[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (8,600)

Search Parameters:
Keywords = interleukin-8

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
31 pages, 5937 KiB  
Article
Oncolytic Tanapoxvirus Variants Expressing mIL-2 and mCCL-2 Regress Human Pancreatic Cancer Xenografts in Nude Mice
by Scott D. Haller and Karim Essani
Biomedicines 2024, 12(8), 1834; https://doi.org/10.3390/biomedicines12081834 - 12 Aug 2024
Abstract
Pancreatic ductal adenocarcinoma (PDAC) is the fifth leading cause of cancer-related death and presents the lowest 5-year survival rate of any form of cancer in the US. Only 20% of PDAC patients are suitable for surgical resection and adjuvant chemotherapy, which remains the [...] Read more.
Pancreatic ductal adenocarcinoma (PDAC) is the fifth leading cause of cancer-related death and presents the lowest 5-year survival rate of any form of cancer in the US. Only 20% of PDAC patients are suitable for surgical resection and adjuvant chemotherapy, which remains the only curative treatment. Chemotherapeutic and gene therapy treatments are associated with adverse effects and lack specificity/efficacy. In this study, we assess the oncolytic potential of immuno-oncolytic tanapoxvirus (TPV) recombinants expressing mouse monocyte chemoattractant protein (mMCP-1 or mCCL2) and mouse interleukin (mIL)-2 in human pancreatic BxPc-3 cells using immunocompromised and CD-3+ T-cell-reconstituted mice. Intratumoral treatment with TPV/∆66R/mCCL2 and TPV/∆66R/mIL-2 resulted in a regression in BxPc-3 xenograft volume compared to control in immunocompromised mice; mCCL-2 expressing TPV OV resulted in a significant difference from control at p < 0.05. Histological analysis of immunocompromised mice treated with TPV/∆66R/mCCL2 or TPV/∆66R/mIL-2 demonstrated multiple biomarkers indicative of increased severity of chronic, active inflammation compared to controls. In conclusion, TPV recombinants expressing mCCL2 and mIL-2 demonstrated a therapeutic effect via regression in BxPc-3 tumor xenografts. Considering the enhanced oncolytic potency of TPV recombinants demonstrated against PDAC in this study, further investigation as an alternative or combination treatment option for human PDAC may be warranted. Full article
Show Figures

Figure 1

Figure 1
<p>Immunocompromised group mean tumor volume. BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immunocompromised female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Total volume measurement was conducted every three days ((length) × (width) × (height) × (П/6)). When tumors reached approximately 200 mm<sup>3</sup>, approximately 27 days post inoculation, animals were randomly assigned to study treatment groups based on tumor volume to achieve an equivalent average group tumor volume as much as possible. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal. Vehicle control average tumor volume is shown in each plot as x symbol plot line (blue) and used as a comparator for each TPV recombinant experimental group: TPV/eGFP shown as hyphen symbol plot line (purple) (<b>A</b>), TPV/∆66R/m-IL-2/mCherry shown as triangle symbol plot line (green) (<b>B</b>), and TPV/∆66R/m-CCL-2/mCherry shown as circle symbol plot line (green) (<b>C</b>). Bars show standard error of the mean (±1 SEM) and, where applicable, an asterisk (*) indicates statistically significant regression in TPV/∆66R/m-CCL-2/mCherry recombinant from vehicle control (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Immune-reconstituted group mean tumor volume. BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immune-reconstituted female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Total volume measurement was conducted every three days ((length) × (width) × (height) × (П/6)). When tumors reached approximately 200 mm<sup>3</sup>, approximately 27 days post inoculation, animals were randomly assigned to study treatment groups based on tumor volume to achieve an equivalent average group tumor volume as much as possible. Each subject was treated with a single, intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal. Four days post vehicle control or virotherapy, each subject received a CD-3+ T-cell adoptive cell transfer (5 × 10<sup>6</sup> cells/subject) via tail vein administration. The vehicle control average tumor volume is shown in each plot as circle symbol plot line (green), and used as a comparator for each TPV recombinant experimental group: TPV/eGFP shown as hyphen symbol plot line (peach) (<b>A</b>), TPV/∆66R/m-IL-2/mCherry shown as square symbol (peach) (<b>B</b>), and TPV/∆66R/m-CCL-2/mCherry shown as square symbol (grey) (<b>C</b>). Bars show standard error of the mean (±1 SEM) and, where applicable, an asterisk (*) indicates a statistically significant greater mean group tumor volume of TPV/∆66R/m-CCL-2/mCherry recombinant from vehicle control (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Immunocompromised and immune-reconstituted vehicle control group mean tumor volume. BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immunocompromised and immune-reconstituted female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Total volume measurement was conducted every three days ((length) × (width) × (height) × (П/6)). When tumors reached approximately 200 mm<sup>3</sup>, approximately 27 days post inoculation, animals were randomly assigned to study treatment groups based on tumor volume to achieve an equivalent average group tumor volume as much as possible. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle. Four days post vehicle control, each immune-reconstituted subject received a CD-3+ T-cell adoptive cell transfer (5 × 10<sup>6</sup> cells/subject) via tail vein administration. Immunocompromised vehicle control average tumor volume is shown in the plot as x symbol plot line (blue) and immune-reconstituted, vehicle control average tumor volume is displayed as circle symbol (green) plot line. Bars show standard error of the mean (±1 SEM) and, where applicable, an asterisk (*) indicates a statistically significant difference in mean group tumor volume between the vehicle-treated control groups (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Immunocompromised group mean tumor [<sup>18</sup>F]-FDG PET/CT-derived standardized uptake value (SUV) and total lesion glycolysis (TLG). BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immunocompetent female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. Baseline [<sup>18</sup>F]-FDG PET/CT images (200 µCi/subject) prior to therapy were acquired 3 weeks post BxPc-3 cell inoculation, and post-virotherapy images were collected at 7–8 weeks post inoculation. Bars show standard error of the mean (±1 SEM); SUV results for each respective group are shown in plot (<b>A</b>) and TLG results are show in plot (<b>B</b>); inter- and intragroup statistical comparisons resulted in a lack of statistical significance for all comparisons.</p>
Full article ">Figure 5
<p>Immune-reconstituted group mean tumor [<sup>18</sup>F]-FDG PET/CT-derived standardized uptake value (SUV) and total lesion glycolysis (TLG). BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immunocompromised or immune-reconstituted female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. Baseline [<sup>18</sup>F]-FDG PET/CT images (200 µCi/subject) prior to virotherapy were acquired 3 weeks post BxPc-3 cell inoculation, and post-virotherapy images were collected at 7–8 weeks post inoculation. Bars show standard error of the mean (±1 SEM); SUV results for each respective group are shown in plot (<b>A</b>) and TLG results are show in plot (<b>B</b>); inter- and intragroup statistical comparisons resulted in a lack of statistical significance for all comparisons.</p>
Full article ">Figure 6
<p>Post-virotherapy [<sup>18</sup>F]-FDG PET/CT Images of BxPc-3 PDAC human tumor xenografts in immunocompromised and immune-reconstituted BALB-c nude mice. BxPc-3 human PDAC cells (5 × 10<sup>6</sup> cells/subject) were inoculated subcutaneously (SC) in the right flank region of immunocompromised or immune-reconstituted female BALB-c nude mice. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. [<sup>18</sup>F]-FDG PET/CT images (200 µCi/subject) were acquired 7–8 weeks post tumor cell inoculation. Select representative PET/CT images are as follows: immunocompromised vehicle control (<b>A</b>), immune-reconstituted vehicle control (<b>B</b>), TPV/eGFP immunocompromised (<b>C</b>), and TPV/eGFP immune-reconstituted (<b>D</b>). Panel (<b>E</b>) (TPV/eGFP immunocompromised subject in sagittal, AP, and transverse planes from left to right) demonstrates the multicompartmental composition of tumor mass observed in many study subjects; white arrows indicate tumor location.</p>
Full article ">Figure 7
<p>Immunocompromised group mean tumor [<sup>18</sup>F]-FLT PET/CT-derived standardized uptake value (SUV) and total lesion proliferation (TLP). BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immunocompetent female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. Baseline [<sup>18</sup>F]-FLT PET/CT images (200 µCi/subject) prior to virotherapy were acquired 3 weeks post BxPc-3 cell inoculation, and post-virotherapy images were collected at 7–8 weeks post inoculation. Bars show standard error of the mean (±1 SEM); SUV results for each respective group are shown in plot (<b>A</b>) and TLP results are show in panel (<b>B</b>). SUV and TLP inter- and intragroup statistical comparisons resulted in a lack of statistical significance for all comparisons except for immunocompromised vehicle-treated subject baseline to post-virotherapy results (<span class="html-italic">p</span> &lt; 0.05, indicated by asterisk (*)).</p>
Full article ">Figure 7 Cont.
<p>Immunocompromised group mean tumor [<sup>18</sup>F]-FLT PET/CT-derived standardized uptake value (SUV) and total lesion proliferation (TLP). BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immunocompetent female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. Baseline [<sup>18</sup>F]-FLT PET/CT images (200 µCi/subject) prior to virotherapy were acquired 3 weeks post BxPc-3 cell inoculation, and post-virotherapy images were collected at 7–8 weeks post inoculation. Bars show standard error of the mean (±1 SEM); SUV results for each respective group are shown in plot (<b>A</b>) and TLP results are show in panel (<b>B</b>). SUV and TLP inter- and intragroup statistical comparisons resulted in a lack of statistical significance for all comparisons except for immunocompromised vehicle-treated subject baseline to post-virotherapy results (<span class="html-italic">p</span> &lt; 0.05, indicated by asterisk (*)).</p>
Full article ">Figure 8
<p>Immune-reconstituted group mean tumor [<sup>18</sup>F]-FLT PET/CT-derived standardized uptake value (SUV) and total lesion proliferation (TLP). BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immune reconstituted female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. Baseline [<sup>18</sup>F]-FLT PET/CT images (200 µCi/subject) prior to virotherapy were acquired 3 weeks post BxPc-3 cell inoculation, and post-virotherapy images were collected at 7–8 weeks post inoculation. Bars show standard error of the mean (±1 SEM); SUV results for each respective group are shown in plot (<b>A</b>) and TLP results are show in plot (<b>B</b>). SUV inter- and intragroup statistical comparisons resulted in a lack of statistical significance for all comparisons. TLP inter- and intragroup statistical comparisons resulted in a lack of statistical significance for all comparisons within immune reconstituted subjects; TLP results for post-virotherapy vehicle and TPV/∆66R/mCCL-2-treated subjects were statistically different (<span class="html-italic">p</span> &lt; 0.05, indicated by asterisk (*)) from immunocompromised vehicle-treated control subjects.</p>
Full article ">Figure 8 Cont.
<p>Immune-reconstituted group mean tumor [<sup>18</sup>F]-FLT PET/CT-derived standardized uptake value (SUV) and total lesion proliferation (TLP). BxPc-3 human PDAC cells were inoculated subcutaneously (SC) in the right flank region of immune reconstituted female BALB-c nude mice. Each subject received a single inoculation of 5 × 10<sup>6</sup> cells/subject. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. Baseline [<sup>18</sup>F]-FLT PET/CT images (200 µCi/subject) prior to virotherapy were acquired 3 weeks post BxPc-3 cell inoculation, and post-virotherapy images were collected at 7–8 weeks post inoculation. Bars show standard error of the mean (±1 SEM); SUV results for each respective group are shown in plot (<b>A</b>) and TLP results are show in plot (<b>B</b>). SUV inter- and intragroup statistical comparisons resulted in a lack of statistical significance for all comparisons. TLP inter- and intragroup statistical comparisons resulted in a lack of statistical significance for all comparisons within immune reconstituted subjects; TLP results for post-virotherapy vehicle and TPV/∆66R/mCCL-2-treated subjects were statistically different (<span class="html-italic">p</span> &lt; 0.05, indicated by asterisk (*)) from immunocompromised vehicle-treated control subjects.</p>
Full article ">Figure 9
<p>Post-virotherapy [<sup>18</sup>F]-FLT PET/CT images of BxPc-3 PDAC human tumor xenografts in immunocompromised and immune-reconstituted BALB-c nude mice. BxPc-3 human PDAC cells (5 × 10<sup>6</sup> cells/subject) were inoculated subcutaneously (SC) in the right flank region of immunocompetent female BALB-c nude mice. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. [<sup>18</sup>F]-FLT PET/CT images (200 µCi/subject) were acquired 7–8 weeks post tumor cell inoculation. Select representative PET/CT images are as follows: immunocompromised vehicle control (<b>A</b>), immune-reconstituted vehicle control (<b>B</b>), TPV/eGFP immunocompromised (<b>C</b>), and TPV/eGFP immune reconstituted (<b>D</b>); white arrows indicate tumor location.</p>
Full article ">Figure 10
<p>[<sup>125</sup>I]-anti-eGFP or [<sup>125</sup>I]-anti-mCherry antibody SPECT/CT images of BxPc-3 PDAC human tumor xenografts in immunocompromised and immune-reconstituted BALB-c nude mice. BxPc-3 human PDAC cells (5 × 10<sup>6</sup> cells/subject) were inoculated subcutaneously (SC) in the right flank region of immunocompetent female BALB-c nude mice. Each subject was treated with a single intratumoral administration of 100 µL PBS formulation vehicle or 100X virus stock of each applicable recombinant virus to result in a virus mass dose of 5 × 10<sup>6</sup> plaque-forming units (pfu) per animal when tumors reached approximately 200 mm<sup>3</sup>. SPECT/CT images were acquired during weeks 7–8, 48 h post imaging agent administration (200 µCi/subject). Select representative SPECT/CT images are as follows: immunocompromised vehicle control (A), immune reconstituted vehicle control (B), TPV/eGFP immunocompromised (C), and TPV/eGFP immune reconstituted (D); white arrows indicate tumor location.</p>
Full article ">Figure 11
<p>Histopathology and immunohistochemistry photomicrograph images of BxPc-3 PDAC human tumor xenografts in immunocompromised and immune-reconstituted BALB-c nude mice. Select representative photomicrograph images as follows: (<b>A</b>) H&amp;E-stained, immunocompromised vehicle control with healthy adenocarcinoma; (<b>B</b>) H&amp;E-stained, immune-reconstituted vehicle control adenocarcinoma with areas of necrosis and inflammatory cell infiltrates (black arrow); (<b>C</b>) H&amp;E-stained, immunocompromised TPV/eGFP adenocarcinoma with areas of necrosis and inflammatory cell infiltrates (black arrow); (<b>D</b>) H&amp;E-stained, immunocompromised TPV/∆66R/mIL-2 adenocarcinoma with areas of necrosis and inflammatory cell infiltrates (black arrow); (<b>E</b>) H&amp;E-stained, immunocompromised TPV/∆66R/mCCL-2 adenocarcinoma with areas of necrosis and inflammatory cell infiltrates (black arrow); (<b>F</b>) caspase-stained, immunocompromised TPV/∆66R/mCCL-2 adenocarcinoma with areas of positive staining (black arrow); (<b>G</b>) CD-3-stained, immunocompromised TPV/∆66R/mCCL-2 adenocarcinoma with areas of positive staining (black arrow); (<b>H</b>) CD-4-stained, immunocompromised TPV/∆66R/mIL-2 adenocarcinoma with areas of positive staining (black arrow); (<b>I</b>) CD-68-stained, immunocompromised TPV/∆66R/mCCL-2 adenocarcinoma with areas of positive staining (black arrow); (<b>J</b>) GFP-stained, immunocompromised TPV/eGFP adenocarcinoma with areas of positive staining demonstrating viral replication and transgene expression (black arrow); (<b>K</b>) mCherry-stained, immunocompromised TPV/∆66R/mCCL-2 adenocarcinoma with areas of positive staining demonstrating viral replication and transgene expression (black arrow); and (<b>L</b>) mCherry-stained, immunocompromised TPV/∆66R/mIL-2 adenocarcinoma with areas of positive staining demonstrating viral replication and transgene expression. All bars = 200 µm, except for image E bar = 400 µm.</p>
Full article ">
8 pages, 9299 KiB  
Case Report
Interleukin-36 Is Highly Expressed in Skin Biopsies from Two Patients with Netherton Syndrome
by Johannes Pawlowski, Tatsiana Pukhalskaya, Kelly Cordoro, Marina Kristy Ibraheim and Jeffrey P. North
Dermatopathology 2024, 11(3), 230-237; https://doi.org/10.3390/dermatopathology11030024 (registering DOI) - 12 Aug 2024
Abstract
Netherton syndrome (NS) is a rare autosomal recessive disorder that occurs due to a loss-of-function mutation in SPINK5; this loss results in significant inflammation, as well as perturbations of the skin barrier’s integrity and functionality. While it is unclear which inflammatory pathways contribute [...] Read more.
Netherton syndrome (NS) is a rare autosomal recessive disorder that occurs due to a loss-of-function mutation in SPINK5; this loss results in significant inflammation, as well as perturbations of the skin barrier’s integrity and functionality. While it is unclear which inflammatory pathways contribute to the development of NS, recent studies have demonstrated the expression of interleukin (IL)-17/IL-36, as well as several Th2 cytokines. Consequently, immunohistochemistry (IHC) with IL-36 may serve as a potential tool for aiding the histopathological diagnosis of this condition. In this case series, we present two cases of NS and capture their immunostaining pattern with IL-36. Both cases demonstrated robust expression of IL-36. This finding bolsters the hypothesis that NS is partially driven by Th17 activation and suggests the potential utility of IL-36 IHC as part of the workup for this rare and diagnostically elusive entity. LEKTI IHC was negative in one biopsy, revealing a limitation of this stain in diagnosing NS. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>–<b>C</b>) Oval and annular erythematous macules, patches, and plaques with slight scale affecting the trunk and extremities.</p>
Full article ">Figure 2
<p>(<b>A</b>,<b>E</b>) Hematoxylin and eosin (H&amp;E)-stained sections demonstrating psoriasiform epidermal hyperplasia, parakeratosis with variable hypergranulosis. Mildly dilated blood vessels and a superficial perivascular lymphocytic infiltrate were also present. (H&amp;E 40×); (<b>B</b>,<b>F</b>) features at high magnification (H&amp;E 100×); (<b>C</b>,<b>G</b>) IL-36 staining grade 4 (IL-36 40×); (<b>D</b>) LEKTI strongly staining the granular layer (LEKTI 40×).</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>C</b>) Numerous erythematous, serpiginous plaques with a double edge scale on the trunk and extremities.</p>
Full article ">Figure 4
<p>(<b>A</b>) H&amp;E demonstrating regular acanthosis, broad parakeratosis with focal neutrophilic inflammation, mild spongiosis, and superficial perivascular inflammation comprising lymphocytes. (H&amp;E 40×); (<b>B</b>–<b>D</b>) features at higher magnification ((<b>B</b>) H&amp;E 100×, (<b>C</b>,<b>D</b>): H&amp;E 400×); (<b>E</b>) IL-36 staining grade 4 (IL-36 40×).</p>
Full article ">
14 pages, 16172 KiB  
Article
Topical Application of Cha-Koji, Green Tea Leaves Fermented with Aspergillus Luchuensis var Kawachii Kitahara, Promotes Acute Cutaneous Wound Healing in Mice
by Yasuhiro Katahira, Jukito Sonoda, Miu Yamagishi, Eri Horio, Natsuki Yamaguchi, Hideaki Hasegawa, Izuru Mizoguchi and Takayuki Yoshimoto
Sci. Pharm. 2024, 92(3), 44; https://doi.org/10.3390/scipharm92030044 (registering DOI) - 12 Aug 2024
Abstract
“Koji” is one of the most well-known probiotic microorganisms in Japan that contribute to the maintenance of human health. Although the beneficial effects of some probiotics on ulcer healing have been demonstrated, there have been no reports on the wound healing effects of [...] Read more.
“Koji” is one of the most well-known probiotic microorganisms in Japan that contribute to the maintenance of human health. Although the beneficial effects of some probiotics on ulcer healing have been demonstrated, there have been no reports on the wound healing effects of koji to date. In the present study, we investigated the effects of “cha-koji”, green tea leaves fermented with Aspergillus luchuensis, on cutaneous wound healing, using a linear incision wound mouse model. Topical application of autoclave-sterilized cha-koji suspension on the dorsal incision wound area healed the wound significantly faster and, notably, with less scarring than did the green tea or the control distilled water treatment. Further in vitro experiments revealed that the accelerated effects of cha-koji could be attributed to its increased anti-bacterial activity, enhanced epidermal cell proliferation and migration, augmented expression of the anti-inflammatory cytokine transforming growth factor-β1, reduced expression of inflammatory cytokine interleukin-6 in macrophages, and decreased endoplasmic reticulum stress. In addition, we conducted a skin sensitizing potential test, which revealed that cha-koji had no adverse effects that posed a sensitizing risk. Thus, cha-koji may have a potent therapeutic effect on cutaneous wound healing, opening up a new avenue for its clinical application as a medical aid. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Autoclave-sterilized cha-koji suspension promotes cutaneous wound healing. A cutaneous incision wound was made on the dorsal skin of the mice; coated with gauze soaked in green tea, cha-koji suspension, or distilled water; and covered with a transparent film dressing, Tegaderm, daily until day 13 (<b>A</b>,<b>B</b>). Photographs of the incision were taken over time. Representative photographs of the healing process are shown (<b>C</b>). The length of the incision wound in each photograph was measured using FIJI, and the relative length at each time point to the initial length of 10 mm was calculated (<b>D</b>). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3) and are representative of five independent experiments. <span class="html-italic">p</span>-values were determined via two-way analysis of variance with Tukey’s multiple comparison test. The red and blue asterisks marked on day 6, 8, 10, and 12 in the figure (<b>D</b>) mean that there are significant differences between cha-koji and green tea and between cha-koji and distilled water, respectively. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001. Scale bar = 5 mm.</p>
Full article ">Figure 2
<p>Autoclave-sterilized cha-koji suspension showing anti-bacterial activity. <span class="html-italic">E. coli</span> was cultured in LB containing various concentrations of autoclave-sterilized cha-koji, green tea (0.5, 1.0, 2.0, 4.0, and 5.0%) or distilled water for 8 h at 37 °C (<b>A</b>). As a positive control, raw green tea was used. Bacterial growth was examined according to the turbidity of the LB medium, which was determined by measuring optical density at 600 nm (OD600) (<b>B</b>). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3) and are representative of three independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Tukey’s multiple comparison test. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Autoclave-sterilized cha-koji suspension promotes skin epidermal cell proliferation and migration. Keratinocyte PAM212 cells (<b>A</b>,<b>B</b>) or fibroblast NIH3T3 cells (<b>C</b>,<b>D</b>) were seeded in 24-well plates and cultured to semi-confluence in DMEM containing 10% FBS. The medium was then exchanged with DMEM containing 0.05% green tea or 0.05% cha-koji and 1.0% FBS. A cross-shaped scratch was made in the center of the well, and photographs were taken over time. Representative photographs are shown (<b>A</b>,<b>C</b>). The remaining wound area size was determined using FIJI, and the ratio relative to the initial wound area on day 0 was calculated (<b>B</b>,<b>D</b>). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 6) and are representative of four (<b>A</b>,<b>B</b>) and three (<b>C</b>,<b>D</b>) independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Tukey’s multiple comparison test. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001. Scale bar = 500 µm.</p>
Full article ">Figure 4
<p>Autoclave-sterilized cha-koji suspension promotes the expression of the anti-inflammatory cytokine TGF-β1. Macrophage RAW264.7 cells were stimulated with 0.3% cha-koji, 0.3% green tea, and distilled water for 24 and 48 h. RT-qPCR was then performed to analyze the mRNA expression of <span class="html-italic">TGF-β1</span> (<b>A</b>,<b>B</b>) and <span class="html-italic">IL-6</span> (<b>C</b>,<b>D</b>). <span class="html-italic">HPRT</span> was used as an internal control, and the relative expression of <span class="html-italic">TGF-β1</span> or <span class="html-italic">IL-6</span> to <span class="html-italic">HPRT</span> was calculated. Data are shown as the mean ± SD (<span class="html-italic">n</span> = 4−6) and are representative of three (<b>A</b>,<b>B</b>) and five (<b>C</b>,<b>D</b>) independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Dunnett’s multiple comparison test. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Autoclave-sterilized cha-koji suspension suppresses ER stress. Fibroblast NIH3T3 cells were stimulated with tunicamycin (1.0 μg/mL) in the presence or absence of 0.25% cha-koji and 0.25% green tea for 12 h. RNA was extracted, and RT-qPCR analysis was performed to examine the mRNA expression of ER stress-related factors, namely, <span class="html-italic">HSPA5</span> (<b>A</b>), <span class="html-italic">XBP1</span> (<b>B</b>), and <span class="html-italic">CHOP</span> (<b>C</b>). <span class="html-italic">HPRT</span> was used as an internal control, and its relative expression of <span class="html-italic">HPRT</span> was calculated. Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3) and are representative of two independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Dunnett’s multiple comparison test. After 48 h, photographs of each well were taken (<b>D</b>), and cell growth activity was determined by measuring the remaining viable cell area relative to the tunicamycin-untreated, distilled water-treated cell area using FIJI (<b>E</b>). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3) and are representative of three independent experiments. <span class="html-italic">p</span>-values were determined via one-way analysis of variance with Dunnett’s multiple comparison test. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001. Scale bar = 250 µm.</p>
Full article ">Figure 6
<p>Autoclave-sterilized cha-koji suspension poses no potential skin sensitizing risk. Human monocytic THP-1 cells were stimulated with cha-koji or green tea suspension (0.25, 1.0, and 4.0%) together with the positive control, conditioned medium from rhododenol-treated melanoma SK-MEL-37 cells. After 24 h, CD86 expression in the THP-1 cells was analyzed via flow cytometry using anti-CD86 (red-shaded histogram) or a control antibody (blue-shaded histogram). Conditioned medium from the rhododenol-treated melanoma (SK-MEL-37) was used as a positive control. Representative histograms for the cell surface expression of CD86 are shown (<b>A</b>). The RFI values of each sample were calculated and compared (<b>B</b>). An RFI higher than 150% is considered to be positive for skin sensitizing potential.</p>
Full article ">
11 pages, 17698 KiB  
Article
IL-10 Overexpression Reduces the Protective Response of an Experimental Chlamydia abortus Vaccine in a Murine Model
by Laura Del Río, Jesús Salinas, Nieves Ortega, Antonio J. Buendía, Jose A. Navarro and María Rosa Caro
Animals 2024, 14(16), 2322; https://doi.org/10.3390/ani14162322 - 12 Aug 2024
Viewed by 56
Abstract
In ovine populations, the enzootic nature of Chlamydia abortus (C. abortus) is attributed to its capacity to establish persistent intracellular infections, which necessitate a cellular immune response mediated by interferon-gamma (IFN-γ) for effective resolution. In both natural hosts and [...] Read more.
In ovine populations, the enzootic nature of Chlamydia abortus (C. abortus) is attributed to its capacity to establish persistent intracellular infections, which necessitate a cellular immune response mediated by interferon-gamma (IFN-γ) for effective resolution. In both natural hosts and murine models, interleukin-10 (IL-10) has been demonstrated to modulate the cellular immune response crucial for the eradication of C. abortus. During gestation, it has also been shown to play a role in preventing inflammatory damage to gestational tissues and foetal loss through the downregulation of pro-inflammatory cytokines. This paradigm can be key for events leading to a protective response towards an infectious abortion. Previous research successfully established a mouse model of chronic C. abortus infection using transgenic mice overexpressing IL-10 (IL-10tg), simulating the dynamics of chronic infection observed in non-pregnant natural host. This study aims to evaluate the efficacy of an experimental inactivated vaccine against C. abortus and to elucidate the immune mechanisms involved in protection during chronic infection using this model. Transgenic and wild-type (WT) control mice were immunized and subsequently challenged with C. abortus. Vaccine effectiveness and immune response were assessed via immunohistochemistry and cytokine serum levels over a 28-day period. Morbidity, measured by daily weight loss, was more pronounced in non-vaccinated transgenic IL-10 mice, though no mortality was observed in any group. Vaccinated control mice eliminated the bacterial infection by day 9 post-infection (p.i.), whereas presence of bacteria was noted in vaccinated transgenic IL-10 mice until day 28 p.i. Vaccination induced an early post-infection increase in IFN-γ production, but did not alter IL-10 production in transgenic mice. Histological analysis indicated suboptimal recruitment of inflammatory cells in vaccinated transgenic IL-10 mice compared to WT controls. In summary, the findings suggest that IL-10 overexpression in transgenic mice diminishes the protective efficacy of vaccination, confirming that this model can be useful for validating the efficacy of vaccines against intracellular pathogens such as C. abortus that require robust cell-mediated immunity. Full article
(This article belongs to the Special Issue Chlamydial Diseases in Animals)
Show Figures

Figure 1

Figure 1
<p>Vaccination and experimental infection. WT and IL-10-tg vaccinated mice received two doses of vaccine (15 μg of <span class="html-italic">C. abortus</span> proteins in 0.2 mL) subcutaneously 44 and 30 days before the infection. Mice were intraperitoneally challenged with 10<sup>6</sup> inclusion-forming units (IFUs) of <span class="html-italic">Chlamydia abortus</span> in 0.2 mL of 0.1M phosphate-buffered saline (PBS), pH 7.2, and monitored daily. Mice were euthanized at 4, 9, and 28 days post-inoculation).</p>
Full article ">Figure 2
<p>Morbidity, measured as daily weight loss, in IL-10-tg and wild-type mice (WT). Weight loss was only significant in vaccinated (V) as compared to control non-vaccinated (c) groups. Weight loss and slower recovery was observed in IL-10-tg-c mice 20 days p.i. No mortality was observed in any of the groups. (*** <span class="html-italic">p</span> &lt; 0.001; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Temporal analysis of <span class="html-italic">C. abortus</span> infection levels in liver tissue after infection as measured by IFUs/g. Although a reduction in infection levels was noted in all mouse groups, only the vaccinated WT group (WT-V) was able to clear the bacteria from the liver by day 9 p.i. At 28 days p.i., <span class="html-italic">C. abortus</span> was detected only in IL-10-tg mice, which was the case for both control (tg-c) and vaccinated (tg-V) groups. The orange line marks the challenge dose at day 0 p.i. (*** <span class="html-italic">p</span> &lt; 0.001; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Inflammatory response and chlamydial antigen levels in WT and IL-10-tg mice at different time points after infection. At day 4 p.i., both WT and IL-10-tg mice developed multifocal hepatitis with high levels of chlamydial antigen, even after vaccination. WT vaccinated (WT-V) mice exhibited inflammation predominantly with macrophages and PMNs, while IL-10-tg (IL-10-tg-V) mice had fewer and smaller foci dominated by macrophages. By day 9 p.i., lymphocytes predominated at inflammatory sites in both strains. By day 28 p.i., the presence of small inflammatory foci with chlamydial antigen in IL-10-tg-V suggested a reduced protective effect of vaccination compared to WT-V mice. Black arrows indicate the presence of chlamydial antigen in the tissue sections.</p>
Full article ">Figure 5
<p>IFN-<span class="html-italic">γ</span> and IL-10 levels in sera at different time points after infection: (<b>A</b>) IFN-<span class="html-italic">γ</span> had significantly higher values in the non-vaccinated groups, particularly in WT mice; (<b>B</b>) IL-10 was significantly elevated in IL-10-tg overexpressing mice as compared to WT mice at 4 and 9 days p.i. in both the control and vaccinated groups (tg-c and tg-V). (Black dots represent individual data points beyond the interquartile range. *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Expression of iNOS at day 4 p.i. in the liver. While iNOS expression remained unaffected in IL-10-tg mice, it was significantly reduced in the WT control group following vaccination ((<b>A</b>) WT non-vaccinated; (<b>B</b>) WT vaccinated). In contrast, the expression of iNOS was strong in the liver macrophages of IL-10-tg mice ((<b>C</b>) IL-10-tg non-vaccinated; (<b>D</b>) IL-10-tg vaccinated).</p>
Full article ">
25 pages, 7897 KiB  
Article
Fufang Muji Granules Ameliorate Liver Fibrosis by Reducing Oxidative Stress and Inflammation, Inhibiting Apoptosis, and Modulating Overall Metabolism
by Lei Men, Zhihong Gu, Enhua Wang, Jiwen Li, Zhongyu Li, Keke Li, Chunbin Li and Xiaojie Gong
Metabolites 2024, 14(8), 446; https://doi.org/10.3390/metabo14080446 (registering DOI) - 11 Aug 2024
Viewed by 288
Abstract
Fufang Muji granules (FMGs) are a prominent modern prescription Chinese patent formulation derived from the Muji decoction. Utilized in clinical practice for nearly four decades, FMGs have demonstrated efficacy in treating liver diseases. However, the precise mechanism of action remains unclear. This study [...] Read more.
Fufang Muji granules (FMGs) are a prominent modern prescription Chinese patent formulation derived from the Muji decoction. Utilized in clinical practice for nearly four decades, FMGs have demonstrated efficacy in treating liver diseases. However, the precise mechanism of action remains unclear. This study investigates the hepatoprotective effects of FMGs against liver fibrosis in rats based on untargeted metabolomics and elucidates their underlying mechanisms. A comprehensive model of liver fibrosis was established with 30% CCl4 (2 mL/kg) injected intraperitoneally, and a fat and sugar diet combined with high temperatures and humidity. Rats were orally administered FMGs (3.12 g/kg/d) once daily for six weeks. FMG administration resulted in improved liver fibrosis and attenuated hepatic oxidative stress and apoptosis. Furthermore, FMGs inhibited hepatic stellate cell activation and modulated transforming growth factor β1/Smad signaling. Additionally, FMG treatment influenced the expression levels of interleukin-6, interleukin-1β, and tumour necrosis factor alpha in the injured liver. Metabolic pathways involving taurine and hypotaurine metabolism, as well as primary bile acid biosynthesis, were identified as mechanisms of action for FMGs. Immunohistochemistry, quantitative reverse transcription polymerase chain reaction (RT-qPCR), and quantitative analysis also revealed that FMGs regulated taurine and hypotaurine metabolism and bile acid metabolism. These findings provide a valuable understanding of the role of FMGs in liver fibrosis management. Full article
13 pages, 5142 KiB  
Article
Spike Protein of SARS-CoV-2 Activates Cardiac Fibrogenesis through NLRP3 Inflammasomes and NF-κB Signaling
by Huynh Van Tin, Lekha Rethi, Satoshi Higa, Yu-Hsun Kao and Yi-Jen Chen
Cells 2024, 13(16), 1331; https://doi.org/10.3390/cells13161331 - 11 Aug 2024
Viewed by 734
Abstract
Background: The spike protein of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is crucial to viral entry and can cause cardiac injuries. Toll-like receptor 4 (TLR4) and NOD-, LPR-, and pyrin-domain-containing 3 (NLRP3) inflammasome are critical immune system components implicated in cardiac fibrosis. [...] Read more.
Background: The spike protein of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is crucial to viral entry and can cause cardiac injuries. Toll-like receptor 4 (TLR4) and NOD-, LPR-, and pyrin-domain-containing 3 (NLRP3) inflammasome are critical immune system components implicated in cardiac fibrosis. The spike proteins activate NLRP3 inflammasome through TLR4 or angiotensin-converting enzyme 2 (ACE2) receptors, damaging various organs. However, the role of spike proteins in cardiac fibrosis in humans and the interactions of spike proteins with NLRP3 inflammasomes and TLR4 remain poorly understood. Methods: We utilized scratch assays, Western blotting, and immunofluorescence to evaluate the migration, fibrosis signaling, mitochondrial calcium levels, reactive oxygen species (ROS) production, and cell morphology of cultured human cardiac fibroblasts (CFs) treated with spike (S1) proteins for 24 h with or without an anti-ACE2 neutralizing antibody, a TLR4 blocker, or an NLRP3 inhibitor. Results: S1 protein enhanced CFs migration and the expressions of collagen 1, α-smooth muscle actin, transforming growth factor β1 (TGF-β1), phosphorylated SMAD2/3, interleukin 1β (IL-1β), and nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB). S1 increased ROS production but did not affect mitochondrial calcium content and cell morphology. Treatment with an anti-ACE2 neutralizing antibody attenuated the effects of S1 on collagen 1 and TGF-β1 expressions. Moreover, NLRP3 (MCC950) and NF-kB inhibitors, but not the TLR4 inhibitor TAK-242, prevented the S1-enhanced CFs migration and overexpression of collagen 1, TGF-β1, and IL-1β. Conclusion: S1 activates human CFs by priming NLRP3 inflammasomes through NF-κB signaling in an ACE2-dependent manner. Full article
(This article belongs to the Special Issue Insight into Cardiomyopathy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>S1 protein enhanced CFs activation. Treatment with S1 protein (5 nM) for 24 h increased cell migration (<b>A</b>) but not cell proliferation (<b>B</b>) in CFs. (<b>C</b>) Additionally, S1 protein also elevated pro-COL1A1 and α-SMA protein expressions. <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 2
<p>S1 protein increased profibrotic signaling. Treatment with S1 protein (5 nM) for 24 h increased protein expressions of TGF-β1 and pSMAD2/3 measured using Immunoblot ((<b>A</b>) <span class="html-italic">n</span> = 4 independent experiments), secretion of TGF-β1 in cultured medium measured using ELISA assays ((<b>B</b>) <span class="html-italic">n</span> = 5 independent experiments), and TGF-β1 mRNA quantified using RT-qPCR in CFs ((<b>C</b>) <span class="html-italic">n</span> = 4 independent experiments).</p>
Full article ">Figure 3
<p>Impact of NLRP3 and TLR4 signaling on S1 protein-mediated CFs migration. Pretreatment with MCC950 (10 µM (<b>A</b>)) but not TAK-242 (1 µM (<b>B</b>)) blocked the effects of S1 protein (5 nM) treatment for 24 h on cell migration. <span class="html-italic">n</span> = 3 independent experiments.</p>
Full article ">Figure 4
<p>The role of NLRP3 signaling on S1 protein-induced fibrotic markers in CFs. MCC950 effectively blocked the effect of S1 on expressions of fibrotic markers including pro-COL1A1, TGF-β1, and its downstream target, pSMAD2/3. <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 5
<p>Role of TLR4 signaling on S1 protein-induced fibrotic markers in CFs. TAK-242 did not change the effect of S1 on pro-COL1A1 and TGF-β1 expressions in CFs (10 µM, B). <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 6
<p>Role of NF-κB on S1 protein-mediated CFs migration and expressions of fibrotic factors. Pretreatment with BAY 11-7082 (an NF-κB inhibitor, 3 µM) completely blocked the effects of S1 protein (5 nM for 24 h) on CFs migration ((<b>A</b>) <span class="html-italic">n</span> = 5 independent experiments) and the protein expressions of IL1-β cleavage, TGF-β1, and pSMAD2/3 ((<b>B</b>,<b>C</b>) <span class="html-italic">n</span> = 4 independent experiments).</p>
Full article ">Figure 7
<p>S1 activated CFs in an ACE2-dependent manner. ACE2 neutralization antibody (5 µM) effectively blocked the enhanced CFs migration induced by the S1 protein (5 nM for 24 h) (<b>A</b>), along with suppressing the protein expressions of pro-COL1A1, TGF-β1, and phosphorylated NF-κB (p-p65) (<b>B</b>). <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 8
<p>Mechanisms underlying S1 protein-induced activation of CFs and promotion of ECM protein synthesis. Key processes involve activation of the NLRP3 inflammasome, ACE2/NF-κB signaling, and ROS formation. Upon binding to ACE2, the S1 protein initiates a signaling cascade that activates NF-κB, a transcription factor promoting the expression of inflammation-related genes, including those required for NLRP3 inflammasome and pro-interleukin (IL)-1β. ROS formation triggers NLRP3 inflammasome activation, leading to processing of pro-IL-1β into its mature form by caspase-1. Mature IL-1β is released extracellularly, binds to its receptor, and initiates a signaling cascade, enhancing TGF-β1 production, promoting CFs activation, and inducing ECM synthesis, ultimately contributing to cardiac fibrosis. Abbreviations: SARS-CoV-2: severe acute respiratory syndrome coronary virus 2, CFs: cardiac fibroblasts, ECM: extra cellular matrix, EMT: epithelial-mesenchymal transition, NLRP3: NLR family pyrin domain containing 3, ACE2: angiotensin converting enzyme 2, NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells, ROS: reactive oxygen species, IL-1β: Interleukin 1 beta.</p>
Full article ">
10 pages, 2807 KiB  
Communication
Long-Term Stability and Efficacy of NCT Solutions
by Gabriel J. Staudinger, Zach M. Thomas, Sarah E. Hooper, Jeffrey F. Williams and Lori I. Robins
Int. J. Mol. Sci. 2024, 25(16), 8745; https://doi.org/10.3390/ijms25168745 (registering DOI) - 10 Aug 2024
Viewed by 292
Abstract
To realize the potential for the use of N-chlorotaurine (NCT) in healthcare, a better understanding of the long-term stability of the compound in water is needed. An array of analytical procedures is required that can measure changes in NCT concentration over time [...] Read more.
To realize the potential for the use of N-chlorotaurine (NCT) in healthcare, a better understanding of the long-term stability of the compound in water is needed. An array of analytical procedures is required that can measure changes in NCT concentration over time and allow for the detection and identification of contaminants and likely degradation end products. We used UV-Vis and NMR spectroscopy, HPLC, and LCMS to establish the stability of NCT in solutions subjected to prolonged ambient and elevated temperatures. Stability proved to be dependent on concentration with half-lives of ~120 days and ~236 days for 1% and 0.5% solutions of NCT at ~20 °C. Regardless of initial pH, all solutions shifted toward and maintained a pH of ~8.3 at 20 °C and 40 °C. NCT at 500 µg/mL and 250 µg /mL inhibited biofilm formation by Pseudomonas aeruginosa and Staphylococcus aureus but did not disperse established biofilms. NCT exposure to the biofilms had profound effects on the viability of both bacteria, reducing live organisms by >90%. Exposure of Interleukin-6 (IL-6) to 11 µM NCT reduced the binding of IL-6 to an immobilized specific antibody by ~48%, which is 5× the amount required for HOCl to bring about the same effect in this test system. Our data demonstrate the potency of the compound as an antimicrobial agent with potential benefits in the management of infected chronic wounds and suggest that NCT may contribute to anti-inflammatory processes in vivo by direct modification of cytokine mediators. Full article
Show Figures

Figure 1

Figure 1
<p>UV-Vis spectroscopy, IR spectroscopy, and <sup>13</sup>C NMR spectroscopy of NCT. (<b>A</b>) UV–Visible spectroscopy wavelength scans of NCT (blue), N,N-dichlorotaurine (black), and taurine (teal). (<b>B</b>) IR spectrum of taurine (black) and NCT (blue). (<b>C</b>) <sup>13</sup>C NMR spectrum of NCT (black, singlets at 50.74 and 48.96 ppm) and taurine (gray, singlets at 47.42 and 35.38 ppm).</p>
Full article ">Figure 2
<p>HPLC and LCMS analysis of NCT. (<b>A</b>) HPLC spectrum of NCT (1 mg/mL) at 252 nm. The area under the curve was calculated and purity was determined to be &gt;95%. (<b>B</b>) UV-Vis wavelength trace of NCT detected by the LCMS. (<b>C</b>) TIC data for the corresponding LCMS peak at 252 nm. (<b>D</b>) MS data for the corresponding TIC peak at 252 nm.</p>
Full article ">Figure 3
<p>Stability of NCT solutions at 1%, 0.5%, and 0.25% at ambient and elevated temperatures. (<b>A</b>) Starting pH 9.5 at room temperature. (<b>B</b>) Starting pH 7 at 40 °C. (<b>C</b>) Starting pH 8 at 40 °C. (<b>D</b>) Starting pH 9.5 at 40 °C. Each condition was repeated three times. Additional stability data for these trials are available in <a href="#app1-ijms-25-08745" class="html-app">Figure S1</a>. All 1% solutions are in blue; 0.5% in green; and 0.25% in black. A single-factor ANOVA was used to test for differences in starting pH values and for differences in concentrations.</p>
Full article ">Figure 4
<p>pH of NCT solutions at 1% (circle), 0.5% (square), and 0.25% (triangle). (<b>A</b>) Solutions starting at a pH value of 7. (<b>B</b>) Solutions starting at a pH value of 8. (<b>C</b>) Solutions starting at a pH value of 9.5. Additional pH data for the three trials are available in <a href="#app1-ijms-25-08745" class="html-app">Figure S2</a>.</p>
Full article ">Figure 5
<p>ELISA test results for IL-6 binding to an IL-6 specific antibody after treatment with concentrations of NCT ranging from 0 to 2.76 mM. A single-factor ANOVA was used to test for differences in binding at all concentrations of NCT tested.</p>
Full article ">Figure 6
<p>Viability of remaining <span class="html-italic">P. aeruginosa</span> (black) and <span class="html-italic">S. aureus</span> (gray) biofilm biomass after treatment with various NCT concentrations. Significant reductions in viability were calculated using ANOVA and Tukey’s post hoc analysis. The significant threshold concentrations were 0.004% for <span class="html-italic">S. aureus</span> and 0.06% for <span class="html-italic">P. aeruginosa</span>.</p>
Full article ">
15 pages, 16273 KiB  
Article
Xanthoxylin Attenuates Lipopolysaccharide-Induced Lung Injury through Modulation of Akt/HIF-1α/NF-κB and Nrf2 Pathways
by Fu-Chao Liu, Yuan-Han Yang, Chia-Chih Liao and Hung-Chen Lee
Int. J. Mol. Sci. 2024, 25(16), 8742; https://doi.org/10.3390/ijms25168742 (registering DOI) - 10 Aug 2024
Viewed by 263
Abstract
Xanthoxylin, a bioactive phenolic compound extracted from the traditional herbal medicine Penthorum Chinense Pursh, is renowned for its anti-inflammatory effects. While previous studies have highlighted the anti-inflammatory and antioxidant properties of Xanthoxylin, its precise mechanisms, particularly concerning immune response and organ protection, [...] Read more.
Xanthoxylin, a bioactive phenolic compound extracted from the traditional herbal medicine Penthorum Chinense Pursh, is renowned for its anti-inflammatory effects. While previous studies have highlighted the anti-inflammatory and antioxidant properties of Xanthoxylin, its precise mechanisms, particularly concerning immune response and organ protection, remain underexplored. This study aimed to elucidate the effects of Xanthoxylin on inflammation and associated signaling pathways in a mouse model of lipopolysaccharide (LPS)-induced acute lung injury (ALI). ALI was induced via intratracheal administration of LPS, followed by intraperitoneal injections of Xanthoxylin at doses of 1, 2.5, 5, and 10 mg/kg, administered 30 min post-LPS exposure. Lung tissues were harvested for analysis 6 h after LPS challenge. Xanthoxylin treatment significantly mitigated lung tissue damage, pathological alterations, immune cell infiltration, and the production of pro-inflammatory cytokines, including tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6). Additionally, Xanthoxylin modulated the expression of key proteins in the protein kinase B (Akt)/hypoxia-inducible factor 1-alpha (HIF-1α)/nuclear factor-kappa B (NF-κB) signaling pathway, as well as nuclear factor erythroid 2-related factor 2 (Nrf2) and oxidative markers such as superoxide dismutase (SOD) and malondialdehyde (MDA) in the context of LPS-induced injury. This study demonstrates that Xanthoxylin exerts protective and anti-inflammatory effects by down-regulating and inhibiting the Akt/HIF-1α/NF-κB pathways, suggesting its potential as a therapeutic target for the prevention and treatment of ALI or acute respiratory distress syndrome (ARDS). Full article
(This article belongs to the Special Issue New Insights in Natural Bioactive Compounds 3.0)
Show Figures

Figure 1

Figure 1
<p>The effect of Xanthoxylin on the viability and pro-inflammatory cytokine levels of RAW 264.7 cells in the presence and absence of LPS. (<b>A</b>) RAW 264.7 cells were treated with various concentrations of DMSO (0.01, 0.05, and 0.5 μL) or Xanthoxylin (0.1, 1, 5, 10, 20, and 50 μM) for 24 h. Results are expressed as a percentage relative to the control group and shown as mean ± SD (<span class="html-italic">n</span> = 6 per group). (<b>B</b>) RAW 264.7 cells were treated with Xanthoxylin (0, 5, and 10 μM) followed by LPS exposure for 48 h to assess cell viability. Results are expressed as a percentage relative to the control group and shown as mean ± SD (<span class="html-italic">n</span> = 12 per group). (<b>C</b>) Pro-inflammatory cytokines IL-1β, IL-6, and TNF-α levels in the supernatants of RAW 264.7 cells were measured after treatment with Xanthoxylin, followed by LPS exposure. ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <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; 0.005 vs. LPS group.</p>
Full article ">Figure 2
<p>General lung appearance after LPS-induced injury and Xanthoxylin treatment. Mice received an intratracheal LPS challenge followed by intraperitoneal administration of Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline. Lungs were collected 6 h post-LPS challenge for analysis.</p>
Full article ">Figure 3
<p>Histological examination of lung tissues stained with H&amp;E after LPS challenge and Xanthoxylin treatment. Mice received LPS intratracheally and were then treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge for H&amp;E staining. Representative images show ALI and histological changes (100× magnification, scar bar = 100 μm). Quantification of histologic lung injury was analyzed according to American Thoracic Society (ATS) scoring system (<span class="html-italic">n</span> = 6 per group). # <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; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001 vs. LPS group.</p>
Full article ">Figure 4
<p>Neutrophil infiltration in lungs following LPS-induced injury and Xanthoxylin treatment. Mice were challenged with LPS intratracheally and treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge and immunostained with Ly6G antibody (200× magnification, scar bar = 50 μm). Quantification of positive cells was analyzed under high power field (HPF). Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 5
<p>Macrophage infiltration in lungs following LPS-induced injury and Xanthoxylin treatment. Mice received LPS intratracheally and were treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge and immunostained with Mac-2 antibody (200× magnification, scar bar = 50 μm). Quantification of positive cells was analyzed under HPF. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 6
<p>Levels of (<b>A</b>) IL-6 and (<b>B</b>) TNF-α in lungs after LPS challenge and Xanthoxylin treatment. Mice were given intratracheal LPS challenge followed by Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge for ELISA. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). # <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; 0.005 vs. control group; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 7
<p>Levels of (<b>A</b>) MDA and (<b>B</b>) SOD in lungs after LPS challenge and Xanthoxylin treatment. Mice received intratracheal LPS challenge and were treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge for oxidative stress assays. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). # <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <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; 0.005 vs. LPS group.</p>
Full article ">Figure 8
<p>Effects of Xanthoxylin on expression of (<b>A</b>) Akt, (<b>B</b>) NF-κB, (<b>C</b>) HIF-1α, and (<b>D</b>) Nrf2 in lungs after LPS challenge. Mice were administered Xanthoxylin (XT, 2.5, 5, and 10 mg/kg) or saline intraperitoneally 30 min post-LPS challenge. Lungs were harvested 6 h later for Western blot analysis. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <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; 0.005 vs. LPS group.</p>
Full article ">Figure 9
<p>Nrf2 expression in lungs after LPS-induced injury and Xanthoxylin treatment. Mice received intratracheal LPS challenge followed by Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge and immunostained with Nrf2 antibody (400× magnification, scar bar = 25 μm). Quantification of positive cells was analyzed under HPF. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 10
<p>A schematic representation of the involvement of Akt/HIF-1α/NF-κB and Nrf2 signaling pathways in the protective effects of Xanthoxylin against LPS-induced lung injury. Xanthoxylin modulates Akt expression, suppresses HIF-1α/NF-κB signaling, and activates Nrf2, thereby reducing cell damage and oxidative stress. It also inhibits TNF-α and IL-6 release from macrophages.</p>
Full article ">
19 pages, 3924 KiB  
Article
Comparative Effectiveness of Various Multi-Antigen Vaccines in Controlling Campylobacter jejuni in Broiler Chickens
by Mostafa Naguib, Shreeya Sharma, Abigail Schneider, Sarah Wehmueller and Khaled Abdelaziz
Vaccines 2024, 12(8), 908; https://doi.org/10.3390/vaccines12080908 (registering DOI) - 10 Aug 2024
Viewed by 309
Abstract
This study was undertaken to evaluate and compare the efficacy of different multi-antigen vaccines, including heat-inactivated, whole lysate, and subunit (outer membrane proteins [OMPs]) C. jejuni vaccines along with the immunostimulant CpG ODN in controlling Campylobacter colonization in chickens. In the first trial, [...] Read more.
This study was undertaken to evaluate and compare the efficacy of different multi-antigen vaccines, including heat-inactivated, whole lysate, and subunit (outer membrane proteins [OMPs]) C. jejuni vaccines along with the immunostimulant CpG ODN in controlling Campylobacter colonization in chickens. In the first trial, 125 μg of C. jejuni OMPs and 50 μg of CpG ODN were administered individually or in combination, either in ovo to chick embryos or subcutaneously (SC) to one-day-old chicks. In the second trial, different concentrations of C. jejuni antigens (heat-killed, whole lysate, and OMPs) were administered SC to one-day-old chicks. The results of the first trial revealed that SC immunization with the combination of CpG ODN and C. jejuni OMPs elevated interferon (IFN)-γ, interleukin (IL)-1β, and IL-13 gene expression in the spleen, significantly increased serum IgM and IgY antibody levels, and reduced cecal C. jejuni counts by approximately 1.2 log10. In contrast, in ovo immunization did not elicit immune responses or confer protection against Campylobacter. The results of the second trial showed that SC immunization with C. jejuni whole lysate or 200 μg OMPs reduced C. jejuni counts by approximately 1.4 and 1.1 log10, respectively. In conclusion, C. jejuni lysate and OMPs are promising vaccine antigens for reducing Campylobacter colonization in chickens. Full article
(This article belongs to the Special Issue Vaccines against Enteric Bacterial Pathogens in Poultry)
Show Figures

Figure 1

Figure 1
<p>Illustration of first experimental design. On embryonic day 18, 136 eggs were randomly divided into four groups (G1–G4), each containing 34 eggs. Embryos of each group were injected intra-amniotic with the assigned immunization: 50 µg CpG ODN or 125 µg <span class="html-italic">C. jejuni</span> OMPs or their combination or PBS (negative control group). On the first day post-hatch, chicks (136) of non-immunized eggs were randomly allocated into four groups (G5–G8) and immunized SC with 50 µg CpG ODN or 125 µg <span class="html-italic">C. jejuni</span> OMPs or their combination or PBS. All chicks received the booster vaccination on day seven, and all groups were challenged with 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span> on day 14. Bursa of Fabricius and spleens were collected for three successive days (<span class="html-italic">n</span> = 8) post-initial vaccination of either SC or in ovo immunized groups. Blood samples were collected weekly, and the cecal contents were collected on day 35 of age (the end of the experiment).</p>
Full article ">Figure 2
<p>Illustration of the second experimental design. On the first day post-hatch, chicks were immunized SC with <span class="html-italic">C. jejuni</span> OMPs (50, 125 or 200 μg) or <span class="html-italic">C. jejuni</span> lysate; low (21.5 μg) or high (43 μg) protein, or heat-killed (10<sup>6</sup> or 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span>). All chicks received the booster vaccination on day 14 of age. The control group was injected at the same age with PBS only. All groups were challenged with 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span> on day 15 of age. Blood samples were collected weekly, and the cecal contents were collected at the fifth week at the end of the experiment.</p>
Full article ">Figure 3
<p><span class="html-italic">C. jejuni</span> CFUs per gram of the cecal content. After primary immunization in ovo or SC on day one of age and booster immunization on day seven of age, chicks in all groups were orally challenged with 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span> on day 14 of age. At 35 days of age (21 days post-challenge), cecal contents were collected for <span class="html-italic">Campylobacter</span> enumeration. Bars marked with different letters (a–c) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the groups, while bars marked with the same letter denote no significant differences between the groups. OMPs = outer membrane proteins. SC = subcutaneous. ODN = synthetic single-stranded oligodeoxynucleotides (ODNs) containing unmethylated CpG motifs.</p>
Full article ">Figure 4
<p>Serum IgY antibody levels. Chicks were immunized in ovo or SC with 125 µg <span class="html-italic">C. jejuni</span> OMPs or 50 µg CpG ODN or their combination or PBS. A booster dose was delivered orally (for those primed in ovo) or SC (for those primed SC) on day seven of age. All the chicks were then orally challenged with 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span> on day 14 of age. Blood samples were collected weekly from all groups, starting from the first week of age through the fifth week of age, with the sera subsequently separated for measuring the IgY antibody (Ab) levels using ELISA. Bars marked with different letters (a–b) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the groups, while bars marked with the same letter denote no significant differences between the groups. OMPs = outer membrane proteins. SC = subcutaneous. ODN = synthetic single-stranded oligodeoxynucleotides (ODNs) containing unmethylated CpG motifs.</p>
Full article ">Figure 5
<p>The correlation between <span class="html-italic">C. jejuni</span> CFUs and serum antibody (Ab) levels using Pearson’s r correlation coefficient. No correlation was observed between the serum IgY Ab levels and cecal counts of <span class="html-italic">C. jejuni</span> in the group immunized SC with the combination of 50 µg CpG ODN and 125 µg OMPs at the fifth week of age.</p>
Full article ">Figure 6
<p>Serum IgM antibody levels. Chicks were immunized in ovo or SC with 125 µg <span class="html-italic">C. jejuni</span> OMPs or 50 µg CpG ODN or their combination or PBS. A booster dose was delivered orally (for those primed in ovo) or SC (for those primed SC) on day seven of age. All the chicks were then orally challenged with 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span> on day 14 of age. Blood samples were collected weekly from all groups, starting from the first week of age through the fifth week of age, with the sera subsequently separated for measuring the IgM antibody (Ab) levels using ELISA. Bars marked with different letters (a–c) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the groups, while bars marked with the same letter denote no significant differences between the groups. OMPs = outer membrane proteins. SC = subcutaneous. ODN = synthetic single-stranded oligodeoxynucleotides (ODNs) containing unmethylated CpG motifs.</p>
Full article ">Figure 7
<p>The correlation between <span class="html-italic">C. jejuni</span> CFUs and serum antibody (Ab) levels using Pearson’s r correlation coefficient. No correlation was observed between the serum IgM Ab levels and cecal counts of <span class="html-italic">C. jejuni</span> in the group immunized with the combination of 50 µg CpG ODN and 125 µg OMPs at the fifth week of age.</p>
Full article ">Figure 8
<p>Relative gene expression of interferon (IFN)-γ (<b>a</b>), interleukin (IL)-13 (<b>b</b>), and IL-1β (<b>c</b>) in the spleen at 24, 48, and 72 h following SC immunization with 125 µg <span class="html-italic">C. jejuni</span> OMPs or 50 µg CpG ODN or their combination or PBS. Data are presented as mean expression (delta CT values) of cytokine mRNA relative to β-actin (housekeeping gene) ± standard error of the mean (SEM). Statistical significance among treatment groups was calculated using 1-way ANOVA followed by Tukey’s comparison test. Bars marked with different letters (<b>a</b>,<b>b</b>) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the groups, while bars marked with the same letter denote no significant differences between the groups. OMPs = outer membrane proteins. SC = subcutaneously. ODN = synthetic single-stranded oligodeoxynucleotides (ODNs) containing unmethylated CpG motifs.</p>
Full article ">Figure 9
<p>Relative gene expression of interferon (IFN)-γ (<b>a</b>), interleukin (IL)-13 (<b>b</b>), and IL-1β (<b>c</b>) in the spleen at 24, 48, and 72 h following in ovo immunization with 125 µg <span class="html-italic">C. jejuni</span> OMPs or 50 µg CpG ODN or their combination or PBS. Data are presented as mean expression (delta CT values) of cytokine mRNA relative to β-actin (housekeeping gene) ± standard error of the mean (SEM). Statistical significance among treatment groups was calculated using 1-way ANOVA followed by Tukey’s comparison test. Bars marked with different letters (<b>a</b>,<b>b</b>) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the groups, while bars marked with the same letter denote no significant differences between the groups. OMPs = outer membrane proteins. SC = subcutaneously. ODN = synthetic single-stranded oligodeoxynucleotides (ODNs) containing unmethylated CpG motifs.</p>
Full article ">Figure 10
<p><span class="html-italic">C. jejuni</span> CFUs per gram of the cecal content. Day-old chicks were immunized SC with different doses of heat-killed or whole lysate or OMPs of <span class="html-italic">C. jejuni</span> and boosted SC with the corresponding vaccine on day 14 of age. On day 15 of age, chicks in all groups were orally challenged with 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span>. On day 35 of age (20-day post-challenge), cecal contents were collected for <span class="html-italic">Campylobacter</span> enumeration. Bars marked with different letters (a–c) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the groups, while bars marked with the same letter denote no significant differences between the groups. OMPs = outer membrane proteins. SC = subcutaneous.</p>
Full article ">Figure 11
<p>Serum IgY antibody levels. On days one and fourteen of age, chicks were immunized SC with different doses of <span class="html-italic">C. jejuni</span> OMPs (50, 125 or 200 μg) or <span class="html-italic">C. jejuni</span> lysate (21.5 μg or 43 μg) or heat-killed <span class="html-italic">C. jejuni</span> (10<sup>6</sup> or 10<sup>7</sup> CFUs). The control group was injected at the same age with PBS only. On day 15 of age, chickens were challenged orally with 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span>. Blood samples were collected weekly from all groups, starting from the first week of age through the fifth week of age, with the sera subsequently separated for measuring the IgY antibody (Ab) levels using ELISA. Bars that are marked with different letters (a–c) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). OMPs = outer membrane proteins. SC = subcutaneous.</p>
Full article ">Figure 12
<p>Serum IgM antibody levels. On days one and fourteen of age, chicks were immunized SC with different doses of <span class="html-italic">C. jejuni</span> OMPs (50, 125 or 200 μg), <span class="html-italic">C. jejuni</span> lysate (21.5 μg or 43 μg) and heat-killed <span class="html-italic">C. jejuni</span> (10<sup>6</sup> or 10<sup>7</sup> CFUs). The control group was injected at the same age with PBS only. On day 15 of age, chickens were challenged orally with 10<sup>7</sup> CFUs of <span class="html-italic">C. jejuni</span>. Blood samples were collected weekly from all groups, starting from the first week of age through the fifth week of age, with the sera subsequently separated for measuring the IgM antibody (Ab) levels using ELISA. Bars marked with different letters (a–b) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the groups, while bars marked with the same letter denote no significant differences between the groups. OMPs = outer membrane proteins. SC = subcutaneous.</p>
Full article ">
13 pages, 1875 KiB  
Article
A Novel Model of Venovenous Extracorporeal Membrane Oxygenation in Rats with Femoral Cannulation and Insights into Hemodynamic Changes
by Fabian Edinger, Thomas Zajonz, Nico Mayer, Götz Schmidt, Emmanuel Schneck, Michael Sander and Christian Koch
Biomedicines 2024, 12(8), 1819; https://doi.org/10.3390/biomedicines12081819 - 10 Aug 2024
Viewed by 200
Abstract
The application of venovenous (VV) extracorporeal membrane oxygenation (ECMO) has gained wide acceptance for the treatment of acute severe respiratory failure. Since no rat model of VV ECMO therapy with femoral drainage has yet been described, although this cannulation strategy is commonly used [...] Read more.
The application of venovenous (VV) extracorporeal membrane oxygenation (ECMO) has gained wide acceptance for the treatment of acute severe respiratory failure. Since no rat model of VV ECMO therapy with femoral drainage has yet been described, although this cannulation strategy is commonly used in humans, this study aimed to establish such a model. Twenty male Lewis rats were randomly assigned to receive a sham procedure or VV ECMO therapy. After the inhalative induction of anesthesia, animals were intubated and the vascular accesses were placed surgically. While venous drainage was achieved through a modified multi-orifice 18 G cannula that was placed in the inferior vena cava through the femoral vein over a guide wire with an ultra-flexible tip, the venous return was realized via a shortened 20 G cannula into the jugular vein. Hemodynamic data were obtained from a tail artery and left ventricular pressure–volume catheter. Repetitive blood gas analyses were carried out, and systemic inflammation was measured using an enzyme-linked immunosorbent assay. While animals in the ECMO group showed adequate oxygenation and decarboxylation, there was no evidence of recirculation. VV ECMO therapy increased stroke volume (SV), cardiac output (CO), and left ventricular end-diastolic volume (LVEDV). ECMO-induced inflammation was reflected in increased levels of tumor necrosis factor alpha. However, no differences in interleukins 6 and 10 were seen. This study describes a frequently used cannulation strategy in humans for a rat model of VV ECMO. Despite successful oxygenation and decarboxylation, the oxygenated blood may reduce pulmonary vascular resistance and lead to an increased LVEDV, which is associated with increased SV and CO. This model allows us to answer research questions about topics such as intestinal microcirculation in further studies. Full article
(This article belongs to the Section Biomedical Engineering and Materials)
Show Figures

Figure 1

Figure 1
<p>Draining ECMO cannula and ultra-flexible wire. Balanced middle weight wire with ultra-flexible tip (0.36 mm; Abbott, Wetzlar, Germany) and modified draining ECMO cannula with 12 side holes (18 G Surflo, Terumo, Eschborn, Germany) are pictured.</p>
Full article ">Figure 2
<p>Time course of (<b>A</b>) arterial and (<b>B</b>) central venous saturation, (<b>C</b>) arterial partial pressure of oxygen (pO<sub>2</sub>) and (<b>D</b>) carbon dioxide (pCO<sub>2</sub>). While no differences were measured between the sham and VV ECMO groups regarding arterial and central venous saturation and pO<sub>2</sub>, animals in the sham group showed significantly elevated concentrations of pCO<sub>2</sub>. The asterisks denote the degree of statistical significance (** <span class="html-italic">p</span> &lt; 0.01). Abbreviations: ECMO = extracorporeal membrane oxygenation; pO<sub>2</sub> = arterial partial pressure of oxygen; pCO<sub>2</sub> = arterial partial pressure of carbon dioxide.</p>
Full article ">Figure 3
<p>Time course of (<b>A</b>) systolic (SAP), (<b>B</b>) mean (MAP), (<b>C</b>) diastolic arterial blood pressure (DAP), and (<b>D</b>) heart rate. During VV ECMO therapy, an elevated SAP was measured compared to sham animals. The asterisks denote the degree of statistical significance (* <span class="html-italic">p</span> &lt; 0.05). Abbreviations: DAP = diastolic arterial blood pressure; ECMO = extracorporeal membrane oxygenation; MAP = mean arterial blood pressure; SAP = systolic arterial blood pressure.</p>
Full article ">Figure 4
<p>Time course of (<b>A</b>) stroke volume (SV), (<b>B</b>) cardiac output (CO), (<b>C</b>) left ventricular end-diastolic volume (LVEDV), and (<b>D</b>) left ventricular ejection fraction (LVEF). During VV ECMO therapy, increased SV, CO, and LVEDV were captured compared to sham animals. Furthermore, no differences were seen regarding LVEF. The asterisks denote the degree of statistical significance (** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001). Abbreviations: CO = cardiac output; ECMO = extracorporeal membrane oxygenation; LVEDV = left ventricular end-diastolic volume; LVEF = left ventricular ejection fraction; SV = stroke volume.</p>
Full article ">Figure 5
<p>Time course of (<b>A</b>) tumor necrosis factor alpha (TNF-α), (<b>B</b>) interleukin 6 (IL-6), and (<b>C</b>) interleukin 10 (IL-10). During VV ECMO therapy, increased concentrations of TNF-α were measured compared to sham animals. Furthermore, no differences were measured regarding IL-6 and IL-10. The asterisks denote the degree of statistical significance (* <span class="html-italic">p</span> &lt; 0.05). Box and whisker plots indicate the median, interquartile range (box), and minimum and maximum (whiskers). Abbreviations: ECMO = extracorporeal membrane oxygenation; IL-6 = interleukin 6; IL-10 = interleukin 10; TNF-α = tumor necrosis factor alpha.</p>
Full article ">
15 pages, 855 KiB  
Article
Genetic Influences of Proinflammatory Cytokines on Pain Severity in Patients with Temporomandibular Disorders
by Marko Zlendić, Ema Vrbanović Đuričić, Koraljka Gall Trošelj, Marko Tomljanović, Kristina Vuković Đerfi and Iva Z. Alajbeg
Int. J. Mol. Sci. 2024, 25(16), 8730; https://doi.org/10.3390/ijms25168730 (registering DOI) - 10 Aug 2024
Viewed by 186
Abstract
This case-control study investigated single nucleotide polymorphism (SNP) genotypes (CXC motif chemokine ligand 8 (CXCL8): rs2227306 and rs2227307 and tumor necrosis factor (TNF): rs1800629) in 85 patients with pain-related temporomandibular disorders (TMDp) and 85 controls to explore their associations [...] Read more.
This case-control study investigated single nucleotide polymorphism (SNP) genotypes (CXC motif chemokine ligand 8 (CXCL8): rs2227306 and rs2227307 and tumor necrosis factor (TNF): rs1800629) in 85 patients with pain-related temporomandibular disorders (TMDp) and 85 controls to explore their associations with TMDp presence, pain intensity (low/high), and the presence of chronic arthralgia/myalgia. TMDp was diagnosed using a validated protocol, and polymorphisms were genotyped from buccal mucosa swabs using TaqMan assays. High pain intensity individuals had an increased risk for carrying minor allele “G” (rs2227307) and “T” (rs2227306) compared to controls (76% vs. 55.3%, p = 0.012; 72% vs. 54.1%, p = 0.030, respectively). Carriers of the minor allele “G” (rs2227307) were more prevalent in TMDp patients with arthralgia compared to controls (70.30% vs. 55.30%, p = 0.037). According to logistic regression, the most important predictors for high pain intensity were minor allele “G” of rs2227307 (OR 2.435, 95% CI 1.123–5.282), increasing age (OR 1.038, 95% CI 1.002–1.075), and female sex (OR 4.592, 95% CI 1.289–16.361). The explored gene polymorphisms were not significant risk factors for TMDp presence. These findings highlight the importance of genetic variations, particularly rs2227307, in understanding the diverse clinical manifestations of temporomandibular disorders. Full article
(This article belongs to the Special Issue New Insights into the Molecular Mechanisms of Chronic Pain)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>c</b>) Genotype distribution in control subjects (CTR) and temporomandibular disorders patients with myalgia (TMDp myalgia). (<b>d</b>–<b>f</b>) Genotype distribution in control subjects (CTR) and temporomandibular disorders patients with arthralgia (TMDp arthralgia).</p>
Full article ">
14 pages, 1510 KiB  
Article
Differences in the Interleukin Profiles in Inattentive ADHD Prepubertal Children Are Probably Related to Conduct Disorder Comorbidity
by Raquel González-Villén, María Luisa Fernández-López, Ana Checa-Ros, Pilar Tortosa-Pinto, Raquel Aguado-Rivas, Laura Garre-Morata, Darío Acuña-Castroviejo and Antonio Molina-Carballo
Biomedicines 2024, 12(8), 1818; https://doi.org/10.3390/biomedicines12081818 - 9 Aug 2024
Viewed by 480
Abstract
Inflammatory cytokines are involved in attention deficit hyperactivity disorder (ADHD), a highly prevalent neurodevelopmental disorder. To quantify the baseline levels of pro- and anti-inflammatory cytokines and their changes after methylphenidate (MPH), a total of 31 prepubertal children with ADHD were recruited and subclassified [...] Read more.
Inflammatory cytokines are involved in attention deficit hyperactivity disorder (ADHD), a highly prevalent neurodevelopmental disorder. To quantify the baseline levels of pro- and anti-inflammatory cytokines and their changes after methylphenidate (MPH), a total of 31 prepubertal children with ADHD were recruited and subclassified into only two ADHD presentations—ADHD attention deficit (n = 13) or ADHD combined (n = 18). The children were also screened for oppositional defiant conduct disorder (ODCD) and anxiety disorder. Blood samples were drawn at 09:00 and after 4.63 ± 1.87 months of treatment. Four pro-inflammatory cytokines (interleukin-1beta (IL-1β), IL-5, IL-6, tumor necrosis factor-alpha (TNF-α)) and three anti-inflammatory cytokines (IL-4, IL-10, IL-13) were measured using a Luminex® assay. For statistics, a factorial analysis was performed in Stata 15.1. Overall, there were no statistically significant differences in the interleukin (IL) values induced by treatment. When grouped by presentation, the differences were present almost exclusively in ADHD-AD, usually with a profile opposite to that observed in ADHD-C, and with interactions between comorbid factors, with IL-1β (p = 0.01) and IL-13 (p = 0.006) being the ones reaching the greatest statistical significance. These differences are probably related to the ODCD factor, and they disappear after treatment. In conclusion, the changes observed in cytokine levels in prepubertal children only in the ADHD-AD presentation are probably related to comorbidities (specifically ODCD) and are mitigated after treatment. Full article
Show Figures

Figure 1

Figure 1
<p>Study design scheme and timeline.</p>
Full article ">Figure 2
<p>Between-patient comparison of the four pro-inflammatory cytokines. Baseline interleukin values are compared between patients without oppositional defiant conduct disorder (ODCD-; in green) and patients with it (ODCD+; in orange) for each ADHD subtype: inattentive (AD) and combined (C). Data are reported as individual datapoints, as box plots (where the black bold line represents the median; hinges show the 25th and 75th percentiles; whiskers represent the 1.5 interquartile ranges beyond the hinges), and as data density estimates following the Gaussian kernel method. (<b>A</b>) Comparisons inf IL-1beta and IL-5; (<b>B</b>) comparisons in IL-6 and TNF-alpha.</p>
Full article ">Figure 2 Cont.
<p>Between-patient comparison of the four pro-inflammatory cytokines. Baseline interleukin values are compared between patients without oppositional defiant conduct disorder (ODCD-; in green) and patients with it (ODCD+; in orange) for each ADHD subtype: inattentive (AD) and combined (C). Data are reported as individual datapoints, as box plots (where the black bold line represents the median; hinges show the 25th and 75th percentiles; whiskers represent the 1.5 interquartile ranges beyond the hinges), and as data density estimates following the Gaussian kernel method. (<b>A</b>) Comparisons inf IL-1beta and IL-5; (<b>B</b>) comparisons in IL-6 and TNF-alpha.</p>
Full article ">Figure 3
<p>Between-patient comparison of the three anti-inflammatory cytokines. Baseline interleukin values are compared between patients without oppositional defiant conduct disorder (ODCD-; in green) and patients with it (ODCD+; in orange) for each ADHD subtype: inattentive (AD) and combined (C). Data are reported as individual datapoints, as box plots (where the black bold line represents the median; hinges show the 25th and 75th percentiles; whiskers represent the 1.5 interquartile ranges beyond the hinges), and as data density estimates following the Gaussian kernel method. (<b>A</b>) Comparisons in IL-4 and IL-10; (<b>B</b>) comparisons in IL-13.</p>
Full article ">
9 pages, 493 KiB  
Communication
Effects of Black Maca Supplementation on Isokinetic Muscle Function and Inflammation in Elite Athletes and Non-Athletes
by Eunjae Lee, Sunghwun Kang and Seung-Taek Lim
Appl. Sci. 2024, 14(16), 7005; https://doi.org/10.3390/app14167005 (registering DOI) - 9 Aug 2024
Viewed by 310
Abstract
This study aimed to examine the effects of black maca supplementation on isokinetic muscle function and inflammatory markers in athletes and to extend these findings to non-athletes. The study involved 24 male participants, including 16 elite athletes (soft tennis and table tennis players) [...] Read more.
This study aimed to examine the effects of black maca supplementation on isokinetic muscle function and inflammatory markers in athletes and to extend these findings to non-athletes. The study involved 24 male participants, including 16 elite athletes (soft tennis and table tennis players) and 8 non-athletes (university students). Participants consumed capsules containing 2.5 g of 100% concentrated black maca extract over a 12-week period. Isokinetic muscle performance and physical fitness (strength, muscular endurance, flexibility, power, agility, cardiovascular endurance) assessments were conducted at baseline and after 12 weeks of supplementation. Two-way within-factor ANOVA showed a significant group × time interaction for 120°/s flexor movements (p < 0.05). Paired t-tests demonstrated significant improvements in 30°/s and 120°/s extensor and flexor movements in both athlete groups (p < 0.05, p < 0.01). Similarly, significant enhancements were observed in the non-athletes for the 30°/s flexor and 120°/s extensor and flexor movements (p < 0.01). Furthermore, reductions in interleukin-6 (from 137.9 ± 8.8 to 132.7 ± 4.6, p < 0.05) and tumor necrosis factor-alpha (from 274.1 ± 13.4 to 264.2 ± 3.2, p < 0.05) were noted in the soft tennis group. The table tennis group also showed significant decreases in interleukin-6 (from 135.9 ± 4.7 to 131.3 ± 2.5, p < 0.01) and tumor necrosis factor-alpha (from 282.1 ± 19.2 to 267.0 ± 6.4, p < 0.05). No significant changes were observed in the non-athlete group. Black maca supplementation may enhance isokinetic muscle function in elite athletes by reducing muscle contraction fatigue and improving anti-inflammatory responses. Full article
(This article belongs to the Special Issue Sports Medicine, Exercise, and Health: Latest Advances and Prospects)
Show Figures

Figure 1

Figure 1
<p>Measurements of inflammation factors by group and time. ST, soft tennis; TT, table tennis; NA, non-athletes. Results were analyzed using a paired <span class="html-italic">t</span>-test: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Group × time interaction: ## <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
13 pages, 8455 KiB  
Article
Starvation and Inflammation Modulate Adipose Mesenchymal Stromal Cells’ Molecular Signature
by Simona Piccolo, Giulio Grieco, Caterina Visconte, Paola De Luca, Michela Taiana, Luigi Zagra, Enrico Ragni and Laura de Girolamo
J. Pers. Med. 2024, 14(8), 847; https://doi.org/10.3390/jpm14080847 (registering DOI) - 9 Aug 2024
Viewed by 226
Abstract
Mesenchymal stromal cells (MSCs) and their released factors (secretome) are intriguing options for regenerative medicine approaches based on the management of inflammation and tissue restoration, as in joint disorders like osteoarthritis (OA). Production strategy may modulate cells and secretome fingerprints, and for the [...] Read more.
Mesenchymal stromal cells (MSCs) and their released factors (secretome) are intriguing options for regenerative medicine approaches based on the management of inflammation and tissue restoration, as in joint disorders like osteoarthritis (OA). Production strategy may modulate cells and secretome fingerprints, and for the latter, the effect of serum removal by starvation used in clinical-grade protocols has been underestimated. In this work, the effect of starvation on the molecular profile of interleukin 1 beta (IL1β)-primed adipose-derived MSCs (ASCs) was tested by assessing the expression level of 84 genes related to secreted factors and 84 genes involved in defining stemness potential. After validation at the protein level, the effect of starvation modulation in the secretomes was tested in a model of OA chondrocytes. IL1β priming in vitro led to an increase in inflammatory mediators’ release and reduced anti-inflammatory potential on chondrocytes, features reversed by subsequent starvation. Therefore, when applying serum removal-based clinical-grade protocols for ASCs’ secretome production, the effects of starvation must be carefully considered and investigated. Full article
Show Figures

Figure 1

Figure 1
<p>ASCs’ immunophenotype: (<b>A</b>) ASCs were positive for MSC markers CD44, CD73, CD90, and CD105, and tissue-resident/early-passage ASC marker CD34. Plots illustrate the results from a representative donor. (<b>B</b>) No difference was observed under the analyzed conditions (N = 3).</p>
Full article ">Figure 2
<p>Correlation between samples and conditions: (<b>A</b>) Intra-group and (<b>B</b>) inter-group correlation analyses for the expression of inflammation-related genes; N = 3, mean ± SD. “<span class="html-italic">r</span>” Pearson values are shown. (<b>C</b>) PCA performed on normalized C<sub>t</sub> values for inflammation genes. The X and Y axes show principal component 1 and principal component 2, which explain 47.3% and 15.3% of the total variance. (<b>D</b>) Hierarchical clustering performed on normalized C<sub>t</sub> values for inflammation genes. Higher Ct means lower amount, and lower Ct means higher amount. (<b>E</b>) Intra-group and (<b>F</b>) inter-group correlation analyses for the expression of mesenchymal stem cell-related genes; N = 3, mean ± SD. “<span class="html-italic">r</span>” Pearson values are shown. (<b>G</b>) PCA performed on normalized Ct values for mesenchymal genes. The X and Y axes show principal component 1 and principal component 2, which explain 35.8% and 13.0% of the total variance. (<b>H</b>) Hierarchical clustering performed on normalized Ct values for mesenchymal genes. Higher Ct means lower amount, and lower Ct means higher amount.</p>
Full article ">Figure 3
<p>Correlation between samples and conditions for significantly modulated genes: (<b>A</b>) PCA performed on ln(FC + 1) values, with FC calculated vs. condition F. The X and Y axes show principal component 1 and principal component 2, which explain 87.7% and 9.7% of the total variance. (<b>B</b>) Hierarchical clustering performed on ln(FC + 1) values, with FC calculated vs. condition F.</p>
Full article ">Figure 4
<p>Quantitative analysis of factors modulated by IL1β and reversed by subsequent starvation. CCL2, CXCL2, IL6, IL8, CSF2, and CSF3 levels detected as pg/mL were measured by ELISA assays. In the absence of plots, the proteins were undetectable or below the lower limit of detection of the assay. Under ANOVA analysis, significance was set for <span class="html-italic">p</span>-value ≤ 0.05 (§ for <span class="html-italic">p</span>-value ≤ 0.1, * ≤ 0.05, ** ≤ 0.01, *** ≤ 0.001 and **** ≤ 0.0001. N = 3).</p>
Full article ">Figure 5
<p>Correlation between conditions for significantly modulated genes in inflamed chondrocytes treated with secretomes: (<b>A</b>) Inter-group correlation analysis of the modulation of OA-related genes in chondrocytes treated with IL1β alone or with IL1β and secretomes with respect to untreated (CTRL) cells. “<span class="html-italic">r</span>” Pearson values are shown. (<b>B</b>) PCA performed on ln(FC + 1) values, with FC calculated vs. condition CTRL. The X and Y axes show principal component 1 and principal component 2, which explain 91.6% and 6.6% of the total variance, respectively. (<b>C</b>) Hierarchical clustering performed on ln(FC + 1) values, with FC calculated vs. condition CTRL. The scale bar’s maximum for ln(FC + 1) values was set to 2.5.</p>
Full article ">
11 pages, 2026 KiB  
Article
Major Adverse Cardiovascular Events: The Importance of Serum Levels and Haplotypes of the Anti-Inflammatory Cytokine Interleukin 10
by Susanne Schulz, Leonie Reuter, Alexander Navarrete Santos, Kerstin Bitter, Selina Rehm, Axel Schlitt and Stefan Reichert
Biomolecules 2024, 14(8), 979; https://doi.org/10.3390/biom14080979 - 9 Aug 2024
Viewed by 216
Abstract
Background: Cardiovascular diseases (CVDs) represent major medical and socio-economic challenges worldwide. There is substantial evidence that CVD is closely linked to inflammatory changes triggered by a complex cytokine network. In this context, interleukin 10 (IL-10) plays an important role as a pleiotropic cytokine [...] Read more.
Background: Cardiovascular diseases (CVDs) represent major medical and socio-economic challenges worldwide. There is substantial evidence that CVD is closely linked to inflammatory changes triggered by a complex cytokine network. In this context, interleukin 10 (IL-10) plays an important role as a pleiotropic cytokine with an anti-inflammatory capacity. In this study (a substudy of ClinTrials.gov, identifier: NCT01045070), the prognostic relevance of IL-10 levels and IL-10 haplotypes (rs1800896/rs1800871/rs1800872) was assessed regarding adverse cardiovascular outcomes (combined endpoint: myocardial infarction, stroke/transient ischemic attack, cardiac death and death according to stroke) within a 10-year follow-up. Patients and methods: At baseline, 1002 in-patients with CVD were enrolled. Serum levels of IL-10 were evaluated utilizing flow cytometry (BD™ Cytometric Bead Array). Haplotype analyses were carried out by polymerase chain reactions with sequence-specific primers (PCR-SSP). Results: In a survival analysis, IL-10 haplotypes were not proven to be cardiovascular prognostic factors in a 10-year follow-up (Breslow test: p = 0.423). However, a higher IL-10 level was associated with adverse cardiovascular outcomes (Breslow test: p = 0.047). A survival analysis considering adjusted hazard ratios (HRs) could not confirm this correlation (Cox regression: adjusted HR = 1.26, p = 0.168). Conclusion: In the present study, an elevated IL-10 level but not IL-10 haplotypes was linked to adverse cardiovascular outcomes (10-year follow-up) in a cohort of CVD patients. Full article
(This article belongs to the Special Issue Biomarkers of Cardiovascular and Cerebrovascular Diseases)
Show Figures

Figure 1

Figure 1
<p>Study design.</p>
Full article ">Figure 2
<p>Relationship between IL-10 serum levels and genetic variants in IL-10 gene: (<b>a</b>) rs1800896; (<b>b</b>) rs1800871; (<b>c</b>) rs1800872; (<b>d</b>) haplotypes of rs1800896/rs1800871/rs1800872 (IQR: interquartile range).</p>
Full article ">Figure 3
<p>Kaplan–Meier survival curve and Breslow test. Impact of IL-10 level (&lt;1.5 pg/mL vs. ≥1.5 pg/mL) on combined endpoint (myocardial infarction, cardiovascular death, stroke/transient ischemic attack and death from stroke) in 10-year follow-up.</p>
Full article ">Figure 4
<p>Kaplan–Meier survival curve and Breslow test. Impact of IL-10 level (IQR25: &lt;0.45 pg/mL vs. IQR75: &gt;1.48 pg/mL) on combined endpoint (myocardial infarction, cardiovascular death, stroke/transient ischemic attack and death from stroke) in 10-year follow-up (IQR: interquartile range).</p>
Full article ">Figure 5
<p>Kaplan–Meier survival curve including Breslow test: evaluation of genetic variants in IL-10 gene as prognostic factors for 10-year combined endpoint (myocardial infarction, cardiovascular death, stroke/transient ischemic attack and death from stroke): (<b>a</b>) rs1800896; (<b>b</b>) rs1800871; (<b>c</b>) rs1800872; (<b>d</b>) haplotypes of rs1800896/rs1800871/rs1800872.</p>
Full article ">
Back to TopTop