Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (64)

Search Parameters:
Keywords = bacterial luciferase

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
11 pages, 2666 KiB  
Article
Increased Autonomous Bioluminescence Emission from Mammalian Cells by Enhanced Cofactor Synthesis
by Theresa Brinker and Carola Gregor
Chemosensors 2024, 12(11), 223; https://doi.org/10.3390/chemosensors12110223 - 25 Oct 2024
Viewed by 569
Abstract
The bacterial bioluminescence system has been successfully implemented in mammalian cell lines, enabling the substrate-free luminescence imaging of living cells. One of the major limitations of the system is its comparatively low brightness. To improve light emission, we aimed to increase the cellular [...] Read more.
The bacterial bioluminescence system has been successfully implemented in mammalian cell lines, enabling the substrate-free luminescence imaging of living cells. One of the major limitations of the system is its comparatively low brightness. To improve light emission, we aimed to increase the cellular production of FMNH2 and NADPH, which serve as cosubstrates in the bacterial bioluminescence reaction. We coexpressed different proteins involved in the synthesis of these two cofactors together with the proteins of the bacterial bioluminescence system in different mammalian cell lines. The combined expression of a riboflavin kinase (RFK) and a constitutively active Akt2 variant (Akt2CA) that participate in the cellular production of FMN and NADP+, respectively, increased bioluminescence emission up to 2.4-fold. The improved brightness allows autonomous bioluminescence imaging of mammalian cells at a higher signal-to-noise ratio and enhanced spatiotemporal resolution. Full article
(This article belongs to the Special Issue Chemiluminescent and Bioluminescent Sensors)
Show Figures

Figure 1

Figure 1
<p>Reactions involved in the light generation by the bacterial bioluminescence system in mammalian cells. (<b>A</b>) Bacterial luciferase catalyzes the oxidation of FMNH<sub>2</sub> and a fatty aldehyde (RCHO) to FMN and a fatty acid (RCOOH) with concomitant photon emission. The two products are regenerated by a flavin reductase and the fatty acid reductase complex, respectively. (<b>B</b>) Riboflavin (RF) is absorbed by the cell and converted into FMN by riboflavin kinase (RFK). (<b>C</b>) Glucose-6-phosphate dehydrogenase (G6PD) catalyzes the reduction of NADP<sup>+</sup> to NADPH. The enzyme is regulated by sirtuin 2 (SIRT2) that deacetylates and thereby activates G6PD. (<b>D</b>) NADP<sup>+</sup> is produced from NAD<sup>+</sup> by NAD kinase (NADK). NADK is activated by phosphorylation by protein kinase B (Akt).</p>
Full article ">Figure 2
<p>The effect of overexpression of different proteins on the bioluminescence emission in mammalian cell lines: (<b>A</b>) LiveLight HEK293; (<b>B</b>) HEK293; (<b>C</b>) HeLa. Cells grown in 24-well plates were transfected with a mixture of 0.4 µg lux plasmids and 0.1 µg of the indicated genes (all constructs in pcDNA3.1(+)). The signal was normalized to the bioluminescence emission of cells transfected with 0.4 µg lux plasmids and 0.1 µg of the empty pcDNA3.1(+) vector (−). For LiveLight HEK293, 0.5 µg of the indicated gene in pcDNA3.1(+) was transfected. Error bars represent standard deviation from 5 wells.</p>
Full article ">Figure 3
<p>Fluorescence ratio of iNap1 in Lux-expressing cells with and without Akt2CA. (<b>A</b>) LiveLight HEK293; (<b>B</b>) HEK293; (<b>C</b>) HeLa cells grown on coverslips were transfected with a mixture of 0.6 µg lux plasmids, 0.2 µg iNap1 pcDNA3.1 Hygro(+) and 0.2 µg Akt2CA pcDNA3.1(+) or the empty pcDNA3.1(+) vector (–). For LiveLight HEK293, 0.2 µg iNap1 pcDNA3.1 Hygro(+) and 0.8 µg Akt2CA pcDNA3.1(+) or the empty pcDNA3.1(+) vector were transfected. iNap1 fluorescence excited at 405 (F405) and 491 nm (F491) was recorded with a custom-built microscope. Error bars represent standard deviation from at least 50 cells. ** and *** represent <span class="html-italic">p</span> values of &lt;0.01 and &lt;0.001, respectively, calculated by a 2-tailed Student’s <span class="html-italic">t</span> test.</p>
Full article ">Figure 4
<p>Effect of combined expression of RFK and Akt2CA on the bioluminescence emission in different mammalian cell lines: (<b>A</b>) LiveLight HEK293; (<b>B</b>) HEK293; (<b>C</b>) HeLa. Cells were grown in 24-well plates and transfected with a mixture of 0.4 µg lux plasmids and 0.1 µg of the indicated constructs (all in pcDNA3.1(+)). For RFK + Akt2CA, two separate plasmids containing RFK and Akt2CA were cotransfected (0.05 µg each). The signal was normalized to cells transfected with 0.4 µg lux plasmids and 0.1 µg of the empty pcDNA3.1(+) vector (−). For LiveLight HEK293, 0.5 µg of the indicated constructs was transfected. Error bars represent standard deviation from 5 wells.</p>
Full article ">Figure 5
<p>MTT assay of cells transfected with Akt2CA-P2A-RFK. (<b>A</b>) LiveLight HEK293; (<b>B</b>) HEK293; (<b>C</b>) HeLa cells grown in 24-well plates were transfected with a mixture of 0.1 µg Akt2CA-P2A-RFK and 0.4 µg of the empty vector or 0.5 µg of the empty vector (−) (all constructs in pcDNA3.1(+)). For LiveLight HEK293, 0.5 µg of the empty vector or the Akt2CA-P2A-RFK plasmid was transfected. Absorbance of cell lysates was measured at 570 nm. Error bars represent standard deviation from 10 wells.</p>
Full article ">Figure 6
<p>Bioluminescence of HeLa cells with and without Akt2CA-P2A-RFK. Cells grown on coverslips were transfected with a mixture of 0.8 µg lux plasmids and (<b>A</b>) 0.2 µg empty pcDNA3.1(+) vector or (<b>B</b>) 0.2 µg Akt2CA-P2A-RFK (all constructs in pcDNA3.1(+)). Bioluminescence emission was recorded using the indicated exposure times. The colormap was adjusted to the minimum and maximum pixel value of each image.</p>
Full article ">
19 pages, 4790 KiB  
Article
Pig Milk Exosome Packaging ssc-miR-22-3p Alleviates Pig Intestinal Epithelial Cell Injury and Inflammatory Response by Targeting MAPK14
by Jie Li, Huihui Hu, Panpan Fu, Qiaoli Yang, Pengfei Wang, Xiaoli Gao, Jiaojiao Yang, Shuangbao Gun and Xiaoyu Huang
Int. J. Mol. Sci. 2024, 25(19), 10715; https://doi.org/10.3390/ijms251910715 - 5 Oct 2024
Viewed by 1073
Abstract
Inflammatory diseases of the intestinal tract in piglets severely impair the economic performance of pig farms. Pig milk exosomes can encapsulate miRNAs which can then enter the piglet intestine to play an immunomodulatory role. Previously, we comparatively analyzed and identified exosomal miRNAs in [...] Read more.
Inflammatory diseases of the intestinal tract in piglets severely impair the economic performance of pig farms. Pig milk exosomes can encapsulate miRNAs which can then enter the piglet intestine to play an immunomodulatory role. Previously, we comparatively analyzed and identified exosomal miRNAs in the colostrum and mature milk of Bamei and Landrace pigs, and we screened for ssc-miR-22-3p, which is associated with inflammation and immune response; however, the role played by ssc-miR-22-3p in the immune response in IPEC-J2 cells is not yet clear. In this study, we first constructed a pig intestinal inflammatory response model using Lipopolysaccharide (LPS) and Polyinosinic-polycytidylic acid (Poly (I:C)), and we investigated the role of ssc-miR-22-3p targeting MAPK14 in the regulation of LPS and Poly (I:C)-induced inflammatory injury in IPEC-J2 cells by RT-qPCR, cell counting kit-8 (CCK-8), EdU staining, lactate dehydrogenase (LDH) activity assay, and dual luciferase reporter gene assay. We successfully established LPS and Poly (I:C)-induced cell damage models in IPEC-J2 cells. The immune response of IPEC-J2 cells was stimulated by induction of IPEC-J2 cells at 10 μg/mL LPS and 20 μg/mL Poly (I:C) for 24 h. Overexpression of ssc-miR-22-3p decreased cytokine expression and promoted cell viability and proliferation. The functional enrichment analysis revealed that ssc-miR-22-3p targets genes enriched in the pathways of negative regulation of inflammatory response and bacterial invasion of epithelial cells. The validity of the binding site of ssc-miR-22-3p to MAPK14 was tested by a dual luciferase reporter gene. Pig milk exosome ssc-miR-22-3p promotes cell viability and proliferation by targeting MAPK14, and it alleviates LPS and Poly (I:C)-induced inflammatory responses in IPEC-J2 cells. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Effects of LPS and Poly(I:C) on IPEC-J2 cell viability and cytokine expression. (<b>A</b>) Effects of different concentrations of LPS on IPEC-J2 viability. (<b>B</b>–<b>F</b>) Effects of different concentrations of LPS on IPEC-J2 <span class="html-italic">TLR4</span>, <span class="html-italic">IL-6</span>, <span class="html-italic">IL-8</span>, and <span class="html-italic">TNF-α</span> mRNA expression, respectively. (<b>G</b>) Effects of different concentrations of Poly(I:C) on IPEC-J2 viability. (<b>H</b>–<b>L</b>) Effect of different concentrations of Poly(I:C) on IPEC-J2 <span class="html-italic">TLR3</span>, <span class="html-italic">IL-6</span>, <span class="html-italic">IL-8</span>, and <span class="html-italic">TNF-α</span> mRNA expression, respectively. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, ns indicates no significant difference.</p>
Full article ">Figure 2
<p>The relative expression levels of ssc-miR-22-3p in LPS and Poly (I:C)-induced IPEC-J2 cells. (<b>A</b>) Sequence conservation analysis of miR-22-3p matrices from different species. (<b>B</b>,<b>C</b>) Relative expression levels of ssc-miR-22-3p after stimulation of IPEC-J2 cells with LPS and Poly (I:C), respectively. ** <span class="html-italic">p</span> &lt; 0.01. The red arrow indicates the seed sequence.</p>
Full article ">Figure 3
<p>ssc-miR-22-3p transfection efficiency assay. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Effect of ssc-miR-22-3p on the effect of LPS (<b>A</b>) and Poly (I:C) (<b>B</b>)-induced inflammatory responses in IPEC-J2 cells. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, ns means no significant difference.</p>
Full article ">Figure 5
<p>Effect of ssc-miR-22-3p on LPS (<b>A</b>) and Poly (I:C) (<b>B</b>)-induced IPEC-J2 cell cytotoxicity. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Analysis of ssc-miR-22-3p on LPS and Poly (I:C)-induced IPEC-J2 cell viability and proliferation. (<b>A</b>,<b>B</b>) indicates the effect of ssc-miR-22-3p on the viability of LPS and Poly (I:C)-induced IPEC-J2 cells, respectively. (<b>C</b>,<b>D</b>) indicates the effect of ssc-miR-22-3p on the number of LPS and Poly (I:C)-induced IPEC-J2 positive cells, respectively. (<b>E</b>,<b>F</b>) EdU cell proliferation assay. Scale bar = 200 µm. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, ns indicates no significant difference.</p>
Full article ">Figure 7
<p>Functional enrichment analysis of ssc-miR-22-3p. (<b>A</b>) ssc-miR-22-3p and mRNA gene interaction network diagram. The ellipse represents mRNA, the triangle represents miRNA, and the color of the ellipse represents the size of the free energy for miRNA to form with the mRNA sequence. (<b>B</b>,<b>C</b>) represents GO and KEGG enrichment analysis of ssc-miR-22-3p target genes, respectively. (<b>D</b>) RT-qPCR verification of immune-related target genes. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, ns indicates no significant difference.</p>
Full article ">Figure 8
<p>Identification of the validity of the targeting relationship between ssc-miR-22-3p and <span class="html-italic">MAPK14</span>. (<b>A</b>) <span class="html-italic">MAPK14</span> recombinant vector enzyme cleavage map. Lane 1: red circle indicates self-labeled empty vector before digestion, yellow indicates recombinant vector before digestion; Lane 2: vector and target fragment after digestion with HindIII; Lane 3: 1 kb ladder. (<b>B</b>) Plot of Sanger sequencing peaks of pmirGLO-<span class="html-italic">MAPK14</span>-WT and pmirGLO-<span class="html-italic">MAPK14</span>-MUT. (<b>C</b>) The targeting relationship between ssc-miR-22-3p and <span class="html-italic">MAPK14</span> was detected by dual luciferase reporter gene. ** <span class="html-italic">p</span> &lt; 0.01, ns indicates no significant difference.</p>
Full article ">Figure 9
<p>Schematic diagram of pmirGLO-<span class="html-italic">MAPK14</span>-3′ UTR recombinant vector construction.</p>
Full article ">
16 pages, 10115 KiB  
Article
Functional Analysis of RMA3 in Response to Xanthomonas citri subsp. citri Infection in Citron C-05 (Citrus medica)
by Mingming Zhao, Rongchun Ye, Yi Li, Lian Liu, Hanying Su, Xianfeng Ma and Ziniu Deng
Horticulturae 2024, 10(7), 693; https://doi.org/10.3390/horticulturae10070693 - 1 Jul 2024
Viewed by 804
Abstract
Citrus bacterial canker disease, caused by Xanthomonas citri subsp. citri (Xcc), poses a significant global threat to the citrus industry. Lateral organ boundaries 1 (Lob1) is confirmed as a citrus susceptibility gene that induces pathogenesis by interaction with the [...] Read more.
Citrus bacterial canker disease, caused by Xanthomonas citri subsp. citri (Xcc), poses a significant global threat to the citrus industry. Lateral organ boundaries 1 (Lob1) is confirmed as a citrus susceptibility gene that induces pathogenesis by interaction with the PthA4 effector of Xcc. Citron C-05 (Citrus medica) is a Citrus genotype resistant to Xcc. However, there is little information available on the regulation of Lob1 in resistant genotypes, which is important for the breeding of citrus cultivars resistant to canker disease. This study aimed to identify upstream regulatory factors of Lob1 in Citron C-05 and to investigate its function in disease resistance. ‘Bingtang’ sweet orange (C. sinensis), a susceptible genotype, was utilized as the control. cDNA yeast libraries of Xcc-induced Citron C-05 and ‘Bingtang’ sweet orange were constructed. The capacities of ‘Bingtang’ and Citron C-05 were 1.896 × 107 and 2.154 × 107 CFU, respectively. The inserted fragments ranged from 500 to 2000 bp with a 100% recombination rate. The promoter of Lob1 was segmented into two pieces and the P1 fragment from both genotypes was used to construct a bait yeast (PAbAi-CsLob1-P1; PAbAi-CmLob1-P1). Through library screening with the bait yeast, upstream regulators interacting with the Lob1-P1 promoter were identified and then validated using Y1H and dual-luciferase tests. The expression analysis of the three transcript factors indicated that RMA3 was upregulated by inoculation with Xcc in the resistant Citron C-05, but not in the susceptible sweet orange. The overexpression of CsRMA3 in ‘Bingtang’ sweet orange led to reduced canker symptoms, with a significantly lower pathogen density in the leaves following Xcc inoculation. When CmRMA3 was silenced by virus-induced gene silencing (VIGS) in Citron C-05, typical canker symptoms appeared on the CmRMA3-silenced leaves at 15 days post-inoculation with Xcc. Further expression analyses revealed that the CmRMA3 transcription factor suppressed the expression of Lob1. These results suggest that RMA3 participates in the resistant reaction of Citron C-05 to Xcc infection, and such a response might be in relation to its suppression of the expression of the pathogenic gene Lob1. Full article
(This article belongs to the Section Plant Pathology and Disease Management (PPDM))
Show Figures

Figure 1

Figure 1
<p>Leaf symptoms on the two citrus genotypes (‘Bingtang’ sweet orange = susceptible, Citron C-05 = resistant) after inoculation with 10<sup>4</sup> CFU/mL <span class="html-italic">Xanthomonas citri</span> subsp. <span class="html-italic">citri.</span> dpi = days post-inoculation.</p>
Full article ">Figure 2
<p>Validation of the interacting protein in Citron C-05 by <span class="html-italic">CmLob1</span>-P1 in Y1H system. pGADT7-AD: empty vector introduced into the same yeast cells as the negative control; pGADT7-p53: plasmid containing the p53 antibody introduced into yeast cells as the positive control; ‘+’: yeast containing Citron C-05 cDNA able to grow normally in SD/-LEU medium, representing the candidate protein interacting with <span class="html-italic">CmLob1</span>-P1; ‘−’: yeast did not grow in SD/-LEU medium, meaning that the protein had no interaction with <span class="html-italic">CmLob1</span>-P1. The concentration of AbA was 200 ng/mL.</p>
Full article ">Figure 3
<p>Identification of the interacting proteins in sweet orange and <span class="html-italic">CsLob1</span>-P1 in the Y1H system. pGADT7-AD: empty vector introduced into the same yeast cells as the negative control; pGADT7-p53: plasmid containing the p53 antibody introduced into yeast cells as the positive control. ‘+’: yeast containing sweet orange cDNA able to grow normally in SD/-LEU medium, indicating that the candidate proteins interact with <span class="html-italic">CmLob1</span>-P1; ‘−’: the same yeast did not grow in SD/-LEU medium, meaning that the protein had no interaction with <span class="html-italic">CmLob1</span>-P1. The concentration of AbA was 200 ng/mL.</p>
Full article ">Figure 4
<p>Identification of proteins interacting with the <span class="html-italic">Lob1-P1</span> promoter by dual-luciferase assay. The CmNAC transcription factors had very low expression, even though they significantly reduced the expression of Lob1. CmTFGTE did not show the suppression of the expression of Lob1. The CmRMA3 and CsRMA3 transcription factors bound to the promoter of <span class="html-italic">Lob1</span>, thereby significantly downregulating the expression of the <span class="html-italic">Lob1</span> gene. CsNAC and CsTFGTE were not analyzed for interactions with <span class="html-italic">CsLob1</span>-P1. ‘**’ indicates a significant difference (<span class="html-italic">p</span> &lt; 0.01). ‘n.s.’ indicates no significant difference.</p>
Full article ">Figure 5
<p>Relative expression of three candidate genes in ‘Bingtang’ sweet orange (susceptible) and Citron C-05 (resistant) following inoculation with 10<sup>5</sup> CFU/mL of <span class="html-italic">Xanthomonas citri</span> subsp. <span class="html-italic">citri</span>.</p>
Full article ">Figure 6
<p>The subcellular localization of CmRMA3 of Citron C-05. Yellow arrows indicate the nucleus; the red arrow points to the cell membrane; PM-RB is a membrane-targeting mark; 35S::GFP: an empty control, localized to the cell membrane and nucleus.</p>
Full article ">Figure 7
<p>Effects on the inhibition of <span class="html-italic">Xanthomonas citri</span> subsp<span class="html-italic">. citri</span> (<span class="html-italic">Xcc</span>) infection following <span class="html-italic">RMA3</span> overexpression in susceptible ‘Bingtang’ sweet orange leaves. On the left, upon inoculation with <span class="html-italic">Xcc</span>, canker symptoms were observed in sweet orange plants with the transient overexpression of the <span class="html-italic">RMA3</span> gene; on the right, quantitative counting plate count data of <span class="html-italic">Xcc</span> bacteria. The concentration of <span class="html-italic">Xcc</span> was 10<sup>5</sup> CFU/mL. ‘**’ indicates a significant difference between ‘EV + <span class="html-italic">Xcc</span>’ and ‘35S::<span class="html-italic">CsRMA3</span> + <span class="html-italic">Xcc</span>’ (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 8
<p>Relative expression of the <span class="html-italic">Lob1</span> gene in citrus leaves following the transient overexpression of <span class="html-italic">RMA3</span> at 3 dpi. (<b>A</b>) Leaf of ‘Bingtang’ sweet orange; (<b>B</b>) leaf of Citron C-05. The concentration of <span class="html-italic">Xanthomonas citri</span> subsp<span class="html-italic">. citri</span> was 10<sup>5</sup> CFU/mL. ‘**’ indicates a significant difference(<span class="html-italic">p</span> &lt; 0.01) between the transient clone and the control.</p>
Full article ">Figure 9
<p>Relative expression of the <span class="html-italic">RMA3</span> gene in Citron C-05 with <span class="html-italic">RMA3</span> silenced by VIGS. CK-1 to CK-4 are clones transferred with the Empty-TRV2 plasmid as controls; VIGS-1 to VIGS-6 are <span class="html-italic">RMA3</span>-silenced plants transferred with <span class="html-italic">RMA3</span>-TRV2. The different letters in the columns indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Relative expression of <span class="html-italic">Lob1</span> in Citron C-05 with <span class="html-italic">RMA3</span> silenced by VIGS. On the left, the plants were not inoculated with <span class="html-italic">Xanthomonas citri</span> subsp. <span class="html-italic">citri</span> sampled at the same time as the inoculated ones; on the right, the plants were inoculated with <span class="html-italic">Xcc</span> and evaluated at 3 days post-inoculation. The concentration of <span class="html-italic">Xcc</span> was 10<sup>5</sup> CFU/mL. CK-1 to CK-4 are clones transferred with the Empty-TRV2 plasmid as controls; VIGS-1 to VIGS-6 are <span class="html-italic">RMA3</span>-silenced plants. The different letters in the columns indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 11
<p>Leaf canker symptom observation of <span class="html-italic">RMA3</span>-silenced Citron C-05 at 15 days post-inoculation with <span class="html-italic">Xanthomonas citri</span> subsp. <span class="html-italic">citri</span>. (<b>A</b>) Leaf canker symptoms of <span class="html-italic">RMA3</span>-silenced Citron C-05. (<b>B</b>) Leaf symptoms of ‘Bingtang’ sweet orange as a positive control for disease susceptibility genotypes. The inoculation concentration of <span class="html-italic">Xcc</span> was 10<sup>5</sup> CFU/mL.</p>
Full article ">
151 KiB  
Abstract
A Lab-on-Paper Biosensor for ATP Quantification via a Chemiluminescent DNA Nanoswitch Assay
by Elisa Lazzarini, Alessandro Porchetta, Donato Calabria, Andrea Pace, Ilaria Trozzi, Martina Zangheri, Massimo Guardigli and Mara Mirasoli
Proceedings 2024, 104(1), 9; https://doi.org/10.3390/proceedings2024104009 - 28 May 2024
Viewed by 343
Abstract
Water is indispensable for life, yet many lack access to clean drinking water, resulting in fatalities from waterborne bacterial infections. Precise assessment of microbial abundance and viability in natural aquatic environments is vital. Adenosine triphosphate (ATP) serves as a parameter for viability assessments [...] Read more.
Water is indispensable for life, yet many lack access to clean drinking water, resulting in fatalities from waterborne bacterial infections. Precise assessment of microbial abundance and viability in natural aquatic environments is vital. Adenosine triphosphate (ATP) serves as a parameter for viability assessments due to its presence in viable bacterial cells as an energy carrier. Traditional ATP detection methods involve chemical or enzymatic extraction, followed by measurement of light emission via the Luciferin–Luciferase complex. However, these methods are costly, present a low stability, require specialized equipment, and entail complex sample pretreatment. To overcome these limitations, we developed a biosensor based on aptamers, nucleic acid sequences with specific target-molecule-binding capabilities. Aptamers offer advantages such as an enhanced stability, a lower cost, and ease of design compared to antibodies. Recently, ATP has been used for aptamer selection testing. Our proposed biosensor utilizes a structure-switching ATP-binding DNA nanoswitch with two functional domains: a catalytic DNA-zyme domain and an ATP-binding aptamer domain. In the presence of ATP, its binding to the aptamer domain triggers the activation of the DNA-zyme domain, which is exploited for chemiluminescence (CL) detection. Integrating functional DNA biosensors with microfluidic paper-based analytical devices (µPADs) holds promise for point-of-care (POC) applications. However, achieving proper DNA binding on paper remains challenging, often requiring solution-based assay protocols, leaving µPADs for final signal readout. Here, we introduce an origami µPAD with preloaded dried reagents, allowing for on-paper assay execution upon sample addition and proper folding. Paper functionalization strategies and assay protocols were optimized to ensure simple and straightforward detection of ATP, employing a portable charge-coupled device (CCD) camera for CL detection. Calibration curves plotted against the logarithm of ATP concentration in the range of 1 to 500 µM facilitated determination of the assay’s limit of detection (LOD), which was found to be 3 µM. Full article
(This article belongs to the Proceedings of The 4th International Electronic Conference on Biosensors)
15 pages, 4162 KiB  
Article
A Variety of Mouse PYHIN Proteins Restrict Murine and Human Retroviruses
by Sümeyye Erdemci-Evin, Matteo Bosso, Veronika Krchlikova, Wibke Bayer, Kerstin Regensburger, Martha Mayer, Ulf Dittmer, Daniel Sauter, Dorota Kmiec and Frank Kirchhoff
Viruses 2024, 16(4), 493; https://doi.org/10.3390/v16040493 - 23 Mar 2024
Viewed by 1359
Abstract
PYHIN proteins are only found in mammals and play key roles in the defense against bacterial and viral pathogens. The corresponding gene locus shows variable deletion and expansion ranging from 0 genes in bats, over 1 in cows, and 4 in humans to [...] Read more.
PYHIN proteins are only found in mammals and play key roles in the defense against bacterial and viral pathogens. The corresponding gene locus shows variable deletion and expansion ranging from 0 genes in bats, over 1 in cows, and 4 in humans to a maximum of 13 in mice. While initially thought to act as cytosolic immune sensors that recognize foreign DNA, increasing evidence suggests that PYHIN proteins also inhibit viral pathogens by more direct mechanisms. Here, we examined the ability of all 13 murine PYHIN proteins to inhibit HIV-1 and murine leukemia virus (MLV). We show that overexpression of p203, p204, p205, p208, p209, p210, p211, and p212 strongly inhibits production of infectious HIV-1; p202, p207, and p213 had no significant effects, while p206 and p214 showed intermediate phenotypes. The inhibitory effects on infectious HIV-1 production correlated significantly with the suppression of reporter gene expression by a proviral Moloney MLV-eGFP construct and HIV-1 and Friend MLV LTR luciferase reporter constructs. Altogether, our data show that the antiretroviral activity of PYHIN proteins is conserved between men and mice and further support the key role of nuclear PYHIN proteins in innate antiviral immunity. Full article
(This article belongs to the Special Issue Innate Sensing and Restriction of Retroviruses)
Show Figures

Figure 1

Figure 1
<p>Mammalian PYHIN gene locus. Syntenic blocks containing PYHIN paralogs from selected mammalian species are shown on the right. Arrows indicate the direction of the open reading frames (ORFs). Genes annotated as orthologs are shown in identical colors. Genes annotated as PYHIN1-, MNDA-, or PYDC3-like genes are marked with an asterisk “*”. Note that identical names do not necessarily indicate true orthologs. In mice, most genes are shown in black as it is not possible to clearly identify orthologs in other species. CADM3 has been included as a reference gene flanking the PYHIN gene locus. Gene lengths, distances, and overlaps are not drawn to scale, and some overlapping and additional ORFs were omitted for clarity. A time-calibrated phylogenetic tree of mammalian evolution generated with TimeTree [<a href="#B26-viruses-16-00493" class="html-bibr">26</a>] is shown on the left. MYA—million years ago.</p>
Full article ">Figure 2
<p>Domain organization of human and murine PYHIN proteins. Most members of the PYHIN protein family are characterized by a Pyrin domain (orange) at their N-terminus and a C-terminal HIN domain. Structural variations within the HIN domain delineate it into three subtypes: HIN-A, HIN-B, and HIN-C (shown in yellow, green, and purple, respectively) and a linker region (gray) with potential nuclear localization signals (NLSs) indicated with circles. Modified from Cridland and colleagues [<a href="#B7-viruses-16-00493" class="html-bibr">7</a>].</p>
Full article ">Figure 3
<p>Expression of PYHIN proteins and inhibition of HIV-1. (<b>a</b>) Expression of human and mouse PYHIN proteins. HEK293T cells were transfected with IRES-BFP constructs expressing the indicated HA-tagged PYHIN proteins. GAPDH was used as loading control. (<b>b</b>) Infectious virus yields in cell supernatants of HEK293T cells cotransfected with vectors expressing the indicated PYHIN proteins and the proviral HIV-1 CH058 construct were determined by TZM-bl cell infection assay at two days post-transfection. Data show mean percentages (±SEM; <span class="html-italic">n</span> = 3). (<b>c</b>) Levels of viral env transcripts in HEK293T cells cotransfected with expression constructs for the indicated PYHIN proteins and the proviral HIV-1 CH058 construct determined by qRT-PCR. Data show mean percentages (±SEM) from 3 experiments, each performed in technical duplicates relative to those detected in the presence of the empty control plasmid (100%). * <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. (<b>d</b>–<b>f</b>) Correlations between the effects of overexpression of PYHIN proteins on infectious HIV-1 yield and (<b>d</b>) the levels of PYHIN proteins detected by Western blot, (<b>e</b>) <span class="html-italic">env</span> RNA transcripts, and (<b>f</b>) p24 antigen levels. Values in panels d and e were derived from the results shown in panels (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 4
<p>Effect of human and mouse PYHIN proteins on basal HIV-1 LTR activity. (<b>a</b>) HEK293T cells were cotransfected with a construct containing the firefly luciferase reporter gene under the control of the HIV-1 NL4-3 LTR. Values show mean percentages (±SEM) relative to those detected in the absence of PYHIN protein (100%) and were derived from three 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; 0.001. (<b>b</b>,<b>c</b>) Correlation between the effects of PYHIN protein overexpression on HIV-1 LTR activity and (<b>b</b>) infectious HIV-1 yield and (<b>c</b>) env RNA levels.</p>
Full article ">Figure 5
<p>Effect of human and mouse PYHIN proteins on MLV transcription and LTR activity. (<b>a</b>) HEK293T cells were cotransfected with either a control vector or plasmids expressing HA-tagged mouse and human PYHIN proteins in a pCG IRES BFP vector together with an env-defective proviral Moloney MLV (M-MLV) eGFP construct. Shown is the eGFP mean fluorescence intensity (MFI) in BFP/GFP double-positive cells. Shown are mean values (±SEM) from three independent experiments in the presence of PYHINs relative to the empty vector control (100%). * <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. (<b>b</b>) HEK293T cells were cotransfected with a Friend MLV (F-MLV) LTR luciferase reporter construct and vectors expressing the indicated PYHIN proteins. Values show mean percentages (±SEM) relative to those detected in the absence of PYHIN protein (100%) and were derived from three or four experiments. (<b>c</b>–<b>f</b>) Correlation between the effects of overexpression of mouse and human PYHIN proteins on M-MLV-driven eGFP expression, F-MLV and HIV-1 LTR activity, and infectious HIV-1 yield as indicated.</p>
Full article ">Figure 6
<p>Sp1 enhances MLV gene expression but is not essential for inhibition by mouse PYHIN proteins. (<b>a</b>) HEK293T cells were cotransfected with increasing amounts of expression constructs for Sp1 IRES BFP (0.001, 0.02, 0.1, and 0.5 µg) and MLV eGFP construct. BFP and eGFP expression was analyzed by flow cytometry. The left panel shows primary FACS data, and the right panel shows the mean fluorescence intensities (MFIs) of eGFP in BFP expressing population. (<b>b</b>) HEK293T cells were cotransfected with a firefly luciferase reporter construct under the control of FV LTR and a vector expressing an increasing amount of Sp1 (0.001, 0.005, 0.02, and 0.1 µg) or an empty control. In addition, a Gaussia luciferase reporter construct was transfected as an internal control. Shown are firefly relative to Gaussia luciferase values measured in the presence of Sp1 overexpression; <span class="html-italic">n</span> = 3 ± SEM. (<b>c</b>) HEK293T cells were cotransfected with an NL4-3 LTR firefly luciferase reporter construct containing mutated Sp1 binding sites and expression constructs for the indicated PYHIN proteins. Values show mean percentages (±SEM) of firefly luciferase expression relative to those detected in the absence of PYHIN protein (100%) and were derived from three 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; 0.001.</p>
Full article ">
16 pages, 2908 KiB  
Article
Delivery of a Hepatitis C Virus Vaccine Encoding NS3 Linked to the MHC Class II Chaperone Protein Invariant Chain Using Bacterial Ghosts
by Yulang Chi, Shikun Zhang and Shouping Ji
Biomedicines 2024, 12(3), 525; https://doi.org/10.3390/biomedicines12030525 - 26 Feb 2024
Cited by 2 | Viewed by 1543
Abstract
Efficient delivery of a DNA plasmid into antigen-presenting cells (APCs) is a potential strategy to enhance the immune responses of DNA vaccines. The bacterial ghost (BG) is a potent DNA vaccine delivery system that targets APCs. In the present work, we describe a [...] Read more.
Efficient delivery of a DNA plasmid into antigen-presenting cells (APCs) is a potential strategy to enhance the immune responses of DNA vaccines. The bacterial ghost (BG) is a potent DNA vaccine delivery system that targets APCs. In the present work, we describe a new strategy of using E. coli BGs as carriers for an Ii-linked Hepatitis C Virus (HCV) NS3 DNA vaccine that improved both the transgene expression level and the antigen-presentation level in APCs. BGs were prepared from DH5α cells, characterized via electron microscopy and loaded with the DNA vaccine. The high transfection efficiency mediated using BGs was first evaluated in vitro, and then, the immune protective effect of the BG-Ii-NS3 vaccine was determined in vivo. It was found that the antibody titer in the sera of BG-Ii-NS3-challenged mice was higher than that of Ii-NS3-treated mice, indicating that the BGs enhanced the humoral immune activity of Ii-NS3. The cellular immune protective effect of the BG-Ii-NS3 vaccine was determined using long-term HCV NS3 expression in a mouse model in which luciferase was used as a reporter for HCV NS3 expression. Our results showed that the luciferase activity in BG-Ii-NS3-treated mice was significantly reduced compared with that in Ii-NS3-treated mice. The CTL assay results demonstrated that BG-Ii-NS3 induced a greater NS3-specific T-cell response than did Ii-NS3. In summary, our study demonstrated that BGs enhanced both the humoral and cellular immune response to the Ii-NS3 DNA vaccine and improved its immune protection against HCV infection. Full article
(This article belongs to the Section Microbiology in Human Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Schematic of the construction of the Ii-linked <span class="html-italic">NS3</span> plasmid. (<b>A</b>) The DNA fragment of the HCV-<span class="html-italic">NS3</span> Th1 epitope (amino acids 1248–1261) was subcloned into the <span class="html-italic">EcoRI/XbaI</span> sites of the pVAX1 plasmid. (<b>B</b>) The mIi-linked <span class="html-italic">NS3</span> was constructed by ligating the mouse invariant chain (mIi, 41 KDa) coding region with the CLIP coding region of mIi substituted by the HCV-<span class="html-italic">NS3</span> Th1 epitope into the <span class="html-italic">HindⅢ</span> and <span class="html-italic">XhoI</span> sites of the pVAX1 plasmid shown in (<b>A</b>).</p>
Full article ">Figure 2
<p>Preparation and morphological characterization of the DH5α BGs. (<b>A</b>) Induction of E protein-mediated cell lysis at 42 °C. (<b>B</b>) Scanning electron microscopy (SEM) of the DH5α BGs, (<b>C</b>) Transmission electron microscopy (TEM) of the DH5α cells, (<b>D</b>) TEM of the DH5α BGs. The magnification of the SEM and TEM images was 24,000×.</p>
Full article ">Figure 3
<p>Fluorescence confocal microscopy and transmission electron microscopy of <span class="html-italic">E. coli</span> ghosts loaded with the pVAX1-Ii-<span class="html-italic">NS3</span> Th1 plasmid. (<b>A</b>–<b>C</b>), Fluorescence confocal microscopy images: (<b>A</b>) MitoTracker-labeled (green) <span class="html-italic">E. coli</span> ghosts, (<b>B</b>) PI-labeled plasmid, (<b>C</b>) Overlay of the differential interference contrast image and fluorescence images (<b>A</b>,<b>B</b>). (<b>D</b>,<b>E</b>) Transmission electron microscopy images of BGs and BGs loaded with plasmids, respectively.</p>
Full article ">Figure 4
<p>FACS analyses of the green fluorescent MitoTracker-labeled BGs and the red fluorescent PI-labeled plasmid. (<b>A</b>) Non-labeled control BGs, (<b>B</b>) BGs labeled with MitoTracker, (<b>C</b>) MitoTracker-labeled BGs loaded with PI-labeled pGL3 plasmids. (<b>D</b>) MitoTracker-labeled BGs loaded with the PI-labeled pVAX1-Ii-<span class="html-italic">NS3</span> Th1 plasmid. The <span class="html-italic">X</span>-axis shows the green fluorescence intensity of the MitoTracker-labeled BGs plasmid, while the <span class="html-italic">Y</span>-axis shows the red fluorescence intensity of the PI-labeled.</p>
Full article ">Figure 5
<p>The uptake of bacterial ghosts by RAW 264.7 cells. (<b>A</b>) Fluorescent signal in the transfected cells, (<b>B</b>) Bright field image of (<b>A</b>), The magnification was 24,000×. (<b>C</b>,<b>D</b>) Fluorescent signal in the control cells and transfected cells, respectively. (<b>E</b>) Overlay of the plots for (<b>C</b>,<b>D</b>). (<b>F</b>) Luciferase activity in RAW 264.7 cells at 24 h after transfection. DNA: cells transfected with free plasmid DNA, BG/DNA: cells transfected with BGs loaded with plasmid DNA. PEI/DNA: cells transfected with PEI. Each set of data represents the mean of three independent experiments with error bars indicating the SD.</p>
Full article ">Figure 6
<p>The endpoint titer of anti-NS3 antibodies in vaccinated mice. BALB/c mice were inoculated with BGs loaded with the Ii-<span class="html-italic">NS3</span> plasmid (BG-Ii-<span class="html-italic">NS3</span>) or with the naked Ii-<span class="html-italic">NS3</span> plasmid (Ii-<span class="html-italic">NS3</span>). Empty BGs and the pVAX1 plasmid were used for the negative control groups. Each titer value represents the mean of three independent experiments with error bars indicating the SD.</p>
Full article ">Figure 7
<p>In vivo validation of the immune response. (<b>A</b>) Schematic of the construction of the plasmidpAA-<span class="html-italic">att</span>B-<span class="html-italic">NS3</span>/4A-Fluc. (<b>B</b>) Bioluminescence imaging of luciferase expression in BALB/C mice immunized with BG-Ii-<span class="html-italic">NS3</span>, Ii-<span class="html-italic">NS3</span> or the empty vector, respectively, 1 d and 10 d post-hydrodynamic injection with pAA-<span class="html-italic">att</span>B-<span class="html-italic">NS3</span>/4A-Fluc and Pphic31φ. (<b>C</b>) In vivo induction of T-cell responses against the NS3-restricted epitopes after immunization with Ii-<span class="html-italic">NS3</span> or BG-Ii-<span class="html-italic">NS3</span>. Splenocytes from BALB/C mice immunized i.p with Ii-<span class="html-italic">NS3</span> or with BG-Ii-<span class="html-italic">NS3</span> were stimulated with the indicated CTL peptides. The CTL activity of the splenocytes was determined using the cytotox96 non-radioactive cytotoxicity assay. The analyses were conducted 14 days after immunization. Each cytotoxicity value represents the mean of three independent experiments with error bars indicating the SD.</p>
Full article ">
13 pages, 2566 KiB  
Article
The Ralstonia solanacearum Type III Effector RipAW Targets the Immune Receptor Complex to Suppress PAMP-Triggered Immunity
by Zhi-Mao Sun, Qi Zhang, Yu-Xin Feng, Shuang-Xi Zhang, Bi-Xin Bai, Xue Ouyang, Zhi-Liang Xiao, He Meng, Xiao-Ting Wang, Jun-Min He, Yu-Yan An and Mei-Xiang Zhang
Int. J. Mol. Sci. 2024, 25(1), 183; https://doi.org/10.3390/ijms25010183 - 22 Dec 2023
Cited by 1 | Viewed by 1516
Abstract
Bacterial wilt, caused by Ralstonia solanacearum, one of the most destructive phytopathogens, leads to significant annual crop yield losses. Type III effectors (T3Es) mainly contribute to the virulence of R. solanacearum, usually by targeting immune-related proteins. Here, we clarified the effect [...] Read more.
Bacterial wilt, caused by Ralstonia solanacearum, one of the most destructive phytopathogens, leads to significant annual crop yield losses. Type III effectors (T3Es) mainly contribute to the virulence of R. solanacearum, usually by targeting immune-related proteins. Here, we clarified the effect of a novel E3 ubiquitin ligase (NEL) T3E, RipAW, from R. solanacearum on pathogen-associated molecular pattern (PAMP)-triggered immunity (PTI) and further explored its action mechanism. In the susceptible host Arabidopsis thaliana, we monitored the expression of PTI marker genes, flg22-induced ROS burst, and callose deposition in RipAW- and RipAWC177A-transgenic plants. Our results demonstrated that RipAW suppressed host PTI in an NEL-dependent manner. By Split-Luciferase Complementation, Bimolecular Fluorescent Complimentary, and Co-Immunoprecipitation assays, we further showed that RipAW associated with three crucial components of the immune receptor complex, namely FLS2, XLG2, and BIK1. Furthermore, RipAW elevated the ubiquitination levels of FLS2, XLG2, and BIK1, accelerating their degradation via the 26S proteasome pathway. Additionally, co-expression of FLS2, XLG2, or BIK1 with RipAW partially but significantly restored the RipAW-suppressed ROS burst, confirming the involvement of the immune receptor complex in RipAW-regulated PTI. Overall, our results indicate that RipAW impairs host PTI by disrupting the immune receptor complex. Our findings provide new insights into the virulence mechanism of R. solanacearum. Full article
(This article belongs to the Special Issue Advances of Plants-Pathogen Interaction 2023)
Show Figures

Figure 1

Figure 1
<p>Generation of <span class="html-italic">RipAW</span>- and <span class="html-italic">RipAW</span><sup>C177A</sup>-transgenic <span class="html-italic">Arabidopsis thaliana</span> and evaluation of disease resistance in these transgenic plants. (<b>A</b>) The growth phenotype of <span class="html-italic">RipAW</span>- and <span class="html-italic">RipAW<sup>C177A</sup></span>-transgenic plants and the expression of transgenes verified by RT-PCR. (<b>B</b>,<b>C</b>) The disease index (<b>B</b>) and survival rate (<b>C</b>) of <span class="html-italic">GFP</span>-, <span class="html-italic">RipAW</span>-, and <span class="html-italic">RipAW<sup>C177A</sup></span>-transgenic plants that were soil-drenching inoculated with GMI1000. (<b>D</b>) The bacterial population in plants that were Petri-dish inoculated with GMI1000. (<b>E</b>) The bacterial population in plants that were leaf-inoculated with DC3000. (<b>F</b>) The bacterial population in plants that were Petri-dish inoculated with GMI1000 <span class="html-italic">HrcV<sup>−</sup></span>. (<b>G</b>) The bacterial population in plants that were leaf-inoculated with DC3000 <span class="html-italic">HrcC<sup>−</sup></span>. Values are means ± standard errors (SEs) (n = 12 for (<b>B</b>,<b>C</b>), and n = 6 for (<b>D</b>–<b>G</b>)). Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>RipAW impairs PTI responses in <span class="html-italic">A. thaliana</span>. (<b>A</b>) The time course curve of flg22-induced ROS burst and total ROS accumulation in <span class="html-italic">RipAW</span>-, <span class="html-italic">RipAW<sup>C177A</sup></span>- and <span class="html-italic">GFP</span>-transgenic plants. (<b>B</b>) Relative expression levels of PTI marker genes, <span class="html-italic">WRKY29</span> and <span class="html-italic">FRK1</span>, in 4-week-old <span class="html-italic">RipAW</span>-, <span class="html-italic">RipAW<sup>C177A</sup></span>-, and <span class="html-italic">GFP</span>-transgenic plants. Leaves were treated with 1 μM flg22 for 3 h, and gene expression levels at 0 h and 3 h were measured by qRT-PCR. <span class="html-italic">ACTIN</span> was used as an internal reference gene. (<b>C</b>) Callose deposition levels of 4-week-old <span class="html-italic">RipAW</span>-, <span class="html-italic">RipAW<sup>C177A</sup></span>-, and <span class="html-italic">GFP</span>-transgenic plants. Leaves were infiltrated with 1 μM flg22 at 12 h before sampling. Scale bars represent 200 μm. (<b>D</b>) The defense priming assay in <span class="html-italic">RipAW</span>-, <span class="html-italic">RipAW<sup>C177A</sup></span>-, and <span class="html-italic">GFP</span>-transgenic plants. Four-week-old plants were pre-treated with 1 μM flg22 12 h before <span class="html-italic">P. syringae</span> DC3000 inoculation. Bacterial population was determined at 2 days post-inoculation. Values are means ± SEs (n = 6). Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>RipAW associates with the immune receptor complex and decreases the protein accumulation of FLS2, XLG2, and BIK1. (<b>A</b>) Split-luciferase complementation assay evaluating the association of RipAW-nLUC with XLG2-cLUC, FLS2-cLUC, or BIK1-cLUC. The combination of BIK1-nLUC and XLG2-cLUC served as a positive control, while the combination of BIK1-nLUC and CPR5-cLUC acted as a negative control. (<b>B</b>) Bimolecular fluorescence complementation assay detecting the interaction between RipAW-YC and FLS2-YN, XLG2-YN, or BIK1-YN. The combination of EV-YN and EV-YC was used as a negative control. YFP indicated the YFP fluorescence and bright-field images showed the plant cell structure. The merged images showed the overlay of YFP fluorescence on the bright field. (<b>C</b>) Association of RipAW with the immune receptor complex detected by co-immunoprecipitation assay. FLS2-HA, XLG2-HA, or BIK1-HA was co-expressed with RipAW-FLAG in 5-week-old <span class="html-italic">N. benthamiana</span> leaves. Samples were collected at 36 h after infiltration, and the protein extracts were subjected to co-immunoprecipitation using anti-FLAG beads. The protein expression levels were checked in the input. The association between RipAW and FLS2, XLG2, or BIK1 was detected by an anti-HA antibody in immunoprecipitation. (<b>D</b>) FLS2 protein accumulation; (<b>E</b>) XLG2 protein accumulation; (<b>F</b>) BIK1 protein accumulation. FLS2-HA, XLG2-HA, or BIK1-HA was co-expressed with RipAW-FLAG, RipAW<sup>C177A</sup>-FLAG, or LTI6b-MYC in 5-week-old <span class="html-italic">N. benthamiana</span> leaves. Samples were collected at 36 h after infiltration.</p>
Full article ">Figure 4
<p>RipAW accelerates degradation of FLS2, XLG2, and BIK1 through the 26S proteasome pathway. (<b>A</b>–<b>C</b>) Protein accumulation of FLS2 (<b>A</b>), XLG2 (<b>B</b>), and BIK1 (<b>C</b>) co-expressed with RipAW or RipAW<sup>C177A</sup> under DMSO or MG132 treatment. (<b>D</b>–<b>F</b>) Ubiquitination levels of FLS2 (<b>D</b>), XLG2 (<b>E</b>), and BIK1 (<b>F</b>) when co-expressed with RipAW, RipAW<sup>C177A</sup>, or LTI6b under MG132 treatment.</p>
Full article ">Figure 5
<p>Expression of FLS2, XLG2, or BIK1 partially reverses the defect in ROS burst caused by RipAW in <span class="html-italic">N. benthamiana.</span> (<b>A</b>) Effect of FLS2 expression on ROS burst. (<b>B</b>) Effect of BIK1 expression on ROS burst. (<b>C</b>) Effect of XLG2 expression on ROS burst. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
15 pages, 2616 KiB  
Article
Development of a Replication-Deficient Bacteriophage Reporter Lacking an Essential Baseplate Wedge Subunit
by Jose Gil, John Paulson, Henriett Zahn, Matthew Brown, Minh M. Nguyen and Stephen Erickson
Viruses 2024, 16(1), 8; https://doi.org/10.3390/v16010008 - 20 Dec 2023
Viewed by 1412
Abstract
Engineered bacteriophages (phages) can be effective diagnostic reporters for detecting a variety of bacterial pathogens. Although a promising biotechnology, the large-scale use of these reporters may result in the unintentional release of genetically modified viruses. In order to limit the potential environmental impact, [...] Read more.
Engineered bacteriophages (phages) can be effective diagnostic reporters for detecting a variety of bacterial pathogens. Although a promising biotechnology, the large-scale use of these reporters may result in the unintentional release of genetically modified viruses. In order to limit the potential environmental impact, the ability of these phages to propagate outside the laboratory was targeted. The phage SEA1 has been previously engineered to facilitate food safety as an accurate and sensitive reporter for Salmonella contamination. In this study, homologous recombination was used to replace the expression of an essential baseplate wedge subunit (gp141) in SEA1 with a luciferase, NanoLuc®. This reporter, referred to as SEA1Δgp141.NL, demonstrated a loss of plaque formation and a failure to increase in titer following infection of Salmonella. SEA1Δgp141.NL was thus incapable of producing infectious progeny in the absence of gp141. In contrast, production of high titer stocks was possible when gp141 was artificially supplied in trans during infection. As a reporter, SEA1Δgp141.NL facilitated rapid, sensitive, and robust detection of Salmonella despite an inability to replicate. These results suggest that replication-deficient reporter phages are an effective method to obtain improved containment without sacrificing significant performance or the ease of production associated with many phage-based diagnostic methods. Full article
(This article belongs to the Special Issue Biotechnological Applications of Phage and Phage-Derived Proteins 4.0)
Show Figures

Figure 1

Figure 1
<p>Homologs of <span class="html-italic">Escherichia</span> phage T4’s essential baseplate wedge subunit and nearby proteins are found in <span class="html-italic">Salmonella</span> phage SEA1. Colors distinguish each set of homologs between T4 (NC_000866) and SEA1 (OQ927978), with arrows indicating calculated protein similarity. Protein similarity was determined using EMBOSS Needle [<a href="#B32-viruses-16-00008" class="html-bibr">32</a>]. The protein product information from T4’s GenBank<sup>®</sup> entry was included below each gene.</p>
Full article ">Figure 2
<p>Homologous recombination was used to generate SEA1∆gp141.NL. Infection of <span class="html-italic">S. enterica</span> carrying the pUC57∆gp141.NL plasmid with SEA1 was performed at a multiplicity of infection of 0.1 to generate the desired recombinant, SEA1∆gp141.NL. After the initial infection, <span class="html-italic">S. enterica</span> carrying the complementation plasmid pUC57.Comp.gp141 was used to supply gp141 in trans to permit isolation and production of SEA1∆gp141.NL. Consistent with <a href="#viruses-16-00008-f001" class="html-fig">Figure 1</a>, genes were assigned a color to aid in visualization. Arrows in the genomic sequence indicate sites where bacterial promoters were added.</p>
Full article ">Figure 3
<p>SEA1∆gp141.NL was confirmed by PCR and was deficient for plaque formation in the absence of gp141. PCR analysis (<b>a</b>) was performed on DNA preparations from SEA1.NL and SEA1∆gp141.NL using primers specific for gp141. Ladder refers to O’GeneRuler 1 kb Plus DNA ladder. An original uncropped gel image is provided elsewhere (<a href="#app1-viruses-16-00008" class="html-app">Supplementary Figure S1</a>). Plaque formation (<b>b</b>) and spotting activity was determined by spotting 4 µL of each serially diluted phage preparation onto a double-layer agar containing <span class="html-italic">S. enterica</span>. The number “0” refers to a working stock concentration of 1 × 10<sup>11</sup> pfu/mL. Ten-fold serial dilutions of this stock are subsequently notated as 10<sup>x</sup>. When indicated, the <span class="html-italic">S. enterica</span> host carried the gp141 complementation plasmid pUC57.Comp.gp141.</p>
Full article ">Figure 4
<p>A one-step growth curve establishes gp141 as essential for SEA1 replication. Infection of a <span class="html-italic">S. enterica</span> with either SEA1.NL (<b>a</b>) or SEA1∆gp141.NL (<b>b</b>) was performed, and plaque-forming units (PFU) were monitored over a 90-min period. When indicated (<b>c</b>), <span class="html-italic">S. enterica</span> expressing gp141 in trans was similarly infected and monitored. To circumvent the deficiency in plaque formation, samples of SEA1∆gp141.NL (<b>b</b>,<b>c</b>) were plated on the gp141-expressing host strain to determine PFU. The latent period was defined as the time prior to an observed increase in phage titer. The burst size was defined as the fold difference between the average PFU during the latent period and average PFU following the plateau of growth. The <span class="html-italic">y</span>-axis was set at the limit of detection for the experiment (100 PFU). Samples that yielded no plaques and are thus below the limit of detection are marked with an arrow and indicated as not detected (ND). Individual plaque counts are provided elsewhere (<a href="#app1-viruses-16-00008" class="html-app">Supplementary Table S1</a>).</p>
Full article ">Figure 5
<p>Both SEA1.NL and SEA1∆gp141.NL are capable of detecting <span class="html-italic">Salmonella</span>. Various colony-forming units (CFU) of <span class="html-italic">S. enterica</span> were infected for (<b>a</b>) 30 min, (<b>b</b>) 60 min, (<b>c</b>) 90 min, or (<b>d</b>) 120 min with either SEA1.NL or SEA1∆gp141.NL before being assessed for NanoLuc<sup>®</sup> production. Luminescence was evaluated as log transformed mean relative light units (Log<sub>10</sub> RLU) generated from triplicate wells. Error bars represent the standard deviation of these log-transformed values. Individual RLU values have been made available (<a href="#app1-viruses-16-00008" class="html-app">Supplementary Table S2</a>). Two-way ANOVAs with Dunnett’s multiple comparison tests were performed using GraphPad Prism. For each phage, asterisks denote a significant difference (<span class="html-italic">p</span> &lt; 0.05) between the luminescence at that CFU and the corresponding background (0 CFU).</p>
Full article ">
13 pages, 2084 KiB  
Article
Acidovorax citrulli Type IV Pili PilR Interacts with PilS and Regulates the Expression of the pilA Gene
by Yuwen Yang, Weiqin Ji, Pei Qiao, Nuoya Fei, Linlin Yang, Wei Guan and Tingchang Zhao
Horticulturae 2023, 9(12), 1296; https://doi.org/10.3390/horticulturae9121296 - 30 Nov 2023
Cited by 1 | Viewed by 1305
Abstract
Acidovorax citrulli can cause bacterial fruit blotch of watermelon, melon, and other cucurbits, and has the potential to cause severe economic losses to growers throughout the world. This article investigated the functions and interactions of the pilR and pilS genes, two important genes [...] Read more.
Acidovorax citrulli can cause bacterial fruit blotch of watermelon, melon, and other cucurbits, and has the potential to cause severe economic losses to growers throughout the world. This article investigated the functions and interactions of the pilR and pilS genes, two important genes in bacterial type IV pili systems, in A. citrulli. For each gene, deletion mutants and complementary strains were constructed via homologous recombination, and their phenotypes were determined. The results showed that the absence of pilR and pilS could significantly reduce the pathogenicity and twitching motility of A. citrulli while increasing the swimming motility, biofilm formation, and in vitro growth. Conversely, complementary strains were no different than the wild-type strain. Using quantitative reverse transcription PCR and promoter activity assays, we confirmed that the deletion of pilR and pilS genes leads to a significant decrease in the transcription level of pilA. Meanwhile, three methods including yeast two-hybrid, glutathione S-transferase pull-down, and luciferase complementation imaging assays were used to verify the direct interaction between the PilR and PilS proteins. These findings revealed the biological function of the pilR and pilS and confirms their regulatory role on pilA. Full article
(This article belongs to the Section Genetics, Genomics, Breeding, and Biotechnology (G2B2))
Show Figures

Figure 1

Figure 1
<p>Pathogenicity assay of <span class="html-italic">Acidovorax citrulli</span> wild-type strain Aac5, as well as ∆<span class="html-italic">pilR</span>, ∆<span class="html-italic">pilS</span>, and ∆<span class="html-italic">pilR</span>comp, ∆<span class="html-italic">pilS</span>comp, and a sterile water control (CK) on watermelon seedlings. (<b>a</b>) Symptoms of watermelon seedlings 10 days after inoculation (DAI). Plants were spray inoculated with a single strain with bacterial suspensions of 10<sup>8</sup> CFU/mL. (<b>b</b>) The disease index of the five treated watermelon seedlings at 10 DAI. Error bars represent standard error of the means of three replicated experiments. Asterisks (*) above a bar indicate strains with significant differences <span class="html-italic">p</span> &lt; 0.05 (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 2
<p>Evaluation of swimming and twitching motility of <span class="html-italic">Acidovorax citrulli</span>. (<b>a</b>) The twitching mobility of the <span class="html-italic">A. citrulli</span> wild-type Aac5, <span class="html-italic">pilR</span> and <span class="html-italic">pilS</span> mutants and complementary strains. The strains were cultured on KB medium plate at 28 °C for 48 h, outer halos around the strains were observed using an optical microscope, and the twitching motility were assessed. (<b>b</b>) The swimming mobility of the <span class="html-italic">A. citrulli</span> wild-type Aac5 strain, <span class="html-italic">pilR</span> and <span class="html-italic">pilS</span> mutants, and complementary strains. The wild type strain Aac5, Δ<span class="html-italic">pilR</span> and Δ<span class="html-italic">pilS</span>, and Δ<span class="html-italic">pilR</span>comp and Δ<span class="html-italic">pilS</span>comp strains were cultured on a 0.3% semi-solid KB medium plate at 28 °C for 36 h, and the swimming halo diameter of each strain was measured. (<b>c</b>) The swimming motility halo average diameter of the five strains. The different letters on the bar chart indicate significant differences in swimming motility between different treated calculated by ANOVA (Duncan, <span class="html-italic">p</span> = 0.05).</p>
Full article ">Figure 3
<p>Evaluation of biofilm formation of <span class="html-italic">Acidovorax citrulli</span> Aac5, the <span class="html-italic">pilR</span> and <span class="html-italic">pilS</span> mutants, and complementary strains. Bacterial suspensions with OD<sub>600</sub> = 0.3 were added to KB medium in a 24-well plate at a <span class="html-italic">v</span>/<span class="html-italic">v</span> ratio of 1:100 and incubated at 28 °C for 48 h. Crystal violet solution (0.1%) was used to stain biofilms for 45 min. The OD<sub>575</sub> value was measured after dissolving with 95% alcohol. (<b>a</b>) Typical biofilms on the inner wall of the culture wells. (<b>b</b>) The absorbance value of each treatment at OD<sub>575</sub>. Error bars represent standard error of the means of three replicated experiments. Treatments with the same letter are not significantly different from each other as calculated by an ANOVA test with an alpha level of 0.05.</p>
Full article ">Figure 4
<p>Evaluation of the growth of <span class="html-italic">Acidovorax citrulli</span> Aac5 wild-type, the <span class="html-italic">pilR</span> and <span class="html-italic">pilS</span> mutants and complementary strains in KB broth. (<b>a</b>) Growth curve of each strain; (<b>b</b>) Statistical analysis of OD<sub>600</sub> values of the 5 strains at different time points. Error bars represent standard error of the means of three replicated experiments. Treatments with an asterisk (*) are significantly different from each other as calculated by an ANOVA test with an alpha level of 0.05.</p>
Full article ">Figure 5
<p>Evaluation of the relative expression of genes in <span class="html-italic">Acidovorax citrulli</span> Aac5 wild-type, and the <span class="html-italic">pilR</span> and <span class="html-italic">pilS</span> deletion mutants. (<b>a</b>) Relative expression of the <span class="html-italic">pilA</span> gene among the wild-type and deletion mutant strains. Relative expression levels were assessed via qRT PCR, and each deletion mutant was statistically compared to the control wild-type strain. Bars represent mean values calculated from three independent replicates. Error bars represent standard deviation. (<b>b</b>) Relative activity of <span class="html-italic">pilA</span> promotor β-Glucuronidase (GUS) in the wild-type strain carrying pBBR-GUS-<span class="html-italic">pilA</span>p (WT-<span class="html-italic">pilA</span>p-GUS), the <span class="html-italic">pilR</span> and <span class="html-italic">pilS</span> deletion mutants carrying pBBR-GUS-<span class="html-italic">pilA</span>p (ΔpilR-<span class="html-italic">pilA</span>p-GUS, Δ<span class="html-italic">pilS</span>-<span class="html-italic">pilA</span>p-GUS), and the wild-type strain carrying pBBRNolacGUS (WT-GUS), which was used as a negative control. The error bar represents the standard deviation. Two asterisks (**) above a bar indicate strains with significant differences <span class="html-italic">p</span> &lt; 0.01 (Student’s <span class="html-italic">t</span>-test); *** on top of the bar indicates strains with significant differences in the GUS (ANOVA, <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Several assays evaluating the interactions of PilR and PilS proteins. (<b>a</b>) Yeast two-hybrid assay. 1. co-expresstion of <span class="html-italic">pilS</span>-pBT3-N and pPR3-N (self-activation verification); 2. co-expression of <span class="html-italic">pilS</span>-pBT3-N and pOst1-NubI (functional verification); 3. co-expression of <span class="html-italic">pilS</span>-pBT3-N and <span class="html-italic">pilR</span>-pPR3-N (experimental group); +. co-expressed with pTSU2-APP and NubG-Fe65 (positive control group); −. co-expression of pTSU2-APP and PPR3-N (negative control group). On double drop-out (DDO) plates, co-expression of <span class="html-italic">pilS-</span>pBT3-N and pPR3-N can grow, indicating the self-activation phenomenon of PilS; on triple drop-out (TDO) plates containing a concentration of 5 mM of 3<span class="html-italic">′</span>AT, co-expression of pilS-pBT3-N and pPR3-N cannot grow, indicating that 5 mM of 3<span class="html-italic">′</span>AT can effectively inhibit self activation without affecting its function (co-expression of pilS-pBT3-N and pOst1-NubI can grow), and co-expression of <span class="html-italic">pilS</span>-pBT3-N and <span class="html-italic">pilR</span>-pPR3-N can grow normally, indicating a direct interaction between PilS and PilR; co-expression of <span class="html-italic">pilS</span>-pBT3-N and <span class="html-italic">pilR</span>-pPR3-N cannot grow normally on quadruple drop-out (QDO) plates. (<b>b</b>) Glutathione-S-transferase (GST) pull-down assay. PilR-GST was incubated with equal amounts of purified PilS-His and precipitated by glutathione agarose. The presence of PilS-His in glutathione agarose-bound protein was detected via anti-His immunoblot. The interactions of GST and PilS-His were evaluated as negative controls. The experiment was independently repeated three times. (<b>c</b>) Luciferase complementation imaging (LCI) assay. The GV3101 <span class="html-italic">Agrobacterium</span> carrying PilS-Nluc or PilR-Cluc were co-injected into <span class="html-italic">N. benthamiana</span> leaves. The leaves were collected after 48 h and observed under a live imaging device (NightSHADE LB985; Berthold).</p>
Full article ">
15 pages, 3488 KiB  
Article
Molecular Cloning and Functional Characterization of the 5′ Regulatory Region of the SLC11A1 Gene from Yaks
by Yuqing Chong, Liping Wang, Bo Wang, Zhendong Gao, Ying Lu, Weidong Deng and Dongmei Xi
Animals 2023, 13(23), 3615; https://doi.org/10.3390/ani13233615 - 22 Nov 2023
Viewed by 1196
Abstract
The solute transport protein family 11 A1 (SLC11A1), also recognized as natural resistance-associated macrophage protein 1 (NRAMP1), represents a transmembrane protein encoded by the SLC11A1 gene. A variety of prior investigations have illuminated its involvement in conferring resistance or [...] Read more.
The solute transport protein family 11 A1 (SLC11A1), also recognized as natural resistance-associated macrophage protein 1 (NRAMP1), represents a transmembrane protein encoded by the SLC11A1 gene. A variety of prior investigations have illuminated its involvement in conferring resistance or susceptibility to bacterial agents, positioning it as a promising candidate gene for breeding disease-resistant animals. Yaks (Bos grunniens), renowned inhabitants of the Qinghai-Tibet Plateau in China, stand as robust ruminants distinguished by their adaptability and formidable disease resistance. Notwithstanding these unique traits, there is scant literature on the SLC11A1 gene in the yak population. Our inquiry commences with the cloning of the 5′ regulatory region sequence of the Zhongdian yak SLC11A1 gene. We employ bioinformatics tools to identify transcription factor binding sites, delineating pivotal elements like enhancers and cis-acting elements. To ascertain the promoter activity of this region, we amplify four distinct promoter fragments within the 5′ regulatory region of the yak SLC11A1 gene. Subsequently, we design a luciferase reporter gene vector containing four site-specific deletion mutations and perform transient transfection experiments. Through these experiments, we measure and compare the activity of disparate gene fragments located within the 5′ regulatory region, revealing regions bearing promoter functionality and discerning key regulatory elements. Our findings validate the promoter functionality of the 5′ regulatory region, offering preliminary insights into the core and principal regulatory segments of this promoter. Notably, we identified single nucleotide polymorphisms (SNPs) that may be associated with important regulatory elements such as NF-1 and NF-1/L. This study provides a theoretical framework for in-depth research on the function and expression regulation mechanism of the yak SLC11A1 gene. Full article
(This article belongs to the Collection Advances in Cattle Breeding, Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Electrophoresis of PCR products. Starting from the top left: <span class="html-italic">SLC11A1</span>-5′, <span class="html-italic">SLC11A1</span>-5′-1, <span class="html-italic">SLC11A1</span>-5′-2, <span class="html-italic">SLC11A1</span>-5′-3, <span class="html-italic">SLC11A1</span>-5′-4.</p>
Full article ">Figure 2
<p>Double electrophoresis of recombinant plasmid with PCR-amplified product electrophoresis. Starting from the left: PGL3/<span class="html-italic">SLC11A1</span>-5′-1, PGL3/<span class="html-italic">SLC11A1</span>-5′-2, PGL3/<span class="html-italic">SLC11A1</span>-5′-3, PGL3/<span class="html-italic">SLC11A1</span>-5′-4.</p>
Full article ">Figure 3
<p>Nucleotide sequence information of the yak <span class="html-italic">SLC11A1</span> gene 5′ regulatory region. +1 represents the transcription start site, the red text represents the predicted promoter region, the boxes represent the predicted transcription factor binding sites, and the uppercase letters represent the exon sequences.</p>
Full article ">Figure 4
<p>Sequence alignment of <span class="html-italic">SLC11A1</span> gene promoter and conservative binding sites of transcription factors in different species. Boxes of different colors are predicted conservative transcription factor binding sites.</p>
Full article ">Figure 5
<p>Map of SNPs’ mutation loci in the 5′ regulatory region of the yak <span class="html-italic">SLC11A1</span> gene. The rectangular box represents different genotypes, the two peaks on the peak map represent heterozygous genotypes, and the single peak represents homozygous genotypes.</p>
Full article ">Figure 6
<p>Relative fluorescence activity of the reporter gene with different length fragments after transfection. The difference between experimental groups with completely different letters is significant (<span class="html-italic">p</span> &lt; 0.05), while the difference between experimental groups with the same letter is not significant (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">
19 pages, 8169 KiB  
Article
CRFB5a, a Subtype of Japanese Eel (Anguilla japonica) Type I IFN Receptor, Regulates Host Antiviral and Antimicrobial Functions through Activation of IRF3/IRF7 and LEAP2
by Tianyu Wang, Peng Lin, Yilei Wang, Xiaojian Lai, Pengyun Chen, Fuyan Li and Jianjun Feng
Animals 2023, 13(19), 3157; https://doi.org/10.3390/ani13193157 - 9 Oct 2023
Cited by 2 | Viewed by 1858
Abstract
IFNAR1, one of the type I IFN receptors, is crucial to mammalian host defense against viral invasion. However, largely unknown is the immunological role of the fish teleost protein IFNAR1, also known as CRFB5. We have successfully cloned the whole cDNA of the [...] Read more.
IFNAR1, one of the type I IFN receptors, is crucial to mammalian host defense against viral invasion. However, largely unknown is the immunological role of the fish teleost protein IFNAR1, also known as CRFB5. We have successfully cloned the whole cDNA of the Japanese eel’s (Anguilla japonica) CRFB5a homolog, AjCRFB5a. The two fibronectin-3 domains and the transmembrane region (238–260 aa) of AjCRFB5a are normally present, and it shares a three-dimensional structure with zebrafish, Asian arowana, and humans. According to expression analyses, AjCRFB5a is highly expressed in all tissues found, particularly the liver and intestine. In vivo, Aeromonas hydrophila, LPS, and the viral mimic poly I:C all dramatically increased AjCRFB5a expression in the liver. Japanese eel liver cells were reported to express AjCRFB5a more strongly in vitro after being exposed to Aeromonas hydrophila or being stimulated with poly I: C. The membranes of Japanese eel liver cells contained EGFP-AjCRFB5a proteins, some of which were condensed, according to the results of fluorescence microscopy. Luciferase reporter assays showed that AjCRFB5a overexpression strongly increased the expression of immune-related genes in Japanese eel liver cells, such as IFN1, IFN2, IFN3, IFN4, IRF3, IRF5, and IRF7 of the type I IFN signaling pathway, as well as one of the essential antimicrobial peptides LEAP2, in addition to significantly inducing human IFN-promoter activities in HEK293 cells. Additionally, RNA interference (RNAi) data demonstrated that knocking down AjCRFB5a caused all eight of those genes to drastically lower their expression in Japanese eel liver cells, as well as to variable degrees in the kidney, spleen, liver, and intestine. Our findings together showed that AjCRFB5a participates in the host immune response to bacterial infection by inducing antimicrobial peptides mediated by LEAP2 and favorably modulates host antiviral immune responses by activating IRF3 and IRF7-driven type I IFN signaling pathways. Full article
(This article belongs to the Section Aquatic Animals)
Show Figures

Figure 1

Figure 1
<p>Japanese eel <span class="html-italic">Aj</span>CRFB5a gene cDNA and its deduced amino acid sequence. The nucleotide and amino acid sequences were numbered on the left. Light gray and dark gray represent fibronectin-3 domains, and the box represents the transmembrane region.</p>
Full article ">Figure 2
<p>Multiple comparisons of IFNAR1 amino acid sequences of Japanese eel <span class="html-italic">Aj</span>CRFB5a of other species. Structural domains are indicated by arrows, identical amino acid residues are denoted by “*”; similar amino acid residues are denoted by “: or .” for similar amino acid residues.</p>
Full article ">Figure 2 Cont.
<p>Multiple comparisons of IFNAR1 amino acid sequences of Japanese eel <span class="html-italic">Aj</span>CRFB5a of other species. Structural domains are indicated by arrows, identical amino acid residues are denoted by “*”; similar amino acid residues are denoted by “: or .” for similar amino acid residues.</p>
Full article ">Figure 3
<p>(<b>A</b>) Three-dimensional ribbon structure of IFN receptor (<b>a</b>) spatial structure simulation of Japanese eel <span class="html-italic">Aj</span>CRFB5a (<b>b</b>) spatial structure simulation of IFN receptor superposition, (<b>c</b>) spatial structure simulation of human IFNAR1, (<b>d</b>) spatial structure simulation of zebrafish CRFB5, (<b>e</b>) spatial structure simulation of Asian arowana CRFB5a. (<b>B</b>) The predicted conserved functional domains of <span class="html-italic">Aj</span>CRFB5a protein. FN-3 denotes the fibronectin type 3 structural domain, with TM as the transmembrane structural domain. The accession numbers are zebrafish (ABJ97310.1), Asian arowana (MW286829.1), and human (NP_000620.2).</p>
Full article ">Figure 4
<p><span class="html-italic">Aj</span>CRFB5a’s phylogenetic tree with other CRFB5a proteins from different animals. The phylogram was built using the Neighbor-joining (NJ) method using the MEGA11 software (Version 11.0.13).</p>
Full article ">Figure 5
<p>Subcellular localization of <span class="html-italic">Aj</span>CRFB5a. EGFP-<span class="html-italic">Aj</span>CRFB5a and EGFP-N1 empty vectors were electrotransfected into Japanese eel liver cells. The cells were stained with DAPI after 24 h of incubation and then examined and photographed under a confocal microscope.</p>
Full article ">Figure 6
<p>Relative expression levels of <span class="html-italic">Aj</span>CRFB5a gene in various tissues of healthy Japanese eel.</p>
Full article ">Figure 7
<p><span class="html-italic">Aj</span>CRFB5a gene expression in vivo in response to immunological challenge. <span class="html-italic">A. hydrophila</span> (<b>A</b>), LPS (<b>B</b>), or poly I:C (<b>C</b>) were intraperitoneally injected into Japanese eels, respectively. RT-qPCR was used to calculate <span class="html-italic">Aj</span>CRFB5a expression levels, and the fold increase was normalized to PBS control (n = 4). Statistical differences were marked with asterisks (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 8
<p>The levels of <span class="html-italic">Aj</span>CRFB5a gene expression in Japanese eel liver cells after PAMP stimulation or after infection with <span class="html-italic">A. hydrophila</span>. Japanese eel liver cells were stimulated using LPS, poly I:C, CpG-DNA, PGN (<b>A</b>), and three different doses of <span class="html-italic">A. hydrophila</span> or PBS as a control (<b>B</b>). <span class="html-italic">Aj</span>CRFB5a expression levels were determined using RT-qPCR, and the fold increase was standardized to PBS control (n = 4). Statistical differences were marked with asterisks (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 9
<p>Human IFNβ response is activated by ectopic expression of <span class="html-italic">Aj</span>CRFB5a in HEK293 cells (<b>A</b>) and FHM cells (<b>B</b>); <span class="html-italic">Aj</span>IFN3 (<b>C</b>) and NF-κB (<b>D</b>) response are activated by ectopic expression of <span class="html-italic">Aj</span>CRFB5a in HEK293 cells. Cells cultured in 48-well plates with 1 × 10<sup>6</sup> cells per well were transfected with 20 ng of pRL-TK reference plasmid, and 80 ng of luciferase reporter plasmid, plus 300 ng of pcDNA3.1His-<span class="html-italic">Aj</span>CRFB5a, or empty pcDNA3.1His as a control. At 24 h post-transfection, cells were harvested to assay for luciferase activity. Statistical differences in expression levels between each sample and the pcDNA3.1His control are indicated by an asterisk (**, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 10
<p>Expression pattern of immune-related genes after overexpression of CRFB5a. Japanese eel liver cells either received pcDNA3.1His-CRFB5a or pcDNA3.1His empty vector (control) during transfection. Western blot analysis using Anti-His antibody was performed to confirm the expression of pcDNA3.1His-<span class="html-italic">Aj</span>CRFB5a and pcDNA3.1His fusion proteins (<b>A</b>). The expression levels of CRFB5a were detected by RT-qPCR (<b>B</b>). mRNA expression levels of IFN1, IFN2, IFN3, IFN4, IRF3, IRF5, IRF7 were detected by RT-qPCR (<b>C</b>). Statistical differences in expression levels between pcDNA3.1His-CRFB5a group and pcDNA3.1His empty vector control group are indicated by asterisks (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 11
<p>Immune-related gene expression patterns following in vitro <span class="html-italic">Aj</span>CRFB5a gene silencing. By using RT-qPCR to measure <span class="html-italic">Aj</span>CRFB5a’s mRNA expression levels in Japanese eel liver cells, the knockdown of the CRFB5a gene (<b>A</b>) was verified. The mRNA expression levels of IFN1, IFN2, IFN3, IFN4, IRF3, IRF5, IRF7 in the cells were detected by RT-qPCR (<b>B</b>). Statistical differences in expression levels between dsCRFB5a group and dsEGFP control are indicated by asterisks (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 12
<p>Immune-related gene expression patterns following in vivo <span class="html-italic">Aj</span>CRFB5a gene silencing. To verify the knockdown of the <span class="html-italic">Aj</span>CRFB5a gene, the levels of mRNA expression of <span class="html-italic">Aj</span>CRFB5a in the kidney, spleen, liver, and intestine were found using RT-qPCR (<b>A</b>). RT-qPCR was used to identify the mRNA expression levels of IFN1, IFN2, IFN3, IFN4, IRF3, IRF5, IRF7 in the kidney (<b>B</b>), spleen (<b>C</b>), liver (<b>D</b>), and intestine (<b>E</b>). Statistical differences in expression levels between dsCRFB5a group and dsEGFP control group are indicated by asterisks (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 13
<p>Expression patterns of the LEAP2 gene following gene overexpression (<b>A</b>) and gene silencing (<b>B</b>). Statistical differences in expression levels between experimental and control group are indicated by asterisks (**, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
15 pages, 2736 KiB  
Article
In Vitro Impact of Fluconazole on Oral Microbial Communities, Bacterial Growth, and Biofilm Formation
by Louise Morais Dornelas-Figueira, Antônio Pedro Ricomini Filho, Roger Junges, Heidi Aarø Åmdal, Altair Antoninha Del Bel Cury and Fernanda Cristina Petersen
Antibiotics 2023, 12(9), 1433; https://doi.org/10.3390/antibiotics12091433 - 11 Sep 2023
Cited by 1 | Viewed by 1975
Abstract
Antifungal agents are widely used to specifically eliminate infections by fungal pathogens. However, the specificity of antifungal agents has been challenged by a few studies demonstrating antibacterial inhibitory effects against Mycobacteria and Streptomyces species. Here, we evaluated for the first time the potential [...] Read more.
Antifungal agents are widely used to specifically eliminate infections by fungal pathogens. However, the specificity of antifungal agents has been challenged by a few studies demonstrating antibacterial inhibitory effects against Mycobacteria and Streptomyces species. Here, we evaluated for the first time the potential effect of fluconazole, the most clinically used antifungal agent, on a human oral microbiota biofilm model. The results showed that biofilm viability on blood and mitis salivarius agar media was increased over time in the presence of fluconazole at clinically relevant concentrations, despite a reduction in biomass. Targeted PCR revealed a higher abundance of Veillonella atypica, Veillonella dispar, and Lactobacillus spp. in the fluconazole-treated samples compared to the control, while Fusobacterium nucleatum was reduced and Streptococcus spp were not significantly affected. Further, we tested the potential impact of fluconazole using single-species models. Our results, using Streptococcus mutans and Streptococcus mitis luciferase reporters, showed that S. mutans planktonic growth was not significantly affected by fluconazole, whereas for S. mitis, planktonic growth, but not biofilm viability, was inhibited at the highest concentration. Fluconazole’s effects on S. mitis biofilm biomass were concentration and time dependent. Exposure for 48 h to the highest concentration of fluconazole was associated with S. mitis biofilms with the most increased biomass. Potential growth inhibitory effects were further tested using four non-streptococcal species. Among these, the planktonic growth of both Escherichia coli and Granulicatella adiacens was inhibited by fluconazole. The data indicate bacterial responses to fluconazole that extend to a broader range of bacterial species than previously anticipated from the literature, with the potential to disturb biofilm communities. Full article
Show Figures

Figure 1

Figure 1
<p>Oral microbiota biofilm assay showing pH (<b>A</b>), biomass (<b>B</b>), CFU counts on blood agar (<b>C</b>) and on mitis salivarius agar (<b>D</b>); 24 h biofilms were exposed to fluconazole at 2.56 μg·mL<sup>−1</sup> (SA—salivary concentration) and 2000 μg·mL<sup>−1</sup> (MO—mouthrinse concentration). Samples without fluconazole were included as control. The horizontal lines represent mean values from three independent experiments with three parallels each. Mean values from each of the parallel samples are represented by colored circles; * indicates statistical difference (<span class="html-italic">p</span> &lt; 0.05) between groups; one-way ANOVA followed by the Tukey’s post hoc test.</p>
Full article ">Figure 1 Cont.
<p>Oral microbiota biofilm assay showing pH (<b>A</b>), biomass (<b>B</b>), CFU counts on blood agar (<b>C</b>) and on mitis salivarius agar (<b>D</b>); 24 h biofilms were exposed to fluconazole at 2.56 μg·mL<sup>−1</sup> (SA—salivary concentration) and 2000 μg·mL<sup>−1</sup> (MO—mouthrinse concentration). Samples without fluconazole were included as control. The horizontal lines represent mean values from three independent experiments with three parallels each. Mean values from each of the parallel samples are represented by colored circles; * indicates statistical difference (<span class="html-italic">p</span> &lt; 0.05) between groups; one-way ANOVA followed by the Tukey’s post hoc test.</p>
Full article ">Figure 2
<p>DNA amount of bacteria and fungi in the oral microbiota biofilms, determined by quantitative RT-PCR. The individual values from three experiments are shown as colored circles. Horizontal bars correspond to the mean values. The horizontal line at 0.00002 ng indicates the threshold recommended for reliable quantification in the RT-PCR reaction (Zymo Femto kit). One-way ANOVA test revealed no significant differences between control and fluconazole-treated samples.</p>
Full article ">Figure 3
<p>Real-time PCR results expressed by log2 relative fold changes (2<sup>−∆∆Ct</sup>) show the impact of fluconazole at 2.56 μg·mL<sup>−1</sup> (SA), 2000 μg·mL<sup>−1</sup> (MO), or without fluconazole exposure (negative control) on microbial composition.</p>
Full article ">Figure 4
<p>Real-time viability and metabolic activity assay showing <span class="html-italic">S. mitis</span> and <span class="html-italic">S. mutans</span> growth evaluated as relative light unit (RLU) by the expression of <span class="html-italic"><sub>p</sub>ldh-luc</span> reporter strains (<b>A</b>,<b>C</b>) and growth by optical density (OD; <b>B</b>,<b>D</b>). Bacterial cells were grown in the presence of fluconazole at 2.56 μg·mL<sup>−1</sup> (SA), 4.39 μg·mL<sup>−1</sup> (1×—peak plasma concentration), 8.78 μg·mL<sup>−1</sup> (2×—twice peak plasma concentration), and 2000 μg·mL<sup>−1</sup> (MO). Samples without exposure to fluconazole were included as negative control.</p>
Full article ">Figure 5
<p>Planktonic growth results assessed by optical density (OD) showed that fluconazole at 2.56 μg·mL<sup>−1</sup> (SA), 4.39 μg·mL<sup>−1</sup> (1×), 8.78 μg·mL<sup>−1</sup> (2×), and 2000 μg·mL<sup>−1</sup> (MO) reduced planktonic growth of <span class="html-italic">E. coli</span> (<b>A</b>), retarded <span class="html-italic">G. adiacens</span> growth (<b>B</b>), and did not influence <span class="html-italic">Lactobacillus salivarius</span> and <span class="html-italic">Lactobacillus crispatus</span> viability (<b>C</b>,<b>D</b>).</p>
Full article ">Figure 6
<p>CFU (<b>A</b>), dry biomass (<b>B</b>) and pH (<b>C</b>)of <span class="html-italic">S. mitis</span> biofilms grown for 6, 24, and 48 h in the presence of fluconazole at 2.56 μg·mL<sup>−1</sup> (SA), 4.39 μg·mL<sup>−1</sup> (1×), 8.78 μg.mL<sup>−1</sup> (2×), and 2000 μg·mL<sup>−1</sup> (MO), or without fluconazole exposure (negative control). Statistical difference * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 compared to the control group using One-Way ANOVA followed by Tukey’s multi-comparison post hoc test.</p>
Full article ">
15 pages, 2125 KiB  
Article
The Mechanism of Antimicrobial Activity of Conjugated Bile Acids against Lactic Acid Bacilli
by Li-Na Chai, Hua Wu, Xue-Jiao Wang, Li-Juan He and Chun-Feng Guo
Microorganisms 2023, 11(7), 1823; https://doi.org/10.3390/microorganisms11071823 - 17 Jul 2023
Cited by 5 | Viewed by 2188
Abstract
The mechanism underlying antimicrobial activity of conjugated bile acids against strains of lactic acid bacilli is not well understood. The purpose of this study was to investigate two typical conjugated bile acids (glycochenodeoxycholic acid and taurochenodeoxycholic acid) for their mechanisms of antimicrobial activity [...] Read more.
The mechanism underlying antimicrobial activity of conjugated bile acids against strains of lactic acid bacilli is not well understood. The purpose of this study was to investigate two typical conjugated bile acids (glycochenodeoxycholic acid and taurochenodeoxycholic acid) for their mechanisms of antimicrobial activity against four strains of different species of lactic acid bacilli at the physiological pH of the small intestine of humans. The bacterial cell membrane integrity, transmembrane potential, and transmembrane pH gradient were examined using the fluorescence probes SYTO 9 plus propidium iodide, 3,3′-dipropylthiadicarbocyanine iodide, and 5(6)-carboxyfluorescein diacetate N-succinimidyl ester, respectively. The intracellular ATP levels were measured by the firefly luciferase-based bioluminescence method. It was found that the antimicrobial activity of conjugated bile acids against the strains of lactic acid bacilli is strain-specific, and glycochenodeoxycholic acid showed significantly greater antimicrobial activity than taurochenodeoxycholic acid against the strains of lactic acid bacilli. The conjugated bile acids inhibited the growth of strains of lactic acid bacilli by disrupting membrane integrity, dissipating transmembrane potential, reducing the transmembrane pH gradient, and depleting intracellular ATP. In conclusion, the antimicrobial activity of conjugated bile acids against lactic acid bacilli is a multifactorial phenomenon. This study will provide valuable information for developing strategies to improve the ability of lactic acid bacilli to tolerate bile in vivo. Full article
(This article belongs to the Section Antimicrobial Agents and Resistance)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Antimicrobial activity of GCDCA and TCDCA against <span class="html-italic">Lactiplantibacillus plantarum</span> ATCC 14917 (<b>a</b>), <span class="html-italic">Lacticaseibacillus casei</span> ATCC 334 (<b>b</b>), <span class="html-italic">Lacticaseibacillus rhamnosus</span> ATCC 53103 (<b>c</b>), and <span class="html-italic">Lactobacillus acidophilus</span> ATCC 700396 (<b>d</b>). Each bile acid was used at a concentration of 5 mM. Data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Means not sharing a common letter differ significantly (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effects of different concentrations of GCDCA and TCDCA on the cellular membrane integrity of <span class="html-italic">Lactiplantibacillus plantarum</span> ATCC 14917 (<b>a</b>), <span class="html-italic">Lacticaseibacillus casei</span> ATCC 334 (<b>b</b>), <span class="html-italic">Lacticaseibacillus rhamnosus</span> ATCC 53103 (<b>c</b>), and <span class="html-italic">Lactobacillus acidophilus</span> ATCC 700396 (<b>d</b>). Data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Means not sharing a common letter differ significantly (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effects of different concentrations of GCDCA and TCDCA on the ΔΨ of <span class="html-italic">Lactiplantibacillus plantarum</span> ATCC 14917 (<b>a</b>), <span class="html-italic">Lacticaseibacillus casei</span> ATCC 334 (<b>b</b>), <span class="html-italic">Lacticaseibacillus rhamnosus</span> ATCC 53103 (<b>c</b>), and <span class="html-italic">Lactobacillus acidophilus</span> ATCC 700396 (<b>d</b>). Data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Means not sharing a common letter differ significantly (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effects of different concentrations of GCDCA and TCDCA on the cellular ΔpH of <span class="html-italic">Lactiplantibacillus plantarum</span> ATCC 14917 (<b>a</b>), <span class="html-italic">Lacticaseibacillus casei</span> ATCC 334 (<b>b</b>), <span class="html-italic">Lacticaseibacillus rhamnosus</span> ATCC 53103 (<b>c</b>), and <span class="html-italic">Lactobacillus acidophilus</span> ATCC 700396 (<b>d</b>). ΔpH = intracellular pH − extracellular pH. Data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Means not sharing a common letter differ significantly (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effects of different concentrations of GCDCA and TCDCA on the intracellular ATP levels of <span class="html-italic">Lactiplantibacillus plantarum</span> ATCC 14917 (<b>a</b>), <span class="html-italic">Lacticaseibacillus casei</span> ATCC 334 (<b>b</b>), <span class="html-italic">Lacticaseibacillus rhamnosus</span> ATCC 53103 (<b>c</b>), and <span class="html-italic">Lactobacillus acidophilus</span> ATCC 700396 (<b>d</b>). Data are expressed as the mean ± SD (<span class="html-italic">n</span> = 3). Means not sharing a common letter differ significantly (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Correlation heatmap analysis between the variables. The correlation was assessed using Pearson correlation coefficients (r). Red represents a positive correlation, and blue represents a negative correlation. In subfigures (<b>a</b>) and (<b>b</b>), GCDCA and TCDCA were used at 5 mM, respectively.</p>
Full article ">
21 pages, 5897 KiB  
Article
The Role of Cosolvent–Water Interactions in Effects of the Media on Functionality of Enzymes: A Case Study of Photobacterium leiognathi Luciferase
by Albert E. Lisitsa, Lev A. Sukovatyi, Anna A. Deeva, Dmitry V. Gulnov, Elena N. Esimbekova, Valentina A. Kratasyuk and Elena V. Nemtseva
Life 2023, 13(6), 1384; https://doi.org/10.3390/life13061384 - 13 Jun 2023
Cited by 2 | Viewed by 1477
Abstract
A complex heterogeneous intracellular environment seems to affect enzymatic catalysis by changing the mobility of biomolecules, their stability, and their conformational states, as well as by facilitating or hindering continuously occurring interactions. The evaluation and description of the influence of the cytoplasmic matrix [...] Read more.
A complex heterogeneous intracellular environment seems to affect enzymatic catalysis by changing the mobility of biomolecules, their stability, and their conformational states, as well as by facilitating or hindering continuously occurring interactions. The evaluation and description of the influence of the cytoplasmic matrix components on enzymatic activity are problems that remain unsolved. In this work, we aimed to determine the mechanisms of action of two-component media with cosolvents of various molecular sizes on the complex multi-stage bioluminescent reaction catalyzed by bacterial luciferase. Kinetic and structural effects of ethylene glycol, glycerol, sorbitol, glucose, sucrose, dextran, and polyethylene glycol on bacterial luciferase were studied using stopped-flow and fluorescence spectroscopy techniques and molecular dynamics simulations. We have found that diffusion limitations in the presence of cosolvents promote the stabilization of flavin substrate and peroxyflavin intermediate of the reaction, but do not provide any advantages in bioluminescence quantum yield, because substrate binding is slowed down as well. The catalytic constant of bacterial luciferase has been found to be viscosity-independent and correlated with parameters of water–cosolvent interactions (Norrish constant, van der Waals interaction energies). Crowding agents, in contrast to low-molecular-weight cosolvents, had little effect on peroxyflavin intermediate decay and enzyme catalytic constant. We attributed specific kinetic effects to the preferential interaction of the cosolvents with enzyme surface and their penetration into the active site. Full article
(This article belongs to the Special Issue Recent Advances in Bioluminescence)
Show Figures

Figure 1

Figure 1
<p>The stages of the reaction catalyzed by bacterial luciferase. L—luciferase, F—flavin mononucleotide, O<sub>2</sub>—molecular oxygen, RCOH and RCOOH—long-chain aldehyde and corresponding carboxylic acid, H<sub>2</sub>O<sub>2</sub>—hydrogen peroxide.</p>
Full article ">Figure 2
<p>Chemical structures of the cosolvents used in the study: (<b>a</b>) ethylene glycol (62 Da), (<b>b</b>) glycerol (92 Da), (<b>c</b>) sorbitol (182 Da), (<b>d</b>) glucose (180 Da), (<b>e</b>) sucrose (342 Da), (<b>f</b>) dextran (70 kDa), (<b>g</b>) polyethylene glycol (4 kDa).</p>
Full article ">Figure 3
<p>Kinetics of the bioluminescent reaction catalyzed by <span class="html-italic">P. leiognathi</span> luciferase in viscous solutions with different cosolvents: ethylene glycol 30 wt % (2.1 cP), glucose 20 wt % (1.95 cP), sorbitol 20 wt % (1.95 cP), Dextran-70k 10 wt % (2.48 cP), and PEG-4k 5 wt % (2.9 cP). The kinetics in buffer without additives (1 cP) is shown by solid black line for comparison. The concentrations were: luciferase—1 μM, FMNH<sub>2</sub>—15 μM, decanal—50 μM. Vertical dotted lines denote the approximate time range for calculating the index of the exponential decay of intensity (k<sub>decay</sub>).</p>
Full article ">Figure 4
<p>The dependence of the kinetic parameters of the reaction catalyzed by <span class="html-italic">P. leiognathi</span> luciferase on solution viscosity in the presence of different cosolvents: (<b>a</b>) the peak intensity (I<sub>max</sub>), (<b>b</b>) the decay constant (k<sub>decay</sub>), (<b>c</b>) the total quantum yield (Q*), (<b>d</b>) the initial velocity (v<sub>0</sub>). Grey markers show the previously obtained data for glycerol and sucrose solutions [<a href="#B19-life-13-01384" class="html-bibr">19</a>]. Aldehyde concentration was 50 μM. Dashed lines are to guide the eyes. In (<b>d</b>), solid lines show the approximations of the experimental data with equation v<sub>0</sub> = A·η<sup>−δ</sup> for sorbitol and glucose solutions, where A is an amplitude coefficient and η is viscosity. The corresponding index δ is indicated.</p>
Full article ">Figure 5
<p>The dependence of the rate constants of FMNH<sub>2</sub> autoxidation k<sub>d</sub> (<b>a</b>), and the dark decay of C(4a)-hydroperoxyflavin intermediate, k<sub>dd</sub> (<b>b</b>), on viscosity in solutions with different cosolvents. Gray markers show previously obtained data for glycerol and sucrose solutions [<a href="#B17-life-13-01384" class="html-bibr">17</a>]. Solid lines refer to the approximation of the experimental data with equation k<sub>d,dd</sub> = A·η<sup>−δ</sup>, where A is an amplitude coefficient and η is viscosity. The corresponding indexes δ are indicated. Dashed lines are to guide the eyes.</p>
Full article ">Figure 6
<p>The dependence of the rate constants on media viscosity in the presence of different cosolvents: (<b>a</b>) binding constants of the reduced flavin k<sub>1</sub>; (<b>b</b>) binding constants of oxygen k<sub>2</sub>; (<b>c</b>) the association constant of aldehyde binding, K<sub>a</sub> = k<sub>3</sub>/k<sub>–3</sub>; (<b>d</b>) catalytic constant k<sub>4</sub>. Gray markers show the previously obtained data for glycerol and sucrose solutions [<a href="#B19-life-13-01384" class="html-bibr">19</a>]. In (<b>a</b>,<b>c</b>,<b>d</b>), the normalized values are shown (divided by the value in the buffer with η = 1 cP). Dashed lines are to guide the eyes. Solid lines refer to the approximation of the experimental data with equation k = A·η<sup>−δ</sup>, where A is an amplitude coefficient and η is viscosity. The corresponding indexes δ are indicated.</p>
Full article ">Figure 7
<p>The structural parameters of <span class="html-italic">P. leiognathi</span> luciferase obtained using the models with water and the mixtures with ethylene glycol (30 wt %), glycerol (40 wt %), sorbitol (40 wt %), glucose (40 wt %), and sucrose (40 wt %); (<b>a</b>,<b>d</b>) the root-mean-square-deviation of C<sub>α</sub>-atoms (RMSD); (<b>b</b>,<b>e</b>) the gyration radius for all protein atoms (R<sub>g</sub>); (<b>c</b>,<b>f</b>) the total solvent accessible surface area (SASA). Panels (<b>a</b>–<b>c</b>) show parameter changes during 100 ns simulations (curves are the average for the three simulation runs (<a href="#app1-life-13-01384" class="html-app">Figure S1</a>)). Panels (<b>d</b>–<b>f</b>) show the values of the parameters for the last 20 ns of the trajectories as mean ± standard deviation. Dashed lines refer to the value in water. Dotted line indicates the range over which the parameters for (<b>d</b>–<b>f</b>) were calculated (80–100 ns).</p>
Full article ">Figure 8
<p>The difference in the root-mean-square fluctuation (RMSF) of C<sub>α</sub>-atoms of <span class="html-italic">P. leiognathi</span> luciferase (catalytic α-subunit) in the presence of cosolvents compared to water (ΔRMSF = RMSF<sub>cos</sub> − RMSF<sub>water</sub>). The area of the functionally important mobile loop is indicated by gray arrows. The crosses refer to the a.r. positions, which are known to be important for luciferase functionality [<a href="#B28-life-13-01384" class="html-bibr">28</a>]. Secondary structure map shows helixes (pink), strands (yellow), and coils/turns (gray) within the enzyme structure after molecular dynamics simulations in water. The mixtures of water with ethylene glycol (30 wt %), glycerol (40 wt %), sorbitol (40 wt %), glucose (40 wt %), and sucrose (40 wt %) were studied.</p>
Full article ">Figure 9
<p>Preferential interaction coefficients for <span class="html-italic">P. leiognathi</span> luciferase with cosolvents (<b>a</b>) and water (<b>b</b>) in the models with the mixtures of water with ethylene glycol (30 wt %), glycerol (40 wt %), sorbitol (40 wt %), glucose (40 wt %), and sucrose (40 wt %). The mean ± standard deviation of three independent runs is shown. The dashed lines refer to the zero level of the coefficients.</p>
Full article ">Figure 10
<p>Two-dimensional representation of the minimum-distance distribution function (MDDF) of cosolvent molecules relative to amino acid residues of the <span class="html-italic">P. leiognathi</span> luciferase active site indicated at the bottom: (<b>a</b>) ethylene glycol 30 wt %; (<b>b</b>) glycerol 40 wt%; (<b>c</b>) sorbitol 40 wt %; (<b>d</b>) glucose 40 wt %; (<b>e</b>) sucrose 40 wt %. The dashed frames refer to the binding sites of flavin (yellow) and aldehyde (blue). Gray arrows indicate the elements of the catalytic gorge architecture. The intensity of the green color indicates the probability of cosolvent appearance at the distance r from the residue. &lt;…&gt; means a break in the sequence.</p>
Full article ">Figure 11
<p>Nonbonded interaction energies in model systems with different cosolvents: (<b>a</b>) of <span class="html-italic">P. leiognathi</span> luciferase: intramolecular (Prot-Prot), with water (Prot-Water), and cosolvent (Prot-Cos); (<b>b</b>) of cosolvent molecules: with protein (Cos-Prot), with water (Cos-Water), and cosolvent (Cos-Cos). Coulomb (filled bars, to the left ordinate) and van der Waals (dashed bars, to the right ordinate) interaction energies are shown. For the models with the mixtures of water with ethylene glycol (30 wt %), glycerol (40 wt %), sorbitol (40 wt %), glucose (40 wt %), and sucrose (40 wt %), the mean ± standard deviation of three independent runs are shown.</p>
Full article ">Figure 12
<p>Fluorescence spectra of <span class="html-italic">P. leiognathi</span> luciferase under excitation at 295 nm in buffer and in the presence of ethylene glycol (40 wt %), glycerol (40 wt %), glucose (40 wt %), sorbitol (40 wt %), sucrose (40 wt %), dextran (15 wt %), and polyethylene glycol (15 wt %). The insert demonstrates the change in the gravity center of the spectrum. Dashed horizontal line in the insert refers to the position of the gravity center in buffer (344.4 nm). Luciferase concentration in the samples was 1.5 µM. The experimental error of the gravity center value was &lt;0.4 nm.</p>
Full article ">Figure 13
<p>(<b>a</b>) Correlation plot between catalytic constant k<sub>4</sub> of bacterial luciferase in solutions with water activity about 0.98 (ethylene glycol 10 wt %, glycerol 10 wt %, sorbitol 20 wt %, glucose 20 wt %, and sucrose 30 wt %) and Norrish constant k<sub>N</sub> of the cosolvents used. The Norrish constants were calculated as mean ± standard deviation using the literature data [<a href="#B41-life-13-01384" class="html-bibr">41</a>]; (<b>b</b>) correlation plot between catalytic constant k<sub>4</sub> of bacterial luciferase in solutions with maximum cosolvent contents (ethylene glycol 30 wt %, glycerol 40 wt %, sorbitol 40 wt %, glucose 40 wt %, sucrose 40 wt %) and van der Waals interaction energies of the cosolvent with water (Cos-Wat, green markers) and itself (Cos-Cos, red markers), as well as the sum of these energies (Cos-(Wat&amp;Cos), blue markers). The energies are calculated per one cosolvent molecule as mean ± standard deviation of three independent runs. The coefficients of linear correlations r are indicated. Dashed lines are linear approximations of the data. Cosolvent markers: ethylene glycol—cross, glycerol—square, sorbitol—circle, glucose—triangle, and sucrose—diamond.</p>
Full article ">
12 pages, 2984 KiB  
Article
In Vivo Inflammation Caused by Achromobacter spp. Cystic Fibrosis Clinical Isolates Exhibiting Different Pathogenic Characteristics
by Angela Sandri, Giulia Maria Saitta, Laura Veschetti, Federico Boschi, Rebeca Passarelli Mantovani, Maria Carelli, Paola Melotti, Caterina Signoretto, Marzia Boaretti, Giovanni Malerba and Maria M. Lleò
Int. J. Mol. Sci. 2023, 24(8), 7432; https://doi.org/10.3390/ijms24087432 - 18 Apr 2023
Cited by 3 | Viewed by 1622
Abstract
Achromobacter spp. lung infection in cystic fibrosis has been associated with inflammation, increased frequency of exacerbations, and decline of respiratory function. We aimed to evaluate in vivo the inflammatory effects of clinical isolates exhibiting different pathogenic characteristics. Eight clinical isolates were selected based [...] Read more.
Achromobacter spp. lung infection in cystic fibrosis has been associated with inflammation, increased frequency of exacerbations, and decline of respiratory function. We aimed to evaluate in vivo the inflammatory effects of clinical isolates exhibiting different pathogenic characteristics. Eight clinical isolates were selected based on different pathogenic characteristics previously assessed: virulence in Galleria mellonella larvae, cytotoxicity in human bronchial epithelial cells, and biofilm formation. Acute lung infection was established by intratracheal instillation with 10.5 × 108 bacterial cells in wild-type and CFTR-knockout (KO) mice expressing a luciferase gene under control of interleukin-8 promoter. Lung inflammation was monitored by in vivo bioluminescence imaging up to 48 h after infection, and mortality was recorded up to 96 h. Lung bacterial load was evaluated by CFU count. Virulent isolates caused higher lung inflammation and mice mortality, especially in KO animals. Isolates both virulent and cytotoxic showed higher persistence in mice lungs, while biofilm formation was not associated with lung inflammation, mice mortality, or bacterial persistence. A positive correlation between virulence and lung inflammation was observed. These results indicate that Achromobacter spp. pathogenic characteristics such as virulence and cytotoxicity may be associated with clinically relevant effects and highlight the importance of elucidating their mechanisms. Full article
(This article belongs to the Collection Microbial Virulence Factors)
Show Figures

Figure 1

Figure 1
<p>IL-8-dependent bioluminescence emitted by WT and KO mice at 24 h after intratracheal challenge with clinical isolates showing (<b>a</b>) high–medium vs. high-medium virulence in <span class="html-italic">G. mellonella</span> larvae, (<b>b</b>) high–medium vs. low–no cytotoxicity in human bronchial epithelial cells, and (<b>c</b>) high–medium vs. low–no biofilm formation. Photon emission is expressed as Folds of Induction (FOI) vs. baseline (before lung challenge). Each value represents the mean of four animals challenged with each isolate. The median ± interquartile range of 16 mice per group is shown. Statistical analysis was performed by Kruskal–Wallis test; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>IL-8-dependent bioluminescence emitted by WT and KO mice after intratracheal challenge with each clinical isolate: 2-4 (<b>a</b>), 7-2 (<b>b</b>), 8-2 (<b>c</b>), 12-2 (<b>d</b>), 16-1 (<b>e</b>), 17-1 (<b>f</b>), 20-1 (<b>g</b>), 23-1 (<b>h</b>). Imaging was performed before lung challenge (0 h) and after 4, 24, and 48 h. FOI = Folds of Induction vs. baseline (0 h). Each value represents the mean ± SEM of four animals. Statistical analysis of treated vs. control mice was performed by repeated measures 2-way ANOVA followed by Tukey’s multiple comparison test; *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Percent survival of WT and KO mice up to 96 h after intratracheal challenge with clinical isolates showing (<b>a</b>,<b>b</b>) high–medium vs. low–no virulence in <span class="html-italic">G. mellonella</span> larvae, (<b>c</b>,<b>d</b>) high–medium vs. low–no cytotoxicity in human bronchial epithelial cells, and (<b>e</b>,<b>f</b>) high–medium vs. low–no biofilm formation. The mean ± SEM of 16 mice per group (<span class="html-italic">n</span> = 4 treated with each isolate) is shown. Statistical analysis was performed by log-rank test; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Percent survival of WT and KO mice up to 96 h after intratracheal challenge with each clinical isolate: 2-4 (<b>a</b>), 7-2 (<b>b</b>), 8-2 (<b>c</b>), 12-2 (<b>d</b>), 16-1 (<b>e</b>), 17-1 (<b>f</b>), 20-1 (<b>g</b>), 23-1 (<b>h</b>). The mean of four mice per group is shown. Statistical analysis was performed by log-rank test; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>CFU recovered from WT mice lungs at 24 h after intratracheal challenge with clinical isolates showing (<b>a</b>) high–medium vs. low–no virulence in <span class="html-italic">G. mellonella</span> larvae, (<b>b</b>) high–medium vs. low–no cytotoxicity in human bronchial epithelial cells, (<b>c</b>) high–medium vs. low–no biofilm formation. Each value represents the mean of three animals challenged with each isolate. The median ± interquartile range of 12 mice per group is shown.</p>
Full article ">Figure 6
<p>CFU recovered from WT mice lungs at 24 h after intratracheal challenge with eight clinical isolates. The bacterial load used for the challenge is indicated (dotted line). The mean ± SD of three mice per group is shown. Statistical analysis of each isolate vs. challenge was performed by 1way ANOVA test; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Correlation between IL-8-dependent bioluminescence emission in WT and KO mice at 24 h after intratracheal challenge with the clinical isolates and virulence in <span class="html-italic">G. mellonella</span> larvae infected with the same isolates. FOI = Fold of Induction vs. baseline (before lung challenge). Each point represents the mean ± SEM of four mice challenged with each isolate. Linear regression ± SEM (dotted lines) is shown.</p>
Full article ">
Back to TopTop