[go: up one dir, main page]

 
 
ijms-logo

Journal Browser

Journal Browser

Reproductive Immunology: Cellular and Molecular Biology

A special issue of International Journal of Molecular Sciences (ISSN 1422-0067). This special issue belongs to the section "Biochemistry".

Deadline for manuscript submissions: closed (31 January 2019) | Viewed by 108733

Special Issue Editors


E-Mail Website
Guest Editor
Department of Obstetrics and Gynecology, Medical School, University of Crete, 71003 Heraklion, Greece
Interests: implantation; receptivity; endometrium; RIF; early pregnancy; trophoblast
Special Issues, Collections and Topics in MDPI journals

Special Issue Information

Dear Colleagues,

Reproductive Immunology in the 21st century deals still with a problem known for decades—the fetus as semi allograft and its response to the maternal immune system. Therefore, there is a strong need to solve problems like spontaneous and recurrent miscarriages and in addition repeated implantation failure.

In addition, socio-economical changes in an aging society are an additional challenge especially for the reproductive medicine specialist. Although highly developed in vitro fertilization techniques are available, many couples still face the problem of childlessness.

A quite new player in the field is the microbiome of the reproductive tract. For decades is was believed that the uterus is sterile but an up to date analyses revealed that we not only have a vaginal microbiome but also a cervical, uterine, male and even a placental microbiome.

Therefore we would like to invite our colleagues to submit articles that deal with cellular and molecular mechanism in the field of reproductive immunology to this Special Issue.

Prof. Dr. Udo Jeschke
Prof. Dr. Antonis Makrigiannakis
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. International Journal of Molecular Sciences is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. There is an Article Processing Charge (APC) for publication in this open access journal. For details about the APC please see here. Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • spontaneous and recurrent miscarriage
  • repeated implantation failure
  • maternal immune cells
  • microbiome
  • assisted reproduction

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Related Special Issues

Published Papers (20 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

Jump to: Review

15 pages, 3962 KiB  
Article
Immunohistochemical Study on the Expression of G-CSF, G-CSFR, VEGF, VEGFR-1, Foxp3 in First Trimester Trophoblast of Recurrent Pregnancy Loss in Pregnancies Treated with G-CSF and Controls
by Fabio Scarpellini, Francesca Gioia Klinger, Gabriele Rossi and Marco Sbracia
Int. J. Mol. Sci. 2020, 21(1), 285; https://doi.org/10.3390/ijms21010285 - 31 Dec 2019
Cited by 23 | Viewed by 4243
Abstract
Background: Recurrent Pregnancy Loss (RPL) is a syndrome recognizing several causes, and in some cases the treatment with Granulocyte Colony Stimulating Factor (G-CSF) may be successful, especially when karyotype of the previous miscarriage showed no embryo chromosomal abnormalities. In order to evaluate the [...] Read more.
Background: Recurrent Pregnancy Loss (RPL) is a syndrome recognizing several causes, and in some cases the treatment with Granulocyte Colony Stimulating Factor (G-CSF) may be successful, especially when karyotype of the previous miscarriage showed no embryo chromosomal abnormalities. In order to evaluate the effects of G-CSF treatment on the decidual and trophoblast expression of G-CSF and its receptor, VEGF and its receptor and Foxp3, specific marker of putative Tregs we conducted an immunohistochemical study. Methods: This study was conducted on three groups of patients for a total of 38 women: in 8 cases decidual and trophoblast tissue were obtained from 8 women with unexplained RPL treated with G-CSF that miscarried despite treatment; in 15 cases the tissue were obtained from 15 women with unexplained RPL no treated; 15 cases of women who underwent voluntary pregnancy termination were used as controls. Tissue collected from these patients were used for immunohistochemistry studies testing the expression of G-CSF, G-CSFR, VEGF, VEGFR-1 and Foxp3. Results: G-CSF treatment increased the concentration of cells expressing Foxp3, specific marker for Tregs, in the decidua, whereas in no treated RPL a reduction of these cells was found when compared to controls. Furthermore, G-CSF treatment increased the expression of G-CSF and VEGF in the trophoblast. Conclusions: Our study showed that G-CSF treatment increased the number of decidual Treg cells in RPL patients as well as the expression of G-CSF and VEGF in villus trophoblast. These finding may explain the effectiveness of this treatment in RPL, probably regulating the maternal immune response through Tregs recruitment in the decidua, as well as stimulating trophoblast growth. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>In decidua of Granulocyte Colony Stimulating Factor (G-CSF) treated samples (<b>A</b>), in the stroma there was a remarkable increase in the number of Foxp3 cells with respect to other two groups, epithelial cells showed a moderate staining (brown color) (400×). In decidua of no treated Recurrent Pregnancy Loss (RPL) samples (<b>B</b>), in the stroma there was a strong reduction in the number of Foxp3 cells, epithelial cells showed a moderate staining (brown color) (400×). In decidua of control samples (<b>C</b>), in the stroma there was a number of cells positive to Foxp3 cells, higher than in no-treated RPL and lower than in G-CSF treated samples, epithelial cells showed a moderate staining (brown color) (400×). The graph (<b>G</b>) showed percent of immunohistochemical positive cells in the stroma to Foxp3 (charts display median and quartiles with whiskers showing the range): There was a statistically significant difference between G-CSF vs. RPL (* <span class="html-italic">p</span> &lt; 0.001), G-CSF vs. Control (** <span class="html-italic">p</span> &lt; 0.01) and RPL vs. Control (*** <span class="html-italic">p</span> &lt; 0.01). Foxp3 expression in decidua and trophoblast of first trimester pregnancy. In trophoblast of GCS-F treated group (<b>D</b>), no treated RPL (<b>E</b>) and Control (<b>F</b>) there was no staining at all for Foxp3 (400×).</p>
Full article ">Figure 2
<p>G-CSF expression in decidua and trophoblast of first trimester pregnancy. In decidua of G-CSF treated group (<b>A</b>), in the no treated RPL group (<b>B</b>) and Control pregnancies (<b>C</b>) there was the same staining levels in the epithelial cells but no in the stroma for G-CSF (brown color) (400×). In trophoblast of G-CSF treated samples (<b>D</b>), the syncytiotrophoblast was positive to G-CSF (brown color) (400×). In trophoblast of RPL no-treated samples (<b>E</b>), the syncytiotrophoblast was weakly positive to G-CSF (400×). In trophoblast of Control samples (<b>F</b>), the syncytiotrophoblast was positive to G-CSF (brown color) similar to G-CSF treated samples (400×). The graph (<b>G</b>) showed immunohistochemical staining semi quantitative HSCORE for G-CSF (Charts display median and quartiles with whiskers showing the range): There was statistically significant differences between G-CSF vs. RPL (* <span class="html-italic">p</span> &lt; 0.001) and RPL vs. Control (** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>G-CSFR expression in decidua and trophoblast of first trimester pregnancy. In decidua of G-CSF treated group (<b>A</b>), in the no treated RPL group (<b>B</b>) and Control pregnancies (<b>C</b>) there was the same staining levels in the epithelial cells but no in the stroma for G-CSFR (brown color) (400×). In trophoblast of G-CSF treated samples (<b>D</b>), the syncytiotrophoblast was moderately positive to G-CSFR (brown colour) (400×). In trophoblast of RPL no-treated samples (<b>E</b>), the syncytiotrophoblast was strongly positive to G-CSFR, more than in other two groups (400×). In trophoblast of Control samples (<b>F</b>), the syncytiotrophoblast was moderately positive to G-CSFR (brown color) similar to G-CSF treated samples (400×). The graph (<b>G</b>) showed immunohistochemical staining semi quantitative HSCORE for G-CSFR (charts display median and quartiles with whiskers showing the range): There was a statistically significant differences between G-CSF vs. RPL (* <span class="html-italic">p</span> &lt; 0.001) and RPL vs. Control (** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>VEGF expression in decidua and trophoblast of first trimester pregnancy. In decidua of G-CSF treated group (<b>A</b>), in the no treated RPL group (<b>B</b>) and Control pregnancies (<b>C</b>) there was the same staining levels in the epithelial cells but no in the stroma for VEGF (brown color) (400×). In trophoblast of G-CSF treated samples (<b>D</b>), the syncytiotrophoblast was strongly positive to VEGF (brown color) (400×). In trophoblast of RPL no treated samples (<b>E</b>), the syncytiotrophoblast was weakly positive to VEGF (400×). In trophoblast of Control samples (<b>F</b>), the syncytiotrophoblast was strongly positive to VEGF (brown color) similar to G-CSF treated samples (400×). The graph (<b>G</b>) showed immunohistochemical staining semi quantitative HSCORE for VEGF (Charts display median and quartiles with whiskers showing the range): There was a statistically significant differences between G-CSF vs. RPL (* <span class="html-italic">p</span> &lt; 0.001) and RPL vs. Control (** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>VEGFR-1 expression in decidua and trophoblast of first trimester pregnancy. In decidua of G-CSF treated group (<b>A</b>), in the no treated RPL group (<b>B</b>) and Control pregnancies (<b>C</b>) there was the same staining levels in the epithelial cells but no in the stroma for VEGFR-1 (brown color) (400×). In trophoblast of G-CSF treated samples (<b>D</b>), the syncytiotrophoblast was moderately positive to VEGFR-1 (brown color) (400×). In trophoblast of RPL no-treated samples (<b>E</b>), the syncytiotrophoblast was strongly positive to VEGFR-1, more than in other two groups (400×). In trophoblast of Control samples (<b>F</b>), the syncytiotrophoblast was moderately positive to VEGFR-1 (brown color) similar to G-CSF treated samples (400×). The graph (<b>G</b>) showed immunohistochemical staining semi quantitative HSCORE for VEGFR-1 (Charts display median and quartiles with whiskers showing the range): There was statistically significant differences between G-CSF vs. RPL (* <span class="html-italic">p</span> &lt; 0.001) and RPL vs. Control (** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
17 pages, 4267 KiB  
Article
Fucoxanthin-Rich Brown Algae Extract Improves Male Reproductive Function on Streptozotocin-Nicotinamide-Induced Diabetic Rat Model
by Zwe-Ling Kong, Sabri Sudirman, Yu-Chun Hsu, Chieh-Yu Su and Hsiang-Ping Kuo
Int. J. Mol. Sci. 2019, 20(18), 4485; https://doi.org/10.3390/ijms20184485 - 11 Sep 2019
Cited by 31 | Viewed by 5176
Abstract
Hypogonadism and oxidative stress are occurring commonly in men with diabetes and associated male infertility. This study aimed to investigate the capability of anti-oxidative and anti-inflammatory properties of fucoxanthin as well as to evaluate its protective effects on male reproduction in diabetic rats. [...] Read more.
Hypogonadism and oxidative stress are occurring commonly in men with diabetes and associated male infertility. This study aimed to investigate the capability of anti-oxidative and anti-inflammatory properties of fucoxanthin as well as to evaluate its protective effects on male reproduction in diabetic rats. The RAW 264.7 macrophage cells were used to evaluate the anti-oxidative and anti-inflammatory activity. Thirty male Sprague-Dawley rats were induced by streptozotocin-nicotinamide for a diabetes model and fed either with three different doses of fucoxanthin (13, 26, and 65 mg/kg) or rosiglitazone (0.571 mg/kg) for four weeks. The fucoxanthin significantly inhibited nitric oxide production and reduced reactive oxygen species level in lipopolysaccharide-induced RAW 264.7 cells. In the animal study, fucoxanthin administration improved insulin resistance, restored sperm motility, decreased abnormal sperm number, and inhibited lipid peroxidation. Moreover, it restored GPR54 and SOCS-3 mRNA expression in the hypothalamus and recovered luteinizing hormone level, as well as the testosterone level. In conclusion, fucoxanthin not only possessed antioxidant and anti-inflammatory properties but also decreased the diabetes signs and symptoms as well as improved spermatogenesis and male reproductive function. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effects of fucoxanthin (FXN) on (<b>A</b>) RAW 264.7 macrophage cells viability, (<b>B</b>) nitric oxide, (<b>C</b>) superoxide, and (<b>D</b>) hydrogen peroxide productions. The cells were stimulated by lipopolysaccharides (LPS, 1 µg/mL). Data are shown as the mean ± S.D. of three independent experiments. The values with different letters (a–f) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) as analyzed by Duncan’s multiple range test. Normal, unstimulated cells; control, untreated cells.</p>
Full article ">Figure 2
<p>Effects of fucoxanthin on plasma glucose after treatment for four weeks. (<b>A</b>) Oral glucose tolerance test (OGTT) and (<b>B</b>) area under the curve (AUC) of plasma glucose. Data are shown as the mean ± S.D. (<span class="html-italic">n</span> = 5). The values with different letters (a–c) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) as analyzed by Duncan’s multiple range test. C, control; DM, diabetes; DMF, diabetes treated with fucoxanthin; DMR, diabetes treated with rosiglitazone.</p>
Full article ">Figure 3
<p>Effects of fucoxanthin on (<b>A</b>) reactive oxygen species (ROS) and (<b>B</b>) superoxide productions in rat sperm after treatment for four weeks. Data are shown as the mean ± S.D. (<span class="html-italic">n</span> = 5). The values with different letters (a–c) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) as analyzed by Duncan’s multiple range test. C, control; DM, diabetes; DMF, diabetes treated with fucoxanthin; DMR, diabetes treated with rosiglitazone.</p>
Full article ">Figure 4
<p>Effects of fucoxanthin on malondialdehyde (MDA) level in rat plasma, testis, and sperm after treatment for four weeks. Data are shown as the mean ± S.D. (<span class="html-italic">n</span> = 5). The values with different letters (a–c) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) as analyzed by Duncan’s multiple range test. C, control; DM, diabetes; DMF, diabetes treated with fucoxanthin; DMR, diabetes treated with rosiglitazone.</p>
Full article ">Figure 5
<p>Effects of fucoxanthin on proinflammatory cytokines after treatment for four weeks. (<b>A</b>) Tumor necrosis factor (TNF)-α level in plasma and testis, and (<b>B</b>) Interleukin (IL)-6 level in plasma and testis. Data are shown as the mean ± S.D. (<span class="html-italic">n</span> = 5). The values with different letters (a–c) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) as analyzed by Duncan’s multiple range test. C, control; DM, diabetes; DMF, diabetes treated with fucoxanthin; DMR, diabetes treated with rosiglitazone.</p>
Full article ">Figure 6
<p>Effects of fucoxanthin on SOCS-3 mRNA expression in rat hypothalamus after treatment for four weeks. Data are showed as the mean ± S.D. (<span class="html-italic">n</span> = 5). The values with different letters (a–c) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) as analyzed by Duncan’s multiple range test. C, control; DM, diabetes; DMF, diabetes treated with fucoxanthin; DMR, diabetes treated with rosiglitazone; SOCS-3, suppressors of cytokine signaling-3.</p>
Full article ">Figure 7
<p>Effects of fucoxanthin on relative (<b>A</b>) Kiss1 and (<b>B</b>) GPR54 mRNA expression in rat hypothalamus after treatment for four weeks. Data are shown as the mean ± S.D. (<span class="html-italic">n</span> = 5). The values with different letters (a–c) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) as analyzed by Duncan’s multiple range test. C, control; DM, diabetes; DMF, diabetes treated with fucoxanthin; DMR, diabetes treated with rosiglitazone; GPR54, G-protein coupling receptor (Kiss1 receptor).</p>
Full article ">Figure 8
<p>Effects of fucoxanthin on (<b>A</b>) testicular morphology and (<b>B</b>) thickness of seminiferous tubule diameter after treatment for four weeks. Data are shown as the mean ± S.D. (<span class="html-italic">n</span> = 5). The values with different letters (a–c) represent significant differences (<span class="html-italic">p</span> &lt; 0.05) as analyzed by Duncan’s multiple range test. Black arrow, Leydig cell; white arrow, Sertoli cell; C, control; DM, diabetes; DMF, diabetes treated with fucoxanthin; DMR, diabetes treated with rosiglitazone.</p>
Full article ">Figure 9
<p>The flowchart of fucoxanthin treatment against streptozotocin-nicotinamide (STZ-NA)-induced diabetes Sprague-Dawley (SD) rat model. DM, diabetes; DMF, diabetes treated with fucoxanthin; DMR, diabetes treated with rosiglitazone.</p>
Full article ">
15 pages, 7292 KiB  
Article
Transcriptomics and Immunological Analyses Reveal a Pro-Angiogenic and Anti-Inflammatory Phenotype for Decidual Endothelial Cells
by Chiara Agostinis, Elisa Masat, Fleur Bossi, Giuseppe Ricci, Renzo Menegazzi, Letizia Lombardelli, Gabriella Zito, Alessandro Mangogna, Massimo Degan, Valter Gattei, Marie-Pierre Piccinni, Uday Kishore and Roberta Bulla
Int. J. Mol. Sci. 2019, 20(7), 1604; https://doi.org/10.3390/ijms20071604 - 31 Mar 2019
Cited by 11 | Viewed by 4406
Abstract
Background: In pregnancy, excessive inflammation and break down of immunologic tolerance can contribute to miscarriage. Endothelial cells (ECs) are able to orchestrate the inflammatory processes by secreting pro-inflammatory mediators and bactericidal factors by modulating leakiness and leukocyte trafficking, via the expression of adhesion [...] Read more.
Background: In pregnancy, excessive inflammation and break down of immunologic tolerance can contribute to miscarriage. Endothelial cells (ECs) are able to orchestrate the inflammatory processes by secreting pro-inflammatory mediators and bactericidal factors by modulating leakiness and leukocyte trafficking, via the expression of adhesion molecules and chemokines. The aim of this study was to analyse the differences in the phenotype between microvascular ECs isolated from decidua (DECs) and ECs isolated from human skin (ADMECs). Methods: DECs and ADMECs were characterized for their basal expression of angiogenic factors and adhesion molecules. A range of immunological responses was evaluated, such as vessel leakage, reactive oxygen species (ROS) production in response to TNF-α stimulation, adhesion molecules expression and leukocyte migration in response to TNF-α and IFN-γ stimulation. Results: DECs produced higher levels of HGF, VEGF-A and IGFBP3 compared to ADMECs. DECs expressed adhesion molecules, ICAM-2 and ICAM-3, and a mild response to TNF-α was observed. Finally, DECs produced high levels of CXCL9/MIG and CXCL10/IP-10 in response to IFN-γ and selectively recruited Treg lymphocytes. Conclusion: DEC phenotype differs considerably from that of ADMECs, suggesting that DECs may play an active role in the control of immune response and angiogenesis at the foetal-maternal interface. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Phenotypic characterization of DECs and ADMECs. (<b>A</b>) Immunofluorescence analysis of vWF, VE-cadherin and vimentin on isolated and cultured DECs and ADMECs. Original magnification: 200×. (<b>B</b>) RT-qPCR of VEGF-A, HGF, IGFBP3, ICAM-2 and ICAM-3 genes differentially expressed by DECs and ADMECs. The data represent the mean ± SD of triplicate samples from five separate experiments, * <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. (<b>C</b>) Evaluation of the production of VEGF-A, HGF and IGFBP3 proteins in the supernatants of a confluent monolayer of DECs and ADMECs after 4 h of culture using a commercial ELISA kit. The data represent the mean ± SD of triplicate samples from five separate experiments. * <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. (<b>D</b>) Cytofluorimetric analysis for the expression of ICAM-2 and ICAM-3 in basal condition of freshly isolated DECs and ADMECs. The ECs were incubated with PE-conjugated mouse anti-human ICAM-2 and ICAM-3 mAb. PE-conjugated isotype-matched IgG2a or IgG1 were used as negative control, respectively. Data are represented as mean ± SD of the Mean Fluorescence Intensity (MFI) of five separate experiments.</p>
Full article ">Figure 2
<p>Permeabilizing activity of endothelial cells to classical vasoactive stimuli. The permeabilizing activity was evaluated kinetically, after 5 (<b>A</b>), 15 (<b>B</b>) and 30 min (<b>C</b>) adding PAF (<b>D</b>), HIS (<b>E</b>) or BK (<b>F</b>), to the upper chamber of the TW, measuring the amount of FITC-labeled BSA that leaked through a monolayer of endothelial cells into the lower chamber. The data represent the mean ± SD of duplicate samples from four separate experiments * <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.</p>
Full article ">Figure 3
<p>Production of intracellular ROS by endothelial cells. ROS production by endothelial cells exposed to TNF-α (<b>A</b>) or HIS (<b>B</b>). Following pre-treatment with TNF-α (100 ng/mL) or histamine (HIS; 0.1 µM), ECs were stained with Ampliflu Red for the evaluation of H<sub>2</sub>O<sub>2</sub> production after 30, 60 120, 180 and 240 min. (<b>C</b>) Histograms represent the production of intracellular ROS by endothelial cells stimulated with TNF-α or HIS after 120 min. The data represent the mean ± SD of triplicate samples from five separate experiments; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Secretion of chemokines and surface-expression of adhesion molecules by ECs stimulated with TNF-α. CXCL8/IL-8 (<b>A</b>), CCL2/MCP-1 (<b>B</b>), CCL3/MIP-1α (<b>C</b>) and CCL5/RANTES (<b>D</b>) production in DEC and ADMEC supernatants after 4 h incubation with TNF-α was measured using a beads-based multiplex immunoassay (Luminex<sup>®</sup>). (<b>E</b>) ELISA on the whole cells for the expression of ICAM-1, VCAM-1 or E-Selectin on DEC and ADMEC plasma membrane after 4 h or 18 h incubation with TNF-α. The data represent the mean ± SD of triplicate samples from five separate experiments * <span class="html-italic">p</span> &lt; 0.01; ** <span class="html-italic">p</span> &lt; 0.005. (<b>F</b>) Trans-endothelial migration of Lympho-Monocytes (LM) across untreated and TNF-α-treated DEC and ADMEC: migration is shown as the number of migrated cells. The data represent the mean ± SD of triplicate samples from three separate experiments * <span class="html-italic">p</span> &lt; 0.05; n.s. not significant.</p>
Full article ">Figure 5
<p>Secretion of chemokines and trans-endothelial migration of LM through EC stimulated with IFN-γ. The production of CXCL10/IP-10 (<b>A</b>) and CXCL9/MIG (<b>B</b>) in DEC and ADMEC supernatants after 4 h incubation with IFN-γ was measured using a beads-based multiplex immunoassay (Luminex<sup>®</sup>). The data represent the mean ± SD of triplicate samples from five separate experiments * <span class="html-italic">p</span> &lt; 0.01. (<b>C</b>,<b>D</b>) Trans-endothelial migration of LM across untreated and IFN-γ-treated DEC and ADMEC. (<b>C</b>) Representative dot plots for flow cytometry of trans-endothelial migrated LM, stained for CD3 and FoxP3. (<b>D</b>) Quantitation of the percent of total migrated LM cells, positive for both CD3 and FoxP3 (CD3<sup>+</sup>FoxP3<sup>+</sup>) by flow cytometry. Migration was presented as the percentage of CD45 migrated cells. The data represent the mean ± SD of duplicate samples from three separate experiments * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
13 pages, 2685 KiB  
Article
Testis-Specific SEPT12 Expression Affects SUN Protein Localization and is Involved in Mammalian Spermiogenesis
by Chung-Hsin Yeh, Ya-Yun Wang, Shi-Kae Wee, Mei-Feng Chen, Han-Sun Chiang, Pao-Lin Kuo and Ying-Hung Lin
Int. J. Mol. Sci. 2019, 20(5), 1163; https://doi.org/10.3390/ijms20051163 - 7 Mar 2019
Cited by 7 | Viewed by 3682
Abstract
Male infertility is observed in approximately 50% of all couples with infertility. Intracytoplasmic sperm injection (ICSI), a conventional artificial reproductive technique for treating male infertility, may fail because of a severe low sperm count, immotile sperm, immature sperm, and sperm with structural defects [...] Read more.
Male infertility is observed in approximately 50% of all couples with infertility. Intracytoplasmic sperm injection (ICSI), a conventional artificial reproductive technique for treating male infertility, may fail because of a severe low sperm count, immotile sperm, immature sperm, and sperm with structural defects and DNA damage. Our previous studies have revealed that mutations in the septin (SEPT)-coding gene SEPT12 cause teratozoospermia and severe oligozoospermia. These spermatozoa exhibit morphological defects in the head and tail, premature chromosomal condensation, and nuclear damage. Sperm from Sept12 knockout mice also cause the developmental arrest of preimplantation embryos generated through in vitro fertilization and ICSI. Furthermore, we found that SEPT12 interacts with SPAG4, a spermatid nuclear membrane protein that is also named SUN4. Loss of the Spag4 allele in mice also disrupts the integration nuclear envelope and reveals sperm head defects. However, whether SEPT12 affects SPAG4 during mammalian spermiogenesis remains unclear. We thus conducted this study to explore this question. First, we found that SPAG4 and SEPT12 exhibited similar localizations in the postacrosomal region of elongating spermatids and at the neck of mature sperm through isolated murine male germ cells. Second, SEPT12 expression altered the nuclear membrane localization of SPAG4, as observed through confocal microscopy, in a human testicular cancer cell line. Third, SEPT12 expression also altered the localizations of nuclear membrane proteins: LAMINA/C in the cells. This effect was specifically due to the expression of SEPT12 and not that of SEPT1, SEPT6, SEPT7, or SEPT11. Based on these results, we suggest that SEPT12 is among the moderators of SPAG4/LAMIN complexes and is involved in the morphological formation of sperm during mammalian spermiogenesis. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Colocalization of SEPT12 and SPAG4 during murine spermiogenesis. The testis collected from C57BL/6 mice (<span class="html-italic">n</span> = 3) were separated according to the male germ cells’ stages. The images at different developmental stages were captured, some of which are presented in the Figure. From left to right: DAPI staining (blue), SEPT12 signal (green), SPAG4 (red), and merging of DAPI, SEPT12, and SPAG4 at the elongating spermatids (<b>A</b>), elongated spermatids (<b>B</b>), and mature sperm (<b>C</b>). (<b>A</b>) Arrows indicate signals within the edges of the acrosome and manchette. Arrowheads indicate signals located at the manchette. (<b>B</b>) Arrowheads indicate signals within the manchette. (<b>C</b>) Arrows indicate signals located at the neck and annulus of mature sperm. Arrowheads indicate SPAG4 signals around the sperm head. Scale bar: 5μm.</p>
Full article ">Figure 2
<p>SEPT12 overexpression disturbs SPAG4 localization around the NE in the NT2/D1 male germ cell line. (<b>A</b>) Western blot analysis of NT2D1 cells transfected with the pEGFP-empty vector (Lane1; GFP-Vector), pEGFP-SEPT12 (Lane2; SEPT12-GFP), pFLAGP-empty vector (Lane3; FLAG-Vector), pFLAG-SPAG4 (Lane4; FLAG-SPAG4), a mixture of the pEGFP-empty vector and pFLAG-SPAG4 vector (Lane5; GFP-Vector + FLAG-SPAG4), and a mixture of the pEGFP-SEPTIN12 vector and pFLAG-SPAG4 vector (Lane6; SEPT12-GFP + FLAG-SPAG4) using the anti-EGFP and anti-FLAG antibodies. (<b>B</b>) Immunofluorescence staining of NT2/D1 cells cotransfected with the pEGFP-SEPTIN12 vector and pFLAG-SPAG4 vector using DAPI (blue) (a), anti-EGFP antibody (green) (b) and anti-FLAG antibody (red) (c); (d) image obtained after merging the images in (a), (b), and (c). Magnification ×400 in (a–d). The arrow indicates the cells transfected with the pEGFP-SEPTIN12 vector (green) and pFLAG-SPAG4 vector (red). The arrowhead indicates the cells that were transfected only with the pFLAG-SPAG4 vector (red). Scale bar: 10 µm. (<b>C</b>) Quantification of the disorganization of SPAG4 in the NT2D1 cells transfected with a mixture of the pEGFP-empty vector and pFLAG-SPAG4 vector (Bar 1; GFP-Vector + FLAG-SPAG4) and a mixture of the pEGFP-SEPTIN12 vector and pFLAG-SPAG4 vector (Bar 2; SEPT12-GFP + FLAG-SPAG4). At least 100 transfected cells were counted in each experiment. Two-tailed Student <span class="html-italic">t</span> test; error bars indicate ± standard error of mean (*** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 3
<p>Effects of SEPT1/6/7/11 on SPAG4 localization around the NE in the NT2/D1 male germ cell line (<b>A</b>) Western blot analysis of NT2/D1 cells transfected with the pEGFP-vector, pEGFP-SEPT1, pEGFP-SEPT7, pEGFP-SEPT11, or pEGFP-SEPT6 with pFLAG-SPAG4 using the anti-GFP antibody (anti-GFP Ab) and anti-FLAG antibody (anti-FLAG Ab). (<b>B</b>) Fold changes in disrupted SPAG4 localization in transfected cells with the pEGFP-vector, pEGFP-SEPT12, pEGFP-SEPT1, pEGFP-SEPT6, pEGFP-SEPT7, or pEGFP-SEPT11 with pFLAG-SPAG4. The height of the boxes represents the mean of values obtained from three independent experiments. At least 100 transfected cells were counted in each experiment (*** <span class="html-italic">p</span> &lt; 0.001, Student <span class="html-italic">t</span> test).</p>
Full article ">Figure 4
<p>SEPT12 overexpression disturbs the morphology of the LAMIN in the NT2/D1 male germ cell line. SEPT12 signals are around NE (<b>A</b>) or cytoplasm (<b>B</b>). Immunofluorescence staining with the anti-LAMIN antibody in the NT2D1 cells cotransfected with the pEGFP-SEPTIN12 and pFLAG-SPAG4 vectors. (a) Anti-EGFP antibody (green), (b) anti-FLAG antibody (pink), (c) and anti-LAMIN antibody (red); (d) image obtained after merging the images in (a), (b), and (c). Scale bar: 10 µm.</p>
Full article ">Figure 5
<p>Effects of SEPT1/6/7/11 overexpression on the morphology of the NE in the NT2/D1 male germ cell line. Immunofluorescence staining with the anti-LAMIN antibody in the NT2D1 cells cotransfected with the pEGFP-empty (<b>A</b>), SEPT1 (<b>B</b>), SEPT6 (<b>C</b>), SEPT7 (<b>D</b>), or SEPT11 (<b>E</b>) vectors and pFLAG-SPAG4 vector. (a) Anti-EGFP antibody (green), (b) anti-FLAG antibody (pink), and (c) anti-LAMIN antibody (red); (d) image obtained after merging the images in (a), (b), and (c). Scale bar: 10 µm.</p>
Full article ">
15 pages, 904 KiB  
Article
Reactive Species Interactome Alterations in Oocyte Donation Pregnancies in the Absence and Presence of Pre-Eclampsia
by Manon Bos, Mirthe H. Schoots, Bernadette O. Fernandez, Monika Mikus-Lelinska, Laurie C. Lau, Michael Eikmans, Harry van Goor, Sanne J. Gordijn, Andreas Pasch, Martin Feelisch and Marie-Louise P. van der Hoorn
Int. J. Mol. Sci. 2019, 20(5), 1150; https://doi.org/10.3390/ijms20051150 - 6 Mar 2019
Cited by 10 | Viewed by 3041
Abstract
In pregnancy, maternal physiology is subject to considerable adaptations, including alterations in cardiovascular and metabolic function as well as development of immunological tolerance towards the fetus. In an oocyte donation pregnancy, the fetus is fully allogeneic towards the mother, since it carries both [...] Read more.
In pregnancy, maternal physiology is subject to considerable adaptations, including alterations in cardiovascular and metabolic function as well as development of immunological tolerance towards the fetus. In an oocyte donation pregnancy, the fetus is fully allogeneic towards the mother, since it carries both oocyte donor antigens and paternal antigens. Therefore, oocyte donation pregnancies result in an immunologically challenging pregnancy, which is reflected by a higher-than-normal risk to develop pre-eclampsia. Based on the allogeneic conditions in oocyte donation pregnancies, we hypothesized that this situation may translate into alterations in concentration of stable readouts of constituents of the reactive species interactome (RSI) compared to normal pregnancies, especially serum free thiols, nitric oxide (NO) and hydrogen sulfide (H2S) related metabolites. Indeed, total free thiol levels and nitrite (NO2) concentrations were significantly lower whereas protein-bound NO and sulfate (SO42−) concentrations were significantly higher in both oocyte donation and naturally conceived pregnancies complicated by pre-eclampsia. The increased concentrations of nitrite observed in uncomplicated oocyte donation pregnancies suggest that endothelial NO production is compensatorily enhanced to lower vascular tone. More research is warranted on the role of the RSI and bioenergetic status in uncomplicated oocyte donation pregnancies and oocyte donation pregnancies complicated by pre-eclampsia. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Indicators of oxidative stress. (<b>A</b>) Total free thiol concentrations were lower in pregnancies complicated by pre-eclampsia compared to controls and uncomplicated oocyte donation pregnancies (all, <span class="html-italic">p</span> &lt; 0.001). (<b>B</b>) Total free thiol concentrations were positively correlated with gestational age in naturally conceived pregnancies complicated by pre-eclampsia (r = 0.453, * <span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Total 8-iso-prostaglandin F<sub>2a</sub> concentrations in serum did not differ between groups. NC, uncomplicated naturally conceived pregnancies; OD, uncomplicated oocyte donation pregnancies; NCPE, naturally conceived pregnancies complicated by pre-eclampsia; ODPE, oocyte donation pregnancies complicated by pre-eclampsia. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Readouts of the nitric oxide (NO) pathway. (<b>A</b>) The concentrations of protein bound NO (RxNO) were higher in pregnancies complicated by pre-eclampsia (** <span class="html-italic">p</span> &lt; 0.01). (<b>B</b>) Diastolic blood pressure was inversely associated with RxNO in uncomplicated naturally conceived pregnancies (r = −0.556, * <span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Nitrite (NO<sub>2</sub><sup>−</sup>) concentrations were lower in naturally conceived pregnancies complicated by pre-eclampsia and oocyte donation pregnancies complicated by pre-eclampsia compared to uncomplicated naturally conceived pregnancies and uncomplicated oocyte donation pregnancies, respectively (** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001). Nitrite levels were higher in uncomplicated oocyte donation pregnancies compared to uncomplicated naturally conceived pregnancies (*** <span class="html-italic">p</span> &lt; 0.001). (<b>D</b>) Patients in the uncomplicated oocyte donation group were divided into two groups based on the nitrite serum levels: One group contained serum nitrite values below the median of 0.168 uM, while the other group contained serum nitrite values above the median. Diastolic blood pressure was found to be higher in women with serum nitrite values below the median compared to women with a serum nitrite values above the median (* <span class="html-italic">p</span> &lt; 0.05). (<b>E</b>) Nitrate (NO<sub>3</sub><sup>−</sup>) concentrations did not differ between groups. NC, uncomplicated naturally conceived pregnancies; OD, uncomplicated oocyte donation pregnancies; NCPE, naturally conceived pregnancies complicated by pre-eclampsia; ODPE, oocyte donation pregnancies complicated by pre-eclampsia. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Readouts of the hydrogen sulfide (H<sub>2</sub>S) pathway. (<b>A</b>) Sulfate (SO<sub>4</sub><sup>2−</sup>) values were higher in naturally conceived pregnancies complicated by pre-eclampsia compared to uncomplicated naturally conceived pregnancies and uncomplicated oocyte donation pregnancies (*** <span class="html-italic">p</span> &lt; 0.001). (<b>B</b>) Sulfate (SO<sub>4</sub><sup>−</sup>) values correlated positively with gestational age in naturally conceived pregnancies complicated by pre-eclampsia (r = 0.416, * <span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Thiosulfate (S<sub>2</sub>O<sub>3</sub><sup>2−</sup>) values did not differ between groups. NC, uncomplicated naturally conceived pregnancies; OD, uncomplicated oocyte donation pregnancies; NCPE, naturally conceived pregnancies complicated by pre-eclampsia; ODPE, oocyte donation pregnancies complicated by pre-eclampsia. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
18 pages, 5418 KiB  
Article
Early Pregnancy Human Decidua is Enriched with Activated, Fully Differentiated and Pro-Inflammatory Gamma/Delta T Cells with Diverse TCR Repertoires
by Antonia Terzieva, Violeta Dimitrova, Lyubomir Djerov, Petya Dimitrova, Silvina Zapryanova, Iana Hristova, Ivaylo Vangelov and Tanya Dimova
Int. J. Mol. Sci. 2019, 20(3), 687; https://doi.org/10.3390/ijms20030687 - 5 Feb 2019
Cited by 33 | Viewed by 6723
Abstract
Pregnancy is a state where high and stage-dependent plasticity of the maternal immune system is necessary in order to equilibrate between immunosuppression of harmful responses towards the fetus and ability to fight infections. TCR γδ cells have been implicated in the responses in [...] Read more.
Pregnancy is a state where high and stage-dependent plasticity of the maternal immune system is necessary in order to equilibrate between immunosuppression of harmful responses towards the fetus and ability to fight infections. TCR γδ cells have been implicated in the responses in infectious diseases, in the regulation of immune responses, and in tissue homeostasis and repair. The variety of functions makes γδ T cells a particularly interesting population during pregnancy. In this study, we investigated the proportion, phenotype and TCR γ and δ repertoires of γδ T cells at the maternal–fetal interface and in the blood of pregnant women using FACS, immunohistochemistry and spectratyping. We found an enrichment of activated and terminally differentiated pro-inflammatory γδ T-cell effectors with specific location in the human decidua during early pregnancy, while no significant changes in their counterparts in the blood of pregnant women were observed. Our spectratyping data revealed polyclonal CDR3 repertoires of the δ1, δ2 and δ3 chains and γ2, γ3, γ4 and γ5 chains and oligoclonal and highly restricted CDR3γ9 repertoire of γδ T cells in the decidua and blood of pregnant women. Early pregnancy induces recruitment of differentiated pro-inflammatory γδ T-cell effectors with diverse TCR repertoires at the maternal–fetal interface. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><span class="html-italic">In situ</span> visualization of γδ T cells (arrows) at the maternal-fetal interface during early pregnancy. (A) Periglandular clusters of γδ T cells; (B) γδ T cells scattered as single cells in decidual stroma; (C) intraepithelial γδ T cells in decidual glands; (D) staining for γδ T cells in human tonsils (positive control), and an inset is shown as a negative control. G: decidual gland.</p>
Full article ">Figure 2
<p>Ex vivo numbers of total γδ T cells and γδ T-cell subsets during pregnancy measured by FACS. (<b>a</b>) An increased γδ T-cell number in the decidua compared to that in the blood (early pregnancy, paired samples); (<b>b</b>) higher number of γδ T cells in early than in term deciduae and comparable γδ T-cell numbers in the peripheral blood of pregnant (PR) and non-pregnant (NP) women (<b>c</b>); (<b>d</b>) higher amount of Vδ1 cells in decidual tissues compared to that in the blood of PR women (paired samples) and predominance of this subset in the decidua at term; (<b>e</b>) conversely, the pathogen-reactive Vδ2 subset dominated the blood of NP women and decreased in the blood of PR women, at MFI Vδ2 cells were in a lower amount being less than 10% of γδ T cells; (<b>f</b>) representative FACS plots showing the number of γδ T cells derived from early and term deciduae and peripheral blood of PR and NP women. The number on the top right corner of each plot denotes the percentage of γδ T cells among CD3+ T cells. Data in the graphs are presented as mean ± s.e., obtained from Mann–Whitney and Wilcoxon matched pairs tests; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Phenotype of γδ T cells during human pregnancy: (<b>a</b>) prevalence of activated (HLA-DR+) γδ T cells at the MFI but not in the blood of both pregnant (PR) and non-pregnant (NP) women; (<b>b</b>) more than half of γδ T cells expressed NKG2D regardless of location and γδ+NKG2D+ cells was lowered in amount in the decidua at term; (<b>c</b>) expansion of pro-inflammatory γδ T-cell effectors (CCR5+) in early pregnancy decidua where naïve/memory CCR7+γδ+ cells were also in a higher amount compared to the peripheral blood counterparts; (<b>d</b>) representative histograms of pro-inflammatory chemokine receptor CCR5 expression on γδ T cells from early and term deciduae and peripheral blood of PR and NP women. The gate is put on CD3+γδ+ cells, FMO: fluorescence minus one control. Graphs show the mean ± s.e., <span class="html-italic">n</span> = 5–8, paired and unpaired <span class="html-italic">t</span>-tests, GraphPad Prism v.4. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Early pregnancy decidua is dominated by fully differentiated γδ T cells: (<b>a</b>) during early gestation, a predominance of fully differentiated effector (CD27−CD28−) γδ T cells in the decidua was found whereas the naïve/memory (CD27+CD28+) γδ T cells were in a negligible amount, in the blood of pregnant women half of γδ T cells were fully differentiated effectors and the remaining γδ T cells were with the naïve/memory and transitional phenotype; (<b>b</b>) the early human decidua was dominated by fully differentiated γδ T-cell effectors and no naïve/memory γδ T cells were detected, whereas the term decidua was populated mainly with γδ T cells with the naïve and transitional phenotype; (<b>c</b>) the differentiation status of peripheral blood γδ T cells during early pregnancy was similar between non-pregnant (NP) and pregnant (PR) women; (<b>d</b>) representative FACS plots show CD27 and CD28 expression for CD3+γδ+-gated cells from early and term deciduae, and from peripheral blood of PR and NP women. Graphs show the mean ± s.e., <span class="html-italic">n</span> = 5–6, paired and unpaired <span class="html-italic">t</span>-tests, GraphPad Prism v.4. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>δ1, δ2, δ3 repertoires (<b>a</b>–<b>c</b>) and γ2, γ3, γ4, γ5 and γ9 repertoires (<b>d</b>–<b>h</b>) in paired blood and decidua (B-D early pregnancy) and in term decidua (PL) samples. Note predominantly polyclonal repertoires of all δ and all γ chains, whereas CDR3 of γ9 chain was highly restricted with enriched CDR3 regions with one and the same length (<b>h</b>, red arrows). Each box represents the spectratyping data of one woman of the indicated CDR3 chain with gestational age in weeks.</p>
Full article ">
17 pages, 2391 KiB  
Article
Placental CX3CL1 is Deregulated by Angiotensin II and Contributes to a Pro-Inflammatory Trophoblast-Monocyte Interaction
by Olivia Nonn, Jacqueline Güttler, Désirée Forstner, Sabine Maninger, Julianna Zadora, András Balogh, Alina Frolova, Andreas Glasner, Florian Herse and Martin Gauster
Int. J. Mol. Sci. 2019, 20(3), 641; https://doi.org/10.3390/ijms20030641 - 2 Feb 2019
Cited by 22 | Viewed by 4551
Abstract
CX3CL1, which is a chemokine involved in many aspects of human pregnancy, is a membrane-bound chemokine shed into circulation as a soluble isoform. Placental CX3CL1 is induced by inflammatory cytokines and is upregulated in severe early-onset preeclampsia. In this study, the hypothesis was [...] Read more.
CX3CL1, which is a chemokine involved in many aspects of human pregnancy, is a membrane-bound chemokine shed into circulation as a soluble isoform. Placental CX3CL1 is induced by inflammatory cytokines and is upregulated in severe early-onset preeclampsia. In this study, the hypothesis was addressed whether angiotensin II can deregulate placental CX3CL1 expression, and whether CX3CL1 can promote a pro-inflammatory status of monocytes. qPCR analysis of human placenta samples (n = 45) showed stable expression of CX3CL1 and the angiotensin II receptor AGTR1 throughout the first trimester, but did not show a correlation between both or any influence of maternal age, BMI, and gestational age. Angiotensin II incubation of placental explants transiently deregulated CX3CL1 expression, while the angiotensin II receptor antagonist candesartan reversed this effect. Overexpression of recombinant human CX3CL1 in SGHPL-4 trophoblasts increased adhesion of THP-1 monocytes and significantly increased IL8, CCL19, and CCL13 in co-cultures with human primary monocytes. Incubation of primary monocytes with CX3CL1 and subsequent global transcriptome analysis of CD16+ subsets revealed 81 upregulated genes, including clusterin, lipocalin-2, and the leptin receptor. Aldosterone synthase, osteopontin, and cortisone reductase were some of the 66 downregulated genes present. These data suggest that maternal angiotensin II levels influence placental CX3CL1 expression, which, in turn, can affect monocyte to trophoblast adhesion. Release of placental CX3CL1 could promote the pro-inflammatory status of the CD16+ subset of maternal monocytes. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>CX3CL1 and AGTR1 mRNA expression in human first trimester placenta. Placental tissue samples (<span class="html-italic">n</span> = 45) from healthy, lean (BMI &lt; 25), non-smoking women with gestational ages ranging from 5 weeks to 10 weeks were analyzed for CX3CL1 (<b>A</b>) and AGTR1 (<b>B</b>) mRNA expression.</p>
Full article ">Figure 2
<p>AngII mediates a transient deregulation of placental CX3CL1. Placental explants were cultured with or without AngII (0.1 µM) in the presence or the absence of the AT1R antagonist Candesartan (0.1 µM) for 3 h (<b>A</b>), 6 h (<b>B</b>), and 24 h (<b>C</b>), respectively. Data are presented as median ± IQR (whiskers are min. to max., in <b>A</b> <span class="html-italic">n</span> = 4, in <b>B</b> <span class="html-italic">n</span> = 7, in <b>C</b> <span class="html-italic">n</span> = 7, * <span class="html-italic">p</span> ≤ 0.05) from different placental tissues.</p>
Full article ">Figure 3
<p>CX3CL1 overexpression in trophoblast cell line SGHPL-4 mediates increased monocyte adherence. SGHPL-4 control cells (<b>A</b>) and CX3CL1 stably overexpressing cells (SGHPL-4-CX3CL1, <b>B</b>) were stained for CX3CL1 by immunocytochemistry. Western blot (<b>C</b>) confirmed CX3CL1-overexpression in SGHPL-4-CX3CL1 cells. ELISA showed abundant release of soluble CX3CL1 in supernatants of SGHPL-4-CX3CL1 after 48 h of culturing (<b>D</b>). Metalloprotease inhibitor Batimastat, at concentrations of 5 µM and 10 µM, decreased the release of soluble CX3CL1, while cell associated CX3CL1 accumulated, when compared to the control after 48 h (<b>E</b>). Adhesion assays were performed with SGHPL-4 control cells (<b>F</b>) and SGHPL-4-CX3CL1 cells (<b>G</b>), which were co-cultured with fluorescence CellTracker Green pre-labeled THP-1 monocytes for 90 min. Monocyte adhesion was assessed by acquisition of trophoblast monolayer areas and bound THP-1 cells in phase contrast and green fluorescence channel, respectively. Pixel areas of bound monocytes were related to pixel areas of trophoblast monolayers and data for SGHPL-4-control was set as one. Data are from three independent experiments, using different cell passages and are presented as mean ± SEM *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 4
<p>Transcriptome analyses of primary human CD16<sup>+</sup> monocytes. Primary human monocytes were incubated with (stimulus, brown) or without (none, green) recombinant human CX3CL1. Thereafter, CD16<sup>+</sup> subtypes were subjected to microarray analysis. The heat map of deregulated genes (<b>A</b>), enrichment heatmap for biological processes from Mammalian Phenotype Ontology and Jensen DISEASES Ontology with color representing fold change of a gene (<b>B</b>), and protein-protein interaction networks according to STRING database with an edge thickness representing the confidence level of the interaction with upregeulated genes in red and downregulated genes in blue (<b>C</b>).</p>
Full article ">
12 pages, 1310 KiB  
Article
Increased HLA-G Expression in Term Placenta of Women with a History of Recurrent Miscarriage Despite Their Genetic Predisposition to Decreased HLA-G Levels
by Moniek H. C. Craenmehr, Iris Nederlof, Milo Cao, Jos J. M. Drabbels, Marijke J. Spruyt-Gerritse, Jacqueline D. H. Anholts, Hanneke M. Kapsenberg, Janine A. Stegehuis, Carin van der Keur, Esther Fasse, Geert W. Haasnoot, Marie-Louise P. van der Hoorn, Frans H. J. Claas, Sebastiaan Heidt and Michael Eikmans
Int. J. Mol. Sci. 2019, 20(3), 625; https://doi.org/10.3390/ijms20030625 - 1 Feb 2019
Cited by 26 | Viewed by 4651
Abstract
Human leukocyte antigen (HLA)-G is an immune modulating molecule that is present on fetal extravillous trophoblasts at the fetal-maternal interface. Single nucleotide polymorphisms (SNPs) in the 3 prime untranslated region (3′UTR) of the HLA-G gene can affect the level of HLA-G expression, which [...] Read more.
Human leukocyte antigen (HLA)-G is an immune modulating molecule that is present on fetal extravillous trophoblasts at the fetal-maternal interface. Single nucleotide polymorphisms (SNPs) in the 3 prime untranslated region (3′UTR) of the HLA-G gene can affect the level of HLA-G expression, which may be altered in women with recurrent miscarriages (RM). This case-control study included 23 women with a medical history of three or more consecutive miscarriages who delivered a child after uncomplicated pregnancy, and 46 controls with uncomplicated pregnancy. Genomic DNA was isolated to sequence the 3′UTR of HLA-G. Tissue from term placentas was processed to quantify the HLA-G protein and mRNA levels. The women with a history of RM had a lower frequency of the HLA-G 3′UTR 14-bp del/del genotype as compared to controls (Odds ratio (OR) 0.28; p = 0.039), which has previously been related to higher soluble HLA-G levels. Yet, HLA-G protein (OR 6.67; p = 0.006) and mRNA (OR 6.33; p = 0.010) expression was increased in term placentas of women with a history of RM as compared to controls. In conclusion, during a successful pregnancy, HLA-G expression is elevated in term placentas from women with a history of RM as compared to controls, despite a genetic predisposition that is associated with decreased HLA-G levels. These findings suggest that HLA-G upregulation could be a compensatory mechanism in the occurrence of RM to achieve an ongoing pregnancy. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Expression of trophoblast cell marker and HLA-G in term placenta. Representative examples of staining for (<b>A</b>) trophoblasts with cytokeratin marker anti-cytokeratin antibody (CAM5.2) and (<b>B</b>) all HLA-G isoforms with marker MEM-G2. Decidual parts of the placenta were annotated to specify the area for analysis.</p>
Full article ">Figure 2
<p>(<b>A</b>) Percentage positivity for trophoblast staining. No difference was observed in trophoblast staining between women with a history of RM and controls. (<b>B</b>) Percentage positivity for HLA-G staining. A higher HLA-G protein expression was observed in the decidual part of the placenta of women with a history of RM compared to controls. (<b>C</b>) HLA-G mRNA expression was measured in the placentas of women with a history of RM and controls. HLA-G mRNA expression was increased in term placenta of women with a history of RM as compared to controls.</p>
Full article ">Figure 3
<p>Amount of HLA-G staining in first trimester miscarriage and elective abortion material. HLA-G positive parts in the extravillous trophoblast (EVT) regions of the placental slides were scored to be (1) minimally, (2) moderately, or (3) intensely stained.</p>
Full article ">
11 pages, 7311 KiB  
Article
Involvement of the PD-1/PD-L1 Co-Inhibitory Pathway in the Pathogenesis of the Inflammatory Stage of Early-Onset Preeclampsia
by Matyas Meggyes, Eva Miko, Adrienn Lajko, Beata Csiszar, Barbara Sandor, Peter Matrai, Peter Tamas and Laszlo Szereday
Int. J. Mol. Sci. 2019, 20(3), 583; https://doi.org/10.3390/ijms20030583 - 29 Jan 2019
Cited by 19 | Viewed by 3306
Abstract
The programmed cell death protein 1 (PD-1) receptor has been reported to downregulate T cell activation effectively via binding to its ligands PD-L1 or PD-L2 in a negative co-stimulatory manner. Little is known about the involvement of PD-1 mediated immunoregulation in pregnancy and [...] Read more.
The programmed cell death protein 1 (PD-1) receptor has been reported to downregulate T cell activation effectively via binding to its ligands PD-L1 or PD-L2 in a negative co-stimulatory manner. Little is known about the involvement of PD-1 mediated immunoregulation in pregnancy and in pregnancy-related disorders. In this work, we investigated the possible role of the PD-1 co-stimulatory pathway in the pathogenesis of the clinical phase of early-onset preeclampsia characterized by a systemic maternal inflammatory response. We performed a cross-sectional study for comparative analysis of phenotypic and functional characteristics of peripheral blood mononuclear cells in women with early-onset preeclampsia and third-trimester healthy pregnant controls. According to our findings, enhanced expression of either PD-1 or its ligand PD-L1, or both, on the cell surface of effector cells (T cells, natural killer (NK) cells, natural killer T (NKT)-like cells) and Tregs could be observed, but PD-1 expression did not correlate with effector cells exhaustion. These results suggest the failure of the axis to downregulate Th1 responses, contributing thereby to the exaggerated immunoactivation observed in early-onset preeclampsia. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Programmed death-1 (PD-1) expression by peripheral lymphocyte populations in 3rd-trimester healthy pregnant women and in women with early-onset preeclampsia. The expression of PD-1 by CD8+ T cells (<b>A</b>), CD4+ T cells (<b>B</b>), Treg cells (<b>C</b>) and NKT-like cells (<b>D</b>) in peripheral blood of women with early-onset preeclampsia and in healthy pregnant women. The solid bars represent medians of 17 and 17 determinations, respectively, the boxes indicate the interquartile ranges, and the lines show the most extreme observations. Differences were considered statistically significant for <span class="html-italic">p</span>-values ≤ 0.05.</p>
Full article ">Figure 2
<p>PD-L1 expression by peripheral lymphocyte populations in 3rd-trimester healthy pregnant women and in women with early-onset preeclampsia. The expression of PD-L1 CD8+ T cells (<b>A</b>), CD4+ T cells (<b>B</b>), NK cells (<b>C</b>), and NKT-like cells (<b>D</b>) in peripheral blood of women with early-onset preeclampsia and in healthy pregnant women. The solid bars represent medians of 17 and 17 determinations, respectively, the boxes indicate the interquartile ranges, and the lines show the most extreme observations. Differences were considered statistically significant for <span class="html-italic">p</span>-values ≤ 0.05.</p>
Full article ">Figure 3
<p>Cytotoxicity of peripheral CD8+ T cells and NKT-like cells in 3rd-trimester healthy pregnant women and in women with early-onset preeclampsia. The expression of CD107a by PD-1 positive and negative CD8+ T cells (<b>A</b> and <b>B</b>) and NKT-like cells (<b>C</b> and <b>D</b>) in peripheral blood of women with early-onset preeclampsia and in healthy pregnant women. The solid bars represent medians of 17 and 17 determinations, respectively, the boxes indicate the interquartile ranges, and the lines show the most extreme observations. Differences were considered statistically significant for <span class="html-italic">p</span>-values ≤ 0.05.</p>
Full article ">Figure 4
<p>Percentage and cytotoxicity of PD-1/NKG2D double positive peripheral CD8+ T (<b>C</b>) and NKT-like cells (<b>D</b>) in 3rd-trimester healthy pregnant women and in women with early-onset preeclampsia. The percentage of PD-1/NKG2D double positive, CD8+ T cells (<b>A</b>), and NKT-like cells (<b>B</b>) in peripheral blood of women with early-onset preeclampsia and in healthy pregnant women. The solid bars represent medians of 17 and 17 determinations, respectively, the boxes indicate the interquartile ranges, and the lines show the most extreme observations. Differences were considered statistically significant for <span class="html-italic">p</span>-values ≤ 0.05.</p>
Full article ">Figure 5
<p>Circulating serum-soluble programmed death-ligand 1 (sPD-L1) levels in 3rd-trimester healthy pregnant women and in women with early-onset preeclampsia. Serum sPD-L1 levels in 16 healthy pregnant women and in 16 women with early-onset preeclampsia. Differences were considered statistically significant for <span class="html-italic">p</span>-values ≤ 0.05.</p>
Full article ">Figure 6
<p>Gating strategy to detect peripheral immune cell populations. Gating technique used to detect immune cell populations in the periphery.</p>
Full article ">
18 pages, 2254 KiB  
Article
Decidual Interleukin-22-Producing CD4+ T Cells (Th17/Th0/IL-22+ and Th17/Th2/IL-22+, Th2/IL-22+, Th0/IL-22+), Which Also Produce IL-4, Are Involved in the Success of Pregnancy
by Federica Logiodice, Letizia Lombardelli, Ornela Kullolli, Herman Haller, Enrico Maggi, Daniel Rukavina and Marie-Pierre Piccinni
Int. J. Mol. Sci. 2019, 20(2), 428; https://doi.org/10.3390/ijms20020428 - 19 Jan 2019
Cited by 39 | Viewed by 5932
Abstract
Trophoblast expressing paternal HLA-C resembles a semiallograft, and could be rejected by maternal T cells. IL-22 seems to be involved in allograft rejection and thus could be responsible for miscarriages. We examined the role of decidual IL-22-producing CD4+ T on human pregnancy. In [...] Read more.
Trophoblast expressing paternal HLA-C resembles a semiallograft, and could be rejected by maternal T cells. IL-22 seems to be involved in allograft rejection and thus could be responsible for miscarriages. We examined the role of decidual IL-22-producing CD4+ T on human pregnancy. In those experiencing successful pregnancy and those experiencing unexplained recurrent abortion (URA), the levels of IL-22 produced by decidual CD4+ T cells are higher than those of peripheral blood T cells. We found a correlation of IL-22 and IL-4 produced by decidual CD4+ T cells in those experiencing successful pregnancy, not in those experiencing URA. The correlation of IL-22 and IL-4 was also found in the serum of successful pregnancy. A prevalence of CD4+ T cells producing IL-22 and IL-4 (Th17/Th2/IL-22+, Th17/Th0/IL-22+, Th17/Th2/IL-22+, and Th0/IL-22+ cells) was observed in decidua of those experiencing successful pregnancy, whereas Th17/Th1/IL-22+ cells, which do not produce IL-4, are prevalent in those experiencing URA. Th17/Th2/IL-22+ and Th17/Th0/IL-22+ cells are exclusively present at the embryo implantation site where IL-4, GATA-3, IL-17A, ROR-C, IL-22, and AHR mRNA are expressed. T-bet and IFN-γ mRNA are found away from the implantation site. There is no pathogenic role of IL-22 when IL-4 is also produced by decidual CD4+ cells. Th17/Th2/IL-22+ and Th17/Th0/IL-22+ cells seem to be crucial for embryo implantation. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Cytokine production by CD4+ T cell clones derived from decidua of those experiencing successful pregnancy and URA and mRNA expression of cytokines and transcription factors in decidual biopsies of successful pregnancy. CD4+ T cell clones were generated from decidual biopsies, and peripheral blood was obtained from those experiencing successful pregnancy and those experiencing unexplained recurrent abortion (URA) (Experiment 1 in <a href="#sec4dot3-ijms-20-00428" class="html-sec">Section 4.3</a>). IL-4, IL-13, IL-5, IL-17A, IL-17F, IL-22, and IFN-γ were measured in the supernatant of the CD4+ T cell clones by a multiplex bead-based assay. The statistical analysis was performed with the Wilcoxon test. The determination of mRNA level for IL-4, GATA-3, IL-17A, ROR-C, IL-22, AHR, T-bet, and IFN-γ in three biopsies of decidua from three pregnant women (with successful pregnancy) was performed by Quantigene 2.0.</p>
Full article ">Figure 2
<p>In those experiencing successful pregnancy, IL-22 is positively correlated with IL-4, whereas, in those experiencing URA, IL-22 produced by CD4+ T cell clones derived from the decidua is positively correlated with IL-17A and IL-17F. The levels of IL-22 and the levels of IL-4, IL-13, IL-5, IL-17A, IL-17F, and IFN-γ measured in the supernatants of the CD4+ T cell clones derived from deciduae of those experiencing unexplained recurrent abortion (URA) and those experiencing successful pregnancy (Experiment 1 in <a href="#sec4dot3-ijms-20-00428" class="html-sec">Section 4.3</a>) have been correlated. NS means that p is not significant.</p>
Full article ">Figure 3
<p>In the serum of successful pregnancy, IL-22 is positively correlated with IL-4. The levels of IL-22, IL-4, IFN-γ, IL-5, IL-13, and IL-17A were measured with a multiplex bead-based assay in the serum of 18 women with successful pregnancy and 18 unexplained recurrent abortion (URA) patients, and these levels were correlated.</p>
Full article ">Figure 4
<p>Prevalence of CD4+ T helper cells producing IL-22 (Th/IL-22+) in the decidua of those experiencing successful pregnancy. The levels of IL-22 and the levels of IL-4, IL-13, IL-5, IL-17A, IL-17F, and IFN-γ have been measured in the supernatants of the CD4+ T cell clones derived from deciduae of those experiencing unexplained recurrent abortion (URA) and those experiencing successful pregnancy (Experiment 1 in <a href="#sec4dot3-ijms-20-00428" class="html-sec">Section 4.3</a>). The percentage of CD4+ T cells producing IL-22 (Th/IL-22+) in the decidua of those experiencing successful pregnancy and the percentage of T helper cells producing IL-22 in the URA deciduae have been calculated. The statistical analysis was performed with chi-square test.</p>
Full article ">Figure 5
<p>T helper subpopulations present in the decidua and in the peripheral blood of those experiencing unexplained recurrent abortion and those experiencing successful pregnancy. To investigate the CD4+ cell subsets that produce IL-22, the percentages of Th1-, Th2-, Th0-, Th17-, Th17/Th1-, Th17/Th2-, and Th17/Th0 cells, which do not produce IL-22 and which also produce IL-22 (Th1/IL-22+, Th2/IL-22+, Th0/IL-22+, Th17/IL-22+, Th17/Th1/IL-22+, Th17/Th2/IL-22+, and Th17/Th0/IL-22+) were analyzed. Cytokines were measured in the supernatants of the CD4+ T cell clones derived from the decidua and peripheral blood of those experiencing successful pregnancy and those experiencing unexplained recurrent abortion (URA) (Experiment 1 in <a href="#sec4dot3-ijms-20-00428" class="html-sec">Section 4.3</a>) by a multiplex bead-based assay. The statistical analysis was performed with the chi-square test.</p>
Full article ">Figure 6
<p>Th17/Th2/IL-22+ and Th17/Th0/IL-22+ CD4+ T cells are exclusively present at the implantation site of ectopic pregnancy. The percentages of Th0/IL-22+, Th2/IL-22+, Th17/Th0/IL-22+, Th17/Th2/IL-22+, and Th17/Th1/IL-22+ cells among the CD4+ T cell clones respectively derived from the implantation site of the embryo and distant from the implantation site in the same Fallopian tube of four women suffering from ectopic pregnancy were evaluated (Experiment 2 according to <a href="#sec4dot3-ijms-20-00428" class="html-sec">Section 4.3</a>). The statistical analysis was performed with the chi-square test. The determination of mRNA level for IL-4, GATA-3, IL-17A, ROR-C, IL-22, AHR, T-bet, and IFN -γ in Fallopian tube tissue taken at the embryo implantation site and tissue sampled distant from the implantation site of an additional woman suffering from ectopic pregnancy was performed by Quantigene 2.0.</p>
Full article ">Figure 7
<p>The determination of mRNA level for IL-4, GATA-3, IL-17A, ROR-C, IL-22, AHR, T-bet, and IFN -γ in Fallopian tube tissue taken at the embryo implantation site and distant from the implantation site of a woman suffering from ectopic pregnancy and in decidual biopsies from three women with successful pregnancy was performed by Quantigene 2.0.</p>
Full article ">Figure 8
<p>Possible positive roles of Th22/Th2/IL-22+ and Th17/Th0/IL-22+ cells at fetal maternal interface.</p>
Full article ">
11 pages, 2243 KiB  
Article
A Proteome Approach Reveals Differences between Fertile Women and Patients with Repeated Implantation Failure on Endometrial Level—Does hCG Render the Endometrium of RIF Patients?
by Alexandra P. Bielfeld, Sarah Jean Pour, Gereon Poschmann, Kai Stühler, Jan-Steffen Krüssel and Dunja M. Baston-Büst
Int. J. Mol. Sci. 2019, 20(2), 425; https://doi.org/10.3390/ijms20020425 - 19 Jan 2019
Cited by 31 | Viewed by 4974
Abstract
Background: The molecular signature of endometrial receptivity still remains barely understood, especially when focused on the possible benefit of therapeutical interventions and implantation-related pathologies. Therefore, the protein composition of tissue and isolated primary cells (endometrial stromal cells, ESCs) from endometrial scratchings of ART [...] Read more.
Background: The molecular signature of endometrial receptivity still remains barely understood, especially when focused on the possible benefit of therapeutical interventions and implantation-related pathologies. Therefore, the protein composition of tissue and isolated primary cells (endometrial stromal cells, ESCs) from endometrial scratchings of ART (Assisted Reproductive Techniques) patients with repeated implantation failure (RIF) was compared to volunteers with proven fertility during the time of embryo implantation (LH + 7). Furthermore, an analysis of the endometrial tissue of fertile women infused with human chorionic gonadotropin (hCG) was conducted. Methods: Endometrial samples (n = 6 RIF, n = 10 fertile controls) were split into 3 pieces: 1/3 each was frozen in liquid nitrogen, 1/3 fixed in PFA and 1/3 cultured. Protein lysates prepared from fresh frozen tissue were processed for mass spectrometric analysis. Results: Three proteins (EPPK1, BCLAF1 and PTMA) showed a statistically altered abundance in the endometrial tissue of RIF patients. Furthermore, pathways like metabolism, immune system, ferroptosis and the endoplasmic reticulum were altered in RIF patients. Remarkably, endometrial tissues of RIF patients showed a significantly higher (p-value = 9 × 10−8) protein intensity correlation (Pearson’s correlation coefficient = 0.95) compared to fertile women (Pearson’s correlation coefficient = 0.88). The in vivo infusion of hCG stimulated proteins of endocytosis, HIF1 signalling and chemokine production. Notably, patients suffering from RIF had a clinical pregnancy rate of 19% after the intrauterine infusion of hCG before embryo transfer (ET) compared to their failed previous cycles. Conclusion: Our study showed for the first time that the endometrial proteome composition of RIF patients differs from fertile controls during the window of implantation. The intrauterine infusion of hCG prior to an embryo transfer might improve the chemokine triggered embryo-endometrial dialogue and intensify the angiogenesis and immune response. From a clinical point of view, the hCG infusion prior to an embryo transfer might increase the pregnancy rate of RIF patients. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Volcano plots of the proteomic analysis of repeated implantation failure (RIF) endometrial tissue (<span class="html-italic">n</span> = 6) vs. proven fertility (PF) (<span class="html-italic">n</span> = 10). All identified proteins are shown with their associated gene names. Proteins showing a significantly altered abundance are highlighted in red.</p>
Full article ">Figure 2
<p>Intensity correlation (log 2 intensities are shown) of protein of endometrial tissue of RIF patients (<span class="html-italic">n</span> = 6) and PF (<span class="html-italic">n</span> = 10). Comparison within the RIF group is marked by a blue triangle and within the PF group by an orange triangle. Circles represent the highest and lowest correlation coefficient of the respective group comparison. The red circles highlight the highest and lowest correlation coefficient between RIF and PF.</p>
Full article ">Figure 3
<p>Volcano plots of cultured RIF decidualized primary endometrial stromal cell (dpESC) incubated for 24 h with and without 100 IU human chorionic gonadotropin (hCG)/mL. All protein names are shown, but only NRAS (marked in red) revealed a significant lower abundance after incubation with hCG.</p>
Full article ">Figure 4
<p>Volcano plot analysis of RIF dpESCs incubated with 100 IU/mL hCG vs. the endometrial tissue of PF (<b>A</b>). All proteins displayed in red show a statistically significant altered abundance. (<b>A</b>) heatmap (low to high abundance is coded by yellow to red color) representing the unsupervised hierarchical cluster analysis is shown in (<b>B</b>).</p>
Full article ">Figure 5
<p>Results of the infusion with 500 IU hCG infusion for 24 h in PF.</p>
Full article ">
16 pages, 2880 KiB  
Article
Melatonin Improves Parthenogenetic Development of Vitrified–Warmed Mouse Oocytes Potentially by Promoting G1/S Cell Cycle Progression
by Bo Pan, Haoxuan Yang, Zhenzheng Wu, Izhar Hyder Qazi, Guoshi Liu, Hongbing Han, Qingyong Meng and Guangbin Zhou
Int. J. Mol. Sci. 2018, 19(12), 4029; https://doi.org/10.3390/ijms19124029 - 13 Dec 2018
Cited by 24 | Viewed by 4646
Abstract
This study aimed to investigate the effect of melatonin on the cell cycle of parthenogenetic embryos derived from vitrified mouse metaphase II (MII) oocytes. Fresh oocytes were randomly allocated into three groups: untreated (control), or vitrified by the open-pulled straw method without (Vitrification [...] Read more.
This study aimed to investigate the effect of melatonin on the cell cycle of parthenogenetic embryos derived from vitrified mouse metaphase II (MII) oocytes. Fresh oocytes were randomly allocated into three groups: untreated (control), or vitrified by the open-pulled straw method without (Vitrification group) or with melatonin (MT) supplementation (Vitrification + MT group). After warming, oocytes were parthenogenetically activated and cultured in vitro, then the percentage of embryos in the G1/S phase, the levels of reactive oxygen species (ROS) and glutathione (GSH), and the mRNA expression of cell cycle-related genes (P53, P21 and E2F1) in zygotes and their subsequent developmental potential in vitro were evaluated. The results showed that the vitrification/warming procedures significantly decreased the frequency of the S phase, markedly increased ROS and GSH levels and the expression of P53 and P21 genes, and decreased E2F1 expression in zygotes at the G1 stage and their subsequent development into 2-cell and blastocyst stage embryos. However, when 10−9 mol/L MT was administered for the whole duration of the experiment, the frequency of the S phase in zygotes was significantly increased, while the other indicators were also significantly improved and almost recovered to the normal levels shown in the control. Thus, MT might promote G1-to-S progression via regulation of ROS, GSH and cell cycle-related genes, potentially increasing the parthenogenetic development ability of vitrified–warmed mouse oocytes. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Typical phase showing nucleolus status of parthenogenetic zygotes. After parthenogenetic activation of mouse MII oocytes followed by in vitro culture for 3 to 4 h, the resulting zygotes ((<b>A</b>) 3 h; (<b>B</b>) 4 h) were observed under a stereomicroscope for determination of their nucleolus status. The embryos with two separate pronuclei and no apparent nucleoli inside (white arrows) remained at the G1 stage, while those with apparent nucleoli (red arrows) had proceeded through the G1 into the S phase. In <a href="#ijms-19-04029-f001" class="html-fig">Figure 1</a>A, the criterion of pronuclei of parthenogenetic zygote in G1 stage is shown in the white rectangle (zoomed-in frame of corresponding zygote with white asterisk). Similarly, in <a href="#ijms-19-04029-f001" class="html-fig">Figure 1</a>B, the criterion of pronuclei of parthenogenetic zygote in the S stage is shown in the red rectangle (zoomed-in frame of corresponding zygote with red asterisk). Original magnification 200×.</p>
Full article ">Figure 2
<p>Reactive oxygen species (ROS) levels in mouse MII oocytes and their parthenogenetic zygotes. The dynamic change of ROS levels in mouse oocytes and their parthenogenetic zygotes (<b>A</b>) ROS staining in oocytes (<b>B</b>,<b>C</b>) and their parthenogenetic zygotes (<b>D</b>) Fluorescence intensities were correlated with intracellular levels of ROS. Different superscripts (a and b) represent treatment differences within panels (<span class="html-italic">p</span> &lt; 0.05). After warming, mouse MII oocytes were in vitro cultured for 0 h (Oocyte-IVC 0 h) or 1 h (Oocyte-IVC 1 h) in M2 medium. The oocytes cultured for 1 h were selected for parthenogenetic activation (PA). During the entire experiment, all the media were supplemented with 10<sup>−9</sup> mol/L (Vitrification + MT group) or 0 mol/L melatonin (Vitrification group). Fresh oocytes without melatonin (MT) treatment were used as controls (Control group). After PA and in vitro culture for 3 h, the resulting zygotes (zygote-IVC 3 h) were used for ROS detection together with mouse oocytes before PA. The values (the relative ROS levels)) are shown as mean ± SEM. The experiment was replicated at least three times. Original magnification 200×.</p>
Full article ">Figure 3
<p>Glutathione (GSH) levels in mouse MII oocytes and their parthenogenetic zygotes. (<b>A</b>) The dynamic change of GSH levels in mouse oocytes and their parthenogenetic zygotes. (<b>B</b>) GSH staining of oocytes and their parthenogenetic zygotes in three groups. Fluorescence intensities were correlated with intracellular levels of GSH. The values (the relative GSH levels) are shown as mean ± SEM. The experiment was replicated at least three times. Different superscripts (a and b) represent treatment differences within panels (<span class="html-italic">p</span> &lt; 0.05). Original magnification 200×.</p>
Full article ">Figure 4
<p>Effect of melatonin on mRNA expression of cell cycle-related genes in parthenogenetic zygotes (G1 stage). (<b>A</b>–<b>C</b>) The relative mRNA expression of cell cycle-related genes (<span class="html-italic">P53</span>, <span class="html-italic">P21</span> and <span class="html-italic">E2F1</span>) in zygotes at G1 stage. The relative expression levels of mRNA were determined by the 2<sup>−△△<span class="html-italic">C</span>t</sup> method and normalized against that of the reference gene <span class="html-italic">GAPDH</span> (glyceraldehyde 3-phosphate dehydrogenase). All data are mean ± SEM from three replicates. Different superscripts (a, b and c) represent treatment differences within panels (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Flowchart of experimental design. Control group: untreated mouse MII oocytes; Vitrification group: oocytes were vitrified by the open-pulled straw method without melatonin (MT) addition; Vitrification + MT group: oocytes were treated with MT at a final concentration of 10<sup>−9</sup> mol/L in all the media used in the entire experiment; PA: parthenogenetic activation; IVC: in vitro culture; MT: melatonin.</p>
Full article ">
12 pages, 2771 KiB  
Article
Gosha-Jinki-Gan Recovers Spermatogenesis in Mice with Busulfan-Induced Aspermatogenesis
by Ning Qu, Miyuki Kuramasu, Yoshie Hirayanagi, Kenta Nagahori, Shogo Hayashi, Yuki Ogawa, Hayato Terayama, Kaori Suyama, Munekazu Naito, Kou Sakabe and Masahiro Itoh
Int. J. Mol. Sci. 2018, 19(9), 2606; https://doi.org/10.3390/ijms19092606 - 3 Sep 2018
Cited by 16 | Viewed by 4643
Abstract
Busulfan is an anti-cancer chemotherapeutic drug and is often used as conditioning regimens prior to bone marrow transplant for treatment of chronic myelogenous leukemia. Male infertility, including spermatogenesis disturbance, is known to be one of the side effects of anticancer drugs. While hormone [...] Read more.
Busulfan is an anti-cancer chemotherapeutic drug and is often used as conditioning regimens prior to bone marrow transplant for treatment of chronic myelogenous leukemia. Male infertility, including spermatogenesis disturbance, is known to be one of the side effects of anticancer drugs. While hormone preparations and vitamin preparations are used for spermatogenesis disturbance, their therapeutic effects are low. Some traditional herbal medicines have been administered to improve spermatogenesis. In the present study, we administered Gosha-jinki-gan (TJ107; Tsumura Co., Ltd., Tokyo, Japan) to mice suffering from severe aspermatogenesis after busulfan treatment to determine whether TJ107 can recover spermatogenesis. Male 4-week-old C57BL/6J mice were administered a single intraperitoneal injection of busulfan, and they were then fed a normal diet for 60 days and then a TJ107 diet or TJ107-free normal diet for another 60 days. After busulfan treatment, the weight of the testes and the epididymal sperm count progressively decreased in the normal diet group. On the other hand, in the TJ107 group, these variables dramatically recovered at 120 days. These results suggest that busulfan-induced aspermatogenesis is irreversible if appropriate treatment is not administered. Supplementation of TJ107 can completely recover the injured seminiferous epithelium via normalization of the macrophage migration and reduction of the expressions of Tool-like receptor (TLR) 2 and TLR4, suggesting that TJ107 has a therapeutic effect on busulfan-induced aspermatogenesis. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Three-dimensional high-performance liquid chromatography profile of Gosha-jinki-gan. Gal: galloyl.</p>
Full article ">Figure 2
<p>Histological examination of the testes (<span class="html-italic">n</span> = 10) at day 120 in each group. Intact seminiferous tubules showing all maturation stages of the germinal epithelium from spermatogonia to spermatozoa are seen in the control group (<b>a</b>) and the control+TJ107 group (<b>b</b>). Atrophic seminiferous tubules with azoospermia are seen in the BSF group (<b>c</b>). Normal-appearing seminiferous tubules are seen in the BSF+TJ107 group (<b>d</b>). Bar = 50 µm.</p>
Full article ">Figure 3
<p>Detection of proliferating cells in the testes using antibodies against Ki67 nuclear antigen at day 120 in the control (<b>a</b>), control+TJ107 (<b>b</b>), BSF (<b>c</b>), and BSF+TJ107 groups (<b>d</b>). The enlarged part (Bar: 40 μm) of the seminiferous tubules is bounded by solid frames and is shown in the top right corner in each testis section. Ten testes from each group were examined. Dark brown spots indicating Ki67-positive nuclei of proliferating spermatogonia are detected in almost all seminiferous tubules in the control (<b>a</b>), control+TJ107 (<b>b</b>), and BSF+TJ107 (<b>d</b>) groups at day 120. Most germ cells are destroyed and few seminiferous tubules with Ki67-positive cells are sporadically observed in the BSF group (<b>c</b>). Bar: 40 μm. (<b>e</b>) Expression of Ki67 in testicular tissues of mice from each group (<span class="html-italic">n</span> = 10). Asterisks indicate <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Histological detection of TUNEL staining in testicular sections of mice from the control (<b>a</b>), control+TJ107 (<b>b</b>), BSF (<b>c</b>), and BSF+TJ107 (<b>d</b>) groups at day 120. Ten testes from each group were examined. Dark brown spots indicating TUNEL-positive nuclei of apoptotic germ cells are seen. Some TUNEL-positive germ cells are detected in the seminiferous tubules in the control (<b>a</b>), control+TJ107 (<b>b</b>), and BSF+TJ107 (<b>d</b>) groups at day 120. Most germ cells are destroyed at day 120 in the BSF group; thus, TUNEL-positive cells (arrowhead) are hardly detected (<b>c</b>). Bar: 40 μm. (<b>e</b>) Expressions of Fas, FasL, Caspase3, Caspase8, and Caspase9 in testicular tissues of mice from each group (<span class="html-italic">n</span> = 10). Asterisks indicate <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>TLR expression and macrophage migration at day 120 in each group. <b>A</b>: Expressions of TLR2 and TLR4 in testicular tissues of mice from each group (<span class="html-italic">n</span> = 10). Asterisks indicate <span class="html-italic">p</span> &lt; 0.05. <b>B</b>: F4/80 antigen-positive cells in testicular sections (<span class="html-italic">n</span> = 10) of mice from the control (<b>a</b>), control+TJ107 (<b>b</b>), BSF (<b>c</b>), and BSF+TJ107 (<b>d</b>) groups. Many macrophages are seen infiltrating the interstitium around the atrophic seminiferous tubules in the BSF group (<b>c</b>). Bar = 20 µm. (<b>e</b>) Expression of the macrophage marker gene in testicular tissues of mice from each group (<span class="html-italic">n</span> = 10). Asterisks indicate <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Time schedule of treatment applied to the mice in each group.</p>
Full article ">
13 pages, 2940 KiB  
Article
Aggregation of Human Trophoblast Cells into Three-Dimensional Culture System Enhances Anti-Inflammatory Characteristics through Cytoskeleton Regulation
by Kotomi Seno, Yasuhisa Munakata, Michiya Sano, Ryouka Kawahara-Miki, Hironori Takahashi, Akihide Ohkuchi, Hisataka Iwata, Takehito Kuwayama and Koumei Shirasuna
Int. J. Mol. Sci. 2018, 19(8), 2322; https://doi.org/10.3390/ijms19082322 - 8 Aug 2018
Cited by 7 | Viewed by 4284
Abstract
Background: Three-dimensional (3D) culture changes cell characteristics and function, suggesting that 3D culture provides a more physiologically relevant environment for cells compared with 2D culture. We investigated the differences in cell functions depending on the culture model in human trophoblast cells (Sw.71). Methods: [...] Read more.
Background: Three-dimensional (3D) culture changes cell characteristics and function, suggesting that 3D culture provides a more physiologically relevant environment for cells compared with 2D culture. We investigated the differences in cell functions depending on the culture model in human trophoblast cells (Sw.71). Methods: Sw.71 cells were incubated in 2D monolayers or simple 3D spheroids. After incubation, cells were corrected to assess RNA-seq transcriptome or protein expression, and culture medium were corrected to detect cytokines. To clarify the role of actin cytoskeleton, spheroid Sw.71 cells were treated mycalolide B (inhibitor of actin polymerization) in a 3D culture. Results: RNA-seq transcriptome analysis, results revealed that 3D-cultured cells had a different transcriptional profile compared with 2D-cultured cells, especially regarding inflammation-related molecules. Although interleukin-6 (IL-6) mRNA level was higher in 3D-culured cells, its secretion levels were higher in 2D-cultured cells. In addition, the levels of mRNA and protein expression of regnase-1, regulatory RNase of inflammatory cytokine, significantly increased in 3D culture, suggesting post-translational modification of IL-6 mRNA via regnase-1. Treatment with mycalolide B reduced cell-to-cell contact to build 3D formation and increased expression of actin cytoskeleton, resulting in increased IL-6 secretin. Conclusion: Cell dimensionality plays an essential role in governing the spatiotemporal cellular outcomes, including inflammatory cytokine production and its negative regulation associated with regnase-1. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Effects of 3D culture conditions on inflammatory cytokines in Sw.71 cells. Sw.71 trophoblast cells were incubated in 2D or 3D culture plates. (<b>A</b>) Observation of structural changes of Sw.71 cells maintained in 3D (×40 magnification). (<b>B</b>,<b>C</b>) After 24 h incubation, <span class="html-italic">IL-8</span> and <span class="html-italic">IL-6</span> mRNA levels were measured using RT-qPCR (<span class="html-italic">n</span> = 4). (<b>D</b>,<b>E</b>) After 24 h incubation, IL-8 and IL-6 concentrations in supernatants (/mL) were determined using ELISA (<span class="html-italic">n</span> = 4). (<b>F</b>,<b>G</b>) After 24 h incubation, IL-8 and IL-6 concentrations in supernatants (/protein mg) were calculated (<span class="html-italic">n</span> = 4). (<b>H</b>,<b>I</b>) After 24 h incubation, IL-8 and IL-6 concentrations in supernatants (/DNA concentration) were calculated (<span class="html-italic">n</span> = 4). (<b>J</b>) After 24 h incubation, cell number-dependent IL-6 concentrations in supernatants (/mL) were determined using ELISA (<span class="html-italic">n</span> = 4). (<b>K</b>) Culture time-dependent IL-6 concentrations in supernatants (/mL) were determined using ELISA (<span class="html-italic">n</span> = 4). Data are expressed as mean ± standard error of the mean (SEM). Significant differences were detected using a <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 (**).</p>
Full article ">Figure 2
<p>Effects of 3D culture conditions on NF-κB system in Sw.71 cells. Sw.71 trophoblast cells were incubated for 24 h in 2D or 3D culture plates. (<b>A</b>) NF-κB p65, phosphor IκB, total IκB, and GAPDH protein levels in the cell lysates were detected using Western blot. Representative data are shown. (<b>B</b>) <span class="html-italic">NF-κB p65</span> mRNA levels were measured using RT-qPCR (<span class="html-italic">n</span> = 4). (<b>C</b>) Active NF-κB p65 expression isolated from nuclei were determined using ELISA (<span class="html-italic">n</span> = 3). Data are expressed as mean ± SEM. Significant differences were detected using a <span class="html-italic">t</span>-test; <span class="html-italic">p</span> &lt; 0.05 (*).</p>
Full article ">Figure 3
<p>Effects of 3D culture conditions on RNase of inflammatory cytokine in Sw.71 cells. Sw.71 trophoblast cells were incubated for 24 h in 2D or 3D culture plates. (<b>A</b>–<b>E</b>) After 24 h incubation, <span class="html-italic">regnase-1</span>, <span class="html-italic">roquin-1</span>, <span class="html-italic">roquin-2</span>, <span class="html-italic">TTP</span>, and <span class="html-italic">AUR1</span> mRNA levels were measured using next-generation sequencing (<span class="html-italic">n</span> = 3). (<b>F</b>) <span class="html-italic">Regnase-1</span> mRNA levels were measured using RT-qPCR (<span class="html-italic">n</span> = 4). (<b>G</b>,<b>H</b>) Regnase-1 and GAPDH protein levels in the cell lysates were detected using Western blot. Representative data are shown. Data are expressed as mean ± SEM (<span class="html-italic">n</span> = 4 in each experiment). Significant differences were detected using a <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 (**).</p>
Full article ">Figure 4
<p>Effects of 3D culture conditions on the actin cytoskeleton in Sw.71 cells. Sw.71 trophoblast cells were incubated for 24 h under 2D or 3D culture plates. (<b>A</b>) After 24 h incubation, β-actin (ACTB) and GAPDH protein levels in the cell lysates were detected using Western blot. Representative data are shown. (<b>B</b>) <span class="html-italic">ACTB</span> mRNA levels were measured using RT-qPCR (<span class="html-italic">n</span> = 4). (<b>C</b>,<b>D</b>) Representative images of F-actin filaments staining after 2D, 3D culture (×400 magnification). Data are expressed as mean ± SEM (<span class="html-italic">n</span> = 4 in each experiment). Significant differences were detected using <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 (**)</p>
Full article ">Figure 5
<p>Effects of mycalolide B and poly I:C on Sw.71 cells. (<b>A</b>) Sw.71 trophoblast cells were incubated for 24 h in 3D culture plates with or without mycalolide B. Observation of structural change of Sw.71 cells under 3D culture plate (×40 magnification). (<b>B</b>) IL-6 concentrations in supernatants (/mL) treated with mycalolide B were determined using ELISA (<span class="html-italic">n</span> = 4). (<b>C</b>) ACTB and GAPDH protein levels in the cell lysates were detected using Western blot. Representative data are shown. (<b>D</b>,<b>E</b>) Representative images of stained F-actin filaments with or without mycalolide B treatment (x 200 magnification). (<b>F</b>,<b>G</b>) IL-6 and sFlt-1 concentrations in supernatants (/mL) from cells treated with poly I:C were determined using ELISA (<span class="html-italic">n</span> = 4). Data are expressed as mean ± SEM. Significant differences were detected using a <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 (**).</p>
Full article ">

Review

Jump to: Research

29 pages, 1005 KiB  
Review
Endometrial Immune Dysfunction in Recurrent Pregnancy Loss
by Carlo Ticconi, Adalgisa Pietropolli, Nicoletta Di Simone, Emilio Piccione and Asgerally Fazleabas
Int. J. Mol. Sci. 2019, 20(21), 5332; https://doi.org/10.3390/ijms20215332 - 26 Oct 2019
Cited by 136 | Viewed by 8718
Abstract
Recurrent pregnancy loss (RPL) represents an unresolved problem for contemporary gynecology and obstetrics. In fact, it is not only a relevant complication of pregnancy, but is also a significant reproductive disorder affecting around 5% of couples desiring a child. The current knowledge on [...] Read more.
Recurrent pregnancy loss (RPL) represents an unresolved problem for contemporary gynecology and obstetrics. In fact, it is not only a relevant complication of pregnancy, but is also a significant reproductive disorder affecting around 5% of couples desiring a child. The current knowledge on RPL is largely incomplete, since nearly 50% of RPL cases are still classified as unexplained. Emerging evidence indicates that the endometrium is a key tissue involved in the correct immunologic dialogue between the mother and the conceptus, which is a condition essential for the proper establishment and maintenance of a successful pregnancy. The immunologic events occurring at the maternal–fetal interface within the endometrium in early pregnancy are extremely complex and involve a large array of immune cells and molecules with immunoregulatory properties. A growing body of experimental studies suggests that endometrial immune dysregulation could be responsible for several, if not many, cases of RPL of unknown origin. The present article reviews the major immunologic pathways, cells, and molecular determinants involved in the endometrial dysfunction observed with specific application to RPL. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the changes occurring in the human endometrium and in local immune cell trafficking in the normal state and in recurrent pregnancy loss (RPL). (<b>a</b>) Endometrium in the secretory phase of the menstrual cycle in the absence of the embryo; (<b>b</b>) endometrium in the presence of a normally implanting embryo; (<b>c</b>) endometrial immune derangements in RPL. M: macrophages; uNK: uterine natural killer cell; iDC: immature uterine dendritic cell; mDC: mature uterine dendritic cell; N: neutrophil granulocyte; Treg: regulatory T cell; MC: mastocyte; Teff: effector T cells; ESC: endometrial stromal cell; BV: blood vessel.</p>
Full article ">
39 pages, 1503 KiB  
Review
Beneficial and Deleterious Effects of Female Sex Hormones, Oral Contraceptives, and Phytoestrogens by Immunomodulation on the Liver
by Luis E. Soria-Jasso, Raquel Cariño-Cortés, Víctor Manuel Muñoz-Pérez, Elizabeth Pérez-Hernández, Nury Pérez-Hernández and Eduardo Fernández-Martínez
Int. J. Mol. Sci. 2019, 20(19), 4694; https://doi.org/10.3390/ijms20194694 - 22 Sep 2019
Cited by 14 | Viewed by 5791
Abstract
The liver is considered the laboratory of the human body because of its many metabolic processes. It accomplishes diverse activities as a mixed gland and is in continuous cross-talk with the endocrine system. Not only do hormones from the gastrointestinal tract that participate [...] Read more.
The liver is considered the laboratory of the human body because of its many metabolic processes. It accomplishes diverse activities as a mixed gland and is in continuous cross-talk with the endocrine system. Not only do hormones from the gastrointestinal tract that participate in digestion regulate the liver functions, but the sex hormones also exert a strong influence on this sexually dimorphic organ, via their receptors expressed in liver, in both health and disease. Besides, the liver modifies the actions of sex hormones through their metabolism and transport proteins. Given the anatomical position and physiological importance of liver, this organ is evidenced as an immune vigilante that mediates the systemic immune response, and, in turn, the immune system regulates the hepatic functions. Such feedback is performed by cytokines. Pro-inflammatory and anti-inflammatory cytokines are strongly involved in hepatic homeostasis and in pathological states; indeed, female sex hormones, oral contraceptives, and phytoestrogens have immunomodulatory effects in the liver and the whole organism. To analyze the complex and interesting beneficial or deleterious effects of these drugs by their immunomodulatory actions in the liver can provide the basis for either their pharmacological use in therapeutic treatments or to avoid their intake in some diseases. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Scheme of the liver and hepatic sinusoid. Liver right lobe (RL), left lobe (LL), and the hepatic sinusoid that is composed by hepatocytes (H), sinusoidal endothelial cells (E), macrophage Kupffer cell (K), natural killer or pit cell (P), Ito or stellate cell (S), space of Disse (D), and lumen (L).</p>
Full article ">Figure 2
<p>Intracellular pathways of estrogens via estrogen receptors (ERs) in hepatocyte cells. (<b>a</b>) Genomic effects of estrogens via nuclear ERs. (<b>b</b>) Non-genomic effects of estrogens via membrane-associated ERs. ERE = Estrogen response element; TF = Transcription Factor.</p>
Full article ">Figure 3
<p>Chemical structure of representative estrogens, progestins, and phytoestrogens.</p>
Full article ">
31 pages, 2353 KiB  
Review
European Patent in Immunoncology: From Immunological Principles of Implantation to Cancer Treatment
by Franziska M. Würfel, Christoph Winterhalter, Peter Trenkwalder, Ralph M. Wirtz and Wolfgang Würfel
Int. J. Mol. Sci. 2019, 20(8), 1830; https://doi.org/10.3390/ijms20081830 - 12 Apr 2019
Cited by 11 | Viewed by 5919
Abstract
The granted European patent EP 2 561 890 describes a procedure for an immunological treatment of cancer. It is based on the principles of the HLA-supported communication of implantation and pregnancy. These principles ensure that the embryo is not rejected by the mother. [...] Read more.
The granted European patent EP 2 561 890 describes a procedure for an immunological treatment of cancer. It is based on the principles of the HLA-supported communication of implantation and pregnancy. These principles ensure that the embryo is not rejected by the mother. In pregnancy, the placenta, more specifically the trophoblast, creates an “interface” between the embryo/fetus and the maternal immune system. Trophoblasts do not express the “original” HLA identification of the embryo/fetus (HLA-A to -DQ), but instead show the non-classical HLA groups E, F, and G. During interaction with specific receptors of NK cells (e.g., killer-immunoglobulin-like receptors (KIR)) and lymphocytes (lymphocyte-immunoglobulin-like receptors (LIL-R)), the non-classical HLA groups inhibit these immunocompetent cells outside pregnancy. However, tumors are known to be able to express these non-classical HLA groups and thus make use of an immuno-communication as in pregnancies. If this occurs, the prognosis usually worsens. This patent describes, in a first step, the profiling of the non-classical HLA groups in primary tumor tissue as well as metastases and recurrent tumors. The second step comprises tailored antibody therapies, which is the subject of this patent. In this review, we analyze the underlying mechanisms and describe the currently known differences between HLA-supported communication of implantation and that of tumors. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Receptor interaction of HLA-E with NKG2A, -B/CD94 and NKG2C/CD94. HLA-E binds to the inhibiting receptors NKG2A, -B and activating receptor NKG2C, belonging to the killer cell lectin-like receptor C1 (KLRC1) family, expressed on NK cells. The NKG2A and -B receptors mediate an inhibitory signal to the NK cell via immunoreceptor tyrosine-based inhibition motifs (ITIMs) [<a href="#B51-ijms-20-01830" class="html-bibr">51</a>]. The activating receptor NKG2C does not possess an intracellular immunoreceptor tyrosine-based activating motif (ITAM), but contains a positively charged transmembrane domain and dimerize with DNAX activation protein 12 (DAP-12), which has an ITAM in its cytoplasmic domain and transmits an activating signal to the cell [<a href="#B51-ijms-20-01830" class="html-bibr">51</a>,<a href="#B54-ijms-20-01830" class="html-bibr">54</a>,<a href="#B55-ijms-20-01830" class="html-bibr">55</a>].</p>
Full article ">Figure 2
<p>Receptor interaction of HLA-F with KIR3DL1, -2, KIR3DS1 and -S4 as well as LILRB1 and -B2. The inhibiting receptors KIR3DL1 and -2 and activating receptors KIR3DS1 and KIR2DS4 belong to the family of killer cell immunoglobulin-like receptors (KIRs), which are expressed on NK cells. The inhibiting receptors KIR3DL1 and -2 have a long cytoplasmatic tail (L) with immunoreceptor tyrosine-based inhibition motifs (ITIMs). The activating receptors KIR3DS1 and KIR2DS4 are classified by their number of extracellular domains (two or three domains, 2D or 3D) and short (S) intracellular cytoplasmatic tail, which contains a charged lysine residue instead of an immunoreceptor tyrosine-based inhibition motif (ITIM). They dimerize with DNAX activation protein 12 (DAP-12), which has an immunoreceptor tyrosine-based activating motif (ITAM). The inhibitory leukocyte-immunoglobulin (Ig)-like receptors (LILR) LILRB1 (also known as Ig-like transcript 2; ILT2) and -B2 (also known as ILT4) are expressed on monocytes, dendritic cells (DCs), as well as on B-, T-, and NK cells and mediate an inhibitory signal via their ITIMs.</p>
Full article ">Figure 3
<p>Receptor interaction of HLA-G with CD8, KIR2DL4, LILRB1, LILRB2 and CD160. CD8 is a marker for cytotoxic T-cells and consists of an extracellular alpha and beta domain [<a href="#B88-ijms-20-01830" class="html-bibr">88</a>]. The receptor KIR2DL4 belongs to the family of killer cell immunoglobulin-like receptors (KIRs), which are expressed on NK cells. The receptor contains two extracellular domains D0 and D1 and has only one immunoreceptor tyrosine-based inhibition motifs (ITIM). A charged arginine residue in its cytoplasmatic tail enables KIR2DL4 to form a complex with Fc fragment receptor γ (FcRγ), which stimulates cytokine production in the NK cell [<a href="#B55-ijms-20-01830" class="html-bibr">55</a>]. The inhibitory leukocyte-immunoglobulin (Ig)-like receptors (LILR) LILRB1 (also known as Ig-like transcript 2; ILT2) and -B2 (also known as ILT4) are expressed on monocytes, dendritic cells (DCs), as well as on B-, T-, and NK cells and mediate an inhibitory signal via their ITIMs [<a href="#B36-ijms-20-01830" class="html-bibr">36</a>,<a href="#B39-ijms-20-01830" class="html-bibr">39</a>,<a href="#B74-ijms-20-01830" class="html-bibr">74</a>]. CD160 is a glycosylphosphatidylinositol-anchored receptor and does not contain an immunoreceptor tyrosine-based activating motif (ITAM) [<a href="#B89-ijms-20-01830" class="html-bibr">89</a>].</p>
Full article ">Figure 4
<p>Hypotheses for an immunological tumor therapy concept (ITTC). The patent focuses on determining the individual expression pattern of the “embryonic” HLA genes on primary tumors to create individual therapy approaches such as antibodies drug conjugates, vaccination/immunization approaches and in situ gene editing to block and downregulate HLA class Ib expression in order to overcome immune evasion mediated by the non-classical HLA groups. The patent also implies the monitoring of the HLA expression patterns on recurrent tumors and metastases for subsequent treatment adaptations.</p>
Full article ">
24 pages, 1756 KiB  
Review
IL-36 Cytokines: Regulators of Inflammatory Responses and Their Emerging Role in Immunology of Reproduction
by José Martin Murrieta-Coxca, Sandra Rodríguez-Martínez, Mario Eugenio Cancino-Diaz, Udo R. Markert, Rodolfo R. Favaro and Diana M. Morales-Prieto
Int. J. Mol. Sci. 2019, 20(7), 1649; https://doi.org/10.3390/ijms20071649 - 3 Apr 2019
Cited by 50 | Viewed by 6853
Abstract
The IL-36 subfamily of cytokines has been recently described as part of the IL-1 superfamily. It comprises three pro-inflammatory agonists (IL-36α, IL-36β, and IL-36γ), their receptor (IL-36R), and one antagonist (IL-36Ra). Although expressed in a variety of cells, the biological relevance of IL-36 [...] Read more.
The IL-36 subfamily of cytokines has been recently described as part of the IL-1 superfamily. It comprises three pro-inflammatory agonists (IL-36α, IL-36β, and IL-36γ), their receptor (IL-36R), and one antagonist (IL-36Ra). Although expressed in a variety of cells, the biological relevance of IL-36 cytokines is most evident in the communication between epithelial cells, dendritic cells, and neutrophils, which constitute the common triad responsible for the initiation, maintenance, and expansion of inflammation. The immunological role of IL-36 cytokines was initially described in studies of psoriasis, but novel evidence demonstrates their involvement in further immune and inflammatory processes in physiological and pathological situations. Preliminary studies have reported a dynamic expression of IL-36 cytokines in the female reproductive tract throughout the menstrual cycle, as well as their association with the production of immune mediators and cellular recruitment in the vaginal microenvironment contributing to host defense. In pregnancy, alteration of the placental IL-36 axis has been reported upon infection and pre-eclampsia suggesting its pivotal role in the regulation of maternal immune responses. In this review, we summarize current knowledge regarding the regulatory mechanisms and biological actions of IL-36 cytokines, their participation in different inflammatory conditions, and the emerging data on their potential role in normal and complicated pregnancies. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>IL-1 superfamily. The IL-1 superfamily is comprised of 11 cytokines with pro-inflammatory or anti-inflammatory activity divided into IL-1, IL-18, and IL-36 subfamilies. All members, except IL-1Ra, are synthesized as long precursor proteins, which are proteolytically cleaved by the indicated enzymes to generate their active mature proteins.</p>
Full article ">Figure 2
<p>IL-36 signaling pathway. IL-36α, IL-36β, and IL-36γ bind to IL-36R leading to its dimerization with IL-1RAcP. The active receptor triggers intracellular signaling cascades involving MyD88, IRAK, and MAPK to induce NF-kB- and AP-1-dependent expression of pro-inflammatory cytokines, chemokines, and secondary mediators of the inflammatory response. IL-36Ra and IL-38 bind to IL-36R inhibiting receptor dimerization and activation.</p>
Full article ">Figure 3
<p>Origin and effects of IL-36 cytokines on different cell types. IL-36 cytokines are released by epithelial cells and processed by neutrophil-derived proteases to generate their active forms. The activated cytokines stimulate dendritic cells to enhance pro-inflammatory mediators responsible for several cellular effects in innate and adaptive immune responses. Some processes regulated on immune and non-immune cells are indicated. Note that there is still no data for trophoblast cells.</p>
Full article ">Figure 4
<p>Illustrative description of the pro-inflammatory (red line) and anti-inflammatory (blue line) balance and IL-36 expression during pregnancy. Deregulated inflammatory reactions (red dotted line) may occur at any time in pregnancy and may overlap with each other. Depending on the stage, distinct complications can arise.</p>
Full article ">
12 pages, 890 KiB  
Review
Effects of Chemotherapy and Radiotherapy on Spermatogenesis: The Role of Testicular Immunology
by Ning Qu, Masahiro Itoh and Kou Sakabe
Int. J. Mol. Sci. 2019, 20(4), 957; https://doi.org/10.3390/ijms20040957 - 22 Feb 2019
Cited by 55 | Viewed by 6416
Abstract
Substantial improvements in cancer treatment have resulted in longer survival and increased quality of life in cancer survivors with minimized long-term toxicity. However, infertility and gonadal dysfunction continue to be recognized as adverse effects of anticancer therapy. In particular, alkylating agents and irradiation [...] Read more.
Substantial improvements in cancer treatment have resulted in longer survival and increased quality of life in cancer survivors with minimized long-term toxicity. However, infertility and gonadal dysfunction continue to be recognized as adverse effects of anticancer therapy. In particular, alkylating agents and irradiation induce testicular damage that results in prolonged azoospermia. Although damage to and recovery of spermatogenesis after cancer treatment have been extensively studied, there is little information regarding the role of differences in testicular immunology in cancer treatment-induced male infertility. In this review, we briefly summarize available rodent and human data on immunological differences in chemotherapy or radiotherapy. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Immunological differences in the testes after cancer treatment.</p>
Full article ">Figure 2
<p>Testes weights and epididymal spermatozoa numbers in normal and busulfan-/irradiation-treated mice.</p>
Full article ">
23 pages, 1074 KiB  
Review
The Impact of Autoantibodies on IVF Treatment and Outcome: A Systematic Review
by Mara Simopoulou, Konstantinos Sfakianoudis, Evangelos Maziotis, Sokratis Grigoriadis, Polina Giannelou, Anna Rapani, Petroula Tsioulou, Agni Pantou, Theodoros Kalampokas, Nikolaos Vlahos, Konstantinos Pantos and Michael Koutsilieris
Int. J. Mol. Sci. 2019, 20(4), 892; https://doi.org/10.3390/ijms20040892 - 19 Feb 2019
Cited by 44 | Viewed by 8857
Abstract
The role of autoantibodies in in vitro fertilization (IVF) has been discussed for almost three decades. Nonetheless, studies are still scarce and widely controversial. The aim of this study is to provide a comprehensive systematic review on the possible complications associated to autoantibodies [...] Read more.
The role of autoantibodies in in vitro fertilization (IVF) has been discussed for almost three decades. Nonetheless, studies are still scarce and widely controversial. The aim of this study is to provide a comprehensive systematic review on the possible complications associated to autoantibodies (AA) impeding the chances of a successful IVF cycle. An Embase, PubMed/Medline and Cochrane Central Database search was performed on 1 December 2018, from 2006 until that date. From the 598 articles yielded in the search only 44 relevant articles ultimately fulfilled the inclusion criteria and were qualitatively analyzed. Five subsets of results were identified, namely, thyroid related AA, anti-phospholipid antibodies, anti-nuclear antibodies, AA affecting the reproductive system and AA related to celiac disease. It may be implied that the majority of auto-antibodies exert a statistically significant effect on miscarriage rates, whereas the effects on clinical pregnancy and live birth rates differ according to the type of auto-antibodies. While significant research is performed in the field, the quality of evidence provided is still low. The conduction of well-designed prospective cohort studies is an absolute necessity in order to define the impact of the different types of autoantibodies on IVF outcome. Full article
(This article belongs to the Special Issue Reproductive Immunology: Cellular and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>PRISMA flowchart.</p>
Full article ">Figure 2
<p>Boxplot regarding CP, LB/OP and miscarriage rates between TAA positive and TAA negative women. Blue color represent CP rate, grey LB/OP rate and magenta represent miscarriage rates. Circles in the figure represent outliers -values outside the 1.5x interquartile range.</p>
Full article ">Figure 3
<p>Boxplot regarding CP, LB/OP and miscarriage rates between aPL positive and aPL negative women. Blue color represent CP rate, grey LB/OP rate and magenta represent miscarriage rate. Circles in the figure represent outliers -values outside the 1.5x interquartile range.</p>
Full article ">Figure 4
<p>Boxplot regarding CP, LB/OP and miscarriage rates between ANA positive and ANA negative women. Blue color represents CP rate, grey LB/OP rate and magenta represents miscarriage rates. Circles in the figure represent outliers -values outside the 1.5x interquartile range.</p>
Full article ">
Back to TopTop