliminator (Eliminator) at MOI = 2 × 10<sup>4</sup>. Two days post infection, the cells were washed with PBS and cocultured with 4 million T cells isolated from HIV-1-infected patient PBMCs in RPMI 1640 medium supplemented with penicillin-streptomycin, L-glutamine, 0.1 nM IL-7, 1 μM tenofovir, 1 μM nevirapine, 1 μM emtricitabine. Four days later, T cells were collected and used for total DNA preparation. Three μL out of 120 μL DNA were used for each ddPCR reaction. Three replicates of ddPCR reactions were performed for each patient sample. (<b>A</b>) AAV-Eliminator reduces latent HIV-1 reservoir in T cells isolated from patient #Lu210. (<b>B</b>) AAV-Eliminator reduces latent HIV-1 reservoir in T cells isolated from patient #Lu107. (<b>C</b>) AAV-Eliminator reduces latent HIV-1 reservoir in T cells isolated from patient #Lu223.</p>
Full article ">
21 pages, 4463 KiB  
Article
Human Stimulator of Interferon Genes Promotes Rhinovirus C Replication in Mouse Cells In Vitro and In Vivo
by Monty E. Goldstein, Maxinne A. Ignacio, Jeffrey M. Loube, Matthew R. Whorton and Margaret A. Scull
Viruses 2024, 16(8), 1282; https://doi.org/10.3390/v16081282 - 10 Aug 2024
Viewed by 863
Abstract
Rhinovirus C (RV-C) infects airway epithelial cells and is an important cause of acute respiratory disease in humans. To interrogate the mechanisms of RV-C-mediated disease, animal models are essential. Towards this, RV-C infection was recently reported in wild-type (WT) mice, yet, titers were [...] Read more.
Rhinovirus C (RV-C) infects airway epithelial cells and is an important cause of acute respiratory disease in humans. To interrogate the mechanisms of RV-C-mediated disease, animal models are essential. Towards this, RV-C infection was recently reported in wild-type (WT) mice, yet, titers were not sustained. Therefore, the requirements for RV-C infection in mice remain unclear. Notably, prior work has implicated human cadherin-related family member 3 (CDHR3) and stimulator of interferon genes (STING) as essential host factors for virus uptake and replication, respectively. Here, we report that even though human (h) and murine (m) CDHR3 orthologs have similar tissue distribution, amino acid sequence homology is limited. Further, while RV-C can replicate in mouse lung epithelial type 1 (LET1) cells and produce infectious virus, we observed a significant increase in the frequency and intensity of dsRNA-positive cells following hSTING expression. Based on these findings, we sought to assess the impact of hCDHR3 and hSTING on RV-C infection in mice in vivo. Thus, we developed hCDHR3 transgenic mice, and utilized adeno-associated virus (AAV) to deliver hSTING to the murine airways. Subsequent challenge of these mice with RV-C15 revealed significantly higher titers 24 h post-infection in mice expressing both hCDHR3 and hSTING—compared to either WT mice, or mice with hCDHR3 or hSTING alone, indicating more efficient infection. Ultimately, this mouse model can be further engineered to establish a robust in vivo model, recapitulating viral dynamics and disease. Full article
(This article belongs to the Special Issue Rhinoviruses and Asthma)
Show Figures

Figure 1

Figure 1
<p>Murine CDHR3 is localized to similar tissues to human CDHR3, yet differs in several key aspects. (<b>A</b>) qPCR of murine CDHR3 expression relative to murine GAPDH by tissue across <span class="html-italic">n</span> = 4 male (M) and <span class="html-italic">n</span> = 4 female (F) mice. Mean +/− standard deviation. (<b>B</b>) Immunohistochemical detection of CDHR3 (green) in a BALB/c mouse lung tissue. Nuclei (DAPI; blue) and cilia (acetylated alpha tubulin; red). White asterisk indicates airway lumen. Scale bar = 20 µm. (<b>C</b>) Table detailing CDHR3 amino acid conservation in chimpanzees and mice. (<b>D</b>) The cryo-EM structure of RV-C15 (green) in complex with the human CDHR3 EC1 (blue) is shown as a Cα trace. The three RV-C15 capsid proteins that interact with EC1 (VP1, VP2, and VP3) are all shown as the same color. The side chains that comprise the RV-C15-EC1 interface are represented as sticks. Residues that differ from the mouse ortholog are labeled. (<b>E</b>) Homology of CDHR3 at position 529 (indicated by the black arrow), performed using ClustalX2.</p>
Full article ">Figure 2
<p>Mouse cells replicate RV-C, and replication is significantly increased upon expression of human STING. (<b>A</b>) Experimental design schematic. HeLa-H1 cells or LET1 cells were transfected in parallel with in vitro-transcribed RV RNA or mock-transfected. (<b>B</b>) In situ hybridization (ISH) probes (dark) identify the presence of negative-sense RV RNA, indicative of replicating RV for RV-C15-transfected LET1 (right) and HeLa-H1 cells (positive control, left), but not in mock or replication-deficient (RV-C15-GAA) genome-transfected cells at 24 hpt. Scale bar = 20 µm. (<b>C</b>) Western blot shows inducible expression of hSTING in modified LET1 cells. (<b>D</b>) Immunofluorescence-mediated detection of dsRNA (red) in HeLa-H1 cells (positive control), LET1, and LET1-hSTING with or without induction after transfection with either RV-C15 or RV-A1A RNA (positive control), but not in mock-transfected wells, confirms RV-C15 replication in a mouse cell background. Staining was performed 24 hpt. Nuclei (Hoechst; blue). Scale bar = 20 µm. (<b>E</b>) CellProfiler quantification of the % dsRNA-positive cells as from panel (<b>D</b>). Each individual point (square, circle, or triangle) represents the mean from an independent experiment. In each experiment, &gt;30 cells were assayed across <span class="html-italic">n</span> = 4 immunofluorescence images taken at predetermined locations within the culture. Error bars represent standard deviation of the mean across experiments. (<b>F</b>) CellProfiler quantification of the average dsRNA intensity in dsRNA-positive cells. Statistical analysis in panels (<b>E</b>,<b>F</b>) was performed using a Mann–Whitney U test (two-tailed; 0.95% confidence interval; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>RV-C produced in mouse cells is infectious and capable of reinfection. (<b>A</b>) Experimental design schematic. HeLa-H1, LET1, or LET1-hSTING cells with or without induction of hSTING were transfected in parallel with in vitro-transcribed RV RNA or mock-transfected. At 24 hpt, supernatants were collected, filtered to remove any cellular material, and passaged to naïve HeLa-E8 cells. Infection of HeLa-E8 cells was confirmed by in situ hybridization or immunofluorescence staining. (<b>B</b>) In situ hybridization (ISH) probes (dark) identify the presence of negative-sense RV RNA, indicative of replicating RV in HeLa-E8 after inoculation with the supernatants described in (<b>A</b>), confirming infectious particle formation in LET1 cells. β-Actin ISH probes validate RNA integrity in the sample. Scale bar = 20 µm. (<b>C</b>) Immunofluorescence-mediated detection of dsRNA (red) in HeLa-E8 cells 24 hpi with supernatants harvested from HeLa-H1 (positive control), LET1, LET1-hSTING (with or without induction) cells previously transfected with either RV-C15 or RV-A1A RNA (positive control) but not mock. Nuclei (Hoechst; blue). Scale bar = 20 µm.</p>
Full article ">Figure 4
<p>RV-C infection of transgenic mice expressing CDHR3 with the addition of human STING results in prolonged viral replication. (<b>A</b>) Human CDHR3 expression in the lungs and trachea of transgenic founder lines, Tg1, Tg2, and Tg3 by qPCR across <span class="html-italic">n</span> = 4 male (M) and <span class="html-italic">n</span> = 4 female (F) mice. Mean +/− standard deviation. ND = not detected. (<b>B</b>) Immunohistochemical detection of human and murine CDHR3 (green) in Tg1 mouse lung tissue compared to WT mice. Nuclei (Hoechst; blue) and cilia (acetylated alpha tubulin; red). Scale bar = 20 µm. (<b>C</b>) Experimental design schematic created with BioRender.com. Transgenic mice expressing human CDHR3 (Tg1) or WT mice were transduced with either human STING, or firefly luciferase (Fluc) via adeno-associated virus (AAV) by intranasal inoculation. On day 10 post-AAV transduction, mice were inoculated intranasally with RV-C15 or UV-inactivated RV-C15. Mice were then sacrificed, and total lungs were collected at 12 and 24 hpi. (<b>D</b>) qPCR detection of RV genome copy numbers at 12 and 24 hpi. Each point represents mean genome copy number from <span class="html-italic">n</span> = 11–14 mice assayed across three independent experiments. Black shapes represent WT mice and red shapes represent hCDHR3 Tg mice. Triangles represent mice receiving AAV-Fluc and circles represent mice receiving AAV-STING. Empty shapes represent UV-inactivated RV-C15 conditions while filled-in shapes represent mice receiving infectious RV-C15. Statistical analysis was performed using a Mann–Whitney U test (two-tailed; 0.95% confidence interval; * <span class="html-italic">p</span> &lt; 0.05). (<b>E</b>) Immunohistochemical detection of dsRNA (J2; red) in paraffin-embedded mouse lung sections 24 hpi. Nuclei (Hoechst; blue). Scale bar = 20 μm.</p>
Full article ">
35 pages, 4052 KiB  
Review
Therapeutic Application and Structural Features of Adeno-Associated Virus Vector
by Yasunari Matsuzaka and Ryu Yashiro
Curr. Issues Mol. Biol. 2024, 46(8), 8464-8498; https://doi.org/10.3390/cimb46080499 - 2 Aug 2024
Viewed by 712
Abstract
Adeno-associated virus (AAV) is characterized by non-pathogenicity, long-term infection, and broad tropism and is actively developed as a vector virus for gene therapy products. AAV is classified into more than 100 serotypes based on differences in the amino acid sequence of the capsid [...] Read more.
Adeno-associated virus (AAV) is characterized by non-pathogenicity, long-term infection, and broad tropism and is actively developed as a vector virus for gene therapy products. AAV is classified into more than 100 serotypes based on differences in the amino acid sequence of the capsid protein. Endocytosis involves the uptake of viral particles by AAV and accessory receptors during AAV infection. After entry into the cell, they are transported to the nucleus through the nuclear pore complex. AAVs mainly use proteoglycans as receptors to enter cells, but the types of sugar chains in proteoglycans that have binding ability are different. Therefore, it is necessary to properly evaluate the primary structure of receptor proteins, such as amino acid sequences and post-translational modifications, including glycosylation, and the higher-order structure of proteins, such as the folding of the entire capsid structure and the three-dimensional (3D) structure of functional domains, to ensure the efficacy and safety of biopharmaceuticals. To further enhance safety, it is necessary to further improve the efficiency of gene transfer into target cells, reduce the amount of vector administered, and prevent infection of non-target cells. Full article
Show Figures

Figure 1

Figure 1
<p>Structure of the wild-type AAV genome. Rep78 and Rep68 are expressed from the p5 promoter and Rep52 and Rep40 from the p19 promoter. VP1, 2, 3 and the assembly-activating protein (AAP) are translated from the p40 transcript encoded by the cap gene.</p>
Full article ">Figure 2
<p><b>(Upper)</b> Amino acid sequence in VP1 and VP2 of AAV type 1 to type 12. (<b>Bottom</b>) Amino acid identity of VP1 and VP2 of AAV type 1 to type 12. Yellows highlight the conserved amino acid motif. BR3 domain is indicated by green.</p>
Full article ">Figure 2 Cont.
<p><b>(Upper)</b> Amino acid sequence in VP1 and VP2 of AAV type 1 to type 12. (<b>Bottom</b>) Amino acid identity of VP1 and VP2 of AAV type 1 to type 12. Yellows highlight the conserved amino acid motif. BR3 domain is indicated by green.</p>
Full article ">Figure 3
<p>AAV2 structures at a 9.7-Å resolution. Example capsid surface regions corresponding to VR-I and VR-IV are indicated by arrows in panels [<a href="#B45-cimb-46-00499" class="html-bibr">45</a>].</p>
Full article ">Figure 4
<p>Transduction model of by AAV vectors via binding AAV to cell surface receptor. AAV was internalized into cytoplasm of cell by endosomal trafficking via interaction with AAV receptor (AAVR). The internalized AAVs entry into nucleus through endosomal escape, and then the AAVs within the nucleus release viral single-stranded DNA via uncoating, which forms nucleoprotein filament complex due to transcription.</p>
Full article ">Figure 5
<p>Allele frequency of HLA class II genes [<a href="#B126-cimb-46-00499" class="html-bibr">126</a>,<a href="#B131-cimb-46-00499" class="html-bibr">131</a>,<a href="#B132-cimb-46-00499" class="html-bibr">132</a>].</p>
Full article ">Figure 6
<p>Construction of AAV variant library.</p>
Full article ">
13 pages, 3193 KiB  
Article
The Neuroprotective Effect of Neural Cell Adhesion Molecule L1 in the Hippocampus of Aged Alzheimer’s Disease Model Mice
by Miljana Aksic, Igor Jakovcevski, Mohammad I. K. Hamad, Vladimir Jakovljevic, Sanja Stankovic and Maja Vulovic
Biomedicines 2024, 12(8), 1726; https://doi.org/10.3390/biomedicines12081726 - 1 Aug 2024
Viewed by 567
Abstract
Alzheimer’s disease (AD) is a severe neurodegenerative disorder and the most common form of dementia, causing the loss of cognitive function. Our previous study has shown, using a doubly mutated mouse model of AD (APP/PS1), that the neural adhesion molecule L1 directly binds [...] Read more.
Alzheimer’s disease (AD) is a severe neurodegenerative disorder and the most common form of dementia, causing the loss of cognitive function. Our previous study has shown, using a doubly mutated mouse model of AD (APP/PS1), that the neural adhesion molecule L1 directly binds amyloid peptides and decreases plaque load and gliosis when injected as an adeno-associated virus construct (AAV-L1) into APP/PS1 mice. In this study, we microinjected AAV-L1, using a Hamilton syringe, directly into the 3-month-old APP/PS1 mouse hippocampus and waited for a year until significant neurodegeneration developed. We stereologically counted the principal neurons and parvalbumin-positive interneurons in the hippocampus, estimated the density of inhibitory synapses around principal cells, and compared the AAV-L1 injection models with control injections of green fluorescent protein (AAV-GFP) and the wild-type hippocampus. Our results show that there is a significant loss of granule cells in the dentate gyrus of the APP/PS1 mice, which was improved by AAV-L1 injection, compared with the AAV-GFP controls (p < 0.05). There is also a generalized loss of parvalbumin-positive interneurons in the hippocampus of APP/PS1 mice, which is ameliorated by AAV-L1 injection, compared with the AAV-GFP controls (p < 0.05). Additionally, AAV-L1 injection promotes the survival of inhibitory synapses around the principal cells compared with AAV-GFP controls in all three hippocampal subfields (p < 0.01). Our results indicate that L1 promotes neuronal survival and protects the synapses in an AD mouse model, which could have therapeutic implications. Full article
(This article belongs to the Section Neurobiology and Clinical Neuroscience)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Schematic depiction of the experimental design. (<b>B</b>) Representative image of the APP/PS1 mouse hippocampus injected with AAV-GFP (green) and immunofluorescently stained for NeuN (red). The overlay (yellow) highlights the transduced cells. (<b>C</b>) Higher magnification of the CA1 hippocampal subfield, with GFP (green) transduced pyramidal cells. (<b>D</b>,<b>E</b>). Immunostaining with L1 555 antibody (green) in wild-type (<b>D</b>) and AAV-L1 injected (<b>E</b>) hippocampus cells. Note the robust transduction of the CA1 pyramidal cells with AAV-L1. CA—cornu ammonis, DG—dentate gyrus, pyr—stratum pyramidale, rad—stratum radiatum. Scale bars: 200 µm (<b>B</b>); 25 µm (<b>C</b>–<b>E</b>).</p>
Full article ">Figure 2
<p>Injection of AAV-L1 reduces the loss of hippocampal NeuN-positive granule neurons in the dentate gyrus of APP/PS1 mice. (<b>A</b>,<b>C</b>,<b>E</b>) Representative images of NeuN-immunostained (NeuN+) neurons in the CA1 (<b>A</b>), CA3 (<b>C</b>), and DG (<b>E</b>) subfields of the hippocampus. Or—stratum oriens, pyr—stratum pyramidale, rad—stratum radiatum, gr—stratum granulosum, hil—hilus of the dentate gyrus. Scale bar: 50 µm. (<b>B</b>,<b>D</b>,<b>F</b>) Densities of the hippocampal NeuN-positive neurons in the pyramidal layer of the CA1 (<b>B</b>), CA3 (<b>D</b>), and granule cells in the DG (<b>F</b>) in wild-type (WT), AAV-L1, and AAV-GFP-injected APP/PS1 mice. Data are shown as mean + standard deviation. Asterisks indicate the differences between treatments, while the hashtag indicates a difference from the wild-type control; one-way ANOVA with Holm–Sidak post hoc, <span class="html-italic">p</span> &lt; 0.05; n = 5 mice/group.</p>
Full article ">Figure 3
<p>Injection of AAV-L1 reduces the loss of hippocampal parvalbumin-positive interneurons in APP/PS1 mice. (<b>A</b>,<b>C</b>,<b>E</b>) Representative images of parvalbumin-immunostained (PV+) interneurons in the CA1 (<b>A</b>), CA3 (<b>C</b>), and DG (<b>E</b>) subregions of the hippocampus. Or—stratum oriens, pyr—stratum pyramidale, rad—stratum radiatum, gr—stratum granulosum, mol—stratum moleculare, hil—hilus of the dentate gyrus. Scale bar: 25 µm. (<b>B</b>,<b>D</b>,<b>F</b>) Densities of PV+ neurons in the CA1 (<b>B</b>), CA3 (<b>D</b>), and DG (<b>F</b>) in the wild-type (WT) and AAV-L1 or AAV-GFP-injected APP/PS1 mice. Data are shown as mean + standard deviation. Asterisks indicate the difference between treatments, the hashtag indicates a difference from the wild-type control; one-way ANOVA with Holm–Sidak post hoc, <span class="html-italic">p</span> &lt; 0.05; n = 5 mice/group.</p>
Full article ">Figure 4
<p>The loss of perisomatic inhibitory synapses on principal neuron cell bodies in the hippocampus of APP/PS1 mice is reduced by AAV-L1 injection. (<b>A</b>,<b>C</b>,<b>E</b>) Representative confocal micrographs of VGAT-(red) and parvalbumin (PV, green)-immunostained perisomatic terminals around CA1 (<b>A</b>), CA3 (<b>C</b>), pyramidal neurons (pyr) and DG granule cells (gr) (<b>E</b>). Scale bar: 10 µm. (<b>B</b>,<b>D</b>,<b>F</b>) Diagrams represent the number of parvalbumin-positive/VGAT-positive (PV+) and parvalbumin-negative/VGAT-positive (PV-) perisomatic terminals per unit length (mm) in the CA1 (<b>B</b>), CA3 (<b>D</b>), and DG (<b>F</b>) of wild-type (WT) mice, and APP/PS1 mice injected with either AAV-L1 or AAV-GFP. Data are shown as mean + SD. Asterisks indicate a difference between the injections (AAV-L1 or AAV-GFP), hashtags indicate a difference between APP/PS1 mice and the wild-type control, in a two-way ANOVA with the factors “parvalbumin expression” and “viral injection”, followed by the Holm–Sidak post hoc method, <span class="html-italic">p</span> &lt; 0.05; n = 5 mice/group.</p>
Full article ">
25 pages, 8663 KiB  
Article
In-Depth Comparison of Adeno-Associated Virus Containing Fractions after CsCl Ultracentrifugation Gradient Separation
by Mojca Janc, Kaja Zevnik, Ana Dolinar, Tjaša Jakomin, Maja Štalekar, Katarina Bačnik, Denis Kutnjak, Magda Tušek Žnidarič, Lorena Zentilin, Dmitrii Fedorov and David Dobnik
Viruses 2024, 16(8), 1235; https://doi.org/10.3390/v16081235 - 31 Jul 2024
Viewed by 1130
Abstract
Recombinant adeno-associated viruses (rAAVs) play a pivotal role in the treatment of genetic diseases. However, current production and purification processes yield AAV-based preparations that often contain unwanted empty, partially filled or damaged viral particles and impurities, including residual host cell DNA and proteins, [...] Read more.
Recombinant adeno-associated viruses (rAAVs) play a pivotal role in the treatment of genetic diseases. However, current production and purification processes yield AAV-based preparations that often contain unwanted empty, partially filled or damaged viral particles and impurities, including residual host cell DNA and proteins, plasmid DNA, and viral aggregates. To precisely understand the composition of AAV preparations, we systematically compared four different single-stranded AAV (ssAAV) and self-complementary (scAAV) fractions extracted from the CsCl ultracentrifugation gradient using established methods (transduction efficiency, analytical ultracentrifugation (AUC), quantitative and digital droplet PCR (qPCR and ddPCR), transmission electron microscopy (TEM) and enzyme-linked immunosorbent assay (ELISA)) alongside newer techniques (multiplex ddPCR, multi-angle light-scattering coupled to size-exclusion chromatography (SEC-MALS), multi-angle dynamic light scattering (MADLS), and high-throughput sequencing (HTS)). Suboptimal particle separation within the fractions resulted in unexpectedly similar infectivity levels. No single technique could simultaneously provide comprehensive insights in the presence of both bioactive particles and contaminants. Notably, multiplex ddPCR revealed distinct vector genome fragmentation patterns, differing between ssAAV and scAAV. This highlights the urgent need for innovative analytical and production approaches to optimize AAV vector production and enhance therapeutic outcomes. Full article
(This article belongs to the Section General Virology)
Show Figures

Figure 1

Figure 1
<p>Representative part of the scAAV heavy fraction micrograph taken with a TEM Philips CM 100 showing different viral particles (blue = partially filled particle, black = damaged particle, white = empty particle and yellow = full viral particle). The full size micrograph can be found in the <a href="#app1-viruses-16-01235" class="html-app">Supplementary Materials (Figure S2)</a>.</p>
Full article ">Figure 2
<p>Distributions of the full, partially filled, empty and damaged viral particles using transmission electron microscopy (TEM). Each sample was simultaneously studied on glow-discharged grids (GD+) (1st panel) and on untreated grids (GD−) (2nd panel). The % of each viral particle was determined based on observation of 2000 viral particles separately for two technical replicates of both viral vectors studied. The number (1 or 2) after the fraction name (H = heavy, F = full, I = intermediate, E = empty) represents the technical replicate.</p>
Full article ">Figure 3
<p>Distribution of the full, partially filled, and empty particles determined using analytical ultracentrifugation (AUC). (<b>A</b>,<b>B</b>) Representation of the normalized results from the analytical ultracentrifugation of the AAV fractions ((<b>A</b>) ssAAV and (<b>B</b>) scAAV) (light blue—heavy fraction, black—full fraction, dark blue—intermediate fraction, yellow—empty fraction); both technical replicates of each fraction were combined due to low sample volume. (<b>C</b>) Calculated distribution of the particles.</p>
Full article ">Figure 4
<p>Evaluation of the vector genome titer using qPCR and ddPCR (CMV and GFP assays). The number (1 or 2) after the fraction name (H = heavy, F = full, I = intermediate, E = empty) represents the technical replicate.</p>
Full article ">Figure 5
<p>Concentration of intact viral particles defined with ELISA. The number (1 or 2) after the fraction name (H = heavy, F = full, I = intermediate, E = empty) represents the technical replicate.</p>
Full article ">Figure 6
<p>The % of full viral capsids present in each fraction evaluated by 5 approaches (TEM, combination of capsid ELISA and 2 different ddPCR assays, AUC, and SEC-MALS). * = the defined % of full viral particles was higher than 100, meaning that the number of capsids was lower than the amount of vector genomes determined in those samples. The number (1 or 2) after the fraction name (H = heavy, F = full, I = intermediate, E = empty) represents the technical replicate. Due to volume constraints, both technical replicates were combined for AUC and therefore both technical replicates have the same value represented in the graph.</p>
Full article ">Figure 7
<p>Single-dye duplex ddPCR results showing the presence of full-length genomes as well as the presence of encapsidated genome fragments. Another way of calculating the genome integrity from ddPCR data, i.e., linkage, is also presented. The analysis was performed on one batch of ssAAV and one batch of scAAV fractions.</p>
Full article ">Figure 8
<p>Host cell impurities evaluated with TEM. (<b>A</b>) Parts of the host cells (marked with black arrow). (<b>B</b>) Huge viral particles aggregated (marked with white arrow). Both types of impurities were observed mostly, but not exclusively, in the intermediate and empty fractions.</p>
Full article ">Figure 9
<p>In-depth analysis of the TEM micrographs showed the presence of different contaminants (e.g., host cell proteins—blue arrow, bigger icosahedral viral particle—black arrow, and potentially AAV viral particles with viral Rep protein complexes attached to them—yellow arrow).</p>
Full article ">Figure 10
<p>Average and normalized values of residual host cell DNA in all the fractions for both viral vectors. The number (1 or 2) after the fraction name (H = heavy, F = full, I = intermediate, E = empty) represents the technical replicate.</p>
Full article ">Figure 11
<p>Average and normalized values of the host cell proteins in each fraction for both viral vectors. The number (1 or 2) after the fraction name (H = heavy, F = full, I = intermediate, E = empty) represents the technical replicate.</p>
Full article ">Figure 12
<p>The presence of viral particles, where smaller and larger aggregates were observed with MADLS in one technical replicate of each scAAV fraction.</p>
Full article ">Figure 13
<p>Read mapping results expressed as the % of total Illumina sequencing reads for each DNase I treated sample and DNase I untreated samples. Additionally, 2nd DNA strand synthesis was performed (+) or not (−). H = heavy, F = full, I = intermediate and E = empty fraction. The product represents the % of total reads mapping to the rAAV genome from ITR to ITR, the plasmids represent the % of total reads mapping to any of the plasmids used in the AAV production, and the hcDNA represents the % of total reads mapping to the human genome. Both technical replicates of each fraction were combined due to the low sample volume.</p>
Full article ">
20 pages, 3544 KiB  
Article
Efficient AAV9 Purification Using a Single-Step AAV9 Magnetic Affinity Beads Isolation
by Kian Chuan Sia, Zhen Ying Fu, Siti Humairah Mohd Rodhi, Joan Hua Yi Yee, Kun Qu and Shu Uin Gan
Int. J. Mol. Sci. 2024, 25(15), 8342; https://doi.org/10.3390/ijms25158342 - 30 Jul 2024
Viewed by 786
Abstract
Adeno-associated viruses (AAVs) have emerged as promising tools for gene therapy due to their safety and efficacy in delivering therapeutic genes or gene editing sequences to various tissues and organs. AAV serotype 9 (AAV9), among AAV serotypes, stands out for its ability to [...] Read more.
Adeno-associated viruses (AAVs) have emerged as promising tools for gene therapy due to their safety and efficacy in delivering therapeutic genes or gene editing sequences to various tissues and organs. AAV serotype 9 (AAV9), among AAV serotypes, stands out for its ability to efficiently target multiple tissues, thus holding significant potential for clinical applications. However, existing methods for purifying AAVs are cumbersome, expensive, and often yield inconsistent results. In this study, we explore a novel purification strategy utilizing Dynabeads™ CaptureSelect™ magnetic beads. The AAV9 magnetic beads capture AAV9 with high specificity and recovery between 70 and 90%, whereas the AAVX magnetic beads did not bind to the AAV9. Through continuous interaction with AAVs in solution, these beads offer enhanced clearance of genomic DNA and plasmids even in the absence of endonuclease. The beads could be regenerated at least eight times, and the used beads could be stored for up to six months and reused without a significant reduction in recovery. The potency of the AAV9-purified vectors in vivo was comparable to that of iodixanol purified vectors. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Adeno-associated virus serotype 9 (AAV9) magnetic affinity beads evaluation. (<b>A</b>) Re-purification of iodixanol-purified AAV9-CAG-GFP, diluted to virus concentrations ranging from 1 × 10<sup>11</sup> to 20 × 10<sup>11</sup> total vector genomes (vg) in 500 µL phosphate buffered saline (PBS). (<b>B</b>) Purification of AAV9-CAG-GFP viral supernatant, containing approximately 3 × 10<sup>10</sup> total vg of AAV9 in serum-free Dulbecco’s Modified Eagle Medium (SF-DMEM). Each purification used 0.5 mL of AAV9 viral solution and 40 µL of magnetic beads. Breakthrough and recovery percentages are indicated in the table below. Data are presented as mean (SD) (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 2
<p>AAV9 magnetic affinity beads purification with and without endonuclease treatment. (<b>A</b>) Comparison of magnetic affinity beads purification of AAV9-CAG-GFP viral supernatant with and without endonuclease treatment. Breakthrough and recovery percentages are indicated in the table below. (<b>B</b>) Total human embryonic kidney 293 (HEK293) genomic DNA (gDNA) contamination. (<b>C</b>) Total plasmid DNA contamination. DNA contamination percentages (normalized to untreated supernatant as 100%) in endonuclease-treated supernatant, flowthrough, and elution are indicated in the table below. Endonuclease treatments: 90 U/mL endonuclease with 2 mM magnesium chloride (MgCl<sub>2</sub>) for 30 min in a 37 °C water bath. Data are presented as mean (SD) (<span class="html-italic">n</span> = 3). Statistical significance was determined using Student’s unpaired <span class="html-italic">t</span>-test, where ‘ns’ indicates not significant, * indicates <span class="html-italic">p</span> &lt; 0.05 and is considered statistically significant, and **** indicates <span class="html-italic">p</span> &lt; 0.0001 and is considered extremely significant.</p>
Full article ">Figure 3
<p>Optimizing crude viral supernatant volume and AAV load for magnetic affinity beads purification. (<b>A</b>) Effect of increasing the volume of AAV9-CAG-GFP viral supernatant from 0.5 to 10.0 mL (containing total AAV9 ranging from about 4 × 10<sup>10</sup> to about 8 × 10<sup>11</sup> total vg). (<b>B</b>) Effect of increasing the volume of viral supernatant from 1.5 to 10.0 mL containing a 1.5 mL fixed volume of AAV9-CAG-GFP (containing about 1.2 × 10<sup>11</sup> total vg of AAV9) that was topped up to the final volume with SF-DMEM. (<b>C</b>) Correlation of AAV load versus breakthrough (top panel) and recovery (bottom panel). A red dashed line marks the AAV load where breakthrough is zero, and the AAV load range is estimated using Prism 6 software (indicated by a black arrow). The blue arrow indicates the corresponding AAV9 recovery when breakthrough is zero, while the green arrow indicates the maximum capacity of AAV9 magnetic affinity beads for crude viral supernatant. (<b>D</b>) Effect of increasing the volume of viral supernatant from 1.0 to 10.0 mL containing a 1.0 mL fixed volume of AAV9-CAG-GFP (containing about 8 × 10<sup>10</sup> total vg of AAV9) that was topped up to the final volume with SF-DMEM. Breakthrough and recovery percentages are indicated in the table below. Data are presented as mean (SD) (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3 Cont.
<p>Optimizing crude viral supernatant volume and AAV load for magnetic affinity beads purification. (<b>A</b>) Effect of increasing the volume of AAV9-CAG-GFP viral supernatant from 0.5 to 10.0 mL (containing total AAV9 ranging from about 4 × 10<sup>10</sup> to about 8 × 10<sup>11</sup> total vg). (<b>B</b>) Effect of increasing the volume of viral supernatant from 1.5 to 10.0 mL containing a 1.5 mL fixed volume of AAV9-CAG-GFP (containing about 1.2 × 10<sup>11</sup> total vg of AAV9) that was topped up to the final volume with SF-DMEM. (<b>C</b>) Correlation of AAV load versus breakthrough (top panel) and recovery (bottom panel). A red dashed line marks the AAV load where breakthrough is zero, and the AAV load range is estimated using Prism 6 software (indicated by a black arrow). The blue arrow indicates the corresponding AAV9 recovery when breakthrough is zero, while the green arrow indicates the maximum capacity of AAV9 magnetic affinity beads for crude viral supernatant. (<b>D</b>) Effect of increasing the volume of viral supernatant from 1.0 to 10.0 mL containing a 1.0 mL fixed volume of AAV9-CAG-GFP (containing about 8 × 10<sup>10</sup> total vg of AAV9) that was topped up to the final volume with SF-DMEM. Breakthrough and recovery percentages are indicated in the table below. Data are presented as mean (SD) (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Reusability of magnetic affinity beads for AAV9 purification. (<b>A</b>) AAV9 breakthrough in flowthrough after 8 rounds of reuse. (<b>B</b>) AAV9 recovery in elution after 8 rounds of reuse. Between different rounds of purification, the beads were subjected to either 1 min of washing with wash buffer or 10 min of washing with 6M Guanidine hydrochloride (HCl), followed by two 1-min washes with wash buffer. Breakthrough or recovery percentages (normalized to supernatant as 100%) for flowthrough or elution, respectively, are indicated in the table below. (<b>C</b>) Effect of magnetic beads storage at 4 °C for 1, 3, and 6 months with AAV9 magnetic affinity beads that were used once and regenerated. (<b>D</b>) Reuse the same magnetic beads that were regenerated at each round. The beads were regenerated with 10 min of washing with 6M Guanidine HCl, followed by two 1-min washes with wash buffer, before storage in a 4 °C fridge. Breakthrough and recovery percentages are presented (normalized to AAV9 load as 100%). Data are presented as mean (SD) (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Scalability of magnetic affinity beads for AAV9 purification. (<b>A</b>,<b>B</b>) Upscaling of AAV9 purification with magnetic affinity beads (breakthrough and recovery percentages are indicated in the table below), (<b>C</b>,<b>D</b>) total HEK293 gDNA contamination, and (<b>E</b>,<b>F</b>) total plasmid DNA contamination (percentages of DNA contamination in flowthrough and elution are indicated in the table below) for small scale: 0.5 mL AAV9 supernatant (left panel: <b>A</b>,<b>C</b>,<b>E</b>) and larger scale: 20 mL AAV9 supernatant (right panel: <b>B</b>,<b>D</b>,<b>F</b>). Data are presented as mean (SD) (<span class="html-italic">n</span> = 3). (<b>G</b>) Silver-stained sodium dodecyl-sulfate polyacrylamide gel electrophoresis (SDS-PAGE) of AAV9 preparations purified by iodixanol ultracentrifugation (lanes 1 and 3) and by magnetic beads (lanes 2 and 4). (<b>H</b>) Typical cryo-electron micrographs (cryo-EM) of AAV9-CAG-GFP particles purified by iodixanol ultracentrifugation (left panel) and by magnetic beads (right panel). Scale bar: 20 nm. Blue arrows indicate AAV full capsids, and red arrows indicate AAV empty capsids.</p>
Full article ">Figure 5 Cont.
<p>Scalability of magnetic affinity beads for AAV9 purification. (<b>A</b>,<b>B</b>) Upscaling of AAV9 purification with magnetic affinity beads (breakthrough and recovery percentages are indicated in the table below), (<b>C</b>,<b>D</b>) total HEK293 gDNA contamination, and (<b>E</b>,<b>F</b>) total plasmid DNA contamination (percentages of DNA contamination in flowthrough and elution are indicated in the table below) for small scale: 0.5 mL AAV9 supernatant (left panel: <b>A</b>,<b>C</b>,<b>E</b>) and larger scale: 20 mL AAV9 supernatant (right panel: <b>B</b>,<b>D</b>,<b>F</b>). Data are presented as mean (SD) (<span class="html-italic">n</span> = 3). (<b>G</b>) Silver-stained sodium dodecyl-sulfate polyacrylamide gel electrophoresis (SDS-PAGE) of AAV9 preparations purified by iodixanol ultracentrifugation (lanes 1 and 3) and by magnetic beads (lanes 2 and 4). (<b>H</b>) Typical cryo-electron micrographs (cryo-EM) of AAV9-CAG-GFP particles purified by iodixanol ultracentrifugation (left panel) and by magnetic beads (right panel). Scale bar: 20 nm. Blue arrows indicate AAV full capsids, and red arrows indicate AAV empty capsids.</p>
Full article ">Figure 6
<p>In vivo bioactivity of AAV9-purified by magnetic affinity beads compared to iodixanol-purified AAV9. Ex vivo liver green fluorescent protein (GFP) imaging: (<b>A</b>) representative images of liver GFP imaging acquired; (<b>B</b>) corresponding GFP total radiant efficiency; and (<b>C</b>) quantification of AAV9-CAG-GFP genome copies. Non-invasive luciferase imaging of whole mice: (<b>D</b>) representative images of luciferase imaging acquired; (<b>E</b>) corresponding luciferase total flux measurements from week 1 to week 9; and (<b>F</b>) quantification of AAV9-CAG-Luc genome copies at endpoint. Individual data points are presented, and mean values are indicated (<span class="html-italic">n</span> = 4).</p>
Full article ">
16 pages, 2862 KiB  
Review
The Expression and Function of the Small Nonstructural Proteins of Adeno-Associated Viruses (AAVs)
by Cagla Aksu Kuz, Shane McFarlin and Jianming Qiu
Viruses 2024, 16(8), 1215; https://doi.org/10.3390/v16081215 - 29 Jul 2024
Viewed by 635
Abstract
Adeno-associated viruses (AAVs) are small, non-enveloped viruses that package a single-stranded (ss)DNA genome of 4.7 kilobases (kb) within their T = 1 icosahedral capsid. AAVs are replication-deficient viruses that require a helper virus to complete their life cycle. Recombinant (r)AAVs have been utilized [...] Read more.
Adeno-associated viruses (AAVs) are small, non-enveloped viruses that package a single-stranded (ss)DNA genome of 4.7 kilobases (kb) within their T = 1 icosahedral capsid. AAVs are replication-deficient viruses that require a helper virus to complete their life cycle. Recombinant (r)AAVs have been utilized as gene delivery vectors for decades in gene therapy applications. So far, six rAAV-based gene medicines have been approved by the US FDA. The 4.7 kb ssDNA genome of AAV encodes nine proteins, including three viral structural/capsid proteins, VP1, VP2, and VP3; four large nonstructural proteins (replication-related proteins), Rep78/68 and Rep52/40; and two small nonstructural proteins. The two nonstructured proteins are viral accessory proteins, namely the assembly associated protein (AAP) and membrane-associated accessory protein (MAAP). Although the accessory proteins are conserved within AAV serotypes, their functions are largely obscure. In this review, we focus on the expression strategy and functional properties of the small nonstructural proteins of AAVs. Full article
(This article belongs to the Special Issue Virology and Immunology of Gene Therapy)
Show Figures

Figure 1

Figure 1
<p>AAV life cycle. AAV is attached to the host cell via glycans. Internalization occurs by various endocytic pathways (clathrin-mediated endocytosis, CLIC/GEEC, or micropinocytosis). Internalized AAV is endocytosed in the early endosome, where the externalization of VP1u occurs. VP1u-externalized AAV then traffics towards the nucleus by following a path through the TGN, Golgi, and ER, or after cytoplasmic escape. Escaped AAV can be degraded after ubiquitination in the proteasome. Entry into the nucleus occurs via nuclear pore complexes (NPC). Once AAV is localized into the nucleus, the capsid disassembles, and viral proteins are expressed following transcription. Viral genome replication follows a rolling-hairpin DNA replication (RHR) model [<a href="#B47-viruses-16-01215" class="html-bibr">47</a>,<a href="#B48-viruses-16-01215" class="html-bibr">48</a>]. The ssDNA genome is extended from the 3′-OH of the ITR to form a monomer turnround replicative form (mtRF). The hairpin is nicked at a terminal resolution site (TRS) by Rep78/68. The hairpinned end is unwound, and the 3′-end formed by Rep cleavage is extended to the end of the template strand. The ends of each strand refold into their alternative self-base pairing hairpin structures, and the full-length DNA synthesis from the 3′-primer at the left end of the genome produces one ssDNA genome and one mtRF DNA, which can each serve as a substrate for an additional round of replication. WDR63 may play a critical role in rAAV genome transcription [<a href="#B41-viruses-16-01215" class="html-bibr">41</a>]. Assembly of progeny virion occurs in the nucleus. Progeny virions then exit from the nucleus and are trafficked to the cytoplasm to be released. Created in <a href="http://Biorender.com" target="_blank">Biorender.com</a> (accessed on 25 July 2024).</p>
Full article ">Figure 2
<p>The ITR sequences and structures of AAV2 (<b>A</b>) and AAV5 (<b>B</b>). Both AAV2 and AAV5 ITRs are diagrammed in a “flip” orientation. The complementary A/A’, B/B’, and C/C’ sequences are labeled. The D-sequence is contiguous with the ssDNA genome body. The terminal resolution site (TRS) consists of the cleavage site (purple arrow) and the flanking nucleotides in the purple line box. Both the AAV2 and AAV5 Rep-binding elements (RBEs) consist of &gt;(GAGC)3, as shown in the purple-lined box. In the ITRs of AAV1–4, 6, and 7 ITRs, multiple transcription start sites (TSSs), indicated by arrows, were found to cluster at the RBE [<a href="#B57-viruses-16-01215" class="html-bibr">57</a>]. The Inr of the AAV5 ITR has one TSS at nt 142 [<a href="#B58-viruses-16-01215" class="html-bibr">58</a>]. The AAV5 ITR is only 58% homologous with the AAV2 ITR.</p>
Full article ">Figure 3
<p>The genetic maps of AAV2 and AAV5. Adapted from [<a href="#B58-viruses-16-01215" class="html-bibr">58</a>,<a href="#B65-viruses-16-01215" class="html-bibr">65</a>,<a href="#B70-viruses-16-01215" class="html-bibr">70</a>]. The ~4.7 kb ssDNA genomes of AAV2 (GenBank accession no. AF043303) (<b>A</b>) and AAV5 (AF085716) (<b>B</b>) are depicted with two ORFs (<span class="html-italic">Rep</span> and <span class="html-italic">Cap</span>); one intron that is spliced at one donor (D); two alternative acceptors, A1 as the minor and A2 the major sites, one (AAV2) or two (AAV5) polyadenylation sites (pA); and two ITRs. Three promoters, P5 (AAV2)/P7 (AAV5), P19, and P40, transcribe three sets of AAV mRNAs, P5/P7 mRNAs, P19 mRNAs, and P40/P41 mRNAs, respectively. For AAV2 (<b>A</b>), P5 mRNAs encode Rep78, Rep68m, and Rep68M; and P19 mRNAs encode Rep52, Rep40m, and Rep40M. For AAV5 (<b>B</b>), P7 mRNA encodes only Rep78, and P19 mRNA encodes Rep40. P40 (AAV2)/P41 (AAV5) mRNAs include P40 unspliced mRNA and alternatively spliced mRNAs at D and A1 or A2, designated as P40/P41-SplA1 and P40/P41-SplA2 mRNA, respectively. SplA1 mRNA encodes VP1, and SplA2 mRNA encodes MAAP, VP2, AAP, and VP3 at the initiation codons as indicated. MAAP and AAP are translated from the +1 frame-shifted ORF. Created in <a href="http://Biorender.com" target="_blank">Biorender.com</a> (accessed on 25 July 2024).</p>
Full article ">Figure 4
<p>Alignment of AAP sequences of AAV1–13. Predicted AAP protein sequences derived from +1 frame-shifted ORF in the <span class="html-italic">Cap</span> genes of AAV1–13 are aligned using MUSCLE and colored with the Clustal X scheme. Consensus amino acids are shown with a threshold of 50%. The CUG codon is conserved as the initiator of translation for all AAPs. Predicted AAP protein sequences are obtained from corresponding cap gene sequences as follows: AAV1-AF063497, AAV2-AF043303, AAV3B-AF028705, AAV4-NC_001829, AAV5-AF085716, AAV6-AF028704, AAV7-NC_006260, AAV8-AF513852, AAV9-AY530579, AAV10-AY631965, AAV11-AY631966, AAV12-DQ813647, and AAV13-EU285562. Domains or motifs of predicted amino acids are underlined. Threonine/serine, T/S, rich. Various motifs are predicted and indicated. Amino acids identical in all 13 proteins are indicated by (*), strongly similar amino acids by (:), and weakly similar by “(.)”.</p>
Full article ">Figure 5
<p>Alignment of MAAP sequences of AAV1–13. Predicted MAAP protein sequences derived from +1 frame-shifted ORF in the <span class="html-italic">Cap</span> gene are aligned using MUSCLE and colored with the Clustal X scheme. Consensus amino acids are shown with a threshold of 50%. Predicted MAAP protein sequences are obtained from corresponding <span class="html-italic">Cap</span> gene sequences as follows: AAV1-AF063497, AAV2-AF043303, AAV3B-AF028705, AAV4-NC_001829, AAV5-AF085716, AAV6-AF028704, AAV7-NC_006260, AAV8-AF513852, AAV9-AY530579, AAV10-AY631965, AAV11-AY631966, AAV12-DQ813647, and AAV13-EU285562. Domains or motifs of predicted amino acids are underlined. Amino acids identical in all 13 proteins are indicated by (*), strongly similar amino acids by (:), and weakly similar by “(.)”.</p>
Full article ">Figure 6
<p>Putative egress routes for AAV2 in nonlytic infection. Assembled AAV2 virions are exported from the nucleus to the endoplasmic reticulum (ER), and then reach the cis-Golgi Network (CGN) to the <span class="html-italic">trans</span>-Golgi Network (TGN), which may be mediated by COPII-coated vesicles, like during MVM egress [<a href="#B107-viruses-16-01215" class="html-bibr">107</a>]. In one way, virions are packaged in Rab8-positive cargos, which are directed to the plasma membrane for release as microvesicles. On the other hand, virions-containing vesicles can reach early endosomes marked with Rab11 and then traffic to recycling endosomes or late endosomes. The virions in the recycling endosomes can be released as microvesicles. The virions in late endosomes marked with Rab11 and CD63 can be released as exosomes. Rab7-marked late endosomes are led to lysosomes, which will later be degraded in acidified lysosomes or potentially released in deacidified lysosomes by lysosomal exocytosis. MAAP, as marked with stars, could be involved in each trafficking pathway from the nucleus to the plasma membrane as marked. MVB, multivesicular body. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 25 July 2024).</p>
Full article ">
17 pages, 2816 KiB  
Article
The Regulation of Frontal Cortex Cholesterol Metabolism Abnormalities by NR3C1/NRIP1/NR1H2 Is Involved in the Occurrence of Stress-Induced Depression
by Rui Shi, Yingmin Li, Weihao Zhu, Hongjian Xin, Huihuang Yang, Xiaowei Feng, Zhen Wang, Shujin Li, Bin Cong and Weibo Shi
Int. J. Mol. Sci. 2024, 25(15), 8075; https://doi.org/10.3390/ijms25158075 - 24 Jul 2024
Viewed by 612
Abstract
Stress-induced alterations in central neuron metabolism and function are crucial contributors to depression onset. However, the metabolic dysfunctions of the neurons associated with depression and specific molecular mechanisms remain unclear. This study initially analyzed the relationship between cholesterol and depression using the NHANES [...] Read more.
Stress-induced alterations in central neuron metabolism and function are crucial contributors to depression onset. However, the metabolic dysfunctions of the neurons associated with depression and specific molecular mechanisms remain unclear. This study initially analyzed the relationship between cholesterol and depression using the NHANES database. We then induced depressive-like behaviors in mice via restraint stress. Applying bioinformatics, pathology, and molecular biology, we observed the pathological characteristics of brain cholesterol homeostasis and investigated the regulatory mechanisms of brain cholesterol metabolism disorders. Through the NHANES database, we initially confirmed a significant correlation between cholesterol metabolism abnormalities and depression. Furthermore, based on successful stress mouse model establishment, we discovered the number of cholesterol-related DEGs significantly increased in the brain due to stress, and exhibited regional heterogeneity. Further investigation of the frontal cortex, a brain region closely related to depression, revealed stress caused significant disruption to key genes related to cholesterol metabolism, including HMGCR, CYP46A1, ACAT1, APOE, ABCA1, and LDLR, leading to an increase in total cholesterol content and a significant decrease in synaptic proteins PSD-95 and SYN. This indicates cholesterol metabolism affects neuronal synaptic plasticity and is associated with stress-induced depressive-like behavior in mice. Adeno-associated virus interference with NR3C1 in the prefrontal cortex of mice subjected to short-term stress resulted in reduced protein levels of NRIP1, NR1H2, ABCA1, and total cholesterol content. At the same time, it increased synaptic proteins PSD95 and SYN, effectively alleviating depressive-like behavior. Therefore, these results suggest that short-term stress may induce cholesterol metabolism disorders by activating the NR3C1/NRIP1/NR1H2 signaling pathway. This impairs neuronal synaptic plasticity and consequently participates in depressive-like behavior in mice. These findings suggest that abnormal cholesterol metabolism in the brain induced by stress is a significant contributor to depression onset. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>Dysregulation of Cholesterol Metabolism in the Mouse Brain under Stress. (<b>A</b>–<b>D</b>) Analysis of stress−induced changes in cholesterol-related genes across different brain regions. Heatmap of DEGs with increasing and decreasing trend. Each column represents a sample, and each gene is visualized in a row. Red indicates a high abundance of the gene, and blue indicates a relatively low abundance of the gene. (<b>E</b>) Changes in body weight of mice. (<b>F</b>) In the SPT, the model group mice showed a significant decrease in the distance traveled in the central zone, and time spent in the central zone. In the TST, the immobility time of the modeling mice significantly increased (from 13 mice, <span class="html-italic">n</span> = 13). (<b>G</b>) Plasma cortisol measured in mice. Values are expressed as the mean ± SEM, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 vs. the control group.</p>
Full article ">Figure 2
<p>Dysregulation of Cholesterol Metabolism in the Mouse Brain under Stress. HE staining and thionine staining revealed edema in the amygdala, hippocampus, prefrontal cortex, and cortex of stressed mice.The enlarged image is located in the bottom right corner boxes. The scale bars in the figure are all 100 µm.</p>
Full article ">Figure 3
<p>Dysregulation of Cholesterol Metabolism in the Mouse Brain under Stress. (<b>A</b>) Representative Western blot images and densitometric quantification of PSD-95 and SYN in the frontal cortex (<span class="html-italic">n</span> = 6). (<b>B</b>) Total cholesterol content in serum and frontal cortex of mice (from 4 mice, repeated twice, <span class="html-italic">n</span> = 8). (<b>C</b>) The mRNA level of HMGCR, LDLR, ABCA1, ACAT1, CYP46A1, APOE (from 4 mice, repeated twice, <span class="html-italic">n</span> = 8). (<b>D</b>,<b>F</b>) Representative micrographs and quantitative analysis of ABCA1 immunohistochemical staining in the frontal cortex (from 3 mice, repeated twice, <span class="html-italic">n</span> = 6). (<b>E</b>,<b>G</b>–<b>I</b>) Representative Western blot images and densitometric quantification of HMGCR, CYP46A1, and ACAT1 in the frontal cortex (from 6 mice, repeated tests were conducted with the smallest statistically significant value of <span class="html-italic">n</span>, including HMGCR (<span class="html-italic">n</span> = 10), CYP46A1 (<span class="html-italic">n</span> = 6), and ACAT1 (<span class="html-italic">n</span> = 8)). Values are expressed as the mean ± SEM, * <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 vs. the control group.</p>
Full article ">Figure 4
<p>NR3C1/NRIP1/NR1H2 Pathway Involved in Depression-Like Behavior Induced by Short-Term Stress. (<b>A</b>) GO pathway enrichment results from intersecting genes. (<b>B</b>) Signal pathway of KEGG enrichment analysis. (<b>C</b>) The enriched terms in the Metascape database. (<b>D</b>) PPI network of 75 cholesterol−related genes in the STRING database. The color of the lines indicates the type of interaction evidence. (<b>E</b>) PPI network of 22 hub target genes.</p>
Full article ">Figure 5
<p>NR3C1/NRIP1/NR1H2 Pathway Involved in Depression−Like Behavior Induced by Short−Term Stress. (<b>A</b>,<b>B</b>) GO and KEGG enrichment analysis in the KOBAS database. (<b>C</b>) Enriched terms for central node genes in the Metascape database. (<b>D</b>) Enriched terms for central node genes in the KOBAS database. (<b>E</b>) PPI network of NR3C1, NRIP1, NR1H2 and ABCA1. (<b>G</b>) The mRNA level of NR3C1 and NRIP1 (from 4 mice, repeated twice, <span class="html-italic">n</span> = 8). (<b>F</b>–<b>H</b>) Representative Western blot images and densitometric quantification of NR3C1, NRIP1, and NR1H2 in the frontal cortex (from 6 mice, repeated tests were conducted with the smallest statistically significant value of <span class="html-italic">n</span>, including NR3C1 (<span class="html-italic">n</span> = 6), NRIP1 (<span class="html-italic">n</span> = 8), and NR1H2 (<span class="html-italic">n</span> = 5)). Values are expressed as the mean ± SEM, * <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 vs. the control group.</p>
Full article ">Figure 6
<p>NR3C1/NRIP1/NR1H2 Pathway Involved in Depression−Like Behavior Induced by Short-Term Stress. (<b>A</b>) Immunofluorescence staining was performed in the frontal cortex after injecting inhibitory virus AAV9−Nr3c1−RNAi. (<b>B</b>) The mRNA level of NR3C1 (from 4 mice, repeated tests were conducted with the smallest statistically significant value of <span class="html-italic">n</span>, <span class="html-italic">n</span> = 7). (<b>C</b>) Total cholesterol content in the frontal cortex of mice (from 4 mice, <span class="html-italic">n</span> = 4). (<b>D</b>,<b>E</b>) In the OFT, the ratio of central activity distance to total distance, as well as the ratio of central activity time to total time in the SPT increased. The immobility time in the TST decreased (from 8 mice, <span class="html-italic">n</span> = 8). (<b>F</b>–<b>H</b>) Representative Western blot images and densitometric quantification of NR3C1, NRIP1, NR1H2, PSD-95, and SYN in the frontal cortex (from 4 mice, repeated tests were conducted with the smallest statistically significant value of <span class="html-italic">n</span>, <span class="html-italic">n</span> = 7). (<b>I</b>,<b>J</b>) Representative micrographs and quantitative analysis of ABCA1 immunohistochemical staining in the frontal cortex (from 4 mice, <span class="html-italic">n</span> = 4). Values are expressed as the mean ± SEM, * <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 vs. the control group.</p>
Full article ">
16 pages, 5268 KiB  
Article
Protective Effects of Hepatocyte Stress Defenders, Nrf1 and Nrf2, against MASLD Progression
by May G. Akl, Lei Li and Scott B. Widenmaier
Int. J. Mol. Sci. 2024, 25(15), 8046; https://doi.org/10.3390/ijms25158046 - 24 Jul 2024
Viewed by 699
Abstract
Progression of metabolic dysfunction-associated steatites liver disease (MASLD) to steatohepatitis (MASH) is driven by stress-inducing lipids that promote liver inflammation and fibrosis, and MASH can lead to cirrhosis and hepatocellular carcinoma. Previously, we showed coordinated defenses regulated by transcription factors, nuclear factor erythroid [...] Read more.
Progression of metabolic dysfunction-associated steatites liver disease (MASLD) to steatohepatitis (MASH) is driven by stress-inducing lipids that promote liver inflammation and fibrosis, and MASH can lead to cirrhosis and hepatocellular carcinoma. Previously, we showed coordinated defenses regulated by transcription factors, nuclear factor erythroid 2-related factor-1 (Nrf1) and -2 (Nrf2), protect against hepatic lipid stress. Here, we investigated protective effects of hepatocyte Nrf1 and Nrf2 against MASH-linked liver fibrosis and tumorigenesis. Male and female mice with flox alleles for genes encoding Nrf1 (Nfe2l1), Nrf2 (Nfe2l2), or both were fed a MASH-inducing diet enriched with high fat, fructose, and cholesterol (HFFC) or a control diet for 24–52 weeks. During this period, hepatocyte Nrf1, Nrf2, or combined deficiency for ~7 days, ~7 weeks, and ~35 weeks was induced by administering mice hepatocyte-targeting adeno-associated virus (AAV) expressing Cre recombinase. The effects on MASH, markers of liver fibrosis and proliferation, and liver tumorigenesis were compared to control mice receiving AAV-expressing green fluorescent protein. Also, to assess the impact of Nrf1 and Nrf2 induction on liver fibrosis, HFFC diet-fed C57bl/6J mice received weekly injections of carbon tetrachloride, and from week 16 to 24, mice were treated with the Nrf2-activating drug bardoxolone, hepatocyte overexpression of human NRF1 (hNRF1), or both, and these groups were compared to control. Compared to the control diet, 24-week feeding with the HFFC diet increased bodyweight as well as liver weight, steatosis, and inflammation. It also increased hepatocyte proliferation and a marker of liver damage, p62. Hepatocyte Nrf1 and combined deficiency increased liver steatosis in control diet-fed but not HFFC diet-fed mice, and increased liver inflammation under both diet conditions. Hepatocyte Nrf1 deficiency also increased hepatocyte proliferation, whereas combined deficiency did not, and this also occurred for p62 level in control diet-fed conditions. In 52-week HFFC diet-fed mice, 35 weeks of hepatocyte Nrf1 deficiency, but not combined deficiency, resulted in more liver tumors in male mice, but not in female mice. In contrast, hepatocyte Nrf2 deficiency had no effect on any of these parameters. However, in the 15-week CCL4-exposed and 24-week HFFC diet-fed mice, Nrf2 induction with bardoxolone reduced liver steatosis, inflammation, fibrosis, and proliferation. Induction of hepatic Nrf1 activity with hNRF1 enhanced the effect of bardoxolone on steatosis and may have stimulated liver progenitor cells. Physiologic Nrf1 delays MASLD progression, Nrf2 induction alleviates MASH, and combined enhancement synergistically protects against steatosis and may facilitate liver repair. Full article
(This article belongs to the Special Issue Chronic Liver Disease and Hepatocellular Carcinoma)
Show Figures

Figure 1

Figure 1
<p>HFFC diet promotes weight gain and MASH. Mice fed control diet or HFFC diet for 24 weeks. (<b>A</b>,<b>B</b>) % change in relative body weight and % liver-to-body weight ratio (<b>A</b>) and levels of triglyceride and cholesterol in liver (<b>B</b>), comparing diets (n = 11–17). (<b>C</b>) Liver qPCR for indicated gene expression, normalized by <span class="html-italic">36b4</span> (n = 11–15). (<b>D</b>) Liver sections were stained with hematoxylin and eosin (H&amp;E) or underwent immunohistochemistry (IHC) with antibody-detecting ki67 or p62, with scale in panel. (<b>E</b>) Steatosis and inflammation in liver determined using H&amp;E-stained sections, and % of ki67 positive cells and % area of p62 determined using IHC (n = 8–17). Data in (<b>A</b>–<b>C</b>,<b>E</b>) are mean ± standard error of the mean, with individual data points shown. <span class="html-italic">p</span>-value was determined by multiple unpaired <span class="html-italic">t</span>-tests, with adjustment for multiple comparisons. In (<b>D</b>), black arrows indicate Ki67<sup>+</sup> nuclei and blue arrows indicate p62<sup>+</sup> area.</p>
Full article ">Figure 2
<p>Effect of hepatocyte deficiency for Nrf1, Nrf2, or both in mice chronically fed HFFC diet. Mice fed HFFC diet for 24 weeks. In (<b>A</b>–<b>C</b>), mice were infected with indicated virus on week 22. In (<b>D</b>), mice were infected on week 16. (<b>A</b>) Liver sections stained with hematoxylin and eosin, with scale indicated, and steatosis and inflammation (n = 5–14). (<b>B</b>) Levels of triglyceride and cholesterol in liver (n = 6–14). (<b>C</b>) qPCR analysis for indicated gene expression, normalized by <span class="html-italic">36b4</span> (n = 5–14). (<b>D</b>) Representative sections for immunohistochemistry (IHC) with antibody-detecting ki67 or p62, with scale indicated, and % of ki67 positive cells and % area of p62 (n = 6–23). Data are mean ± standard error of the mean, with individual data points shown (males = circles; females = triangle). In (<b>A</b>–<b>C</b>), <span class="html-italic">p</span>-value was determined by two-way analysis of variance, with Sidak post-test. In (<b>D</b>), <span class="html-italic">p</span>-value was determined by one-way analysis of variance, with Dunnett post-test. In (<b>D</b>), black arrows indicate Ki67<sup>+</sup> nuclei and blue arrows indicate p62<sup>+</sup> area.</p>
Full article ">Figure 3
<p>Effect of hepatocyte deficiency for Nrf1, Nrf2, or both on liver tumor development. (<b>A</b>) Study design showing that mice were fed HFFC diet with 2% cholesterol for 52 weeks and infected on weeks 16, 32, and 48 with indicated virus. Liver tumor analysis was performed at the endpoint. (<b>B</b>,<b>C</b>) % incidence of liver tumors (<b>B</b>) and the number of tumors per liver (<b>C</b>) in males (n = 5–7) and females (n = 5–6). (<b>D</b>) Volume of largest liver tumor in males (n = 5–7) and females (n = 5–6) and representative tumor images (black arrows indicate tumors). Data are mean ± standard error of the mean, with individual data points shown. The <span class="html-italic">p</span>-value was determined by two-way analysis of variance, with Sidak post-test.</p>
Full article ">Figure 4
<p>Combined effect of Nrf2-inducing drug bardoxolone and Nrf1 overexpression on MASH-linked fibrosis. (<b>A</b>) Study design showing C57bl/6J mice that were fed HFFC diet with 2% cholesterol for 24 weeks. Mice were injected with carbon tetrachloride once per week from week 0–15 to induce liver fibrosis. On weeks 16–24, mice were treated as indicated with modulators of Nrf1 and Nrf2 activity. Liver analysis was performed at the endpoint. (<b>B</b>) Liver sections stained with hematoxylin and eosin, with scale indicated in panel, and corresponding score for steatosis and inflammation in liver (n = 12–15). (<b>C</b>) Liver sections stained with fibrosis detecting sirius red, with scale indicated in panel, and corresponding % sirius red<sup>+</sup> area (n = 10–14). (<b>D</b>) Liver qPCR analysis for indicated gene expression, normalized by ribosomal protein <span class="html-italic">36b4</span> (n = 11–15). Data are mean ± standard error of the mean, with individual data points shown (males = circles; females = triangles). The <span class="html-italic">p</span>-value was determined by one-way analysis of variance, with Dunnett post-test.</p>
Full article ">
20 pages, 7429 KiB  
Article
MC4R Localizes at Excitatory Postsynaptic and Peri-Postsynaptic Sites of Hypothalamic Neurons in Primary Culture
by Haven Griffin, Jude Hanson, Kevin D. Phelan and Giulia Baldini
Cells 2024, 13(15), 1235; https://doi.org/10.3390/cells13151235 - 23 Jul 2024
Cited by 1 | Viewed by 793
Abstract
The melanocortin-4 receptor (MC4R) is a G protein-coupled receptor (GPCR) that is expressed in several brain locations encompassing the hypothalamus and the brainstem, where the receptor controls several body functions, including metabolism. In a well-defined pathway to decrease appetite, hypothalamic proopiomelanocortin (POMC) neurons [...] Read more.
The melanocortin-4 receptor (MC4R) is a G protein-coupled receptor (GPCR) that is expressed in several brain locations encompassing the hypothalamus and the brainstem, where the receptor controls several body functions, including metabolism. In a well-defined pathway to decrease appetite, hypothalamic proopiomelanocortin (POMC) neurons localized in the arcuate nucleus (Arc) project to MC4R neurons in the paraventricular nuclei (PVN) to release the natural MC4R agonist α-melanocyte-stimulating hormone (α-MSH). Arc neurons also project excitatory glutamatergic fibers to the MC4R neurons in the PVN for a fast synaptic transmission to regulate a satiety pathway potentiated by α-MSH. By using super-resolution microscopy, we found that in hypothalamic neurons in a primary culture, postsynaptic density protein 95 (PSD95) colocalizes with GluN1, a subunit of the ionotropic N-methyl-D-aspartate receptor (NMDAR). Thus, hypothalamic neurons form excitatory postsynaptic specializations. To study the MC4R distribution at these sites, tagged HA-MC4R under the synapsin promoter was expressed in neurons by adeno-associated virus (AAV) gene transduction. HA-MC4R immunofluorescence peaked at the center and in proximity to the PSD95- and NMDAR-expressing sites. These data provide morphological evidence that MC4R localizes together with glutamate receptors at postsynaptic and peri-postsynaptic sites. Full article
(This article belongs to the Special Issue Advances in Neurogenesis: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Hypothalamic neurons in primary culture express GluN1 localized at postsynaptic sites. (<b>A</b>) An orthogonal projection of six confocal images of a primary neuron expressing endogenous GluN1 and PSD95 visualized with the rabbit anti-GluN1 (AGC-001) and anti-PSD95 K28/43 antibody, respectively. The white rectangles indicate the inset regions shown at higher magnification in (<b>C</b>), and the yellow asterisk indicates the neuron of interest. (<b>B</b>) Left and right panels: Single confocal images from different positions of a Z-stack of 6 µm thickness showing the punctate distribution of PSD95 (left) and the negative staining of the nucleus (right) in the cell soma of the neuron shown in (<b>A</b>). (<b>C</b>) Enlarged images of the inset regions in (<b>A</b>). Yellow arrowheads indicate branching points. (<b>D</b>) A representative SR image of a point of PSD95 fluorescence with a representative yellow dashed line was used to determine the average diameter of PSD95 points. The graph on the right illustrates the average +/− SEM fluorescence (gray value) of PSD95 along the line, determined by measuring 180 spots of PSD95 fluorescence from 6 neurons. (<b>E</b>,<b>F</b>) SR images of a neurite expressing GluN1 and PSD95. The white arrow indicates a spot of PSD95 and GluN1 colocalization, and the green/magenta arrows indicate spots where PSD95 and GluN1 are adjacent. The regions indicated by the arrows are shown enlarged in (<b>F</b>). Yellow lines indicate those drawn for line segment analyses, with corresponding graphs displayed as means +/− SEM. (<b>G</b>) The distance (nm) between GluN1 and PSD95 fluorescence peaks was determined by the line segment analyses from 6 neurons from 2 independent experiments (n = 46 colocalizing points, n = 17 adjacent points), displayed as means +/− SD. The green dotted line is placed at the 216 nm cutoff that discriminates spots at which GluN1 and PSD95 colocalize (distance of fluorescence peaks &lt; 216 nm) and those where GluN1 and PSD95 are adjacent (distance between fluorescence peaks &gt; 216 nm and &lt;450 nm), as indicated by the models.</p>
Full article ">Figure 2
<p>Murine endogenous MC4R and human epitope-tagged HA-MC4R have a similar distribution in primary hypothalamic neurons. (<b>A</b>) Confocal images of Neuro2A cells that were transiently transfected with mouse MC4R and were incubated with MC4RIIIab that was either pre-adsorbed with antigenic peptide (left) or not pre-adsorbed (right). Scale bar: 100 µm. (<b>B</b>) Top panel: Illustration of untagged mouse MC4R and the MC4RIIIab antibody binding to the third intracellular loop of the receptor. Bottom panel: Confocal image of a Neuro2A cell transiently transfected with mouse MC4R and immune-stained with MC4RIIIab. (<b>C</b>) Top panel: Illustration of human MC4R containing the HA tag at the N-terminus and the HA antibody binding to the tag. Bottom panel: Confocal image of a Neuro2A cell transiently transfected with HA-MC4R and immune-stained with the HA antibody. (<b>D</b>) Confocal images of primary neurons incubated with (right) or without (left) MC4RIIIab and visualized by tyramide amplification. Top panel: 647 nm fluorescence only; bottom panel: 647 nm fluorescence merged with the brightfield. The region indicated by the red arrow is enlarged to indicate individual spots of endogenous murine MC4R fluorescence (yellow arrowhead). (<b>E</b>) Diagram of the AAV2-Syn-HA-MC4R plasmid. (<b>F</b>) Gel electrophoresis of the AAV2-Syn-HA-MC4R plasmid, digested with HindIII and BamHI, showing the vector backbone migrating at the 4 kb band of DNA ladder, and the HA-MC4R insert, migrating at the 1 kb band of DNA ladder. (<b>G</b>) Live cell confocal imaging of a neuron transduced with AAV2-Syn-HA-MC4R and incubated with mouse anti-HA Tag (F-7) conjugated to Alexa Fluor 647 for 30 min before imaging. The region indicated by the red arrow is enlarged to indicate individual spots of exogenous HA-MC4R fluorescence (yellow arrowhead). (<b>H</b>) Line segment analyses of n = 49 spots of endogenous MC4R immunofluorescence and n = 42 spots for HA-MC4R immunofluorescence, displayed as means +/− SEM. (<b>I</b>) CAMPER neurons transduced with AAV2-synapsin-Cre expressing cytosolic TEpacVV, appearing as diffuse fluorescence in the cell body and (green arrow) in the primary process is shown enlarged in the right panel. A secondary process is indicated by a red arrow.</p>
Full article ">Figure 3
<p>HA-MC4R expressed at neuronal processes colocalizes with PSD95. Confocal images of neurons treated without AAV2 (left panel) and neurons transduced with AAV2 Syn-HA-MC4R (middle and right panels). Neurons were immunostained with rabbit anti-HA C29F4 antibodies against the HA tag and mouse anti-PSD95 MA1-046 against PSD95. The yellow asterisks indicate the cell body, the white rectangle indicates the region shown enlarged in the right panels, and the white arrows indicate points of colocalization.</p>
Full article ">Figure 4
<p>HA-MC4R localizes at the postsynaptic and peri-postsynaptic sites. (<b>A</b>) Confocal images (20 µm scale bar) and SR images (5 µm scale bar) of a HA-MC4R-expressing neuron, immunostained with rabbit anti-HA C29F4 antibodies and mouse anti-PSD95 MA1-046. The white rectangle indicates the enlarged region. The white arrow indicates a point of colocalization, and the green/magenta arrows indicate a spot where PSD95 and HA-MC4R are adjacent. (<b>B</b>) Enlarged images of the neurite are shown above, where HA-MC4R and PSD95 colocalize (white arrow). (<b>C</b>) Enlarged images of the neurite above where HA-MC4R and PSD95 are adjacent (magenta and green arrows). (<b>B</b>,<b>C</b>) The yellow dashed lines indicate those drawn for the segment analyses, with the corresponding graphs displayed as means +/− SEM on the right. (<b>D</b>) The distance (nm) between HA-MC4R and PSD95 fluorescence peaks was determined by line segment analyses from 5 neurons from 3 independent experiments (n = 52 colocalizing points, n = 21 adjacent points) displayed as the means +/− SD. The green dotted line in the graph is as in <a href="#cells-13-01235-f001" class="html-fig">Figure 1</a>. Models indicate the range of distances between colocalizing and adjacent peaks of HA-MC4R and PSD95 fluorescence, respectively.</p>
Full article ">Figure 4 Cont.
<p>HA-MC4R localizes at the postsynaptic and peri-postsynaptic sites. (<b>A</b>) Confocal images (20 µm scale bar) and SR images (5 µm scale bar) of a HA-MC4R-expressing neuron, immunostained with rabbit anti-HA C29F4 antibodies and mouse anti-PSD95 MA1-046. The white rectangle indicates the enlarged region. The white arrow indicates a point of colocalization, and the green/magenta arrows indicate a spot where PSD95 and HA-MC4R are adjacent. (<b>B</b>) Enlarged images of the neurite are shown above, where HA-MC4R and PSD95 colocalize (white arrow). (<b>C</b>) Enlarged images of the neurite above where HA-MC4R and PSD95 are adjacent (magenta and green arrows). (<b>B</b>,<b>C</b>) The yellow dashed lines indicate those drawn for the segment analyses, with the corresponding graphs displayed as means +/− SEM on the right. (<b>D</b>) The distance (nm) between HA-MC4R and PSD95 fluorescence peaks was determined by line segment analyses from 5 neurons from 3 independent experiments (n = 52 colocalizing points, n = 21 adjacent points) displayed as the means +/− SD. The green dotted line in the graph is as in <a href="#cells-13-01235-f001" class="html-fig">Figure 1</a>. Models indicate the range of distances between colocalizing and adjacent peaks of HA-MC4R and PSD95 fluorescence, respectively.</p>
Full article ">Figure 5
<p>HA-MC4R localizes together with and in the proximity of GluN1. (<b>A</b>) Confocal images (30 µm scale bar) and SR images (5 µm scale bar) of an HA-MC4R-expressing neuron, immunostained with antibodies with mouse anti-HA Tag (F-7) and rabbit anti-GluN1 (AGC-001) to visualize endogenous GluN1. The white rectangle indicates the enlarged region. The white arrow indicates a point of colocalization, and the cyan/magenta arrows indicate a site where GluN1 and HA-MC4R are adjacent. (<b>B</b>) Enlarged images of the neurite above where HA-MC4R and GluN1 colocalize (white arrow). (<b>C</b>) Enlarged images of the neurite above where HA-MC4R and GluN1 are adjacent (magenta and cyan arrows). (<b>B</b>,<b>C</b>) The yellow dashed lines indicate those drawn for the segment analyses, with the corresponding graphs displayed as means +/− SEM on the right. (<b>D</b>) The distance (nm) between HA-MC4R and GluN1 fluorescence peaks was determined by line segment analyses from 3 neurons from 2 independent experiments (n = 34 colocalizing points, 21 adjacent points). The green dotted line in the graph is as in <a href="#cells-13-01235-f001" class="html-fig">Figure 1</a>. Models indicate the range of distances between colocalizing and adjacent peaks of HA-MC4R and GluN1 fluorescence, respectively.</p>
Full article ">Figure 5 Cont.
<p>HA-MC4R localizes together with and in the proximity of GluN1. (<b>A</b>) Confocal images (30 µm scale bar) and SR images (5 µm scale bar) of an HA-MC4R-expressing neuron, immunostained with antibodies with mouse anti-HA Tag (F-7) and rabbit anti-GluN1 (AGC-001) to visualize endogenous GluN1. The white rectangle indicates the enlarged region. The white arrow indicates a point of colocalization, and the cyan/magenta arrows indicate a site where GluN1 and HA-MC4R are adjacent. (<b>B</b>) Enlarged images of the neurite above where HA-MC4R and GluN1 colocalize (white arrow). (<b>C</b>) Enlarged images of the neurite above where HA-MC4R and GluN1 are adjacent (magenta and cyan arrows). (<b>B</b>,<b>C</b>) The yellow dashed lines indicate those drawn for the segment analyses, with the corresponding graphs displayed as means +/− SEM on the right. (<b>D</b>) The distance (nm) between HA-MC4R and GluN1 fluorescence peaks was determined by line segment analyses from 3 neurons from 2 independent experiments (n = 34 colocalizing points, 21 adjacent points). The green dotted line in the graph is as in <a href="#cells-13-01235-f001" class="html-fig">Figure 1</a>. Models indicate the range of distances between colocalizing and adjacent peaks of HA-MC4R and GluN1 fluorescence, respectively.</p>
Full article ">Figure 6
<p>HA-MC4R localizes together with and in the proximity of GluN2A. (<b>A</b>) Confocal images (20 µm scale bar) and SR images (5 µm scale bar) of an HA-MC4R-expressing neuron, immunostained with mouse anti-HA Tag (F-7) and anti-NR2A (N327/95) to visualize endogenous GluN2A. The white rectangle indicates the relative region of the enlarged images in the panel below. The white arrow indicates a point of colocalization, and the cyan/magenta arrows indicate a site where GluN2A and HA-MC4R are adjacent. (<b>B</b>) Enlarged images of the neurite above where HA-MC4R and GluN2A colocalize (white arrow) and where they are adjacent (cyan/magenta arrow). The yellow dashed lines indicate those drawn for the segment analyses, with the corresponding graphs displayed as means +/− SEM on the right. (<b>C</b>) The distance (nm) between HA-MC4R and GluN2A fluorescence peaks was determined by line segment analyses from 5 neurons from 2 independent experiments (n = 39 colocalizing points, n = 19 adjacent points). The green dotted line in the graph is as in <a href="#cells-13-01235-f001" class="html-fig">Figure 1</a>. Models indicate the range of distances between colocalizing and adjacent peaks of HA-MC4R and GluN2A fluorescence, respectively.</p>
Full article ">
34 pages, 2835 KiB  
Review
Navigating the CRISPR/Cas Landscape for Enhanced Diagnosis and Treatment of Wilson’s Disease
by Woong Choi, Seongkwang Cha and Kyoungmi Kim
Cells 2024, 13(14), 1214; https://doi.org/10.3390/cells13141214 - 18 Jul 2024
Viewed by 1048
Abstract
The clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein (Cas) system continues to evolve, thereby enabling more precise detection and repair of mutagenesis. The development of CRISPR/Cas-based diagnosis holds promise for high-throughput, cost-effective, and portable nucleic acid screening and genetic disease diagnosis. In [...] Read more.
The clustered regularly interspaced short palindromic repeats (CRISPR)/CRISPR-associated protein (Cas) system continues to evolve, thereby enabling more precise detection and repair of mutagenesis. The development of CRISPR/Cas-based diagnosis holds promise for high-throughput, cost-effective, and portable nucleic acid screening and genetic disease diagnosis. In addition, advancements in transportation strategies such as adeno-associated virus (AAV), lentiviral vectors, nanoparticles, and virus-like vectors (VLPs) offer synergistic insights for gene therapeutics in vivo. Wilson’s disease (WD), a copper metabolism disorder, is primarily caused by mutations in the ATPase copper transporting beta (ATP7B) gene. The condition is associated with the accumulation of copper in the body, leading to irreversible damage to various organs, including the liver, nervous system, kidneys, and eyes. However, the heterogeneous nature and individualized presentation of physical and neurological symptoms in WD patients pose significant challenges to accurate diagnosis. Furthermore, patients must consume copper-chelating medication throughout their lifetime. Herein, we provide a detailed description of WD and review the application of novel CRISPR-based strategies for its diagnosis and treatment, along with the challenges that need to be overcome. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Copper uptake and homeostasis pathway in the human body. Copper is absorbed mainly in the duodenum and partly in the stomach via proteins such as high-affinity copper uptake protein 1 (CTR1) and divalent metal-ion transporter 1 (DMT1). Before transportation, cupric ions (Cu<sup>2+</sup>) are reduced to cuprous ions (Cu<sup>+</sup>) with the assistance of six transmembrane epithelial antigen of the prostate (STEAP) reductase, ascorbate, or duodenal cytochrome b (DCYTB). Upon absorption by and entry into an enterocyte, copper is transported by key copper chaperones, predominantly cytochrome c oxidase copper chaperone (COX17), superoxide dismutase (CCS), and antioxidant-1 (ATOX1). COX17 transfers copper to mitochondrial proteins SCO1 and SCO2, which are crucial for cytochrome c oxidase biogenesis. CCS delivers copper to superoxide dismutase (SOD1), which is a major antioxidant maintaining cytoplasmic ROS homeostasis. Copper is transferred by ATOX1 to the trans-Golgi network (TGN), where it interacts with copper-transporting P-type ATPase 1 (ATP7A) and P-type ATPase 2 (ATP7B). The ATP7A facilitates the delivery of Cu<sup>+</sup> to copper-dependent enzymes such as peptidyl-α-monooxygenase, tyrosinase, and lysyl oxidase, while also exporting excess copper into the extracellular fluid. In contrast, the ATP7B stores Cu<sup>+</sup> in the vesicles. Once exported, Cu<sup>+</sup> is oxidized to Cu<sup>2+</sup> by utilizing dissolved oxygen in the blood, where it binds to proteins such as albumin, histidine, etc., and then is transported via the portal vein to the liver. Hepatic uptake is facilitated by CTR1 post-reduction by copper reductase. Copper is transported within hepatocytes by key copper chaperones such as COX17, CCS, and ATOX1, likely entering via enterocytes. However, hepatocytes exhibit a high expression level of ATP7B and a low expression level of ATP7A. Consequently, ATP7B serves as the primary transporter in hepatocytes. The illustration was created using <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>The structure of ATP7B. The A-domain has a TGEA motif, the P-domain has a DKTGT motif, and the N-domain has a TGDN and a SEPHL motif. There are six metal-binding domains (MBDs) containing the core conserved sequence GMXCXXC at the N-terminus located intracellularly, each of which can bind to one Cu<sup>+</sup> ion. Among the transmembrane domains (TMDs), TMD6 has a CPC motif, which makes it characteristic of heavy metal-transporting P-type ATPase. The illustration was created using <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 3
<p>Prevalent mutation variants on the <span class="html-italic">ATP7B</span> gene that induce Wilson’s disease (WD). The most common mutations in the <span class="html-italic">ATP7B</span> genomic DNA causing WD are highlighted in blue (p.R778L/c.2333G&gt;T; Asia) and yellow (p.H1069Q/c.3207C&gt;A; Europe and North America). No predominant mutation was found to be associated with South America [<a href="#B28-cells-13-01214" class="html-bibr">28</a>]. The p.H1069Q mutation occurs in the SEHPL motif in the N-domain, while the p.R778L mutation affects transmembrane transportation activity. The red dots indicate the locations of the mutation variants shown in <a href="#cells-13-01214-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Schematics of the mechanisms of the CRISPR/Cas system types. The CRISPR/Cas system is classified into two classes based on the effector molecules involved. Class I is characterized by multi-unit effectors, while Class II exhibits a single effector. Each class is further divided into three types distinguished by their catalytic domain and target nucleotide specificity. Type I, represented by Cas3, forms a cascade complex with Cas6, Cas7, Cas5, Cas11, and Cas8. Upon crRNA binding, Cas3 generates a single-strand nick on the unwinding target DNA. Utilizing the Cas10-Cmr/Cms complex, the Type III system degrades nascent RNA before enzymatically cleaving complementary DNA. Type IV comprises csf1, csf2, csf3, an endoribonuclease (csf5), and a helicase (dinG), although its mechanism is not completely understood. The most widely used Type II system employs a single effector, Cas9, which induces a double-strand break (DSB) to target the double-stranded DNA (dsDNA) alongside the guide RNA (crRNA and tracrRNA). Types V and VI possess both target and collateral cleavage activities. Cas12 functions in conjunction with crRNA to generate staggered dsDNA breaks and nonspecifically cleaves the single-stranded DNA (ssDNA) collateral present in the vicinity of a dsDNA target. Cas13, in conjunction with crRNA, frequently cleaves single-stranded RNA (ssRNA) at multiple sites and exhibits nonspecific collateral ssRNA cleavage in the presence of the ssRNA target. The illustration was created using <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 5
<p>CRISPR/Cas system-based diagnosis of <span class="html-italic">ATP7B</span> mutations. (<b>A</b>) Diagnosing <span class="html-italic">ATP7B</span> mutations in WD patients depends on the specific type of mutation in the <span class="html-italic">ATP7B</span> gene and the type of nucleic acid (genomic DNA (gDNA) or mRNA) extracted from the patient. The red star represents the mutation point. The appropriate Cas protein must be selected and applied with these factors in mind. (<b>B</b>) A simple example of how the CRISPR/Cas system could potentially be used for diagnosing four distinct <span class="html-italic">ATP7B</span> mutations. The yellow highlight represents PAM sequences for SpCas9 (5′-NGG) and AsCas12a (5′-TTTN). The blue highlight represents the crRNA sequence in sgRNA to recognize the target sequence. The asterisk on the amino acid mutation means the stop codon. (<b>C</b>) The figure shows the locations in the genomic DNA (gDNA) and cDNA where the c.2072G&gt;T (p.G691V) and c.2108G&gt;A (p.C703Y) mutations can occur. The yellow highlight represents PAM sequences for SpCas9. The blue highlight represents potential sgRNA candidates for SpCas9, which are in proximity to the p.G691V and p.C703Y mutations. (<b>D</b>) The RGEN-RFLP method distinguishes between WT and mutant sequences by cutting the WT using sgRNA that matches the WT sequence. The upper figure shows possible scenarios after the treatment of SpCas9 with sgRNA. The blue highlight represents potential sgRNA candidates for SpCas9. The yellow marker indicates the location of the PAM sequence. The red letter on the DNA sequence represents the mutation site that induces p.R778L of <span class="html-italic">ATP7B</span>. The figure below shows the expected results of the RGEN-RFLP method resolved by agarose gel electrophoresis.</p>
Full article ">Figure 6
<p>Schematics of the mechanism of four different CRISPR/Cas system-based treatment strategies. (<b>A</b>) To induce a homology-directed repair (HDR) event mediated by the Cas9 protein, it is necessary to provide donor DNA, which should include the wild-type sequence of the <span class="html-italic">ATP7B</span> gene. The Cas9 nuclease then induces a DSB in the vicinity of the mutation on the <span class="html-italic">ATP7B</span> gene. It is then necessary to select a suitable repair pathway for the treated cell. (<b>B</b>) Adenosine deaminase within the adenine base editor (ABE) system catalyzes the conversion of adenine (A) to inosine (I) on the strand not bound by the sgRNA. The nCas9 enzyme induces a nick in the DNA, initiating the DNA repair mechanism. Subsequently, during replication of the target DNA, a conversion of a thymine (T)/A base pair to a cytosine (C)/guanine (G) base pair is accomplished. (<b>C</b>) The cytosine base editing (CBE) is similar to that of ABE. In this case, a G/C-to-A/T base pair conversion is achieved via a uracil (U)-containing intermediate. Uracil glycosylase inhibitor (UGI) provides stability to the U/G base pair. (<b>D</b>) Unlike base editing, the prime-editing system utilizes a specific nCas9 variant with the H840A mutation, which silences the HNH domain. Reverse transcriptase (RT) then synthesizes a new DNA sequence from the end of the truncated strand using prime-editing guide RNA (pegRNA) containing both the template sequence and editing sequence. The illustration was created using <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">
16 pages, 2615 KiB  
Article
Viral Vector Based Immunotherapy for Peanut Allergy
by Miguel Gonzalez-Visiedo, Roland W. Herzog, Maite Munoz-Melero, Sophia A. Blessinger, Joan M. Cook-Mills, Henry Daniell and David M. Markusic
Viruses 2024, 16(7), 1125; https://doi.org/10.3390/v16071125 - 13 Jul 2024
Viewed by 1048
Abstract
Food allergy (FA) is estimated to impact up to 10% of the population and is a growing health concern. FA results from a failure in the mucosal immune system to establish or maintain immunological tolerance to innocuous dietary antigens, IgE production, and the [...] Read more.
Food allergy (FA) is estimated to impact up to 10% of the population and is a growing health concern. FA results from a failure in the mucosal immune system to establish or maintain immunological tolerance to innocuous dietary antigens, IgE production, and the release of histamine and other mediators upon exposure to a food allergen. Of the different FAs, peanut allergy has the highest incidence of severe allergic responses, including systemic anaphylaxis. Despite the recent FDA approval of peanut oral immunotherapy and other investigational immunotherapies, a loss of protection following cessation of therapy can occur, suggesting that these therapies do not address the underlying immune response driving FA. Our lab has shown that liver-directed gene therapy with an adeno-associated virus (AAV) vector induces transgene product-specific regulatory T cells (Tregs), eradicates pre-existing pathogenic antibodies, and protects against anaphylaxis in several models, including ovalbumin induced FA. In an epicutaneous peanut allergy mouse model, the hepatic AAV co-expression of four peanut antigens Ara h1, Ara h2, Ara h3, and Ara h6 together or the single expression of Ara h3 prevented the development of a peanut allergy. Since FA patients show a reduction in Treg numbers and/or function, we believe our approach may address this unmet need. Full article
(This article belongs to the Special Issue Virology and Immunology of Gene Therapy)
Show Figures

Figure 1

Figure 1
<p>Characterization of in-house generated peanut extract. (<b>A</b>) Raw values of relative Ara h1, Ara h2, Ara h3, and Ara h6 levels measured by Ara h specific ELISA (Indoor Biotechnologies) in peanut extracts from peanut flour and dry roasted peanuts at different pHs (7.2 and 8.5) and a commercial peanut extract (Stallergenes Greer). (<b>B</b>) Relative abundance of Ara h1, 2, 3, and 6 in peanut extracts obtained from different sources. (<b>C</b>) SDS-PAGE of two different peanut protein extracts and purified Ara h1, 2, 3, and 6 peanut proteins (Indoor Biotechnologies).</p>
Full article ">Figure 2
<p>Validation of peanut allergy sensitization in the FT<sup>+/−</sup> model and assessment of allergic responses following two separate challenges. (<b>A</b>) Experimental timeline followed. (<b>B</b>) Changes in core body temperature from baseline to 30 min post challenge. (<b>C</b>) Symptom score (see <a href="#viruses-16-01125-t001" class="html-table">Table 1</a> for definitions) following challenge with peanut extract after sensitization and four weeks later. (<b>D</b>) Peanut-specific levels of IgE and (<b>E</b>) IgG1 measured after peanut sensitization and four weeks later. Data are presented as single data points and means ± standard deviation. Statistical testing was conducting using unpaired T-test for all the data sets excluding panel (<b>C</b>), which was performed using the Mann–Whitney test. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &gt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>AAV Ara h cocktail prevents peanut sensitization. (<b>A</b>) Experimental timeline followed. (<b>B</b>) Changes in core body temperature and (<b>C</b>) symptom score (see <a href="#viruses-16-01125-t001" class="html-table">Table 1</a> for definitions) following challenge with peanut extract. (<b>D</b>) Peanut-specific levels of IgE and (<b>E</b>) IgG1. (<b>F</b>) Ara h vector genome (vg) copy numbers and (<b>G</b>) Ara h mRNA expression in liver samples. (<b>H</b>) Expression of Ara h proteins in liver and (<b>I</b>) plasma. Data are presented as single data points and means ± standard deviation. Statistical testing was conducting using unpaired T-test for panels (<b>B</b>–<b>E</b>), one-way ANOVA statistic test was used for panels (<b>F–H</b>) and Kruskal–Wallis test was used for panel (<b>I</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Single AAV-Ara h vectors are less effective for PE prophylaxis. (<b>A</b>) Experimental timeline followed. (<b>B</b>) Changes in core body temperature and (<b>C</b>) symptom score (refer to symptom score table here) following challenge with peanut extract. (<b>D</b>) Peanut-specific levels of IgE and (<b>E</b>) IgG1. (<b>F</b>) Representation of Ara h VG copy numbers and (<b>G</b>) Ara h mRNA expression in liver samples. (<b>H</b>) Expression of Ara h proteins in liver and (<b>I</b>) plasma. Data are presented as single data points and means ± standard deviation. Statistical testing was conducting using one-way ANOVA statistic test for panels (<b>B</b>,<b>D</b>–<b>H</b>), and Kruskal–Wallis was used for panels (<b>C</b>,<b>I</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt;0.001 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Increased dose of a single AAV-Ara h3 vector prophylactic effect was comparable to the AAV cocktail. (<b>A</b>) Experimental timeline followed. (<b>B</b>) Changes in core body temperature and (<b>C</b>) symptom score (refer to symptom score table here) following challenge with peanut extract. (<b>D</b>) Peanut-specific levels of IgE and (<b>E</b>) IgG1. (<b>F</b>) Representation of Ara h VG copy numbers and (<b>G</b>) Ara h mRNA expression in liver samples. (<b>H</b>) Expression of Ara h proteins in liver and (<b>I</b>) plasma. Data are presented as single data points and means ± standard deviation. Statistical testing was conducting using one-way ANOVA statistic test for panels (<b>B</b>,<b>D</b>,<b>F</b>,<b>G</b>), and Kruskal–Wallis test was used for panels (<b>C</b>,<b>E</b>,<b>H</b>,<b>I</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
Back to TopTop