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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,181)

Search Parameters:
Keywords = heterologous

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 1906 KiB  
Article
Probiotic Spores of Shouchella clausii SF174 and Displayed Bromelain Show Beneficial Additive Potential
by Rowena Corona, Valeria Bontà, Loredana Baccigalupi and Ezio Ricca
Int. J. Mol. Sci. 2025, 26(3), 942; https://doi.org/10.3390/ijms26030942 - 23 Jan 2025
Viewed by 52
Abstract
Probiotics have health-beneficial properties mainly due to either a direct action on the host or the modulation of the host microbiota. Health-beneficial properties have also been associated with a variety of plant-derived molecules, widely used as dietary supplements. This study explores the possibility [...] Read more.
Probiotics have health-beneficial properties mainly due to either a direct action on the host or the modulation of the host microbiota. Health-beneficial properties have also been associated with a variety of plant-derived molecules, widely used as dietary supplements. This study explores the possibility of combining the actions of probiotics and of plant-derived molecules by developing beneficial, probiotic-carrying, heterologous molecules. To this extent, spores of SF174, a well-characterized probiotic strain of Shouchella clausii (formerly Bacillus clausii), were used to bind bromelain, a plant-derived mixture of endopeptidases with beneficial effects. Probiotic spores displaying bromelain maintained their antioxidant activity and acquired the endopeptidase activity of the heterologous molecule. The endopeptidase activity was stabilized by the interaction with the spore and largely preserved from degradation at simulated gastric conditions. Under conditions mimicking those encountered in the intestine, as well as upon spore germination, active bromelain was released from the spore surface. The in vitro results reported in this study support the idea that probiotics carrying beneficial heterologous molecules combine the health properties of the probiotic with those of the delivered molecule and pave the way for the development of a novel class of functional probiotics. Full article
Show Figures

Figure 1

Figure 1
<p>Bromelain endopeptidase activity before fractionation (total units, grey bars) and after fractionation by centrifugation, with pellet (adsorbed units, orange bars) and supernatant (Free units, blue bars). (<b>A</b>) Adsorption reaction performed in citrate buffer at pH 4.5. (<b>B</b>) Adsorption reaction performed in phosphate or acetate buffer at pH 4.5. (<b>C</b>) Adsorption reaction performed in acetate buffer at various pH values. The data represent the mean of three independent experiments, and the error bars are the standard error of the mean. <span class="html-italic">p</span>-values were calculated by using two-tailed <span class="html-italic">T</span>-test with ns indicating <span class="html-italic">p</span> &gt; 0.05, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001, respectively.</p>
Full article ">Figure 1 Cont.
<p>Bromelain endopeptidase activity before fractionation (total units, grey bars) and after fractionation by centrifugation, with pellet (adsorbed units, orange bars) and supernatant (Free units, blue bars). (<b>A</b>) Adsorption reaction performed in citrate buffer at pH 4.5. (<b>B</b>) Adsorption reaction performed in phosphate or acetate buffer at pH 4.5. (<b>C</b>) Adsorption reaction performed in acetate buffer at various pH values. The data represent the mean of three independent experiments, and the error bars are the standard error of the mean. <span class="html-italic">p</span>-values were calculated by using two-tailed <span class="html-italic">T</span>-test with ns indicating <span class="html-italic">p</span> &gt; 0.05, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001, respectively.</p>
Full article ">Figure 2
<p>Endopeptidase activity of free (blue line) and spore-adsorbed (orange line) bromelain after various times of exposure at SGF at pH 2.5. The data are reported as a percentage of the bromelain activity, considering as 100% the activity before incubation at SGF conditions. The data represent the mean of three independent experiments, and the error bars are the standard error of the mean. <span class="html-italic">p</span>-values were calculated by using a two-tailed <span class="html-italic">T</span>-test with *** and **** indicating <span class="html-italic">p</span> &lt; 0.001 and <span class="html-italic">p</span> &lt; 0.0001, respectively.</p>
Full article ">Figure 3
<p>Endopeptidase activity of bromelain before fractionation (total units, grey bar) and after SIF treatments and fractionation leading to the pellet (adsorbed enzyme, orange bars) and the supernatant (free enzyme, blue bars). The data represent the mean of three independent experiments, and the error bars are the standard error of the mean. <span class="html-italic">p</span>-values were calculated by using a two-tailed <span class="html-italic">T</span>-test with ** and *** indicating <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.001, respectively.</p>
Full article ">Figure 4
<p>(<b>A</b>) Efficiency of germination monitored by OD loss of <span class="html-italic">B. subtilis</span> SF106 (orange line) and <span class="html-italic">S. clausii</span> (blue line) spores (1 × 10<sup>7</sup>). (<b>B</b>) Efficiency of germination monitored by OD loss of <span class="html-italic">S. clausii</span> spores (1 × 10<sup>9</sup> spores) adsorbed (orange line) or not (blue line) with bromelain. The arrows indicate the time of collection of samples for panel C. (<b>C</b>) Endopeptidase activity of bromelain at the induction of germination (T0) and at the indicated times after the induction of germination. Samples of each time point were fractionated, and fractions were independently assayed: pellet (adsorbed enzyme, orange bars) and supernatant (free enzyme, blue bars). The data represent the mean of three independent experiments, and the error bars are the standard error of the mean. <span class="html-italic">p</span>-values were calculated by using a two-tailed <span class="html-italic">T</span>-test with ns indicating <span class="html-italic">p</span> &gt; 0.05, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and **** <span class="html-italic">p</span> &lt; 0.0001, respectively.</p>
Full article ">Figure 5
<p>Antioxidant activity of spores (dark green bars), bromelain (dark blue bars) and spore adsorbed bromelain (dark orange bars). Light bars report samples analyzed after SGF treatment. (<b>A</b>) Hydrogen peroxide and (<b>B</b>) free radicals scavenging activities. The data represent the mean of three independent experiments, and the error bars are the standard error of the mean. <span class="html-italic">p</span>-values were calculated by using a two-tailed <span class="html-italic">T</span>-test with ns indicating <span class="html-italic">p</span> &gt; 0.05, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001, respectively.</p>
Full article ">Figure 6
<p>Endopeptidase activity of food-grade bromelain before fractionation (grey bars) or after fractionation by centrifugation, with pellet (adsorbed enzyme, orange bars) and supernatant (free enzyme, blue bars). <span class="html-italic">p</span>-values were calculated by using a two-tailed <span class="html-italic">T</span>-test with ns indicating <span class="html-italic">p</span> &gt; 0.05 and **** <span class="html-italic">p</span> &lt; 0.0001, respectively.</p>
Full article ">
22 pages, 5625 KiB  
Article
Genetic and Physiological Characterization of the Pentose Phosphate Pathway in the Yeast Kluyveromyces lactis
by Laura-Katharina Bertels, Stefan Walter and Jürgen J. Heinisch
Int. J. Mol. Sci. 2025, 26(3), 938; https://doi.org/10.3390/ijms26030938 - 23 Jan 2025
Viewed by 70
Abstract
The pentose phosphate pathway (PPP) is essential for human health and provides, amongst others, the reduction power to cope with oxidative stress. In contrast to the model baker’s yeast, the PPP also contributes to a large extent to glucose metabolism in the milk [...] Read more.
The pentose phosphate pathway (PPP) is essential for human health and provides, amongst others, the reduction power to cope with oxidative stress. In contrast to the model baker’s yeast, the PPP also contributes to a large extent to glucose metabolism in the milk yeast Kluyveromyces lactis. Yet, the physiological consequences of mutations in genes encoding PPP enzymes in K. lactis have been addressed for only a few. We here embarked on a systematic study of such mutants, deleting ZWF1, SOL4, GND1, RKI1, RPE1, TKL1, TAL1, and SHB17. Interestingly, GND1, RKI1, and TKL1 were found to be essential under standard growth conditions. Epistasis analyses revealed that a lack of Zwf1 rescued the lethality of the gnd1 deletion, indicating that it is caused by the accumulation of 6-phosphogluconate. Moreover, the slow growth of a tal1 null mutant, which lacks fructose-1,6-bisphosphate aldolase, was aggravated by deleting the SHB17 gene encoding sedoheptulose-1,7-bisphosphatase. A mitotically stable tetOFF system was established for conditional expression of TAL1 and TKL1, encoding transaldolase and transketolase in the non-oxidative part of the PPP, and employed in a global proteome analysis upon depletion of the enzymes. Results indicate that fatty acid degradation is upregulated, providing an alternative energy source. In addition, tal1 and tkl1 null mutants were complemented by heterologous expression of the respective genes from baker’s yeast and humans. These data demonstrate the importance of the PPP for basic sugar metabolism and oxidative stress response in K. lactis and the potential of this yeast as a model for the study of PPP enzymes from heterologous sources, including human patients. Full article
(This article belongs to the Special Issue Yeasts: Model Systems for Molecular Research)
Show Figures

Figure 1

Figure 1
<p>Routes of glucose catabolism in <span class="html-italic">Kluyveromyces lactis</span>. Simplified overview of the enzymes involved in the initial steps of glycolysis (green letters) and those of the pentose phosphate pathway (PPP) investigated in this work (bold blue letters). The oxidative part of the PPP is shaded in pale blue, and the non-oxidative part in pale yellow. Arrows between metabolites indicate their interconversion, with double arrows designating reversible, and pointed arrows the irreversible reactions. Abbreviations of enzymes follow the gene nomenclature and are as follows: Hxk1 = hexokinase, Pgi1 = phosphoglucose isomerase, PFK = phosphofructokinase, a heterooctameric complex composed of four alpha- and four beta-subunits [<a href="#B5-ijms-26-00938" class="html-bibr">5</a>], Fba1 = fructose-1,6-bisphosphate aldolase, Tpi1 = triosephosphate isomerase, Zwf1 = glucose-6-phosphate dehydrogenase (“Zwischenferment”), Sol4 = 6-phosphogluconolactonase, Gnd1 = 6-phosphogluconate dehydrogenase, Rki1 = ribosephosphate ketol isomerase, Rpe1 = ribulosephosphate epimerase, Tal1 = transaldolase, Tkl1 = transketolase, and Shb17 = sedoheptulose-1,7-bisphosphatase. Note that for reasons of clarity, the non-glycolytic reaction of Fba1 is depicted in the lower right corner. TCA designates the tricarboxylic acid cycle. Metabolite abbreviations are as follows: Glc = glucose, G6P = glucose-6-phosphate, F6P = fructose-6-phosphate, F1,6P<sub>2</sub> = fructose-1,6-bisphosphate, DHAP = dihydroxyacetone-3-phosphate, GAP = glyceraldehyde-3-phosphate, Pyr = pyruvate, EtOH = ethanol, 6PGL = 6-phosphogluconolactone, 6PGA = 6-phosphogluconate, Ru5P = ribulose-5-phosphate, Ri5P = ribose-5-phosphate, Xu5P = xylulose-5-phosphate, E4P = erythrose-4-phosphate, S7P = sedoheptulose-7-phosphate, and S17P<sub>2</sub> = sedoheptulose-1,7-bisphosphate.</p>
Full article ">Figure 2
<p>Phenotypes associated with the lack of specific PPP enzymes. Diploid strains heterozygous for the gene deletions indicated were sporulated and subjected to tetrad analyses. Only some representative tetrads are shown from each cross, with a minimum of 18 tetrads analyzed in each case. If not stated otherwise, segregants were obtained on standard YEPD medium with 2% glucose as a carbon source and incubated for the times indicated at 28 °C. Viable segregants carrying the deletion marker are highlighted by yellow circles. Where the deletion proved to be lethal (no circles), the two colony-forming segregants invariably lacked the deletion marker, as expected. Blue circles in the analysis of <span class="html-italic">rpe1::SpHIS3</span> also being heterozygous for the <span class="html-italic">rho5::kanMX</span> deletion designate the wild-types, and yellow circles the single <span class="html-italic">rpe1</span> deletions. Note that in an independent analysis of another heterozygous <span class="html-italic">rpe1</span> deletion being wild-type for <span class="html-italic">RHO5</span>, the slower growing variants from 10 tetrads analyzed invariably carried the deletion marker, confirming the results shown here. Growth phenotypes were also confirmed in similar analyses employing a different selection marker for the deletion of all genes, with the exception of <span class="html-italic">RPE1</span>.</p>
Full article ">Figure 3
<p>Epistasis analyses of different PPP null mutant combinations (<b>a</b>,<b>b</b>). For the preparation of master plates, segregants were picked according to their colony sizes, starting with the larger colonies and followed by the smaller ones. Numbers written with the first segregants reflect the tetrad number from each cross. Master plates were generated on YEPD and incubated overnight at 28 °C, before replica-plating onto the indicated media. Images were taken after 1–2 days of incubation at 28 °C. Strains employed were KHO528 (<b>a</b>) and KHO526 (<b>b</b>). Yellow circles in (<b>a</b>) highlight <span class="html-italic">gnd1</span> deletions carrying the plasmid, and blue circles the double deletions with <span class="html-italic">zwf1</span>. (<b>c</b>) The strains designated below each plate were sporulated, subjected to tetrad analysis on YEPD plates, and incubated for the times indicated at 28 °C. Colony sizes were assessed by determination of the pixel areas using the ImageJ program 1.53c. Columns show the mean colony sizes with error bars indicating the standard deviations from the total number of colonies (n) examined for each genotype.</p>
Full article ">Figure 4
<p>Establishment of the <span class="html-italic">tetOFF</span> system for conditional expression of transketolase and transaldolase in <span class="html-italic">K. lactis</span>. (<b>a</b>) Genomic integration of the <span class="html-italic">tTA</span> expression cassette and of a selectable <span class="html-italic">tetOFF</span> promoter construct for PCR-based one-step substitutions. The <span class="html-italic">TKL1</span> promoter is depicted as an example for conditional expression of any desired target gene. (<b>b</b>) Kinetics of transketolase depletion followed by determination of specific enzyme activities (<b>left</b>) and effect on growth (<b>right</b>). (<b>c</b>) Kinetics of transaldolase depletion followed by determination of specific enzyme activities (<b>left</b>) and effect on growth (<b>right</b>) on synthetic complete medium (SCD).</p>
Full article ">Figure 5
<p>Comparative proteome analyses after depletion of transketolase (<b>a</b>) or transaldolase (<b>b</b>). Volcano plots show the decrease (green dots) and increase (red dots) in proteins upon incubation of the strains KHO382-1B (<b>a</b>) and KHO551-3B (<b>b</b>) with the relevant genotypes depicted in the gray shaded bars with doxycycline to inhibit gene expression. As indicated, cut-offs were chosen at a factor of 2 relative to the untreated control and at a significance of 20. Tables below the plots are numbered according to the plots, with downregulation on pale green and upregulation on a pale red background. Peptides identified by mass spectrometry were assigned using basically the annotations in the <span class="html-italic">Saccharomyces</span> genome database (<a href="https://www.yeastgenome.org" target="_blank">https://www.yeastgenome.org</a>; last accessed on 20 December 2024) and annotated to the <span class="html-italic">K. lactis</span> genes in the KEGG database (<a href="https://www.genome.jp" target="_blank">https://www.genome.jp</a>; last accessed on 20 December 2024).</p>
Full article ">Figure 5 Cont.
<p>Comparative proteome analyses after depletion of transketolase (<b>a</b>) or transaldolase (<b>b</b>). Volcano plots show the decrease (green dots) and increase (red dots) in proteins upon incubation of the strains KHO382-1B (<b>a</b>) and KHO551-3B (<b>b</b>) with the relevant genotypes depicted in the gray shaded bars with doxycycline to inhibit gene expression. As indicated, cut-offs were chosen at a factor of 2 relative to the untreated control and at a significance of 20. Tables below the plots are numbered according to the plots, with downregulation on pale green and upregulation on a pale red background. Peptides identified by mass spectrometry were assigned using basically the annotations in the <span class="html-italic">Saccharomyces</span> genome database (<a href="https://www.yeastgenome.org" target="_blank">https://www.yeastgenome.org</a>; last accessed on 20 December 2024) and annotated to the <span class="html-italic">K. lactis</span> genes in the KEGG database (<a href="https://www.genome.jp" target="_blank">https://www.genome.jp</a>; last accessed on 20 December 2024).</p>
Full article ">Figure 6
<p>Tetrad analyses of strains heterozygous for the <span class="html-italic">tkl1</span> or <span class="html-italic">tal1</span> deletion carrying complementing genes integrated at the <span class="html-italic">leu2</span> locus. Designations of the strains employed and times of incubation at 28 °C on standard YEPD medium are indicated below each plate. For KHO577 and KHO579, a rich medium containing only 0.3% glucose was used for tetrad analysis, additionally supplemented with ammonium sulfate and uracil. The very small colonies not circled represent <span class="html-italic">tkl1</span> deletions not carrying the integrated <span class="html-italic">ScTKL2</span> allele. Blue circles designate the colonies carrying the integrated constructs in the background of the respective deletion.</p>
Full article ">Figure 7
<p>Overview on the connection between peroxisomal fatty acid degradation and mitochondrial turnover of acetyl-CoA with the pathways for glucose degradation in <span class="html-italic">K. lactis</span> (adapted from [<a href="#B46-ijms-26-00938" class="html-bibr">46</a>,<a href="#B47-ijms-26-00938" class="html-bibr">47</a>,<a href="#B48-ijms-26-00938" class="html-bibr">48</a>]). The reactions corresponding to the enzymes upregulated upon depletion of transketolase or transaldolase are shown with regard to their intracellular distribution in peroxisomes (dark blue line), mitochondria (green line), or the cytosol (compare <a href="#ijms-26-00938-t001" class="html-table">Table 1</a> for functions and fold-changes in protein concentrations). Enzyme names in red designate those increased upon depletion of both transketolase or transaldolase. Lpx1 (dark blue) is only increased if transketolase is depleted, Fox3 (light blue) only upon transaldolase depletion. Note that Yat1 localization at the outer mitochondrial membrane is not shown here for reasons of clarity in presentation. Depicted metabolic pathways (gray shaded) include the pentose phosphate pathway (PPP with depletion of the two key enzymes indicated by the downward arrows), glycolysis, and the tricarboxylic acid cycle (TCA). Other abbreviations are as follows: TAG = triacylglycerol, FA = fatty acid, and PDH = pyruvate dehydrogenase complex.</p>
Full article ">
18 pages, 5974 KiB  
Article
Lineage 7 Porcine Reproductive and Respiratory Syndrome Vaccine Demonstrates Cross-Protection Against Lineage 1 and Lineage 3 Strains
by Hsien-Jen Chiu, Shu-Wei Chang, Hongyao Lin, Yi-Chun Chuang, Kun-Lin Kuo, Chia-Hung Lin, Ming-Tang Chiou and Chao-Nan Lin
Vaccines 2025, 13(2), 102; https://doi.org/10.3390/vaccines13020102 - 21 Jan 2025
Viewed by 328
Abstract
Background/Objectives: Porcine reproductive and respiratory syndrome virus (PRRSV) has a major impact on swine productivity. Modified-live vaccines (MLVs) are used to aid in control. We investigated the cross-protection provided by a lineage 7 PRRSV MLV against a lineage 1 isolate under laboratory [...] Read more.
Background/Objectives: Porcine reproductive and respiratory syndrome virus (PRRSV) has a major impact on swine productivity. Modified-live vaccines (MLVs) are used to aid in control. We investigated the cross-protection provided by a lineage 7 PRRSV MLV against a lineage 1 isolate under laboratory conditions and a lineage 3 challenge under field conditions in Taiwan. Methods: In the first study, thirty PRRS antibody-negative conventional piglets were vaccinated via the intramuscular (IM) or the intradermal (ID) route, with the control group receiving a placebo. Four weeks after immunization, all groups were challenged with a Taiwanese lineage 1 strain. The standard protocol for detection of reversion to virulence was applied to the vaccine strain in the second study, using sixteen specific pathogen-free piglets. In the third study, on an infected pig farm in Taiwan (lineage 3 strain), three hundred piglets were randomly selected and divided into three groups, each injected with either the PrimePac® PRRS vaccine via the IM or the ID route, or a placebo. Results: In the first study, both vaccinated groups demonstrated reduced viraemia compared to the control group. The second study demonstrated that the MLV strain was stable. In the third study, piglet mortality, average daily weight gain, and pig stunting rate were significantly improved in the vaccinated groups compared to the control group. Conclusions: PrimePac® PRRS is safe to use in the field in the face of a heterologous challenge, successfully providing cross-protection against contemporary lineage 1 and lineage 3 PRRSV strains from Taiwan. Full article
(This article belongs to the Special Issue Animal Virus Infection, Immunity and Vaccines)
Show Figures

Figure 1

Figure 1
<p>Scheme of the study design. (<b>A</b>) Laboratory efficacy trial; (<b>B</b>) laboratory reversion to virulence trial; (<b>C</b>) field efficacy trial.</p>
Full article ">Figure 2
<p>Daily average rectal temperature after vaccination per experimental group. The dotted line indicates the threshold for a clinical definition of fever (40.5 °C).</p>
Full article ">Figure 3
<p>Daily average rectal temperature after the challenge. The dotted line indicates the threshold for a clinical definition of fever (40.5 °C). Different letters indicate significant differences between each group at a given date (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Macroscopic (<b>A</b>) and microscopic (<b>B</b>) lung lesion scores. * Indicates significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Genomic PRRSV load in the serum of pigs after vaccination (day post vaccination, DPV) and after challenge (day post challenge, DPC). Different letters indicate significant differences between each group (<span class="html-italic">p</span> &lt; 0.05) on a given date. The dotted line is the detection limit of PRRSV RT-qPCR.</p>
Full article ">Figure 6
<p>Elisa PRRSV-specific antibody values in the serum of pigs after vaccination (day post vaccination, DPV) and after challenge (day post challenge, DPC). Different letters indicate significant differences between each group at a given date (<span class="html-italic">p</span> &lt; 0.05). The dotted line is the PRRSV ELISA interpretation threshold (0.4); any point below is considered negative.</p>
Full article ">Figure 7
<p>Comparison of the viral genomic load in lung tissue and hilar lymph nodes of pigs in each experimental group. Red: IM group; blue: ID group; black: control group. *** Indicates significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Comparison of the weekly number of dead pigs and cumulative mortality rate in each group during the nursery stage. The <span class="html-italic">y</span>-axis on the left is the number of dead animals, and the <span class="html-italic">y</span>-axis to the right is the cumulative mortality rate among each group.</p>
Full article ">Figure 9
<p>Comparison of the liveweight of pigs in each group at 2, 4, and 12 weeks of age. ****, *** Indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Comparison of the amount of PRRSV in the serum of pigs of different groups at different ages. The <span class="html-italic">y</span>-axis on the left is the genomic load, and the <span class="html-italic">y</span>-axis to the right is the virus positivity rate among each group. The dotted line is the limit of detection of the PRRSV RT-qPCR.</p>
Full article ">Figure 11
<p>Elisa PRRSV-specific antibody values in the serum of field pigs. The dotted line is the PRRSV ELISA interpretation threshold (0.4); any point below is considered negative.</p>
Full article ">
19 pages, 8275 KiB  
Article
Adenoviral Vector-Based Vaccine Expressing Hemagglutinin Stem Region with Autophagy-Inducing Peptide Confers Cross-Protection Against Group 1 and 2 Influenza A Viruses
by Wen-Chien Wang, Ekramy E. Sayedahmed, Marwa Alhashimi, Ahmed Elkashif, Vivek Gairola, Muralimanohara S. T. Murala, Suryaprakash Sambhara and Suresh K. Mittal
Vaccines 2025, 13(1), 95; https://doi.org/10.3390/vaccines13010095 - 20 Jan 2025
Viewed by 500
Abstract
Background/Objectives: An effective universal influenza vaccine is urgently needed to overcome the limitations of current seasonal influenza vaccines, which are ineffective against mismatched strains and unable to protect against pandemic influenza. Methods: In this study, bovine and human adenoviral vector-based vaccine platforms were [...] Read more.
Background/Objectives: An effective universal influenza vaccine is urgently needed to overcome the limitations of current seasonal influenza vaccines, which are ineffective against mismatched strains and unable to protect against pandemic influenza. Methods: In this study, bovine and human adenoviral vector-based vaccine platforms were utilized to express various combinations of antigens. These included the H5N1 hemagglutinin (HA) stem region or HA2, the extracellular domain of matrix protein 2 of influenza A virus, HA signal peptide (SP), trimerization domain, excretory peptide, and the autophagy-inducing peptide C5 (AIP-C5). The goal was to identify the optimal combination for enhanced immune responses and cross-protection. Mice were immunized using a prime-boost strategy with heterologous adenoviral (Ad) vectors. Results: The heterologous Ad vectors induced robust HA stem-specific humoral and cellular immune responses in the immunized mice. Among the tested combinations, Ad vectors expressing SP + HA stem + AIP-C5 conferred significant protection against group 1 (H1N1 and H5N1) and group 2 (H3N2) influenza A viruses. This protection was demonstrated by lower lung viral titers and reduced morbidity and mortality. Conclusions: The findings support further investigation of heterologous Ad vaccine platforms expressing SP + HA stem + AIP-C5. This combination shows promise as a potential universal influenza vaccine, providing broader protection against influenza A viruses. Full article
(This article belongs to the Special Issue Advances in Vaccines against Infectious Diseases)
Show Figures

Figure 1

Figure 1
<p>Diagrammatic representation of adenoviral (Ad) vector constructs and experimental plan of Study #1. (<b>A</b>) The expression cassettes containing the hemagglutinin domain 2 (HA2) gene of A/Vietnam/1203/2004(H5N1) conjugated with/without HA signal peptide (SP) and trimerization motif (Tri) were used for HAd5 (human Ad type 5) vectors. The resultant HAd vectors are listed. CMV, cytomegalovirus promoter; PA, bovine growth hormone (BGH) polyadenylation signal; ITR, inverted terminal repeats. (<b>B</b>) Expression of the HA2-containing protein in each Ad vector was confirmed by immunoblotting. Cell extract with empty vector infection was used as a negative control. (<b>C</b>) Eight-week-old BALB/c mice were vaccinated intranasally (i.n.) with a single dose of HAd vectors for the immunogenicity evaluation. Three weeks post-inoculation, mice were euthanized, and the blood, lungs, spleen, and mediastinal lymph node (MLN) were collected to evaluate the development of humoral and cell-mediated immune responses. To assess the protection efficacy, the mice were challenged with 100 mouse infectious dose 50% (MID<sub>50</sub>) of A/Vietnam/1203/2004(H5N1)-PR8/CDC-RG, and the lung viral titers were determined three days post-infection (dpi).</p>
Full article ">Figure 2
<p>HA-specific antibody responses in the serum and lung washes of mice following adenoviral vector vaccine inoculation in Study #1. Eight-week-old BALB/c mice (5 animals/group) were vaccinated intranasally with 3 × 10<sup>7</sup> PFU (plaque-forming units) of HA2-based HAd vectors as outlined in <a href="#vaccines-13-00095-f001" class="html-fig">Figure 1</a>C. Three weeks post-vaccination, the blood and lung washes were collected. The serum samples were used for determining HA-specific IgG, IgG<sub>1</sub>, and IgG<sub>2a</sub> titers (<b>A</b>), while lung washes were examined for HA-specific IgG, IgG<sub>1</sub>, IgG<sub>2a</sub>, and IgA titers (<b>B</b>) using ELISA. The ELISA data are expressed in the area under curve (AUC) with a cut-off value determined by the average of blank wells. Each symbol represents an individual animal, and error bars indicate SD. Data were analyzed using one-way analysis of variance (ANOVA) with Dunnett’s post hoc test. Statistical significance compared to the empty vector group is denoted as follows: *, significant at <span class="html-italic">p</span> ≤ 0.05; **, significant at <span class="html-italic">p</span> ≤ 0.01; ***, significant at <span class="html-italic">p</span> ≤ 0.001; and ****, significant at <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 3
<p>HA2-specific cellular immune responses and protective efficacy in mice vaccinated with adenoviral vector vaccines in Study #1. Eight-week-old BALB/c mice (5 animals/group) were inoculated intranasally (i.n.) with 3 × 10<sup>7</sup> PFU (plaque-forming units) of HA2-based HAd vectors as shown in <a href="#vaccines-13-00095-f001" class="html-fig">Figure 1</a>C. At 3 weeks after vaccination, splenocytes, mediastinal lymph node (MLN) cells, and lung mononuclear (MN) cells were collected and stimulated with HA2 overlapping peptide array. The number of HA2-specific cytokine-expressing T cells was quantified using ELISpot assay. The number of HA2-specific T cells expressing IFN-γ (<b>A</b>) or IL-2 (<b>B</b>) in splenocytes, MLN cells, or lung MN cells is presented. To assess the protective efficacy, animals (5 animals/group) were challenged with 100 mouse infectious dose 50% (MID<sub>50</sub>) of A/Vietnam/1203/2004(H5N1)-PR8/CDC-RG influenza virus. Lung tissues were collected three days post-infection for viral load titration (<b>C</b>). Each symbol represents an individual animal, and error bars indicate SD. Data were analyzed using one-way analysis of variance (ANOVA) with Dunnett’s post hoc test. Statistical significance compared to the empty vector group is denoted as follows: *, significant at <span class="html-italic">p</span> ≤ 0.05; **, significant at <span class="html-italic">p</span> ≤ 0.01; ***, significant at <span class="html-italic">p</span> ≤ 0.001; and ****, significant at <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 4
<p>Diagrammatic representation of adenoviral (Ad) vector constructs and experimental plan of Study #2. (<b>A</b>) The expression cassettes containing the hemagglutinin stem region (HA stem) gene of A/Vietnam/1203/2004(H5N1) conjugated with immunoglobulin excretory signal peptides (EP) or HA signal peptide (SP) and with/without repetitive extracellular domains of matrix protein 2 (4M2e) were used for BAd3 (bovine Ad type 3) or HAd5 (human Ad type 5) vectors. An expressing cassette containing whole HA region (Full HA) was also generated. The resultant BAd and HAd vectors are listed. TND, transmembrane domain; CD, cytoplasmic domain; CMV, cytomegalovirus promoter; PA, bovine growth hormone (BGH) polyadenylation signal; ITR, inverted terminal repeats. (<b>B</b>) Expression of the HA stem-containing protein in each BAd and HAd vector was confirmed by immunoblotting. Cell extracts with empty vector infection were used as the negative control. (<b>C</b>) For the immunogenicity evaluation, 8-week-old BALB/c mice were vaccinated intranasally (i.n.) with BAd vectors and boosted with HAd vectors after 4 weeks. Two weeks after the second-dose inoculation, mice were euthanized, and the blood, lungs, spleen, and mediastinal lymph node (MLN) were collected to evaluate the development of humoral and cell-mediated immune responses. To assess the protection efficacy, the mice were challenged with 100 mouse infectious dose 50% (MID<sub>50</sub>) of A/Vietnam/1203/2004(H5N1)-PR8/CDC-RG, and the lung viral titers were determined three days post-infection (dpi).</p>
Full article ">Figure 5
<p>HA-specific antibody responses in the serum and lung washes of mice after adenoviral vector vaccine inoculation in Study #2. Eight-week-old BALB/c mice (4 animals/group) were vaccinated intranasally with 10<sup>7</sup> PFU (plaque-forming units) of HA stem-based BAd vectors and boosted with 10<sup>8</sup> PFU of HA stem-based HAd vectors as outlined in <a href="#vaccines-13-00095-f004" class="html-fig">Figure 4</a>C. The blood samples and lung washes were collected 2 weeks post-second dose. The serum samples were analyzed for HA-specific (A) or M2e-specific (B) IgG, IgG<sub>1</sub>, and IgG<sub>2a</sub> titers, while lung washes were examined for HA-specific IgG, IgG<sub>1</sub>, IgG<sub>2a</sub>, and IgA titers (C) by ELISA. The ELISA data are expressed as the area under curve (AUC), with a cut-off value determined by the average of blank wells. Each symbol represents an individual animal, and error bars indicate SD. Data were analyzed using one-way analysis of variance (ANOVA) with Dunnett’s post hoc test. Statistical significance compared to the empty vector group is denoted as follows: *, significant at <span class="html-italic">p</span> ≤ 0.05; **, significant at <span class="html-italic">p</span> ≤ 0.01; ***, significant at <span class="html-italic">p</span> ≤ 0.001; and ****, significant at <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 6
<p>HA2-specific cellular immune responses and protective efficacy in mice vaccinated with adenoviral vector vaccines in Study #2. Eight-week-old BALB/c mice (4 animals/group) were vaccinated intranasally with 10<sup>7</sup> PFU (plaque-forming units) of HA stem-based BAd vectors and received a booster with 10<sup>8</sup> PFU of HA stem-based HAd vectors as shown in <a href="#vaccines-13-00095-f004" class="html-fig">Figure 4</a>C. Two weeks post-booster dose, splenocytes, mediastinal lymph node (MLN) cells, and lung mononuclear (MN) cells were collected and stimulated with an HA2 overlapping peptide array. The number of HA2-specific cytokine-expressing T cells was quantified using ELISpot assay. The number of HA2-specific T cells expressing IFN-γ (<b>A</b>) or IL-2 (<b>B</b>) in splenocytes, MLN cells, or lung MN cells are presented. To assess the protective efficacy, vaccinated animals (5 animals/group) were challenged with 100 mouse infectious dose 50% (MID<sub>50</sub>) of A/Vietnam/1203/2004(H5N1)-PR8/CDC-RG influenza virus. Lung tissues were collected three days post-infection for viral load titration (<b>C</b>). Each symbol represents an individual animal, and error bars indicate SD. Data were analyzed using one-way analysis of variance (ANOVA) with Dunnett’s post hoc test. Statistical significance compared to the empty vector group is denoted as follows: *, significant at <span class="html-italic">p</span> ≤ 0.05; **, significant at <span class="html-italic">p</span> ≤ 0.01; ***, significant at <span class="html-italic">p</span> ≤ 0.001; and ****, significant at <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 7
<p>Experimental plan of Study #3 and morbidity, mortality, and lung viral load in the immunized mice challenged with homologous and heterosubtypic influenza viruses. (<b>A</b>) Vaccine candidates were selected based on the findings from Study #1 and Study #2. Eight-week-old BALB/c mice (5 animals/group) were immunized intranasally (i.n.) with HAd vectors, followed by a booster dose with BAd vectors at 3-week intervals. Immunized mice were challenged with 5 mouse lethal dose 50% (MLD<sub>50</sub>) of A/Vietnam/1203/2004(H5N1)-PR8/CDC-RG [VN/1203/04], A/Puerto Rico/8/1934(H1N1) [PR8], or A/HongKong/1/68(H3N2) [HK68] to evaluate the protection efficacy. The morbidity (weight loss) (<b>B</b>,<b>E</b>,<b>H</b>) and mortality (survival rates) (<b>C</b>,<b>F</b>,<b>I</b>) were monitored over 14 days post-challenge. The animals with weight loss exceeding 20% were euthanized and recorded as deceased. Data are shown as the mean with the error bars representing SD. The lungs were collected three days post-challenge for virus titration (<b>D</b>,<b>G</b>,<b>J</b>), expressed as Log<sub>10</sub> TCID<sub>50</sub> (50% tissue culture infectious dose), with a detection limit of 0.5 Log<sub>10</sub> TCID<sub>50</sub>. Each symbol represents an individual animal, and error bars indicate SD. Data were analyzed using one-way analysis of variance (ANOVA) with Dunnett’s post hoc test. Statistical significance compared to the empty vector group is denoted as follows: **, significant at <span class="html-italic">p</span> ≤ 0.01; and ****, significant at <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">
12 pages, 2290 KiB  
Article
Heterologous Expression and Functional Analysis of Exiguobacterium Algin Lyase Gene by Pichia pastoris
by Hanwen Wu, Kai Hou, Yutong Jiang, Mingjian Luan, Yuxia Sun, Xi He and Xiangzhong Zhao
Fermentation 2025, 11(1), 34; https://doi.org/10.3390/fermentation11010034 - 16 Jan 2025
Viewed by 450
Abstract
Algin is the most abundant substance in alga. Alginate lyase degrades algin and produces algin monosaccharides, disaccharides, and oligosaccharides, which are widely used in bioenergy, food, medicine, and other fields. In this study, one Exiguobacterium strain isolated from rotten kelp exhibited a robust [...] Read more.
Algin is the most abundant substance in alga. Alginate lyase degrades algin and produces algin monosaccharides, disaccharides, and oligosaccharides, which are widely used in bioenergy, food, medicine, and other fields. In this study, one Exiguobacterium strain isolated from rotten kelp exhibited a robust ability to degrade the alga. The sequencing of this strain revealed the presence of three different types of algin alginate lyase. Nevertheless, the expression of three genes in Escherichia coli revealed a lower alginate lyase activity compared to that of the original strain. After codon optimization, the gene with the highest activity of the three was successfully expressed in Pichia pastoris to produce recombinant EbAlg664. The activity of the recombinant enzyme in 5 L high-density fermentation reached 1306 U/mg protein, 3.9 times that of the original Exiguobacterium strain. The results of the enzymatic analysis revealed that the optimal temperature and the pH range of recombinant EbAlg664 were narrower compared to the original strain. Additionally, the presence of Cu2+ and Co2+ enhanced the enzymatic activity, whereas Mg2+ and Fe3+ exhibited inhibitory effects on the recombinant alginate lyase. The study offers a theoretical and practical foundation for the industrial-scale production of engineered Pichia pastoris with high alginate lyase activity. Full article
(This article belongs to the Section Microbial Metabolism, Physiology & Genetics)
Show Figures

Figure 1

Figure 1
<p>Growth and alginate lyase activity of <span class="html-italic">Exiguobacterium</span> in different media. (<b>a</b>): The growth of <span class="html-italic">Exiguobacterium</span> in different media; (<b>b</b>): enzyme activity of <span class="html-italic">Exiguobacterium</span> in different media.</p>
Full article ">Figure 2
<p>Amino acid sequence of EbAlg660, EbAlg664, and EbAlg665. ‘’*” indicates a common termination sequence.</p>
Full article ">Figure 3
<p>Alginate lyase activity of different strains. <span class="html-italic">E. coli</span> and <span class="html-italic">Exiguobacterium</span> are the original untransformed strains; <span class="html-italic">E. coli</span> (660), <span class="html-italic">E. coli</span> (664), <span class="html-italic">E. coli</span> (665), and <span class="html-italic">P. pastoris</span> (664) belong to the recombinant strains containing different algin lyases gene. “*” indicates <span class="html-italic">p</span> &lt; 0.05 and “**” indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Enzyme activity at different times and SDS-PAGE of <span class="html-italic">P. pastoris</span> recombinant. (<b>a</b>): Activity of alginate lyase of recombinant <span class="html-italic">P. pastoris</span> at different times; (<b>b</b>): SDS-PAGE of <span class="html-italic">P. pastoris</span> fermentation, Lane M, low-molecular-weight standard; Lane 1, crude enzyme extraction of recombinant Alg664; Lane 2, fermentation supernatant of recombinant <span class="html-italic">Pichia pastoris</span>; Lane 3, fermentation supernatant of original <span class="html-italic">P. pastoris</span>.</p>
Full article ">Figure 5
<p>The optimum temperature and pH of original <span class="html-italic">Exiguobacterium</span> alginate lyase and recombinant <span class="html-italic">P. pastoris</span> EbAlg664. (<b>a</b>): Optimum temperature of original <span class="html-italic">Exiguobacterium</span> alginate lyase and recombinant <span class="html-italic">P. pastoris</span> EbAlg664; (<b>b</b>): optimum pH of original <span class="html-italic">Exiguobacterium</span> alginate lyase and recombinant <span class="html-italic">P. pastoris</span> EbAlg664.</p>
Full article ">Figure 6
<p>Effects of different ions on enzyme activity of recombinant EbAlg664 at different concentrations. “*” indicates <span class="html-italic">p</span> &lt; 0.05 and “**” indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
13 pages, 2108 KiB  
Article
Genomic, Evolutionary, and Pathogenic Characterization of a New Polerovirus in Traditional Chinese Medicine Viola philippica
by Yuanling Chen, Gaoxiang Chen, Jiaping Yu, Yali Zhou, Shifang Fei, Haorong Chen, Jianxiang Wu and Shuai Fu
Viruses 2025, 17(1), 114; https://doi.org/10.3390/v17010114 - 15 Jan 2025
Viewed by 357
Abstract
Viola philippica, a medicinal herbaceous plant documented in the Chinese Pharmacopoeia, is a promising candidate for research into plant-derived pharmaceuticals. However, the study of newly emerging viruses that threaten the cultivation of V. philippica remains limited. In this study, V. philippica plants [...] Read more.
Viola philippica, a medicinal herbaceous plant documented in the Chinese Pharmacopoeia, is a promising candidate for research into plant-derived pharmaceuticals. However, the study of newly emerging viruses that threaten the cultivation of V. philippica remains limited. In this study, V. philippica plants exhibiting symptoms such as leaf yellowing, mottled leaves, and vein chlorosis were collected and subjected to RNA sequencing to identify potential viral pathogens. A novel polerovirus, named Viola Philippica Polerovirus (VPPV), was identified in V. philippica. VPPV possesses a linear, positive-sense, single-stranded RNA genome consisting of 5535 nucleotides (nt) and encodes seven highly overlapping open reading frames (ORFs). Two potential recombination events were identified within ORF2, ORF3a, and ORF3, providing insights into the genetic diversity and evolution history of this novel polerovirus. An infectious cDNA clone of VPPV was successfully constructed and shown to infect Nicotiana benthamiana. Using a PVX-based heterologous expression system, the VPPV P0 protein was shown to trigger a systemic hypersensitive response (HR)-like reaction in N. benthamiana, indicating that P0 functions as the main pathogenicity determinant. These findings contributed to the detection and understanding of pathogenic mechanisms and control strategies for VPPV in V. philippica. Full article
(This article belongs to the Special Issue Emerging and Reemerging Plant Viruses in a Changing World)
Show Figures

Figure 1

Figure 1
<p>Symptoms of the diseased <span class="html-italic">Viola philippica</span> plants, RT-PCR detection of VPPV contigs and characteristics of the VPPV genome. (<b>a</b>,<b>b</b>) The symptomatic <span class="html-italic">V. philippica</span> plants show symptoms of leaf yellowing, mottling, and vein chlorosis. (<b>c</b>) Healthy <span class="html-italic">V. philippica</span> plants. (<b>d</b>) RT-PCR detection of contig 1345 and contig 21,591 in the symptomatic <span class="html-italic">V. philippica</span> plants and healthy plants (CK-). (<b>e</b>) Schematic diagram of the genome organization of VPPV. The positions of −1 ribosomal frameshift site in ORF1 and stop codon readthrough are marked in the schematic diagram.</p>
Full article ">Figure 2
<p>Recombination and phylogenetic analysis of VPPV with other related poleroviruses in the family <span class="html-italic">Solemoviridae</span> and <span class="html-italic">Tombusviridae</span>. Results of the recombination analyzed by GENECONV method (<b>a</b>) and MaxChi method (<b>b</b>). Lines indicate the percentage of similarity per alignment, and the pink area indicates the recombinant part of the sequence. The phylogenetic trees were constructed based on RdRp (<b>c</b>) or CP protein (<b>d</b>) using the maximum likelihood method with 1000 bootstraps. VPPV is shown in red. The bootstrap values are indicated adjacent to the nodes. Detailed sequence information of these viruses is shown in <a href="#app1-viruses-17-00114" class="html-app">Table S4</a>.</p>
Full article ">Figure 3
<p>Agroinfiltration inoculation of the established VPPV infectious cDNA clone and identification of pathogenicity determinants of VPPV. (<b>a</b>) RT-PCR detection of VPPV in the systemic leaves of <span class="html-italic">N. benthamiana</span> plants infiltrated with <span class="html-italic">A. tumefaciens</span> culture carrying the pCB301-VPPV (line 1 and 2) or carrying the empty pCB301 vector (CK-). pCB301-VPPV plasmid was used as PCR positive control (CK+). (<b>b</b>) Symptoms of <span class="html-italic">N. benthamiana</span> plants infiltrated with <span class="html-italic">A. tumefaciens</span> culture carrying the pCB301-VPPV vector (<b>left</b>) or carrying the empty pCB301 vector (the CK- control, <b>right</b>). (<b>c</b>) Symptoms of <span class="html-italic">N. benthamiana</span> plants expressing VPPV P0, P1, RdRp, P3a, P3, P4, or RTP using PVX heterologous expression system. White arrows indicate necrotic patches. (<b>d</b>) Western blot analysis of PVX CP accumulation in systemic leaves of PVX inoculated <span class="html-italic">N. benthamiana</span> at 10 days post inoculation (dpi). The Rubisco stained by Ponceau S was used as the loading control.</p>
Full article ">
22 pages, 7210 KiB  
Article
Single Dose of Attenuated Vaccinia Viruses Expressing H5 Hemagglutinin Affords Rapid and Long-Term Protection Against Lethal Infection with Highly Pathogenic Avian Influenza A H5N1 Virus in Mice and Monkeys
by Fumihiko Yasui, Keisuke Munekata, Tomoko Fujiyuki, Takeshi Kuraishi, Kenzaburo Yamaji, Tomoko Honda, Sumiko Gomi, Misako Yoneda, Takahiro Sanada, Koji Ishii, Yoshihiro Sakoda, Hiroshi Kida, Shosaku Hattori, Chieko Kai and Michinori Kohara
Vaccines 2025, 13(1), 74; https://doi.org/10.3390/vaccines13010074 - 15 Jan 2025
Viewed by 689
Abstract
Background/Objectives: In preparation for a potential pandemic caused by the H5N1 highly pathogenic avian influenza (HPAI) virus, pre-pandemic vaccines against several viral clades have been developed and stocked worldwide. Although these vaccines are well tolerated, their immunogenicity and cross-reactivity with viruses of different [...] Read more.
Background/Objectives: In preparation for a potential pandemic caused by the H5N1 highly pathogenic avian influenza (HPAI) virus, pre-pandemic vaccines against several viral clades have been developed and stocked worldwide. Although these vaccines are well tolerated, their immunogenicity and cross-reactivity with viruses of different clades can be improved. Methods: To address this aspect, we generated recombinant influenza vaccines against H5-subtype viruses using two different strains of highly attenuated vaccinia virus (VACV) vectors. Results: rLC16m8-mcl2.2 hemagglutinin (HA) and rLC16m8-mcl2.3.4 HA consisted of a recombinant LC16m8 vector encoding the HA protein from clade 2.2 or clade 2.3.4 viruses (respectively); rDIs-mcl2.2 HA consisted of a recombinant DIs vector encoding the HA protein from clade 2.2. A single dose of rLC16m8-mcl2.2 HA showed rapid (1 week after vaccination) and long-term protection (20 months post-vaccination) in mice against the HPAI H5N1 virus. Moreover, cynomolgus macaques immunized with rLC16m8-mcl2.2 HA exhibited long-term protection when challenged with a heterologous clade of the HPAI H5N1 virus. Although the DIs strain is unable to grow in most mammalian cells, rDIs-mcl2.2 HA also showed rapid and long-lasting effects against HPAI H5N1 virus infection. Notably, the protective efficacy of rDIs-mcl2.2 HA was comparable to that of rLC16m8-mcl2.2 HA. Furthermore, these vaccines protected animals previously immunized with VACVs from a lethal challenge with the HPAI H5N1 virus. Conclusions: These results suggest that both rLC16m8-mcl2.2 HA and rDIs-mcl2.2 HA are effective in preventing HPAI H5N1 virus infection, and rDIs-mcl2.2 HA is a promising vaccine candidate against H5 HA-subtype viruses. Full article
Show Figures

Figure 1

Figure 1
<p>Protective efficacy of recombinant vaccinia virus LC16m8 strain (rLC16m8) expressing H5 hemagglutinin (HA) protein against lethal infection with highly pathogenic avian influenza (HPAI) H5N1 virus in mice. (<b>A</b>–<b>D</b>) Vaccination and infection studies in naïve BALB/c mice. (<b>A</b>) Experimental schedule of vaccination and H5N1 influenza virus infection in mice. Eight-week-old female BALB/c mice were inoculated intradermally with 1 × 10<sup>7</sup> PFU of rLC16m8-mcl2.2 HA, rLC16m8-mcl2.3.4 HA, or rLC16m8-empty; 5 weeks after vaccination, animals were infected intranasally with 1 × 10<sup>4</sup> PFU of H5N1 A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1). rLC16m8-mcl2.2 HA, rLC16m8 encoding the H5 HA protein (clade 2.2, A/Qinghai/1A/05); rLC16m8-mcl2.3.4 HA, rLC16m8 encoding the H5 HA protein (clade 2.3.4, A/Anhui/1/05); rCL16m8-empty, rCL16m8 harboring only the ATI/p7.5 hybrid promoter sequence. (<b>B</b>) Body weight was monitored daily after the H5N1 virus infection. Values are shown as mean ± SD. (<b>C</b>) Survival rate was observed until 9 days post-infection (dpi). (<b>D</b>) Pulmonary virus titer was determined in four mice per group at each time point after H5N1 influenza virus infection. Values are shown as geometric mean ± geometric SD. <span class="html-italic">p</span> values are calculated using two-tailed non-paired one-way ANOVA followed by Turkey’s test. (<b>E</b>,<b>F</b>) Vaccination and infection studies in BALB/c mice sensitized with vaccinia virus (VACV). Eight-week-old female BALB/c mice were inoculated intradermally with either 1×10<sup>7</sup> PFU of rLC16m8-empty or culture medium (vehicle). Four weeks after sensitization with VACV, these mice were further inoculated with 1 × 10<sup>7</sup> PFU of either rLC16m8-mcl2.2 HA or rLC16m8-empty. Four weeks after vaccination with rLC16m8-based viruses, animals were infected with 1 × 10<sup>4</sup> PFU of H5N1 A/whooper swan/Hokkaido/1/2008. * <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>E</b>) Body weight was monitored daily after H5N1 virus infection. Values are shown as mean ± SD. (<b>F</b>) Survival rate was observed until twelve dpi. Survival rates were compared in panels (<b>C</b>,<b>F</b>) using the Gehan-Breslow-Wilcoxon method.</p>
Full article ">Figure 2
<p>Rapid protective effects induced by rLC16m8-H5 HA against lethal infection with the HPAI H5N1 virus in mice. (<b>A</b>) Experimental schedule of vaccination and H5N1 influenza virus infection in BALB/c mice. Female BALB/c mice were inoculated intradermally with 1 × 10<sup>7</sup> PFUs of rLC16m8-mcl2.2 HA, rLC16m8-mcl2.3.4 HA, or rLC16m8-empty (<b>B</b>) 2 weeks, (<b>C</b>) 1 week, or (<b>D</b>) 3 days before intranasal infection with 1 × 10<sup>4</sup> PFUs of H5N1 A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1). (<b>B</b>) BALB/c mice were infected with the H5N1 HPAI virus 2 weeks after vaccination with rLC16m8-H5 HA. After infection with the virus, body weight was monitored daily (left panel), and the survival rate was recorded until 9 days post-infection (dpi) (right panel). (<b>C</b>) One week after vaccination, the rapid protective efficacy of rLC16m8-H5 HA against H5N1 HPAI virus infection was evaluated by monitoring body weight daily (left panel) and recording survival until 9 dpi (right panel). (<b>D</b>) Three days after vaccination, the rapid protective efficacy of rLC16m8-H5 HA against H5N1 HPAI virus infection was evaluated by monitoring body weight daily (left panel) and recording survival until 9 dpi (right panel). Survival rates are compared in panels (<b>B</b>–<b>D</b>) using the Gehan–Breslow–Wilcoxon method. NS: not significant. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Long-term protection by rLC16m8-H5 HA against lethal infection with the HPAI H5N1 virus in mice. Female BALB/c mice were inoculated intradermally with 1 × 10<sup>7</sup> PFUs of rLC16m8-mcl2.2 HA or rLC16m8-empty [<span class="html-italic">n</span> = 3–5 in each group except for <span class="html-italic">n</span> = 2 in rLC16m8-mcl2.2 HA in (<b>E</b>)]. (<b>A</b>) The time course of production of immunoglobulin G (IgG) specific to H5 HA (clade 2.2) after vaccination was evaluated by an ELISA. (<b>B</b>) The HI titer of antisera from mice immunized with rLC16m8-mlc.2.2 HA (<span class="html-italic">n</span> = 4) or rLC16m8-empty (<span class="html-italic">n</span> = 4) 1 and 6 months after immunization. HI titers were determined against HPAI H5N1 A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1; Hok/1/08), A/whooper swan/Mongolia/3/2005 (clade 2.2; Mon/3/05), and A/Vietnam/UT3040/2004 (clade 1; UT3040/04) using 0.75% guinea pig erythrocytes. (<b>C</b>–<b>E</b>) Vaccinated mice were challenged with 166 × 50% mouse lethal dose (MLD<sub>50</sub>) of HPAI H5N1 A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1) 12 (<b>C</b>) or 18 (<b>D</b>) months after vaccination. (<b>E</b>) Vaccinated mice were challenged with 150 × MLD<sub>50</sub> of H5N1 A/Vietnam/UT3040/2004 (clade 1) 20 months after vaccination. Left panels show body weight changes when monitored daily after the H5N1 virus challenge. Right panels show the survival rate when assessed until 9 days post-infection. ND: not determined (rLC16m8-mcl2.2 HA, <span class="html-italic">n</span> = 2; rLC16m8-empty, <span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Long-term protection by rLC16m8-H5 HA against lethal infection with the HPAI H5N1 virus in cynomolgus macaques. Female cynomolgus macaques were inoculated intradermally with 1 × 10<sup>7</sup> PFUs of rLC16m8-mcl2.2 HA (<span class="html-italic">n</span> = 3) or rLC16m8-empty (<span class="html-italic">n</span> = 2) on their upper arms. Twelve months after vaccination, HPAI H5N1 virus A/whooper swan/Hokkaido/1/2008 was inoculated into the nostrils, oral cavity, and trachea of each macaque. (<b>A</b>) The time course of production of IgG specific to the H5 HA protein (clade 2.2) after vaccination was evaluated using an ELISA. (<b>B</b>) The mean value of the body temperature of individual macaques from 8 p.m. to 8 a.m. every night was calculated from the temperature recorded every 5 min. Body temperature changes of individual macaques on each day after virus infection were compared with mean temperature changes from 8 p.m. on day 1 to 8 a.m. on day 0 before virus infection. (<b>C</b>) Cumulative temperature increase, calculated as the area under the curve (AUC) from the data recorded 3 days post-infection (dpi) to 7 dpi in (<b>B</b>). (<b>D</b>) Temporal changes in viral titers in nasal (upper panel), oral (middle panel), and tracheal (lower panel) swab samples were determined by a 50% tissue culture infectious dose (TCID<sub>50</sub>) assay using Madin–Darby canine kidney (MDCK) cells. The dpis are indicated in navy (1 dpi), red (3 dpi), green (5 dpi), and yellow (7 dpi). The lower limit of detection (1.7 log units) is indicated by a horizontal dashed line.</p>
Full article ">Figure 5
<p>Histopathology and viral load in the lungs of cynomolgus macaques after HPAI H5N1 virus infection. (<b>A</b>) Representative lung sections (hematoxylin and eosin staining; section thickness 4 μm) at 7 days post-infection (dpi); original magnification was 100×. Bar, 200 μm. The number indicates animal ID. (<b>B</b>) Histopathological scores were obtained for each of the 15 defined regions of the lung lobe of each animal (RU, right upper; RM, right middle; RL, right lower; LU, left upper; LM, left middle; LL, left lower) 7 dpi with rLC16m8-empty or rLC16m8-mcl2.2 HA. Red horizontal bars indicate the mean pathological score in each group. <span class="html-italic">p</span> values were calculated using the Mann–Whitney U test. (<b>C</b>) Viral load in all lung lobes was determined by reverse transcription–quantitative PCR. The central horizontal value represents the geometric mean, and the whiskers indicate the geometric SD. <span class="html-italic">p</span> values were calculated by two-tailed non-paired Student’s <span class="html-italic">t</span>-tests. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Rapid and long-term protection by a single dose of replication-deficient rDIs-mcl2.2 HA against lethal infection with H5N1 HPAI virus in mice. Female BALB/c mice were inoculated intradermally with 1 × 10<sup>7</sup> PFU of rDIs-mcl2.2 HA or DIs. (<b>A</b>) Vaccinated mice (rDIs-mcl2.2 HA, <span class="html-italic">n</span> = 4; DIs, <span class="html-italic">n</span> = 4) were infected with 1 × 10<sup>4</sup> PFUs of H5N1 A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1) 5 weeks after vaccination. Left: body weight was monitored daily after H5N1 virus infection. Right: survival rate was observed until 12 days post-infection (dpi). (<b>B</b>,<b>C</b>) The speed of protection by rDIs-mcl2.2 HA against the HPAI H5N1 virus was investigated. Two weeks (<b>B</b>) or one week (<b>C</b>) after vaccination, mice were infected intranasally with 166 × MLD<sub>50</sub> of A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1) and then monitored daily for changes in their body weight (left panel) and survival rate until 12 dpi (right panel). (<b>D</b>,<b>E</b>) Long-term immunity by a single dose of rDIs-mcl2.2 HA was investigated. (<b>D</b>) The time course of production of IgG specific to the HA protein (clade 2.2) was measured by an ELISA. One thousand-fold diluted murine sera were used. (<b>E</b>) The HI titer of antisera from mice immunized with rDIs-mlc.2.2 HA (<span class="html-italic">n</span> = 4) or DIs (<span class="html-italic">n</span> = 4) 1 and 6 months after immunization. HI titers were determined against HPAI H5N1 A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1; Hok/1/08), A/whooper swan/Mongolia/3/2005 (clade 2.2; Mon/3/05), and A/Vietnam/UT3040/2004 (clade 1; UT3040/04) using 0.75% guinea pig erythrocytes. (<b>F</b>) Vaccinated mice were infected intranasally with 1 × 10<sup>4</sup> PFUs of H5N1 A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1) after 20 months. Survival rates are compared in data shown in panels (<b>A</b>–<b>C</b>,<b>F</b>) using the Gehan–Breslow–Wilcoxon method. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Protective efficacy of rDIs-mcl2.2 HA against the HPAI H5N1 virus in VACV-sensitized mice. (<b>A</b>) Experimental schedule. To investigate the protective efficacy of a single dose of rDIs-mcl2.2 HA against lethal infection with the HPAI H5N1 virus in VACV-sensitized mice, female BALB/c mice were sensitized intradermally with 1 × 10<sup>7</sup> PFUs of the VACV LC16m8 strain and then immunized intradermally with rDIs-mcl2.2 HA (1 × 10<sup>7</sup> PFUs, 3 × 10<sup>7</sup> PFUs, or 1 × 10<sup>8</sup> PFUs) 22 weeks after VACV sensitization (<b>D</b>). Age-matched naïve mice were used as controls (<b>C</b>). (<b>B</b>) A total of 21 weeks after VACV sensitization, the neutralization titer (50% neutralization) against LC16m8 was measured. Dashed lines denote the limits of detection. (<b>C</b>) Naïve mice were inoculated with rDIs-mcl2.2 HA and then infected intranasally with 166 × MLD<sub>50</sub> of A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1) 5 weeks after vaccination. Body weight (left panel) and survival rate (right panel) were monitored daily. (<b>D</b>) VACV-sensitized mice were inoculated with rDIs-mcl2.2 HA and then infected intranasally with 166 × MLD<sub>50</sub> of A/whooper swan/Hokkaido/1/2008 (clade 2.3.2.1) 5 weeks after vaccination. Body weight (left panel) and survival rate (right panel) were monitored daily. Survival rates are compared in panels (<b>C</b>,<b>D</b>) using the Gehan–Breslow–Wilcoxon method. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Th1/Th2 immune reponse to rVACV-mcl2.2 HA vaccine. (<b>A</b>) IgG1 (left) and IgG2a (right) responses against H5 HA clade 2.2 in mice vaccinated with rLC16m8-mcl2.2 HA (<span class="html-italic">n</span> = 4; red) or rDIs-mcl2.2 HA (<span class="html-italic">n</span> = 4) 1 and 6 months after immunization. Antisera from rLC16m8-empty-immunized (<span class="html-italic">n</span> = 4; white) and DIs-immunized (<span class="html-italic">n</span> = 4) mice were used as negative controls. The dashed line indicates a minimal dilution rate (1:1000) of antisera used in the ELISA. The endpoint titers of negative controls were defined as 500. (<b>B</b>) Th1/Th2 skewing responses in mice vaccinated with rLC16m8-mcl2.2 HA (<span class="html-italic">n</span> = 4) or rDIs-mcl2.2 HA (<span class="html-italic">n</span> = 4) 1 and 6 months after immunization. The IgG2a/IgG1 ratio was calculated using the respective endpoint titer values. (<b>C</b>) IFN-γ (left) or IL-4 (right) levels were measured via an ELISpot assay using the splenocytes of mice immunized with rDIs-mcl2.2 HA or DIs 1 month after immunization. Values are shown as the mean ± SD. Statistical analysis was performed using two-tailed one-way ANOVA with post hoc Tukey’s multiple comparison test.</p>
Full article ">
20 pages, 1270 KiB  
Review
Current Understanding on the Heterogenous Expression of Plastic Depolymerising Enzymes in Pichia pastoris
by Shuyan Wu, David Hooks and Gale Brightwell
Bioengineering 2025, 12(1), 68; https://doi.org/10.3390/bioengineering12010068 - 14 Jan 2025
Viewed by 475
Abstract
Enzymatic depolymerisation is increasingly recognised as a reliable and environmentally friendly method. The development of this technology hinges on the availability of high-quality enzymes and associated bioreaction systems for upscaling biodegradation. Microbial heterologous expression systems have been studied for meeting this demand. Among [...] Read more.
Enzymatic depolymerisation is increasingly recognised as a reliable and environmentally friendly method. The development of this technology hinges on the availability of high-quality enzymes and associated bioreaction systems for upscaling biodegradation. Microbial heterologous expression systems have been studied for meeting this demand. Among these systems, the Pichia pastoris expression system has emerged as a widely used platform for producing secreted heterologous proteins. This article provides an overview of studies involving the recombinant expression of polymer-degrading enzymes using the P. pastoris expression system. Research on P. pastoris expression of interested enzymes with depolymerising ability, including cutinase, lipase, and laccase, are highlighted in the review. The key factors influencing the heterologous expression of polymer-degrading enzymes in P. pastoris are discussed, shedding light on the challenges and opportunities in the development of depolymerising biocatalysts through the P. pastoris expression system. Full article
(This article belongs to the Special Issue Synthetic Biology and Bioprocess Engineering for High-Value Compounds)
Show Figures

Figure 1

Figure 1
<p>Highlighted factors influencing the effective functional overexpression of plastic-depolymerising enzymes in <span class="html-italic">P. pastoris</span>.</p>
Full article ">Figure 2
<p>Overview of the heterogenous expression of depolymerising enzymes in the <span class="html-italic">P. pastoris</span> system (generated by BioRender).</p>
Full article ">
13 pages, 867 KiB  
Article
Comparing Antibody Responses to Homologous vs. Heterologous COVID-19 Vaccination: A Cross-Sectional Analysis in an Urban Bangladeshi Population
by Kazi Istiaque Sanin, Mansura Khanam, Azizur Rahman Sharaque, Mahbub Elahi, Bharati Rani Roy, Md. Khaledul Hasan, Goutam Kumar Dutta, Abir Dutta, Md. Nazmul Islam, Md. Safiqul Islam, Md. Nasir Ahmed Khan, Mustufa Mahmud, Nuzhat Nadia, Fablina Noushin, Anjan Kumar Roy, Protim Sarker and Fahmida Tofail
Vaccines 2025, 13(1), 67; https://doi.org/10.3390/vaccines13010067 - 13 Jan 2025
Viewed by 518
Abstract
Background: Vaccination has played a crucial role in mitigating the spread of COVID-19 and reducing its severe outcomes. While over 90% of Bangladesh’s population has received at least one COVID-19 vaccine dose, the comparative effectiveness of homologous versus heterologous booster strategies, along with [...] Read more.
Background: Vaccination has played a crucial role in mitigating the spread of COVID-19 and reducing its severe outcomes. While over 90% of Bangladesh’s population has received at least one COVID-19 vaccine dose, the comparative effectiveness of homologous versus heterologous booster strategies, along with the complex interplay of factors within the population, remains understudied. This study aimed to compare antibody responses between these booster approaches. Methods: This cross-sectional study enrolled 723 adults in urban Dhaka who had received COVID-19 booster doses within the last six months. Participants were grouped based on homologous or heterologous booster vaccination. Data were collected through structured household surveys, and 2 mL blood samples were collected for measuring antibody titers. Results: Heterologous booster recipients showed higher median antibody titers (8597.0 U/mL, IQR 5053.0–15,482.3) compared to homologous recipients (6958.0 U/mL, IQR 3974.0–12,728.5). In the adjusted analysis, the type of booster dose had no significant impact on antibody levels. However, the duration since the last booster dose was significantly associated with antibody levels, where each additional month since receiving the booster corresponded to approximately a 15–16% reduction in antibody levels (Adj. coeff: 0.85, 95% CI: 0.81, 0.88; p < 0.001). Participants over 40 years demonstrated higher antibody levels than younger individuals (Adj. coeff: 1.23, 95% CI: 1.07, 1.43; p = 0.005). Sex, BMI, and prior COVID-19 infection showed no significant associations with antibody levels after adjustment. Conclusion: The results underscore the complexity of immune responses across different demographic groups and suggest potential benefits of ongoing heterologous booster strategies in sustaining immunity. Full article
(This article belongs to the Section COVID-19 Vaccines and Vaccination)
Show Figures

Figure 1

Figure 1
<p>Antibody titer level by type of vaccination.</p>
Full article ">Figure 2
<p>Antibody responses with time of receiving booster by homologous and heterologous group.</p>
Full article ">
16 pages, 4771 KiB  
Article
Heterologous and High Production of Ergothioneine in Bacillus licheniformis by Using Genes from Anaerobic Bacteria
by Zhe Liu, Fengxu Xiao, Yupeng Zhang, Jiawei Lu, Youran Li and Guiyang Shi
Metabolites 2025, 15(1), 45; https://doi.org/10.3390/metabo15010045 - 12 Jan 2025
Viewed by 479
Abstract
Purpose: This study aimed to utilize genetically engineered Bacillus licheniformis for the production of ergothioneine (EGT). Given the value of EGT and the application of Bacillus licheniformis in enzyme preparation production, we cloned the key enzymes (EanA and EanB) from Chlorbium limicola. [...] Read more.
Purpose: This study aimed to utilize genetically engineered Bacillus licheniformis for the production of ergothioneine (EGT). Given the value of EGT and the application of Bacillus licheniformis in enzyme preparation production, we cloned the key enzymes (EanA and EanB) from Chlorbium limicola. Through gene alignment, new ergothioneine synthase genes (EanAN and EanBN) were identified and then expressed in Bacillus licheniformis to construct strains. Additionally, we investigated the factors influencing the yield of EGT and made a comparison with Escherichia coli. Methods: The relevant genes were cloned and transferred into Bacillus licheniformis. Fermentation experiments were conducted under different conditions for yield analysis, and the stability of this bacterium was also evaluated simultaneously. Results: The constructed strains were capable of producing EGT. Specifically, the yield of the EanANBN strain reached (643.8 ± 135) mg/L, and its stability was suitable for continuous production. Conclusions: Genetically engineered Bacillus licheniformis demonstrates potential in the industrial-scale production of EGT. Compared with Escherichia coli, it has advantages, thus opening up new possibilities for the application and market supply of EGT. Full article
(This article belongs to the Section Microbiology and Ecological Metabolomics)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">N. crassa</span> pathway versus comparison of the <span class="html-italic">C. limicola</span> pathway. Blue represents biosynthetic pathways of EGT in <span class="html-italic">N. crassa</span> [Egt1: Bifunctional enzymes (SAM-dependent histidine methyltransferase and mononuclear non-heme iron enzyme); Egt2: PLP-mediated C-S lyase]. Red represents biosynthetic pathways of EGT in <span class="html-italic">C limicola</span> [EanA: Methyltransferase; EanB: rhodanese-like sulfur transferase] (<b>A</b>) and amino acid sequence alignment of methyltransferases (EanA and EanAN) and sulfurtransferases (EanB and EanBN) from two different sources [Dark blue indicates sequence consistency] (<b>B</b>).</p>
Full article ">Figure 2
<p>Ergothioneine biosynthesis pathway from <span class="html-italic">Chlorbium limicola</span> (<b>A</b>) and Ergothioneine production of two recombinant strains (<b>B</b>) are shown.</p>
Full article ">Figure 3
<p>Ergothioneine is produced using the whole-cell transformation method. The yield of ergothioneine from recombinant bacteria EanAB at different temperatures (<b>A</b>), the yield of ergothioneine from recombinant bacteria EanANBN at different temperatures (<b>B</b>), and the detection of changes in pH and glucose concentration during the transformation process (<b>C</b>) are shown.</p>
Full article ">Figure 4
<p>Effect of different initial pH at 25, 30, and 37 degrees on ergothioneine conversion. Initial OD and EGT yields of two recombinant bacteria at 25 degrees Celsius (<b>A</b>,<b>B</b>), initial OD and EGT yields of two recombinant bacteria at 30 degrees Celsius (<b>C</b>,<b>D</b>), and initial OD and EGT yields of two recombinant bacteria at 37 degrees Celsius (<b>E</b>,<b>F</b>) are shown.</p>
Full article ">Figure 5
<p>Yield of EGT by two recombinant bacteria under different conditions. Changes in extracellular amino acid concentration and EGT yield after exogenous addition of a single amino acid precursor (<b>A</b>,<b>B</b>), changes in extracellular amino acid concentration and EGT yield after exogenous addition of four mixed amino acid precursors (<b>C</b>,<b>D</b>), effect of the addition of different groups of precursor amino acids on EGT yield (<b>E</b>), and final yield of EGT in recombinant shake flask culture (<b>F</b>,<b>G</b>) are shown.</p>
Full article ">
15 pages, 3655 KiB  
Article
Truncated NS1 Influenza A Virus Induces a Robust Antigen-Specific Tissue-Resident T-Cell Response and Promotes Inducible Bronchus-Associated Lymphoid Tissue Formation in Mice
by Anna-Polina Shurygina, Marina Shuklina, Olga Ozhereleva, Ekaterina Romanovskaya-Romanko, Sofia Kovaleva, Andrej Egorov, Dmitry Lioznov and Marina Stukova
Vaccines 2025, 13(1), 58; https://doi.org/10.3390/vaccines13010058 - 10 Jan 2025
Viewed by 475
Abstract
Background: Influenza viruses with truncated NS1 proteins show promise as viral vectors and candidates for mucosal universal influenza vaccines. These mutant NS1 viruses, which lack the N-terminal half of the NS1 protein (124 a.a.), are unable to antagonise the innate immune response. This [...] Read more.
Background: Influenza viruses with truncated NS1 proteins show promise as viral vectors and candidates for mucosal universal influenza vaccines. These mutant NS1 viruses, which lack the N-terminal half of the NS1 protein (124 a.a.), are unable to antagonise the innate immune response. This creates a self-adjuvant effect enhancing heterologous protection by inducing a robust CD8+ T-cell response together with immunoregulatory mechanisms. However, the effects of NS1 modifications on T-follicular helper (Tfh) and B-cell responses remain less understood. Methods: C57bl/6 mice were immunised intranasally with 10 μL of either an influenza virus containing a truncated NS1 protein (PR8/NS124), a cold-adapted influenza virus with a full-length NS1 (caPR8/NSfull), or a wild-type virus (PR8/NSfull). Immune responses were assessed on days 8 and 28 post-immunisation by flow cytometry, ELISA, and HAI assay. Results: In this study, we demonstrate that intranasal immunisation with PR8/NS124 significantly increases tissue-resident CD4+ and CD8+ T cells in the lungs and activates Tfh cells in regional lymph nodes as early as day 8 post-immunisation. These effects are not observed in mice immunised with caPR8/NSfull or PR8/NSfull. Notably, PR8/NS124 immunisation also leads to the development of inducible bronchus-associated lymphoid tissue (iBALT) in the lungs by day 28, characterised by the presence of antigen-specific Tfh cells and GL7+Fas+ germinal centre B cells. Conclusions: Our findings further underscore the potential of NS1-truncated influenza viruses to drive robust mucosal immune responses and enhance vaccine efficacy. Full article
(This article belongs to the Special Issue The Recent Development of Influenza Vaccine: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Safety and reproductive activity of PR8/NSfull, caPR8/NSfull, and PR8/NS124 viruses. Study design and sampling plan (<b>A</b>). Under light ether anaesthesia, groups of ten 6–8-week-old female C57bl/6 mice were inoculated with 10 μL of either PR8/NS124 or caPR8/NSfull at a dose of 6.0 Lg EID<sub>50</sub>/animal or with PR8/NSfull at a sublethal dose of 3.0 Lg EID<sub>50</sub>/animal. Body weight dynamics were monitored for 8 days after immunisation. Nasal turbinates (NT) and lungs were collected on days 2 and 4 p.im. to assess virus shedding. T-cellular response was evaluated in draining lymph nodes (LN) and lungs on days 8 and 28 p.im. Serum and nasal wash (NW) samples for humoral response assessment were collected on day 28 p.im. Virus shedding in nasal turbinates (<b>B</b>) and lungs (<b>C</b>). Viral loads in 10% nasal turbinate (NT) and lung suspensions were determined on days 2 and 4 post-immunisation in ECE. Virus titres are expressed as Lg EID<sub>50</sub>/mL. The detection threshold was 1.7 Lg EID<sub>50</sub>/mL (dotted line). If no virus was detected, a value of 1.2 Lg EID<sub>50</sub>/mL was assigned. Mice with a viral load &lt;1.7 Lg EID<sub>50</sub>/mL were considered not infected. Body weight dynamics (<b>D</b>) was assessed for 8 days p.im. and expressed as a percentage of the initial body weight (Mean ± SD). Data were considered statistically significant at <span class="html-italic">p</span> &lt; 0.05, as determined by two-way ANOVA followed by Tukey’s multiple comparison test (*: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 2
<p>Local and systemic humoral immune responses. Nasal wash IgA titre (<b>A</b>), total serum IgG (<b>B</b>), anti-hemagglutinating antibodies (<b>C</b>), IgG1 (<b>D</b>), IgG2b (<b>E</b>), IgG3 (<b>F</b>). Nasal wash and serum antibodies were measured by ELISA or HAI. Data are presented as individual log<sub>2</sub> titres with a geometric mean (horizontal line). Data were considered statistically significant at <span class="html-italic">p</span> &lt; 0.05, as determined by one-way ANOVA followed by Tukey’s multiple comparison test (*: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 3
<p>Antigen-specific CD8+ Trm response in the lungs. Trm response in the lungs was assessed at 8 (<b>A</b>,<b>B</b>) and 28 (<b>C</b>,<b>D</b>) d.p.im. by intracellular cytokine staining after 6 h of in vitro stimulation with PepTivator<sup>®</sup> Influenza A (H1N1) NP supplemented with the NP<sub>366–374</sub> peptide. The total percentage of cytokine-producing CD8+ Trm lymphocytes (<b>A</b>,<b>C</b>) and the percentage of CD8+ Trm producing any combination of IFNγ, IL2, or TNFα (<b>B</b>,<b>D</b>) are shown as box and whiskers plots (min and max with individual values and the median indicated). Data were considered statistically significant at <span class="html-italic">p</span> &lt; 0.05, as determined by one-or two-way ANOVA followed by Tukey’s multiple comparison test (*: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>Antigen-specific CD4+ Trm response in the lungs. Trm response in the lungs was assessed at 8 (<b>A</b>,<b>B</b>) and 28 (<b>C</b>,<b>D</b>) d.p.im. by intracellular cytokine staining after 6 h of in vitro stimulation with PepTivator<sup>®</sup> Influenza A (H1N1) NP. The total percentage of cytokine-producing CD4+ Trm lymphocytes (<b>A</b>,<b>C</b>) and the percentage of CD4+ Trm producing any combination of IFNγ, IL2, or TNFα (<b>B</b>,<b>D</b>) are shown as box and whiskers plots (min and max with individual values and the median indicated). Data were considered statistically significant at <span class="html-italic">p</span> &lt; 0.05, as determined by one-or two-way ANOVA followed by Tukey’s multiple comparison test (*: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 5
<p>Tfh responses in lymph nodes. Tfh (CXCR5+ or CXCR5+/ CXCR3+) subpopulations were evaluated in lymph nodes at 8 d.p.im. The percentage of Tfh cells in the parent population (CD4+ Tem) (<b>A</b>,<b>C</b>), ICOS+ Tfh (<b>B</b>,<b>D</b>), and Tfh1/Tfh2/Tfh17 subsets among lymph node Tfh em (<b>E</b>) are shown as box and whiskers plots (min and max with individual values and the median indicated). Data were considered statistically significant at <span class="html-italic">p</span> &lt; 0.05, as determined by one- or two-way ANOVA followed by Tukey’s multiple comparison test (*: <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; 0.0001).</p>
Full article ">Figure 6
<p>Tfh and germinal centre B-cell responses in the lungs. Tfh subpopulations were evaluated in the lungs at 28 d.p.im. The percentage of Tfh (CXCR5+ CXCR3+) cells in the parent population (CD4+ Tem) (<b>A</b>), the percentage of CL7+Fas+ GC B cells (<b>B</b>), Tfh1/Tfh2/Tfh17 subsets among lung Tfh em (<b>C</b>), and percentages of antigen-specific cytokine-producing Tfh cells are shown as box and whiskers plots (min and max with individual values and the median indicated) (<b>D</b>). Data were considered statistically significant at <span class="html-italic">p</span> &lt; 0.05, as determined by one-or two-way ANOVA followed by Tukey’s multiple comparison test (*: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">
17 pages, 1391 KiB  
Systematic Review
Autologous and Heterologous Minor and Major Bone Regeneration with Platelet-Derived Growth Factors
by Gianna Dipalma, Angelo Michele Inchingolo, Valeria Colonna, Pierluigi Marotti, Claudio Carone, Laura Ferrante, Francesco Inchingolo, Andrea Palermo and Alessio Danilo Inchingolo
J. Funct. Biomater. 2025, 16(1), 16; https://doi.org/10.3390/jfb16010016 - 9 Jan 2025
Viewed by 431
Abstract
Aim: This review aims to explore the clinical applications, biological mechanisms, and potential benefits of concentrated growth factors (CGFs), autologous materials, and xenografts in bone regeneration, particularly in dental treatments such as alveolar ridge preservation, mandibular osteonecrosis, and peri-implantitis. Materials and Methods. A [...] Read more.
Aim: This review aims to explore the clinical applications, biological mechanisms, and potential benefits of concentrated growth factors (CGFs), autologous materials, and xenografts in bone regeneration, particularly in dental treatments such as alveolar ridge preservation, mandibular osteonecrosis, and peri-implantitis. Materials and Methods. A systematic literature search was conducted using databases like PubMed, Scopus, and Web of Science, with keywords such as “bone regeneration” and “CGF” from 2014 to 2024. Only English-language clinical studies involving human subjects were included. A total of 10 studies were selected for qualitative analysis. Data were processed through multiple stages, including title and abstract screening and full-text evaluation. Conclusion: The findings of the reviewed studies underscore the potential of the CGF in enhancing bone regeneration through stimulating cell proliferation, angiogenesis, and extracellular matrix mineralization. Autologous materials have also demonstrated promising results due to their biocompatibility and capacity for seamless integration with natural bone tissue. When combined with xenografts, these materials show synergistic effects in improving bone quantity and quality, which are crucial for dental implant success. Future research should focus on direct comparisons of different techniques, the optimization of protocols, and broader applications beyond dental medicine. The integration of CGFs and autologous materials into routine clinical practice represents a significant advancement in regenerative dental medicine, with the potential for improved patient outcomes and satisfaction. Full article
(This article belongs to the Special Issue Advanced Biomaterials for Bone Tissue Engineering)
Show Figures

Figure 1

Figure 1
<p>PRISMA flowchart.</p>
Full article ">Figure 2
<p>Risk of bias [<a href="#B8-jfb-16-00016" class="html-bibr">8</a>,<a href="#B12-jfb-16-00016" class="html-bibr">12</a>,<a href="#B17-jfb-16-00016" class="html-bibr">17</a>,<a href="#B21-jfb-16-00016" class="html-bibr">21</a>,<a href="#B27-jfb-16-00016" class="html-bibr">27</a>,<a href="#B37-jfb-16-00016" class="html-bibr">37</a>,<a href="#B52-jfb-16-00016" class="html-bibr">52</a>,<a href="#B53-jfb-16-00016" class="html-bibr">53</a>,<a href="#B54-jfb-16-00016" class="html-bibr">54</a>,<a href="#B55-jfb-16-00016" class="html-bibr">55</a>].</p>
Full article ">
18 pages, 4979 KiB  
Article
The Functional Identification of the CYP2E1 Gene in the Kidney of Lepus yarkandensis
by Dingwei Shao, Ke Sheng, Bing Chao, Yumei Tong, Renjun Jiang and Jianping Zhang
Int. J. Mol. Sci. 2025, 26(2), 453; https://doi.org/10.3390/ijms26020453 - 8 Jan 2025
Viewed by 422
Abstract
This study aims to identify the function of the cytochrome P450 2E1 (CYP2E1) gene in the kidneys of Lepus yarkandensis. CYP2E1 is a significant metabolic enzyme involved in the metabolism of various endogenous and exogenous compounds and is associated with [...] Read more.
This study aims to identify the function of the cytochrome P450 2E1 (CYP2E1) gene in the kidneys of Lepus yarkandensis. CYP2E1 is a significant metabolic enzyme involved in the metabolism of various endogenous and exogenous compounds and is associated with the occurrence and progression of multiple diseases. Given L. yarkandensis’s ability to survive in the extremely arid L. yarkandensis, we hypothesize that CYP2E1 in its kidneys plays a crucial role in adaptability. Through molecular cloning and sequence analysis, we discovered that the CYP2E1 gene of Lepus yarkandensis encodes a protein of 493 amino acids. The 493-amino acid protein encoded by the Lepus yarkandensis CYP2E1 gene shows 13 amino acid variation sites compared to the homologous protein in Oryctolagus cuniculus. The protein is primarily localized to the endoplasmic reticulum membrane and lacks transmembrane structures. In the yeast expression system, the heterologous expression of the CYP2E1 gene enhanced the yeast’s tolerance to drought, salinity, and high temperatures, achieved by increasing antioxidant enzyme activity and reducing levels of oxidative stress markers. Additionally, this study identified a “Yeast Oxidative Stress Lethal Threshold (Yeast OSLT)” under specific stress conditions. Once this threshold is exceeded, the cell’s antioxidant defense system can no longer maintain cellular homeostasis, leading to massive cell death. Although CYP2E1 did not change this threshold, it contributed to cell survival to some extent. These findings not only reveal the function of L. yarkandensis CYP2E1 in stress adaptation but also provide valuable molecular insights into its survival strategy in extreme environments. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Cloning of the <span class="html-italic">CYP2E1</span> gene from <span class="html-italic">L. yarkandensis</span> and a sequence comparison with the amino acids of <span class="html-italic">O. cuniculus</span>. (<b>A</b>) Agarose gel electrophoresis detection results of <span class="html-italic">L. yarkandensis CYP2E1</span> gene PCR products. (<b>B</b>) Comparison of individual amino acid numbers between <span class="html-italic">L. yarkandensis</span> and <span class="html-italic">O. cuniculus</span> CYP2E1 proteins. (<b>C</b>) Amino acid sequence alignment of <span class="html-italic">L. yarkandensis</span> and <span class="html-italic">O. cuniculus</span> CYP2E1 proteins.</p>
Full article ">Figure 2
<p>Alignment and phylogenetic analysis of CYP2E1. (<b>A</b>) Multiple sequence alignment analysis of CYP2E1. (<b>B</b>) Phylogenetic tree of CYP2E1 protein and homologous sequences (neighbor joining, NJ). The values on the branches of the phylogenetic tree are bootstrap values. Abbreviated names and GenBank accession numbers are as follows: <span class="html-italic">Bos taurus</span> (AAY83882.1), <span class="html-italic">Cervus elaphus hippelaphus</span> (OWK08011.1), <span class="html-italic">Ovis aries</span> (ADZ11094.1), <span class="html-italic">Sus scrofa</span> (NP_999586.1), <span class="html-italic">Felis catus</span> (NP_001041475.1), <span class="html-italic">Panthera pardus</span> (XP_019287846.2), <span class="html-italic">Homo sapiens</span> (NP_000764.1), <span class="html-italic">Macaca mulatta</span> (AAT49269.1), <span class="html-italic">Mus musculus</span> (NP_067257.1), and <span class="html-italic">O. cuniculus</span> (XP_002718818.1).</p>
Full article ">Figure 3
<p>Heterologous expression of <span class="html-italic">CYP2E1</span> gene in yeast. (<b>A</b>) Double digestion identification of pYES2 and pYES2-<span class="html-italic">CYP2E1</span>. (<b>B</b>) Western blotting detection of CYP2E1 protein expression levels in INVSc1-pYES2 and INVSc1-pYES2-<span class="html-italic">CYP2E1</span>.</p>
Full article ">Figure 4
<p>Physiological effects of heterologous expression of <span class="html-italic">CYP2E1</span> gene on yeast cells under mannitol stress. (<b>A</b>) Colony formation of yeast cells (INVSc1-pYES2 and INVSc1-pYES2-<span class="html-italic">CYP2E1</span>) under different concentrations of mannitol treatment. (<b>B</b>) Changes in SOD activity. (<b>C</b>) Changes in CAT activity. (<b>D</b>) Changes in ROS levels. (<b>E</b>) Changes in Pro levels. (<b>F</b>) Changes in MDA levels. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Physiological effects of heterologous expression of <span class="html-italic">CYP2E1</span> gene on yeast cells under NaCl stress. (<b>A</b>) Colony formation of INVSc1-pYES2 and INVSc1-pYES2-<span class="html-italic">CYP2E1</span> under different concentrations of NaCl treatment. (<b>B</b>) Changes in SOD activity. (<b>C</b>) Changes in CAT activity. (<b>D</b>) Changes in ROS levels. (<b>E</b>) Changes in Pro levels. (<b>F</b>) Changes in MDA levels. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Physiological effects of heterologous expression of <span class="html-italic">CYP2E1</span> gene on yeast cells under high-temperature stress. (<b>A</b>) Colony formation of INVSc1-pYES2 and INVSc1-pYES2-<span class="html-italic">CYP2E1</span> under different temperature treatments. (<b>B</b>) Changes in SOD activity. (<b>C</b>) Changes in CAT activity. (<b>D</b>) Changes in ROS levels. (<b>E</b>) Changes in Pro levels. (<b>F</b>) Changes in MDA levels. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Visualization of stress response steps in <span class="html-italic">CYP2E1</span>-transformed yeast.</p>
Full article ">
11 pages, 3220 KiB  
Article
A Novel Nitrite Reductase from Acinetobacter haemolyticus for Efficient Degradation of Nitrite
by Xiao-Yan Yin, Emmanuel Mintah Bonku, Jian-Feng Yuan and Zhong-Hua Yang
Biomolecules 2025, 15(1), 63; https://doi.org/10.3390/biom15010063 - 4 Jan 2025
Viewed by 367
Abstract
Nitrite reductases play a crucial role in the nitrogen cycle, demonstrating significant potential for applications in the food industry and environmental remediation, particularly for nitrite degradation and detection. In this study, we identified a novel nitrite reductase (AhNiR) from a newly [...] Read more.
Nitrite reductases play a crucial role in the nitrogen cycle, demonstrating significant potential for applications in the food industry and environmental remediation, particularly for nitrite degradation and detection. In this study, we identified a novel nitrite reductase (AhNiR) from a newly isolated denitrifying bacterium, Acinetobacter haemolyticus YD01. We constructed a heterologous expression system using E. coli BL21/pET28a-AhNir, which exhibited remarkable nitrite reductase enzyme activity of 29 U/mL in the culture broth, substantially higher than that reported for other strains. Structural analysis of AhNiR revealed the presence of [Fe-S] clusters, with molecular docking studies identifying Tyr-282 and Ala-289 as key catalytic sites. The enzymatic properties of AhNiR demonstrated an optimal pH of 7.5 and an optimal catalytic temperature of 30 °C. Its kinetic parameters, Km and vmax, were 1.53 mmol/L and 10.18 mmol/min, respectively, fitting with the Michaelis–Menten equation. This study represents the first report of a nitrite reductase from a denitrifying bacterium, providing a new enzyme source for nitrite degradation applications in the food industry and environmental remediation, as well as for biosensing technologies aimed at nitrite detection. Full article
(This article belongs to the Section Enzymology)
Show Figures

Figure 1

Figure 1
<p>Phylogenetic tree of A. haemolyticus YD01 based on 16SrDNA gene sequence and other related sequences.</p>
Full article ">Figure 2
<p>SDS-PAGE of heterologous expression <span class="html-italic">Ah</span>NiR. (<b>A</b>) Expressed <span class="html-italic">Ah</span>NiR in <span class="html-italic">E. coli</span> BL21 (Lane M: Protein Marker; Lane 1: <span class="html-italic">E. coli</span> BL21/pET28a-<span class="html-italic">Ah</span>NiR; Lane 2: <span class="html-italic">E. coli</span> BL21/pET28a); (<b>B</b>) Purified <span class="html-italic">Ah</span>NiR by Ni-NTA Purose 6 Fast Flow column affinity chromatography (Lane M: Protein Marker; Lane 1: <span class="html-italic">Ah</span>NiR crude enzyme solution; Lane 2: Tube 1 eluent; Lane 3: Tube 2 eluent; Lane 4: Tube 3 eluent).</p>
Full article ">Figure 3
<p><span class="html-italic">Ah</span>NiR enzymatic properties. (<b>A</b>) Effect of reaction temperature on the <span class="html-italic">Ah</span>NiR activity; (<b>B</b>) effect of reaction pH on the <span class="html-italic">Ah</span>NiR activity; (<b>C</b>) effect of ion on the <span class="html-italic">Ah</span>NiR activity; (<b>D</b>) Kinetic parameter simulation of <span class="html-italic">Ah</span>NiR; (<b>E</b>) thermal stability; (<b>F</b>) pH stability.</p>
Full article ">Figure 4
<p>Sequence alignment and secondary structure analysis of <span class="html-italic">Ah</span>NiR and <span class="html-italic">Mt</span>NiR.</p>
Full article ">Figure 5
<p><span class="html-italic">Ah</span>NiR structure by homologous modeling. (<b>A</b>) <span class="html-italic">Ah</span>NiR and (<b>B</b>) 1Zj9.</p>
Full article ">Figure 6
<p>Model of interaction between <span class="html-italic">Ah</span>NiR and substrate NO<sub>2</sub><sup>−.</sup></p>
Full article ">
18 pages, 2085 KiB  
Review
Lipoprotein Signal Peptide as Adjuvants: Leveraging Lipobox-Driven TLR2 Activation in Modern Vaccine Design
by Muhammad Umar, Haroon Afzal, Asad Murtaza and Li-Ting Cheng
Vaccines 2025, 13(1), 36; https://doi.org/10.3390/vaccines13010036 - 2 Jan 2025
Viewed by 881
Abstract
Toll-like receptor 2 (TLR2) signaling is a pivotal component of immune system activation, and it is closely linked to the lipidation of bacterial proteins. This lipidation is guided by bacterial signal peptides (SPs), which ensure the precise targeting and membrane anchoring of these [...] Read more.
Toll-like receptor 2 (TLR2) signaling is a pivotal component of immune system activation, and it is closely linked to the lipidation of bacterial proteins. This lipidation is guided by bacterial signal peptides (SPs), which ensure the precise targeting and membrane anchoring of these proteins. The lipidation process is essential for TLR2 recognition and the activation of robust immune responses, positioning lipidated bacterial proteins as potent immunomodulators and adjuvants for vaccines against bacterial-, viral-, and cancer-related antigens. The structural diversity and cleavage pathways of bacterial SPs are critical in determining lipidation efficiency and protein localization, influencing their immunogenic potential. Recent advances in bioinformatics have significantly improved the prediction of SP structures and cleavage sites, facilitating the rational design of recombinant lipoproteins optimized for immune activation. Moreover, the use of SP-containing lipobox motifs, as adjuvants to lipidate heterologous proteins, has expanded the potential of vaccines targeting a broad range of pathogens. However, challenges persist in expressing lipidated proteins, particularly within heterologous systems. These challenges can be addressed by optimizing expression systems, such as engineering E. coli strains for enhanced lipidation. Thus, lipoprotein signal peptides (SPs) demonstrate remarkable versatility as adjuvants in vaccine development, diagnostics, and immune therapeutics, highlighting their essential role in advancing immune-based strategies to combat diverse pathogens. Full article
(This article belongs to the Special Issue State-of-the-Art Vaccine Design)
Show Figures

Figure 1

Figure 1
<p>Signal peptides: structure and function in protein targeting (Created in <a href="https://BioRender.com" target="_blank">https://BioRender.com</a>—accessed on 15 November 2024).</p>
Full article ">Figure 2
<p>Lipidation and immune recognition: the role of Lgt, LspA, and Lnt in protein modification. Lipidation begins with Lgt adding lipids, resulting in diacylation in Gram-positive and triacylation in Gram-negative bacteria. LspA cleaves the signal peptide, and Lnt adds more lipids. These lipidated proteins activate immune signaling by engaging TLR2 with TLR1 or TLR6 (created with BioRender.com).</p>
Full article ">Figure 3
<p>Activation of TLR2 signaling pathway by bacterial lipoproteins (BLPs). The TLR2 signaling pathway is triggered by TLR2 forming heterodimers with TLR1 or TLR6 upon binding bacterial lipoproteins, recruiting adaptor proteins like MyD88 and TIRAP. This activates IRAKs and downstream NF-κB, driving pro-inflammatory cytokine production (created in <a href="https://BioRender.com" target="_blank">https://BioRender.com</a>—accessed on 15 November 2024).</p>
Full article ">Figure 4
<p>Signal sequence prediction by SignalP-6.0. The signal sequence was analyzed using bioinformatics, distinctly identifying the n, h, and c regions, along with precise marking of the cleavage site and lipobox (Created with <a href="https://services.healthtech.dtu.dk/services/SignalP-6.0/" target="_blank">https://services.healthtech.dtu.dk/services/SignalP-6.0/</a>, accessed on 10 November 2024).</p>
Full article ">Figure 5
<p>Lipidation of proteins: the critical role of the signal peptide lipobox. Signal peptide contains a lipobox which consists of a conserved cysteine residue at the +1 position, with [LVI][ASTVI][GAS]. This structural motif is essential for the attachment of lipid moieties. The E3 protein from the dengue virus has been cloned and expressed in an <span class="html-italic">E. coli</span> with SP lipobox, resulting in a lipidated form of the protein (Created in <a href="https://BioRender.com" target="_blank">https://BioRender.com</a>—accessed on 15 November 2024).</p>
Full article ">
Back to TopTop